Vous êtes sur la page 1sur 534

ERRATUM

Authors corrections after printing:


Foundation Engineering by S. Hansbo
ISBN 0.444.88549.8

REFERENCE LIST; COMPLEMENTARY ADDITION

Alphan, I., 1967. The empirical evaluation of the coefficient K0 and KQR. Jap. Soc. Soil
Mech. Found. Eng., Soil and Found., Vol. 7, No. 1, 31-40.
American Concrete Pipe Association, ACPA, 1988. Concrete pipe handbook. ISBN 0-90-
38681-6, Vienna, USA.
Bergado, D. T., Chai, C. T., Alfaro, M. C. & Balasubramaniam, A. S., 1992. Improvement
techniques of soft ground in subsiding and lowland environment. Asian Inst, of Technology,
Bangkok.
Berggren, B., 1981. Grvplar p friktionsjordsttningar och brfrmaga (Bored piles
on non-cohesive soilssettlement and bearing capacity). Ph. D. Thesis, Chalmers
University of Technology, Gothenburg.
Bustamante, M. G. & Gianeselli, L., 1981. Readjustment des paramtres des calculs des
pieux. Proc. 10th Int. Conf. Soil Mech. Found. Eng., Vol. 2, 643-646.
Cambefort, H., 1967. Injection des sols. Eyrolles, Vol. 1 and 2.
Chambosse, G. & Dobson, T., 1982. Stone columns IEstimation of bearing capacity and
expected settlement in cohesive soils. GKN Keller Inc., Tampa, Florida.
Caquot, ., Kerisel, J. & Absi, F., 1973. Tables de bute et de pousse. Gauthier-Villars,
Paris-Bruxelles-Montral.
Donchev, P., 1980. Compaction of loess by saturation and explosion. Int. Conf. on Com-
paction, Paris, Vol 1,313-317.
Esrig, M. J., 1968. Pore pressures, consolidation, and electrokinetics. Proc. ASCE, J. Soil
Mech. Found. Eng., Vol. 94, SM 4.
Hansbo, S., 1962. Ny konapparat fr bestmning av lerors skrhallfasthet. (A new cone
apparatus for determination of the shear strength of clays). Byggmstaren No. 10, 215-
220.
Hansbo, S., Hofmann, E. & Mosesson, J., 1973. stra Nordstaden, Gothenburg. Experiences
concerning a difficult foundation problem and its unorthodox solution. Proc. 8th Int.
Conf. Soil Mech. Found. Eng., Moscow, Vol. 2.2, 105-110.
Hansbo, S. & Jendeby, L., 1983. A case study of two alternative foundation principles:
conventional friction piling and creep piling. Vag- och Vattenbyggarcn, No.7-8,29-31.
Hansbo, S. & Kllstrm, R., 1983. Creep pilesa cost-effective alternative to conventional
friction piles. Vg- och vattenbyggaren No. 7-8, 23-27.
Hansbo, S., Pramborg, B. & Nordin, P. O., 1977. The Vnern terminal. Illustrative example
of dynamic consolidation of hydraulically placed fill of organic silt and sand. Sols Soils,
No. 25,5-11.
Hardin, B. O. & Black, W. L., 1969. Vibration modulus of normally consolidated clay.
(Closure). Proc. ASCE, J. Soil Mech. Found. Div., Vol. 95, No. SM 6, 1531-1537.
Hardin, B. O. & Richart, F. E. Jr., 1963. Elastic wave velocities in granular soils. Proc.
ASCE, J. Soil Mech. Found. Div., Vol. 89, No. SM 1, 35-65.
Jamiolkowski, M., Ladd, C. C , Germaine, J. T. & Lancelotta, R., 1985. New developments
in field and laboratory testing of soils. Proc. 11th Int. Conf. Soil Mech. Found. Eng., San
Francisco, Vol 1, 57-154.
Janbu, ., 1965. Consolidation of clay layers based on non-linear stress-strain. Proc. 6th
Int. Conf. Soil Mech. Found. Eng., Montreal, Vol. 2, 83-87.
Karol, R. H., 1960. Field tests for evaluating the effectiveness of a grouting operation. Am.
Cyan. Co Expl. and Min. Chem. Dep.
Larsson, R., 1975. Konsolidering av lera med elektroosmos. (Consolidation of clay by
means of electro-osmosis). Byggforskningen R45: 1975.
Larsson, R., 1977. Basic behaviour of Scandinavian soft clays. Swedish Geotech. Institute,
Report No. 4.
Liedberg, S., 1991. Earth pressure distribution against rigid pipes under various bedding
conditions. Ph. D. Thesis, Chalmers University of Technology, Gothenburg.
Littlejohn, G. S., 1992. Chemical grouting. In: M. P. Moseley (Editor), Ground improve-
ment, Blackie Academic & Professional, CRC Press, Inc., 100-129.
Maag, E., 1938.ber die Verfestigung und Dichtung des Baugrundes (Injektionen).
Erdbautechnik, .
Mesri, G. & Godlewski, P. M.,1977. Time- and stress-compressibility interrelationship.
ASCE, J. Geotech. Eng. Div., GT 5, 417-430.
Olander, H. C , 1950. Stress analysis of concrete pipe. US Bureau Reel. Eng. Monographs,
No. 6.
Pramborg, B. & Albertsson, B., 1992. Underskning av kalk/cementpelare. (Investigation
of lime/cement columns). SBUF-Anslag Projekt 1075.
Pusch, R., Hansbo, S., Berg, G. & Henricson, E., 1974. Brighet och sttningar vid
grundlggning p berg. (Bearing capacity and settlement when founding on rock).
Svenska Byggnadsentreprenrfreeningen, Report No. 11.
Schneider, P. J., 1963. Temperature response charts. New York/ London, Wiley.
Schmertmann, J. H., 1955. The undisturbed consolidation behaviour of clay. Transactions
ASCE, Vol. 120.
Schmertmann, J. H., Hartmann, J. P. & Brown, P. R., 1978. Improved strain influence
factor diagrams. ASCE, J. Soil Mech. Found. Div., No. GT 8.
Schulter, A. & Wagener, H., 1989. Improvement of clay and silt by dewatering with a new
anchoring technology. Proc. 12th Int. Conf. Soil Mech. Found. Eng., Vol. 2,1409-1414.
Simonini, P. & Sorenzo, M, 1987. Design and performance of piles driven into a soft
cohesive deposit. Proc. Int. Symp. on Geot. Eng. of Soft Soils, Mexico, Vol. 1,371-378.
Smith, W. W., 1978. Stresses in rigid pipe. ASCE, Transp. Eng. J., Vol. 104,No. TE 3.
Spangler, M. G., 1948. Underground conduitsAn appraisal of modem research. ASCE,
Trans., Vol. 113.
Svensson, P. L., 1991. Soil-structure interaction of foundations on soft clayExperience
during the last ten years. Proc. 10th European Conf. Soil Mech. Found. Eng., Florence,
Vol II, 583-586.
Szchy, ., 1965. Der Grundbau, Vol. 2, Part 1Die Baugrube. Springer-Verlag, Wien/
New York.
Sllfors, G., 1975. Preconsolidation pressure of soft high-plastic clays. Ph. D. Thesis,
Chalmers University of Technology, Gothenburg.
Terzaghi, K., 1923. Die Berechnung der Durchlssigkeitsziffer des Tones aus dem Verlauf
der hydrodynamischen Spannungserscheinungen. Akad. Wissensch., Wien, Sitzungsber.
Bd. 132, H. 1, Mat.-Naturwissensch. Klasse.
Terzaghi, K., 1925. Erdbaumechanik. Leipzig-Wien.
Terzaghi, K., 1943. Theoretical soil mechanics. New York.
stedt, ., Weiner, L. & Holm, G., 1990. Friktionsplar. Brfrmagans tillvxt med tiden.
(Friction piles. Increase in bearing capacity with time). Preprint, Swedish Geotech. Inst.,
Linkping.
CORRIGENDA

Cross - refe rences : Delete:


page; line from bottom - line from top + p. 247, bottom
65; 8 - (see paragraph 6.3) 'Furthermore, the maximum can be
91; 10- in par. 6.2 established.'
99; 3 - Section 5.3
p. 248, top
127,11+ Eq. (110)
mz{)o +cz0(u +kzo = QQ (298)
6+ (cf. pp. 275-277)
7- (p. 324) p. 227, 2 - (Poulus, 1990)
7- (seep. 179)
7+ Fig. 138 Other corrections:
16- Eq. (206) page
7- p. 192 77,3+ (Schmertmann, 1955)
2+ (Figs. 165-166) 79, 1- Mesri et al (1990)
5+ Eq. (10) 125, Eq. (107)
6+ Eqs. (299-300)
11- Fig. 105c ^-ds + Ipia^'^-ds + /sin(v + ')
os as
2- Eq. (286) 125, Fig. 92.
15+ Eq. (147) dv ,dv
11+ (Fig. 196) Replace 2pds by 2/?tan</> ds
ds os
18- Fig. 187
, ^ dv . , _ ,,v .
12+ (Fig. 218) and ipdrby 2/?tan0 dr
or or
15+ (p. 127)
i38;Ex. 13 1 (vert.): 1.5 (hor.)
8+ (Fig. 230)
195; Fig. 139 qc (in MPa)
11- Fig. 232
214, Eq. (250)
8- Fig. 80
4- (p. 116) da y
=o
11+ Fig. 80
1+ (Fig. 226) 217,9+ (Davisson et ai, 1965)
3- (Figs. 269-270) 220; Fig. 157 equal to 4bp
8+ Eqs. 365-366 221; Table 26 Initial ID%
5+ Eq. (12) 223; 12+ 60 mm by 60 mm
9- Eq. (12) 223; Fig. 159 equal to 4bp
9- Eq. (510) 224; Fig. 160 equal to Sbp
6- Eq. (510) and Fig. 320 228; 3+ Hansbo et ai, 1973
12- Eq. (513) 231, 4 - Randolph and Clancy, 1993
4- Eq. (513) 266; 2+
17+ (Fig. 353) The displacement amplitude at D is obtained
1- Eq. (545) when the force vector is pointing in the
11+ Fig. 353 direction.
342, Eq. (447)
Add: M s = Rlrl = RB(c'+ a'tantfO/^,
p. 113, after (Fig. 82). 344,6- F C ^ 1.05-1.06= 1.11
Count the number of tetragons covered by 353,5- FCQ= 1.05-1.06= 1.11
the loaded area. The stress is obtained by the
expression = 0.00\nq. 479, 5 -
The standardised form of normal distribution

(p(x)= = exp(- )
ERRATUM

Authors corrections after printing:


Foundation Engineering by S. Hansbo
ISBN 0.444.88549.8

REFERENCE LIST; COMPLEMENTARY ADDITION

Alphan, I., 1967. The empirical evaluation of the coefficient K0 and K0R. Jap. Soc. Soil
Mech. Found. Eng., Soil and Found., Vol. 7, No. 1, 31-40.
American Concrete Pipe Association, ACPA, 1988. Concrete pipe handbook. ISBN 0-90-
38681-6, Vienna, USA.
Bergado, D. T., Chai, C. T., Alfaro, M. C. & Balasubramaniam, A. S., 1992. Improvement
techniques of soft ground in subsiding and lowland environment. Asian Inst, of Technology,
Bangkok.
Berggren, B., 1981. Gravplar p friktionsjordsttningar och brfrmaga (Bored piles
on non-cohesive soilssettlement and bearing capacity). Ph. D. Thesis, Chalmers
University of Technology, Gothenburg.
Bustamante, M. G. & Gianeselli, L., 1981. Readjustment des paramtres des calculs des
pieux. Proc. 10th Int. Conf. Soil Mech. Found. Eng., Vol. 2, 643-646.
Cambefort, H., 1967. Injection des sols. Eyrolles, Vol. 1 and 2.
Chambosse, G. & Dobson, T., 1982. Stone columns IEstimation of bearing capacity and
expected settlement in cohesive soils. GKN Keller Inc., Tampa, Florida.
Caquot, ., Kerisel, J. & Absi, F., 1973. Tables de bute et de pousse. Gauthier-Villars,
Paris-Bruxelles-Montral.
Donchev, P., 1980. Compaction of loess by saturation and explosion. Int. Conf. on Com-
paction, Paris, Vol 1, 313-317.
Esrig, M. J., 1968. Pore pressures, consolidation, and electrokinetics. Proc. ASCE, J. Soil
Mech. Found. Eng., Vol. 94, SM 4.
Hansbo, S., 1962. Ny konapparat for bestmning av lerors skrhallfasthet. (A new cone
apparatus for determination of the shear strength of clays). Byggmstaren No. 10, 215-
220.
Hansbo, S., Hofmann, E. & Mosesson, J., 1973. stra Nordstaden, Gothenburg. Experiences
concerning a difficult foundation problem and its unorthodox solution. Proc. 8th Int.
Conf. Soil Mech. Found. Eng., Moscow, Vol. 2.2, 105-110.
Hansbo, S. & Jendeby, L., 1983. A case study of two alternative foundation principles:
conventional friction piling and creep piling. Vg- och Vattenbyggaren, No.7-8,29-31.
Hansbo, S. & Kllstrm, R., 1983. Creep pilesa cost-effective alternative to conventional
friction piles. Vg- och vattenbyggaren No. 7-8, 23-27.
Hansbo, S., Pramborg, B. & Nordin, P. O., 1977. The Vnern terminal. Illustrative example
of dynamic consolidation of hydraulically placed fill of organic silt and sand. Sols Soils,
No. 25,5-11.
Hardin, B. O. & Black, W. L., 1969. Vibration modulus of normally consolidated clay.
(Closure). Proc. ASCE, J. Soil Mech. Found. Div., Vol. 95, No. SM 6, 1531-1537.
Hardin, B. O. & Richart, F. E. Jr., 1963. Elastic wave velocities in granular soils. Proc.
ASCE, J. Soil Mech. Found. Div., Vol. 89, No. SM 1, 35-65.
Jamiolkowski, M., Ladd, C. C , Germaine, J. T. & Lancelotta, R., 1985. New developments
in field and laboratory testing of soils. Proc. 11th Int. Conf. Soil Mech. Found. Eng., San
Francisco, Vol 1,57-154.
Janbu, ., 1965. Consolidation of clay layers based on non-linear stress-strain. Proc. 6th
Int. Conf. Soil Mech. Found. Eng., Montreal, Vol. 2, 83-87.
Karol, R. H., 1960. Field tests for evaluating the effectiveness of a grouting operation. Am.
Cyan. Co Expl. and Min. Chem. Dep.
Larsson, R., 1975. Konsolidering av lera med elektroosmos. (Consolidation of clay by
means of electro-osmosis). Byggforskningen R45: 1975.
Larsson, R., 1977. Basic behaviour of Scandinavian soft clays. Swedish Geotech. Institute,
Report No. 4.
Liedberg, S., 1991. Earth pressure distribution against rigid pipes under various bedding
conditions. Ph. D. Thesis, Chalmers University of Technology, Gothenburg.
Littlejohn, G. S., 1992. Chemical grouting. In: M. P. Moseley (Editor), Ground improve-
ment, Blackie Academic & Professional, CRC Press, Inc., 100-129.
Maag, E., 1938.ber die Verfestigung und Dichtung des Baugrundes (Injektionen).
Erdbautechnik, .
Mesri, G. & Godlewski, P. M.,1977. Time- and stress-compressibility interrelationship.
ASCE, J. Geotech. Eng. Div., GT 5, 417-430.
Olander, H. C , 1950. Stress analysis of concrete pipe. US Bureau Reel. Eng. Monographs,
No. 6.
Pramborg, B. & Albertsson, B., 1992. Underskning av kalk/cementpelare. (Investigation
of lime/cement columns). SBUF-Anslag Projekt 1075.
Pusch, R., Hansbo, S., Berg, G. & Henricson, E., 1974. Brighet och sttningar vid
grundlggning p berg. (Bearing capacity and settlement when founding on rock).
Svenska Byggnadsentreprenrfreeningen, Report No. 11.
Schneider, P. J., 1963. Temperature response charts. New York/ London, Wiley.
Schmertmann, J. H., 1955. The undisturbed consolidation behaviour of clay. Transactions
ASCE, Vol. 120.
Schmertmann, J. H., Hartmann, J. P. & Brown, P. R., 1978. Improved strain influence
factor diagrams. ASCE, J. Soil Mech. Found. Div., No. GT 8.
Schulter, A. & Wagener, H., 1989. Improvement of clay and silt by dewatering with a new
anchoring technology. Proc. 12th Int. Conf. Soil Mech. Found. Eng., Vol. 2,1409-1414.
Simonini, P. & Sorenzo, M, 1987. Design and performance of piles driven into a soft
cohesive deposit. Proc. Int. Symp. on Geot. Eng. of Soft Soils, Mexico, Vol. 1,371-378.
Smith, W. W., 1978. Stresses in rigid pipe. ASCE, Transp. Eng. J., Vol. 104,No. TE 3.
Spangler, M. G., 1948. Underground conduitsAn appraisal of modem research. ASCE,
Trans., Vol. 113.
Svensson, P. L., 1991. Soil-structure interaction of foundations on soft clayExperience
during the last ten years. Proc. 10th European Conf. Soil Mech. Found. Eng., Florence,
Vol II, 583-586.
Szchy, ., 1965. Der Grundbau, Vol. 2, Part 1Die Baugrube. Springer-Verlag, Wien/
New York.
Sllfors, G., 1975. Preconsolidation pressure of soft high-plastic clays. Ph. D. Thesis,
Chalmers University of Technology, Gothenburg.
Terzaghi, K., 1923. Die Berechnung der Durchlssigkeitsziffer des Tones aus dem Verlauf
der hydrodynamischen Spannungserscheinungen. Akad. Wissensch., Wien, Sitzungsber.
Bd. 132, H. 1, Mat.-Naturwissensch. Klasse.
Terzaghi, K., 1925. Erdbaumechanik. Leipzig-Wien.
Terzaghi, K., 1943. Theoretical soil mechanics. New York.
stedt, ., Weiner, L. & Holm, G., 1990. Friktionsplar. Brfrmagans tillvxt med tiden.
(Friction piles. Increase in bearing capacity with time). Preprint, Swedish Geotech. Inst.,
Linkping.
CORRIGENDA

Cross - refe rences : Delete:


page; line from bottom - , line from top + p. 247, bottom
65; 8 - (see paragraph 6.3) 'Furthermore, the maximum can be
91; 10- in par. 6.2 established.'
99; 3 - Section 5.3
127 11 + Eq.(llO) p. 248, top
134 6+ (cf. pp. 275-277) mzQuP- +cz0co +kzo = QQ (298)
154 7 - (p. 324) p. 227, 2 - (Poulus, 1990)
162 7 - (see p. 179)
192 7+ Fig. 138 Other corrections:
201 16- Eq. (206) page
229 7 - p. 192 77, 3+ (Schmertmann, 1955)
234 2+ (Figs. 165-166) 79, 1- Mesri et al (1990)
241 5+ Eq.(10) 125, Eq. (107)
248 6+ Eqs. (299-300)
251 11- Fig. 105c %ds + 2 / ? t a n 0 ' ^ d s + / sin(v + ')
os ds
259 2 - Eq. (286) 125, Fig. 92.
264 15+ Eq. (147)
280 11+ (Fig. 196) Replace 2pds by 2/?tan0 ds
os ds
294 18- Fig. 187
, ^ v , , ^ ,v ,
305 12+ (Fig. 218) and zpdrby 2/?tan0 dr

314 15+ (p. 127) or or


320 8+ 138; Ex. 13 1 (vert.): 1.5 (hor.)
(Fig. 230)
321 11- Fig. 232 195; Fig. 139 ^ c(inMPa)
323 8 - Fig. 80 214, Eq. (250)
323 4 - (p.116) =0
329 11+ Fig. 80
334 1+ (Fig. 226) 217, 9+ (Davisson et al, 1965)
376 3 - (Figs. 269-270) 220; Fig. 157 equal to 4bp
389 8+ Eqs. 365-366 221; Table 26 Initial ID%
400 5+ Eq.(12) 223; 12+ 60 mm by 60 mm
401 9 - Eq.(12) 223; Fig. 159 equal to 4bp
434 9 - Eq. (510) 224; Fig. 160 equal to Sbp
444 6 - Eq. (510) and Fig. 320 228; 3+ Hansboeia/., 1973
445 12- Eq. (513) 231, 4 - RandolphandClancy, 1993
449 4 - Eq. (513) 266; 2+
488 17+ (Fig. 353) The displacement amplitude at D is obtained
489 1- Eq. (545) when the force vector is pointing in the
490 11+ Fig. 353 direction.
342, Eq. (447)
Add: MS = #/ = //(c'+ a'tan^O/^
p. 113, after (Fig. 82). 344,6- Fc(p~ 1.05-1.06 = 1.11
Count the number of tetragons covered by 353, 5 - F C ^ 1.05-1.06= 1.11
the loaded area. The stress is obtained by the
expression = 0.00 \nq. 479, 5 -
The standardised form of normal distribution

(p(x) = - 7 = e x p ( - 2)
72
ERRATUM

Authors corrections after printing:


Foundation Engineering by S. Hansbo
ISBN: 0.444.88549.8

R E F E R E N C E LIST; C O M P L E M E N T A R Y A D D I T I O N

Alphan, I., 1967. The empirical evaluation of the coefficient K0 and K0R. Jap. Soc. Soil Mech.
Found. Eng., Soil and Found., Vol. 7, No. 1, 31-40.
American Concrete Pipe Association, ACPA, 1988. Concrete pipe handbook. ISBN 0-90-
38681-6, Vienna, USA.
Bergado, D. T., Chai, C. T., Alfaro, M. C. & Balasubramaniam, A. S., 1992. Improvement
techniques of soft ground in subsiding and lowland environment. Asian Inst, of Technology,
Bangkok.
Berggren, B., 1981. Gravplar p friktionsjordsttningar och brfrmaga (Bored piles on
non-cohesive soilssettlement and bearing capacity). Ph. D. Thesis, Chalmers University
of Technology, Gothenburg.
Bustamante, M. G. & Gianeselli, L., 1981. Readjustment des paramtres des calculs des
pieux. Proc. 10th Int. Conf. Soil Mech. Found. Eng., Vol. 2, 643-646.
Cambefort, H., 1967. Injection des sols. Eyrolles, Vol. 1 and 2.
Chambosse, G. & Dobson, T., 1982. Stone columns IEstimation of bearing capacity and
expected settlement in cohesive soils. GKN Keller Inc., Tampa, Florida.
Caquot, ., Kerisel, J. & Absi, F., 1973. Tables de bute et de pousse. Gauthier-Villars,
Paris-Bruxelles-Montral
Esrig, M. J., 1968. Pore pressures, consolidation, and electrokinetics. Proc. ASCE, J. Soil
Mech. Found. Eng., Vol. 94, SM 4.
Hansbo, S. & Jendeby, L., 1983. A case study of two alternative foundation principles:
conventional friction piling and creep piling. Vg- och Vattenbyggaren, N o . 7 - 8 , 2 9 - 3 1 .
Hansbo, S. & Kllstrm, R., 1983. Creep pilesa cost-effective alternative to conventional
friction piles. Vg- och vattenbyggaren No. 7-8, 23-27.
Hansbo, S., Pramborg, B. & Nordin, P. O., 1977. The Vnern terminal. Illustrative example
of dynamic consolidation of hydraulically placed fill of organic silt and sand. Sols Soils,
No. 25, 5 - 1 1 .
Hardin, B. O. & Black, W. L., 1969. Vibration modulus of normally consolidated clay.
(Closure). Proc. ASCE, J. Soil Mech. Found. Div., Vol. 95, No. SM 6, 1531-1537.
Hardin, B. O. & Richart, F. E. Jr., 1963. Elastic wave velocities in granular soils. Proc.
ASCE, J. Soil Mech. Found. Div., Vol. 89, No. SM 1, 35-65.
Karol, R. H., 1960. Field tests for evaluating the effectiveness of a grouting operation. Am.
Cyan. Co Expl. and Min. Chem. Dep.
Larsson, R., 1977. Basic behaviour of Scandinavian soft clays. Swedish Geotech. Institute,
Report No. 4.
Liedberg, S., 1991. Earth pressure distribution against rigid pipes under various bedding
conditions. Ph. D. Thesis, Chalmers University of Technology, Gothenburg.
Littlejohn, G. S., 1992. Chemical grouting. In: M. P. Moseley (Editor), Ground improve-
ment, Blackie Academic & Professional, CRC Press, Inc., 100-129.
Mesri, G. & Godlewski, P. M., 1977. Time- and stress-compressibility interrelationship.
ASCE, J. Geoteh. Eng. Div., GT 5,417-430.
Olander, H. C , 1950. Stress analysis of concrete pipe. US Bureau Reel. Eng. Monographs,
No. 6.
Schneider, P. J., 1963. Temperature response charts. New York/ London, Wiley.
Schmertmann, J. H., 1955. The undisturbed consolidation behaviour of clay. Transactions
ASCE, Vol. 120.
Schmertmann, J. H., Hartmann, J. P. & Brown, P. R., 1978. Improved strain influence factor
diagrams. ASCE, J. Soil Mech. Found. Div., No. GT 8.
Simonini, P. & Sorenzo, M, 1987. Design and performance of piles driven into a soft
cohesive deposit. Proc. Int. Symp. on Geot. Eng. of Soft Soils, Mexico, Vol. 1,371-378.
Smith, W. W., 1978. Stresses in rigid pipe. ASCE, Transp. Eng. J., Vol. 104,No. TE 3.
Spangler, M. G., 1948. Underground conduitsAn appraisal of modem research. ASCE,
Trans., Vol. 113.
Svensson, P. L., 1991. Soil-structure interaction of foundations on soft clayExperience
during the last ten years. Proc. 10th European Conf. Soil Mech. Found. Eng., Florence,
Vol II, 583-586.
Szchy, ., 1965. Der Grundbau, Vol. 2, Part 1Die Baugrube. Springer-Verlag, Wien/
New York.
Sllfors, G., 1975. Preconsolidation pressure of soft high-plastic clays. Ph. D. Thesis,
Chalmers University of Technology, Gothenburg.
Terzaghi, K., 1923. Die Berechnung der Durchlssigkeitsziffer des Tones aus dem Verlauf
der hydrodynamischen Spannungserscheinungen. Akad. Wissensch., Wien, Sitzungsber.
Bd. 132, H. 1, Mat.-Naturwissensch. Klasse.
Terzaghi, K., 1925. Erdbaumechanik. Leipzig-Wien.
Terzaghi, K., 1943. Theoretical soil mechanics. New York.
stedt, ., Weiner, L. & Holm, G., 1990. Friktionsplar. Brfrmagans tillvxt med tiden.
(Friction piles. Increase in bearing capacity with time). Preprint, Swedish Geotech. Inst.,
Linkping.
CORRIGENDA

Cross-references: Delete:

page; line from bottom - , line from top + p. 247, bottom


65; 8 - (see paragraph 6.3) 'Furthermore, the maximum can be
91; 1 0 - in par. 6.2 established.'
99; 3 - Section 5.3
p. 248, top
201; 16- Eq. (206)
mz^fi +cz0Q) +kz0 = o (298)
229; 7 - p. 192
234; 2+ (Figs. 165-166) p. 2 2 7 , 2 - (Poulus, 1990)
241; 5+ Eq. (10)
248; 6+ Eqs. (299-300)
251; 1 1 - Fig. 105c Other corrections:
259; 2 - Eq. (286)
page
264; 15+ Eq. (147)
77, 3+ (Schmertmann, 1955)
280; 11+ (Fig. 196)
79, 1 - M e s n e / / . (1990)
294; 1 8 - Fig. 187
138; Ex. 13 1 (vert.): 1.5 (hor.)
305; 12+ (Fig. 218)
195; Fig. 139 qc (in MPa)
314; 15+ (p. 127)
217, 9+ (Davisson et al, 1965)
320; 8+ (Fig. 230)
220; Fig. 157 equal to 4bp
321; 1 1 - Fig. 232
221; Table 26 Initial ID %
323; 8 - Fig. 80
223;12+ 60 mm by 60 mm
323; 4 - (p. 116)
223; Fig. 159 equal to 4b
329; 11+ Fig. 80 equal to Sbp
224; Fig. 160
334; 1+ (Fig. 226) RandolphandClancy, 1993
231,4-
376; 3 - (Figs. 269-270)
214, Eq. (250)
389; 8+ Eqs. 365-366
400; 5+ Eq.(12) W 0
401; 9 - Eq. (12) dX ~ * %/ +
434; 9 - Eq. (510) 266; 2+
444; 6 - Eq. (510) and Fig. 320 The displacement amplitude at D is obtained
445; 1 2 - Eq.(513) when the force vector is pointing in the
449; 4 - Eq. (513) direction.
488; 17+ (Fig. 353)
489; 1 - Eq. (545)
490; 11+ Fig. 353

Add:

p. 113, after (Fig. 82).


Count the number of tetragons covered by
the loaded area. The stress is obtained by the
expression = O.OOlnq.
D e v e l o p m e n t s in G e o t e c h n i c a l E n g i n e e r i n g , 7 5

Foundation Engineering

Sven Hansbo

Lyckov2, Stocksund, S-18274, Sweden

ELSEVIER

Amsterdam - London - New York - Tokyo 1994


Further titles in this series:
Volumes 2, 3, 5-7, 9 10, 12, 13, 15, 16A, 22 and 26 are out of print

I. G. SANGLERAT THE PENETROMETER A N D SOIL EXPLORATION


4. R. SILVESTER COASTAL ENGINEERING, 1 AND 2
8. L.N. PERSEN ROCK DYNAMICS AND GEOPHYSICAL EXPLORATION
Introduction to Stress Waves in Rocks
II. H.K. GUPTA AND B.K. RASTOGI DAMS A N D EARTHQUAKES
14. B. VOIGHT (Editor) ROCKSLIDES AND AVALANCHES, 1 and 2
17. A.P.S. SELVADURAI ELASTIC ANALYSIS OF SOIL-FOUNDATION INTERACTION
18. J . FEDA STRESS IN SUBSOIL AND METHODS OF FINAL SETTLEMENT CALCULATION
19. . KZDI STABILIZED EARTH ROADS
20. E.W. BRAND A N D R.P. BRENNER (Editors) SOFT-CLAY ENGINEERING
21. A. MYSLIVE A N D Z. KYSELA THE BEARING CAPACITY OF BUILDING FOUNDATIONS
23. P. BRUUN STABILITY OF TIDAL INLETS Theory and Engineering
24. Z. BAZANT METHODS OF FOUNDATION ENGINEERING
25. . KZDI SOIL PHYSICS Selected Topics
27. D. STEPHENSON ROCKFILL IN HYDRAULIC ENGINEERING
28. P.E. FRIVIK, N. J A N B U , R. SAETERSDAL A N D L.I. FINBORUD (Editors) G R O U N D FREEZING 1980
29. P. PETER CANAL AND RIVER LEVES
30. J . FEDA MECHANICS OF PARTICULATE MATERIALS The Principles
31. Q. ZRUBA AND V. MENCL LANDSLIDES A N D THEIR CONTROL
Second completely revised edition
32. I.W. FARMER (Editor) STRATA MECHANICS
33. L. HOBST AND J. ZAJC ANCHORING IN ROCK AND SOIL
Second completely revised edition
34. G. SANGLERAT, G. OLIVARI AND B. C A M B O U PRACTICAL PROBLEMS IN SOIL MECHANICS AND
FOUNDATION ENGINEERING, 1 and 2
35. L. RTHTI GROUNDWATER IN CIVIL ENGINEERING
36. S.S. VYALOV RHEOLOGICAL FUNDAMENTALS OF SOIL MECHANICS
37. P. BRUUN (Editor) DESIGN AND CONSTRUCTION OF M O U N D S FOR BREAKWATERS AND COASTAL
PROTECTION
38. W.F. CHEN AND G.Y. BALADI SOIL PLASTICITY Theory and Implementation
30. E T . HANRAHAN THE GEOTECTONICS OF REAL MATERIALS: THE e g E k METHOD
40. J. ALDORF AND K. EXNER MINE OPENINGS Stability and Support
41. J.E. GILLOT CLAY IN ENGINEERING GEOLOGY
42. A.S. C A K M A K (Editor) SOIL DYNAMICS AND LIQUEFACTION
43. A.S. C A K M A K (Editor) SOIL-STRUCTURE INTERACTION
44. A.S. C A K M A K (Editor) G R O U N D MOTION AND ENGINEERING SEISMOLOGY
45. A.S. C A K M A K (Editor) STRUCTURES, UNDERGROUND STRUCTURES, DAMS, AND STOCHASTIC
METHODS
46. L. RTHTI PROBABILISTIC SOLUTIONS IN GEOTECTONICS
47. B.M. DAS THEORETICAL FOUNDATION ENGINEERING
48. W. DERSKI, R. IZBICKI, I. KISIEL AND Z. MROZ ROCK AND SOIL MECHANICS
49. T. ARIMAN, M. HAMADA, A.C. SINGHAL, M A . HAROUN AND A.S. C A K M A K (Editors) RECENT A D V A N -
CES IN LIFELINE EARTHQUAKE ENGINEERING
50. B.M. DAS EARTH A N C H O R S
51. K. THIEL ROCK MECHANICS IN HYDROENGINEERING
52. W.F. CHEN AND X . L LIU LIMIT ANALYSIS IN SOIL MECHANICS
53. W.F. CHEN AND E. MIZUNO NONLINEAR ANALYSIS IN SOIL MECHANICS
54. F.H. CHEN FOUNDATIONS ON EXPANSIVE SOILS
55. J . VERFEL ROCK GROUTING AND DIAPHRAGM W A L L CONSTRUCTION
56. B.N. WHITTAKER AND D.J. REDDISH SUBSIDENCE Occurrence, Prediction and Control
57. E. NONVEILLER GROUTING, THEORY AND PRACTICE
58. V.KOUXR AND I. EM EC MODELLING OF SOIL-STRUCTURE INTERACTION
59A. R.S. SINHA (Editor) UNDERGROUND STRUCTURES Design and Instrumentation
59B. R.S. SINHA (Editor) UNDERGROUND STRUCTURES Design and Construction
60. R.L. HARLAN, K.E. K O L M AND E.D. GUTENTAG WATER-WELL DESIGN AND CONSTRUCTION
61. I. KASDA FINITE ELEMENT TECHNIQUES IN GROUNDWATER FLOW STUDIES
62. L. FIALOVSZKY (Editor) SURVEYING INSTRUMENTS AND THEIR OPERATIONAL PRINCIPLES
63. H. GIL THE THEORY OF STRATA MECHANICS
64. H.K. GUPTA RESERVOIR-INDUCED EARTHQUAKES
65. V.J. LUNARDINI HEAT TRANSFER WITH FREEZING AND THAWING
66. T.S. NAGARAI PRINCIPLES OF TESTING SOILS, ROCKS A N D CONCRETE
67. E. JUHSOV SEISMIC EFFECTS ON STRUCTURES
68. J . FEDA CREEP OF SOILS A n d Related Phenomena
69. E. D U U \ C S K A SOIL SETTLEMENT EFFECTS O N BUILDINGS
70. D. MILOVIC STRESSES A N D DISPLACEMENTS FOR SHALLOW FOUNDATIONS
71. B.N. WHITTAKER, R.N. SINGH AND G. SUN ROCK FRACTURE MECHANICS Principles, Design and
Applications
72. M.A. MAHTAB AND P. GRASSO - GEOMECHANICS PRINCIPLES IN THE DESIGN OF TUNNELS AND
CAVERNS IN ROCK
73. R.N. YONG, A.M.O. M O H A M E D AND B.P. WARKENTIN - PRINCIPLES OF CONTAMINANT TRANSPORT
IN SOILS
74. H. BURGER (Editor) - OPTIONS FOR TUNNELLING 1993
ELSEVIER SCIENCE B.V.
Sara Burgerhartstraat 25
P.O. Box 2 1 1 , 1000 AE Amsterdam, The Netherlands

Library of Congress Catalog1ng-1n-PublIcatIon Data

Hansbo, Sven, 1924-


Foundatlon engineering / Sven Hansbo.
p. cn. (Developments In g e o t e c h n 1 ca1 engineering ; 75)
Includes bibliographical references and Index.
ISBN 0-444-88549-8
1. Foundations. I. T i t l e . I I . Series.
TA775.H36 1994
624. 1' 5dc20 93-43290
CIP

ISBN 0-444-88549-8

1994 Elsevier Science B.V. All rights reserved.

No part of this publication may be reproduced, stored in a retrieval system of transmitted in any
form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the
prior written permission of the publisher, Elsevier Science B.V., C o p y r i g h t and Permissions
Department, P.O. Box 5 2 1 , 1000 AM Amsterdam, The Netherlands.

Special regulations for readers in the U.S.A. - This publication has been registered with the
Copyright Clearance Center Inc. (CCC), Salem, Massachusetts. Information can be obtained from
the CCC about conditions under which photocopies of parts of this publication may be made in
the U.S.A. All other copyright questions, including photocopying outside of the U.S.A., should be
refered to the publisher.

No responsibility is assumed by the publisher for any injury and/or damage to persons or property
as a matter of products liability, negligence or otherwise, or from any use or operation of any
methods, products, instructions or ideas contained in the material herein.

This book is printed on acid-free paper.

Printed in The Netherlands


Preface xvii

PREFACE

This b o o k is a synthesis of m o r e than 30 years of experience as professor of Geotechnical


engineering at C h a l m e r s University of Technology and as geotechnical chief consultant
of J & W, o n e of the largest consulting firms in civil engineering in Scandinavia. It includes
a general review of those parts in soil and rock mechanics that are d e e m e d necessary in
the geotechnical design of foundations. Its main purpose, however, is to shed light on the
most important design aspects encountered in foundation engineering and to present
basic design principles. There is an abundance of design methods presented in the
literature, and it w o u l d lead too far to include but a restricted n u m b e r of them. I have
chosen to include only those which, from my personal experience, give acceptable
agreement with practice. Of course, other methods of tackling the foundation design than
those included here m a y be just as reliable and my choice cannot be regarded as an
intention to belittle possible alternative solutions. Thus, local experience of a certain
design m e t h o d can certainly justify its application.
D u e to increasing urbanisation there is an increasing d e m a n d for building sites, and
ground with very p o o r soil properties m a y have to be utilised for building purposes with
heavy initial capital costs. T h e costs, however, can generally be greatly reduced by the
use of ground modification techniques. In consequence, these have b e c o m e an integral
part of foundation engineering and have to be considered as possible m e a n s of reducing
capital costs in the building industry. In this textbook, the most important design and
practical aspects of soil i m p r o v e m e n t h a v e been included. M u c h of the material included
is based on m y personal experience.
T h e b o o k can b e used as a textbook for senior undergraduate and graduate students.
It can also serve as a c o m b i n e d text- and handbook for professional engineers working
in the field of geotechnical engineering.
All the line d r a w i n g s in the b o o k are d r a w n by hand. Photos presented in the b o o k h a v e
been a s s e m b l e d from time to time since the beginning of the 1960s and it is n o w im-
possible for m e to give credit to all the photographers in question. M a n y photos h a v e been
received from S w e d i s h and European contractors. A m o n g Swedish contractors, not
particularly m e n t i o n e d in the text, w h o have contributed photos I w o u l d like to
xviii Preface

a c k n o w l e d g e the assistance received from Dynapac, Geodynamik, Grundfrstrkningar,


Hagconsult, N C C and Stabilator.
I apologise to all those w h o have contributed in one way or the other without being
mentioned.
I express m y sincere thanks to my son Jonas w h o taught m e h o w to write computational
programs and h o w to use the layout program PageMaker. I also express m y thanks to
Kjell Ntterdahl for pulling copyproofs of all the line drawings and to all m y other
colleagues at the Geotechnical Department, Chalmers University of Technology, for
fruitful discussions and support. The assistance received from m y colleagues at J & W is
gratefully acknowledged.
Finally, I express m y thanks to Tina Foulder, w h o has proof-read the text from a
linguistic point of view.

Sven Hansbo
Introduction 1

INTRODUCTION
(i) Soil-structure interaction. T h e aim of foundation engineering is to find solutions to the
foundation p r o b l e m s which safeguard structural stability and life-long serviceability of
the building in a c c o r d a n c e with the o w n e r ' s requirements. In the design of a foundation,
it has to b e realised that the supporting soil, the foundation and the superstructure form
one single unit. Therefore, the structural engineer and the geotechnical engineer must
work in close cooperation to find the best possible solution under the given circumstances.
In order to avoid excessive settlements, causing d a m a g e to the building or cutting down
its serviceability, the building will h a v e to be founded on soil strata that are strong enough
to carry the building load. A c o m m o n task for the geotechnical and structural engineers
is to choose a foundation system that does not cause unacceptable deformations of the
superstructure. T h e choice b e t w e e n a shallow foundation on, for e x a m p l e , footings
placed directly in the soil at shallow depth, and a deep foundation by, for e x a m p l e , the
use of piles depends on the geotechnical characteristics of the subsoil and on the
architectural and functional requirements placed on the building. Evidently, a thorough
k n o w l e d g e of the geotechnical and geological circumstances at the building site is
imperative. Only then can a reliable analysis be m a d e of the bearing capacity of a certain
type of foundation and of the settlements to b e expected. T h e m o r e reliable the geotech-
nical basis for analysis, the better the possibilities of choosing the best foundation
solution from a technical and e c o n o m i c a l point of view.
In cases w h e r e piling m a y s e e m necessary, soil i m p r o v e m e n t w h i c h m a k e s shallow
foundation possible directly in the soil m a y b e the m o s t cost-effective solution.
T h e loads acting on conventional buildings are mainly vertical by nature. Horizontal
loads can easily b e taken care of by suitable design of the superstructure. Their influence
on foundation design is usually negligible. However, in the case of bridge and quay
foundations, and m o s t certainly in the case of retaining walls, the influence of horizontal
forces on the foundation design cannot b e neglected.
Building activities often m e a n raising the soil level around the prospective building.
If the soil at the building site is highly compressible, such an operation m a y entail serious
settlement p r o b l e m s , both d u e to the load increase on the subsoil and to long-term
settlement of the fill itself. B y the application of soil i m p r o v e m e n t m e t h o d s , this kind of
problem can b e avoided.
2 Introduction

(ii) Creating a reliable basis for design. One of the most important parts in the design is
to establish a reliable picture of the soil conditions at the building site, both from a
geotechnical and a geological point of view. Therefore, the planning and the realisation
of the site investigation are vital for successful design. This fact is often neglected. From
the client's viewpoint, the money spent on soil investigations is unprofitable and should
therefore be minimised. This has entailed a common procedure of inviting tenders for soil
investigations, which doubtless hazards the information needed. It is but natural that the
economic pressure exerted on the field crew in an investigation received by tender can
have a negative influence on the results obtained. This may be the case even where the
skill of the field personnel is beyond question. The extent of the investigation, and even
the method of investigation, may also have to be modified with regard to the results
obtained. These facts speak against tendering; they also speak for a very close co-
operation between the geotechnical expert on the one hand and the field investigation
crew on the other. In reality, for liability reasons, the field crew and the geotechnical
expert, who is responsible for the interpretation of the results obtained, should preferably
belong to the same organisation.
A correct determination of the strength and deformation properties of the soil is in fact
one of the most difficult tasks in geotechnical engineering. Their determination has to be
coupled with just the type of problem that is encountered in the design. In most cases in
practice, in situ investigations are preferable to laboratory investigations. However, for
determination of long-term deformation properties of cohesive soils, laboratory inves-
tigations are preferable to in situ ones. Sampling and laboratory investigations are also
required for the reason of soil classification which is necessary for the final assessment
of the geomechanical properties of the soil. Unfortunately, the classification systems vary
in different parts of the world, and this situation most probably will persist due to local
tradition and local soil conditions.

(iii) Execution. The execution of the foundation often entails the need for excavations
being carried out to a great depth. The problems connected with excavation for founda-
tion purposes often represent the most difficult and dangerous part of the j o b . Deep
excavations, with regard to slope and bottom stability, support of vertical cuts, etc. are
therefore just as important a part of the design as that of the foundation itself.
Quite often, provisional structures, necessary for the support of vertical cuts, can also
be utilised as structural members of the building itself. This is, for example, the case with
diaphragm walls. The borderline between the method of foundation in itself and the
execution of the excavation for the foundation has more or less vanished.
Evidently, to be a competent foundation engineer, broad knowledge of soil and struc-
tural mechanics is imperative. Furthermore, knowledge of geology is extremely important
for a correct assessment of the possible variation in geotechnical properties to be
expected at a building site.
Introduction 3

(i ) International aspects. F r o m an international point of view, the solution to foundation


problems depend on local cost of labour, tradition, available building material, level of
geotechnical education and codes of practice. It is therefore impossible to give, in a text
book, a complete list of foundation and foundation design m e t h o d s that w o u l d satisfy all
readers with their various backgrounds.
In this text book, m o d e r n geotechnical investigation m e t h o d s and their interpretation
are exemplified. T h e foundation m e t h o d s are representative of the developed part of the
world. T h e theoretical approach is influenced by the results of research carried through
at Swedish universities and research institutes and by experience gained as a geotechnical
consultant for m o r e than 35 years.
T h e design of foundations has to b e carried out in accordance with the prescriptions
presented in the building code of the country concerned. In Europe, a new code for geo-
technical design, the so-called E u r o c o d e 7, is now ready for publication. T h e philosophy
behind this code will b e presented in the last chapter of this book. In order not to create
confusion, safety aspects will only be treated in exceptional cases and then in a m o r e
traditional way. However, the safety philosophy, in the Eurocode, for e x a m p l e , can be
easily introduced without affecting the essence of the text.

(v)Aim of the book. In this textbook, those parts of soil mechanics are included which are
deemed to b e specially important in relation to the problems encountered in foundation
engineering. Certain parts are included for the purpose of increasing the r e a d e r ' s physical
understanding of the mechanical behaviour of soil and rock. T h e intention is to m a k e it
easier for the reader to grasp the context of the b o o k without need of looking into refer-
ence books.
N o w a d a y s , c o m p u t e r analyses (finite element analysis, b o u n d a r y element analysis or
finite difference analysis) have b e c o m e popular and widely used in geotechnical design.
Results obtained by back-analysis, presented in the literature, are usually in very close
agreement with real behaviour. As yet, however, prediction of the results to b e expected
in foundation design by computer analysis is rarely m o r e reliable than the results obtained
by simple ' h a n d calculations'. With regard to the uncertainties in soil modelling for com-
puter analyses, these will not b e treated in the present textbook. T h o s e interested are
referred to a text-book by B o w l e s (1988).
With m o d e r n scientific calculators it is very simple to solve seemingly complicated
implicit formulae which in fact reduces the need of diagrams. All the same, diagrams are
helpful under certain circumstances, so in this textbook both formulae (with derivation,
if deemed necessary) and corresponding diagrams are given in parallel.
4 Fundamentals

FUNDAMENTALS

1. STRUCTURAL FEATURES OF SOILS

Basic k n o w l e d g e of the features of soil structure is the key to understanding the


mechanical behaviour of soils. Of course, considerable structural variations of different
origin and nature exist in different parts of the world. T h o r o u g h k n o w l e d g e of local
geological conditions is therefore imperative. H e r e only s o m e general aspects of the
influence of soil structure on the mechanical behaviour and consequently on the
foundation analysis will be presented.
Soil is a three-phase material with a skeleton of particles in a certain structural
arrangement and open spaces (voids) filled with water and gas. F r o m a structural view-
point, w e differ between macro-structure, visible to the naked eye, and micro-structure
which can only be observed through a magnifying glass or a microscope.

1.2 Micro-structure

T h e possibility of studying the influence of micro-structure on the mechanical behaviour


of soils has greatly i m r o v e d due to the development of electron microscopes, particularly
the scanning and the transmission types. These, in combination with new techniques of
sample preparation, h a v e m a d e it possible to penetrate into the previously imagined, but
still u n k n o w n , interior of the soil structure. T h e scanning m i c r o s c o p e is well adapted for
studies of particle shape and particle arrangement (Barden and Sides, 1971), while the
transmission m i c r o s c o p e has been a valuable tool in the study of micro-structure of clay
and fine-grained organic soils (Pusch, 1967). In the former case, the specimen is first
deep-frozen and the section surface to be studied is then covered with an ultrathin film
of gold. In the latter case, the pore water is replaced by acrylate plastic with the aid of a
catalyst. In a diffusion process the plastic replaces the pore water, and after the plastic has
hardened, the specimen is cut by means of a m i c r o t o m e in ultrathin slices, which can be
penetrated by the transmission microscope.
T h e micro-structural features of coarse-grained soils can also be studied on plastic-
e m b e d d e d samples. In this case optical microscopes are generally used.

(i) Clay minerals. T h e m o s t c o m m o n clay minerals are kaolinite, illite, m o n t m o r i l l o n i t e


(smectite) and chlorite. T h e y are characterised by a monoclinic crystalline system with
eminently basal cleavage. Montmorillonite has a crystal lattice that is similar to that of
Fundamentals

Fig. 1. Scanning microscope pictures of clay particles. Glacial quick clay (top) and lacustrine clay.
6 Fundamentals

illite but the attraction between the crystal mono-layers of montmorillonite is weaker and
therefore allows expansion of the interspace between the layers in connection with water
uptake. Montmorillonite clay therefore behaves differently from other clays in that it is
prone to swelling w h e n unloaded or exposed to water.
As a result of their crystalline structure, particles of clay and m i c a minerals have a m o r e
or less flat, flake-like form with very irregular contours. T h e edges of the particles are thin

Fig. 2. Ultrathin cuts of marine (top) and lacustrine clays photographed using a transmission electron
microscope and schematic pictures of their structure (after Pusch, 1970).
Fundamentals 7

and w e e k (Fig. 1). T h e micro-structure of clay is dependent on the environmental


conditions during its formation, especially the salt content of the water in which the clay
has been deposited and on the overburden pressure during its consolidation.
F r o m a geotechnical viewpoint, the most important feature of the mi crostructural build
of illite clays is the formation of aggregates linked together via bridges of small particles
in a three-dimensional network (Fig. 2). This microstructural pattern has a dominating
influence on the deformation and strength characteristics. In clays that h a v e been heavily
overloaded, for e x a m p l e , w e find groups of particles with an ordered orientation
("domains"), b e t w e e n aggregates with a r a n d o m orientation. T h e high consolidation
pressure has forced the aggregates close together and resulted in the bridges being
strongly deformed and sheared. This process resembles the failure d e v e l o p m e n t in shear.
In lacustrine clays, particles a b o v e 1 in size show a tendency towards horizontal
orientation irrespective of the overburden pressure, whereas marine clays show such a
tendency only w h e n the overburden pressure exceeds 200 kPa (Pusch, 1967).
In montmorillonite clays, the mi crostructural build is different from that of illite clays.
S o d i u m montmorillonite, deposited in a marine environment, is characterised by having
extremely thin, flake-like particles with a m o r e or less horizontal orientation, while

Fig. 3. Microphotograph of sand showing cementation in the contact surface between grains.
8 Fundamentals

calcium montmorillonite is characterised by having m o r e r a n d o m l y oriented groups of


clay particles parallel between themselves in the groups.
Pictures taken by m e a n s of the scanning microscope technique show that particles of
clay-size (< 2) often adhere to the surface of coarser grains, usually as a kind of coat
(Fig. 3). This m e a n s that even a small clay content m a y have a strong influence on the
creep properties of the soil. At a higher clay content, the clay minerals will be decisive
for the physical characteristics of the soil. Already at as low a clay content as 1 5 - 2 5
weight%, the clay fraction forms a continuous mass that separates coarser grains from
each other.

(ii) Crystalline, rock-forming materials. Crystals of feldspar and quartz, which are
crystalline, rock-forming materials, h a v e anisotropic strength characteristics and have
therefore a shape which is associated with that of a prism. Particles of feldspar and quartz
less than 2 in size (belonging to the clay fraction) exist in great number also in coarse-
grained soils (Fig. 4) but the effect of these particles on strength and deformation is not
similar to the effect exerted by the clay particles.

(iii) Organic material. Besides the organic substance visible to the naked eye, such as
fruits, leaves, roots, seeds, etc., w e find microscopic remainders of spores, pollen, micro-

Fig. 4. Clay-size quartz and feldspar particles in sand from a sand pit in Sweden.
Fundamentals 9

scopic animals, algae, germs and viruses as well as organic molecules and c o m p o u n d s .
Living as well as dead organisms exist.
T h e organic matter is b o u n d to the surface of the clay particles via hydrophilic groups
which, depending on the electrolytic composition, m a y give rise either to a protecting
coat or to the formation of a cementing gel complex. Generally speaking, the organic
substance and its bonds are a m o r p h o u s in nature, which implies a strong tendency
towards creep deformations on loading. This effect is amplified by the fact that the degree
of order of the water in the region of contact b e t w e e n the grains is low. Organic matter
contributes to the creation of a very large v o l u m e of voids in the soil. This is also the case
as to coarse-grained soils. Organic soils are therefore generally very compressible.

1.2 Macro-structure

In a geological formation every type of stratum forming an integral part of the soil mass
has its specific mechanical characteristics depending upon grain size, grain shape, grain
origin, v o l u m e of voids, etc. T h e revelation of those structural features that m a y have a
decisive influence on foundation design is one of the foremost aims of soil investigations.
T h e extent and choice of m e t h o d of soil investigation is thus very m u c h dependent on the
geological situation at the site.

(i) Structural anisotropy. In sedimentary soils, the strata are typically horizontal or nearly
horizontal. Coarse-grained layers are often e m b e d d e d in fine-grained soils and vice versa.
Undetected layers with different characteristics from those of the soil mass as a whole are
often the cause of unsuccessful foundation design and even of disastrous events. For

Fig. 5. Structural anisotropy of a glaciofluvial deposit.


10 Fundamentals

e x a m p l e , the landslide at Furre in the centre of N o r w a y took place along a thin quick clay
layer, e m b e d d e d in a deep deposit of mainly sand and silt and sloping at an angle of about
6 degrees towards the river N a m s e n (Hutchinson, 1961). This thin clay layer was not
discovered until the slide had taken place.
Impervious layers in coarse-grained soil m a y serve as a watertight lid and cause
problems in connection with deep excavations (for e x a m p l e hydraulic uplift).
Coarse-grained layers e m b e d d e d in clay deposits are of great i m p o r t a n c e for the
overall permeability in the horizontal direction. In hilly surroundings, continuous coarse-
grained layers e m b e d d e d in clay deposits are often fed with water under artesian pressure,
which leads to a condition of h y d r o d y n a m i c equilibrium.
So-called erratic strata, representing very irregular and unpredictable stratification,
are of particular concern in foundation engineering. Disturbed or distorted stratification
is quite c o m m o n . An e x a m p l e is given in Fig. 5.

(ii) Non-homogeneities. There is almost no soil deposit that can be considered fully
h o m o g e n e o u s . T h e deformation and strength characteristics are, for e x a m p l e , normally
different in the vertical and the horizontal directions, not only because of variations in soil
structure but also because of the effect of stress history. Of course, the stratification in
itself gives rise to n o n - h o m o g e n e o u s conditions.
Till is a typical e x a m p l e of n o n - h o m o g e n e o u s soil: a conglomeration of grains with a
very large variation in size and shape, most of which are sharp-edged. Coarser grains are
normally e m b e d d e d in a fine-grained matrix that governs mechanical behaviour. In
boulder clay (clay till), fissure systems form planes of weakness.
Clay deposits often contain vertical channels formed by roots or by gas evolution. This
is especially the case at shallow depth, for e x a m p l e in dry crust clay. Open root channels,
which for some species of clover can extend to depths of about 5 m, are recognised by
a change in colour and stiffness. Moreover, deep-going fissures normally extend through
the dry crust d o w n into the underlying soft clay.
In foundation engineering, local non-homogeneities are of particular danger. For
e x a m p l e , lenses of highly compressible soils e m b e d d e d in deposits, which as a whole can
be considered to b e very satisfactory under any proposed foundations, are often the cause
of detrimental, differential settlement of buildings.

1.3 Voids

(i) Content of matter. T h e voids in the soil are filled with either water or gas or both. T h e
groundwater always contains dissolved elements, both ionised or non-ionised, and salts
as well as suspensions of mineral particles, h u m u s gel or other organic matter, gases, etc.
T h e content of organic matter can be roughly determined by vaporisation of the ground-
water after which the remainders are oven-dried at 180C for one hour and then weighed.
T h e result is c o m p a r e d with the electrical conductivity of the groundwater.
Fundamentals 11

T h e substances usually dissolved in the water are different kinds of gases, such as
oxygen, nitrogen, h y d r o g e n sulphide, m e t h a n e and carbon dioxide; different elements,
such as silicate, iron, calcium, m a g n e s i u m , sodium and potassium; different kinds of
salts, such as carbonates and bicarbonates, sulphates, chlorides, nitrates, h u m t e s and
tannins. T h e concentration of dissolved matter can b e subjected to strong variations, from
about 0.01%o in rain water and snow to over 3 0 % in certain salt lakes. T h e total
concentration of suspended matter is very rarely above 0.5%c (0.005%o in respect of
mineral particles and, except for organic soils, 0.15%o in respect of organic matter). T h e
influence of dissolved and suspended matter on the unit weight of g r o u n d w a t e r m a y have
to be taken into consideration in geotechnical analysis.
T h e presence of gas in the soil voids is of great importance in geotechnical design. T h e
most c o m m o n gases are carbon dioxide above the groundwater level and m e t h a n e below.
T h e solubility of gases in the water is directly proportional to the water pressure and
inversely proportional to the temperature. T h e concentration of dissolved gases in the
groundwater varies generally from 0.001 to 0.1%o.

(ii) Volume of voids. In soil m e c h a n i c s , the v o l u m e of voids in the soil is either expressed
in terms of porosity or in terms of void ratio. T h e porosity is defined as the ratio of v o l u m e
of voids to total v o l u m e and is designated by the symbol , while the void ratio is defined
as the ratio of v o l u m e of voids to v o l u m e of solids and is designated by the symbol e.
Expressed in t h e terms given in Fig. 6, w e thus have:

n = VpIV (1)

e = Vp/Vs (2)

volume mass

ma
Va gas
density pa

V vw
water
density pw

,
- ' ^'"*^ VJ; ^:::.7.
,
ms
solids:,
density'^','" '." ' '''"",'

Fig. 6. A schematic picture of a soil element divided into its constituents: solids, water and gas.
12 Fundamentals

T h e correlation between and e is given by:


e = (3)
l-n

T h e porosity of coarse-grained soils is strongly dependent on the grain size distribution


and the shape of the grains. If the grains w e r e all equal in size and altogether spherical,
the range of variation would be from = 2 6 . 0 % in the densest state, to = 4 7 . 6 % in the
loosest state. As a result of fines filling the voids b e t w e e n coarser grains, the porosity can
be considerably lower than 2 6 % . In tills, for e x a m p l e , which contain most of the grain
fractions, the porosity can b e as low as about 10%.

1.4 Water content

T h e water content w is defined by the relation:

w - mwlms (4)

where mw is the mass of water and ms is the mass of solids.

(i) Natural water content. T h e natural water content of a specimen is determined by


weighing the specimen, first in its natural state and then again after having kept it in a
drying oven at 110 5C for 24 hours, or, alternatively, in a m i c r o w a v e oven (for details,
see Gilbert, 1991). T h e determination of the water content is part of the routine
investigation of fine-grained soils.

(ii) Degree of water saturation. T h e degree of water saturation Sr is obtained by the relation :

Sr=Vw/Vp (5)

w h e r e Vw is the v o l u m e of water and Vp is the v o l u m e of voids.


T h u s , for a soil w h o s e degree of water saturation is Sr, the natural water content is:

w -mwlms = eSrpwlps (6)

where pw and ps as defined below (p. 14).


Normally, both the water content and the degree of water saturation are expressed in
percentages.
Fundamentals 13

1.5 Density

(i) Specific density. The specific density (i.e. the density of solid material) is defined by
the relation:
V
Ps = s' s (7)

where ms = mass of the solid material,


Vs = volume of solid material.
The specific density of a soil material is more of academic than of practical interest (see
below grain density). Its numerical value, which depends on the molecular structure,
gives an indication of the mineral compounds. The ps value normally ranges from 2.65
3 3
to 2.70 t / m for coarse-grained soils, and from 2.70 to 2.80 t / m for fine-grained mineral
soils (clay).

(ii) Grain (particle) density. The grains themselves are seldom completely solid but have
a certain porosity and, therefore, from a practical viewpoint, it is better to make use of
grain density instead of specific density. The grain density is defined by the relation:

P g = mg/Vg (8)

where mg= mass of the grain,


Vg = volume of the grain.
The grain density is very nearly equal to the specific density. However, in the case of

TABLE 1.
Grain (particle) density of some different minerals

3
Mineral Grain density, t/m Comments

Amphibole 2.8-3.4 Rock-forming minerals, mainly constituting coarser


Biotite 2.7-3.1 grains. However, quartz and feldspar sometimes con-
Calcite 2.7 stitute more than 50% of clay fraction.
Quartz 2.65
Feldspar 2.5-2.9
Mica 2.8-2.9
Muscovite 2.8-3.0
Pyrite 5.0-5.1
Pyroxene 3.1-3.6
Illite 2.6-2.8 Clay minerals, mainly constituting clay fraction.
Kaolinite 2.6-2.7
Montmorillonite 2.4-2.8
Chlorite 2.6-3.0

3
Remarks: According to Jelinek (1966), the average value of pg can be assumed equal to 2.65 t/m for
3
sand and gravel and 2.75 t / m for clay and clayey silt
14 Fundamentals

TABLE 2.
Typical bulk density values

3
Soil type Density (t/m )

Water-saturated Above groundwater surface

Peat 1.0-1.1 Often water-saturated


Dy and gyttja 1.2-1.4
Clay and silt 1.4-2.0
Sand and gravel 2.0-2.3 1.6-2.0
Till 2.1-2.4 1.8-2.3
Rock fill 1.9-2.2 1.4-1.9

coarse grains of sandstone and limestone, for example, there may be a noteworthy
difference between the respective numerical values.
Typical grain densities are presented in Table 1.

(iii) Bulk density. The bulk density is defined by the relation:

= m/V (9)
where m = total mass,
V = total volume.
The bulk density can also be calculated on the basis of porosity and degree of water
saturation Sr according to the relation:

p = (l -n)pg + Srnpw (10)

where pw = density of pore water,


pg - density of grains,
or from the void ratio e according to the relation:

P = - ^ - ()
1+ e
Typical bulk density values are presented in Table 2.

(iv) Dry density. The dry density pd is the mass of solid matter per unit volume, i.e. the
mass that the soil would have per unit volume if the water in its voids were removed
without volume change taking place. Its numerical value can be obtained by either of the
relations:

Pd=~f = - = P*(l-n) - (12)


V 1+ w 1+ e
Fundamentals 15

As can b e seen, the dry density is directly correlated to the porosity (void ratio) of the
soil. It is therefore c o m m o n l y used as a m e a s u r e of the result achieved by compaction of
soil (p. 400).

Example 1: Determine the dry density, the degree of water saturation and the water content of a soil
3
with bulk density = 1.7 t/m and void ratio e = 0.8.

3
Solution: Assuming that the grain density pg = 2.65 t/m w e find:
3
pd = p j / ( l + e) = 2.65/1.8 = 1.47 t/m ,
1.7 = = (pg + eSr)/(l +e) = (2.65 + 0.8S r)/1.8,
whence Sr = 0.51,
w = mwlms = eSrpw/pg = 0.8-0.51-1.0/2.65 .= 0.15 (15%).

2. STRUCTURAL FEATURES OF ROCK

It is often quite difficult to distinguish b e t w e e n soil and rock. T h u s the rock surface m a y
be subjected to severe weathering, converting rock to soil and m a k i n g the border line
between r o c k and soil indistinct. Since excavation costs for soil is generally m u c h lower
than for rock, this uncertainty m a y entail controversies between contractors and clients.
It is obviously important to h a v e a clear definition of what is m e a n t with rock in civil
engineering. T h e definition accepted in S w e d e n can serve as an e x a m p l e . Accordingly,
rock is defined as that part of the earth 's crust which is characterised by high hardness and
low porosity and which cannot normally be dislodged by excavation.

2.1 Micro-structure

T h e micro-structural features of rock concern the matrix of individual crystals and their
atomic arrangement. Investigations h a v e indicated that the strength and deformation
properties of the r o c k material are governed by micro-structural anisotropy. T h u s , micro-
structural features m a y be important for the shear strength of small rock samples tested
on a laboratory scale, but they are not significant of strength and deformation properties
of a rock m a s s . T h e s e are governed by macro-structural features, such as planes of
weakness, w e a k zones, joints, etc.

2.2 Macro-structure

T h e fabric of the three m a i n rock categories has distinctive, differentiating features. T h u s ,


igneous and other crystalline rocks (such as granite, gabbro and basalt) are characterised
by the patterns produced by non-uniform arrangement of grains, crystals and groundmass;
metamorphic rocks (such as gneiss, schist and slate) by the patterns produced by
schlieren, foliation and lineation; sedimentary rocks (such as sandstone, m u d s t o n e , shale
and limestone) by stratification.
16 Fundamentals

T h e genetic origin of rock is of great i m p o r t a n c e for the strength and deformation


properties of intact rock. T h e overall strength of u n w e a t h e r e d and t e c h n i c a l l y unaffected
or only slightly affected rock is generally governed by joints or latent, invisible cleavage
planes. In m a s s i v e rocks, such as granite and certain types of limestone, the orientation
of the joints is discordant and quite irregular. Typically, the joints intersect each other at
right angles. Layered sedimentary rock is characterised by t w o m o r e or less parallel joint
systems perpendicular to each other and reminding of a brickwork.
Other types of structural discontinuities are shear zones, crushed zones, apertures,
bedding planes and solution cavities. T h e s e discontinuities are of main concern in
foundation engineering. F r o m the point of view of foundation engineering, the mechanical
characteristics of rock to b e considered are not those of the intact rock but instead those
of w e a k zones and other discontinuities.
In limestone formations, so-called karst developments often take place. Karst denotes
cave and gorge formations due to solution in carbonate rocks. Large cave systems are
k n o w n in, for e x a m p l e , South-East Asia and South Africa. Building activities which
influence the groundwater situation m a y lead to new karst formations with serious
consequential d a m a g e . Severe calamities h a v e occurred w h e n karst roofs h a v e collapsed
in housing areas.
In connection with shear zones and crushed zones there is a great probability that the
rock is decomposed or disintegrated to soil. Argillaceous zones in seemingly h o m o g e n e o u s
bed-rock, containing the clay mineral smectite, which b e c o m e e x p o s e d to air due to
blasting for the foundation w o r k or for tunnelling, m a y b e subjected to swelling and long-
term transformation to soft clay by suction of water (Fig. 7). A possible existence of such
zones has to b e k n o w n in order to avoid unexpected p r o b l e m s after the building
construction period.
H u m a n activities in rock also play an important role. For e x a m p l e , in places where
mining operations take place, problems m a y occur both due to unsatisfactory stability of

Fig. 7. Clay-altered zone has become fluid after being exposed to air.
Fundamentals 17

Fig. 8. Collapse of the rock cover of an old iron mine.

the rock cover (Fig. 8) and to long-term subsidence of the ground due to creep in the rock
material.

3. INFLUENCE OF GROUNDWATER

3 . 1 Natural groundwater condition

Groundwater conditions and water saturation play an extraordinarily important role in


soil mechanics. T h e r e are three groundwater zones that can b e discerned a b o v e the free
groundwater surface, n a m e l y the vadose, seepage and capillary zones.

(i) The vadose and seepage zones. Nearest to the ground surface w e find t w o z o n e s t h e
vadose z o n e and the seepage z o n e i n which the humidity varies b e t w e e n nearly
complete dryness and almost full capillary saturation.
In the top z o n e t h e v a d o s e z o n e t h e soil humidity is subjected to large variations
depending on external conditions. It is affected by the infiltration of rainwater, evaporation
from vegetation, evaporation during dry periods, formation of dew at night, frost activity,
and so on. T h e water is mainly concentrated at points of contact b e t w e e n grains.
In the seepage zone, the humidity normally increases with depth until it reaches almost
full saturation in the i m m e d i a t e vicinity of the capillary z o n e below. H e r e the groundwater
is in steady motion under the influence of gravitational forces. Pervious layers in low-
p e r m e a b l e soil can b e filled with flowing water, whereas the water saturation in the soil
m a y otherwise b e low. Local bodies of water can also b e found.

(ii) The capillary zone. In the capillary zone, nearest a b o v e the g r o u n d w a t e r surface, the
water is held b y capillary forces. T h e surface of the capillary z o n e is irregular due to
variations in the diameter and shape of the voids, and its level varies with changes in the
18 Fundamentals

groundwater level. T h e thickness of the zone is dependent on the capillary rise of the soil
in question (grain size and density).
T h e capillary rise b e c o m e s smaller in the case w h e r e the g r o u n d w a t e r level rises than
in the case w h e r e it falls. T h e reason is that the upper part of the capillary zone contains
air-filled pockets which set b o u n d s for the m o v e m e n t of water. In the lower part of the
capillary zone, however, the voids are completely filled with water. Here the water m o v e -
ment is governed by the same physical laws as b e l o w groundwater level. T h e height of
the capillary z o n e in the case of a sinking groundwater surface is given in Table 3.
T h e water in the capillary zone, as well as in the unsaturated zones above, is bound
mainly by surface tension in the interfaces between water, soil particles and gas. Other
types of binding forces, such as sorption forces, and chemical bindings also contribute.

TABLE 3.
Approximate capillary rise, in m (sinking groundwater level)

Soil type Grain size range, mm Loose state Dense state

Coarse sand 2-0.6 0.03-0.12 0.04-0.15


Medium sand 0.6-0.2 0.12-0.50 0.35-1.10
Fine sand 0.2-0.06 0.30-2.0 0.40-3.5
Silt 0.06-0.002 1.5-10 2.5-12
Clay < 0.002 > 10

(iii) The groundwater level T h e groundwater level (groundwater surface) is defined as


the level w h e r e the hydrostatic pressure equals the atmospheric pressure (the lower
b o u n d of the capillary zone). T h e physical m e a n i n g of the given definition is illustrated
in Fig. 9.
Naturally, the groundwater level is influenced by the natural supply of water to the
ground. Since there is a great seasonal variation in the supply of rain-water to the ground
and loss of groundwater due to transpiration and evaporation, the g r o u n d w a t e r level can
be expected to vary with climatic changes during the years. T h e r e are also minor
influences on the g r o u n d w a t e r level resulting from changes in atmospheric pressure,
variations in gravity (tides, earthquakes), etc. which are not often taken into account.
The climatological variations, of course, have a dominating influence on the groundwater
situation. A n illustrative e x a m p l e of the seasonal variations that m a y take place in a clay
deposit is s h o w n in Fig. 10. As can b e seen from Fig. 10, the hydraulic heads observed
at different levels in the aquifer below the clay, and in the clay layer itself, are almost
completely synchronised. W e also notice that there is a time lag b e t w e e n rainfall and its
effect on the g r o u n d w a t e r level.
C o m p l e t e saturation is probably rare in natural soils. A certain a m o u n t of gas s e e m s
always to b e present, although its quantity from a practical point of view is negligible.
T h e ion content of p o r e water is of great i m p o r t a n c e for the m e c h a n i c a l b e h a v i o u r of
Fundamentals 19

Fig. 9. Definition of groundwater level. In the capillary zone, the pore pressure differs from the
atmospheric pressure by the capillary suction hc - a. Below the groundwater level, the pore pressure
differs from the atmospheric pressure by the pressure exerted by a water column of height h.

-2 I I I
Hydraulic head in relation to ground level, m

Uta L J fil 1977


IT rJhrrr _ i
1978
JlilLf
Precipitation, m m

Fig. 10. Seasonal variation of pore water pressure at different depths in a clay deposit in Gothenburg,
Sweden. (After Berntson, 1983).
20 Fundamentals

clay materials. A reduction of the ion content by leakage of m a r i n e clays, or an e x c h a n g e


of ions through h u m u s additives, can lead to the creation of quick clays, i.e. clays which
in a r e m o u l d e d state b e c o m e liquid (see Fig. 16).

3.2 Total and effective stresses

T h e concept of total and effective stresses constitutes the basis of soil mechanics,
although it does not always seem to b e fully appreciated.
By definition, the normal and shear stresses acting on a section surface of a material
are a s s u m e d to b e evenly distributed over a solid area. However, a section surface of a
soil consists partly of intersected grains and partly of intersected voids filled with water
or gas, or a combination of both. T h e s u m of all internal stresses on a unit area of the
section surface (intersected grains as well as intersected voids) is k n o w n as total stress.
T h e latter can be considered as carried partly by direct contact be tw e en the grains and
partly by pore water or pore gas, or both combined. T h e part of the total normal stress a
that is carried in the points of contact between grains is called effective stress a ' a n d the
part carried by pore water (gas) is called pore water pressure uw (pore gas pressure ua ) .
In the case of water-saturated soils, the word pore water pressure is usually replaced by
pore pressure and the symbol uw is replaced by the symbol u.
T h e correlation between total stress and effective stress can be expressed by the
relation:
='+ u x+u (\-x)
w a (13)

w h e r e is a function of the degree of saturation (0 < <1). F r o m a practical viewpoint,


can be put equal to the degree of saturation Sr
For a water-saturated soil w e immediately realise that:

G=o'+uw (14)

Since water and gas cannot take up shear stress, total shear stress is equal to effective
shear stress.

Example 2: Determine the effective vertical stress 3 m below the ground surface.The capillary rise in
the soil is 2 m and the groundwater level is at 4 m depth. The soil in the capillary zone is assumed to
3
be water saturated. The density of non-saturated soil above the capillary zone is 1.8 t/m and of saturated
3
soil 2.1 t/m .

Solution: The total vertical stress at 3 m depth is equal to = 9.81(1.8-1 + 2.1-2) = 58.9 kPa and the pore
water pressure uw = - 10 kPa. Thus, the effective stress becomes equal to '\, = 58.9 + 10 = 68.9 kPa.
Fundamentals 21

3.3 Hydrostatic and hydrodynamic condition

In a hydrostatic condition, the pore water pressure at any depth is equal to the hydrostatic
pressure m e a s u r e d from the groundwater level. Moreover, the p o r e water pressure at one
and the s a m e level is the s a m e all over the place. T h u s , in both the vertical and the
horizontal directions w e h a v e a state of equal hydraulic head, i.e. the water is in a state
of static equilibrium and consequently n o water flow is taking place. This m a y be thought
of as the natural state in a virgin area but frequently is not. In reality, the pore water
pressure distribution quite often deviates from the hydrostatic o n e and has d o n e so for
ages (Fig. 11). In this case w e h a v e to deal with a h y d r o d y n a m i c condition; a condition
characterised by a steady state flow. T h e direction of flow d e p e n d s upon whether the
hydraulic head increases or decreases in the direction considered. T h e flow is governed
by the hydraulic gradient:

= // (15)

w h e r e Al - distance between the points of observation,


Ah = difference in hydraulic head at the points of observation.

(i) Darcy's law. In the middle of the 19th century, the French engineer Darcy discovered
that the rate of water flow in water-saturated sand followed the simple relationship
(DARCY, 1856):

GW
\ GW

clay

\ clay

m \ \ \
G
g \ / R

10 10 /
1 \

clay <
J3 *

15
V \
\
15
sand * ^ -~ -

" " - sand /


1
\
1
- %-:
\
\
. range of pore pressure variation,

~ Aug.-76 to May-82
20
pore pressure, Aug. 21 - 7 9

Fig. 11. Equilibrated pore pressure distribution with depth in two clay layers after groundwater lowering
in underlying sand, taking place around 4000 years ago (left) and 1000 years ago (Torstensson, 1975).
22 Fundamentals

= ki (16)

w h e r e k is termed the hydraulic conductivity or, m o r e commonly, the permeability of the


soil.
D a r c y ' s law is generally considered to be valid for laminar flow in all types of soils.
3
[Laminar flow can be expected when the hydraulic gradient / < 0.1 / J H , , where dw = grain
diameter in a soil having one single grain size and the s a m e total grain surface area as in
real condition (normally dl0<dw<d40, s e e p . 41)]. However, deviations from D a r c y ' s law
have been found to occur in clay soils at small hydraulic gradients. T h u s , in some
investigations, a threshhold gradient has been noticed below which no flow is taking
place. Considering the character of the internal forces in pore water (electric double layer
forces, sorption forces, etc.), it is logical to assume that the porosity of the clay would
seemingly increase, up to a certain point, with increasing h y d r o d y n a m i c forces. This fact
has in reality been observed in several experiments.
T h e rate of flow according to this concept could then be assumed to follow the relation-
ship (Hansbo, 1960; Dubin and Molin, 1986), see Fig. 12:

n
= k] i when / < ii
(17)
= k2 (i - z'o) when / > il

11
w h e r e kx and k2 are coefficients of permeability (k2- nkx if ' ),
n> 1,
i0= ii ( -\)ln.

(ii) Seepage pressure. Seepage of water gives rise to a h y d r o d y n a m i c force in the


direction of seepage equal to igpw per unit v o l u m e of the soil, w h e r e i is the hydraulic
gradient vector and pw is the density of water. T h u s , in the case of vertical seepage
(hydraulic gradient in the vertical direction equal to i ), the effective density of the soil
p' is either increased by ipw ( d o w n w a r d seepage) or decreased by ipw (upward seepage).
Obviously, an u p w a r d flow of water will lead to liquefaction w h e n ipw> p' (''> 0).

Example 3: The groundwater level in a sand deposit is at 0.5 m depth. Determine the effective vertical
3
pressure at 3 m depth in a soil with porosity = 30% and grain density pg = 2.65 t/m if the pore pressure
is increasing linearly with depth below groundwater level to 70 kPa at 5 m depth. The degree of water
saturation above groundwater is equal to 10%.

Solution: There are two alternative solutions, one based on seepage pressure and the other on the basic
relation between total and effective stress.
Using seepage pressure as a basis of analysis w e find:
i = (70/9.81 - 4 . 5 ) / 4 . 5 = 0.59
The bulk density above groundwater is:
Fundamentals 23

= k2(i - i0)

Hydraulic gradient i

Fig. 12. Observed deviation from Darcy's law at low hydraulic gradients in clay (Hansbo, 1960).

3
= (1 - 0.3)2.65 + 0.1-0.3-1.0 = 1.89 t/m
and the submerged density:
3
p ' = ( l - 0 . 3 ) - 2 . 6 5 = 1.16 t/m
The effective vertical pressure becomes:
<?v = 9.81[1.89-0.5 + (1.16 - 0.59)2.5] = 23.4 kPa
Using instead total vs.effective stress concept we find:
& v = 9.81(1.89-0.5 + 2.16-2.5) - 70-2.5/4.5 = 23.4 kPa

3.4 Pore pressure induced by loading

T h e stress increments induced in the soil due to loading can b e considered as total stress
increments. In soils with a low hydraulic conductivity the i m m e d i a t e effect of loading is
the creation of an excess pore pressure that will carry part of the load until the soil skeleton
has had time to rearrange itself and b e c o m e strong enough to carry the load on its own
(the effective stress increase is b e c o m i n g equal to the total stress increase). Several pore
pressure equations h a v e been suggested for the prediction of excess p o r e pressure in-
duced in the soil due to loading. A m o n g these, the ones k n o w n as S k e m p t o n ' s and
H e n k e l ' s are probably best k n o w n . T h e results obtained by the different equations are
very nearly the s a m e irrespective of the loading condition. A c c o r d i n g to S k e m p t o n ' s pore
pressure equation, w h i c h is less complicated than the others, the excess pore pressure Au
is obtained b y the relation:
Au = B [ 3 + ( - 3) ] (18)

where A and are the so-called S k e m p t o n pore pressure coefficients (Skempton, 1954),
and 3 are the major and m i n o r principle stress i n c r e m e n t s due to loading.
For water-saturated soil, = 1 while for non-saturated soils < 1. T h e A value depends
24 Fundamentals

30

4
depth 5.5r depth 4 m

"


depth 5 m

2 03
^ External load, kN/m

2
External load, kN/m

Fig. 13. Results of excess pore pressure observations during water-filling of a steel tank on water-
saturated glacial medium-sensitive clay (left) and organic postglacial low-sensitive clay (Sllfors,
1975). Dimensions of tank: diameter 9.6 m; height 5 m. Observations yield A values between the limits
0.14 below and 1.3 above the preconsolidation pressure for the glacial clay and 0.58 below and 1.07
above the preconsolidation pressure for the postglacial clay.

a m o n g other things on the preloading history of the soil. Generally speaking, A = 1/3 for
an elastic material, A > 1/3 for a contractive material, and A < 1/3 for a dilatant material.
T h e A parameter is not a constant but is subjected to a continuous c h a n g e during load
application. Its m a g n i t u d e below the preconsolidation pressure (the m a x i m u m past
pressure) w h e r e the soil skeleton, through earlier consolidation processes, is strong
e n o u g h to carry the load on its o w n is considerably lower than above.
E x a m p l e s of h o w A varies in soft high-plasticity clays in the course of loading are given
in Fig. 13.
Fundamentals 25

4. DEFORMATION CONCEPTS

4.1 Deformation moduli

T h e deformation properties of soils can b e expressed in terms of two basic moduli,


n a m e l y the b u l k m o d u l u s ^ ( r e f e r r i n g to c h a n g e of v o l u m e under isotropic stress increase
= da- ldev Fig. 14a, left) and the shear m o d u l u s G (referring to angular change
under p u r e s h e a r G = Fig. 14a, right). In practice, however, these basic moduli
are not usually applied in settlement analysis. D u e to the historical b a c k g r o u n d to
deformation analysis, the theory of elasticity is still widely used, w h i c h requires
k n o w l e d g e of the m o d u l u s of elasticity (Young's m o d u l u s ) and Poisson 's ratio v.
T h e correlations b e t w e e n the parameters and and the basic moduli are given by the
expressions:
E= (19)
1+G/3K

1-2G/3K
= (20)
; v
2 + 2G/3K
Equations (19) and (20) yield:

= 2G(l+v) (21)

For a m o r e or less incompressible material (in which case the value b e c o m e s very
large), for instance water-saturated clay in undrained condition (mineral particles and
water can be considered incompressible in comparison with the soil skeleton), w e find
= 3 G a n d = 0.5.
In the case of highly compressible soils, the deformation m o d u l u s usually applied in
settlement analysis is determined by m e a n s of compression tests in which lateral strains
are prevented, so-called o e d o m e t e r tests (Fig. 14b).
In classical soil m e c h a n i c s (Terzaghi, K., 1923, 1925), t w o moduli derived from the
results of o e d o m e t e r tests w e r e introduced, namely the coefficient of theoretical
compressibility av = Ae/Aav\ where Ae is the decrease in void ratio due to a vertical stress
increase ' , and the coefficient of v o l u m e compressibility mv - av/(\+e0) = / ' ,
where e0is the initial void ratio and is the relative compression. As the c h a n g e in void
ratio of clay in the virgin state, i.e. for a load a b o v e the past m a x i m u m load on the clay,
was found to b e directly proportional to the logarithm of effective vertical stress increase,
the primary c o m p r e s s i o n index Cc was later introduced, defined as:

Cc = AelA{\o%av') (22)

These " m o d u l i " are all foreign to the traditional moduli used in neighbouring branches
26 Fundamentals

Fig. 14a. Stress/strain conditions for determination of bulk modulus (left) and shear modulus G.

Fig. 14b. Stress/strain conditions for determination of oedometer modulus M.

of science and will, therefore, in the following b e replaced by a m o d u l u s related to the


mechanics of materials : =' / = 1 lmv (which is identical with for a Poisson ratio
v = 0).
T h e correlation between M and the basic moduli is given by the expression:

4G
M=K(\+ ) (23)

or, alternatively: 2G(l-v)


M= (24)
l-2v

4.2. Strain dependence

T h e shear m o d u l u s of soil depends to a high degree on the shear strain level the lower
the strain, the higher the m o d u l u s of deformation. T h e strains occurring in static loading
Fundamentals 27

TABLE 4
Soil parameters and ( = number of loading cycles)

Type of soil

Clean dry sand -0.5 0.16


Clean water-saturated sand -0.21og(AO 0.16
Water-saturated cohesive soil l+0.251og(AO 1.3

are generally of another order of m a g n i t u d e than in d y n a m i c loading (Fig. 15). This strain
dependence has to b e taken into account in the analysis of foundation p r o b l e m s .
T h e relation b e t w e e n secant m o d u l u s G and shear strain y(G = /) can b e expressed
by the relation (Hardin and Drnevich, 1972):

G = (25)
1+7/*

5
w h e r e G 0 = shear m o d u l u s with reference to a shear strain of about 10~ ,

yh =^[i+aexp(-j8^)],
7r 7r
yr - reference value of shear strain corresponding to r m ax / G 0,
and are soil parameters.
T h e and values presented by Hardin and Drnevich are given in Table 4.
To determine the reference shear strain yr w e h a v e to k n o w the initial shear m o d u l u s
G 0 and the shear stress T m ax (Fig. 16).
T h e shear stress leading to failure is governed by the M o h r - C o u l o m b failure criterion,
If w e a s s u m e that, initially, the vertical stress in the soil is ' and the horizontal stress
' = 0', w e find (Fig. 16):

2 2
[(a + - 5 cfv) '] - ( - 5 cfv) (26)

Machine foundations

Traffic vibrations
Off-shore gravity structures

Strong earthquakes

'Static' loading tests

( 5 4 :
io- - " 10"
Shear strain y

Fig. 15. Range of shear strain amplitudes for different types of loading conditions. (After Andrason, 1979)
28 Fundamentals

1
^{P failure c o n d i t i o n / * ^

/in situ c o n d i t i o n ^ \
max \


' ( '

Fig. 16. The T m ax value to be used for determination of yrcorresponds to the shear stress at failure in the
horizontal shear plane.

4.3 Stress dependence

T h e preloading history of the soil has a great influence on the deformation properties. One
of the m o s t important parameters in soil mechanics, which has to be investigated in order
to m a k e possible a correct evaluation of the settlement in loading, is the m a x i m u m past
pressure or, as it is generally termed, the preconsolidation pressure. A correct evaluation
of the preconsolidation pressure is of p a r a m o u n t importance both with regard to the
prediction of total settlement and to settlement vs. time relationship. A m o s t widely
utilised m e t h o d of determining the preconsolidation pressure is based on the semi-
logarithmic p l o t v o i d ratio e (or strain ) vs. log. according to C a s a g r a n d e (Fig. 17).
However, this m e t h o d can give a preconsolidation pressure that does not exist in reality
(see Fig. 5 1 , p. 78). T h e value obtained also depends on the choice of logarithmic scale.
Therefore, the result should always be checked by using instead a linear plote (or )
vs. a'. The preconsolidation pressure in this case (Fig. 17) can be chosen according to
Sllfors (1975).

4.4 Time dependence.

As is well k n o w n from c o n t i n u u m m e c h a n i c s , deformations h a v e t w o causes: one related


to c h a n g e in v o l u m e and the other to c h a n g e in shape. In m o s t materials, particularly in
soil, these deformations are time-dependent. Therefore, it is imperative to pay attention
to time effects.

(i) Time-dependent volume change. T h e t i m e - d e p e n d e n c e of v o l u m e changes is very


important in the case of fine-grained soils with low permeability and is d u e to p o r e water
being gradually squeezed out of (or sucked u p into) the voids in the soil, so-called
hydrodynamic time-lag. This p h e n o m e n o n , termed primary consolidation, is continuing
as long as part of the stress increase i m p o s e d on the soil is carried by p o r e water p r e s s u r e .
M e a n w h i l e , creep is taking place in the soil skeleton.
Fundamentals 29

Effective pressure ' , in kPa (log. scale)

40 I 1 1 1 1 1 I I 1 I L_J
0 100 200 400 600 800 1000
Effective pressure \ in kPa (lin. scale)

Fig. 17. Oedometer curve presented in lin.-log. (full line) and lin.-lin. scales (broken line). Lin.-log.
curve used for determination of the preconsolidation pressure according to Casagrande. Lin.-lin. curve
used for detewrmination fo the preconsolidation pressure according to Sllfors.

T h e long-term changes in v o l u m e due to an effective stress increase in the soil can be


expressed through the following constitutive relation (Taylor and M e r c h a n t , 1940):

d_e = d * + de ) ( 2 7

dt d& dt

w h e r e de/dcf = c h a n g e in void ratio due to c h a n g e in effective stress,


dcf /dt = c h a n g e in effective stress with time ( h y d r o d y n a m i c time-lag),
de/dt = c h a n g e in void ratio with time that is independent of effective stress.
In this expression, of course the void ratio e can be replaced by the relative compression .
In classical soil m e c h a n i c s the v o l u m e c h a n g e caused by creep p h e n o m e n a is
considered to begin at the end of primary consolidation, and is termed secondary
consolidation (Fig. 18). Although physically incorrect, this artificial partition of the
consolidation process into primary and secondary periods of consolidation is most
practical. S e c o n d a r y consolidation can b e very important in organic soils.

(ii) Primary consolidation. T h e consolidation properties of the soil are mostly expressed
in terms of one-dimensional (vertical) strain. T h e p a r a m e t e r g o v e r n i n g the duration and
30 Fundamentals

rate of consolidation is termed the coefficient of consolidation. If, in the consolidation


process, pore water is squeezed out (sucked up) in the vertical direction w e speak of the
coefficient of consolidation in vertical pore water flow, cv. Correspondingly, we speak
of the coefficient of consolidation in horizontal pore water flow, ch. T h e consolidation
coefficients are related to the oedometer modulus M and the permeability of the soil
through the relation:
cv = kvMlyw
(28)
c =
h k M/r
h w

where kv mdkh are permeabilities (hydraulic conductivities) in the vertical and horizontal
directions, respectively,
y w = unit weight of water.
The primary consolidation process is generally defined according to Terzaghi (1923,
1925). T h e basic assumptions behind Terzaghi's theory are:
the soil is water-saturated,
pore water flow is taking part only in the vertical direction (one-dimensional
consolidation),
Darcy's law is valid,
the decrease in excess pore water pressure equals the increase in ef fective stress (Acf = - Au),
there is a unique relationship between void ratio and effective stress.
On these assumptions w e can derive the consolidation equation:
2
du M d du du
V =
~V^V^ vTT c
(29)
L
dt Ywdz dz dz
where du/dt is the rate of excess pore water pressure dissipation,
is the depth coordinate (distance from the drained surface),
t is the consolidation time.
In consequence with Terzaghi's theory, the degree of primary consolidation U is
enunciated in the form of remaining excess pore pressure Au in relation to initial excess
pore pressure Au0:
U=l-Au/Au0 (30)

Alternatively, the degree of primary consolidation is expressed as the ratio of relative


strain , having occurred in reality, to relative strain ep to b e expected after full primary
consolidation:
=/
(31)

In reality, the permeability is a function of the strain, and consequently by the excess
pore water pressure, and therefore the consolidation equation is more correctly written:
Fundamentals 31

Void ratio e

Slope = Ca

Primary consolidation stage Secondary


consolidation stage
da' da'
0 =0
dt dt

logr

Fig. 18. Decrease in void ratio with time of loading of a clay specimen. Primary and secondary con-
solidation periods.

2
du du M dk du 2
2
- = c y L + _ _ ( ) (32)
dt dz ywdu dz

Alternatively, the course of consolidation can be expressed in terms of remaini ng primary


consolidation, er = - (Janbu, 1965):
2
cfe r 3 dev 9 2
+ (33)
dt dz 9 dz

T h e error introduced by ignoring the last term in this equation is less important than in
the prevous case, the reason being that the c v value does not usually vary as m u c h with
changes in as the k value with changes in u (in a").

(iii) Secondary consolidation. After, or rather at the end of, the primary consolidation
period, the relative compression (the change in void ratio) is directly proportional to the
logarithm of time elapsed. We have entered into the period named secondary consolidation.
T h e rate of void ratio decrease during secondary consolidation is usually expressed by
the relation (Fig. 18):
Ae = C a A ( l o g i ) (34)

where Ca = secondary compression index.


Instead of void ratio, secondary consolidation is often related to relative compression
by the relation:
A e = a vA ( l o g / ) (35)
32 Fundamentals

w h e r e cxs(l+e0) = Ca.

(iv) Creep due to shear. Shear deformations are the p r i m e cause of changes in shape. T h e
main part of these deformations take place instantaneously w h e n the load is applied.
However, even under working load condition the shear deformations caused by long-
term creep are mostly of the same order of m a g n i t u d e as those taking place immediately
in connection with load application. T h e rate of creep deformation Ascr/At in this case
is directly proportional to the increase in logarithm of time A(log t) and is a function of
the deviatoric stress level - 3 ; the higher the deviatoric stress, the higher the rate of
creep deformation. Eventually, creep m a y lead to failure (see par. 5.3).
T h e creep deformations under a given time, due to shear, are normally increasing
linearly with the applied load up to a certain limit, the so-called creep limit. W h e n the load
increases above the creep limit, creep deformations b e c o m e increasingly important.
Therefore, from the point of view of settlement, the creep limit is of great importance.

5. STRENGTH CONCEPTS

5.1 Strength parameters

(i) Effective and total strength. F r o m a physical viewpoint, the shear strength of soil is
built up of either of two c o m p o n e n t s or both: one that is only dependent on void ratio but
independent of the effective stress (cohesion) and the other that is dependent on the
effective stress (friction). In practice, the shear strength of soil is either expressed in terms
of effective stress by the so-called effective strength parameters c (the effective cohesion
intercept) and ' (the effective angle of internal friction,) or in terms of total stress by the
so-called total strength parameters c (the total cohesion intercept) and (the total angle
of internal friction). In the first case, the shear strength has been determined in a way that
gives us full k n o w l e d g e of the excess pore pressure developed during the test, whereas
in the second case the pore pressure is u n k n o w n . We also speak of the drained strength
parameters cd and in the sense that the shear strength has been determined in drained
condition with complete excess pore pressure dissipation.

' 3 ' \

Fig. 19. The Mohr-Coulomb failure criterion is represented by the envelope to the Mhr failure circles.
Fundamentals 33

c,

/
cr3 = ' 3 + 3 = + Au3

Fig. 20. Failure criterion for water-saturated clay in undrained condition.

T h e strength parameters of the soil constitute the basis for the analysis of the bearing
capacity of the ground. Failure is reached w h e n the shear stresses in the soil reach an
upper limit represented by the envelope to the M h r failure circles (Fig. 19). T h e Mohr-
C o u l o m b failure criterion is given by the relation:

Tf= c' + c/tan ' - (a + c/)tan ' (36)

w h e r e = cVtan ' stands for "attraction".

(ii) Undrained strength . In practice, the risk of failure is often g o v e r n e d by the shear
strength of the soil in undrained condition, i.e. w h e n the water content along the failure
surface remains unchanged. T h e undrained shear strength is particularly important in the
case of cohesive soils. T h e strength parameters in undrained condition are termed cu and
. In water-saturated cohesive soil, the effective stresses b e c o m e unaffected by changes
in total stress (Fig. 20) and h e n c e = 0. T h e failure condition is governed by the relation

(iii) 'True'strength. T h e so-called true strength parameters refer to the M o h r - C o u l o m b


failure criterion at constant void ratio. T h e true shear strength can b e expressed as the sum
of two c o m p o n e n t s : o n e I independent of , the other D d e p e n d e n t on ( S c h m e r t m a n n
and Osterberg, 1961). T h e independent c o m p o n e n t Ie can be expressed as ieo'c and the
dependent c o m p o n e n t De as ' w h e r e ie represents the strength of the microstructural
bridges and de the frictional resistance between the aggregates (p. 7; H a n s b o , 1975).

5.2 Quick clay and quicksand

T h e undrained shear strength of a cohesive specimen is generally m u c h smaller in


remoulded than in undisturbed condition. T h e soil is said to b e sensitive to disturbance
and the ratio of undrained shear strength in undisturbed condition to that in remoulded
condition is defined as sensitivity St. Ion leaching in m a r i n e clays or ion e x c h a n g e in clay
due to the effect of h u m u s acids m a y give rise to an instable clay skeleton, reminding of
a house of cards, so-called quick clays (Fig. 21).
34 Fundamentals

Fig. 21. Quick clay in undisturbed (left) and remoulded state. The disturbance of the clay structure due
to remoulding has turned it into a liquid. (By courtesy of AAB).

T h e condition of r e m o u l d e d quick clay reminds of another p h e n o m e n o n occurring in


loose water-saturated fine sand. Shear of loose sand will result in a decrease of its porosity.
When the rate of shear of the sand is great the effective stresses will tend to zero (the sand grains
will be floating in a surplus of water that has to escape in order to admit a decrease in the porosi ty )
and a state of liquefaction is obtained. A sand having these characteristics is named quicksand.
Quicksand phenomena can be expected in sand with rounded grains having a relative density
below 0.5 and a uniformity coefficient below 5. Shear caused by earthquakes may lead to
quicksand phenomena (liquefaction) in loose to medium dense sand (Seed and Idriss, 1971).

5.3 Creep failure

As pointed out in paragraph 4.4 , creep in cohesive soils induced by deviatoric stresses
of a lower m a g n i t u d e than those conventionally representing the failure condition may
end in failure (Fig. 22). T h e shear stress leading to long-term failure in undrained
condition is defined as creep strength ccr
T h e ratio of deviatoric stress level leading to long-term failure to the deviatoric stress
level leading to failure in standard testing procedures decreases with increasing content
of organic matter and increasing range of plasticity (cf. Fig. 70).

5.4 Strength anisotropy

Anisotropy of soil deposits is easily observed by the m a c r o s c o p i c features of the soil.


However, even seemingly h o m o g e n e o u s soils, such as glacial and postglacial clays, show
Fundamentals 35

Fig. 22. Observed axial rate of strain vs. logarithm of time in undrained triaxial test under two different
stress conditions. Bckebol clay with natural water content w=80% and consistency limits wL =75% and
wP = 30%. Undrained shear strength Tj= cu = 17 kPa. The creep strength ccr ~ 0.7cu.

structural anisotropy ( m a d e evident for e x a m p l e by the fact that the permeability in the
horizontal direction is generally higher than in the vertical direction). Moreover,
unisotropic strength properties derive from the fact that the overburden pressure in the
vertical and horizontal directions differ from each other. This entails that every soil
element in its natural state is subjected to a shear stress vector (Fig. 23) which adds to the
shear stress vector induced by the i m p o s e d load.

Fig. 23. Stress condition for soil element in situ.


36 Fundamentals

0.4
4
f
1>
< ) O
0.3 cS
I
<! )
O
} ;
)

& 0.2
<

0.1 '

O Sk- Edeby clay ( = 48%; cfCL&Q = 1.2 )


/
DraiTimen clay (/ p = 2 9 % ; :/ =15)

-90 -60 -30 0 30 60 90

Angle of inclination of failure plane

Fig. 24. Correlation between angle of inclination of failure surface and normalised undrained shear
strength cu/a'0 ( ' 0 = effective overburden pressure). After Bjerrum (1973).

A n e x a m p l e of strength anisotropy is given in Fig. 24. T h e influence of the inclination


of the failure surface can b e e x p r e s s e d by the relation ( H a n s b o , 1975):

2 2

CUC(/<JQ = (1 - K0)sinacosa+ 0.6[de(cos a + AT 0sin a) (37)


- ( 1 -K0) sin a COS a] + icfc/ (f0
w h e r e K0 is the earth p r e s s u r e coefficient at rest (see p p . 2 7 1 - 2 7 2 ) ,

Slow test
Quick test


j1

0.2 0.4 0.6 0.8 1.0 1.2


Plate diameter, m

Fig. 25. Volume dependence of undrained shear strength exemplified for Swedish boulder clay. Results
of plate loading tests. (Harden, 1974).
Fundamentals 37

= effective overburden pressure,


i and d according to p. 3 3 .
In the cases shown in Fig. 24, i-0.10-0.14 and d=0.29-0.20 w h e r e the former values
refer to D r a m m e n clay and the latter to S k - E d e b y clay.

5.5 Volume dependence

T h e influence on the shear strength of zones or planes of w e a k n e s s is of utmost


importance. T h e m o s t obvious proof of this fact is found in rock m e c h a n i c s w h e r e
hydraulic conditions and mechanical properties are totally governed by fissures and
crush zones. A test result obtained on a laboratory scale on small rock samples does not
give any information of use for the design of rock caverns or tunnels or foundation works.
T h e influence of the size of v o l u m e involved in a given p r o b l e m m u s t b e taken into
account in cases w h e r e the soil can b e considered n o n - h o m o g e n e o u s (Fig. 25).

5.6 The yield surface

F r o m what has been said in paragraph 4 . 2 - 3 w e realise that the soil structure starts
yielding w h e n it enters into a state of shear failure or w h e n the stresses exceed the
preconsolidation pressure G'C. In a case w h e r e the principal normal stresses are acting in
the vertical and horizontal directions, the yield surface will thus b e represented b y the
M o h r - C o u l o m b failure lines on the one hand (the 0'lines, Fig. 26) and the preconsolidation
pressure in the vertical and horizontal directions on the other. Inside the yield surface
border lines, the soil can b e a s s u m e d to b e h a v e m o r e or less as an elastic material.

2o'c

0.4t-

/
Fig. 26. Example of yield stresses and theoretical yield border lines. Medium plastic Drammen clay ( 0
= 30; K0 = 0.5). Inside the border lines, the clay can be assumed to behave elastically. Test results
according to Berre (1975). After Larsson (1977).
38 So/7 classification

SOIL C L A S S I F I C A T I O N

1. O B J E C T

Soil classification is a very important part of the identification process necessary for the
solution of problems involved in foundation engineering. For classification purposes
routine testing p r o g r a m m e s have been m a d e up, the extent of which depend upon the type
of soil to be classified. In foundation engineering classification with regard to geological
background, to particle size ranges and distribution, to plastic properties and organic
constituents and to strength and deformation properties is of principal interest.
Various classification systems exist depending upon tradition and regional geological
conditions. In this textbook, the main outlines of the classification system proposed by
ISO (the International Standardisation Organisation) and C E N (Comit Europen de
Normalisation) will be presented. Other classification systems of current interest until the
ISO and C E N standards have been finally adopted will be mentioned briefly.

2. G E O L O G I C A L B A C K G R O U N D

As regards the geological origin, we differ between sedentary and sedimentary m o d e s of


formation. Sedentary soils have been formed on the spot while sedimentary soils have
been transported in one way or the other to the spot. A m o n g sedentary formations w e find
residual soils (such as weathered rock and weathered gravel) and alluvium. Organic soil
formations created on the spot (such as peat) also belong to this group. A m o n g sedi-
mentary formations we find the water-deposited glaciofluvial and postglacial sediments
(such as clay, silt and sand, esker formations, and sludge) and the wind-deposited
sediments (such as dunes and loess). Till formations which belong to a special type of
glacial sediment are characterised by their wide grain size distribution and the irregular
shape of the grains. Often the grains are characterised by having sharp edges. In regions
subjected to glaciation, till formations usually form a hard cover to bed-rock. T h e hard-
ness depends upon the fact that the till has generally been subjected to a heavy overburden
pressure of inland ice. In consequence, till usually forms an excellent basis for the foun-
dation of structures.
It is often difficult to clearly distinguish the boundary line between rock and soil. From
contractual point of view, soil can be defined as the part of the ground that can be dis-
lodged by excavation while rock cannot.
T h e geological origin has a very great influence on the geomechanical properties of
So/7 classification 39

soil and represents, therefore, a very important part of soil identification and classification.
Profound k n o w l e d g e of the geological features at the building site is also of great value
for the evaluation of which investigations should b e carried out and to what extent. If the
geotechnician has a lack of such k n o w l e d g e he ought to w o r k in close co-operation with
geologists.

3. C L A S S I F I C A T I O N A C C O R D I N G T O C O M P O S I T I O N

3.1 Grain size and grain size distribution

Every type of mineral soil contains a mixture of grains (particles) of different sizes and
shapes. Grain size and grain size distribution form the fundamental basis for designating
mineral soils using soil fractions to distinguish the major classes of geotechnical behav-
iour. Grain size, therefore, serves as a basis for soil classification. Unfortunately, as yet
the fractional limits, which are applied w h e n classifying a soil, are not the same in
different countries. This must be paid attention to in cases of soil classifications which
are presented without reference to any specific classification system.

(i) Fractional limits of mineral soils. As regards grain size, mineral soils are divided into
the following main fractional groups: boulders and cobbles (very coarse fraction), gravel
and sand (coarse fraction) and silt and clay (fine fraction). T h e main fractional groups and
their subdivisions are presented in Table 5.
M o s t of the European countries follow the principles established by I S O / C E N , with

TABLE 5.
Fractional limits according to ISO/CEN. The fractional limits according to the Nordic classification
system, wherever they differ from the ISO/CEN limits, are given in parenthesis.

Main groups Grain size, mm Sub-groups Grain size, mm

Boulders >200
(> 600) (Large boulders > 2000)
Cobbles 200-60
(600-60) (Large cobbles 600-200)
(Small cobbles 200-60)
Gravel 60-2 Coarse gravel 60-20
Medium gravel 20-6
Fine gravel 6-2
Sand 2-0.06 Coarse sand 2-0.6
Medium sand 0.6-0.2
Fine sand 0.2-0.06
Silt 0.06-0.002 Coarse silt 0.06-0.02
Medium silt 0.02-0.006
Fine silt 0.006-0.002
Clay < 0.002 Fine clay < 0.0006
40 Soil classification

only a few exceptions. For example, the N o r d i c countries ( D e n m a r k , Finland, Norway,


and S w e d e n ) h a v e raised the limit boulders/cobbles to 6 0 0 m m . This limit has been
considered m o r e suitable from the point of view of ease of excavation, and it also agrees
better with the general conception a m o n g laymen of the size of a boulder. In UK, the limit
boulders/cobbles is 256 m m , while in the U S A , the limit gravel/sand is 3.2 m m (1/8 in.)
and the limit sand/silt is 0.074 m m . Moreover, according to the unified soil classification
system ( U S C S , see p. 52), the limit silt/clay is determined by the consistency properties
of the soil, and not by the grain size. T h e reason is that soils, largely consisting of quartz
particles less than 2 in size, m a y not possess clay characteristics, although they should
be termed clay according to a classification system based on grain size.
We also differ between boulder and cobble soils (content of boulders and cobbles > 40
w t % of total material), coarse-grained soils (content of fines, comprising clay and silt, <
15 w t % of material < 60 m m ) , composite soils (content of fines 1 5 - 4 0 w t % of material
< 60 m m ) , and fine-grained soils (content of fines > 4 0 w t % of material < 60 m m ) .

(ii) Grain-size distribution. T h e grain-size distribution of soil with grain size between 60
and 0.06 m m is determined using a series of sieves with different m e s h widths. T h e same
set of sieves is used in International standards (ISO/R 565) and in G e r m a n standards (DIN
4188), except that International standards do not include sieves with m e s h widths 0.2,
0.63, and 6.3 m m . F r o m a practical point of view, any set of sieves can b e used to give
a good picture of the grain-size distribution.
Guiding values for the division of coarse-grained and fine-grained mineral soils are
given in Table 6.

TABLE 6.
Guiding values for soil classification on a basis of the contents of various fractions.

Fraction Content of fraction Content of clay Name of soil


in wt% of material in wt% of material Modifier Main term
< 60 mm < 0.06 mm

Gravel 20-40 gravelly


>40 gravel
Sand 10-20 sandy
>20 sand
Fines 5-15 <20 somewhat silty
(silt+clay) >20 somewhat clayey
15-40 <20 silty
>20 clayey
>40 < 10 silt
10-20 clayey silt
20-40 silty clay
>40 clay
So/7 classification 41

T h e grain-size distribution of material with grain size < 0.06 m m must be determined
by sedimentation analysis, while the size of boulders and cobbles is determined in the
field by sieving through a grating or by direct measurement.
T h e grain size is indicated by the symbol d (Fig. 27). T h e size of those grains which
correspond to 6 0 % , 4 0 % , etc. on the grain-size distribution curve is denoted d60, d40, etc.
T h e inclination of the grain-size distribution curve, also called the grading curve, is
indicated by the so-called uniformity coefficient C0 = d6Q/dl0. According to the ISO
proposal, the soil is called poorly graded if Cv < 6 and well-graded if Cv > 6.
T h e uniformity coefficient is sometimes not representative of the grading. This is the
case w h e n o n e or m o r e intermediate fractions are strongly under-represented. Such soils
2
are termed gap-graded. They are characterised by a coefficient of curvature Cc=d^Q /dQ d
c o m m o n l y b e l o w 1 or above 3.

(iii) Designation of soils for engineering purposes. T h e classification of mineral soils is


simplified by n o m o g r a m s which vary according to the country in which they are drawn
up (see Wiegers, 1974). T h e n o m o g r a m drawn up by the Swedish Geotechnical Society
for general classification of mineral soils (Fig. 28) is simple to use and provides an
u n a m b i g u o u s identification of the soil.

(iv) Visual observations. In the case of coarser soil material a fairly good estimate can be
m a d e of the size of the grains based on the size of m o r e familiar objects. For e x a m p l e ,
cobbles are larger than h o c k e y balls and gravel larger than lead shots. Sand, at its lower
limit, contains particles that are hardly visible to the naked eye.
In the case of silt and clay where the particles are invisible to the naked eye certain
methods can be utilised to help in the identification. Coarse silt particles can be felt between

Fines Coarse fractions Boulder and cobble fractions


clay silt sand gravel I cobbles , boulders

with boulders

0,002 0,02 0,06 0,2 0,6 20 60 200 600 2000


0,006
Grain size d, mm 1.85 mm d*60
t 170 mm CU = 92
i / 3 0 = 20 mm C c= 1 . 3

Fig. 27. Examples of grain size distribution curves and how to determine the uniformity coefficient.
42 So/7 classification

Fig. 28. Example of nomogram used for classification of sedimentary soils (Karlsson & Hansbo, 1984).
In accordance with the exemplified fraction percentages the soil should be designated sandy, silty day.

Example 4: Use the nomogram in Fig. 28 for designation of soil whose content of clay is (a) 20% [(b)
4%] of fines (a) 70% [(b) 45%] and of sand (a) 10% [(b) 22%]. All figures are given in wt.% of material
<60 mm.

Solution: The designation of the soil in example (a) is silty clay and in (b) sandy, gravelly silt.
Soil classification 43

the fingers and give a rough feeling. In the wet state the particles stick together but can
easily b e m o v e d in relation to each other. If the hands are r u b b e d with wet coarse silt, the
material dries rapidly to a p o w d e r which can easily be brushed off. T h e colour is normally
grey.
Medium silt and fine silt in a dry state h a v e a floury character and feel smooth between
the fingers. In wet state the material is plastic and at high water content, sticky. T h e
material can easily b e w a s h e d off the hands. T h e colour of dry material is normally white-
grey.
Clay in a wet state is plastic and sticky. It cannot easily be w a s h e d off the h a n d s without
a brush. O n drying, the samples turn into hard l u m p s that cannot b e crushed between the
fingers. T h e colour is normally grey, b r o w n - g r e y or red-grey.

3.2 Lime content.

On a basis of the calcium carbonate content in w t % of material < 0.06 m m , fine-grained


soils with a lime content > 8 0 % can be classified as highly calcareous, those with a lime
content of 4 0 - 8 0 % as very calcareous (clayey or silty marl), those with a lime content
of20-40% as moderately calcareous (very marly or very calciferous clay or silt) and those with
a lime content of 5 - 2 0 % as slightly calcareous (marly or calciferous clay or silt).
According to the I S O proposal, the carbonate content is assessed on the basis of the
reaction of soil to dilute hydrochloric acid, HCl. T h e soil is t e r m e d non-calcareous if there
is no effervescence, slightly calcareous if there is a slight, but not sustained, effervescence
and highly calcareous if there is a strong and sustained effervescence.

3.3 Organic content.

Quite often mineral soils are mingled with h u m u s or contain layers of h u m u s . H u m u s is


a m o u l d e r e d organic matter consisting partly of m a c r o s c o p i c formations, such as leaves,
rootlets, seeds, and larger animal r e m a i n s , and partly of m i c r o s c o p i c formations, such
as remains of m i c r o s c o p i c animals and insects, spores, algae, bacteria and viruses and
organic m o l e c u l e s or c o m p o u n d s . B o t h living and dead organisms are present.
In the process of decomposition, large quantities of organic sulphides are often formed
which h a v e a similar effect on the properties of the soil as that caused by h u m u s .
As organic matter has quite a negative influence on the m e c h a n i c a l properties of soil,
its possible occurrence should always b e recognised. Therefore, classification b a s e d on
the content of organic matter is equally important as classification with regard to grain
size and grain-size distribution. A s in the case of grain size, different classification
systems exist throughout the world. O n e of these, r e c o m m e n d e d b y the S w e d i s h G e o -
technical Society (Karlsson and H a n s b o , 1984), is given in Table 7.
Colour, even in the case of small organic content, facilitates the distinction between
44 Soil classification

organic and mineral soils. T h e colour is dark and b e c o m e s black w h e n the organic content
exceeds about 5 % . Organic soils also have a distinctive odour.
Peat is a soil formed from vegetable remains, deposited in fens ( l o w m o o r peat) or in
raised b o g s (highmoor peat), with an organic content that exceeds 20 w t % of dry matter
(grain size < 2 m m ) , see Table 7. Depending upon the degree of decomposition, peat is
classified as fibrous peat, pseudo-fibrous peat and a m o r p h o u s peat, according to the von
Post classification system (von Post, 1921).
Fibrous peat has a low degree of decomposition, a fibrous structure and an easily
recognisable plant structure (primarily white m o s s ) . W h e n squeezing a sample of fibrous
peat in the p a l m of the hand, only water, and no solid matter, passes between the fingers.
Pseudo-fibrous peat has an intermediate degree of decomposition and a recognisable
plant structure. W h e n squeezed, less than half of the solid matter passes between the
fingers.
Amorphous peat has a high degree of decomposition, no visible plant structure, and a
mushy consistency. W h e n squeezed, more than half of the solid matter passes between
the fingers without free water being separated.
An exact determination of the content of organic matter is rather difficult to achieve.
T h e most c o m m o n methods are the ignition loss method and the colorimetric method. In
the first case, the soil sample should gradually be brought to 9 5 0 C (550C as recommended
in the A S T M Standards seems not always sufficient) and held until completely ashed (not
less than one hour), while in the second case, the organic matter is determined through
rapid wet combustion, as described by Walkley and Black (1934), followed by a
colorimetric test using a light filter for wavelengths of about 6 2 0 p m .
T h e ignition loss method is preferably used for a determination of the organic content
of highly organic soils, especially peat, provided that the lime content of the soil is not
too high (< 2 0 % ) . T h e lime content must be determined and the ignition loss corrected

TABLE 7
Terms for the designation of organic content

Term Organic content in wt% Examples


of dry material (< 2 mm)

Slightly organic 2-6 Gyttja-bearing clay soils


Dy-bearing silt
Humus-bearing clayey sand
Moderately organic 6-20 Clayey gyttja
Silty dy
Humus-rich sand
Highly organic >2() Gyttja
Dy
Peat
Humus-rich topsoil
So/7 classification 45

accordingly. Corrections should also b e m a d e with regard to the clay content. T h e


colorimetric method is applicable to soils with an organic content less than 6 0 % . It
normally gives m o r e accurate results ( 1 percentage point) than the ignition loss method.

4. CLASSIFICATION ACCORDING TO GEOTECHNICAL PROPERTIES

Classification systems based on the geotechnical properties of the soil are applied in order
to serve as a m e t h o d o l o g y in the appraisal of foundation problems. Normally, the soil is
classified with regard to relative density, strength characteristics, sensitivity, consistency
limits, consolidation characteristics and frost activity.

4.1 Relative density

T h e relative density is defined by the density index:


ej - e
ID =
e
(38)
L - eD

w h e r e eL is the void ratio in the loosest possible state,


eD is the void ratio in the densest possible state,
e is the existing void ratio.
T h e relative density can only be determined for coarse-grained soils with a low content
of fines (< 10%).
T h e eL value is determined by letting the soil sprinkle gently into a cylinder (generally
with a v o l u m e of 1 litre and an inner diameter of 102 m m ) through a sprinkling tube. T h e
sprinkling tube should b e held vertically with the orifice in a central position, m a x i m u m
20 m m a b o v e the filled-in soil surface.
T h e eD value can b e determined by compacting the s a m e quantity of the soil in
question, in a water-saturated state, until the smallest possible void ratio is achieved. T h e
compaction can b e carried out by vibration, or by s o m e other m e t h o d that does not infer
crushing of the grains.
T h e relative density of soil is a good indication of the strength and deformation

TABLE 8.
Classification of coarse-grained soils according to relative density and penetration resistance

Designation density index CPT SPT WST


? c( M P a ) halftums/0.2m
h
^30

Very loose <0.25 <2.5 <4 < 10


Loose 0.25-0.50 2.5-5.0 4-10 10-30
Medium dense 0.50-0.75 5.0-10.0 10-30 30-60
Dense 0.75-0.90 10.0-20.0 30-50 60-100
Very dense 0.90-1.0 >20 >50 > 100
46 So/7 classification

properties. Normally, it is estimated from the penetration resistance (pp. 6 1 - 6 4 ), for


e x a m p l e , as indicated in Table 8 (Bergdahl & Eriksson, 1983).
A m a n u a l examination in the field can also give an indication of the relative density
of coarse-grained soil. T h u s , loose soil can b e excavated with a spade, and a 50 m m
w o o d e n peg can b e easily driven. D e n s e soil requires a pick for excavation, and a 50 m m
w o o d e n p e g is hard to drive.

4.2 Strength properties.

As regards strength properties w e differentiate b e t w e e n non-cohesive soils, cohesive


soils and intermediary soils which do not belong to either of the t w o aforementioned
groups.

(i) Non-cohesive soils. In non-cohesive soils, the shear strength is derived partly from
friction between the grains and partly from dilation energy. T h e dilation energy which is
c o n s u m e d in the shearing of dense material has a great influence on shear resistance. In
m e d i u m dense material, the influence of dilation energy is negligible. In such a case, the
soil is said to be in a state of critical density. In loose soils w e h a v e to deal with negative
dilation.
Coarse-grained soils are typically non-cohesive. However, cementation between the
grains as well as capillary forces can give rise to apparent cohesion. A l t h o u g h the effect
of apparent cohesion m a y be very useful, it rarely can b e taken into account in the design
of foundations.

(ii) Cohesive soils. Cohesive soils are characterised by the fact that the shear strength is
due to both friction (between coarser grains and between aggregates that are formed by
clay particles) and cohesion within the material (through the action of sorption forces or
organic c o m p o n e n t s ) .
Clay and organic mineral soils are typically cohesive. Since it is not possible to
determine the relative density of these types of soil, w e c h o o s e instead to group them
according to shear strength properties, determined under undrained condition, the so-
called undrained shear strength cu (or Tfu). T h u s , the soil can b e said to b e very soft if cu
< 20 kPa, soft if 2 0 < cu < 4 0 kPa, firm if 4 0 < cu < 75 kPa, stiff if 75 < cu < 150 kPa and
very stiff if cu > 150 kPa.
Cohesive soils are also classified according to their sensitivity, i.e. the ratio between
the undrained shear strength of a specimen in undisturbed and in r e m o u l d e d states.
Sensitivity is very important for the estimation of h o w m u c h the undrained shear strength
may decrease in a case of disturbance, for instance d u e to piling. T h e sensitivity is m o s t
easily determined by m e a n s of the fall-cone test. Soils are termed slightly sensitive w h e n
the sensitivity St < 8, moderately sensitive w h e n 8 < St < 30 and highly sensitive w h e n
St > 30. Clays with St > 50 are called quick clays (cf. Fig. 2 1 , p. 34).
Soil classification 47

(iii) Intermediary soils. Silt and composite soils (content of fines 15-40 wt% of total
material < 60 m m ) occupy, from the point of view of strength, an intermediate position
between non-cohesive and cohesive soils. T h u s , the shear strength is due to both
frictional and cohesive resistance.

4.3 Consistency

(i) Consistency limits. T h e mechanical properties of r e m o u l d e d fine-grained soils is to a


great extent governed by the water content. If, for e x a m p l e , the water content of a clay,
originally in a liquid state, is gradually reduced it passes through a plastic into a firmer,
brittle state in which it easily crumbles. T h e water content limits, within which the soil
has a plastic consistency in the remoulded state, cannot be given exactly since the
transition from liquid/plastic consistency, on the one hand, into plastic/brittle, on the
other, takes place gradually. T h e consistency limits are therefore a matter of definition.
T h e m e t h o d s generally applied for their determination were originally suggested by
Atterberg, and in c o n s e q u e n c e these limits are alternatively called the Atterberg limits.
Atterberg p r o p o s e d the testing procedures for the determination of three limits: the liquid
limit w L , the plastic limit Wp, and the shrinkage limit ws. A full and detailed description
of the testing procedures applied for determination of the consistency limits, and their
historical back-ground, can be found in a d o c u m e n t presented by the Laboratory
C o m m i t t e e of the Swedish Geotechnical Society (Karlsson, 1977).
The liquid limit, which defines the transition from liquid to plastic state, is usually
determined according to C a s a g r a n d e ' s method, representing a final d e v e l o p m e n t of the
method originally suggested by Atterberg. T h e liquid limit is defined as the water content
at which the soil, in a r e m o u l d e d state, when placed in b o w l - s h a p e d c u p and parted into
two halves b y a V-shaped groove as shown in Fig. 29, flows together about 13 m m at the
b o t t o m of the cup after the cup has been dropped freely 25 times from a height of 10 m m
against a b o t t o m plate of micarta or ebonite.

Fig. 29. Determination of the percussion liquid limit. To the left, the groove has just been formed. To
the right, the groove has flown together 13 mm lengthwise.
48 So/7 classification

Lately, another m e t h o d based on the fall-cone test has b e c o m e increasingly popular


and seems to b e gradually replacing the C a s a g r a n d e method. In this case, a fall-cone is
placed so that the tip of the cone touches the levelled surface of the r e m o u l d e d sample
(Fig. 30). T h e c o n e is then released and the depth of penetration into the sample is
measured. In Sweden, where the fall-cone m e t h o d was invented and first applied in
practice, a cone with an apex angle of 60 and a weight of 60 g is utilised. T h e n the liquid
limit is defined as the water content at which the depth of penetration is 10 m m . This
corresponds to an undrained shear strength of about 1.8 kPa. (According to British
standards, a cone with an apex angle of 30 and a weight of 80 g should be used. Then
the liquid limit is defined as the water content at which the depth of penetration is 20 m m .
This yields very nearly the same result as the m e t h o d described above).
In order to m a k e a distinction between the t w o methods, the liquid limit values thus
determined are given different names: Casagrande's m e t h o d yields the percussion liquid
limit and the fall-cone m e t h o d yields the fall-cone liquid limit.
The plastic limit, which defines the transition from plastic to semi-solid state, is the
lower water content limit at which the sample can b e rolled to a thread, 3 m m in thickness,
without crumbling. T h e test is e a r n e d out by hand by rolling the sample in a plastic state
repeatedly on a water-absorbing paper until the given requirement is fulfilled.
The shrinkage limit represents the m a x i m u m water content at which the soil undergoes
no further shrinking w h e n its water content is reduced. It can also be defined as the
m a x i m u m water content at which the soil transforms from semi-solid to solid state.
Evidently, the soil in a remoulded state has a plastic consistency w h e n its water content
is between wL and wF T h e difference:

Fig. 30. Determination of the fall-cone liquid limit is generally performed by using a 60g/60 cone. To
the left, the cone is adjusted before being dropped. To the right, after the cone has been released and
penetrated into the soil.
Soil classification 49

TABLE 9
Classification of fine-grained soils on a basis of plasticity

Designation Liquid limit wL% Plasticity index IP%

Non-plastic < 1
Low plasticity <35 1-7
Intermediate plasticity 35-50 7-17
High plasticity 50-70 17-35
Very high plasticity 70-90 35-50
Extremely high plasticity >90 >50

IP = wL- wP (39)
is called the plasticity index,
and the ratio: W ~Wp W- Wp
(40)
WL Wp

the liquidity index,


and the ratio: Wj - w
(41)
k= = 1 - 4 w
the consistency index.
A plastic soil has a liquidity index 0 < IL< 1 and a corresponding consistency index 1>
/c>0.
T h e consistency index is sometimes used as an alternative basis for characterisation
of the strength properties of silts and clays. T h u s , according to I S O , soils are characterised
as very soft if Ic< 0.05, soft if 0.05 < Ic< 0.25, firm if 0.25 < Ic< 0 . 7 5 , stiff if 0.75 < Ic
< 1.0 and very stiff if Ic> 1.0.
O n the basis of the plasticity characteristics, soils are divided into four groups as shown
in Table 9.
A statistical study of the consistency limit values of m a n - m a d e h o m o g e n e o u s clays
(one and the s a m e for each test series) arrived at in different S w e d i s h geotechnical lab-
oratories (Figs. 3 1 - 3 2 ) s h o w e d a large coefficient of variation (Karlsson, 1977). T h e
most consistent values w e r e obtained with regard to the fall-cone liquid limit (Fig. 31).

(ii) Classification on the basis of plasticity chart. A s pointed out in Section 3 . 1 , the
division of fine-grained soils is often m a d e internationally according to plasticity
properties on a basis of the relation b e t w e e n the liquid limit and the plasticity index.
According to C a s a g r a n d e (1947), the mineral and organic soils fall on either side of the
so-called A Une in the plasticity chart IP vs. wL (Fig. 33).The A line follows the relation
Ip - 0 . 7 3 ( w L - 20), w h e r e wL (the percussion liquid limit) and IP in %.
50 So/7 classification

Low-plasticity clay High-plasticity clay

10r
Average value: 54%
(a) (b) Standard deviation: 3.0%

_ .is
>

MM J3

15i Average value: 54%


(c) (d) Standard deviation: 1.2%

a
10

20 25 30 35 45 50 55 60 65

Fig. 31. Statistical variation of the consistency limit w 7 determined in different laboratories, (a) and (b)
represent the percussion liquid limit, (c) and (d) the fall-cone liquid limit.

Low-plasticity clay High-plasticity clay


1 1
Average value: 18% Average value: 28%
Standard deviation: 1.3% Standard deviation: 3.2%

10 15 20 25 20 25 30 35 40
(
Test results, wp

Fig. 32. Statistical variation of the consistency limit wP determined in different laboratories.
So/7 classification 51

Soils with the same geological origin seem to fall within narrow zones, closely parallel
to the A line. However, the A line function itself is not unique but changes with regard
to geological conditions. For Scandinavian fine-grained soils, for example, the upper
boundary of organic clay and silt seems to follow the relation IP = 0 . 9 5 ( w L - 26) while
that of inorganic clay follows the relation IP = 0 . 8 ( w L - 8), where wL = fall-cone liquid
limit in %.
The plasticity index divided by the clay content, the so-called activity number, is a
measure of the colloidal activity of clay. The activity number is first of all dependent on
the ion exchange capacity and the specific surface of the clay minerals and on the content
of organic colloids.

4.4 The unified soil classification system

In 1952, the US Corps of Engineers and Bureau of Reclamation agreed on a soil desig-
nation system, the so-called Unified Soil Classification System (USCS), which has
become the most well-known and most widely adopted soil designation system throughout
the world. According to this system, the soil should be identified by a group name and
a group symbol, based on certain criteria using results of laboratory tests. According to
U S C S , soils are designated coarse-grained if more than 5 0 % by dry weight of material
with grain size < 75 m m is retained on No 200 sieve (mesh width 0.074 mm) and fine-
grained if the corresponding figure is less than 5 0 % . Thus, just as in other classification
systems, coarse-grained soils represent sand and gravel while fine-grained soils represent clay
and silt. A more detailed picture is given in Table 10.

Fig. 33. The Casagrande plasticity chart. Soils of equal composition lie in zones more or less parallel
with the A line. The soil classification symbols are those of the USCS (cf. Table 10) determined in
different laboratories.
52 Soil classification

TABLE 10.
USCS. Criteria for assigning group symbols and group names.

Symbol Name

Gravels, Clean gravels, C M> 4 a n d 1 < C C < 3 GW Well-graded gravel


more than 50% < 5% fines Cu < 4 and/or 1 > Cc > 3 GP Poorly graded gravel
> 4.76 mm Gravels with Fines classify as ML or MH GM Silty gravel
> 12% fines Fines classify as CL or CH GC Clayey gravel

Sands, Clean sands, C w> 6 a n d 1 < C C < 3 SW Well-graded sand


50% or more < 5% fines Cu < 6 and/or 1 > Cc > 3 SP Poorly graded sand
< 4.76 mm Sands with Fines classify as ML or MH SM Silty sand
>12% fines Fines classify as CL or CH SC Clayey sand

Silts and clays, Inorganic I > 1 and on or above "A" line CL Lean clay
wL < 50% Ip < 4 and below "A" line ML Silt
Organic ^P-oven dried^P-not dried <0.75 OL Organic clay/Organic silt
Silts and clays, Inorganic Ip on or above "A" line CH Fat clay
wL > 50% Ip below "A" line MH Elastic silt
Organic ^P-oven dried^P-not dried <0-75 OH Organic clay/Organic silt

Highly organic soils Primarily organic matter,dark PT Peat


in colour,and organic odour

The symbol O L means organic clay when IP > 4 and on or above the " A " line in the
/ p / w L - d i a g r a m and organic silt when IP< 4 and below the " A " line, while the symbol OH
means organic clay when IP is on or above the " A " line and organic silt when IP is below
the " A " line.
A full description of U S C S can be found in, for example, the 1986 Annual of A S T M
Standards under Designation: D 2487.

4.5 Frost activity

Frost penetration into the soil leads to formation of ice crystals which create a pore water
underpressure in the adsorption layers surrounding the particles. As a result of the
underpressure, pore water is sucked up from the environs and concentrated as ice lenses
and/or ice layers in the soil. The inflow of water continues as long as water can penetrate
into the adsorption layers between the particles and the growing ice crystals. The ice
crystals continue to grow as long as the inflow of water can keep pace with the ice
formation. The thickness of the water film between the grains and the ice crystals in which
the water is moving is independent of the grain size. Therefore, the total flow area is much
smaller in frozen coarse-grained material than in frozen fine-grained material, which
reduces the possibility of water flow into the frozen zone. Consequently, coarse-grained
soils are not subjected to frost heave due to suction of water from the environs and
So/7 classification 53

formation of ice lenses or ice layers. Fine-grained soils, on the other hand, and silt in
particular, are very disposed to frost heave.
T h e frost h e a v e to b e expected obviously depends on the permeability of the soil
beneath the frozen z o n e and on the grain size in the frozen z o n e as well as on the rate of
penetration of the frost.
Classification of soils with regard to frost activity can be m a d e as follows:
G r o u p I - Frost-insusceptible soils, characterised by insignificant frost h e a v e and ice
formation, and b y insignificant softening during thawing. R o u g h l y speaking, this group
covers coarse-grained mineral soils and organic soils.
Group - Moderately frost-susceptible soils, characterised by m o d e r a t e frost heave
in normal conditions, and b y m o r e or less softening of the soil during thawing. R o u g h l y
speaking, this group covers the mixed-grain soils and clays, except silty clays.
G r o u p III - Strongly frost-susceptible soils, characterised by normally significant
frostheave and by severe softening during thawing. R o u g h l y speaking, this group covers
silt soils and silty clays.
A diagram for the determination of frost-susceptibility based on grain size distribution
has been presented in a w o r k report 1 9 8 5 - 1 9 8 9 by the I S S F M E Tecnical C o m m i t t e e on
frost (Fig. 34).

0.002 0.006 0.02 0.06 0.2 0.6

Grain size d, mm

Fig. 34. Determination of frost-susceptibility of a soil on the basis of the grain size curve. Region 1L
corresponds to group II (moderately frost-susceptible soils) and region 1 to group III (strongly frost-
susceptible soils). If the grain size curve falls inside either of regions 2-4 the soil belongs to group II
(frost-insusceptible soils). However, if the lower part of the grain size curve permanently passes the the
next region on the finer side, the soil belongs to group II-III. (ISSMFE Technical Committee on Frost).
54 So/7 classification

5. D I S C O N T I N U I T I E S A N D B E D D I N G

5.1 Discontinuities

Discontinuities m a y significantly affect the engineering behaviour of the soil m a s s . There


are two main groups of discontinui ties: natural discontinuities whichwereformeddur'mg
the deposition of the soil, such as bedding planes, and mechanical discontinuities which
include physical breaks in the soil due to shrinkage and/or ice loading in the Pleistocene
or tectonic disturbance. Fissures, faults and shears are other e x a m p l e s of mechanical
discontinuities.
T h e frequency of discontinuity occurrance is expressed by noting their spacing. T h e
terms r e c o m m e n d e d by ISO are given in Table 11.

TABLE 11
Terms for the designation of discontinuity spacing.

Term Mean spacing, mm

Very widely > 2000


Widely 2000-600
Medium 600-200
Closely 200-60
Very closely 60-20
Extremely closely < 20

5.2 Bedding

Bedding is another factor of great importance for the engineering behaviour of soil. As
was the case with the Furre slide event, undetected fine-grained layers e m b e d d e d in
coarse-grained soil m a y entail unforeseen risks of failure. Coarse-grained layers in fine-
grained soil represent another e x a m p l e . T h e s e generally provide drainage paths which
govern the rate of settlement induced by loading.
T h e terms r e c o m m e n d e d by ISO for the designation of bedding are given in Table 12.

TABLE 12
Terms for designation of bedding thickness.

Term Mean thickness, mm

Very thickly bedded > 2000


Thickly bedded 2000-600
Medium bedded 600-200
Thinly bedded 200-20
Thickly laminated 60-6
Thinly laminated < 6
Rock classification 55

ROCK CLASSIFICATION

T h e most important aspects of rock classification and identification in foundation


engineering concerns rock types and the structure of the rock mass. As in the case of soil
classification, a rock classification system is being worked out by the International
Standardisation Organisation, ISO (Price, 1992). Here, a short s u m m a r y of the proposal
put forward to I S O will be presented.

1 ROCK IDENTIFICATION

T h e first step is to define whether the rock is igneous, m e t a m o r p h i c or sedimentary.


Igneous and m e t a m o r p h i c rocks are crystalline; crystal surfaces reflect light, some
crystals show geometric forms. Igneous rocks show no sharp layer boundaries and are
massive. M e t a m o r p h i c rocks show layering, are often b o u n d e d by w a v y surfaces and are
described as foliated. Sedimentary rocks are clastic, /. e. they contain fragments or particles
belonging to sedimentary soils or rock that is not formed in situ. T h e y are mostly composed
of mineral grains cemented together and show bedding planes marking the boundaries
between sediment layers.
An aid to rock identification for engineering purposes is given in Table 13.

TABLE 13
Rock identification for engineering purposes.

Grain size Bedded rocks Foliated Massive and crystalline


mm Sedimentary Metamorphic Igneous

Coarse Conglomerate Gneiss Granite Gabbro


>2 (rounded particles
in a finer matrix)
Breccia
(angular particles
in a finer matrix)

Medium Sandstone Schist Microgranite Dolerite


0.06 - 2

Fine Mudstone Slate Rhyolite Basalt


0.002 - 0.006 Shale

< 0.002 Flint, chert Mylonite Obsidian


56 Rock classification

Chalk, clastic limestone, crystalline limestone and dolomite, evaporites, coal and
lignite also belong to the group of sedimentary rocks. T h e s e h a v e a grain size in the whole
range shown in Table 13.
Fossils m a y b e found in sedimentary rocks. T h e mineral calcite in calcareous rocks
m a y b e scratched with knife and will react with dilute hydrochloric acid. Quarts scratches
steel. B r o k e n crystals in crystalline rocks reflect light.
T h e r o c k to b e identified is best seen in outcrop or as large fragments showing broken
surfaces.

2 ROCK MASS

2.1 Mass structure

T h e terms to b e used in standard geological practice are given in Table 14.

TABLE 14
Examples of terms used in the description of rock mass structure.

Sedimentary Metamorphic Igneous

Bedded Cleaved Massive


Interbedded Foliated Flowbanded
Laminated Schistose
Folded Banded
Massive Lineated
Graded

2.2 Weathering

T h e following s c h e m e of terms to describe weathering grades of rock material can be


utilised:
G r a d e I Fresh. N o visible sign of weathering of the rock material; perhaps slight
discolouration on major discontinuity surfaces.
G r a d e II Slightly weathered. Discolouration indicates weathering of rock material
and discontinuity surfaces.
G r a d e III Moderately weathered. Less than half of the rock is d e c o m p o s e d or
disintegrated to soil. Fresh or discoloured rock is present either as a discontinuous
framework or as core-stones.
G r a d e IV Highly weathered. M o r e than half of the rock is is d e c o m p o s e d or
disintegrated to soil. Fresh or discoloured rock is present either as a discontinuous
framework or as core-stones.
Grade V Completely weathered. All rock material is d e c o m p o s e d and/or
disintegrated to soil. T h e original mass structure is still largely intact.
Rock classification 57

G r a d e V I Residual soil. All rock material is converted to soil. T h e m a s s structure


and material fabric are destroyed. T h e r e is a large c h a n g e in v o l u m e , but the soil has not
been significantly transported.

2.3 Strength of rock material

With reference to the compressive strength of rock material, determined by m e a n s of


unconfined compression tests, rock is subdivided in a scale according to Table 15.

TABLE 15
Classification with regard to compressive strength of rock material.

Term Unconfined compressive strength, MPa

Very weak < 1.25


Weak 1.25-5
Moderately weak 5-12.5
Moderately strong 12.5-50
Strong 50-100
Very strong 100-200
Extremely strong >200

2.4 Discontinuities

(i) Discontinuity spacing. F r o m the point of view of foundation engineering, the


discontinuities in the rock mass govern its mechanical behaviour. As in the case of soil
there are two m a i n groups of discontinuities: genetic discontinuities (such as bedding and
foliation) and mechanical discontinuities (such as joints). B e d d i n g and foliation spacing
are termed according to Table 16 and mechanical discontinuity spacings according to
Table 17.

TABLE 16
Terms to describe bedding and foliation spacing.

Term Spacing, mm

Very thick bedded (foliated) > 2000


Thick bedded (foliated) 2000-600
Medium bedded (foliated) 600-200
Thin bedded (foliated) 200-60
Very thin bedded (foliated) 60-20
Thickly laminated (closely foliated) 20-6
Thinly laminated (very closely foliated) < 6
58 Rock classification

TABLE 17
Terms to describe mechanical discontinuity spacing,

Term Spacing, mm

Very widely spaced >2000


Widely spaced 2000-600
Medium spaced 600-200
Closely spaced 200-60
Very closely spaced 60-20
Extremely closely spaced <20

(ii) Aperture. T h e aperture, i.e. the p e r p e n d i c u l a r d i s t a n c e t h e g a p b e t w e e n walls


of discontinuities (Fig. 35), is t e r m e d according to Table 18.

TABLE 18
Terms for description of discontinuity aperture.

Aperture Aperture size term Feature description term

< 0.1 mm Very tight


0.1-0.25 mm Tight 'Closed' features
0.25-0.5 mm Partly open

0.5-2.5 mm Open
2 . 5 - 1 0 mm Moderately wide 'Gapped' features
> 10 mm Wide

10-100 mm Very wide


0.1-1 m Extremely wide 'Open' feature
> 1 m Cavernous

Fig. 35. Extremely wide aperture with infillings of sand.


Rock classification 59

(iii) Infilling. Identification of the infilling material in the apertures and determination of
its shear strength are important for the j u d g e m e n t of the stability of the rock mass.

(iv) Water seepage. Possible existence of free moisture or water flow is another important
factor to b e taken into account and has, therefore, to be reported.
60 So/7 investigations

SOIL I N V E S T I G A T I O N S

1. OBJECT

T h e object of soil investigation is to establish a reliable picture of the building site


conditions with regard to geological and geotechnical characteristics, necessary as a
basis for design. This purpose, which, of course, is likely to be in the m i n d of all
geotechnicians, is not always fulfilled. Not only do w e h a v e the inevitable pressure of
competitive tendering, which may result in a reduction of the level of site investigations
but also the p r o b l e m that the client, often inexperienced in the value of thorough and
professional site investigations, prefers good e c o n o m y to good engineering, ignorant of
the fact that insufficient information about the subsoil conditions m a y result in very poor
e c o n o m y in the long run. T h e client m a y also feel suspicious of the value of thorough and
expensive soil investigations since there are a great n u m b e r of geotechnicians w h o carry
out soil investigations, the results of which are m o r e or less irrelevant to the problems met
with in foundation design but serving merely as a display of their a c a d e m i c k n o w l e d g e
of soil mechanics.
W h a t kind of information, then, do w e need to be able to arrive at an o p t i m u m design?
First of all w e h a v e to form an opinion of the geological characteristics of the site, for
instance on the basis of geological maps. T h e s e m a p s , and the geological features
observed by inspection, give us a b a s i s for t h e p l a n n i n g of direct soil investigations. These
generally begin with some type of penetration testing, the results of which may give
valuable information about h o w the subsoil conditions vary at the site. M o r e advanced
methods of soil investigation can then b e determined with due regard to soil type,
variation in sounding results and type of project in question.
T h e m o s t vital information about the subsoil required for the design concerns the
strength and deformation properties and their variation at the site and the in situ stress
distribution. Although w e m a y do our best, on the basis of sounding results, to take
representative samples or to m a k e adequate in situ tests for determination of these prop-
erties, w e m u s t realise that, after all, it is merely a r a n d o m survey. In consequence,
statistical m e t h o d s h a v e been applied in order to account for the variations normally
occurring.
In the next paragraphs, the most c o m m o n m e t h o d s of investigation, and the relevance
of the results obtained for the solution of foundation engineering p r o b l e m s , will b e briefly
discussed.
So/7 investigations 61

2. P E N E T R A T I O N T E S T S

An extensive d e v e l o p m e n t and mechanisation of different sounding e q u i p m e n t s have


taken place in the last few decades. T h e most advanced sounding m e t h o d today is the
CPT-test (the c o n e penetration test), w h i c h has b e c o m e very popular in Europe. In the
U S A and in m a n y other parts of the world, the S P T (the standard penetration test) is still
extensively used due to long familiarity and experience. Other sounding m e t h o d s , such
as d y n a m i c probing (reminding of S P T ) and weight sounding, are also quite c o m m o n .
T h e a i m behind the innovations in sounding e q u i p m e n t has been to achieve a m a x i m u m
of information about the soil at the lowest possible cost. E r g o n o m i e considerations have
also played an important role.
All the penetration m e t h o d s mentioned a b o v e h a v e been standardised under the
auspices of the International Society for Soil M e c h a n i c s and Foundation Engineering
( I S S M F E ) . T h e r e c o m m e n d e d standard is publicised in Volume 3 of the Proceedings of
the International Conference on S M F E , held in Tokyo in 1977 (pp. 1 0 1 - 1 2 0 ) and in Vol-
u m e 4 of the Proceedings of the International Conference on S M F E , held in S t o c k h o l m
in 1981 (pp. 1 2 0 - 1 2 1 ) . National standards also exist.
In the following, a general description of various penetration m e t h o d s will be given.
For those w h o w a n t a m o r e detailed description, reference is given to the standards
presented in the Proceedings mentioned above.

2.1 Dynamic penetration tests

(i) SPT In the standard penetration test, a split-barrel sampler is driven from the b o t t o m
of a prebored h o l e into the soil by m e a n s of a 63.5 kg h a m m e r , d r o p p e d freely from a
height of 0.76 m. T h e sampler (Fig. 36) has a length of 457 m m , an outer diameter of 51
m m and an inner diameter of 35 m m . T h e diameter of the prebored hole varies normally
b e t w e e n 6 0 and 2 0 0 m m . If the h o l e does not stay open by itself, casing or drilling m u d
should b e used. T h e sampler is first driven to a depth of 0.15 m b e l o w the b o t t o m of the
prebored hole, then the n u m b e r of blows required to drive the sampler another 30 c m into
the soil, the so-called N30 count, is recorded. T h e rods used for driving the sampler should
h a v e sufficient stiffness. Normally, w h e n sampling is carried out to depths greater than
around 15 m, 5 4 m m rods are used.
T h e S P T has the advantage over other sounding m e t h o d s in that it provides samples,
certainly disturbed but still m a k i n g a classification of the soil possible and it suits
practically all types of soils. However, it m a y b e t i m e - c o n s u m i n g and e x p e n s i v e to
perform unless c h e a p labour is available.

(ii) Dynamic probing. D y n a m i c p r o b i n g includes several m e t h o d s and those w h i c h h a v e


been standardised b y I S S F M E are s u m m a r i s e d in Table 19. D y n a m i c probing, type Hf A,
standardised through the Swedish Geotechnical Society, is also presented in Table 19.
62 Soil investigations

511 mm

coupling

4 vents, 13 mm

diameter, min.

steel ball, 25 mm diam

split tube

driving shoe M

1.6 mm 351 mm

Fig. 36. The SPT sampler.

As can be seen from Table 19, the main difference between the various methods is the
driving energy. T h e r e c o m m e n d e d driving rate is 30 blows per m i n u t e but up to 60 blows
per minute can b e used in non-cohesive soils. In cohesive soils, however, the rate should
not exceed 30 blows per minute. T h e n u m b e r of blows required for every 0.2 m of
penetration is recorded in the site log.

TABLE 19
Data on various dynamic probing methods.

Method DPA DPB DPL

Point diam., mm 62 51 35.7 45


Point length, mm 62 51 71.4 90
Point angle, degr. 90 90 90 90
Rod diam., mm 40-45 32 22 32
Insertion preboring no preboring
Hammer, kg 63.5 63.5 10 63.5
Drop height, m 0.75 0.75 0.50 0.50
So/7 investigations 63

32 m m

320.3 m m

solid or
hollow

45 m m

510.2 m m

fixed or "lost"
I
\ /
90

5 mm
~f

Fig. 37. Dynamic probing equipment, types DPB (left) and HfA.

D y n a m i c s o u n d i n g w i t h o u t p r e b o r i n g can be carried out relatively quickly.


Consequently, it is m o r e cost-effective than the SPT. On the other hand, it does not include
any sampling to help in classifying the soil.
T h e points and rods to match d y n a m i c probing, types D P B and HfA, are s h o w n in Fig.
37. Rotational resistance provides additional, valuable information about the soil.

2.2 Static penetration tests

(i) WST. Weight sounding is a fairly old-fashioned m e t h o d but is still used, particularly
where there is too little space for other sounding m e t h o d s . In weight sounding, a screw-
shaped point (Fig. 38) fixed to sounding rods is pushed into the soil by m e a n s of weights,

350.2 m m

22

Fig. 38. Weight sounding point.


64 So/7 investigations

pushrod

Seal

friction sleeve

Fig. 39. The CPT point.

m o u n t e d on the top of the sounding rod by the aid of a c l a m p . T h e diameter of the


sounding rod is 22 m m and the weights comprise one 5 kg c l a m p , t w o 10 kg weights, and
three 25 kg weights in all 100 kg. T h e load should b e applied in steps and adjusted to
give a rate of penetration of about 50 m m per second and the m i n i m u m force required to
cause self-penetration is registered. W h e n the force has reached 1 kN (100 kg weight
applied) and penetration has stopped, the rod is rotated and the n u m b e r of half-turns of
the sounding rod required for every 0.2 m of penetration is registered.
Originally, weight sounding was always carried out by hand. N o w a d a y s , the sounding
is mostly mechanised and the load measured with a d y n a m o m e t e r attached to the ma-
chine. T h e results obtained in this case differ from the results obtained b y manual weight
sounding. A n important part of weight sounding, n o w a d a y s often neglected in its
mechanised form, was to register the sound produced by the screw-point while rotating
the rods. This sound gives a good indication of the soil type.

(ii) CPT. T h e cone penetration test, in its most modern design, offers excellent possibilities
to classify the subsoil from different aspects in spite of the fact that it does not include
sampling. Several types of c o n e penetrometers exist on the market. A c o m m o n feature
Soil investigations 65

is that the tip is cone-shaped with an apex angle of 6 0 (or, m o r e seldom, 90) and a diameter
of 35.7 m m (Fig. 39). It is usually provided with stress transducers which register
electrically the tip resistance and the friction against a friction sleeve. Hydraulic and
mechanical m e a s u r i n g devices exist. According to the standard, the penetrometer tip
should h a v e the s a m e diameter as the cone over a length of 1000 m m a b o v e the cone base.
Moreover, the friction sleeve should be placed immediately above the base of the cone
2
and h a v e a surface area of 150 c m . O n e has to remember, however, that there are
divergencies from the standard r e c o m m e n d e d , both with regard to the diameter of the
cone and to the location and surface area of the friction sleeve.
In its most m o d e r n design, the tip is also provided with a pore pressure transducer
which registers the pore water pressure during penetration the so-called piezocone. In
this case, the penetration test is usually referred to as the C P T U test. T h e p i e z o c o n e offers
excellent possibilities of identifying soil type and soil stratification. It is also a promising
method, based on local experience, of finding realistic values of strength parameters and
deformation moduli as well as consolidation characteristics (see, for e x a m p l e , Senneset
etal, 1989 and p. 103).
T h e cone tip is connected to a push-rod, with the s a m e diameter as the tip (or slightly
smaller) and is pushed into the soil by m e a n s of a thrust m a c h i n e at the r e c o m m e n d e d rate
of penetration of 20 m m per second (1.2 m per minute) T h e pushing force required for
penetration varies, of course, with the soil type, and a n u m b e r of thrusting rigs of various
powers exist.
Based on the C P T results, a report on soil type, relative density and undrained shear
strength in cohesive soils is often included. H o w well these parameters represent the truth
is as yet an open question.

3. GEOPHYSICAL METHODS

Geophysical investigation methods, for e x a m p l e seismic m e t h o d s , electrical resistivity


methods and echo sounding, are often utilised for subsoil survey. Seismic m e t h o d s are
of particular interest in foundation engineering. T h u s , by the use of seismic m e t h o d s , soil
properties can b e determined which are of interest for the solution of p r o b l e m s connected
with dynamically loaded foundations (see paragraph 5.3). In this paragraph, the presentation
of the m e t h o d s is restricted to subsoil survey.

3.1 Seismic methods

If the soil is subjected to a d y n a m i c shock, e.g. b y a detonation, three types of w a v e s will


be induced into the soil, n a m e l y compression waves (P waves), shear w a v e s (S waves)
and Rayleigh w a v e s (R waves). T h e soil particle m o v e m e n t s in the different m o d e s of
w a v e propagation are illustrated in Fig. 40. In conventional subsoil survey, it is generally
the wave propagation that is utilised.
66 Soil investigations

Source of vibration

S wave J R wave
Spherical propagation Circular-cylindrical propagation
along ground surface

Fig. 40. The different modes of soil particle movement induced by a shock at the ground surface. The
longitudinal and the transversial S waves spread spherically while the R wave moves in circles along
the ground surface.

T h e most c o m m o n method, used for the determination of the internal boundary


between an upper softer layer and an underlying harder layer, for instance between soil
and bedrock, is based upon the fact that the rate of w a v e propagation is a function of the
elastic properties of the soil, (cf. p p . 8 6 - 8 7 ) . T h u s , at the interface between two media
with different elastic properties, the propagating w a v e front is partly reflected and partly
refracted. By analogy with optics, the angle of refraction is obtained by the relation (Fig. 41):

(42)
s i n a /sin/3 = v lv
x 2
w h e r e = angle of incidence,
= angle of refraction,
v 1 = w a v e velocity in m e d i u m 1,
v 2 = w a v e velocity in m e d i u m 2.
If in a three layer m e d i u m 1-3 (Fig. 41) the w a v e velocities v 3 > v 2 > vh then for
certain limiting angles of incidence oc lr and o^r > the values in m e d i a 2 and 3 b e c o m e

Source of vibration Geophones

Medium 3

Fig. 41. Principles of the seismic refraction method in a three layer medium with v 3 > v 2 > v t . The fastest
wave propagation lines from the shock point to the respective geophones indicated.
So/7 investigations 67

equal to 90, w h e n c e s i n a l r = 2 and s i n o ^ = v 2 / v 3 . In this case, the refracted waves


will run parallel with the respective interfaces with velocities v 2 and v 3 . A s a result of
particle motions in the planes of refraction, new w a v e s with velocities vx and v 2 respec-
tively, and refraction angles equal to the limiting angles of incidence are created in the
upper m e d i a 1 and 2.
T h e m o d e of p r o c e d u r e is as follows: a shock is p r o d u c e d at the ground surface,
normally by firing an explosive in a shallow hole, and the time it takes for the reflected
and refracted w a v e s to arrive at a n u m b e r of sensors (so-called g e o p h o n e s ) , placed at
different distances from the explosive, is registered. T h e arrival times form the basis of
interpretation.
Evidently, according to the interrelation b e t w e e n velocity and m o d u l u s of elasticity,
the b o u n d a r y b e t w e e n a soft m e d i u m below a stiff m e d i u m (a m e d i u m with low m o d u l u s
of elasticity b e l o w a m e d i u m with high m o d u l u s of elasticity) cannot b e detected since,
in such a case, the angle of refraction will b e smaller than the angle of incidence.
Since the w a v e is a compression w a v e its velocity b e l o w water is m u c h higher than
above water. Therefore, the seismic m e t h o d can also b e used for location of the water
table.

3.2 Electric resistivity method

By the resistivity m e t h o d , a direct-current field is applied b e t w e e n electrodes installed


in the soil. T h e depth of penetration can b e regulated by varying the distance b e t w e e n the
electrodes, the potential difference b e t w e e n which gives an indication of the soil layer
sequence. However, in the case of water-saturated soils or soil b e l o w the g r o u n d w a t e r
level, the results cannot b e interpreted in a reliable way. O n the other hand, it is superior
to the seismic m e t h o d in the case of a soft layer underlying a stiff upper layer.

3.3 Echo sounding

E c h o sounding is often used for the determination of the soil layer s e q u e n c e beneath sea
or lake b o t t o m s . It provides not only direct information about the b o t t o m profile but also,
to a certain extent, the transition from one soil layer to another. It m a y also b e possible
to discern the b e d r o c k profile if it is not too deep below the soil surface. T h e depth of
penetration and resolving p o w e r depend upon the frequency applied. With m o d e r n
instruments, and under favourable conditions, changes in soil strata can b e observed to
depths of 2 0 - 3 0 m .

4. PORE PRESSURE MEASUREMENTS

Ever since the evolution of the effective stress concept, p o r e pressure prognostications
and m e a s u r e m e n t s h a v e been considered imperative. Nevertheless, soil investigations
and geotechnical reports h a v e contained very little information about groundwater
68 Soil investigations

conditions. T h e m e a s u r e m e n t s have generally been confined to the soil investigation


period, rarely lasting m o r e than a few w e e k s , and it has been taken for granted that the
pore pressure, in a steady state condition, can b e a s s u m e d to b e hydrostatically
distributed. This certainly can entail fatal misjudgements of stability conditions and
foundation p r o b l e m s on the one h a n d as well as exaggerated and u n e c o n o m i c a l solutions
on the other due to too pessimistic an approach to the effect of building activities on the
long-term groundwater conditions. Doubtless, reliable and long-term groundwater
observations, as well as possibilities of observing d y n a m i c pore pressure response, are
imperative in the analysis of the geotechnical, and foundation, and environmental
problems met with in practice.
An e x a m p l e of a m o d e r n piezometer is the BAT probe, s h o w n in Fig. 4 2 . It consists
of a flush d i a p h r a g m electric pressure transducer equipped with a h y p o d e r m i c needle.
W h e n the needle penetrates the rubber disc a direct connection is created between the
sensor and the filter. T h u s the sensor can be temporarily connected to the piezometer filter
tip w h e n taking the reading and then removed.
Other examples of m o d e r n piezometers are the Borro and N G I probes which measure
the pore water pressure by m e a n s of a vibrating wire. In the case of the B o n o probe, the
length of the wire, and consequently its frequency, is influenced by a spindle placed inside

Wire or electrical cable

_ One-in. pipe

. Test adaptor

_ Hypodermic needle

- Disc of resilient material

. Filter tip

Fig. 42. Piezometer probe of type BAT.


Soil investigations 69

a solenoid and attached to a bellows connected with the filter stone. T h e frequency of the
string is registered by m e a n s of sound w a v e s through the sounding rods and thus no
connecting cables are required. T h e results are stored digitally in a solid state memory.
In practice, in most occasions, satisfactory information about the pore pressure
situation is obtained by the use of open standpipes with filter tips. T h e m o r e advanced
piezometers are required w h e n it is important to study q u i c k pore pressure changes.

5. S A M P L I N G

Sampling is normally carried out on the basis of the results of penetration tests
(sounding). Sampling is necessary to ensure a m o r e accurate soil classification than that
which is possible on the basis of sounding alone. Sampling is also d o n e for the purpose
of laboratory investigations of the strength and deformation characteristics. This is
particularly important in the case of cohesive soils w h e r e consolidation tests on
undisturbed samples are the only m e a n s of investigating the characteristics required for
analysis of long-term settlement. In cohesionless soils, undisturbed sampling is not
possible. It is, therefore, questionable if laboratory testing of samples of cohesionless
soils will yield strength and deformation parameters of any practical use.
Details about different samplers and methods of sampling from all over the world, with
special emphasis on their application to undisturbed sampling of soft cohesive soils, are
given in an international m a n u a l , prepared by an I S S M F E s u b - c o m m i t t e e on soil
sampling and published by Tokai University Press, Tokyo, in 1981.

5.1 Undisturbed sampling.

(i) Intermittent sampling. T h e special samplers designed for the p u i p o s e of causing the
least possible disturbance during sampling share a c o m m o n feature. T h e y are so-called
thin-walled samplers with fixed pistons. T h e s e consist essentially of a sampler head, a
piston, piston extension rods and a sampling tube with a cutting edge. T h e sampler is
either pushed into the soil to the intended depth or lowered to the b o t t o m of a prebored
hole. M a n y different m e t h o d s of preboring exist, such as rotatory drilling, percussion
drilling, flight augers and hand augers. T h e choice of m e t h o d depends on the soil
characteristics and the depth of sampling.
T h e principles of the sampling operation in itself are m o r e or less the s a m e for all types
of samplers. In Fig. 4 3 , the sampling operation using the Swedish Standard Piston Sam-
pler I is shown as an e x a m p l e . W h i l e the sampler is placed in position for sampling, the
piston is fixed to the sampling tube (Fig. 43a). T h e piston is then released from the
sampling tube and rigidly fixed in relation to the ground surface (Fig. 43b). Finally, the
sampling tube is advanced until it is completely filled with soil (Fig. 43c).
In order to reduce friction between the sample and the walls of the sampling tube duri ng
the sampling operation, the inner diameter of the cutting e d g e can be m a d e slightly
Fig. 43. Principle of undisturbed sampling using the Swedish Standard Piston Sampler I:
(a) Piston locked to the sampling tube while the sampler is inserted into the soil.
(b) Piston rigidly fixed vertically during the sampling operation.

smaller than the inner diameter of the tube k n o w n as inside clearance. Therefore, it is
usually necessary to let the sampler stay in the soil after sampling for about 5 to 10
minutes to give the sample time to swell. (Swelling is d u e to the horizontal p r e s s u r e
release obtained w h e n the sample is punched into the tube). In this w a y e n o u g h adhesion
will b e developed for the sample to be torn off at the tip of the sampler w h e n it is
Soil investigations 71

edge taper angle

Fig. 44. Principles of sampler design.

withdrawn. T h e sample can also be sheared off at the tip by turning the sampler before
it is withdrawn. In s o m e cases, however, it m a y be necessary to use s o m e kind of sample
retainer or to h a v e s o m e arrangement for releasing the v a c u u m effect w h e n the sampler
is withdrawn from the soil. Retainers should be avoided if possible, as they always cause
some disturbance.
T h e inner diameter of the sampling tube varies with the type of sampler from 50 m m
to 100 m m and a length of 8 - 1 2 times the inside diameter is c o m m o n . T h e Swedish
Standard Piston Samplers h a v e an inner tube divided into three centrally placed 170 m m
lengths and t w o outer 85 m m lengths, i.e. a total sample length of 6 8 0 m m , or 13.6 times
the sample diameter. However, the soil samples in the two outer tubes, the one nearest
to the piston and the other nearest to the cutting edge, are considered as being disturbed.
T h e tubes are m a d e of reinforced plastic and can be used repeatedly. Using the Swedish
method, cutting of the tube in the laboratory is avoided.
A good sampler should fulfil certain requirements with regard to inside clearance ratio,
2 2 2
(^-), arearatio, (D2 -Dl )/Di , and e d g e taper angle (Fig. 44). T h e r e c o m m e n d e d
inside clearance ratio depends on the diameter of the sampling tube and on the type of soil.
In s o m e countries, inside clearance is not considered necessary. T h e r e c o m m e n d e d edge
taper angle depends on the area ratio; the higher the area ratio, the smaller the angle. M o s t
of the samplers h a v e an area ratio of about 1 0 - 2 0 % and an e d g e taper angle of 5 - 1 0
degrees. (The Swedish piston sampler, for e x a m p l e , which is considered to yield high-
quality samples of soft cohesive soils, has an area ratio of 21 % and an e d g e taper angle
of 5 degrees. T h e inside clearance ratio is 0.4%).

(ii) Continuous sampling. T h e frictional resistance in the sample/sampling tube interface,


which can entail serious disturbance, is the foremost obstacle to the possibility of taking
long, undisturbed samples. Therefore, in this case, friction has to be eliminated in one
way or another.
T h e first sampler, designed to fulfil the zero-friction requirement, w a s the Swedish foil
sampler. Here, the friction b e t w e e n the soil and the inner walls of the sampler is
eliminated b y m e a n s of foil strips of steel that are locked to the piston and unrolled from
a foil m a g a z i n e (placed in the sampler head) w h e n the sampler is a d v a n c e d into the soil
(Fig. 45). Consequently, the only friction obtained during sampling is the friction caused
by tensile strain in the foils. T h e inner diameter of the cutting e d g e of the foil sampler is
72 So/7 investigations

. Fastening of foils

. Piston (held stationary


during driving)

-Foil strips (unrolled as


as cutting edge is advanced)

. Foil magazine

Fig. 45. Principles of the Swedish metal foil sampler.

either 67 m m or 4 4 m m . Core lengths of u p to 25 m can be taken, but lengths of 10 m at


a time are m o s t c o m m o n .
In the Netherlands, a sampler, functioning principally in the s a m e way as the foil
sampler, w a s invented by B e g e m a n n . Here, the foil strips are replaced by a nylon stocking
with a total length of 18 m. In this case, the friction is eliminated by m e a n s of a liquid
3
lubricant with a density of about 1.7 t / m , placed in the space between the stocking and
the surrounding plastic tube. T h e inner diameter of the B e g e m a n n sampler is either 66
m m or 29 m m . T h e latter is not considered to yield undisturbed samples and is therefore
only used for classification purposes.
Continuous sampling offers a great advantage over intermittent sampling in that it
gives a detailed picture of the soil strata. Such a detailed picture can be of great
importance w h e n erratic soil conditions h a v e to b e dealt with. However, the sample
quality is not as good as in the case of a high-quality piston sampler being used.

5.2 Disturbed sampling.

If the only purpose of sampling is to classify the soil, undisturbed sampling m a y not b e
required. It is also m o r e or less impossible to carry out such sampling by conventional
m e t h o d s in the case of non-cohesive soils.
Soil investigations 73

T h e r e are a great n u m b e r of m e t h o d s used for taking disturbed samples, one of the most
c o m m o n m e t h o d s being the standard penetration test. A u g e r boring and core boring are
other c o m m o n methods of obtaining disturbed samples. Representative disturbed samples
can also b e obtained by the use of thick-walled rugged piston samplers. For shallow
sampling of hard soil, digging test pits is often the easiest and m o s t e c o n o m i c solution.

6. D E T E R M I N A T I O N O F D E F O R M A T I O N P R O P E R T I E S

6.1 Introductory remarks.

T h e loading conditions are very important for the response characteristics of the ground.
Normally, soil investigations are carried out with the a i m of determining the properties
of the soil in static loading conditions. However, the u n d e r g r o u n d r e s p o n s e in d y n a m i c
loading is m o s t probably very different from the response in static loading. T h e under-
lying causes of deviation between soil behaviour in static and d y n a m i c loadings are
differences in m a g n i t u d e of deformations and time effects. On o n e hand, the m a g n i t u d e
of the amplitude of shear deformation in d y n a m i c loading is generally m u c h smaller than
in static loading and, on the other hand, the loading conditions are quite different (gen-
erally cyclic loads of short duration).
In static loading, both strength and deformation properties are governed by consolidation
(hydraulic time lag) and creep p h e n o m e n a . L o n g - t e r m effects on strength and strain in
d y n a m i c loading are mainly due to the n u m b e r of loading cycles in combination with the
m a g n i t u d e of the amplitudes.
T h e soil can b e either strain-softening or strain-hardening, or b e h a v e like a m o r e or less
ideally elasto-plastic material. It can dilate or contract in shear d e p e n d i n g u p o n the den-
sity of its structure. Water-saturated granular soils in a loose state easily liquefy in shear.
Liquefaction can also take place in relatively dense soils subjected to long-term d y n a m i c
loading.
Investigations of strength and deformation properties of soil are chosen with due
regard to the the loading condition (static or dynamic), the permeability of the soil and
the possibility of taking undisturbed samples. If the latter is considered possible then
laboratory testing is often chosen, especially in cases w h e r e the soil characteristics are
governed by time-bound p h e n o m e n a (e.g. primary and secondary consolidation). If, on
the other hand, undisturbed sampling is not possible, in situ m e t h o d s are preferable.
T h e determination of the strength parameters is associated with a n u m b e r of uncertain
factors. In the case of n o n - c o h e s i v e soils, there is n o possibility of taking undisturbed
samples. Even if it w e r e possible to re-establish the in situ stress condition and the in situ
void ratio, possible cementation forces w o u l d h a v e been destroyed, the soil skeleton
subjected to rearrangement, ageing effects eliminated, etc. M o r e o v e r , considering the
size of soil v o l u m e tested and all the possible heterogeneities that can b e expected in the
underground, the limitations in the results obtained are obvious.
74 Soil investigations

6.2 Laboratory investigations

(i) Determination ofK. T h e bulk m o d u l u s can be determined by m e a n s of triaxial tests.


In the triaxial test, a circular-cylindrical soil specimen, enclosed by a rubber m e m b r a n e ,
is placed in a pressure cell filled with water or paraffin oil. T h e height of the sample is
normally twice the diameter, the latter being mostly 37 or 5 0 m m . T h e specimen is loaded
in the radial and the axial directions by applied cell pressure. In order to allow a change
of v o l u m e to take place during the test, it is carried out in a drained condition. Moreover,
the apparatus m u s t b e provided with s o m e arrangement which allows an accurate
m e a s u r e m e n t of the change in v o l u m e of the specimen.
T h e determination of the bulk m o d u l u s m a y s e e m simple but is quite complicated
unless the soil sample is completely water-saturated. Even the slightest a m o u n t of gas in
the voids will h a v e a great influence on the results of the test.

(ii) Determination ofG. T h e determination of the shear m o d u l u s can b e m a d e by means


of a shear test apparatus of the type shown in Fig. 4 6 . In order to prevent a v o l u m e change
during shearing, the specimen is surrounded by a n u m b e r of interspaced parallel metal
rings (or s o m e other similar arrangement which admits unprevented shearing), and the
vertical distance b e t w e e n the top piston plate and the b o t t o m plate is kept constant (direct,
simple shear test). Normally, the specimen is first consolidated in the apparatus under a
vertical pressure equal to that prevailing in situ before it is subjected to shear. T h e
inclination of the shearing stress vs. shearing strain curve yields the (tangent) shear
m o d u l u s G.

(ii) Determination of and v. Geotechnical laboratory investigations usually aim at


determining the elastic (or rather the pseudo-elastic) parameters J5and v, well-known from

Fig. 46. Direct shear test apparatus. Consolidation phase (a) and shearing phase (b). Before shearing,
the oedometer ring is removed and replaced by the rubber membrane and the metal rings (or by a
reinforced rubber membrane). The horizontal load is applied in steps until failure takes place.
So/7 investigations 75

classical m e c h a n i c s , rather than determining the AT and G values.The m o d u l u s of elasticity


and Poisson's ratio v c a n be determined by m e a n s of triaxial tests (Fig. 47). In this case,
the s p e c i m e n is loaded in t h e radial direction by applied cell pressure and in the axial
direction b y a combination of cell pressure and an additional external axial load. During
the test, the cell pressure is kept constant while the axial pressure is successively
increased.
As in this case of the bulk m o d u l u s , the v o l u m e c h a n g e of the s a m p l e has to be
measured with e x t r e m e accuracy and this is only possible if the s a m p l e is fully saturated.
T h e tangent m o d u l u s of elasticity can b e deduced from the results of the triaxial test
from the relation: (-or)
E = ^ ^ (43)

w h e r e ( - or) = c h a n g e in deviatoric stress,


dea = corresponding c h a n g e in axial strain.
Investigations by D u n c a n and C h a n g ( 1 9 7 0 ) h a v e s h o w n that the correlation between
deviatoric stress aa - Gr ( kept constant) and axial strain is satisfied by the equation:

a-<y = ^r-
r <>44

a + bea
w h e r e a and b are inverted moduli w h o s e determination is demonstrated in Fig. 4 8 .
Derivation of Eq. (44) in respect of yields the tangent m o d u l u s :

Fig. 47. Triaxial test apparatus.


76 Soil investigations

0.5-10-3

- kPa 'Tarct m b
+

0 V
0 12
Axial strain ,%

Fig. 48. Determination of the parameters and b in Eq. (44).

2
(45)
(\ + eaE0b)

where Eq=1/ is its initial value.


Also the value of Poisson's ratio can be determined by the s a m e type of triaxial test
where the cell pressure is kept constant. We have:
- Aer
- (46)
2

w h e r e - increase in axial strain,


- + 2 = decrease in v o l u m e ( positive for radial decrease).

(iii) Determination of consolidation characteristics. As stated previously, the modulus


=/&- 1 lmv is representative of the v o l u m e change obtained under a vertical stress
increase when deformations in the horizontal direction are prevented. It is normally
determined by oedometer tests or, m o r e seldom, by triaxial tests in which no radial
deformation is allowed to take place.
In the oedometer test, the specimen is inserted in an o e d o m e t e r ring, generally made
out of stainless steel, and placed in a loading device between t w o filter stones at the top
and bottom. T h e oedometer ring is either free to m o v e during the test(floating ring) or
fixed to the bottom plate (Fig. 49). In order to reduce friction between ring and specimen
in the best possible way, the interior part of the ring is m a d e very smooth and, furthermore,
is generally lubricated with, for example, silicon grease or m o l y b d e n u m disulphide.
T h e oedometer test is mostly used to investigate the consolidation characteristics of
fine-grained soils, especially of clays and organic soils, b u t sometimes also for
determination of the compression modulus of coarse-grained soils. In the former case, the
height of the specimen, according to the standard procedure, is chosen at half its diameter.
T h e load is applied step by step and each load step is kept constant for 24 hours or until
the pore pressure induced by loading has dissipated. Generally, the load applied in two
Soil investigations 77

Loading plate
Porous s t o n e -
Floating ring
Soil specimen

0 V< ; Porous stone


Bottom plate

Fig. 49. Sketch of oedometers with fixed and floating rings.

consecutive steps is doubled. T h e M value represented by the inch nation of the oedometer
virgin curve is called the primary compression modulus. It m a y have to be corrected with
regard to the influence of sample disturbance (Schmertmann, 1985).
T h e M value is very m u c h dependent on whether or not the effective stresses exceed
the preconsolidation pressure of the soil specimen (Fig. 50). A correct evaluation of the
preconsolidation pressure is of paramount importance and an uncritical determination of
G'C according to the Casagrande method must be warned against. T h u s , the outlook of the
compression curve in the semi-logarithmic plot of a disturbed clay specimen may
indicate the existence of a preconsolidation pressure although in reality no sign of a
preconsolidation pressure exists (Fig. 51).
As mentioned previously (p. 25), the primary compression index Cc - AelA(\ogo\),
introduced by Terzaghi, is frequently used in settlement analysis instead of M. Still another
deformation modulus, often used in settlement analysis, is 2 representing the compression
obtained w h e n doubling the consolidation pressure. T h e relation between these moduli
CR = Cc/(l+e0) = 2/\og2 is usually termed the compression ratio.

2.5 1 1 1 1
&

' L - 'c 1 _ (' - ' )m _ ,
<
+
+ ln[l + ] ' = 0 + '
/ XiL m ML
M0 1 1
03 4
0. &c - ' - &c ,
/
= - + if &c < ' < a'L
I

C/5 I 1

\
D
I _ ' - '0
'B
I if ' = 0 + ' <&c

\
WH
1
<u >r m

<U

1
1
! ^ ' 1
50 / 150 200
Effective vertical pressure ', kPa

Fig. 50. Correlation between M and effective overburden pressure "(Sllfors and Larsson, 1981 ; Larsson
and Sllfors, 1986).
78 Soil investigations

Effective pressure ' , in kPa (log. scale)


10 20 40 80 160 320 640 1280
>F 1 1 1 1 1 1 1

1 1
0.8 L- I 1 1 1
0 20 40 80 160 320
Effective pressure & v, in kPa (lin. scale)

Fig. 51. Oedometer test result taken from a consultant's report. The consultant evaluated the precon-
solidation pressure in the semi-log. plot at 80 kPa (according to Casagrande's method). By rewriting
the oedometer curve in linear scale we realise that there is no visible sign of a preconsolidation pressure.

In the conventional oedometer of the type shown in Fig. 4 9 , the soil layer submitted
to consolidation is drained at both top and bottom. Moreover, it can b e a s s u m e d that the
initial excess pore pressure u0 is constant throughout the specimen. A s s u m i n g that the
specimen has a thickness of 2h, the average consolidation degree {7V av can b e expressed
2
in terms of the time factor Tv- cv tlh through the relation:

t/ v > av = 2/7^7 (47)

w h e n Tv < 0.2,

f/ v av = 1 _ 0.81 l e x p ( - 2 . 4 6 8 T v ) (48)

when Tv > 0.2.

This gives us a basis for the determination of the coefficient of consolidation cv (Fig. 52).
N o w a d a y s , oedometer tests are often carried out by m e a n s of "constant rate-of-strain"
oedometer tests, so-called C R S tests. In this case the specimen is undrained at the b o t t o m ,
and the pore pressure at the impervious base of the specimen, induced during consolidation,
is registered. T h e coefficient of consolidation is evaluated from the relation:
So/7 investigations 79

_dd_\?_
(49)
dt 2ub

w h e r e dcfldt = rate of effective stress increase ( & a s s u m e d equal to - 2ub),


h = height of the specimen,
ub = p o r e pressure at the impervious b a s e of the specimen.
T h e coefficient of consolidation in horizontal p o r e water flow ch is not so easily
determined b y laboratory testing. It is therefore often d e t e r m i n e d in situ b y m e a n s of
piezo-cone sounding according to the m e t h o d described on p . 6 5 . T h e m a g n i t u d e of ch
in particular is very m u c h dependent on the possible existence of layers that are consid-
erably m o r e pervious than the soil as a w h o l e . E v e n in seemingly h o m o g e n e o u s clay, the
permeability in the horizontal direction is generally t w o to three times higher than in the
vertical direction. This fact justifies the use of in situ m e t h o d s .
T h e secondary compression index Ca = ocs(l + e0) can b e found from the inclination of
the tail of the o e d o m e t e r curve as shown in Fig. 18, p. 31 (cf. c o m m e n t s by Zeevaert, 1986).
According to Mesri and Godlewski (1977), the secondary c o m p r e s s i o n index Ca is
directly proportional to the primary compression index Cc. T h e most typical values of Ca/Cc
for inorganic soft clays and organic soft clays are 0.04 0.01 and 0.05 0 . 0 1 , respec-
tively. For sand, Mesri (1990) reports Ca/Cc = 0.02 0 . 0 1 .

Time of consolidation t
24 h
Load steps
0 - 1 0 kPa
1 0 - 2 0 kPa
2 0 - + 4 0 kPa

4 0 - 80 kPa

- 1 6 0 kPa

Fig. 52. Evaluation of the coefficient of consolidation from the results of oedometer tests with step by
2
step increase of consolidation pressure. For Uv = 50%, Tv = 0.197 which yields c v = 0 . 1 9 7 / i 5 O / i 5 O.
80 So/7 investigations

6.3 Field investigations

Direct in situ determinations of the deformation characteristics of soils can be carried out
in several w a y s , for instance by pressuremeter tests or dilatometer tests, and of course
by full-scale or half-scale loading tests. T h e deformation properties of soil under dynamic
loading conditions are c o m m o n l y determined by seismic m e t h o d s .

(i) The pressuremeter test. A m o n g the different in situ m e t h o d s , the pressuremeter has
proven to b e a most reliable tool for the determination of the deformation characteristics
of various types of coarse-grained soil and rock as well as overconsolidated fine-grained
soil. T h e pressuremeter (for details, see Baguelin et al, 1978), which was developed as
a practical tool by the French engineer Louis Mnard, consists of a cylindrical body with
originally three cellsa central measuring cell and two guard cells (Fig. 53). N o w a d a y s ,
pressuremeters with only one cell (a measuring cell), long enough to ensure that the end
effects are negligible, are also utilised.
After the pressuremeter is installed in the soil, the pressure in the cell(s) is increased,
which brings about a state of cylindrical expansion of the soil surrounding the measuring
cell. T h e radial deformation of the outer boundary of the measuring cell, according to the
M n a r d procedure, is obtained directly from the amount of water that is inflated into the
cell. T h e cell pressure is increased in steps and kept constant during each step for 2

Pressure gauge

Fig. 53. Sketch of the Mnard pressuremeter.


Fig. 54. Presentation and interpretation of the Mnard pressuremeter test.

minutes. Readings are taken after 30, 60, and 120 seconds. T h e readings h a v e to be
corrected with regard to the stiffness of the measuring device itself. T h e corrected
pressure vs. creep deformation from 30 to 120 seconds is plotted together with the
corrected pressure vs. total deformation (after 120 seconds) in a diagram of the type
shown in Fig. 54.
T h e results of pressuremeter investigations are greatly influenced by the installation
technique. To avoid disturbance in the best possible way, the M n a r d pressuremeter is
generally installed in the soil in a carefully prebored hole of the s a m e diameter as the
pressuremeter. However, in difficult soil conditions direct insertion inside a driven
slotted tube m a y b e necessary. A comparison between the pressuremeter moduli obtained
with and without the use of the slotted tube shows quite different results above, but no
significant difference beneath, the groundwater level. In dense and m e d i u m dense sand,
the moduli obtained without the use of the slotted tube h a v e been shown to vary from
about 40 to 7 5 % of the moduli obtained by its use (Hansbo and Pramborg, 1990). In the
literature it is clearly stated that direct insertion inside a slotted tube should be resorted
to only after all other m e t h o d s of installation h a v e failed. This, in m y opinion, is an
overstatement considering the costs involved in each unsuccessful trial. In Sweden, for
example, direct insertion inside a slotted tube has b e c o m e the rule rather than the
exception, and the results obtained have yielded settlement values that h a v e shown
acceptable agreement with those observed in practice.
In the evaluation of the pressuremeter test, the soil is assumed to b e h a v e as an elastic
medium. Since, in such a case, the stresses induced in the soil are of deviatoric character
the pressuremeter test e a r n e d out in a prebored hole yields the shear m o d u l u s :
82 Soil investigations

Gpr = (Vc+Vm)Ap/AV (50)

Vq = v o l u m e of the measuring cell at the beginning of the straight part of the


pressuremeter curve,
Ap/AV is the inclination of the straight part of the pressuremeter curve.
If the test is carried out inside a slotted tube the shear m o d u l u s is obtained from the
relation:

Av +v )(v + v )
c m t m (51)

w h e r e Vc as a b o v e and Vt = corresponding v o l u m e initially occupied by the slotted tube.


3
For a standard M n a r d pressuremeter w e h a v e Vc = 535 c m (outer diameter of cell
equal to either 60 or 44 m m ) . Using a slotted tube with an outside diameter of 64.8 m m
3
(meant for the 4 4 m m pressuremeter), w e have Vt =\ 160 c m .
Since, in fact, the soil does not deform elastically, the shear m o d u l u s derived from Eq.
(50) does not represent the shear modulus to be applied directly in formulae based on the
theory of elasticity. A recent study by Briaud (1992) on the effect on the pressuremeter
+
'first-load' m o d u l u s exerted by different ratios of elastic m o d u l u s in compression E to
+
elastic m o d u l u s in tension E~ indicates that, in reality, E is 2 to 3 times larger than the
pressuremeter first-load modulus. In the M n a r d interpretation this has been taken into
account by introducing a set of rheological coefficients with regard to soil type and
loading history.
Usually, the shear modulus determined by the pressuremeter test is replaced by the so-
called pressuremeter m o d u l u s Epr, assuming that Poisson's ratio of the soil is v v = 1/3,
i.e.:

Epr=2Gpr(l+vs) = &GprB (52)

In order to minimise disturbance effects due to installation, so-called self-boring


pressuremeters h a v e been developed both at C a m b r i d g e U n i v e r s i t y t h e Camcometer
(Fig. 5 5 ) a n d at Ponts et Chausses in ParisPAF (pressiomtre autofaureur). These
tools can only be used in relatively fine-grained soils and have, therefore, a restricted field
of application in comparison with the preboring pressuremeters of M n a r d type.

(ii) Dilatometer tests. T h e flat dilatometer (Marchetti, 1975) consists of a steel plate with
a thickness of 15 m m , a width of 96 m m , and a length of 2 4 0 m m (Fig. 56). On one of
its sides it is provided with a circular steel m e m b r a n e , 6 0 m m in diameter, flush with the
side. Inside the steel m e m b r a n e there is a pressure c h a m b e r and a distance g a u g e for
Soil investigations 83

V A

Fig. 55. Sketch of the Camcometer.

m e a s u r e m e n t of the m o v e m e n t of the m e m b r a n e when the pressure inside is changed. It


is inserted into the soil by the aid of sounding rods provided with an inner hole, 16 m m
in diameter, to m a k e r o o m for the connecting cable of the reading device.
T h e dilatometer is inserted into the soil by m e a n s of jacking. It is easily d a m a g e d and

96 m m

Fig.56. Principle outline of dilatometer test.


84 Soil investigations

is therefore mainly used in sand and silt. Under difficult conditions it m a y be advanced
by driving, but then the results should be treated with caution.
T h e evaluation of the dilatometer test is based on two limiting pressure values, namely
px representing the pressure required to produce 1.10 m m m o v e m e n t of the m e m b r a n e ,
and p0 being the pressure at zero m o v e m e n t of the m e m b r a n e . A s s u m i n g that the soil
behaves elastically, the dilatometer modulus can be deduced from the relation ( see
discussion by Ekstrm, 1989):

ED = 48A(Pl-Po) (53)

T h e dilatometer m o d u l u s is mostly used as a m e a n s of determining the tangent


compression (oedometer) modulus M of sand, silt and clayey silt. Experience has shown
that a good estimate of M is obtained by the relation:

M=\.\RmED (54)

where for sand (ID > 3):

m 0 . 5 + 21ogtfD > 0 . 8 5 ,

and for silt and silty sand (0.6 < ID <3):

Rm - [0.14 + 0 . 1 5 ( / D - 0.6)](1 - log KD) + 2.51og KD > 0.85,

and for clayey silt:

/ ? m 0 . 1 4 + 2 . 3 6 1 o g t f D > 0.85.

ID is a material index given by the relation:

, _P\ -Po
iD - ,
Po - "o
and u0 = pore water pressure at rest (no excess pore pressure),

Ovo
= effective overburden pressure.

If KD > 10, then RM = 0.32 + 2.181og Kt


Soil investigations 85

T h e ID value varies from about 0.6 to 1.8 for silt and is above 1.8 for sand. For normally
consolidated soil KD ~ 2.5. If KD > 2.5 it is r e c o m m e n d e d that the ID value b e corrected.
T h e correction to b e m a d e depends upon the depth of investigation.
For depths < 2 m :

IiD(corr) = / D - 0 . 0 7 5 ( t f D - 2 . 5 ) .

For depths > 2 m

D(corr) = / D- 0 . 0 3 5 ( ^ D- 2 . 5 ) .

T h e correction of the m e a s u r e d ID value in the case of sand and silt is generally of minor
importance.
T h e dilatometer is sometimes utilised as a m e a n s of determining the deformation prop-
erties of overconsolidated clays. However, the results are difficult to interpret as they
represent a partly u n d r a i n e d condition.

(iii) Plate loading tests. Plate loading tests, to full or r e d u c e d scale, are sometimes
considered the best m e a n s of determining the deformation characteristics of soils, but are
only used in exceptional cases b e c a u s e of the costs involved. T h e best w a y of performing
these tests is to use a testing procedure in w h i c h the load is applied in steps of equal
duration. First the failure load m a y be estimated on a theoretical basis. T h e n the load m a y
be applied in steps of about 5 - 1 0 % of the theoretical failure load. R e a d i n g s of the
settlement m a y b e taken 1, 2, 4, 8, and 16 minutes after the application of each new load
step. B y plotting the creep settlement from 1 to 16 minutes against the load, the critical
load can b e interpreted in a w a y similar to that used in the p r e s s u r e m e t e r test (see Fig. 54,
p. 81). T h e determination of the critical load which leads to excessive creep settlement
is j u s t as important as the determination of the failure load.
T h e settlement s observed at a certain load per unit area q (below the critical load) in
a plate loading test can b e used for the determination of the pseudo-elastic m o d u l u s of
elasticity of the soil according to the relation:
2
K(\-V )qD
E 55
= 4s <>

for a circular plate with diameter D ,

2
-0.815(1- v )qb/s (56)

for a square plate with width b.


Although Eqs. ( 5 5 - 5 6 ) are based on the assumption of a semi-infinite, ideally elastic
half-space, the m o d u l u s thus obtained can b e a s s u m e d representative of the soil v o l u m e
86 Soil investigations

nearest below the plate (to a depth of approximately five times the plate width or plate
diameter, cf. pp. 1 4 1 - 1 4 4 ) .
Plate loading tests using small-scale plates are frequently e a r n e d out. With regard to
normal variations in soil properties, the width of such plates should not be below 0.6 m
(cf. Fig. 2 5 , p. 3 6 ) . T h e so-called field compressometer developed at the N o r w e g i a n
Technical University (Janbu and Senneset, 1973) belongs to a special category of plate
loading tests that can b e carried out at various depths below the ground surface.

(iv) Seismic investigations. Seismic investigations are c o m m o n l y used as a means of


determining the moduli of deformation to be applied in the case of small strains of the
-5
order of m a g n i t u d e of ~ 1 0 or less, such as the strain amplitudes caused by vibrating
m a c h i n e foundations. T h e moduli thus determined (usually called d y n a m i c moduli) can
be considered approximately equal to G 0 , defined by Eq. (25).
T h e w a v e velocity vP which is generally used as a basis for determination of ground-
water level and depth to underlying harder layers (see p. 66 ) can also be applied to deter-
m i n e the d y n a m i c compression modulus, according to the relation:

M 0 = p v p2 (57)

w h e r e is the total bulk density of the soil.


However, since the w a v e is a compression wave, the value of M thus obtained will
be highly influenced by the presence of water (water can be considered as nearly incom-
pressible). Therefore, in the case of water-saturated soil, it is not representative of the
compression m o d u l u s of the soil skeleton. T h e results obtained yield too high a modulus
value.
T h e d y n a m i c shear m o d u l u s of soil can be determined from the shear w a v e velocity
vs . As water cannot sustain shear stresses, the shear w a v e velocity is independent of

Oscilloscope Oscilloscope

Input rg Trigger
Trigger
geophonc Hammer impulse
Impulse

Horizontal v e l o c i t y String and


transducers cable I m p u l s e rod

* S c r e w plate
Receiver geophones

Fig. 57. Down-hole and hole-to-hole methods for determination of the dynamic shear modulus of soil.
So/7 investigations 87

whether the soil is above or below the groundwater level. T h e shear m o d u l u s is obtained
from the relation:

2
G0 = pvs (58)

where as above.
T h e shear w a v e velocity can be determined directly in boreholes by so-called down-
hole (alternatively up-hole) or cross-hole techniques (Fig. 57). All these m e t h o d s com-
prise a shock i m p u l s e generator (anything from an ordinary h a m m e r to an explosive), a
pick-up (for e x a m p l e m o u n t e d in a seismic cone of the type developed at the University
of British C o l u m b i a ) and a registration instrument (normally an oscilloscope). By
generating shock w a v e s with inverted w a v e amplitudes, the shear w a v e velocity can be
determined with great accuracy.
T h e d y n a m i c shear m o d u l u s can also be obtained by measuring the surface w a v e (the
Rayleigh w a v e ) velocity from the relation:

G0~U5pvR* (59)

T h e surface w a v e is of particular interest since it represents the major part of the


oscillation energy (nearly 7 0 % ) and has very m u c h the s a m e velocity as the shear wave.
It m o v e s along the ground surface in a layer of about one wavelength in thickness.
Therefore, by changing the w a v e length of the surface wave, the variation of G 0 with depth
can be estimated.
T h e surface w a v e velocity can be determined by placing an i m p u l s e generator on the
ground and operating it at a certain frequency. T h e R w a v e velocity is then obtained from
the relation:

v*=A? (60)

w h e r e / = frequency (Hz) and = wavelength.

6.4 Empirical correlations with reference to deformation properties

(i)Stress correlations. As previously shown (Fig. 17, p. 29), the stress history of the soil
is of utmost importance for its deformation characteristics in static loading condition. T h e
most important parameter which has to be k n o w n in order to m a k e possible an accurate
prediction of foundation settlement is the preconsolidation pressure oc'. A drastic change
in long-term volumetric strain takes place w h e n the preconsolidation pressure is
exceeded. Of course, the stress history has also a similar effect upon the long-term shear
strains taking place in loading, i.e. upon the shear m o d u l u s to b e applied in the analysis.
88 So/7 investigations

Consequently, it is important to k n o w both the vertical and the horizontal stress histories
in relation to the actual vertical and horizontal stress levels.
T h e effective octahedral stress level c f o c i has also a certain influence on the magnitude
of the shear m o d u l u s G 0 (see, for example, Hardin and Richart, 1963; Hardin and Black,
1969). Hardin and Richart investigated the shear w a v e velocity for sands with different
void ratios. For sands with rounded and sharp-edged grains, respectively, the following
correlations w e r e established:

m
vs = 160(2.17 -e)() (61)
O r

1 M
vs = 110(2.97 - 0 ( ) (62)

w h e r e or represents a reference stress equal to 100 kPa.


3
For sand whith dry density pd - 2 . 6 5 / ( l + e ) t / m these values of shear w a v e velocity
yield the shear moduli:
<1
G ~690 f
0
(2
(63)
1+e

for sand with round grains, and

( 2 7 g )
G0 - 3 2 0 ^ ~ \/& (64)
1+ e

for sand with sharp-edged grains.


Similar investigations on water-saturated sand do not seem to exist. However, it can
be assumed that the G 0 value is not influenced by the degree of water saturation.
For clay soils Hardin (1978) suggests the following relation:

r & oa
G 0 = 625(OCR)*' ^ (65)

w h e r e O C R is the overconsolidation ratio //',


k' is a function of the plasticity index IP (Fig. 58).
T h e G 0 value given by Hardin is r e c o m m e n d e d by Larsson and M u l a b d i c ' (1991) to
be used in the case of low-plasticity clays and organic clays.
T h e preconsolidation pressure of cohesive soils which determines the O C R value can
be estimated on the basis of the undrained shear strength, either with regard to the liquid
limit (Hansbo, 1957):
Cu
0.45wL (66)
Soil investigations

Plasticity index IP (%)

Fig. 58. Overconsolidation adjustment parameter k' as a function of plasticity index IP.

or with regard to the plasticity index (Skempton, 1954):

/= ^ (67)
0.11+0.37/p

(ii) Strength correlations. For normally and lightly overconsolidated clays of high to
m e d i u m plasticity, Larsson and M u l a b d i c ' (1991) r e c o m m e n d the empirical relation:

10000

100
50

101 I I I I
7 6 5 4 3
- - - - io-
Shear strain amplitude

Fig. 59. Empirical correlation between G/cu and ^according to Larsson ( 1986) and Larsson & Mulabdic (1991).
90 So/7 investigations

G 0 - (208//p + 2 5 0 ) c u (68)
or, alternatively,

G0=504c> L
(69)

w h e r e cu - undrained shear strength of the clay,


G 0 = shear m o d u l u s for strains ~ 1 0 A
T h e correlation between normalised Glcu values (G representing secant modulus) and
shear strain according to different investigators has been analysed by Larsson & M u l a b d i c '
(1991). T h e result of their analysis is summarised in Fig. 59.
Recent investigations in Sweden (Gereben and Pramborg, 1990) of the Rayleigh w a v e
velocity in very stiff dry-crust clay (140 kPa < cu < 300 kPa) indicate the correlation:

2
G 0 - 6cu + 500cM (70)

(iii) Correlation with sounding resistance. T h o s e w h o h a v e a keen interest in the different


aspects related to the possibilities of estimating G 0 from sounding resistance are referred
to L o P r e s t i and Lai (1989). T h e empirical studies e a r n e d out on this matter are mainly
based on comparisons between shear w a v e velocities obtained by seismic methods
(cross-hole and seismic cone tests) and penetration resistance by the use of the S P T and
C P T sounding m e t h o d s . T h e results obtained, which mainly depend upon geological

24
a i
\ \ -* (X

20

16
\ OCR= 10
V \
12 \ \
\ \ \
s

OCR= 1

0
200 300 500 1000 2000 3000

Fig. 60. Correlation between G 0 and cone penetration resistance qc for quartz sand (Jamiolkowski and
Robertson, 1988).
Soil investigations 91

origin, grain size, cementation effects and stress history of the soil, show great dispersion,
particularly with reference to the S P T penetration resistance.
According to Imai et al. (1982), a rough estimate of the shear m o d u l u s G 0 , based on
comparison b e t w e e n S P T resistance and shear w a v e velocities, can b e obtained from the
relation (Imai et al, 1982):
G 0 14.1(N 3 0)0-68MPa (71)

T h e correlation b e t w e e n G 0 and the penetration resistance in the cone penetration test


can b e estimated from Fig. 60.
In a recent investigation of the correlation b e t w e e n seismic w a v e velocities and the
M n a r d pressuremeter moduli in sand and stiff clay, Gereben and P r a m b o r g ( 1990) found
the correlation:
G0 = 20Epr (72)

7. DETERMINATION OF STRENGTH PROPERTIES

T h e strength properties of soil are either determined in the laboratory by tests on m o r e


or less disturbed soil samples or directly in the field by in situ tests. Laboratory testing
has been very popular, especially a m o n g academics, but the results obtained are very
much dependent on the quality of the sample and on the testing m e t h o d . Therefore,
n o w a d a y s in situ testing has b e c o m e increasingly popular a m o n g consultants and con-
tractors as well as academics. T h e results obtained by in situ testing may, of course, also
be affected by disturbance of the soil, but the risk of disturbance is only related to the
installation of the testing e q u i p m e n t (similar to the disturbance due to the installation of
the sampling e q u i p m e n t ) while other sources of disturbance are avoided.

7.1 Laboratory investigations

(i) Determination of effective strength parameters. T h e effective strength parameters are


generally determined by m e a n s of triaxial and/or shear tests.
T h e execution of the triaxial test (Fig.47) was described in par. 5 . 1 . A s m e n t i o n e d
there, the s p e c i m e n is loaded in the radial direction by applied cell pressure and in the
axial direction by a combination of cell pressure and an additional external axial load. T h e
total axial stress applied can b e either higher (active test) or lower (passive test) than the
radial stress (Fig. 61).
In order to k n o w the effective stress level at failure, the tests h a v e to b e carried out
either in a drained condition or with pore pressure m e a s u r e m e n t s . Moreover, the speci-
men has to b e water-saturated. As regards fine-grained soils, the tests are generally per-
formed in an undrained condition with simultaneous observation of the pore pressure
induced due to loading.
92 Soil investigations

a) b)

Fig. 61. Active (a) and passive (b) triaxial tests. In the active test = { and Gr - 3 while in the passive
test - 3 and =

T h e test procedure is as follows.


1. A m i n i m u m of three specimens is selected.
2. T h e specimens are consolidated at three different stress levels.
3. In the active test, i.e. when the sample is in axial compression, the axial (major
principle) stress aa is increased either in steps (each step about 1/15 of the estimated
ultimate stress) or increased continuously. T h e rate of loading is chosen so as to obtain
complete pore pressure dissipation (drained test) or, if possible, to admit an equalisation
of the excess pore pressure induced due to loading (undrained test).
In the passive test, i. e. when the sample is in axial extension, the axial (minor principle)
stress is reduced in a corresponding way and the test otherwise e a r n e d out as described
above.
T h e test results are generally presented in a diagram showing the effective stress path
followed during the test. In the diagram shown in Fig. 62, the M o h r - C o u l o m b failure
criterion yields the relation:

c
l = ^(

- = [a + ^(cfa + efr)] sin ' (73)

w h e r e = (major principle stress)and = 3 (minor principle stress) in the active test,


while = 3 and or - in the passive test.
S o m e geotechnicians prefer to plot the stress path in a q vs. diagram, where q as above
and p'=(o'a - 2\). In such a case the M o h r - C o u l o m b failure criterion yields the
relation:

3 sin '
q = (p' + a) (74)
3 - sin '

in the active test, and:


3 sin '
q = (p' + a) /
(75)
3-fsin0

in the passive test.


Again, others prefer to plot the results in a q vs. a\ diagram, in which case the Mohr-
C o u l o m b failure criterion yields the relation:
- 100 .

Fig. 62. Effective stress paths obtained in active and passive triaxial tests on organic sulphide silt. The
marks on the curves represent axial strains 0 , 0 . 2 , 0 . 5 , 1, 2 and 3%. The 'failure lines' correspond to an
internal angle of friction of '= 35 (a = 0).

sine'
q = (&r + ah "", (76)
1 - sin
in the active test, and:

sine'
q = fr + a)- ^ (77)
1 + sin 0

in the passive test.


In the active shear test, the theoretical angle of inclination of the failure surface is 45
+ 0 7 2 and in the passive state 45 - 072. As the strength properties m a y vary with the
direction of shear strain (strength unisotropy), both active and passive triaxial tests are
often carried out, especially as regards clay specimens.
Besides triaxial testing, shear strength is also determined by direct shear tests using
special shear apparatuses, of which the C a s a g r a n d e shear box (Fig. 63) is c o m m o n l y used
in the case of coarse-grained soils, and, for example, the G e o n o r or the SGI shear
apparatus (Fig. 46, p. 74) is used in the case of fine-grained soils. In the C a s a g r a n d e shear
box, shear failure takes place along a horizontal plane, while in the G e o n o r and the SGI
shear apparatuses, it takes place in simple shear of the w h o l e specimen. In both cases,
94 Soil investigations



}
" Failure surface

Fig. 63. The Casagrande shear apparatus.

deformations at right angles to the direction of shear failure are prevented and consequently
w e have to deal with a plain strain condition which is often in better a g r e e m e n t with real
conditions in practice.
T h e testing principle is as follows: T h r e e specimens are selected and consolidated
under three different normal (vertical) stresses and then sheared to failure at a rate of
strain that has to b e adjusted with regard to the permeability of the soil (the rate of excess
pore pressure dissipation).
During the test, the principal stresses are undergoing a c h a n g e in both direction and
magnitude which m a k e s the interpretation of the test extremely difficult. T h u s , the i n t e r -
pretation of the results in respect of angle of internal friction will depend on the quotient
= \\ at failure which is u n k n o w n . A s s u m i n g that the main principle stress rotates
the angle a, the failure condition can b e expressed by the relation (Hansen, 1961):

sin0 ' s i n 2 a
(78)
a + <fv 1 + sin ' c o s 2

where is the horizontal shear stress at failure and a'v is the normal vertical stress.
Expressed in terms of the quotient K, the friction angle is obtained from the relation:

sin0' = (79)
a +
If in the simple shear test the occurrence of failure is not clear, failure can be assumed
to take place at a relative displacement between the piston plate and the b o t t o m plate of
about 5 % of the height of the specimen.
As mentioned above, the rate of shear deformation must be adjusted with regard to the
permeability of the soil, so in the case of clays it has to be kept very low. For e x a m p l e ,
for a high-plasticity clay specimen with a height of 20 m m and a diameter of 5 0 m m , the
Soil investigations 95

'

Fig. 64. Determination of failure criterion on the basis of the Mhr stress circles.

rate of relative displacement between the bottom plate and the top piston plate should not
3
exceed mm/min.

(ii) Determination of total strength parameters. If the tests are performed in such a way
that the effective stresses cannot b e determined then the results can only b e used as a
m e a s u r e of the so-called total strength parameters c and . This is, for e x a m p l e , the case
when the soil is not fully water-saturated, because if the soil contains pore gas the
effective stresses will be affected by surface tension forces of u n k n o w n m a g n i t u d e in the
menisci of the p o r e water. T h e total strength parameters are of practical interest when
failure can be assumed to take place under undrained condition. The most common
laboratory m e t h o d s of determination are the triaxial test, the unconfined compression
test, and the fall-cone test. By using either of the unconsolidated undrained triaxial test,
the unconfined compression test or the fall-cone test, only the undrained shear strength
parameter cu (or Tj-U), valid for the prevailing consolidation pressure of the specimen, can
be determined. If, for s o m e reason, the undrained shear strength at various consolidation
pressures is needed, then a n u m b e r of specimens first have to be consolidated in the
triaxial test cells at different pressures and then loaded to failure u n d e r undrained
conditions in the triaxial apparatus.
In the active unconsolidated undrained triaxial test, the specimen is loaded axially to
failure under undrained conditions with a cell pressure preferably equal to in situ
horizontal pressure. In the passive test, the specimen is loaded radially to failure under
undrained conditions with an axial pressure equal to in situ vertical pressure. Since the
effective stress path is u n k n o w n , the results are generally used as a basis for drawing the
M h r circles at failure (Fig. 64). T h e inclination of the failure envelope represents the
apparent angle of friction , and the intercept on the deviatoric stress axis represents the
apparent cohesion c. For water-saturated soil, w e find = 0 and in this case c is usually
denoted cu.
In the unconfined compression test (Fig. 65), the specimen, generally with its height
twice its diameter, is placed between a top and a b o t t o m plate and loaded axially to failure.
96 Soil investigations

h =2d

///////////,

Fig. 65. Unconfined compression test.

T h e test is merely a special case of the triaxial test with the radial pressure or equal to zero
and can evidently only b e used in the case of cohesive soils, preferably clay. It can either
be e a r n e d out with a constant rate of axial deformation or by continuously increasing
axial loading. T h e rate of axial strain ought to b e approximately 2.5%/min. In order to
prevent dessication of the specimen, it m a y be necessary to h a v e it enclosed by a sub-
stance such as paraffin oil.
T h e axial failure load is normally obtained from the m a x i m u m value of the load/
deformation curve. If there is no such m a x i m u m value, failure can b e assumed to have
taken place at an axial strain of 10%. T h e undrained shear strength, as given by the M h r
circle, is cu = 12.
T h e most simple and also the quickest testing m e t h o d is the fall-cone test (Fig. 66). T h e
test w a s developed by the Geotechnical C o m m i s s i o n of the Swedish State Railways (SJ:
Geotekniska K o m m i s s i o n e n , 1 9 1 4 - 1 9 2 2 ) . It is performed by letting a cone-shaped
weight (the fall-cone) fall freely from a certain height on to the plane surface of a cohesive
specimen. T h e penetration of the cone into the specimen is a m e a s u r e of the undrained
shear strength.
T w o standard cones are generally used for the determination of the undrained shear
strength of undisturbed samples, namely a 100g/30-cone (i.e. with a m a s s of 100 g r a m m e s
and an apex angle of 3 0 degrees), and a 60g/60-cone. S o m e t i m e s , in the case of very stiff
clays, a 4 0 0 g / 3 0 - c o n e is utilised. For determination of the r e m o u l d e d shear strength, a
10g/60-cone is also used. T h e cone is normally placed in a position with its tip j u s t in
Soil investigations 97

Fig. 66. The Swedish fall-cone test. A 100 g/30 cone placed in position to be dropped.

contact with the p l a n e surface of the specimen (Fig. 66). T h e c o n e is then dropped freely
into the clay and the depth of penetration measured.
T h e undrained shear strength of clay can be obtained by the following relation
(Hansbo, 1957 and 1962):

Km g h
(80)
lL l

0.08 0.1 0.2 0.3 0.4 0.5 1.0


Parameter

Fig. 67. Relation between the parameter and the apex angle of the fall-cone.
98 So/7 investigations

w h e r e m - m a s s of the cone,
g = acceleration of gravity,
i - depth of penetration (indentation),
h - distance between cone apex and soil surface when the cone is dropped (nor-
mally h - 0)
= function of the apex angle of the cone.
If calibrated against the field vane test, for samples taken by m e a n s of the Swedish
standard piston sampler can be taken from Fig. 67.
T h e fall-cone test is frequently used as a standard procedure in the determination of the
sensitivity of cohesive soils. T h e results are generally in good agreement with the results
obtained by the field vane test.

7.2 Field investigations

As mentioned in the introductory remarks to this paragraph, in situ testing has many
advantages over laboratory testing and has b e c o m e increasingly popular in applied
geotechnical engineering. In situ testing can either be used for determination of the
strength parameters to be used as input data in bearing capacity formulae, as was the case
in laboratory testing, or for direct determination of bearing capacity. Tests which allow
the latter type of interpretation are usually preferable in practical applications. A m o n g the
most c o m m o n in situ m e t h o d s w e find the field vane test, used to determine the undrained
shear strength of cohesive soils, the pressuremeter test, the results of which can be used
both for determining strength parameters and the bearing capacity of foundations, and the

Fig. 68. Field vane test apparatus, type Nilcon (left) and type SGI with protective housing for vane
during installation. Standard vane, 130 mm in height and 65 mm in width.
9
Soil in vestigations

dilatometer toi which can be used both for soil classification purposes and for determining
strength parameters.

(i) Determination of effective strength parameters. Determination of effective strength


parameters (normally the angle of internal friction) in situ is usually carried out by m e a n s
of the pressuremeter or the dilatometer. However, there is seldom a need for translating
the results of these tests into effective strength parameters. In foundation engineering this
is an unnecessarily r o u n d a b o u t method. T h e results can b e used directly for calculation
of the bearing capacity of foundations (see p p . 1 3 5 - 1 3 9 and 1 9 0 - 1 9 2 ).

(ii) Determination of undrained strength parameters. T h e undrained shear strength is


generally determined by m e a n s of the field vane test (Cadling and Odenstad, 1950).
In the field vane test, a fourbladed vane attached to a sounding rod (Fig. 68) is pushed
into the soil to the depth of strength determination. T h e assembly is then rotated until
failure is reached. T h e relation between torque and angular rotation of the vane is
generally registered on a diagram. Investigations h a v e shown that the soil fails along a
cylindrical surface with the same diameter as the width of the v a n e and along plane
horizontal circular surfaces at the top and b o t t o m of the vane.
T h e standard vane is 130 m m in height and 65 m m in width. Vanes with heights of 80
m m and 170 m m and with width of 40 m m and 80 m m , respectively, are used in cases
where the standard v a n e is not applicable. According to the standard procedure, the vane
is left in the soil o n e minute before rotating it to failure, which should occur a minute later.
T h e undrained shear strength, according to the field v a n e test, can b e calculated from the
observed m a x i m u m torque, M m a x, according to the relation:

2M
u 2 ) ( 8 1
nd h(\+d/3h)

w h e r e h = height of the vane blade,


d = the width (diameter) of the vane.
T h u s , if h = 2d as for the standard vane, w e have:

cu = 0.213Mmax/di (82)

If the shear strength is different in the vertical (= cuv) and in the horizontal directions
(= cuh ), w e have: j
2
Mmai = -nd h(cuv +cuhd/3h) (83)

(iii) Correction with regard to anisotropy and time to failure. As shown in Section 5.5,
the shear strength mobilised in clay in an undrained condition is dependent on the strain
rate the shorter the time to failure, the higher the mobilised strength (Fig. 69). This is
100 Soil investigations

0 10 20 30 50 70
Torsional angle, degrees

Fig. 69. Shear stress vs. strain relationship in soft high-plasticity clay determined by field vane tests with
different rates of vane rotation. Times to failure: (1) 1.2 s, (2) 7 s, (3) 1 min, (4) 10 min, (5) 100 min,
(6) 1000 min and (7) 10000 min (Torstensson, 1973).

particularly the case with clays of high plasticity. Moreover, the strength m a y vary with
the direction of shear deformation because of anisotropic properties of the clay. T h e
undrained shear strength will therefore have to be corrected with regard both to time and
to anisotropy. For soils of extremely high plasticity, the undrained shear strength will only
have to b e corrected with regard to time (cf. Eq. 84). T h e lower limit of the shear strength

Plasticity index Ip (%)


0 20 40 60 80 100

0 40 80 120 160 200


Liquid limit wL (%)

Fig. 70. Correction of undrained shear strength with regard to liquid limit (Andrasson, 1974)
continuous lineand plasticity index (Bjerrum, 1973)dashed line.
Soil investigations 101

prevalent in long-term loading is defined as the creep strength ccr (cf. p. 34). T h e correction
factor ^ i s generally related to the plasticity index or, alternatively, to the liquid limit (Fig.
70). T h e value presented as a function of the plasticity index by Bjerrum was based
on results of test e m b a n k m e n t s loaded to failure and on case records. T h e liquid limit has
served as a basis for correction in S w e d e n since the 1950s.
T h e undrained shear strength determined by the field vane test can be corrected with
regard to both time and anisotropy according to the relation (Larsson, 1977):

2 2
c c o s 0 + ( 0 . 1 7 + 0.7vv L) s i n 0
^,corr = : (84)
0.45wL 3
w h e r e 0 i s the inclination of the main principal stress to the vertical [ 0 i s 0in the active
test, 90 in the passive test and 60 in the direct shear test (' assumed equal to 30)].

7.3 Empirical correlations with reference to strength properties

(i) Correlations with preconsolidation pressure. Several empirical correlations between


the undrained shear strength of clay and the preconsolidation pressure have been
suggested. For instance, in a case w h e r e the preconsolidation pressure of the clay is
known, the undrained shear strength can be estimated on the basis of Eqs. ( 6 6 - 6 7 ) .

(ii) Correlations with penetration resistance. In m a n y cases, the determination of the


shear strength parameters by laboratory testing will yield unreliable results, partly due
to sample disturbance (breakdown of original structure) and partly to the effect of
changes in stress history in comparison with in situ conditions. Field investigations adapted
for direct m e a s u r e m e n t s of the shear strength m a y be considered too expensive or, for
s o m e reason or other, be impossible to use. In such cases, local experience can be used

ol I I I2 I3 I4 I 5
o.i
Time to failure, min.

Fig. 71. Normalised undrained shear strength vs. time to failure for Drammen clay (Bjerrum, 1973).
The creep strength ccr 0.75c M.
102 So/7 investigations

to establish fairly reliable correlations between the results of penetration tests and the
strength parameters.
A m o n g the various penetration methods, the cone penetration test with a friction
sleeve and pore pressure transducers mounted at the tip (the C P T U penetration test)
seems to b e quite a useful tool for determination of the effective strength parameters
(Sandven et , 1988; Senneset et al, 1989). A pore pressure parameter Bq is introduced,
determined from the measured excess pore water pressure AuT and net c o n e resistance
qn (= qT - ' 0) according to the relation:

Bq=AuT/qn (85)

and a cone resistance parameter Nm, determined from the effective overburden pressure
' ^ and the net cone resistance according to the relation:

Nm = qnK^v0 + a) (86)

w h e r e a = attraction.
A s s u m i n g that the failure zone beneath the cone tip is in a g r e e m e n t with the Prandtl
solution for a shallow footing, the cone resistance parameter can b e expressed by the
relation (Sandven et ai, 1988):

Nm = (Nq-l)/(l+NuBq) (87)

2
w h e r e Nq = tan (45+072)exp(7Ctan0O,
N M 6 t a n 0 ' ( l + tan0O-
Introducing the values of Bq and Nm into Eq. (87), the angle of internal friction ' can
be determined (Fig. 72).
If the soil is h o m o g e n e o u s , the attraction a can be obtained directly from the penetration
resistance curve b y extrapolating it to zero overburden pressure. For i n h o m o g e n e o u s
soils, the following values of a (in kPa) m a y b e applied (Sandven et ai, 1988):
5 - 1 0 for soft clay, 1 0 - 2 0 for m e d i u m clay and 2 0 - 5 0 for stiff clay,
0 - 5 for soft silt, 5 - 1 5 for m e d i u m silt and 1 5 - 3 0 for stiff silt,
0 for loose sand, 1 0 - 2 0 for m e d i u m sand and 2 0 - 5 0 for dense sand.

8. PRESENTATION

T h e results of geotechnical investigations ought to b e presented in a w a y to facilitate


comprehension for the people concerned and to enable t h e m to grasp the context. D r a w -
ings should comprise plans showing the situation of the soil investigation points a n d other
objects of interest; sections showing the ground level and all the results of field
investigations (results of penetration tests, groundwater observations, soil s a m p l e
Soil investigations 103

Pore pressure parameter

1 2 5 10 20 50 100 150
Cone resistance parameter Nm

Fig. 72. Assessment of effective angle of internal friction from the results of CPTU penetration tests

classification, shear strength values, consistency limits, natural water content, etc.) as
exemplified in Fig. 7 3 .
Presentation of laboratory test results should include all information needed for the
appreciation of the testing routines applied and of the deformation and strength parameters
stated in the report. It should always b e possible to check the results presented. This is
a vital part of a quality assurance system.
An interpretation of the results of the geotechnical investigation should b e presented
in a special geotechnical report aimed at forming a basis for solving the p r o b l e m s which
arise in the foundation project in question.
In order to avoid misinterpretations of the information given in the report, the final
solution to the foundation problems should b e w o r k e d out in close cooperation with the
structural engineers involved in the project as well as other representatives of the client.
T h e situation previously experienced, w h e r e the geotechnical engineer, after having
delivered his report, had no connection at all with the structural engineer or the con-
tractor, or the client for that matter, until something w e n t w r o n g with preparatory foun-
dation works or d a m a g e caused by excessive settlement took place, has hopefully c o m e
to an end.
104 So/7 investigations

Fig. 73. Example of a borehole section and corresponding inteipretation of subsoil stratification. By
courtesy of J&W. Abbreviations for names of soil according to Karlsson and Hansbo (1981).
Spread foundation 105

SPREAD FOUNDATIONS

1. INTRODUCTION

Spread foundations comprise footings placed at shallow depth for the support of
individual structural columns and walls. A footing that supports a single c o l u m n is called
an individual footing whilst one that supports a group of columns is a c o m b i n e d footing
and one that supports a wall or a row of columns is a strip footing. In soft soil conditions,
and also in s o m e other special cases w h e r e there is a need to reduce the m a x i m u m
foundation pressure, it is advantageous to replace foundation on footings by a c o m m o n
raft designed to carry the total load of the building. In both types of foundation, the load
of the building will h a v e to be carried by direct contact stresses at the footing/soil
interface which places strong requirements on the geotechnical characteristics and
behaviour of the subsoil.

1.1 Depth of foundation

(i) Regional requirements. T h e depth of foundation is not only dependent on the strength
and deformation characteristics of the soil but also on the climatological conditions and
the response to these conditions exerted by the soil. In s o m e places w e encounter soils
that can give rise to serious foundation difficulties, for e x a m p l e soils susceptible to frost
action, soils disposed to swelling, collapsible soils, chemically unstable soils, etc.
In places with frost-susceptible soils, the footings have to be placed below the depth
of frost penetration, or in regions with permafrost below the depth of thawing.
Different m e t h o d s of reducing frost penetration and the depth of thawing can be used,
mostly by insulation or by feeding heat or cold into the soil.
In places with swelling soils, such as smectite, the footings must be located at a depth
that is not influenced by drying. S o m e t i m e s it is necessary to anchor the foundation in
deeper layers to prevent the building from being d a m a g e d when water is sucked up by
the soil during rain seasons. Efficient drainage around the building and protection agai nst
water infiltration is needed.
Residual soils, such as latrites, are often covered by a collapsible top layer. This layer
is ordinarily firm and stable. However, when it b e c o m e s saturated, the soil skeleton
undergoes a complete collapse. Therefore, the footings h a v e to b e placed b e l o w the col-
lapsible top layer or the top layer has to be compacted.
Buildings with shallow foundations on clay frequently suffer d a m a g e by differential
106 Spread foundation

Fig. 74. Footings cast on the surface of compacted rock fill.

settlements caused by water-suction from nearby trees, such as poplars, oaks, birches,
elms and linden trees. In order to avoid d a m a g e , trees should not be planted closer to a
building than its expected full-grown height.
Building activities often affect the infiltration of rain water and m a y cause a lowering
of the groundwater table. In consequence, settlements can be induced both by dry crust
formation and by long-term increase of effective vertical stresses throughout the soil
profile. T h e dry crust formation is mainly governed by vegetation and other sources of
evaporation. Withering of vegetation starts at a water binding pressure of around 1.5
MPa. This entails that the shrinkage limit is decisive for the m a x i m u m v o l u m e decrease
entailed by dry crust formation.

(ii) Depth of foundation with regard to geotechnical requirements. T h e choice between


shallow and deep foundations of a structure is last of all depending upon the depth to
bearing strata and/or the preloading history of the subsoil. In a n u m b e r of cases, a deep
foundation is c h o s e n a l t h o u g h not n e e d e d b e c a u s e of the designer's belief that,
otherwise, settlement would exceed the permissible value. A m o r e detailed or advanced
soil investigation than the one at hand m a y often result in a m o r e cost-effective solution
with shallow foundation, or with a combination of shallow and deep foundations.
K n o w l e d g e of the deformation characteristics and of the loading history of the subsoil
is of p a r a m o u n t importance in the design.
Lack of suitable ground for building purposes, especially in densely populated areas,
makes it necessary to optimise the design of the building by increasing the n u m b e r of
basement floors. In consequence, the weight of the soil that has to be excavated for the
basement often exceeds the weight of the building itself. We then h a v e to deal with a so-
called compensated foundation. T h e building can be c o m p a r e d to a boat which is floating
Spread foundation 107

due to the uplift caused by the combined effect of effective pressure acting at the soil/raft
interface and pore water pressure. Obviously, in this case a raft foundation is always
possible, although in s o m e situations, for instance under heavily loaded c o l u m n s , the raft
may h a v e to b e supported by piles.

2. INDUCED PRESSURE DISTRIBUTION

2.1 Foundation on footings

(i) Contact pressure. T h e distribution of contact pressure at the footing/soil interface


depends on the mechanical characteristics of the soil and the bending rigidity of the
footing. Regarding individual and continuous footings, the bending rigidity is generally
very large in comparison with the deformability of the soil. T h e analysis of the contact
pressure distribution can therefore be m a d e on the assumption that the footing is infinitely
rigid. However, confined rigidity may h a v e to be given certain considerations in the case
of c o m b i n e d footings.
A characteristic of an infinitely rigid footing is that there is equal vertical deformation
at the soil/footing interface. A s s u m i n g the soil to b e h a v e as an elastic m e d i u m , the contact
pressure distribution beneath a strip footing is governed by the relation (Fig. 75):

(88)

where = line load acting along the centre of the footing,


b = width of the footing,
= distance from the centre line of the footing.
For a circular footing, the corresponding relation is:

(89)


0.5b

P/b

Fig. 75. Contact pressure distribution beneath a rigid strip footing on an ideally elastic medium
108 Spread foundation

Fig. 76. Observed contact pressure distribution at various loads in stiff clay. Rigid footing, 0.3 m in
diameter, placed on ground surface. (Faber, 1933)

w h e r e Q - point load acting in the centre of the footing,


D = diameter of the footing,
= radius vector.
We find that > oo when x-b/2- and when p->D/2-. This can b e c o m p a r e d with the
pressure distribution beneath a uniformly distributed flexible load w h e r e the contact
pressure at a given point corresponds directly to the load intensity at the point in question.
In reality, there is no m e d i u m that can resist an infinitely large pressure. T h e m a x i m u m
possible contact pressure will be governed by the strength characteristics of the soil.
Therefore, a redistribution of the contact pressure in the direction towards the centre of
the footing will take place. T h e extent of the redistribution is dependent on the soil
characteristics and the loading conditions.
For a footing on clay (Fig. 76), the upper limit of e d g e pressure will b e governed by
the shear strength of the clay. Since clay, if subjected to a small e n o u g h load, behaves as
an elastic m e d i u m , the pressure distribution will follow closely the distribution given by
Eqs. ( 8 8 - 8 9 ) with the exception that the outer edge zone will yield. T h e higher the load,
the broader the zone of yielding until, finally, failure will take place. T h e contact pressure
distribution will thus depend on the intensity of the load and the shear strength of the clay
(undrained in rapid loading; drained in long-term loading).
For a footing on cohesionless soil the pressure distribution will b e quite different. In
this case the shear strength depends on the effective stress condition. T h u s , for a footing
placed on the ground surface, the effective stresses just outside the e d g e of the footing
are dependent on possible existence of capillary forces, possible loading of the ground
surrounding the footing, etc. If there is no such loading, and the g r o u n d w a t e r level is j u s t
below the ground surface, the effective stresses at the foundation level, just outside the
footing, will b e very close to zero. Therefore, unless w e h a v e to deal with c e m e n t a t i o n
effects, the shear strength of the sand will b e close to zero and, in c o n s e q u e n c e , the contact
Spread foundation 109

Fig.77. Observed contact pressure distribution at various loads in sand. Rigid footing, 0.3 m in diameter,
2
placed on ground surface without and with a surrounding surcharge of 7 k N / m . (Faber, 1933).

pressure along the e d g e of the footing also close to zero. T h e pressure distribution will
follow the trend s h o w n in Fig. 77.
Usually, b e c a u s e of the footings being placed at a certain foundation depth, whether
this is related to the ground surface or to the b a s e m e n t floor, a certain overburden pressure
is always acting at the foundation level outside the footings. This allows for a higher edge
pressure also for footings on cohesionless soil (Fig. 77, right).
T h e contact pressure distribution for footings on silt is in b e t w e e n the distributions
given for clay and sand (Fig. 78).
B e c a u s e of the various factors that influence the contact pressure distribution, not least
the unforeseen variations in soil properties, effect of capillary forces, cementation, etc.,
it is extremely difficult to predict the real contact pressure distribution induced b y the
load. In practice, however, the structural design of a footing can generally b e e a r n e d out
on the assumption of evenly distributed contact pressure.

(ii) Pressure distribution with depth based on theory of elasticity. T h e variation with depth
of the pressure caused by the load applied on the footing is usually calculated on the basis
of the theory of elasticity. Practical experience has s h o w n that this can be done without
serious errors. Solutions also exist w h i c h are based on statistics w h e r e the soil is
considered as a particulate m e d i u m (see p. 115).
E v e n though the contact pressure distribution under a rigid footing is subjected to large
variations, the difference in stress condition u n d e r n e a t h a rigid footing on the one hand
and a flexible footing on the other is m o r e or less negligible at depths e x c e e d i n g the width
of the footing (Table 20). Therefore, the stress increase occurring at greater depths can
be calculated on the assumption that the contact pressure is equivalent to an evenly
distributed flexible load. O n this basis, the analysis of the stress increase is generally
performed according to Steinbrenner (1936) or N e w m a r k (1942).
110 Spread foundation

Fig. 78. Observed contact pressure distribution at various foundation loads in silt. Rigid footings, 0.31
m in diameter, placed on ground surface and at different depths df. (Helenelund, 1965).

The principal stresses induced by a flexible strip load can be found by the relation (Fig. 79):

q
= - ( y / + s i n ) (90)

T A B L E 20
Stress distribution at depth underneath rigid and flexible strip footings with width b according to the
theory of elasticity (Szechy, 1965).

zJb xlb = 0 xlb =0.5 xlb = 1 xlb = 2


rigid flexible rigid flexible rigid flexible rigid flexible

0 0.637 1.000 oo 0.500 0 0 0 0


0.25 0.683 0.960 0.710 0.493 - -
0.5 0.676 0.818 0.535 0.480 0.104 0.084 0.006 0.005
1.0 0.513 0.550 0.407 0.409 0.186 0.185 0.031 0.029
1.5 0.383 0.396 0.329 0.334 0.215 0.211 0.061 0.059
2.0 0.300 0.306 0.271 0.275 0.209 0.205 0.085 0.083
3.0 0.206 0.208 0.196 0.198 0.170 0.170 0.103 0.103
Spread foundation 111

Fig. 79. Geometric determination of the principal stress at a given point below a flexible strip load.

3 = ~(- sin ) (91)



whence, since & = 0:

2q
2 - (92)

T h e principal shear stresses b e c o m e :

T U = ( - 3 ) / 2 = - sin (93)

A stress distribution corresponding to that induced by a flexible load is encountered


when dealing with tankfoundations. Tanks are large-diameter cylindrical steel containers
filled with liquids and can therefore, from a geotechnical view-point, be considered as
circular, evenly distributed surface loads. T h e stress distribution caused by a circular,
evenly distributed load on an elastic half-space was first solved by Love (1929). The
results are presented in diagrammatic form in Fig. 80.
By Steinbrenner's method, the stress increase can be found in a fairly simple way at
an arbitrary point underneath, or outside, a rectangular load (or a load that can be divided
into several rectangular part areas). This method is based on the following reasoning.
Assuming a rectangular load of length 21 and width 2b, this can b e divided up into four
congruent part rectangles / b. T h e stress increase underneath the centre of the load can
now, according to the principle of superposition, b e calculated as the sum of the stress
increase obtained below the corner of each o n e of the four part rectangles (in the centre
of the load). T h e stress increase under a corner of a rectangular load of length / and width
b is thus equal to one fourth of the stress increase under the centre of a rectangular load
of length 21 and width 2b.
Accordingly, w e find the following expression for the stress increase at depth below
the corner of a rectangular load with length / and width b\
9 9
q mn(\+m +2n ) J
m n / rx w
= [ +arcsin-7= 1 (94)
2 2 2 2 2 2 2 2
2 (i + n )(l + n ) / l + w + n /(m + n )(l + )

w h e r e m = lib,
- zlb.
N o w , considering a loaded area that can be divided up into a n u m b e r of part rectangles,
Eq. (94) m a k e s it possible to calculate the vertical stress induced b y the load for any point
and for any depth.
T h e pressure distribution under a load area of arbitrary shape can be obtained by the
Spread foundation 113

aid of Newmark's influence diagrams, shown in Figs. 8 1 - 8 2 . T h e following procedure


is used:
Let us a s s u m e that w e want to find the stress increase at depth below a given point
A underneath, or outside, the loaded area. Draw a figure of the loaded area to such a scale
that the depth b e c o m e s equal to the length designated in the diagram. T h e figure of
the loaded area is then to be placed on the diagram in such a way that point A coincides
with the centre of the diagram. As regards the vertical stress increase it is irrelevant how
the figure is oriented (Fig. 81 ). On the other hand, as regards the horizontal stress increase,
the figure has to be oriented so that the direction of the horizontal stress looked for
coincides with the vertical axis of the diagram (Fig. 82).

Fig. 81. Newmark's influence diagram for determination of vertical stresses under a load area of
arbitrary shape.
114 Spread foundation

Fig. 82. Newmark's influence diagram for determination of horizontal stresses under a load area of
arbitrary shape.

Example 5: Calculate the vertical stress at 2 m depth below points A, and C due to an evenly distributed
load shown on top of p. 115.

Solution: The stress below point A is the sum of the stresses below corner A of the rectangles GHIA (m
=1.5; = 0.5), IJDA (m = 1.0; =0.5) and DEFA (m = 1.0; =0.5) whence:
azA/q = 0.238 + 2-0.232 = 0.702
The stress below point is the sum of the stresses below corner of the rectangles RGLB (m = 4.0;
= 1.0) plus LHVB (m = 4.0; = 1.0) plus QFKB (m = 3.0; = 1.0) plus PEQB (m = 3.0; = 1.0) plus
VJPB (m = 1.0; = 1.0) minus RAKB (m = 1.0; = 1.0), the effect of which would otherwise be included
twice. Hence:
/ = 0.204-2 + 0.203-2 + 0.175 - 0.175 = 0.814
Spread foundation 115

" ^~ ' I
, HI V J1
g

X
G

////////
F
Q 2 m
U - -^
6m 4m

The stress below point C is the sum of the stresses below corner C of the rectangles SGMC (m = 2.2;
= 0.4) plus TFUC (m = 1.8; = 0.4) minus SAUC (m = 1.0; = 0.4) minus TE NC (m = 9.0; - 2.0)
minus OHMC (m = 11 ; - 2.0) plus OJNC (in - 1.0; = 2.0) which would otherwise be deducted twice.
ozClq = 0.244 + 0.244 - 0.240 - 0.137 - 0.137+ 0.084 = 0.058
The stress below point is the sum of the stresses below corner of the rectangles RGLB (in = 4.0;
= 1.0) plus LHVB (m = 4.0; = 1.0) plus QFKB (m = 3.0; = 1.0) plus PEQB (in = 3.0; = 1.0) plus
VJPB (m = 1.0; = 1.0) minus RAKB (m = 1.0;= 1.0), the effect of which would otherwise be included
twice.
azB/q = 0.204-2 + 0.203-2 + 0.175 - 0.175 = 0.814
The stress below point C is the sum of the stresses below corner C of the rectangles SGMC (m = 2.2;
= 0.4) plus TFUC (m = 1.8; = 0.4) minus SAUC (m = 1.0; = 0.4) minus TENC (m = 9.0; = 2.0)
minus OHMC (m = 11 ; = 2.0) plus OJNC (m = 1.0; = 2.0) which would otherwise be deducted twice.
ozClq = 0.244 + 0.244 - 0.240 - 0.137 - 0.137+ 0.084 = 0.058

(iii) Probabilistic approach. In the probabilistic approach (cf. p . 479), the soil is consid-
ered as a particulate m e d i u m . In such a soil, is zero at any given point located in a void.
T h e real value of in relation to the expected value E(az) can b e calculated b y the
probabitity relation:
/ E ( c )
7 n(-lnn)^ ^
p[ (95)
() [/()]\

w h e r e is the porosity of the soil.


T h e expected vertical stress value at depth below a corner of a rectangular load area
of length / and width b can b e obtained b y the correlation (Harr, 1977):

E(az)=q(-^3)(-/3) (96)

w h e r e () represents the area below the normal distribution curve from 0 to . They/(x)
value can b e found in standard mathematical tables.
T h e expected vertical stress at any point below a load area of arbitrary shape that can
be divided up into a n u m b e r of rectangular areas can n o w b e calculated by superposition
according to Steinbrenner's m e t h o d as described above.
116 Spread foundation

Example 6: Calculate for the load according to Example 5 the vertical stress to be expected at 2 m depth
below points A and C in a particulate medium.

Solution: The stress below point A is the sum of the stresses below corner A of the rectangles GHIA (//
= 3; biz =2) plus IJDA (l/z = 2; biz =2) plus DEFA (llz = 2; biz =2) whence:
2 2
E(ol )lq = v ( 3 / 3 ) V(2/3) + 2[y/(2/3)] = 0.5 0.5 + 2 0.5 = 0.750
The stress below point C is the sum of the stresses below corner C of the rectangles SGMC (Il = 5.5;
biz = 2.5) plus TFUC (llz = 4.5; biz = 2.5) minus SAUC (l/z = 2.5; biz = 2.5) minus TENC (llz = 4.5;
biz = 0.5) minus OHMC (llz = 5.5; biz = 0.5) plus OJNC (llz = 0.5; biz = 0.5) which would otherwise
be deducted twice. Hence:
2
E(GzC)lq = y/(9.53)y/(4.33) + v<7.79)v<4.33) - [y/(4.33)] - y/(7.79)y/(0.866) - y/(9.53)y/(0.866)
2 2 2
+ [y/(0.866)] = 0.25 + 0.25 - 0.5 - 2-0.5-0.307 + (0.307) = 0.037

(iv) Empirical method. T h e stress increase below a load area can also be determined
by approximate m e t h o d s . O n e of the most c o m m o n m e t h o d s for determining the stress
increase at various depths below the centre of a load area is the so-called 2:1 method.
According to this method the v o l u m e of stress influence is b o u n d e d by planes inclined
2 (vertical): 1 (horizontal) as shown in Fig. 8 3 . T h e stress distribution in each horizontal
section of the 'stress p y r a m i d ' is assumed to be constant. T h u s , the vertical stress increase
at depth below the centre of a rectangular area with length / and width b under an external
load q - Q/bl is governed by the condition of equilibrium:

7 = (97)
(1 + zll)(\ +zlb)

T h e horizontal stress increase in the y direction can be determined approximately by


the relation:


AOy = (98)
3
(l+z/b)

b +z

Fig. 83. Stress pyramid used for approximate calculation of vertical and horizontal stress increase under
load area.
Spread foundation 111

1 0
y)
y
/A

r
M/
ff
iJb

h
1
II
II
II

Fig. 84. Vertical stress distribution with depth below the centre of square (left) and strip loads. Full line:
2:1 method. Dash line: Rigid footing, theory of elasticity. Dash-dot line: Weak load, theory of elasticity.

A comparison of the stress increase below the centre of a square and a strip footing
according to the empirical method to that obtained according to the theory of elasticity
is m a d e in Figs. 8 4 - 8 5 .
T h e empirical m e t h o d of stress analysis gives results in close a g r e e m e n t with the stress
increase b e l o w the critical point (see p. 142) of a footing according to the theory of elasti-
city, Fig. 86.

2.2 Raft foundation

(i) Contact pressure. In the case of raft foundations, the limited bending rigidity of the
raft will h a v e to b e considered in the analysis of the foundation pressure distribution. It
is also necessary to consider the influence of the deflection of the raft on the internal

Fig. 85. Horizontal stress distribution with depth below the centre of square (left) and strip load areas
according to approximate and strict calculation methods.
118 Spread foundation

Fig. 86. Vertical stress distribution with depth according to theory of elasticity (full lines)below the
characteristic point of a square (left) and a strip footing. Broken lines represents vertical stress
distribution according to the empirical 2:1 method.

m o m e n t s and forces in the superstructure (as is the case with the influence of differential
settlements of individual footings).
M a n y attempts have been m a d e to solve, in a m o r e or less exact way, the pressure
distribution at the soil/raft interface. T h e problem is quite difficult for several reasons:
the pressure distribution is a function not only of the bending rigidity of the raft but
also of the rigidity of the superstructure,
the soil m a y be heterogeneous with varying deformation properties below different
parts of the raft,
the rigidity of the structure is subjected to gradual changes during the time of
construction,
stress changes in the subsoil during the time of construction has a strong influence
on the resulting, final stress distribution.
In respect of all the difficulties involved, we are reduced to finding approximate
solutions to the problem. T h e oldest approach is based on the use of a subgrade coeffi cient
while the m o d e r n approach is based on the utilisation of finite element m e t h o d s and
computer-based solutions.

(i) The subgrade coefficient theory. In the subgrade coefficient approach, it is assumed
that the soil can be replaced by a bed of vertical, elastic springs without horizontal
coupling. In a physical sense, this assumption means that the load does not entail shear
stresses in the soil and that the contact stresses are directly proportional to the vertical
settlements. The contact stress p a t a given point is thus given by the relation:
Spread foundation 119

P = KS (99)

where Ks = the coefficient of subgrade reaction,


= the vertical m o v e m e n t of the point under consideration.
According to this definition, the contact pressure against a raft acted upon by an
equally distributed, constant load will be constant and equal to the applied load
independent of the bending rigidity of the raft which is not in agreement with reality. A
c o m m o n m e t h o d to correct this anomaly is to a s s u m e an increase in the spring stiffness
along the edges of the raft in order to simulate the influence of shear stresses in the soil.
T h e fact that Ks is not simply a material parameter but also dependent on the dimensions
and stiffness of the load area is a serious drawback of the subgrade coefficient theory. T h e
choice of subgrade coefficient must be m a d e with due consideration to this fact.
In spite of all the uncertainties involved in the use of a subgrade coefficient, the theory,
if applied with caution, can yield satisfactory accuracy. In the analysis, the raft is divided
up in strips in s o m e suitable way. These strips can then be treated as b e a m s on a resilient
bed (Vesic, 1971).

(ii) The subsoil is considered as an elastic medium. In this case, solutions to the problem
are obtained by the aid of the finite element m e t h o d and computer-aided calculations. T h e
soil at the foundation level is often divided into finite surface elements which are coupled
to the connecting finite raft elements and to the superstructure. Every raft element is acted
upon by a uniform raft load and the point load exerted by the superstructure. T h e s e loads
are uniformly distributed over the element. T h e elements are first di sconnected from each
other and a settlement analysis is performed. T h e settlement of the element is obtained
by s u m m a t i o n of the vertical deformation of the different layers in the subsoil with due
consideration to the settlement contribution obtained from nearby elements. In the
following step of the analysis, enforced forces and m o m e n t s are applied to the elements
in order to maintain the continuity of the raft in the nodes between the elements. T h e
additional settlements thus obtained are considered in the next step of the analysis, and
so on in an iterative process.
T h e superstructure and the raft and the subsoil can also b e treated as a continuous
system divided into substructures that are analysed b y themselves and then connected
with each other to fulfil the the continuity requirements. T h e soil is then divided into a
n u m b e r of rectangular, trapezoidal or triangular elements in a two-dimensional analysis,
or in prismatic or tetrahedral elements in a three-dimensional analysis. Now, if the stress/
deformation characteristics can b e correctly described, a correct solution to the problem
is theoretically possible. In practice, however, these calculations require a very extensive
computer m e m o r y which is not usually available. In consideration of the difficulties
previously m e n t i o n e d (time effects, plastic yield, etc.), this type of analysis is as yet only
of interest in p a r a m e t e r studies and research.
120 Spread foundation

Fig. 87. Contact pressure distribution caused by point and line loads.

(iii) Approximate evaluation. In m a n y cases the p r o b l e m of stress distribution can b e


solved by approximate m e t h o d s . Beigler (1976) s h o w e d that the contact pressure dis-
tribution under a point and a line load can b e approximated to a cone and a w e d g e ,
respectively, with a base width ry given by the relation (Fig. 87):

r=\3Kr (100)

w h e r e Kr = (,/5),
t = thickness of the raft,
Er = elastic m o d u l u s of the raft,
Es = elastic m o d u l u s of the soil.
T h e peak pressure under a point load Q can be approximated to:

3
Po =
7ir
:z (101)

and under a line load P:


=P/r
Pl (102)

For a Une load acting at distance a from the e d g e of the raft that is below the value of
r (r < bl2), the contact pressure along the edge, according to Beigler (1976), can be
obtained from the relation (Fig. 88a):

r-a
Pe=P\ (103)
a
If the value pe thus obtained exceeds the plastic yield pressure of the soil pcn two outcomes
m a y occur:

P/pcr < 2a
Spread foundation 121

Fig. 88. Contact pressure distribution caused by a line load near the edge of a raft, (a) Edge pressurep e
< pcr (b) Edge pressure pe = pcr and < 2apcr (c) Edge pressure pe = pcr and > 2apcr.

F r o m the conditions of equilibrium, the e d g e pressure is found to b e constant and equal


to pcr up to a distance x0 from the e d g e according to the relation (Fig. 88b):
P-a{pcr + Px)
x0 = (104)
Pcr "Pi
a + 0.5r> Plpcr> 2a
In this case w e find x0 by solving the equation (Fig. 88c):

2 2pcrr 2Pr
(a + r-x0) JC0 = (105)
Pi Pi

T h e critical e d g e pressure pcr is governed b y the external effective pressure at the


foundation level and the strength parameters of the soil c ' a n d '. A s the width of the yield
zone is generally small, pcr can b e a s s u m e d to b e constant.
In the case of evenly distributed load q on the raft, the contact stress distribution can
be considered constant from the centre of the raft u p to a distance r from the e d g e of the
raft w h e r e the pressure can b e a s s u m e d to increase towards the e d g e according to the
relation, Fig. 89 (Beigler, 1976):

4
<r * .-a
4 ' **'.<>. '
- L i : -V, i
>-V III {
11
Px Pm P,
II i l l . .

b/2 - r !
r
1
4^
Fig. 89. Contact pressure distribution under a raft acted upon by evenly distributed load q.
122 Spread foundation

0 6
Px^Pm+iPe-PmK-^)" O )
r-x 0

where = 2.SpCfJq,
pm = q(\-mx)(\+m2),
Pe -Pm + -5qb(n + l)[mx - ra2(l - mx)]/r <pcr,
mx = 0 . 3 6 3 / [ l - 0 . 0 5 f e / r + 0 . 4 ( f e / r ) 3 ] ,
m 2 = 0.826qmx/pcr if x0> 0,
m2 = 0 if x0 < 0,

b{n + \)(q-pm) r
XQ = - - > 0 .
n
2n(Pcr-Pm)

(If r > fe/2 then put r = b/2, and if x 0 < 0 then put JC 0 = 0 ) .

T h e e d g e p r e s s u r e p e (at = r - x0) is limited by the plastic yield pressure pcr.


If the raft is acted u p o n by a c o m b i n a t i o n of line loads, point loads and uniformly
distributed load the contact pressure is obtained by s i m p l e superposition. A case where
the c o m b i n e d action results in an e d g e pressure in excess of the critical ground pressure
can be solved by applying as critical pressure the r e m a i n i n g bearing capacity potential
a b o v e the e d g e pressure for the load case first deduced.

Example 7: Determine the contact pressure distribution beneath a raft foundation, 15 m in width, acted
upon by a line load Px = 100 kN/m, 0.5 m from the outer edge of the raft, and a central line load P2 =
2
200 kN/m. The raft is subjected to a uniformly distributed load of 10 kN/m . The raft, which is of
3
concrete with unit weight 24 kN/m , is founded at 0.5 m depth in sand. It has a thickness of 0. 5 m and
3
a modulus of elasticity Er = 20 GPa. The sand has a unit weight = 18 kN/m , an internal angle of friction
'= 30 and a modulus of elasticity Es = 25 MPa. The groundwater level is 1 m below the ground surface.

Solution'. For a friction angle of 30 the bearing capacity factor Nq = 18.4 (Eq. 112). The critical edge
pressure (which is obtained by the relation pcr = df)Nq, where dj= 0.5 m and = 0.5 m) b e c o m e s p c r =
166 kPa.
Influence of the line load near the edge of the raft:
3 1 /3
We find r = 1.30.5(20 1(> /25) = 6.0 m. Furthermore, the a value is 0.5 m. Consequently, the
contact pressure below the load becomes equal to:
px = 100/6 = 16.7 kPa
and below the edge pe = 16.7(6 - 0.5)/0.5 = 183.7 kPa > pcr
The width of the zone of plastic yield
1 0 0 - 0 . 5 ( 1 6 6 + 16.7)
in = = 0.06 m
166-16.7
Influence of central line load:
We find px = 200/6 = 33.3 kPa
2
Influence of uniform load consisting of surface load 10 k N / m and self-weight of the raft 0.5-24 = 12
2 2
kN/m , i.e. in total 22 kN/m :
We find:
77 = 2.8-166.5/22 = 8.4
Spread foundation 123

3
mx = 0.363/[l - 0 . 0 5 - 1 5 / 6 + 0 . 4 ( 1 5 / 6 ) ] = 0.051
m 2 = 0.826-22-0.051/166 = 0.0056 (see below)
pm = 22(1 - 0.051)-(1 + 0.0056) = 21.0
15-45-(22-21) 6
XQ = - < 0
2-44.(166-21) 8.4
which yields ra2 = 0.
Thus pm = 22(1 - 0.051) = 20.9 kPa
pe = 20.9 + 22-9.4-0.051-15/12 = 34.1 kPa
Considering the effect of the line load near the edge (which caused yield) the uniform load will extend
the width of the zone of yield to a distance from the edge of JC0 = 0.06 + (0.5 - 0.06)-34.1/(166 - 16.7)
= 0.16 m
Superposition yields:
Contact pressure at raft center:
pm = 20.9 + 33.3 - 54 kPa
57
Contact pressure 0.5 m from e d g e p x ( 55 m) = 16.7 + 20.9 + (34.1 - 20.9)-[5.5/(6 - 0 . 1 6 ) ] - 47 kPa
57
Contact pressure 0.3 m from e d g e p x (5 7 m) = 20.9 + (34.1 - 20.9)[5.7/(6 - 0 . 1 6 ) ] + (0.24/0.44)(166
-16.7) + 1 6 . 7 - 1 3 1 kPa
Contact pressure 0.16 m from edge p^o.iom) = Per=166 kPa
The contact pressure distribution across the raft is shown below.

100 KN/M 200KN/M

0.06
0.16 M
-JFTO-i

A case record of contact stress distribution, observed by m e a n s of 2 0 Gltzl pressure


cells installed at the raft/soil interface in two cross-sections of a three-storey nursing
h o m e on lightly overconsolidated clay (preconsolidation pressure > 4 0 kPa higher than
the in-situ effective stress), is presented in Fig. 90.
A n analytical e x a m p l e of the influence on the contact stress distribution of the stiffness
of the superstructure is given in Fig. 9 1 . In this case w e h a v e to deal with a small house
founded on an elastic m e d i u m with =25 M P a and = 0.4. T h e raft has a thickness of
0.1 m, a m o d u l u s of elasticity = 10.5 G P a and = 0.15 and the walls a thickness of 0.15
2
m, = 7.85 G P a and = 0.2. T h e load on the floor is 2.5 k N / m and the load from the
walls 5 kN/m.
124 Spread foundation

Fig.90. Observed contact stress distribution in two cross-sections of a concrete raft foundation, 400 mm
in thickness. Column spacing lengthwise of the buildings 7.2 m. In shelter section, the column load 490
kN (to the left) is replaced by a wall load of 52 kN/m. The internal wall load of the shelter = 42 kN/m.
Building completed 2/7 -74 and in use 17/7 -75.
Spread foundation 125

3. BEARING CAPACITY

3.1 Individual footings

(i) Stress condition on failure surface. Let us a s s u m e that w e have to deal with a two-
dimensional case and that the force vector p, acting on a unit length of the failure surface
at a certain point, is k n o w n as well as the inclination of the failure surface at the point
in question (Fig. 92). T h e position of the point considered is determined by the arc length
s, set out from a fixed point. T h e parameters p, v, s and $ ' a r e positive w h e n they have
the directions shown in Fig. 92. Consider an element A B C D in the failure surface. Draw
the lines D C and B C parallel to A B and A D . D u e to the rotation of the sides B C and D C
the force vectors acting against D C ' and B C will change in size as shown in Fig. 92.
Now, from the equilibrium conditions for soil element A B C D an expression for the
stress vector can b e derived, the so-called Kotier 's equations. Neglecting terms of minor
significance, the equilibrium condition in the s direction yields:

+ 2p tan ' + / sin(v + 0 ' ) = O (107)


ds

Integrating over s, the stress vector is given by the general expression:

p = -e 2 v t a n 0 \_^^' + > 1
( ) 5 )( 1 0 8

w h e r e C is a constant of integration w h o s e value depends on the b o u n d a r y conditions.


For a plane failure surface the angle of inclination vis constant ( v = v 0 ) and the expression
takes the form:

p=/?o-/ssin(vo+0O (109)

Fig. 92. Failure surface, affected by the effective stress vectorp, inclined </>'to the normal of the failure
surface (left) and soil element ABCD in the failure surface. AB C D represents a parallelogram.
126 Spread foundation

where p0 is the stress vector at s = 0.


For a weightless soil ( / = 0), Ktter's equation has the following general solution,
irrespective of the shape of the failure surface:

/
p = jp oe x p [ 2 ( v o- v ) t a n 0 ] (110)

where p0 and v 0 are values at a given point of the failure surface.


A zone failure is characterised by two groups of failure surfaces with an intersection
angle of 9 0 '. T h e intersection angle b e t w e e n the failure surfaces and the major
principal stress is 4 5 - '/2 and between the failure surfaces and the minor principal stress
45+ 072.

(ii) The bearing capacity factors. Consider a strip footing of width b resting on the ground
surface and surrounded by an evenly distributed load q0 (without internal friction). T h e
strip footing is carrying a centrical vertical load Q per length unit. T h e effective unit weight
of the soil is / .
Now, if Q is increased up to soil failure and the soil is weightless, it is obvious that the
effective stress equals the contact pressure. T h e unit weight of the soil has no influence
on the cohesion intercept c\ only on the frictional resistance. Considering only the
influence of the weight of the soil beneath the footing, the effective stresses increase from
a zero value along the edges of the footing to a m a x i m u m value below the centre of the
footing (Fig. 93). In consequence of this reasoning, w e can express the bearing capacity
of the footing as the s u m of three bearing capacity factors: related to the external load
q0, Nc related to the cohesion intercept ( / a n d N r r e l a t e d to the unit weight / o f the soil
beneath the footing. On the basis of the stress distribution at failure, shown in Fig. 9 3 , w e
have:
qf= Qj/b = 0.5bYNY+ N
qo q + c7V c (HI)

T h e m i n i m u m value of the bearing capacity of a footing subjected to a vertical load is

/
qNq+c Nc

assumed
b

Fig. 93. Idealised stress distribution below a footing at failure.


Spread foundation 127

Fig. 94. Rankine and Prandtl failure zones developed below the frictionless base of a strip footing.

obtained by assuming a zone of failure c o m p o s e d of an active and a passive Rankine


failure zone and an intermediate Prandtl failure zone. T h e R a n k i n e zones are characterised
by plane failure surfaces and the Prandtl zone by a group of logarithmic failure surfaces
and another group of plane failure surfaces through the centre of the logarithmic spiral
(Fig. 94). T h e intersection angle between the t w o groups of failure surfaces is 90 '.
Assuming weightless soil, the stress vector in the active Rankine zone becomes:
/ /
P l = (qf+ c cot(/> )tan(45 - 072)

and in the passive R a n k i n e zone:


/ /
p2 = (qQ + c cot0 )tan(45 + 2)

(In these relations, c ' c o t 0 ' c a n be exchanged for the attraction value a).
T h e relation b e t w e e n px and p2 is found from Eq.(109 ):

Pi ~ PiexpC-TCtan^O

Inserting the expressions for p2 and px w e find:

2 / / 2
qf= 4 0 t a n ( 4 5 + 072)exp(7itan0O + c c o t 0 [ t a n ( 4 5 + 072)exp(7rtan0O - 1]

This yields:

2
Nq = t a n ( 4 5 + 072)exp(7rtan0O (112)

Nc = (Nq-l)cot<l>' (113)

In the special case of ' = 0 (i.e. w h e n failure is governed by the undrained shear
strength cu) w e find ^ 0 = 1 and = 2+.
T h e s e solutions for Nq and Nc, originally presented b y Prandtl and Reissner (Fig. 95),
are most c o m m o n l y applied in the analysis of the bearing capacity of strip footings.
While there is a general acceptance regarding Nq and Nc, there is quite a disagreement
128 Spread foundation

about the evaluation of Ny. According to the German code DIN 4017, Ny is expressed
by the relation (Fig. 95):

NY=2{Nq-\)\smQ' (114)

which can b e approximated to (' degrees):

# y= 0 . 0 8 e x p ( 0 . 1 8 f ) (115)

while, for e x a m p l e , according to Meyerhof (1963)

yVr=(^-l)tan(1.40O (116)

T h e bearing capacity factors given above are only relevant for strip footings placed
either on the ground surface or at a depth w h e r e the material above the foundation level
has neither friction nor cohesion. For individual pad footings, for footings buried in the
soil and for certain special loading conditions, the bearing capacity factors will have to

10001 1 1 1 1 1 1 1 1 1 1

500

0 10 20 30 40 SO-
Effective angle of friction '

Fig. 95. The bearing capacity factors Nc and vs. angle of internal friction.
Spread foundation 129

be corrected. This is done by multiplying the bearing capacity factors with coefficients
w h o s e m a g n i t u d e depend on geometry of the footings, depth of foundation, loading
conditions, and so forth.

(iii) Shape and depth coefficients. In practice, w e mostly h a v e to deal with the problem
of determining the bearing capacity of individual pad footings placed at a certain depth
below the ground surface or the b a s e m e n t floor. Failure will then take place in the
direction towards the lowest adjoining ground surface or floor. Consequently, the depth
coefficients dq, dc and dy should b e based on the foundation depth that is critical with
regard to failure. Both the depth coefficients and the shape coefficients sq, sc and sy
r e c o m m e n d e d in the literature are determined on an empirical basis.
T h e shape and depth coefficients r e c o m m e n d e d in the G e r m a n code D I N 4017 are:

sq= 1 + ysin0' (117)

b
s = l-03y
Y (118)

1
s _ W (119)

dq = dc = dY=l (120)

For ' = 0 the shape factor tends to:

b
scQ = l+02- (121)

T h e variation of Nc0 with regard to shape and depth of the footing can b e taken from
Fig. 96.
Results of large-scale plate model tests on sand (Du Thinh, 1984) indicate a better
agreement with the shape and depth factors r e c o m m e n d e d by M e y e r h o f (1961; 1963)
than with those r e c o m m e n d e d in D I N 4017. According to M e y e r h o f w e have:

b 2

s = s = 1 +0.1 y tan (45+ 0 7 2)


q y (122)

s = 1 + 0.2y tan (45 + 7 2 )


c
2
(123)
130 Spread foundation

10
1
b/I=\
9 1
1
/3// = 0.5
8
a? /?// = 0
t-l

0 1 2 3 4
Depth to width ratio dflb

Fig. 96. Bearing capacity Nc0 as a function of shape and depth of foundation.

dq = dy = 1 + 0.1 ^ tan(45 + 0 7 2 ) (124)

4 = 1 + 0 . 2 ^ tan(45 + 0 7 2 ) (125)

M e y e r h o f (1961) points out that the angle of internal friction to b e used in his
correction coefficients should be determined by direct shear tests (plain strain t e s t s ) ' I n
case the angle of friction is determined by triaxial tests, M e y e r h o f suggests that the value
thus obtained should b e replaced by:

^(U-O.lyWiri. (126)

( M e y e r h o f ' s correction of the triaxial value is based on results of true triaxial testing
under plain strain condition. T h u s , the angle of friction d e t e r m i n e d under plain strain
condition has been shown to be around 1 0 % higher than the angle of friction determined
in the conventional active triaxial test).
M e y e r h o f ' s correction coefficients given in Eqs. (122) through (126) cannot be ap-
plied for values of '< 10.
T h e influence of shape and depth of foundation on the bearing capacity of a cohesive
soil in undrained condition can be obtained from the d i a g r a m s h o w n in Fig. 96.
In case the groundwater level is situated below the foundation level, / i s replaced by
/ a b o v e the groundwater level and / b e l o w the groundwater level, from the foundation
level d o w n to a depth below the foundation level equal to the width of the footing.
Spread foundation 131

Example 8: Calculate the bearing capacity according to DIN 4017 on one hand and according to Meyer-
hof on the other for a centrically loaded square footing with width 2.5 m founded at a depth of 1 m below
the ground surface. The soil consists of sand with an internal friction angle ''= 35. The groundwater
3
level is 1 m below the foundation level. The density of the sand is 1.8 t/m above the groundwater level
3
and 1.1 t/m below the groundwater level. The influence of capillary forces and surface tension on
effective stresses can be ignored.

1.0 m

GW 2.5 m 1.0m

Solution: The bearing capacity factors according to DIN are Nq = 33 and Ny = 45 and according to
Meyerhof Nq = 33 and Ny= 37. According to DIN we have sq = 1.574 and sy= 0.7 and dq = dy= 1 while
2
according to Meyerhof sq=sy -1.369 and dq-dy-\ .077. Furthermore, q0 = 18 k N / m and / a v = ( 1 .() 18
3
+1.5-11)72.5 = 13.8 kN/m -
Thus, according to DIN:
2 2
Vf= ft2
q = (0.5-2.5-13.8-45-0.7 + 18-33-1.574)-2.5 = (543 + 935)-2.5 = 9240 kN

and according to Meyerhof:


2 2 2
Vf= qjb = (0.5-2.5-13.8-37 + 18-33)-1.369-1.077-2.5 = (941 + 876)-2.5 = 11355 kN

(iv) Eccentric loading. In reality, the load acting on the footing is often eccentric. T h e
bearing capacity is then calculated on the assumption of a fictive, centrically loaded
footing with r e d u c e d dimensions. If the eccentricity of the load is eb in the direction of
the width of the footing and el in the length direction (Fig. 97), then the fictive footing
is assumed to h a v e the width b - 2eb and the length / - 2el. A footing of arbitrary shape
is replaced by a fictive rectangular footing with the same b o t t o m area and the same
principal axes and length/width relations as the reduced shape of the real footing. T h e
analysis of the reduced, fictive footing is carried out in the s a m e w a y as shown above.

(v) Inclined load resultant. W h e n the load resultant is inclined, the z o n e of failure will
b e shallower than w h e n the load resultant is vertical. In consequence, this loading case
b e c o m e s m o r e critical the higher the ratio of the horizontal c o m p o n e n t H to the vertical
c o m p o n e n t V(Fig. 98). In a case w h e r e the horizontal c o m p o n e n t is achngperpendicular
to the length direction of the footing, the correction coefficients p r o p o s e d by Brinch
Hansen (1967) can be applied:

l-2e,

Fig. 97. Fictive footings to replace real footings under eccentric loading conditions.
132 Spread foundation

OJH 3

= ( 1 ) 1 2 7
^ " v + Wc'cot^ < >

^ - T ^ i ( 1 2 8 )

i = 1 1 2 9
r ( -TT-77^ ( )
' V + blc cot

T h e s e correction coefficients are also r e c o m m e n d e d in the G e r m a n c o d e D I N 4 0 1 7 .


According to M e y e r h o f (1963), the correction coefficients for inclined load can be
written:

i i = c = ( l - f i / 9 0 - ) 2 (130)

=(\-& (131)

w h e r e = arctan(///V)
In the case of '= 0, Brinch H a n s e n (1967) p r o p o s e s the correction coefficient:

-W'-^ <132)

Example 9: Calculate according to DIN 4017 and Meyerhof the bearing capacity of the square footing
in Ex. 8 if the load resultant goes through the centre of the base of the footing and has an inclination of
3(vertical): 1 (horizontal).
Spread foundation 133

| 1m 0
..xf&\
52m
GW U J I 1.0 m

3 3
Solution: According to DIN we have = ( 1 - 1 / 3 ) = 0.296 and iq = ( 1 - 0 . 7 / 3 ) = 0.451, while according
2 2
to Meyerhof we have i y = [1 - arctan(l/3)/35] = 0.224 and iq = [1 - arctan(l/3)/90] = 0.632.
The bearing capacity according to DIN is:
2 2
Rf= qfb //3 = (0.296-543 + 0.451935)2.5 /T 13 = 3834 kN
and, according to Meyerhof:
2 2
Rf= qfb JW/3 =(0.224-941 + 0.632-876)-2.5 /TO 13 = 5036 kN

Example 10: A square footing under a vertical column load of 200 kN is subjected to a horizontal load
/ / k N and a rotational moment M = 2 / / k N m . The footing has a width of 2.5 m and is placed at 1 m depth.
The groundwater level is 0.5 m below the foundation level. Determine according to DIN 4017 the value
of//, leading to foundation failure at pj- 500 kPa, for a soil with an effective friction angle 0 ' = 32 and
3 3
a density above the groundwater level of = 1 . 8 t/m and below the groundwater level of p ' = 1.1 t/m .

= 2 0 0 kN

IH *- iti 3?
1.0m

GW 0.5 m

2.5 m

Solution: The eccentricity of the load is eb = 2///200 = 0 . 0 1 / / which gives a reduced footing width of
b' = 2.5 - 0 . 0 2 / / m. The correction factors become:
2.5-0.02//
^ = 1+ sin*/)'
2.5
2.5-0.02//
7= 1 - 0 . 3
2.5
0.7//
200

7
200
The bearing capacity factors for '- 32 become Nq - 23.2 and = 27.7 and the average unit weight
= = 3
7av Pav 12.2 k N / m . The relation becomes:
500 = 0.5-(2.5 - 0.02//) 1 2 . 2 - 2 7 . 7 - 5 ^ + 1 8 - 2 3 . 2 - ^
which yields / / = 37.5 kN.

In case the footing is acted u p o n by a horizontal force c o m p o n e n t parallel to its length


direction, the correction coefficients iq and / 7 c a n b e simplified to:


(133)
l q l y
~ ~ " V + Wc'cot0 '
134 Spread foundation

Fig. 99. Zone of failure for footing in immediate vicinity of slope.

(vi) Footings next to slope. For a footing that is placed near a slope with inclination
to the horizontal, Brinch Hansen (1967) suggested that the bearing capacity factors in the
case of a cohesionless soil be corrected by the coefficients:

2
(1 - s i n < / ) ' ) c o s ^
= e x p K
Si Zr = ,o *>s - 0 - 2 + 20) tan 0 '] (134)
1 - sin sin(2u + ) 2

w h e r e u represents the angle of inclination of the failure surface to the down-slope (Fig.
99) which by conditions of equilibrium (cf. p. 2 7 3 - 2 7 4 ) is obtained from the relation:

2u - arccos(sin/sin</0 + - ' (135)

These correction coefficients can be approximated to:

g ? = g y= ( l - 0 . 7 t a n j 3 ) 3 ;i36)

T h e approximate expression generally yields results on the safe side but when the
angle of inclination of the slope tends to the angle of internal friction the difference in
results can be up to 4 0 % on the unsafe side.
In the case of an unloaded slope in cohesive soil with 0, the correction coefficient
can b e written ( in radians):

2
ScO =
1
- (137)
2+

Example 11: Determine according to Brinch Hansen how much the bearing capacity of the square
footing in Ex. 8 will be reduced if the footing is placed near aslope with inclination 1 (vertical):4(horizontal).


5
Spread foundation

Solution: We have = arctan(0.25) = 14 and 2u = arccos(sinl4/sin35) + 1 4 - 35 = 44


Thus:
= ( l - s i n 3 5 < ) c o s 2 14o
6 q 76 F
l-sin35sin79 180
which yields: gq = gY = 0.57
3
The approximate relation gives: g = gy = (1 - 0.7tanl4) = 0.56
We find that the bearing capacity is reduced to approximately 60% of the bearing capacity obtained
when the ground surface is horizontal.

(vii) Inclined base of the footing. In the case of weightless soil, the stresses along the
failure surface change in accordance with the exponential function of 2 t a n $ ' times the
angular change v. Consequently, the correction factor bq is given by:

bq = e x p ( - 2 v t a n 0 O (138)

The weight of the soil can be considered by the correction factor (Brinch Hansen, 1967):

by- exp(-2.7vtan0O (139)

By comparison with the earth pressure against a vertical, smooth retaining wall, Brinch
Hansen proposed that the correction factor in the case of inclined loading be approximated
to (notations given in Fig. 100):


U =[1-(1- ) (140)
300 + blc'cot0'

In the case of cohesive soil with ' - 0, w e h a v e (v in radians):

2v
bc0 = 1 (141)
2+

(viii) The pressuremeter method. A m o r e reliable estimation of the bearing capacity of


footings can be obtained by in-situ determination of the strength characteristics, for
instance, on the basis of the limit pressure pt determined in the pressuremeter test (see p.

Fig. 100. Zone of failure under footing with inclined base.


136 Spread foundation

80). M e t h o d s for calculation of the bearing capacity h a v e been w o r k e d out by M n a r d


(see Baguelin et , 1978). T h e bearing capacity can b e determined by the relation
r e c o m m e n d e d in the Canadian M a n u a l on Foundation Engingeering (1985):

,
/= 0 + / (142)

w h e r e k = ks[I^ + (& S q u a re kstIl^)b/l,


' 0 = effective overburden pressure at the foundation level,
= = n et
* Pi~Po liroit pressure,
p0 = horizontal in situ stress at the foundation level.
T h e k values can be obtained from the diagram, shown in Fig. 1 0 1 .
In i n h o m o g e n e o u s soil conditions, the variations in p{ should be taken into account
within a layer extending from 1.5/? above, to \5b below, the foundation level. T h e {
values (n in n u m b e r ) determined within this layer are transformed to:

/>, = 6 MPA

2 MPA

1 MPA

0.5 MPA
-pt = 6 MPA

2 MPA

0 . 4 MPA

0 1 2 3

'Pi = 3 MPA
1 MPA
0 . 5 MPA

pt = 3 MPA
1 MPA
0 . 5 MPA

Ratio of equivalent depth of foundation to width of footing dfJb

Fig. 101. & values recommended for silt, sand and gravel
Spread foundation

Un
Pie = ('Pn'Pi2-'PB-'PiJ (143)

In this case the real depth of foundation d^is also replaced by an equivalent depth of
f o u n d a t i o n ^ determined by the relation:

dfe = ll(AzrPil/yie) (144)

where ; ; = dj.
T h e pressuremeter method can also b e used to determine the so-called creep load, i.e.
the load that leads to excessive creep deformations. This is done by using the creep
pressure pcr in the bearing capacity formula instead of the limit pressure/?/ (for the definition
of the creep and limit pressures, see p . 81).
For eccentric loading, a fictive footing with reduced length and width according to the
principles shown in Fig. 97 is used, i.e. with length / - 2e{ and width b - 2eh w h e r e e{ and
eb are the load eccentricities in the length and width directions, respectively.
For inclined loading, the reduction factor il applied to the bearing capacity is:

w h e r e 0 = arctan(///V),
-
Xd= \-d/bfoi'0<df/b<\,
Xd = 0fovd/b>\,
Xm - 1 - m for 0 < m < 1,
Xm = 0 for m > 1,
m = ( at depth df)/(*pl at depth dj+ b).
If / b e c o m e s negative the footing must b e e m b e d d e d deeper.

Fig. 102. Footing near a slope.


138 Spread foundation

For footings near a slope, the reduction factor gs is obtained from Eq. (145) with re-
placement of 0 by the angle ', defined according to Fig. 102.
For concentric inclined loading near a slope w h e r e the inclined load is directed towards
the slope, the reduction factor is is calculated by using Eq. (145) after replacing by + '.
If the inclined load is directed away from the slope, failure m a y take place either towards
the slope (replace 0 b y 0 + ') or away from the slope (keep 0). T h e t w o cases will have
to be considered separately.
For eccentric inclined loading, the reduction factor depends on whether the load is
inclined in the direction away from the centre of the footing or towards the centre of the
footing. In the former case, the footing is assumed to h a v e the width b + 2eb while, in the
latter, the footing is assumed to h a v e the width b - 2eb.
T h e pressuremeter m e t h o d is especially suitable for footings on sand, gravel and till,
where undisturbed sampling for laboratory testing is m o r e or less impossible. It is also
a useful tool for the determination of the bearing capacity of foundations on clays and
silts.

Example 12: Determine the bearing capacity of a 2 m wide strip footing acted upon by an inclined load
with V/H = 5 and founded at 1 m depth, 0.5 m from a slope 1:3 (1 vertical to 3 horizontal). The load is
inclined in the direction towards the slope. The pressure limit has been determined to 1.0 MPa at 1
%
m depth and to 1.4 MPa at 3 m depth. The equivalent net limit pressure is ple = 1.2 MPa.

Solution: The bearing capacity of the footing under a vertical centric load is obtained from:
Pj~~ 1.4*1.2 1.68 MPa
We have:
= 1 - 1 / 2 = 0.5
= 1 - 1.0/1.4 = 0.286
= 0 . 5 0 . 2 8 6 = 0.143
The angle = arctan(0.2) = 11.3 and the angle '= arctan[2/(l+3-3)] = 11.3 whence + '= 22.6.
This yields:
2
is = (1 - 22.6/90) (l - 0.143) + (1 - 22.6/20)*0.143 = 0.46
Thus the reduced bearing capacity is Vj-= 0.46*1.68*2 = 1.5 MN/m

Example 13: Calculate the bearing capacity of a square footing with 2.5 m by 2.5 m base area, placed
at 1.4 m depth 2.5 m away from a slope with inclination 1 (vertical):3(horizontal) in a soil with pt = 1.5
MPa. A horizontal load, equal to one third of the vertical load, is applied 0.9 m above the base of the
footing in the direction towards the slope.
Spread foundation 139

Solution: The inclination of the resultant R is = arctan(l/3) = 18.4 and the eccentricity eb = 0.9/3 =
0.3 m. The reduced width of the footing is 2.5 - 2-0.3 = 1.9 m. This yields '= arctanfl .9/(2.5+3.3- 1.5)]
= 14.3 from which + '= 32.7. Further we have m=\J.e. Xm = 0 and, consequently, = 0. The reduc-
tion factor becomes:
2
i 5 = ( l - 3 2 . 7 / 9 0 ) = 0.40
The bearing capacity factor for d/b = 1.4/1.9 = 0.74 and bll =1.9/2.5 = 0.76 is k = 1.9. The reduced
2
base area is 1.9-2.5 = 4.75 m .
The reduced bearing capacity of the footing is thus 0.4-1.9-1.5-4.75 = 5.3 M N

3.2 Raft foundations

T h e bearing capacity of raft foundations is seldom a problem. D u e to the size of a raft,


a possible failure z o n e involving the raft as a w h o l e will extend to great depth. Such
failures h a v e occurred in s o m e exceptional cases of very heavy structures founded on
deep clay layers but usually only local failure will take place. T h u s , as shown in the
section on contact stress distribution, plastic yield in the soil m a y occur along the edge
of a raft but this can b e easily handled b y increasing the b e n d i n g rigidity of the raft.
T h e danger of exceeding, in the global sense, the bearing capacity of a raft can be
e x a m i n e d in the s a m e way as was shown for individual footings. T h u s , the bearing
capacity formulae are identical and it is only a matter of differences in size parameters.

4. SETTLEMENTS

4.1 Introductory remarks

Settlement calculations are generally based on the presumption that plastic yield in the
soil is not taking place or has a negligible effect on the settlement. T h e reliability of
conventional settlement analyses decreases successively the nearer the load is to failure.
Reliable settlement calculations can, of course, also b e carried out in such cases, for
instance b y the u s e of data technique, but these mostly require very a d v a n c e d computer
capacities and are, therefore, mainly utilised in connection with research or for parameter
studies. Before conventional settlement calculations are carried out it is certainly
necessary to m a k e sure that the factor of safety against failure is satisfactory.
The course of settlement is very m u c h dependent on the hy drauli c conductivity and the
creep properties of the soil. Settlement will not terminate until the soil skeleton has
140 Spread foundation

readjusted itself and b e c o m e strong enough to carry the effective stress increase induced
in the soil by the load of the footing. In water saturated soils with low permeability, such
as clays, the load applied is originally, wholly or partially, e a r n e d by excess pore water
pressure. T h e load is then gradually transferred to the soil skeleton with a corresponding,
simultaneous excess pore water pressure dissipation. For footings on clay, the pore
pressure m a y certainly dissipate rather quickly due to three-dimensional or two-
dimensional consolidation. However, spread loading of the ground in the vicinity of the
foundation, for e x a m p l e by placement of fill, m a y cause a very drawn-out settlement
process. Even after complete dissipation of excess pore pressure dissipation, long-term
settlement can be expected due to creep p h e n o m e n a .
In coarse-grained material pore pressure dissipation takes place simultaneously with
the application of the load. In coarse-grained soil, settlement d u e to creep often
amounts to 3 0 - 4 0 % of the 'initial' settlement.
T h e settlement of footings on rock are governed by open fissures and other weaknesses
which lead to inelastic behaviour. W h e n the fissures have b e c o m e closed by the ground
pressure the rock behaves more or less as an elastic m e d i u m (Fig. 103). T h e bearing
capacity of footings on rock can be estimated from pressuremeter tests or from the
strength characteristics presented in Table 15.
T h e r e are two main causes of settlement that have to b e considered: ( 1 ) settlement due
to isotropic stress changes (mainly leading to a change in v o l u m e and, consequently, to
pore water being squeezed out of the soil) and (2) settlement due to deviatoric stress
changes (mainly leading to a change in shape). T h e volume change is governed by the
bulk m o d u l u s and the change in shape by the shear m o d u l u s G of the soil. As regards
settlement of footings, settlements due to deviatoric stress changes dominate over
settlements due to isotropic stress changes.

Average contact pressure, MPa


- 0 10 20 30 40 50 60
D = 1.0 m
D = 0.5 m
D = 0.7 m
2

Fig. 103. Results of plate loading test on fissured gneiss rock of bad quality (After Pusch et <-//., 1974)
Spread foundation 141

T h e use of deformation moduli determined by in situ m e a s u r e m e n t s is preferable to the


use of deformation moduli determined on a laboratory scale. T h e only exception to this
general rule concerns settlements caused by primary and secondary consolidation of clay
at stress increments in excess of the preconsolidation pressure. In this case, settlement
calculations are generally based on the oedometer m o d u l u s in combination with the
coefficient of consolidation and the permeability. Foundation on footings h o w e v e r is
seldom chosen in cases w h e r e consequential long-term consolidation settlements can be
expected.

4.2 Settlement of footings

if) Analysis based on theory of elasticity. As the bending rigidity of a footing is very high
in comparison with the rigidity of the soil, the settlement of the footing can b e considered
equal at all points. According to the theory of elasticity, there are certain so-called
characteristic points at which the surface settlement of an elastic half-space is independent
of the bending rigidity of the footing. In order to find the settlement of a rigid footing, w e
then h a v e to calculate the settlement at the characteristic point. According to Steinbrenner
(1936) this can b e done by superposition of the settlements obtained b e l o w corner I of
the four part rectangles E A F I y F B G I , G C H I and H D E I , corner I being situated at the
characteristic point of load area A B C D (Fig. 104).
Now, the settlement of a corner sc of a rectangular or square area with evenly di stributed
load q, acting on the surface of an elastic m e d i u m with limited thickness d, can be calculated
by the relation (Steinbrenner, 1936):

2 2
sjb = q[(\ - v )fx + (1 - v-2v )f2]/E (146)
where

/ | n [(i + AJVq + c J- 2 ] + ,n[ ( * + y q ) / Q ,


/l =
b (l/b)(l + yjCi + Q - 1 ) llb + yjQ+Cd-\

1 d lib
f2 --arctan [ . =
2nb (dlb)yj Q+Q-X
2 2
Q = l + ( | ) a n d Q = l + A -
b b

A s s u m i n g that the characteristic point is independent of the thickness of the elastic


m e d i u m (an assumption that can be accepted with regard to the required accuracy of
settlement calculations), the settlement of the footing can b e calculated as indicated
above by superposition of the corner settlements sc of the given part areas, as derived from
Eq. (145). Accordingly, the settlement of the footing will be dependent on Poisson's ratio.
For calculation of settlements of a footing on cohesionless soil, w e can a s s u m e v- 1/3.
142 Spread foundation

0.42D

B F

Fig. 104. The position of the characteristic point. The settlement at the characteristic point I is obtained
by superposition of the corner settlement of the rectangles GCHI, HDEI, EAFI and FBGI.

For calculation of the initial settlement of a footing on water saturated clay, which is due
only to shear deformations, w e have to a s s u m e = 0.5.
T h e settlements obtained by superposition of the sc values according to Eq. (146 ) can
be expressed by the relation:

s = reqb/E (147)

0.1

- \ -
0.2

0.5

1
f

< - -
>
I
2

PC
\ \ \ \
\\ \ N

10
- \\y -

20
- , V \
0.5 1 1.5 2 2.5

Settlement coefficient re for = 0

Fig. 105a. Settlement coefficient Fe for = 0 (relevant for - M).


Spread foundation 143

w h e r e q = q - q0 = net load increase at the foundation level,


b = width of the footing,
Fe = settlement coefficient according to Figs. 105 a - c ,
= m o d u l u s of elasticity of the soil.
In case the subsoil is built up of layers with different moduli of elasticity, the settlement
can be obtained from the relation:

s/b = ql (r -r _ )/E
i e9i EJ L I
(148)

w h e r e Fei and Fej_x represent the settlement coefficients according to Figs. 105 a - c for
the upper (depth d{_{) and lower (depth d ) t boundaries of the respective layers.

Example 14: Determine the settlement at midpoint of the long side of a rectangular area with length 10
2
m and width 2.5 m acted upon by a uniformly distributed, flexible load q - 20 k N / m . The soil consists
of sand underlain by bedrock at 4 m depth. The sand is assumed to behave as an elastic medium with
Es = 20 MPa and = 0.3.

Solution: The area is cut into two equal halves, each one with length 5m and width 2.5 m. For these we
have C/ = 5 and Cd = 3.56. Thus:

( 1 + ) 2 (2 + : , ) l i
/ , = l { 2 . 1 n t ^ ^ ^ ] + ln[ / - ^ : ] } =( . 2 2 9
2(1 -h V5 + 3 . 5 6 - 1 ) 2 + V5 + 3 . 5 6 - 1

0.1
-
0.2

0.5
-

-
2

\\\
\ \\\YV

10
W \

20' 1 \ \ 1
0.5 1 1.5

Settlement coefficient T0 for = 0.5

Fig. 105b. Settlement coefficient Fe for = 0.5 (relevant for incompressible soil)
144 Spread foundation

u.i

Fig. 105c. Settlement coefficient Te for = 0.3

2
/ 2 = 1.6 arctan[ ] = 0.109
2
1.6 V5 + 3 . 5 6 - 1
The settlement becomes equal to:
2 2 3
j = 2-2.5-20[(l - 0 . 3 ) 0 . 2 2 9 + (1 - 0.3 - 20.3 )0.109]/20000 = 1.3-l()- m

Example 15: Determine the settlement of a square footing, 2.5 m by 2.5 m, founded at 1 m depth in sand
3
and subjected to a column load of 700 kN. The sand is assumed to have a unit weight of 18 kN/m , a
modulus of elasticity of 20 MPa to a depth of 5 m and below that 30 MPa.

Solution: The settlement analysis is carried out in two steps. For the layer with Es = 20MPa we have
dej-i - 0 and dei = 4 m from which we find Te -{_x = 0 and Te t = 0.52. For the underlying sand we have
fe%i_x = 0 . 5 2 and = 0.77.
The settlement, solved by Eq. (148), becomes equal to:
0 7 7 5 2
, ^ - 1 1 8 ) 2 . 5 ( - = = , + - ) = 0.008
2 3 3
2.5 20 10 30 10

(ii) The pressuremeter method. A s p r e v i o u s l y s h o w n , the shear m o d u l u s of soil can b e


d e t e r m i n e d in situ b y m e a n s of the p r e s s u r e m e t e r test. F r o m the p o i n t of v i e w of
settlement analysis b a s e d on the results of p r e s s u r e m e t e r tests, a c o n s i d e r a b l e a m o u n t of
practical e x p e r i e n c e exists in various soil conditions. T h e p r e d o m i n a n t e x p e r i m e n t a l and
practical e v i d e n c e of h o w to interpret the results of p r e s s u r e m e t e r tests in t e r m s of
Spread foundation 145

* D
-

Fig. 106. Assumed stress distribution around a rigid half-sphere embedded in soil.

settlement is related to the Mnard pressuremeter. U p to now, the self-boring pressuremeter


has been used mainly for research.
In the settlement analysis according to Mnard, two contributions to the total
settlement of a circular footing are considered, sd and st\
sd caused by elastic shear deformations (mainly within the zone w h e r e deviatoric
stresses induced by the load are dominating),
s caused b y a decrease in soil v o l u m e (mainly within the zone w h e r e isotropic stresses
induced by the load are dominating).
T h e contribution to s of sd is calculated on the assumption that the soil below the
footing, enclosed in a half-sphere with the s a m e diameter as the footing, is rigid. As was
shown by D e Josselin de Jong (1957), the settlement of a rigid half-sphere with diameter
D, subjected to a load q per unit area, is equal to:

sd = H (149)
12G

w h e r e G = shear m o d u l u s of the soil.


T h e contribution to s of s, is a s s u m e d to b e caused b y v o l u m e decrease of the half-
sphere. T h e stress distribution around the half-sphere is a s s u m e d to fulfil the relation
2
= qsin co (Fig. 106). B y integration over the surface area of the half-sphere w e find an
average value of 2q/3. A s s u m i n g further that 2q/3 is an isotropic stress acting against
the half-sphere w e find the vertical deformation (the settlement s,-) equal to:

qD
s , = - (150)
where = b u l k m o d u l u s of the soil.
Introducing a rheological coefficient - EIK and assuming that Poisson's ratio v = 1/3,
w e find the total settlement:
146 Spread foundation

TABLE 21
The value of = El to be applied in different soil and rock conditions.

Soil type Clay Silt Sand Gravel


EpSPi Epr/'Pl EprfPl Eprhpt

Heavily over- >16 1 >14 2/3 >12 1/2 >1() 1/3


consolidated
Normally 9 - 16 2/3 8 - 14 1/2 7 - 12 1/3 6-10 1/4
consolidated
Weathered and/or 7-9 1/2 1/2 1/3 1/4
remoulded

Rock type Extremely Other Slightly fractured or


fractured extremely weathered
cc= 1/3 = 1/2 =2/3

TABLE 22
Shape coefficients A.

lib 1 2 3 5 20
circle square

(1 1 1.12 1.53 1.78 2.14 2.65


A, 1 1.10 1.20 1.30 1.40 1.50

1 1 D
s = qD(
H
+ ) = ( 2 + ) (151) J K
12G 9K 9E

For an elastic m e d i u m with = 1/3 w e h a v e E = K, w h e n c e = 1. By choosing other


values of a , consideration can be paid to the special behaviour of different soils.
Replacing the m o d u l u s of elasticity by the pressuremeter m o d u l u s Epr and introducing
correction coefficients A with regard to shape of the footing, the following half-empirical
settlement relation is obtained:

y bt
^=^(1.2-0.2-f)[-^(-^)^ + - ^ - l
prd b t 0 pri
(152)

w h e r e q = net load increase at the foundation level,


dj = depth of foundation,
b - width of the footing (or diameter if circular),
b0 = 0.6 m i s a reference width (diameter) of the footing,
indices d and / refer to the 'deviatoric' and 'isotropic t e r m s ' ,
7
Spread founda tion

d/b<l.
If b < 0.6 m (which is rarely the case in practice), then b0 in Eq. (152) should be put
equal to b.
T h e rheological coefficient to be applied in different soil conditions is given in Table
21 and the shape coefficient in Table 22.
T h e practical u s e of Eq. (152) for settlement analysis described in the following is in
agreement with the routine method presented by Baguelin et al. (1978).
For the determination of the pressuremeter moduli to b e applied in the deviatoric and
the isotropic terms, the subsoil is divided into five layers as s h o w n in Fig. 107 (Baguelin
etaL, 1978).
For each one of these layers, the h a r m o n i c m e a n of the m e a s u r e d pressuremeter moduli
is determined. T h u s , if the pressuremeter values measured in a layer are Eh E2, E3,-En,
the h a r m o n i c m e a n EnrY is obtained from the relation:

Eprx = HM(El/E2/E3I- /) = (153)

Now, the harmonic mean of the pressuremeter moduli in layer 1 Eprl is taken as the

(I) 1.5/7

_0.5fc
(ID 1.5/7

(3) 1.5/7

(4) 1.5/?

(5) 4/7

Fig. 107. Division of subsoil into layers for determination of pressuremeter moduli to be applied in
deviatoric and isotropic terms of Eq. (152).
148 Spread foundation

pressuremeter modulus to be used in the isotropic term Epri. The pressuremeter modulus
to b e used in the deviatoric term Eprd is obtained from the relation:

-L =V ! _ + - L + _ ! ! _ , ( | 5 4 )

Eptd 4 Epri 0.85Epr2 Epr3 2.5Epr4 2.5Epr5

If there is a layer of thickness with considerably lower pressuremeter m o d u l u s EprZ


than in the surrounding soil, the settlement is first calculated on the assumption that the
layer has a fictitious pressuremeter m o d u l u s Epnn equal to the m e a n m o d u l u s of the
suiTounding soil. T h e n an additional settlement is added to the previously calculated
settlement according to the relation:

1 1
As = cczZqz(--) (155)

w h e r e ocz = the rheological coefficient of the layer.


T h e pressure increase qz in the centre of this layer is calculated on the basis of the
theory of elasticity or by using the 2:1 method (p. 116).
With regard to the fact that the pressuremeter is a short-term test it cannot be used for
determination of consolidation settlements in cohesive, l o w - p e r m e a b l e soils. However,
according to Mnard, Eq. (152) takes into account the influence on settlement of 10 years
of creep deformations.

Example 16: A square footing, 2.7 m in width and founded on sand 1 m below floor level, shall be
designed for a column lod of 3 MN. Pressuremeter investigations have given the results shown below.
Determine the ultimate load and the settlement for the design load.

PRESSUREMETER MODULUS Epr, MPA LIMIT PRESSURE/;,, MPA


0 10 20 30 0 2 4
Spread foundation 49

Solution: The ultimate load is governed by the equivalent pressure limit and the bearing capacity factor
k. In our case there is no pressuremeter value above the foundation level. The equivalent value of pj is
thus a function only of the values obtained to a depth of 1.5/?, including four pressuremeter values. We
1 /4
find *ple = (2.30.80.80.6) = 0.97 MPa. The equivalent depth in this case becomes equal to the
foundation depth, 1 m. We find the ratio djb = 1/2.7 = 0.37. From Fig. 101 the bearing capacity factor
k can be estimated at 1.2.
2 3
The ultimate load Qf = ( 18 1 + 1.2970)2.7 = 8.6 10 kN
For the determination of the settlement of the footing we have to calculate the harmonic means of the
pressuremeter moduli in the different layers 1 -4. These become E { = HM( 16.5/5.2) = 7.9 MPa; E 2 = 3.8
MPa; E 3 = HM(4.3/7.3/13.8) = 6.8 MPa and E 4 = HM( 16.2/12.8) = 14.3 MPa. The value of E 5 is assumed
equal to E 4 . We have also to consider the loose layer at depths 7 and 8 m with z = HM(4.0/2.1) =
2.8 MPa.
The moduli to be introduced in the deviatoric and isotropic settlement terms are equal to:
1
Entd = f-( + + + = 5.5 MPa
^ 4 7.9 0 . 8 5 - 3 . 8 6.8 14.3
E p n = E { = 7.9 MPa
For determination of the additional settlement due to the loose layer, E p r m is chosen as the mean value
of 14.3 and 13.8 MPa, i.e. E p n n = 14.0 MPa.
The shape coefficients for a square foundation are Xd = 1.12 and Xt - 1.1. In our case, the ratio EptJp/ is
in the range of 4 to 10. The rheological coefficients can thus be estimated at 1/3. The factor ( 1.2 - 0 2d jib)
= 1.126
The settlement is obtained from the relation:
H ^ ^ 2 3 - 1.5 _L_J_ 0 0 30 m
2 2
9-2.7 5.5 0.6 3-7.9 3(2.7 + 7 . 5 ) 2.8 14.0

(iii) Empirical methods. A large n u m b e r of empirical methods have been developed by


which the settlement of footings can be be roughly estimated. Most of these m e t h o d s are
connected with S P T t h e standard penetration testor C P T t h e c o n e penetration test.
The interpretation of the penetration resistance in terms of settlement is very m u c h
dependent on the loading history of the soil. Thus, experiments have shown that the ratio
of the m o d u l u s of pseudoelasticity to the penetration resistance of C P T is m a n y times
higher ( 5 - 1 0 times) in overconsolidated, than in normally consolidated sand (Jamiol-
kowski t al, 1985). Therefore, if the relations between penetration resistance and set-
tlement are not connected with the preconsolidation pressure, the results can be strongly
misleading. In fact, almost all existing empirical relations do not take into account the
influence of possible overconsolidation. Nevertheless, empirical relations can be useful
for a rough, and generally conservative, estimate of the settlements to be expected.
On the basis of the results of 48 case records, Schultze and Sherif (1973) developed
an empirical correlation between penetration resistance of SPT and settlement at the
characteristic point of a load area with length / and width b. T h e settlement is obtained
by the relation:

f
s=q (156)
(N^f^il + OAdf/b)
150 Spread foundation

Settlement parameter/, m/MPa

Fig. 108. Settlement parameter/for various length to width ratios lib of the footing.

w h e r e / = settlement parameter according to Fig. 108,


dj-= depth of foundation,
N30 = blow count for 30 c m of penetration.
T h e Schultze-Sherif relation presumes that the thickness of the soil, characterised by
the given penetration resistance, is large in comparison with the width of the footing. If
this is not the case, it may be preferable to choose empirical relations between results of
S P T and the m o d u l u s of pseudoelasticity. Such correlations h a v e been suggested by, for
example, D ' A p p o l o n i a et (1970) and Parry (1971).
D ' A p p o l o n i a et differ between normally consolidated and overconsolidated soil
conditions. T h e following correlations are proposed:
for normally consolidated soil

E= 19.6 + 0 . 7 9 ^ 3 0 M P a (157)

for overconsolidated soil

= 4 1 . 6 + 1.09 N 3 0 M P a (158)

Parry proposes the simple correlation:

= 5 N 3 0M P a (159)

Example 18: Determine the settlement of a rectangular footing with a bottom area of 2 m by 3 m
subjected to a column load of 3 MN if the footing is founded at 1.5 m depth in sand with an average
3
standard penetration resistance of 7V 30 = 35.The density of the sand is 1.8 t/m . The groundwater level
is below the foundation depth.

Solution: The settlement will be evaluated on the basis of the empirical relations given above.
1
Spread foundation

Schultze-Sherif: We have dj/b = 0.75 and lib = 1.5. This y i e l d s / 0.57 m/MPa. The net foundation
pressure is q = 0.5 - 1.5-0.018 = 0.47 MPa. The settlement is obtained as:
0: 57
' = 0.47 8 7 = 0.009 m
35 (1 +0.4-0.75)

D,Appolonia et al. : Assuming that the soil is normally consolidated we have = 19.6 + 0.79-35 =
47.3 MPa. The Te value (Fig. 105c), taking into account a soil layer with a thickness of 8 m (= 4b) is
0.77. The settlement is obtained as:
s = 0.47-2-0.77/47.3 = 0.015 m
Assuming the soil to be overconsolidated we Find = 41.6 + 1.09-35 = 79.7 MPa which yields:
s = 0.47-2.0.77/79.7 = 0.009 m
Parry: The modulus of elasticity = 5-35 = 175 MPa. The settlement becomes:
s = 0.47-2-0.77/175 = 0.004 m

Probably the m o s t w e l l - k n o w n correlation between results of CPT and settlement is


the one p r o p o s e d by S c h m e r t m a n n et al (1978). For calculation of the settlement,
S c h m e r t m a n n et al. suggest the use of a triangular influence function Iz according to Fig.
109 with a p e a k value at a depth varying from 0.5b for a square footing to b for a strip
footing. In the settlement calculation, the soil is divided into layers with thickness Az in
which the cone resistance can b e considered m o r e or less constant. For each o n e of these
layers the area (IzIMz)Az is calculated. S u m m a t i o n of the part areas thus obtained yields
the settlement s according to the correlation:

s = ClC2qZ-^-Az (160)
Mz

w h e r e q = net load increase at the foundation level


Mz = settlement m o d u l u s varying from 2.5qcz for a square footing to 3.5qcz for a
strip footing (for heavily overconsolidated soil, these values should m o s t probably be
replaced by about \0qc and I3qc, respectively).
qcz = cone resistance at depth
Cx = \-0.5(qlq-\)
C 2 = 1 + 0.21og(10i)
t = loading time in years
T h e p e a k value of Iz is obtained from the relation:

(161)

where ' = vertical stress increase at depth = 0.5b for a square footing and =b
for a strip footing
152 Spread foundation

0 0.2 0.4 0.6 0.8 1.0

0.5b

b
y
y
y
y
y
2b y
y
y
y
y
y
y
y
y
y'
4b

Fig. 109. Influence function Iz for square and strip footings. Function Iz for a strip footing can be used when
/ > 20b. When b <l< 20b the influence value can be obtained by interpolation between Iz for square
and strip footings.

3
Example 19: A static cone penetration test in silty sand with a unit weight of 19 k N / m has given the
result shown below. The groundwater level is at 4 m depth. Determine by Schmertmann's method the
10 years settlement of a strip footing, 0.8 m in width, subjected to a line load = 130 kN/m. The footing
is founded at a depth of 1.0 m.

1.0 m

3.2 m

Solution: The average cone resistance qc varies from 1.2 MPa at the foundation level to 1.5 MPa at a
depth of 3.2 m (4b) below the foundation level. Thus, the settlement modulus Mean be assumed to vary
from 3.5-0.12 = 4.20 MPa to 3.5-0.15 = 5.25 MPa. At 0.8 m below the foundation level we have M =
4.20 + (0.8/3.2)(5.25 - 4 . 2 0 ) = 4.46 MPa. Furthermore, we have ' - ( 1.0 + 0.8) 19 = 34 kPa at a depth
Spread foundation 153

of 0.8 m ) below the foundtaion level and q = 130/0.8 - 1.0 19 = 143.5 kPa. These values of & z and
1 /2
q yield7 Z m ax = 0.5 + 0 . 1 ( 1 4 3 . 5 / 3 4 ) = 0.705.
The parameter CX = \ - 0.5(162.5/143.5) = 0.434 and the parameter C 2 = 1 + log(10-10) = 3.0.
Dividing the subsoil into 0.4 m thick layers, we find for the respective layers - 4.21, 4.40, 4.53,
4 . 6 6 , 4 . 7 9 , 4 . 9 2 , 5.05 and 5.18 MPa and the corresponding Iz values 0 . 3 2 6 , 0 . 5 7 9 , 0 . 6 4 6 , 0 . 5 2 9 , 0 . 4 1 1 ,
0.294, 0.176 and 0.059. The settlement of the footing becomes:
s = 0.4343.00.14350.4[(0.326/4.27 + 0.579/4.40 + 0.646/4.53 + 0.529/4.66 + 0.411/4.79 +
0.294/4.92 + 0.176/5.05 + 0.059/5.18) = 0.05 m.
The settlement can also be calculated in an approximate way as:
s = 0.434-3.0-0.1435[(0.2/4.2 + 0.705/4.46)0.8/2 + (0.705/4.46)-(3.2 - 0.8)/2] = 0.05 m

4.3 Settlement of rafts

T h e settlement analysis in the case of raft foundations is an intricate matter and is very
much dependent on the location of columns and load-carrying walls and the stiffness of
the raft and superstructure. Therefore, settlement calculations are generally carried out
by the aid of computers (see Bowles, 1988).
Mostly, calculations can be performed in a fairly simple manner. T h u s , for raft
foundations on soft, compressible soil (such as normally consolidated, or lightly
overconsolidated clay) w h e r e settlement is mainly a matter of primary and secondary
consolidation, the overall average settlement can b e calculated in the s a m e analytical way
as for individual footings, for e x a m p l e by using M values determined by the oedometer
test and the 2:1 m e t h o d for calculation of the stress increase in the subsoil. T h e settlement
distribution underneath the building will then have to be estimated with due consideration
to the gradual build-up of the rigidity of the raft and superstructure during construction
of the building.

Example 20: A building with a bottom area of 15 m by 20 m is founded at 2m depth in normally


consolidated clay using a raft foundation with high bending rigidity. The load of the building is 50 kN/
2 3
m on the average. The clay is underlain by sand at 17 m depth. It has a density of 1.6 t/m and a compression
modulus ML = 0.17 MPa, prevailing in the stress range G*L - o'c - 20 kPa {cf. Fig. 50). The coefficient
2
of consolidation cv = 0.5 m /year. The groundwater level is situated at 1 m depth below the ground sur-
face. Determine the average settlement of the building after 10 years.

Relative compression, %
0 5 10 15
is 1 1 0
154 Spread foundation
2
Solution: The load release due to excavation is equal to 2 16 = 32 kN/m . Thus, the load increase under
2
the weight of the building is 50 - 32 = 18 kN/m . The stress increase in the soil is below 20kPa and,
therefore, the ML value governs the compression throughout the clay layer. The determination of the
primary compression is exemplified for 5, 10 and 15 m depths.
3 2
5 m depth: = 18/[(1+5/15)(1+5/20)0.1710 ] = 6.4-10'
3
10 m depth: = 18/[(1+10/15)(1+10/20)0.1710 ] = 4.2 K H
3 2
15 m depth: = 18/[(1+15/15)(1+15/20)0.1710 ] = 3.0-10"
The total primary settlement to be expected, which is the area of the primary compression graph, is
sp = 0.85 m. (This settlement corresponds to a CR value of 0.35, see p. 77).
The remaining primary settlement (cf. p. 31)after 10 years is determined by Helenelund's method (p.
2
323). Choosing a layer thickness Az = 3m we have At=(Az) /(4cv) = 2.25/0.5=4.5 years. The construction
is carried out as shown on the previous page.
After a loading time of 2At (9 years) we have a remaining primary settlement sr = 3(0.026 + 0.056
+ 0.052 +0.038 +0.014) = 0.56 m and after a loading time of 3At (13.5 years) sr = 3(0.021 + 0.048 + 0.049
+0.035 +0.012) = 0.50 m. The remaining primary settlement after 10 years can be estimated by
interpolation which yields sr = 0.56 - (0.56 - 0.50)/4.5 = 0.55 m.
Thus, the settlement after 10 years can be estimated at 0.85 - 0.55 = 0.30 m.
Deep foundations 155

DEEP FOUNDATIONS

1. INTRODUCTION

By deep foundations w e consider all the different types of foundation m e t h o d s w h e r e the


building load is carried at great depth beneath the level of the b a s e m e n t of the building
by piles, secant walls, diaphragm walls, e t c F r o m a geotechnical viewpoint, the
analysis of the problems involved in all the mentioned deep foundation m e t h o d s are of
equal character and, therefore, from a theoretical standpoint differ only with regard to
geometry. S o m e t i m e s it may b e difficult to draw a clear border line b e t w e e n shallow and
deep foundations, for instance when the building load is partly carried by direct contact
pressure against a piled footing (or piled raft) and partly by the piles; a kind of mixed
shallow and deep foundation. In this context, the latter type of foundations will be treated
as deep foundations also.

2. PILES

2.1 Common pile types

(i) Timber piles. Timber piles represent the earliest type of piles and w e r e m o r e or less
the only ones used until the end of the nineteenth century. Old buildings on piles,
therefore, p r e s u m a b l y rest on timber piles, often via a mattress of horizontal planks, laid
out in order to achieve an evened-out stress distribution underneath the footings.
N o w a d a y s , timber piles are mainly used in places with deep cohesive soil layers and a
high groundwater level.
T h e m o s t obvious p r o b l e m with timber piles is the deterioration caused by fungi
attacks above and in the vicinity of a fluctuating groundwater level. Rotting of the part
of timber piles that sticks out of the groundwater is very frequent and m a y cause serious
problems. T h u s , as the degree of rotting underneath the different footings is generally
subjected to large variations, rotting usually leads to large differential settlements
followed by severe cracking of the buildings. This is a serious p r o b l e m in m a n y old cities.
Piles, originally submerged, may gradually rise above groundwater level due to withdrawal
of groundwater or, as is the case in Scandinavia, due to the land h e a v e taking place after
the ice age. Typical damages caused by rotting of piles and mattresses are shown in Fig. 110.
Investigations performed by w o o d researchers indicate that bacteriological activity
m a y cause l o n g - t e r m deterioration also below the g r o u n d w a t e r level. In practice,
however, the effect of such possible attacks of bacteria can b e disregarded. A good
156 Deep foundations

Fig. 110. Damages to buildings caused by rotting of piles and matresses.

e x a m p l e of the preservation of submerged w o o d is the C a m p a n i l e of St. M a r k ' s in Venice.


W h e n , in 1902, the Campanile fell because of structural defects, the then 1000 years old
timber piles of its foundation were found to be in such a good condition that they could
be used to support the new structure.
Timber piles in open sea water are generally attacked by m a r i n e borers of various
kinds. A c o m p r e h e n s i v e description of the different species of m a r i n e borers and the
environmental conditions for their activity as well as of different m e t h o d s of preservation,
is given by Chellis (1951).

(ii) Concrete piles. T h e use of concrete piles has m a n y practical advantages and,
therefore, h a v e b e c o m e the m o s t c o m m o n type of piles used in foundation engineering.
T h e m a i n reason for their popularity is that there are almost no restraints in respect of
shape or installation m e t h o d s .
T h e m e t h o d s of installation and production of piles vary with regard to subsoil
conditions and local tradition. In s o m e countries precast piles d o m i n a t e the market,
while, in other countries, piles cast in place are prevailing. T h e m e t h o d of installation has
Deep foundations 157

Fig. 111. Pile shoe of steel for protection of pile tip from being damaged during pile driving.

a great influence on the geotechnical properties to be used in the analysis of the per-
formance of the pile when loaded.
The concrete to be used for the production of precast piles has to be of very high quality
in order to withstand the severe treatment of pile driving. In particular, the toughness of
the concrete is important as the number of blows of the pile hammer required to reach the
expected bearing capacity can be very large and lead to destruction of the pile. Moreover,
in order to protect the tip of the pile from being damaged during installation, the pile is
often provided with a pile shoe of steel (Fig. 111).
The width, or diameter, of precast piles is usually between 0.25 and 0.35 m and the
length limited to 1214 m in order to facilitate transport of the piles from the pile factory
to the building site as well as installation. Special splices have been constructed which
have the capacity of taking up equally large tension forces and bending moments as the
pile itself (Fig. 112). Moreover, the splice connecting two pile segments should be
constructed so as to have the same rigidity as the pile segments themselves.
Piles cast in place exist in many different forms among which the most common ones
are bored and auger piles formed by removing the soil. These are also referred to as non-
displacement piles in contrast to displacement piles which represent all the precast and
cast in place piles which are driven into the ground.
A m o n g the driven cast-in-place piles, the Raymond and the Franki piles are probably
the most well-known.
In some places, concrete can be subjected to considerable corrosion due to aggressive
constituents in the groundwater. Certain molluscs exist in sea water which burrow in
concrete or secrete a substance that has a solvent action even on rock.
158 Deep foundations

Fig. 112. Example of splice for concrete piles.

(iii) Steel piles. Steel piles represent an alternative to precast concrete piles in places
w h e r e the subsoil conditions entail large risks of the concrete piles being d a m a g e d during
pile driving. Rolled steel shapes, for e x a m p l e H piles, which have high bending rigidity
are c o m m o n l y utilised, but rolled bar irons of circular, square or X-shaped cross-section

Concrete pile
with fixed
steel case
--+. J

Fig. 113. Example of splice for composite timber/concrete pile.


Deep foundations 159

are also in use. Steel piles are generally driven into firm, bearing strata often bedrock
or hard till.
Steel piles mostly require some kind of protection against corrosion.

(iv) Composite piles. Composite piles refer to piles which are formed by a suitable
combination of concrete and steel, or concrete and timber.
A conical shape of the piles is advantageous with regard to pile bearing capacity.
Therefore, and of course also for economical reasons, timber piles are still of interest in
foundation engineering. Then, if the head of the timber pile can be expected to stick out
of the groundwater surface, which would entail fungi attacks and rotting of the timber,
composite piles have to be chosen where the timber pile segment is spliced with a
concrete pile segment on top. The main obstacle to the use of these composite piles is the
difficulty of constructing a splice that is rigid enough and can resist the tension forces
obtained during pile driving (Fig. 113).
Tubular-steel piles are another type of composite piles, usually consisting of steel
casing utilised during pile installation and later filled with reinforced concrete. In other
cases, piles are made of a steel core enclosed by reinforced concrete. So-called micro-

Fig. 114. Micro-piles, in this case so-called steel-plastic piles, being installed in room with low ceiling
height. The steel-plastic piles consists of a steel tube covered by 1.8 mm of polyethylene plastic. The
pile segments are spliced by means of galvanised steel tube sleeves which are pressed upon the pile after
that the plastic cover has been softened by heating. The diameter varies from 76 to 102 mm.
160 Deep foundations

piles consist of small-diameter steel tubes filled with cement and provided with some
kind of corrosion protection, for example a plastic cover (Fig. 110).

2.2 Displacement piles

A displacement pile is defined as a solid pile, or a hollow pile driven with its tip closed,
which displaces an equivalent soil volume by compaction or by lateral or vertical
displacement of the soil. Most of the displacement piles are precast concrete piles, steel
piles, or timber piles which by necessity have to be driven into the soil. However, driven
concrete piles can also be cast in place. In this case a closed-ended casing is driven into
the soil whereupon the pile is cast inside the casing. The casing can either be left in the
soil, which is the case with the Raymond pile, or withdrawn from the soil, as is the case
with the Franki pile. The Raymond pile is tapered which improves its bearing capacity.
The Franki pile can be installed in several ways (Fig. 115). There is, however, a common
feature. The casing is filled before installation with 'dry' concrete to a height of 1 to 2 m.
The concrete plug, thus formed in the casing, serves as a water-tight pile tip during
driving. The pile is then driven into the soil by means of 4 - 5 m long hammer dropped
inside the casing. When the pile tip has reached its design level, the casing is fixed to the
ground surface and the concrete plug is driven out of the casing, thereby causing
compaction of the soil around the pile tip. In this way the Franki pile b e c o m e s
dynamically preloaded.

(i) Driving equipment. Displacement piles are driven into the soil, generally by means of
gravity hammers but also by the use of steam/air hammers, diesel hammers of different
kinds or vibratory drivers. The gravity hammer is simply a weight (generally t) which
is lifted a certain distance with a hoist line and then released to fall and strike a drive cap
(Fig. 116).

(a) (b) (c) (d) (e) (f) (g)

Fig. 115. Schematic picture of hammer and driving tube and installation methods for three different
types of Franki piles: a+b+c = standard Franki pile, a+b+d+e = vibrated Franki pile and a+b+f+g =
composite Franki pile.
Deep foundations 161

Fig. 116. Gravity hammer and steel cap.

T h e steam/air h a m m e r s in use are of type single-acting, double-acting, differential-


acting and c o m p o u n d h a m m e r s . T h e single-acting steam/air h a m m e r is similar to the
gravity h a m m e r except that it has a cylinder and a piston to lift the r a m weight instead
of a hoist line. T h e lift of the h a m m e r is produced by compressed air or steam (motive
fluid). T h e double-acting steam/air h a m m e r and the c o m p o u n d steam/air h a m m e r differ
from the single-acting h a m m e r in that the m o t i v e fluid is also introduced over the piston
to accelerate the r a m in its d o w n w a r d stroke. T h e differential-acting steam/air h a m m e r
is another type of double-acting hammer. M o t i v e fluid is introduced b e t w e e n large and
small piston heads to lift the r a m to the top of its stroke. T h e n m o t i v e fluid is introduced
over the large piston head to accelerate the r a m in its d o w n w a r d stroke.
Diesel h a m m e r s are of two different kinds: open-end (single-acting) and closed-end
(double-acting) h a m m e r s . Both these diesel h a m m e r s are one-cylinder diesel engines in
which the potential energy of diesel fuel combustion is converted through the m o v i n g
piston to kinetic i m p a c t energy delivered at the anvil block. T h e difference between the
closed-end and the open-end diesel h a m m e r s is that the gravity-controlled flight of the
r a m of the open-end h a m m e r is shortened in the closed-end h a m m e r . This is produced by
decelerating the r a m ' s u p w a r d stroke by trapping air to form an air spring. This m a k e s
the closed-end diesel h a m m e r operate at about twice the b l o w rate of an o p e n - e n d diesel
h a m m e r of c o m p a r a b l e size.
Vibratory drivers get their driving capacity by the centrifugal force produced by
162 Deep foundations

hammer

--
pile

J w +

Fig. 117. Stresses acting on pile element due to the impact of the pile hammer.

eccentrics in the driver which can be rotated at a steady-state frequency w h e n loaded with
an oscillating pile. Their driving efficiency depends on amplitude, frequency, centrifugal
force, vibrating weight and non-vibrating weight. European vibratory drivers usually
operate at a frequency of 1 0 - 3 0 H z while, for instance, the A m e r i c a n B o d i n e Resonant
Driver operates at a frequency of 5 0 - 1 5 0 Hz.

(ii) Influence of stress wave during pile driving. T h e stress w a v e caused by the impact of
the r a m is travelling d o w n w a r d s the pile until it reaches the pile tip w h e r e it is reflected,
travelling u p w a r d until it reaches the pile head, reflected again and travelling d o w n w a r d s ,
and so on. Stress w a v e analysis is used as a basis for determination of the bearing capacity
of driven piles (see p. 175).
T h e effect of the stress w a v e on the pile can be studied by simply assuming that
damping of the stress w a v e by the surrounding soil can b e disregarded.Considering the
forces acting u p o n a pile element (Fig. 117) under the assumption of no influence of the
surrounding soil, w e h a v e ( N e w t o n ' s 2nd law):

w
Ap = -Appp 2
(162)

w h e r e pp is the density of the concrete pile.


Deep foundations 163

But = Epz= - Epdw/dz, whence:

2 2
dw 2
jd w
2 c - ^ (163)
dt dz

w h e r e c = yj' Epl pp = the velocity of the stress w a v e .


Eq. (163) has the solution:

w(z,0 = M z - et) +f2(z + et) (164)

w h e r e f \ ( z - ci) a n d / 2 U + ci) are constants of integration.


T h e first of these terms represents the stress w a v e travelling d o w n w a r d s (z increases
when t increases) while the second represents the reflected w a v e (z decreases w h e n t
increases).

T h e derivative of Eq. (164) takes the form:

dw
(, 0 = -cf\ (z-ct) + cf'2(z + ct) (165)
at

N o w , for the initial (primary) stress w a v e (the first term in Eq. 164) w e find the
following correlation:

dw , E n dw
= -E (z,t)
p = -EJ^z-ct) = ^ ( z , t ) (166)
az c m

Hence:
dw dw
Oz = {z,t)y/Eppp = (Z IA )(z,t) (167)
at p
at
where -E^A^fc is the i m p e d a n c e ; alternatively n a m e d the d y n a m i c stiffness of the pile
(cf. the static stiffness EpAp/lp).
In the case of the gravity h a m m e r , the velocity enforced on the pile w h e n it hits the pile
h e a d is:

= a/gh (168)
dt
where h = free fall-height of the h a m m e r (cf p. 178, h a m m e r efficiency),
g = acceleration of gravity.
164 Deep foundations

1 _ 1

~ l + E J^~ \+z /zp h

(index refers to the pile and index h to the pile h a m m e r ) .


T h e b o u n d a r y condition for = 0 yields:

= ^/ / = aZy/lghlAp
(169)

If E - > w e find a = 1.
If the moduli of elasticity and the cross-sectional areas and the densities of the pile and
the pile h a m m e r are the same, i.e. Eh- Ep\ Ah- Ap\ rh- rp, w e find a = 0.5. In the latter
case the stress w a v e b e c o m e s rectangular with a length equal to double the length of the
pile hammer.
T h e shape of the stress w a v e can be adjusted by choosing a suitable cushion block
placed in the drive cap under the hammer. T h e length of the w a v e is independent of the
drop height.
F r o m Eq. (169) w e find that the normal stress induced in the pile by the impact of the
h a m m e r is dependent on the drop height but independent of the weight of the hammer.
In case the pile tip is driven d o w n to rigid material (big blocks or bedrock), the reflected
stress w a v e will be a compressive stress w a v e with the same intensity as the stress wave
travelling d o w n w a r d s . T h u s , for a short m o m e n t the stresses at the pile tip will
superimpose which often leads to crushing of the pile tip (Fig. 118).
In the case of rectangular stress waves, the m a x i m u m compressive stress at the pile tip
will be:

Fig. 118.Crushing of pile tip due to overlapping of the compressive stress wave caused by pile driving.
Deep foundations 165

Fig. 119. Tensile cracks in concrete pile caused by the tensile stress wave obtained in pile driving.

(170)

Normally, however, the m a x i m u m compressive stress reached at pile refusal does not
exceed 1.5-1.8 times the initial stress.
If, on the other hand, the soil resistance at the pile tip is negligible, the stress w a v e will
stretch the pile tip and the reflected w a v e travelling u p w a r d s will be a tensile stress w a v e
of equal intensity as the compressive stress w a v e reaching the tip, /. e.\

(171)

In this case, tensile cracks in the pile m a y be obtained (Fig. 119). If the pile is under
water, water will be sucked into the cracks and then forced out of the cracks by the
following compressive stress w a v e , a jetting process that m a y cause erosion of concrete
around the opening of the cracks.
T h e stress w a v e velocity in h o m o g e n e o u s (uncracked) piles is 3 2 0 0 to 3 9 0 0 m/s,
depending upon the quality of the concrete, and in steel about 5 1 0 0 to 5 2 0 0 m/s.
The driving of the piles b e c o m e s most efficient w h e n the stress w a v e has a rectangular
shape. Therefore, the design of the gravity h a m m e r and the drive cap has been carried out
with the attempt of achieving a stress w a v e that b e c o m e s as closely rectangular as
166 Deep foundations

Fig. 120. Example of cup spring drive cap.

possible. For e x a m p l e , a cup spring drive cap (Fig. 120) has been developed w h i c h limits
the peak value of the stress w a v e and, consequently, m a k e s it independent of the height
of fall of the hammer. An increase of the height of fall only increases the length (the
duration) of the stress w a v e . By using the cup spring drive cap, crushing of piles driven
by m e a n s of steam/air h a m m e r s can also be avoided.

(iii) Influence on soil properties of pile driving. In c o n s e q u e n c e of the soil displacements


taking place during pile driving, the properties of the soil are subjected to changes,
sometimes for the better, sometimes for the worse. T h e vibrations induced in the soil and
the enforced soil displacements during pile driving m a y give rise to large and destructive
settlements of adjacent buildings (Fig. 121). This is certainly to be expected w h e n piles
are driven into loose granular soil. In fine-grained loose soils, pile driving m a y induce
high residual excess pore water pressures. As the n u m b e r of piles increases, w e m a y end
up with a state of liquefaction, or the shear strength of the soil m a y decrease to such an
extent that the stability of the place is put at stake. In particular, this peril has to be taken
into account in sloping clay regions with silt and sand layers. Two such cases of failure
due to piling w e r e reported in the beginning of the 1970s (see H a n s b o , 1987). In both
cases, the sites had fairly similar geological conditions: smooth clay slopes with a very
low inclination. In one case, after driving of only 5 piles, a 2 0 - 3 0 c m w i d e crack, 5 5 0
m in length, developed causing severe d a m a g e to three houses. In the other c a s e , pile
driving for s o m e terrace houses induced a 5 - 1 0 c m w i d e crack, 2 5 0 m in length, causing
s o m e d a m a g e to streets and houses and breaking two water conduits.
Deep foundations 167

Fig. 121. Damage to building caused by pile driving for-foundation of adjacent building.

M o s t probably, piling also triggered the slide disaster at Surte in 1950 comprising an
area of around 25 h a (Fig. 122). In the slide, one person was killed and two badly hurt
while 31 houses were m o r e or less destroyed.
In dense, saturated granular soil, pile driving by m e a n s of a gravity h a m m e r or a steam/
air h a m m e r m a y b e very time-consuming and troublesome and entail considerable risk
of concrete piles fatigue. However, the rate of penetration of the piles can b e increased
considerably by the use of vibratory drivers. In that case, the bearing capacity and the
settlement behaviour of the pile will b e dependent on the frequency of the driver during
installation.
So-called false refusal, i.e. w h e n the resistance observed during pile driving will be
strongly reduced after termination of driving, represents a p r o b l e m often encountered,
particularly in fine-grained cohesionless soils. This p h e n o m e n o n is m o s t probably due
to an increase in soil strength by pore water underpressure (suction) i n d u c e d during pile
driving.
T h e compaction of loose granular soil caused by pile driving m a y obstruct the
possibility of driving the piles to equal depth in a pile group. T h u s , the piles first driven
reach m u c h deeper before the required resistance is obtained than the piles driven at the
168 Deep foundations

Fig. 122. The slide at Surte north of Gothenburg on Sept. 29, 1950, is assumed to have been triggered
by pile driving in the area.

end. This entails bad e c o n o m y since the piles not driven to full depth will h a v e to b e cut.
Moreover, the load/settlement characteristics of the piles first driven m a y b e quite
different from those last driven. This p r o b l e m can be avoided by providing the piles with

Fig. 123. Installation of piles into fine-grained till by water jetting.


Deep foundations 169

Fig. 124. Coring operation in soft clay. To the left: clay core just withdrawn from the soil. To the right:
clay forced out of the tube by the aid of compressed air.

central tubes and special arrangements for the application of water jetting (Fig. 123).
W h e n the piles h a v e reached the intended depth, the pile installation can then be
finalised b y driving the pile to greater depth with a pile h a m m e r .
Pile driving in soft clays usually entails a heave and a reduction (sometimes considerable,
for e x a m p l e in highly sensitive clays) of the undrained shear strength w h i c h has to be
taken into account in stability analyses. In order to reduce the disturbance effects, clay
cores are taken w h e r e the piles are to b e driven. T h e coring operation is carried out by
m e a n s of a hollow tube with a vent at the top w h i c h is kept open w h e n the tube is driven
into the soft clay to release trapped air. T h e tube is driven to a depth of 6 to 8 m. T h e vent
is closed and c o m p r e s s e d air is injected into the tube at its b o t t o m to eliminate v a c u u m
during withdrawal. After withdrawal, the clay is forced out of the tube b y m e a n s of
compressed air (Fig. 124).

2.3 Non-displacement piles

N o n - d i s p l a c e m e n t piles are piles formed by boring or other m e t h o d s of excavation.

(i) Installation methods. B o r e d piles can b e installed by a great variety of m e t h o d s , using


sludge p u m p s , h a m m e r grabs, augers, rotary buckets, continuous augers, percussive or
rotary boring with direct or reverse fluid circulation, coring b u c k e t s , etc. (Fig. 125). T h e
170 Deep foundations

Fig. 125. Examples of installt of borec' ^iles: To the le. use of auger boring. To the right: use of
rotary boring (by courtecy of Bachy).

reinforcement, in the form of a prefabricated cage, is usually placed in the hole prior to
concreting. In the case of continuous auger piles, the reinforcement cage is placed
immediately after the auger has been r e m o v e d (Fig. 126).

Fig. 126. Reinforcement cage under preparation (by courtecy of Bachy).


Deep foundations 171

Fig. 127. Underreamer (by courtecy of Bachy).

In many cases the borehole for the pile can be formed in stable ground condition where
no support of the sides of the hole is required. In other cases, however, the employment
of casings or, more commonly, bentonite slurry may be required to ensure the stability
of the borehole.
Casings are generally used only as a temporary measure to provide stability during
boring and concreting. Permanent casings may be necessary in weak soils which cannot
sustain the lateral pressure of the fresh concrete.
If the borehole is stabilised by means of bentonite suspension a short collar casing is
almost always used. The use of bentonite does not seem to have any negative influence
on the quality or integrity of the pile. On the contrary, it seems safer to use bentonite for
borehole stabilisation than casing (Sliwinski & Fleming, 1984)
Bored piles with a diameter ranging from 0.3 to 0.6 m are generally referred to as small
diameter piles and those with diameters above 0.6 m as large diameter piles or sometimes
as caissons. With modern methods, bored piles can be installed with a m a x i m u m rake of
4:1 (vertical:horizontal). The diameter is in the range of 0.45 to 3.5 m. By means of a
special tool, the piles can be underreamed to a tip diameter of up to 5.4 m (Figs. 127-128).
The diameter of continuous flight auger piles is in the range of 0.35 to 1.5 m.

(ii) Installation problems. In difficult soil conditions, for instance when the ground
contains a great number of boulders, blasting may have to be carried out below the tip of
the casing. The shock waves produced by blasting may induce excess pore water pres-
sures and cause compaction of the soil in a similar way as during pile driving. Careless
excavation in sand and silt soils below the groundwater level involves a great risk of
bottom erosion (piping) which can badly affect the ground conditions in the vicinity ofthe
172 Deep foundations

Fig. 128. Underreamed cave under inspection. Underreaming can only be used in the case of stiff or very
stiff clays.

excavation. In this case, the borehole for the pile has to be stabilised by m e a n s of bentonite
suspension or by filling the borehole with water to a higher level than that of the sur-

Fig. 129. Damage to building caused by installation of bored piles.


Deep foundations 173

\fs
(negligible)

m
qt (negligible)
Fig. 130. End bearing pile (left) and friction pile.

rounding groundwater. As in the case of driven piles, the installation of bored piles may
lead to severe d a m a g e to nearby buildings (Fig. 129).
Before the excecution of holes for piles in soft clay, the shear strength of the clay has
to b e k n o w n . T h e existence of possible sand or silt layers is a reason for taking special
precautions. T h e release of overburden pressure during excavation can lead to a drastic
decrease in the shear strength below the b o t t o m of the hole and adventure stability even
though the original shear strength might h a v e seemed satisfactory.
T h e integrity of bored piles can be jeopardised by segregation of concrete, inclusion
of soil or slurry, cavities in the concrete, displacement of the reinforcement cage, etc.
Therefore, possible sources of defects m u s t be anticipated and measures taken to ensure
the integrity of the piles.

2.4 Load transfer pile/soil

T h e load carried by the pile is transferred to the soil b y frictional resistance along the pile
shaft and/or b y pile tip resistance (Fig. 130). If the major resistance to load is derived
merely by side friction (or by adhesion in the pile/soil interface), w e speak offriction piles
(or, less commonly, floating piles). On the other hand, if the major resistance to load is
exerted by the pile tip, w e speak of point bearing piles or, alternatively, end bearing piles.
A friction/end bearing pile is a pile that carries the load by both frictional resistance and
point resistance.
T h e bearing capacity of end bearing piles is of course dependent on the characteristics
of the soil on which the pile tip is resting and on tip dimensions. In h o m o g e n e o u s soil,
the end bearing capacity increases with depth until the pile tip reaches the so-called
'critical depth' below which no further increase can be noticed. Dynamic preloading
174 Deep foundations

improves the bearing capacity of the soil. Therefore, driven piles can usually c a n y a
higher load than non-displacement piles with the same tip area.
T h e bearing capacity of friction piles depends on the angle of internal friction in the
soil, the coefficient of friction between soil and pile and the normal pressure against the
sides of the pile. A considerable n u m b e r of pile loading tests h a v e s h o w n that the bearing
capacity of friction piles in cohesionless soils generally increases with time after pile
installation (stedt et , 1990). This is most probably due to postdensification effects,
similar to those observed in connection with soil i m p r o v e m e n t by blasting or vibratory
compaction (Mesri et , 1990).
In clay soils, the bearing capacity is governed by the long-term shear strength of the
clay. T h e roughness of the sides of the pile and pile material h a v e certainly a considerable
influence on the frictional resistance in granular soils but s e e m negligible in cohesive
soils. W h e n , in the latter case, the pile is driven, the soil around the pile will be remoulded
within a zone that extends to at least one pile diameter from the pile surface. T h e resulting
excess pore water pressure around the pile m a y reach and even locally exceed the total
overburden pressure. Reconsolidation of the r e m o u l d e d clay leads to an increase in the
undrained shear strength of the clay. T h e reconsolidated, r e m o u l d e d clay adheres to the
pile surface and rupture takes place in the clay with lower shear strength outside the pile
surface (cf. B r o m s & H a n s b o , 1981).
T h e shape of the pile is important for the bearing capacity of friction piles. Conic piles,
for instance, such as timber piles with a pile tip diameter considerably smaller than the
diameter of the pile head, are very advantageous from the point of view of frictional
resistance.

3. BEARING CAPACITY O F AXIALLY LOADED SINGLE PILES

In foundation engineering m u c h interest has been devoted to the p r o b l e m of predicting


the ultimate bearing capacity of single piles. In consequence, a great m a n y different
methods h a v e been advanced, theoretical on the basis of fundamental soil parameters as
well as half empirical to merely empirical. T h e prediction of the ultimate bearing capacity
is also part of the 'ultimate limit state design' philosophy included in the Eurocode. T h e
presentation herein includes methods of historical interest and m e t h o d s which in the
A u t h o r ' s opinion form a sound and reliable basis for design.

3.1 Pile loading tests

In the case of piled foundations, full-scale loading tests are often prescribed as a check
on the bearing capacity predicted according to the m e t h o d of analysis applied. Full-scale
loading test h a v e the advantage of providing information not only of the ultimate pile load
but also of the w h o l e load vs. settlement relationship.
T h e ultimate pile load by definition is the load that leads to failure. T h e r e are, however,
Deep foundations 175

Fig. 131. Determination of the ultimate load by the Polish method (Mazurkiewicz, 1972), (left), and by
Brinch Hansen's 80% criterion (Brinch Hansen, 1963).

cases w h e r e the ultimate load according to this definition is difficult to e s t e e m from the
shape of the load/settlement curve. Therefore, rules h a v e been w o r k e d out to ease the
determination of the ultimate load irrespective of the shape of this curve. Unfortunately,
however, such rules exist in great n u m b e r s and they do not give us o n e and the same
answer (This has been thoroughly discussed in Sellgren, 1981).
Also the w a y the pile loading test is carried out affects the result obtained. Certainly,
the best w a y of performing the test is to use the loading procedure described previously
in the case of plate loading tests (see p. 79), a so-called " M L T " (maintained load test), in
which the load is increased in defined increments and each load level is held at equal
length of time. This gives us not only the ultimate load but also the critical pile load, also
called the creep load, which represents the load w h i c h causes a sudden increase in the rate
of creep settlement of the pile (in this case equal to the creep that takes place in each load
step).
T h e ultimate load can preferably b e interpreted on the basis of the so-called Polish
m e t h o d or by Brinch H a n s e n ' s 8 0 % criterion (Fig. 131).

3.2 Methods of analysis

(i) Pile driving energy. In this approach it is a s s u m e d that the pile and the h a m m e r are both
totally stiff (cf. stress w a v e analysis).
Let us a s s u m e that the pile is driven by m e a n s of a gravity h a m m e r . Let us further m a k e
the following assumptions:
176 Deep foundations

Fig. 132. Assumed correlation between pile load and pile movement. The shadowed area represents
consumed energy by the impact of the hammer.

m a s s of gravity h a m m e r mh
mass of pile mp
velocities of hammer pile
at the impac vh 0
immediately after Vh v'p
coefficient of restitution pile/hammer e.
T h e correlation between the velocities of the gravity h a m m e r and the pile then
becomes:
v'u = v'p-evh (172)

T h e law of m o m e n t u m yields:

mhvh = mhvh + m/p = mhv'h + mp(v h + evh) (173)

whence
m e m
* - P (174)
v'h = /
mh + nip
and

v'p = v h - ^ - ( \ + e ) (175)
mh 4- nip

T h e total energy c o n s u m e d by the impact of the h a m m e r (provided mh > emp) is:

w L
= -(mhvh +m pv p ) = (176)
I 2 m h + mn
Deep foundations 111

2
or, since vh = 2gh, w h e r e h is the drop height of the h a m m e r and g is the acceleration of
gravity,
2
mh + e mp
W = gmhh ^ (177)
mh+mp
This energy is transferred into pile m o v e m e n t . A s s u m i n g that the elastic r e b o u n d of the
pile head is sel and that the remaining, plastic (irrecoverable) m o v e m e n t is spl (the set),
w e have (Fig. 132):

W= Qf(spl+-sel) (178)

This yields the w e l l - k n o w n Hiley's pile formula:

2
m m h h m h + e m p

+
s m m
Spl 2d h + p
w h e r e is the h a m m e r efficiency.
T h e m a g n i t u d e of the coefficient of restitution depends on whether or not a drive cap
is used and, in such a case, on the properties of the drive cap unit (see, for e x a m p l e ,
Chellis, 1951).
A s regards the h a m m e r efficiency w e can a s s u m e = 1 for free-fall h a m m e r s and
7] = 0.75 for single hoist line h a m m e r s .
In the S w e d i s h pile code, a modified H i l e y ' s formula, especially useful w h e n the pile
is driven b y the aid of a mandrel, has been accepted which takes into account results of
stress w a v e m e a s u r e m e n t s :

Qf= * 0.8(1-0.1-^) (180)


Spi +sel/2 m h

where

A E EA 9
Pp p P ff
mh>3t,
Ep = elastic m o d u l u s of pile,
Ap = cross-sectional area of pile,
Pp = density of pile,
If = length of follower,
Ef- elastic m o d u l u s of follower,
Af= cross-sectional area of follower.
178 Deep foundations

Example 21: Determine the ultimate pile load according to Hiley's pile formula, on the one hand, and
according to the modifed pile formula of the Swedish pile code, on the other, if the permanent set spl
= 3 mm per blow using a 3 1 single hoist line hammer and a drop height of 0.3 m. The rebound sd is estimated
at 10 mm. The pile is 12 m in length and has a cross-sectional area of 0.25 m by 0.25 m and a modulus
3
of elasticity of 3 GPa. The density of the pile is 2.4 t/m . A 2 m long steel mandrel with the same cross-
sectional area as the pile is utilised for the pile driving. The elastic modulus of the steel mandrel is 20
GPa. The coefficient of restitution e = 0.5.

2
Solution: The mass of the pile mh = 120.25 2.4 = 1.8 t. For a single hoist line hammer the hammer
efficiency = 0.75.
Hiley's pile formula yields:
0 . 7 5 . 9 . 8 1 . 3> 0.5 3 + 0.52-1.8
^ 0.003 + 0 . 0 1 0 / 2 3+1.8
According to the modified pile formula, the rebound sel is expressed as a function of and need
not be measured. Substituting sel in Eq. (180), Qf can be solved analytically from a second degree
equation. It can also be found by trial and error:
Assuming Qj= 992 kN, we find:
3
Sei =992( -4A -6 + 2 6
) = 10.74- 10" m
2 . 4 0 . 2 5 3 0 . 0 . 2 5 2 0 0
whence
0 . 7 5 . 9 . 8 1 . 3 . 0.5 1 0

^ 0.003 + 0 . 0 1 0 7 4 / 2
i.e. the same value as that obtained by Hiley's formula.

Hiley's pile formula can be applied also for single-acting steam/air h a m m e r s . In the
case of double-acting steam/air h a m m e r s or diesel h a m m e r s being used, the impact
velocity of the r a m vr is obtained from the relation:

2
vr = 27] Wlmr (181)

w h e r e W i s the impact energy, is the h a m m e r efficiency and mr is the mass of the ram.
This yields the pile formula:

Of = -
S

E
(182)
Spl + 2 el Wr + Mp

For the m o s t c o m m o n types of steam/air h a m m e r s , M c Kiernan Terry, is reported


equal to 0.85. Information about the rated impact energy W and the h a m m e r efficiency
should b e accessible before a certain h a m m e r is put into use.
In the case of vibratory pile driving, the energy input for o n e cycle of oscillation is
analogous to o n e blow of an impact h a m m e r (Davisson, 1970). Consequently, the set
should b e put equal to the rate of pile penetration rp (m/s) divided by the driving frequency
/ ( H z ) . A s s u m i n g that the effect of the driver is , the energy input per cycle of oscillation
equals Elf and the ultimate pile capacity can b e estimated by the relation:
Deep foundations 179

E+gmvrn
P
Qf= , 083)
rp +fsL

w h e r e mv = m a s s of the vibratory driver,


sL = loss factor.
According to Davisson (1970) the loss factor sL for the Bodine Resonant Driver can
be a s s u m e d equal to about 0.25 m m / c y c l e for loose silt, sand and gravel, 0.75 m m / c y c l e
for m e d i u m dense sand or sand and gravel and 2.5 m m / c y c l e for dense sand and gravel.
For -piles, however, sL should be put equal to - 0.2 m m / c y c l e in loose silt, sand and
gravel.

(ii) Stress wave analysis. T h a n k s to m o d e r n measuring technique, the stress w a v e in-


duced by the impact of the h a m m e r (cf. p. 162) can be used as a m e a n s to determine the
bearing capacity of the pile. By the aid of a stress transducer and an accelerometer, placed
on the pile j u s t b e l o w the pile head, the force and the particle velocity at the measuring
point is registered as a function of time. T h e data thus obtained are analysed in different
w a y s , m o s t c o m m o n l y by the so-called Case and C A P W A P m e t h o d s .
According to Eq. (167), the initial compressive force induced by the impact of the
h a m m e r can b e expressed by the relation:

= ^ ( , 0 = (184)
dt

w h e r e = EpAplc is the pile i m p e d a n c e ,


= particle velocity.
T h e deduction of Eq. (167) w a s based on the assumption that the pile w a s unaffected
by the surrounding soil during pile driving. In reality the soil will offer d y n a m i c resistance
against the pile motion. T h e influence of this resistance must b e taken into consideration
when predicting, b y m e a n s of stress w a v e analysis, the load/settlement characteristics of
the pile.
In the Case analysis, the d y n a m i c resistance is represented by the so-called d a m p i n g
factor Jc. It is a s s u m e d that Jc is a linear function of the particle velocity of the pile. By
the aid of the stress transducer and the accelerometer, the force Q and the particle velocity
are m e a s u r e d j u s t b e l o w the pile head. T h e total pile force is the s u m of the force
travelling d o w n w a r d s and the force travelling u p w a r d s , /. e.\

Q = IQ + Q (185)
180 Deep foundations

T h e particle velocity is a function of d y n a m i c force and can be expressed in a corre-


sponding w a y by the relation (velocity in d o w n w a r d s direction positive):

v = dw/dt = lQ/Z-Q/Z (186)

B e t w e e n the stress w a v e travelling d o w n w a r d s and the stress w a v e travelling upwards


there is a p h a s e shift in time of 211c, w h e r e / = distance b e t w e e n the m e a s u r i n g point and
the tip of the pile and c = w a v e velocity. T h e reflected stress w a v e has been affected by
the frictional resistance along the pile shaft and the force m e a s u r e d is the sum of the tip
resistance and the shaft resistance.
T h e bearing capacity of the pile, derived from the results of the stress w a v e measurements
and divided u p into two c o m p o n e n t s , one directed d o w n w a r d s and the other u p w a r d s , can
thus b e written:

lQ = [Q(t) + v(t)Z]/2
= [(*+2Z/c) - v(f+2//c)Z]/2

which yields the total force:

() + (0 Q(t + 2l/c)-Zv(t + 2l/c)


tot = 2 + (188)

N o w the d y n a m i c resistance can be expressed as:

dyn = ^ - T ) (189)
or
r r (r)+Zv(Q ( /+ 2 / / c ) - Z v ( ; + 2 / / c ) n
dyn=4l - - ~ J U^U;

which can also b e written:

dyn = ^ [ ( 0 + Zv(i)- J
t o (191)

T h e static bearing capacity of the pile is the difference b e t w e e n total and d y n a m i c


capacities, i.e.:

stat = t o t - J c m ) + () - t ]
o t (192)

T h e m a g n i t u d e of the d a m p i n g factor Jc depends on the soil type and the relative


density of the soil. T h e Jc values for different soils can b e a s s u m e d to lie b e t w e e n the
following r a n g e s :
Deep foundations 181

sand 0.05-0.20
silty sand 0.15-0.20
sandy silt 0.20-0.30
silt 0.20-0.45
silty clay 0.40-0.70
clay 0.60-1.10

T h e m a g n i t u d e of the d a m p i n g factor Jc is of little i m p o r t a n c e in the case of end


bearing piles but is of great i m p o r t a n c e in the case of friction piles. Therefore, in case of
friction piles being used, the Case m e t h o d is often utilised as a c o m p l e m e n t to other
computational m e t h o d s , such as the C A P W A P and the S V I D Y N m e t h o d s .

Example 22: Determine by the Case method the static bearing capacity of a driven concrete pile, 20 m
in length and 235 mm in width, for which the stress wave measurements has given the results shown
below. The measuring point is 1 m below the pile head. The pile tip is in silt. The elastic modulus of
concrete can be assumed equal to 30 GPa.

Solution: In this case, 7 c c a n be assumed equal to 0.3. The stress wave velocity is obtained by the relation:
9
c = y/Ep/pp = ^ 3 0 - 1 0 / 2 4 0 0 = 3536 m/s
The time from the initiation of the stress wave to the arrival of the reflected stress wave:
t = 2lp/c = 2-19/3536 = 10.7 ms
which is in good agreement with the stress wave interpretation that yields t = 11 ms.
By setting off the time iy c from the time of the measured force maximum ( m ax = 1610kN) we find
the magnitude of the reflected force (vZ). We obtain:
Ou* = (1610 + 550)/2 + [(1610 - ( - 167)]/2 = 1970 kN
whence:
s t a=
t 1 9 7 0 - 0 . 3 ( 2 - 1 6 1 0 - 1 9 7 0 ) - 1600 kN

In the CAPWAP analysis (Rausche et , 1972) the pile and the h a m m e r are represented
by finite rigid e l e m e n t s (Fig. 133) j o i n e d together by springs representing the stiffness
of the pile and the h a m m e r while the soil resistance is represented by elastic springs and
rigid plastic s p elements in combination with viscous dashpots. A s s u m i n g that the total
dynamic and static soil resistance per unit length of the pile is Rd w e h a v e the stress w a v e
equation (cf. Eq. 163):
182 Deep foundations

viscous dashpot

Fig. 133. Pile/soil model used in the CAPWAP analysis. Elements 1 and 2 represent the hammer and
the pile cap, elements 3-11 the pile itself. The resistance of the surrounding soil is represented by springs
in parallel with viscous dashpots.

2 D2
dw 2 W
= c (193)
A
dz Pp p

w h e r e pp- density of the pile,


Ap= cross-sectional area of the pile.
Finite difference modelling of this equation enables the drivability and the bearing
capacity of the pile to be examined. T h e computation starts with estimated values of the
distribution of soil resistance, damping and quake. By modifying the assumed parameters
until the m e a s u r e d and the calculated stress waves concur, a good correlation is generally
achieved between predicted and calculated bearing capacities and pile settlements under
working load condition (Fig. 134).
T h e soil model used i the SVID FN program (Nguyen, 1987) represents an improvement
of the model used in the C A P W A P p r o g r a m in that the input parameters are directly
correlated with the conventional soil parameters determined by field and laboratory
investigations.
In the analysis, the static and dynamic soil resistance against the displacement w of a
pile element can be expressed by the relation:

Rd Al = ksw + (Cy+ Crf- CR)v (194)

w h e r e Al = length of the pile element,


Deep foundations 183

Force, kN

2000
Measured
1500 L Calculated

1 1 1
' ' '
10 15 20 2 5 ^ " 30 35 40 45 50
Time, ms

Fig. 134. Result of computerised CAPWAP analysis. Good agreement has been obtained between the
measured and calculated force vs. time function by successively readjusting the spring and damping
coefficients of the soil.

ks = spring stiffness of the soil element,


Cv= viscous damping,
CH= hysteretic damping,
CR = radiation damping,
= dw/dt = particle velocity of the pile element.
Regarding the shaft resistance the soil parameters can b e expressed b y the relations:

ks = nGAl (195)

w h e r e G = shear m o d u l u s of the soil, see Eq. (25).


T h e elastic behaviour is only valid u p to a certain displacement limit, the so-called
quake value wq, representing shear failure in the soil. T h e q u a k e value can b e estimated
from the relation:

Dn^mzx 5/(1-)
w
9 =-
4G
^
P

Dp
P
I+ >
2
(196)

w h e r e lp = length of the pile,


D p = pile diameter,
vp = Poisson 's ratio of the pile,
T
max according to Fig. 16 and Eq. (26).
T h e hysteretic and radiation d a m p i n g s can b e estimated by the relations:

CH = TiDpMDyfp~G (197)
184 Deep foundations

CR =nDpAl/pG (198)

w h e r e D = d a m p i n g ratio,
pp = density of the pile,
= soil density.
T h e d a m p i n g ratio D can be obtained by the the relation:

= 7 ^ - (199)
1+7/*

y by
w h e r e yh = [1 + a e x p ( - ) ] .
Yr 7r

T h e parameters a, b and D m ax for sand are given in Table 2 3 .

TABLE 23
Parameters and b and maximum damping ratio D m ax according to Hardin & Drnevich (1972).

Soil type a b

1/6
Clean, dry sand 0.6yV- -l 1_/12 1
3 3 - 1.5-logW )
1 /6 m2
Clean, water saturated sand 0.54/V- - 0.9 0.65(1 -N~ ) 2 8 - 1.5-logW

^ = the number of loading cycles (in this case the blow counts for the last m of penetration)

As viscous d a m p i n g Cv is usually negligible in comparison with CH and CR, it can be


assumed equal to zero in order to reduce the computation time (Nguyen, 1987).
As regards the pile tip resistance, w e can a s s u m e the correlations:

k^lGD^l-v) (200)

w h e r e = Poisson's ratio of the soil.


In this case the q u a k e value is obtained from the relation:

w = Qp/ks (201)

w h e r e Qpf represents the tip resistance of the pile.


T h e hysteretic and radiation damping can b e assumed to b e governed by the relations:
Deep foundations 185

0.85Z)/
CR = (203)
1-v

w h e r e v i s the Poisson ratio of the soil.


As p r e v i o u s l y s h o w n on, see p. 2 4 , Eq. (25), the relation b e t w e e n shear strain and shear
m o d u l u s of the soil can b e e x p r e s s e d as:

1
G = G 0 ( l + 7/>)-
5
w h e r e G 0 = shear m o d u l u s for strains y~ l(h ,

with and a c c o r d i n g to Table 4, p. 27 (/Vin Table 4 represents in this case the b l o w count
for the last m e t r e of penetration).
T h e shear strain a m p l i t u d e can be taken as the ratio of the a v e r a g e particle velocity
of the pile and the shear w a v e velocity vs of the soil, i.e.:


(204)

As the shear strain a m p l i t u d e s , p r o d u c e d by pile driving, are quite large, both shear
m o d u l u s and d a m p i n g factors m u s t b e d e t e r m i n e d as functions of the i n d u c e d strain in

Example 23: Determine the input values of shear modulus, soil stiffness, soil rigidity and damping
factors to be applied in the SVIDYN analysis for a pile shaft element at 10 m depth in saturated sand
3
with density = 1.8 t/m , angle of internal friction 32 and Poisson's ratio = 0 . 4 5 . The shear wave
3
velocity in the sand is 200 m/s. The steel pile has a diameter Dp = 0.5 m and a density pp = 7.78 t/m .
The observed number of blow counts per m of penetration = 500. The average particle velocity can
be assumed equal to 1 m/s.

Solution: The shear modulus can be determined from Eq. (25). According to Table 4 we have a - -
3
0.2-log500 = - 0.540 and = 0.16. From the shear wave and particle velocities we find y= 1/200 - 5 10
2 3
and G 0 = 1.8-200 =72 10 kPa. The r m ax value depends on the depth in the soil. At a depth of 10 m, c\
2
can be estimated at 110 kPa. Choosing K0 = 0.5 in Eq. (26) we then find T m xa = [(0.75 1 1 0 s i n 3 2 ) -
2 1 /2 4
( 0 . 2 5 - 1 1 0 ) ] = 34 kPa. This yields yr = 0.034/72 = 4.7 10" and ylyr = 10.6 whence yh = (10.6)[1 -
3 3
0.540-exp(-0.16 10.6)] = 9.6 and G - 7 2 - 1 0 / 1 0 . 6 = 6.8 1 0 kPa. Maximum damping ratio, according
to Hardin & Drnevich, becomes D m ax = (28 - 1.5-log500)/100 = 0.24. With regard to damping we have
yh = (10.6)11 - 0.708-exp(- 0.263-10.6)] = 10.1. Thus, D = 0.24-10.1/11.5 = 0.21. This leads to the
following results for a pile shaft element at 10 m depth:
soil stiffness ks = nG = 21.3 MPa.
soil resistance T m xa = 34 kPa.
3 1 /2
hysteretic damping CH = -0.50.21(7,78-6.8-10 ) = 76 kNs/m
3 1 /2
radiation damping CR = 0 . 5 ( 1 . 8 6 . 8 1 0 ) = 173 kNs/m
186 Deep foundations

the soil. A consequential modification of the soil parameters will h a v e to m a d e . Different


methods of determining the G 0 value are presented on p p . 8 7 - 9 1 .
T h e load/settlement relationship predicted by the use of C A P W A P or S V I D Y N anal-
ysis is generally in good agreement with the results of pile loading tests. Therefore,dynamic
pile loading tests are n o w utilised as a cost-effective alternative to full-scale, static pile
loading tests.

(iii) Geostatical analysis. T h e traditional approach to the bearing capacity is based on


the strength parameters of the soil. In the case of pile foundations this type of analysis,
however, is connected with several difficulties. T h u s , as previously mentioned, pile
installation in itself m a y h a v e an unpredictable influence on the soil properties and it is
therefore hard to know which properties should be used to get the right answer. T h e basis
for a successful application of geostatic analysis obviously has to b e carefully considered
from case to case.
As previously mentioned, the bearing capacity of the pile is built u p of shaft resistance
and tip resistance, the respective contribution of which depends on the soil conditions and
the pile length in the soil. T h e bearing capacity Qjof a single pile in cohesionless soils
is generally expressed in the form:

1
Pp

0 20 40 45
Angle of internal friction '

Fig. 135. Bearing capacity factors Nq and Nc and critical depth for different lplDp ratios. lp = pile length
in soil; Dp = pile diameter. (Meyerhof, 1976).
Deep foundations 187

Qf = ptfAt+fsfAs (205)

w h e r e ptf= c'0tNq<pcr,
a'ot = effective overburden pressure at the pile tip level,
Nq = bearing capacity factor (Fig. 135),
At = pile tip area,
K
fsf= s'osten<fcr,
o'0s = average effective overburden pressure along the shaft,
Ks = earth pressure coefficient (average value),
= angle of friction at the soil/shaft interface,
As - shaft area.
If the soil consists of layers with different characteristics the total shaft resistance is
obtained b y s u m m a t i o n of the contributions given by each separate layer.
B e l o w a certain depth dcr (Fig. 135), the tip resistance does not increase with increasing
depth. T h e limits reached at the critical depth is denoted with pcr a n d / c r.
T h e diagrams given in Fig. 135 show that Nq reaches its m a x i m u m value at about half
the critical depth. O n the other hand, the product Nqo'0 increases until the critical depth
is reached.
T h e choice of the special parameters required for the determination of the shaft resis-
tance ( S a n d Ks) is quite delicate and m o r e difficult than the choice of '. T h e parameters
are very m u c h dependent on the relative density, initial state of stress, shape and diameter
of the pile and the m e t h o d of pile installation. For e x a m p l e , in a n o n - c o h e s i v e soil with
a given value of the angle of internal friction, the bearing capacity of a b o r e d pile is
normally only o n e third to o n e half the bearing capacity of a driven pile.
If the piles are installed in normally consolidated or lightly overconsolidated cohesive
soils, then the tip resistance is usually negligible in comparison with the shaft resistance.
T h e tip resistance will not b e mobilised until the shaft resistance has passed its m a x i m u m
value (Fig. 136) and, furthermore, its value will b e comparatively very small. Therefore,
piles in soft c o h e s i v e soils can b e considered to b e h a v e as friction piles (since the term
friction piles seems to m a k e reference to cohesionless soils, piles in soft cohesive soils
are also referred to as cohesion piles or floating piles).
Also in the case of b o r e d u n d e r r e a m e d piles in stiff clay the shaft resistance will be
mobilised long before the base resistance. T h u s the b a s e resistance develops very slowly
and is s e l d o m fully mobilised until the settlement reaches 1 0 - 2 0 % of the b a s e diameter
(Burland, 1986). O n the other h a n d the shaft resistance is fully mobilised w h e n the
settlement reaches only about 0 . 5 % of the shaft diameter (or about 5 m m ) and then
remains m o r e or less constant with increasing settlement.
T h e total bearing capacity of a floating pile can b e determined on the basis of the
undrained shear strength of the soil according to the relation:
188 Deep foundations

Fig. 136. The maximum shaft resistance in clay is reached at very small relative displacement between
clay and pile, long before tip resistance is mobilised. The influence on shaft resistance of pile material
is negligible. (Torstensson, 1973).

Qf =<xcusAs + cut Nc0At (206)

w h e r e oc= empirical factor w h o s e value depends on soil characteristics, shape of pile, pile
diameter, type of pile (driven or bored), installation method, time to failure, time after pile
installation, etc.
cus = undrained shear strength along the shaft (average value),
cut = undrained shear strength at the pile tip,
= the bearing capacity factor,
Deep foundations 189

As = shaft area,
At = tip area.
T h e bearing capacity factor in clay can be a s s u m e d equal to 9. For driven timber
piles in soft clays can be put equal to 1 for short-term loading and 0.7 for long-term
loading ( a = 0.7 can b e considered representative of the 'creep failure' of the pile, i.e.
the load that leads to excessive creep settlement). For piles with constant cross-section
driven into in soft clays (such as concrete piles), has to be taken s o m e w h a t lower ( 0 . 8 -
0.9 of the values given above). For bored straight-shafted piles and bored u n d e r r e a m e d
piles in stiff clays with cu determined by unconsolidated, undrained triaxial tests on 38
m m diameter samples, a c a n b e assumed equal to 0.45 ( S k e m p t o n , 1 9 5 9 ; B u r l a n d , 1986).
In layered soil with varying characteristics, the total shaft resistance is obtained by
summation of the contributions given by the various layers.
In the case of a pile w h o s e diameter varies linearly with depth and which is installed
in clay with an undrained shear strength increasing linearly with depth (cu = cu0 + kz), the
bearing capacity can be determined by the relation:

1 1
Qf = ccnlp[-(Dph + Dpt )cu0 + - (Dph + 2Dpt )klp] (207)
6

w h e r e Dph = pile h e a d diameter,


Dpt - pile tip diameter,
lp - pile length in clay.
If Dph < Dpt w h i c h happens for the upper pile segment of a spliced timber pile, then,
for the u p p e r pile segment, the cu value has to be reduced with regard to the soil displace-
ment caused by the installation of the lower pile s e g m e n t and the unfavourable shape of
the pile (decreasing diameter in the u p w a r d direction). T h u s , for the upper pile segment
it is r e c o m m e n d e d to put = 0 to a depth of 3 m and = 1/3 below 3 m depth.
Using, as an alternative, effective stress analysis for the determination of the bearing
capacity, this can b e d o n e according to M e y e r h o f (1976):

Qf = As&0s(l - sin ' ) \ / O o s / & c tan = As<fQs (208)

where ' = average effective stress along the shaft,


5
G'C - corresponding preconsolidation pressure.
Again, if the soil consists of layers with different characteristics the total bearing
capacity is obtained by s u m m a t i o n of the contributions given by each separate layer.
/
T h e value of 0 t o b e applied in this case m a y vary from about 20 to about 30. T h e
skin friction <5 can normally b e a s s u m e d equal to '.
Empirical studies of the bearing capacity of piles in soft, normally consolidated or
lightly overconsolidated clays show that decreases with increasing pile length from
190 Deep foundations

Example 24: A 15 m long timber pile with a tip diameter of 150 mm is driven into a clay deposit with
an undrained shear strength c M= 10 + 1. 8z kPa, where z, in m, denotes the depth below the ground surface.
3
The clay is normally consolidated and has a density of 1.6 t/m . The groundwater level is 2 m below the
ground surface and the pore pressure distribution with depth is hydrostatic.
Determine the ultimate load of the pile if the diameter of the pile increases by 8 mm per m from the
pile tip upwards. Use both undrained and drained analysis. In the latter case the angle of internal friction
can be taken as '' - 20. The friction angle in the pile/soil interface <5can be assumed equal to '.

Solution: The pile head diameter Dph = 150 + 8-15 = 270 mm.
Undrained analysis: Qf= 15[(0.27 + 0 . 1 5 ) 4 0 / 2 +(0.27 + 2 0 . 1 5 ) 1.8 15/6] = 220 kN. For long-
term loading can be put equal to 0.7 which yields / = 154 kN
Drained analysis: The effective stress distribution increases linearly from about 20 kPa (= 2-1-9.81)
at the ground surface to 31 kPa (= 1.6-2-9.81) at 2 m depth and from there linearly to 108 kPa (= 31 +
/
13-0.6-9.81) at 15 m depth. Since the clay is normally consolidated o {j&c - G'^G'V = 1.
We find:
Qf= [2(0.27 + 0.254)(20 + 31)/4 + 13-(0.254 + 0.15)(31 + 108)/4](1 - sin 20)-tan 20 = 147 kN.

about 0 . 5 - 0 . 2 5 w h e n lp < 15 m to about 0 . 2 5 - 0 . 1 for very long piles. M o r e o v e r , is higher


for piles with a d i a m e t e r decreasing with depth than for piles with constant diameter.

(i v) The pressuremeter method. T h e tip resistance of the pile can b e calculated on the basis
of p r e s s u r e m e t e r test results according to the s a m e p r i n c i p l e as p r e v i o u s l y s h o w n (pp. 145
- 1 4 8 ) . In this c a s e w e h a v e :

(a)

(b)

0
0 2 4 6 8 10
Depth of e m b e d m e n t d/Dp

Fig. 137. values vs. pile embedment for driven piles (Bustamente et al, 1981).
(a) Very dense sand and gravel.
(b) Medium dense sand, firm silt and firm clay.
(c) Loose silt, clayey sand and soft clay.
Deep foundations 191


Ptf=
+
* (209)

where '0t = effective overburden pressure at the pile tip level,


*Pi = Pi~ P{) - net limit pressure,
p0 = horizontal in situ stress at the pile tip level,
K = parameter w h o s e size depends on soil characteristics, m e t h o d of pile instal-
lation and length of pile e m b e d d e d in soil.
T h e m a x i m u m values for driven piles (Fig. 137) are about 2 0 % higher than those for
bored piles.
If the soil is heterogeneous an equivalent limit pressure is determined as the geometrical
m e a n value of pl from \ .5Dp below, to \ .5Dp above, the pile tip, i.e.:

Pie ^yJpixPllPn '"Pin (210)

T h e correlation between shaft resistance/ v ^and the net limit pressure *p{ can be obtained
from Fig. 138. T h e shaft resistance depends not only on the net limit pressure but also on
the m e t h o d of pile installation and the pile material.

150

120 ( a ) _
*
* /
Ou,
Jt
/ /
90
/
' / _ - v u/

i 60 / /
(d)
/

30

0.5 1 1.5 2 2.5


Net limit pressure *pt, MPa

Fig. 138 (a-c). Shaft resistance vs. net limit pressure.


Driven piles: (a) Very dense sand and gravel, (b) Medium dense sand and gravel, (c) Silt, silty sand, clay
and clayey silt. (Bustamante etal., 1981). (d) Non-cohesive soil. (Sellgren, 1985).
Non-displacement piles: (c) Any soils. (Baguelin et al, 1978)
192 Deep foundations

Example 25: Determine the bearing capacity of a concrete pile with a cross-sectional area of 0.25 m
by 0.25 m driven to a depth of 8 m in a sand deposit whose properties, determined by pressuremeter tests,
are given in the figure shown below. The groundwater level is at 1 m depth. The unit weight of the sand
3 3
is 18 k N / m above and 11 k N / m below the groundwater table.

Pressuremeter modulus Ept L i m i t p r e s s u r e pt, M P a

0 10 20 0 2 4
16.5

,
_ _
/ 4 1)
1

r j 16

Solution: The point resistance is governed by the bearing capacity factor K~ 2.7 and the pressure limit
ple - 0.5 MPa. The shaft resistance is governed by the average of pt from the pile head to the tip which
is equal to (2.3 + 0.8 + 0.8 +0.6 + 1.2 + 1.8 +1.0 + 0.5)/8 = 1 . 1 MPa. From Fig. 133, curve (b), we find
the average v a l u e ~ 80 kPa, and from curve ( d ) , / ^ ~ 110 kPa.
2
U s i n g / 9 /= 80 kPa, the bearing capacity of the pile becomes Qf= 0.25 ( 18 + 7 11 +2.7-500) + 804-0.25-8
= 730 kN, while, using fSf= 110 kPa, the bearing capacity becomes Qj - 970 kN.

K n o w i n g the values of a n d / ^ t h e bearing capacity is obtained by the relation:

Qf=(Ot + K*Pi)At+f5fAs (211)

w h e r e At and As as before (cf. Eq. 205).

(v) Empirical methods. T h e soil resistance to pile driving reminds of the soil resistance
obtained in different penetration methods used in soil investigations. It is therefore but
natural that m a n y attempts have been m a d e throughout the world to establish relations
between the resistance to penetration of different sounding tools and the bearing capacity
of piles. S o m e of these m e t h o d s will b e presented.
For driven piles, M e y e r h o f (1976) proposes the following relation b a s e d on the results
2
of S P T (Qf in kN and At and As in m ) :

/ = 4 0 0 ^ ^ + 2 ^ 3 ^ (212)

w h e r e 3 0 represents the /V 3 0 count at the pile tip level and N30s the average / V 3 0 count
along the shaft of the pile.
Deep foundations 193

Dcourt (1982) suggests a corresponding relation based on S P T results (Qf in kN and


2
At and As in m ) :

Qf=KN30tAt + 1 0 ( t f 3 0 / 3 +1)A, (213)

w h e r e = 2 0 0 for clayey silt, 300 for sandy silt and 4 0 0 for sand.
T h e tip resistance should b e limited to a m a x i m u m of about 10 M P a .
B u s t a m a n t e and Gianeselli (1982) p r o p o s e the following relation based on the results
of C P T investigations:

Qf=qcpkcAt+fcAs (214)

w h e r e qcp = average point resistance in C P T from 1.5D t above the pile tip to 1.5D, below
the pile tip,
kc = parameter w h o s e size depends on pile type and soil characteristics,
fc = qcs/cx,
qcs = average point resistance along the pile shaft.
T h e kc and values r e c o m m e n d e d by B u s t a m a n t e & Gianeselli (1982) for two pile
categories, n a m e l y category I A (including plain bored, m u d - b o r e d , hollow auger-bored,
cast screwed piles, micro-piles grouted under low pressure and piers) and category IIA
(including driven precast piles, prestressed tubular piles and j a c k e d concrete piles) are
given in Table 24.
A very c o m p l e t e literature survey of the various m e t h o d s of analysis based on C P T
results, including pile installation effects, has been presented by Van I m p e (1991).
For bored piles in cohesionless soil, Berggren (1981) presented a d i a g r a m for the
determination of the critical pile point load (the creep load) b e l o w which settlement can

TABLE 24.
Values of bearing capacity factor kc and coefficient for pile categories IA and IIA.

Type of soil <7 c(MPa) /cm


k P
( a)
IA IIA

Silt and loose sand <5 0.40 0.50 60 35


1
Compact to stiff clay >5 0.45 0.55 60 35 ( 8 0 )
and compact silt
Moderately compact 5 - 12 0.40 0.50 100 80(120)
sand and gravel
Compact to very com- > 12 0.30 0.40 150 120 (150)
pact sand and gravel
Soft chalk <5 0.20 0.30 100 35
Weathered to frag- >5 0.20 0.40 150 120 (150)
mented chalk

The values in parenthesese refer to very careful pile installation, involving minimum disturbance of soil.
194 Deep foundations

Example 26: Pile loading tests were carried out on concrete piles with square cross-section and a side
length of0.235 m. Three piles, driven to depths of 15 and 19 m, were loaded to failure. Results of piezo-
cone soundings showed that the soil consisted of clay underlain by sand at 5 m depth. A typical result
of the piezo-cone sounding and the pile tip level of the three tested piles are presented below. Calculate
the bearing capacity of the piles on the basis of the penetration resistance according to Bustamente &
Gianeselli.

Point resistance Pore pressure S l e e v e friction


MPa MPa MPa
0 10

Solution: The CPT point resistance at the pile tip qc ~ 8 MPa for all the piles tested. As regards the pile
shaft we find qc 0.6 MPa to a depth of 5 m, qc 6 MPa, on the average, from 5 m to 15 m depth and
qc ~ 7.5 MPa from 15 m to 19 m depth. The piles belong to Category IIA. Since there is no value for
clay, the contribution to the bearing capacity of the clay layer will be based upon the value for silt.
The bearing capacity of the different piles can now be calculated.
2
Pile PI: Qf= 0.5-8-0.235 + ( 5 0 . 6 / 6 0 + 10-6/100)40.235 = 0.88 MN
Piles P2 and P3: Qf = 0.88 + 4 ( 7 . 5 / 1 0 0 ) 4 0 . 2 3 5 = 1.16 MN
According to the loading tests we have:
P i l e PI: / = 0 . 6 8 MN
P i l e P2: Qf = 1.07 M N
P i l e P3: Qf= 1.16 M N
The ratio of calculated to observed ultimate pile loads varies from 1.0 to 1.3. The agreement between
theory and reality in this case is quite satisfactory.
Deep foundations 195

25 30 40 50 60
Design parameter Ns

0 5 10 15
Relative depth of penetration lxIDp

Fig. 139. Top: Relation between sounding resistance and design parameter Ns. Range of validity for
different penetration methods shaded. Bottom : Critical tip resistance vs. pile penetration depth for different
values of Ns (After Berggren, 1981).

be a s s u m e d to increase linearly with increasing load (Fig. 139). In the top diagram, a
design parameter Ns is determined on the basis of the results of different penetration tests.
Then knowing Ns, the critical tip r e s i s t a n c e p t c is obtained from the b o t t o m diagram and
the corresponding pile tip settlement from Fig. 140.
196 Deep foundations

60

201 I I I I I I I I I I
0.05 0.10 0.15
Ratio of elastic settlement limit to pile diameter

Fig. 140. Pile tip settlement at critical point resistance as function ofNs and uniformity coefficient Cv.

In m a n y cases, it can b e economically advantageous to i m p r o v e the bearing capacity


of an in-situ pile by d y n a m i c preloading. A typical e x a m p l e of d y n a m i c preloading is the
Franki pile. Special equipment, suitable for d y n a m i c preloading of bored piles, has been
in use now for a long time past and by this e q u i p m e n t it has been possible to considerably

Fig. 141. Equipment used for dynamic preloading of bored piles, (a) 3.8 t hammer, (b) Guide casing,
(c) Measuring cable, (d) Casing for work, (e) Damper, (f) Measuring cell, (g) Pile point of steel.
Deep foundations 197

increase the bearing capacity of the piles, thereby reducing the required length of the piles
and in c o n s e q u e n c e leading to cost-effective solutions.
T h e r e c o m m e n d e d procedure of d y n a m i c preloading is as follows (Berggren &
Bengtsson, 1985). A special set-up for d y n a m i c preloading is utilised, consisting of a 3.8
t h a m m e r , a guide casing and a 'pile point' equipped with accelerometer, strain gauges
and displacement transducers, Fig. 1 4 1 . T h e diameter of the 'pile point' should not be
below 0.8 times the pile diameter. D y n a m i c preloading is carried out in several steps by
gradually increasing the drop height of the hammer, e.g. 0.2 m, 0.5 m, 1.0 m, 2.0 m and
4.0 m, and so that the ratio of remaining plastic settlement spi to m a x i m u m settlement s m a x
caused by the impact of the h a m m e r decreases in each step. T h e increase in drop height
can take place when the ratio spi/smx< 0.2 (or possibly 0.3).

Example 27: Determine the critical point load of a bored pile, 0.9 m in diameter, by the aid of the graphs
shown in Fig. 139. The pile tip is at 1 m depth in gravel with a cone resistance qc - 8 MPa. The gravel
is overlain by a soft clay deposit, 10 m in thickness. The groundwater level is situated at a depth of 2
3 3
m. The unit weight of the clay is 16 k N / m and of water-saturated gravel 21 kN/m .

Solution: From Fig. 139 we find the design value Ns - 35. The depth of penetration into the gravel is
1 m which yields l\IDp = 1 . 1 . From the bottom diagram we find ppj~ 22'0 = 22 (2-16 + 8-6 + 1-11)
3 2
= 2-10 kPa. Thus, the critical point load is equal to 20.9 /4 = 1.3 MN. The elastic settlement limit
according to Fig. 140 is around 0.09 to 0.14 m depending on theuniformity coefficient of the gravel.

4. SETTLEMENT OF AXIALLY LOADED SINGLE PILES

W h i l e the bearing capacity is decisive of the ultimate state, the s e t t l e m e n t a b o v e all the
differential settlementis decisive of the serviceability state. T h e possibility of predicting
pile settlement with satisfactory accuracy is limited. T h e r e are several reasons for this.
Pile installation generally alters the stress/strain conditions in the soil. In soft, sensitive
clays disturbance effects m a y give rise to a decrease of the preconsolidation pressure
followed by consolidation. In loose cohesionless soil pile driving causes compaction
which will i m p r o v e the settlement characteristics. Driven piles will b e subjected to
residual stresses which m a y influence the load/settlement relationship. M a n y other
factors not mentioned here can increase the difficulties of settlement prediction. Therefore,
in practice the permissible pile load is mostly coupled with a certain factor of safety
against pile failure.
A s a general rule it can b e stated that in-situ methods for determination of the deformation
characteristics are preferable in the case of cohesionless soils w h e r e a s laboratory meth-
ods are generally adaptable in the case of cohesive soils. Stress-wave m e a s u r e m e n t s and
analyses can b e used not only for determination of the bearing capacity of piles but also
for prediction of pile settlement under working load condition.
T h e column and wall loads in a building are mostly too large to b e carried by single
piles and, therefore, pile groups h a v e to b e installed. E v e n if the settlement of a single pile
198 Deep foundations

can be predicted with satisfactory accuracy, the settlement of a pile group can b e entirely
different. This is important and has to b e recognised.
T h e settlements occurring in the serviceability state are limited which, in the case of
friction piles, m e a n s that load will be carried m o r e or less merely by shaft resistance. T h e
tip resistance will either not be mobilised at all or will b e of negligible m a g n i t u d e . On the
other hand, piles driven to refusal, or with the pile tip in layers of high bearing capacity,
such as hard till, will carry the service load almost entirely by tip resistance. D y n a m i c
preloading of the pile tip, as is sometimes utilised with bored piles, represents another
case w h e r e the service load is carried mainly by tip resistance.

(i) The pressuremeter method. T h e settlement of a friction pile, in the serviceability state
w h e n the shaft resistance is not yet fully mobilised, will mainly b e governed by the shear
m o d u l u s of the soil. T h e foremost aim of the pressuremeter is to determine the shear
m o d u l u s of the soil and the results of pressuremeter tests therefore serve as a reliable basi s
for pile settlement calculations. A strict analysis based on the results obtained by the
M n a r d pressuremeter, taking into account both shaft and tip resistance, was carried out
by Cassan (1966). In C a s s a n ' s analysis it is a s s u m e d that the shaft resistance fs = Bs(z)
and the tip r e s i s t a n c e p p = Sp/Dp where s(z) and sp represent pile settlement at, respecti vely,
depth and pile point level and and represent soil resistance. C a s s a n ' s analysis, later
modified by Sellgren (1981), yields the following relation:

4 \ + [/()]1<() Q_ _ Q
= (215)
nDp + ) \-QIQf \-QIQf


where
EpDp
= GP/L$ with Lq = 0.3 m for driven piles and L q = 0.9 m for bored piles,
= 12GS with GS = 3GPR for driven piles and GS - GPR for bored piles (GPRS =
average of GPR along the pile shaft and GPRT = GPR at the pile tip),
EP = m o d u l u s of elasticity of pile,
Dp = pile diameter.
Using instead of GPR the pressuremeter m o d u l u s EPR = 2Gpr(l+vs) w e find for driven
1 1
piles = 1.25p r n r and = 1 3 . 5 E P R and for bored piles = 0 i r r and =4.5EPR
For piles with square cross-sectional area, nDP/4 should b e e x c h a n g e d for bp w h e r e
bp is the width of the pile.
A comparison of the values for bored piles to those for driven piles is m a d e in Table
25.
In m a n y cases bored piles are installed through w e a k soil into hard b o t t o m layers, such
as till. T h e pile load is then carried by tip resistance and shaft resistance can b e neglected.
Deep foundations 199

TABLE 25
Examples of values for bored and driven piles, assuming Ep = 30 GPa.

lp (m) Epr (MPa) (m m/M )


bored piles driven piles
Dp=03m Dp =0.6 m Dp=()3 m Dp - 0 . 6 m

5 10 34.2 16.8 12.2 5.8


10 22.0 10.5 8.6 3.8
20 14.6 6.5 6.8 2.7

5 50 7.7 3.6 3.3 1.4


10 5.8 2.5 2.9 1.1
20 5.0 1.9 2.8 1.0

For these end-bearing piles the settlement is obtained by the relation:

9APEPR D0

w h e r e Q = Ap(q - ^ ) = net increase of pile tip load,


D0 - reference diameter = 0.6 m ,
Xd and are shape factors ( = 1 for circular-cylindrical piles and 1.12 and 1.1,
respectively, for piles with square cross-section, cf. Table 22, p . 146,
ad and at are rheological coefficients according to Table 2 1 , p . 146,
Dp > D0. (If Dp < D0 it should b e put equal to D0).

Example 28: Determine the settlement of the pile, the bearing capacity of which was determined in
Example 25, under a permissible load of one third of the calculated failure load (730 kN). The pile has
a cross-sectional area of 0.25m by 0.25 m and a length of 8 m. The results of pressuremeter investigations
are shown in the figure below. The elastic modulus of the pile is 30 GPa.

P r e s s u r e m e t e r m o d u l u s Epr, MPa L i m i t p r e s s u r e Pj, M P a


200 Deep foundations

Solution: The average value of the pressuremeter moduli along the pile shaft is Epr = 6.7 MPa. At the
pile point, the pressuremeter value is taken as the harmonic mean of 2.1 and 16.2 MPa, i.e. Epr = 3.7
MPa. This yields = 1.25-6.7 = 8.4 MPa/m and = 13.5-3.7 = 50.2 MPa. The equivalent pile diameter
3 1 /2 1
Dp = 40.25/ = 0.318 m, whence = 2-[8.4/(30-10 0.318)1 = 0.0593 m" .
The value according to Eq. (214) becomes:
50.2
1 1+- tanh(0.0593- 8)
0.0593-30-103-0.318
a- - 3 : 0.0139 ra/MN,
0.25 50.2 + 0.0593 30 - 1 0 0.318 tanh(0.0593 8)
from which:
s = 0.0139(0.73/3)/(1 - 1/3) = 0.005 m

(ii) Empirical methods. T h e settlement of friction piles can also b e d e t e r m i n e d with good
a p p r o x i m a t i o n by the aid of the d i a g r a m given in Fig. 142. T h e d i a g r a m is valid for
calculation of settlement up to half the ultimate load. T h e settlement is obtained by the
relation:

s=fs/Ks (217)

w h e r e fs = a v e r a g e frictional resistance along the shaft,


K = a v e r a g e pile d i s p l a c e m e n t m o d u l u s .
T h e Ks v a l u e is d e t e r m i n e d from Fig. 142. T h e shear m o d u l u s of the soil Gs can be

5 5
10 2*10

Fig. 142. Diagram for determination of settlement of friction pile (Hansbo & Bengtsson, 1979). Legend:
Ep= Elastic modulus of pile; Dp= pile diameter; lp= pile length in soil; G= shear modulus of soil; K =
pile displacement modulus.
Deep foundations 201

a s s u m e d e q u a l to 150c M for c o h e s i v e soils and e q u a l to Gpr (= Epr/2.6) for cohesionless


soils.
In c o h e s i v e soils, the settlement at 8 5 % of the u l t i m a t e load and at failure can be
estimated at respectively 2.4 and 4 times the the settlement at half the u l t i m a t e load
(Torstensson, 1973).
In c o h e s i o n l e s s soil, the settlement at 7 5 % of the u l t i m a t e load and at failure can be
estimated at respectively 2 a n d 5 times the settlement at half the u l t i m a t e load (Sellgren,
1985).

Example 29: In order to investigate the load vs. settlement behaviour of piles in soft clay, test loading
was carried out on two piles, one with a diameter of 0.33 m and the other with a diameter of 0.5 m, both
driven to a depth of 25 m through a top layer of peat, 3 m in thickness, underlain of a 10 m thick layer
of volcanic clay on alluvial, silty clay (Simonini & Soranzo, 1988). A finite element analysis was carried
out, based on elastic properties determined by means of unconfined compression tests at half failure
load. The undrained shear strength was determined by means of field vane tests. The mean value of the
undrained shear strength was 13 kPa in the volcanic clay and 26 kPa in the alluvial clay. The result of
the test loading and the finite element analysis is shown below.
Determine the load vs. settlement behaviour of the piles by means of Eq. (217) and Fig. 142.

Pile load, kN Pile load, kN

Solution: The failure loads of the two piles is determined according to Eq. (207). The contribution to
the bearing capacity of the peat layer is neglected since the load/deformation characteristics of peat are
quite different from those of the clay.
For the pile, 0.33 m in diameter, we have (a assumed equal to unity):
2
Qf= 0.33(10 13 + 12-26) + 9260.33 /4 = 478 kN
and for the pile, 0.5 m in diameter
2
Qf= 0.5(10 13 + 12-26) + 9260.5 /4 = 740 kN
The mean value of the shear modulus along the shaft of the piles can be estimated at:
Gs = 1 5 0 ( 1 0 13 + 12-26)/22 = 3000 kPa
4
Assuming that the elastic modulus of the pile is 30 GPa we find Ep/Gs = 1 0 . For the pile, 0.33 m in
diameter , we have lplDp = 22/0.33 = 67 which, according to Fig. 142, yields KsDpIGs = 0.33 and thus
Ks = 3.0 MPa/m
The settlement at half the failure load for the pile, 0.33 m in diameter, ( g = 239 kN) becomes:
3 3
s = 239/(310 0.3322) = 3.5-10~ m
and at 85% of the failure load ( g = 406 kN)
3 3
s = 2.4-3.5-10- = 8 . 4 - 1 0 - m
202 Deep foundations

For the pile, 0.5 m in diameter, w e have lplDp = 44 whence KsDp/Gs = 0.41. In this case we find Ks
= 2.5 MPa/m. Thus the settlement at half the failure load( = 370 kN) is equal to:
3 3
s = 370/(2.5 10 0.522) = 4 . 3 1 0 - m
and at 85% of the failure load ( = 629 kN)
3 3
s = 2.4-4.3-10- = 10.3-10~ m
Assuming a safety factor of 3, the calculated and measured results are as follows:

Pile diameter, m Working load, kN Settlement, mm


Approximate FEM Measured

0.33 160 2.3 2.5 1.4


0.50 245 2.9 2.4 1.5

Example 30: Determine by the approximate method, used in Example 30, the settlement of the pile
according to Example 28.

Solution: We have IJD? = 8/0.318 = 25. The Gs value can be put equal to Gpr which yields Gs = 6.7/2.6 =
4
2.6 MPa. Thus EJGS - 30-103/2.6 =1.15 10 which gives ^=0.46. Hence we find Ks=0.46-2.6/0.318
= 3.8 MPa. The settlement for a load of one third of the failure load 800 kN becomes:
s = 0.8/(3-3.8) = 0.007 m
which is 14% above the value obtained in Example 28.

In U K , large-diameter u n d e r r e a m e d bored piles are n o w the m o s t widely used type of


pile in stiff clay. By experience, u n d e r r e a m e d piles settle m o r e u n d e r w o r k i n g load than
straight-shafted bored piles (Burland, 1986). Moreover, the practical consequences of the
underreaming operation m a y alter the settlement characteristics of the clay in an
unpredictable w a y and, therefore, m a k e settlement prediction unreliable. A conservative
estimate of the settlement s under w o r k i n g load is obtained by the relation (Burland,
1986):
s=0.02Dpb/Fs (218)

w h e r e Dpb = the b a s e diameter of the pile,


Fs = QflQ = the factor of safety.

5. LATERALLY LOADED PILES

5.1 Ultimate resistance

W h e n the pile is subjected to a horizontal load, it will m o v e either m o r e or less in parallel


(if very short and restrained at top) or around a point of rotation until the counteracting
earth pressure is distributed in such a way that the equilibrium condition b e c o m e s
satisfied. Obviously, the earth pressure distribution along the pile ( and, where appropriate,
the point of rotation) will depend on the size of the horizontal force applied and its point
of action, possible restraint upon the pile head and the length and strength of the pile. T h e
Deep foundations 203

Fig. 143. Pile movement and ultimate earth pressure distribution against horizontally loaded short pile
in cohesionless soil. Restrained pile.

lateral limiting capacity will b e reached either w h e n the earth pressure reaches its upper
limit or w h e n the pile is broken in bending.

(i) Non-cohesive soils. Investigations on model piles, loaded laterally in a centrifuge


(Barton, 1982), indicate that the limiting earth pressure against the pile, below a depth
of about 1.5 pile diameters, follows the relation K^Dp\, w h e r e Kp is the passive earth
2
pressure coefficient [Kp= t a n ( 4 5 + 072)], ' is the effective overburden pressure and
Dp is the pile diameter. Nearest to the ground surface, the limiting pressure was found to
b e about KpDpa'v (Fig. 143). This assumption differs s o m e w h a t from the assumption
m a d e b y B r o m s (1964) w h e r e the limiting pressure is put equal to S A ^ a ^ i n s t e a d of Kj-o'v
Let us a s s u m e that the point of rotation of a short pile subjected to a horizontal force
at height e a b o v e the ground surface is at depth lp-2a (Fig. 144a). Let us further assume
that the limiting earth pressure against the rotating pile is distributed according to Brinch
Hansen (1953) which can b e simplified according to Fig. 144c (cf. Fig. 205).

Fig. 144. Pile movement and ultimate earth pressure distribution against horizontally loaded short pile
in cohesionless soil. Unrestrained pile, (b) Probable earth pressure distribution, (c) Assumed earth
pressure distribution.
204 Deep foundations

Fig. 145. Nomogram for determination of lateral capacity of short piles in cohesionless soil.

T h e lateral capacity of an unrestrained, short pile can now be calculated from the
conditions of equilibrium:

2
Hf = Kp YD/-^--a(lp--2)] (219)

2
9 a j2 (L-P 3a)
Hfe = [( - - ) - 3 ) (220)

Hfis obtained by elimination of a in the equation system.


For a restrained, short pile (Fig. 143), the lateral capacity will b e dependent on
whether the pile is yielding or not.
If the pile is not yielding (no hinge) the condition of equilibrium is satisfied by the
relation:
2
I

Hf - KpjDp-^ (221)

while, if it is yielding (one hinge with a yielding m o m e n t Mpf, Eq. (221) takes the form:
Deep foundations 205

Fig. 146. Assumed limiting earth pressure distribution against long piles in cohesionless soil.

2 (L - 3) 2 a 2?
M pf = Kp YDp[^-^-a(lp--) ] (222)

In this case, a is expressed in M ^ a n d introduced into Eq. (219).


Accordingly, the lateral capacity of a short pile can b e obtained from the n o m o g r a m
shown in Fig. 145.
In the case of longer piles, the limiting horizontal force is determined by the fact that
it leads to bending failure of the pile (Fig. 146). A s s u m i n g that the pile is broken in
bending at d e p t h / ( o n e hinge), the conditions of m a x i m u m m o m e n t (transversal force
equal to zero) and equilibrium yield:

Fig. 147. Nomogram for determination of the lateral capacity of long piles in cohesionless soil.
206 Deep foundations

Example 3 1 : In a test carried out by Koskinen (1991), the lateral bearing capacity of a steel pile, 273
mm in diameter, embedded in sand was investigated. The sand had the following characteristics:
3
internal angle of friction '= 37 and unit weight / = 9 k N / m . The length of embedment of the pile
was lp = 4 m and the lever arm of the horizontal load e - 0.8 m. The yield moment Mpj of the pile was
assumed equal to 92.7 kNm. Determine the lateral capacity on the basis of Eqs. (219-220) and (223-224).

Solution: We have e ID = 0.8/0.273 = 2.93 and lpID =4/0.273 = 14.65. The earth pressure coefficient
2 2 4
Kp = tan (45 + 3 7 7 2 ) = 4.02. The value Mpfl(Kp YD ) = 114.7. From the diagram, Fig. 147 (the pile
2 3
is assumed to be broken in bending), we find H^k/yD^) - 17 whence Hf= 174.02 90.273 = 50
kN (an exact solution yields Hj= 49.9 kN). From the diagram, Fig. 145 (the pile is not broken), we find
2 3
Hfl(Kp YDp ) 20 whence Hf= 59 kN (an exact solution yields Hf= 58.6 kN). In reality the pile was
loaded up to 60 kN which can be assumed to correspond to the failure load (the occurrence of failure
is not distinct). The horizontal displacement of the top of the pile at = 54 kN was 0.13 m.
2
f
Hf - KpjDp (223)

2 J
Hfe = Mpf-K rDp - (224)

If the pile is restrained (two hinges), Eq. (223) will be replaced by:

2 J
Mpf=Kp YDp - (225)

Hf,which is obtained as a function of Mpjby e l i m i n a t i n g / i n the equation system, can


be obtaind from the diagram shown in Fig. 147.

(ii) Cohesive soils. In the case of cohesive soils the lateral earth pressure against a short,
restrained pile varies normally as shown in Fig. 148 (left). It can be a s s u m e d (Fig. 148,
right) that the lateral resistance, below a depth of 1.5 pile diameters, is equal to 9 times
the undrained shear strength of the soil multiplied by the pile diameter (Broms, 1964).
If the undrained shear strength cu is constant with depth the conditions of equilibrium
of a short, unrestrained pile at failure will b e fulfilled by the relations:

Hf = 9cu Dp[(z0 - 1.5D p) - (lp -z0)] (226)

Z 5 D 2
Hfie + zo) = 9 c u d / - [ ^ + ^ - ] (227)

Hfis obtained by eliminating z 0 in the equation system.


For a short, restrained pile, the lateral capacity of an unyielding pile can b e calculated
by the relation:
Deep foundations 207

9c,pn
up

%cj}p-\2cj>r

Fig. 148. Limiting earth pressure distribution against a short restrained pile in cohesive soil (left) and
assumed limiting earth pressure against unrestrained, short pile (Broms, 1964).

Hf = 9cuDp(lp-\5Dp) (228)

If the pile is yielding (one h i n g e with a yielding m o m e n t Mpj, Eq. (227) takes the form:

(229)

In this case z 0 is expressed in and introduced into Eq. (226).


T h e lateral capacity according to these latter relations is given in n o m o g r a p h i c form
in Fig. 150.
For longer piles w h e r e the bending capacity of the pile is decisive of its lateral capacity,
the conditions of equilibrium of an unrestrained pile (one hinge) b e c o m e :

Hf=9cuDp(f-l.5Dp) (230)

Fig. 149. Assumed limiting earth pressure distribution against a long pile in cohesive soil.
208 Deep foundations

Fig. 150. Nomogram for determination of lateral capacity of short piles in cohesive soil.

Fig. 151. Nomogram for determination of lateral capacity of long piles in cohesive soil.
Deep foundations 209

2
(f-\.5D)
pJ
Mpf = Hf(e +f)-9cuD/ ^ (231)

If the pile is restrained (two hinges), Eq. (231) is replaced by:

2 2
f -{\.5DD)
Mpf = 9cuDp (232)

/ ^ i s obtained as a function of M ^ b y e l i m i n a t i n g / i n the equation system.


T h e result is given in n o m o g r a p h i c form in Fig. 1 5 1 .
T h e lateral resistance of a pile in soft cohesive soil is generally g o v e r n e d by the dry
crust strength which is considerably higher than the strength of the underlying clay. In
such a case the equilibrium conditions of the laterally loaded pile will h a v e to be corrected
in accordance with the variation with depth of the shear strength.

5.2 Deflection

A s s u m i n g that a laterally loaded pile is elastically supported, the deflection y at depth


can b e obtained by the differential equation:

4
dy psDD
J L
A = ^ (233)
dz Eplp

w h e r e Eplp = b e n d i n g rigidity of the pile,


ps - lateral reaction against the pile deflection,
Dp = pile diameter (width).
Eq. (233) is valid only if the pile is straight e n o u g h for the a s s u m p t i o n s behind the
elementary b e a m theory to b e satisfied. For the equation to be solved analytically it is
required that the b e n d i n g rigidity of the pile and the lateral reaction against pile deflection
are i n d e p e n d e n t of depth or follow s o m e simple well-defined relationship. For e x a m p l e ,
assuming a long pile with constant bending rigidity E^p subjected to a lateral force H at
level e a b o v e the g r o u n d surfaceand a corresponding soil reaction ps = Klsy, w h e r e Kls =
coefficient of lateral subgrade reaction, w e find the deflection:

2Hk

y = |>/c(cos Kz - sin ) + cos ] e x p ( - K 2 ) (234)

w h e r e kis = Kls Dp = m o d u l u s of lateral s u b g r a d e reaction


210 Deep foundations

Transverse force, kN Bending moment, k N m Displacement, mm


-5 0 5 10 0 2 4 6 0 10 20

Fig. 152. Internal forces and deflection of a concrete pile subjected to a horizontal load H= 10 kN at
a height of 0.2 m above ground surface. Pile with square cross section, 0.35m in width. Modulus of
lateral subgrade reaction kls - 40 MPa. Elastic modulus of pile Ep - 30 GPa. (= 5.7).

_ 4 / KisDp _ 4 / k[s
k
~ \ a e p I p ~ \ a e p I P

T h e b e n d i n g m o m e n t is governed by the relation:


M - [eK(cosKz + sinKz) + SINK*z] e x p ( - K 2 ) (235)

and the transverse force by:

T=H [COSKZ - (2 + l ) s i n r c ] e x p ( - x z ) (236)

T h e pile can b e considered as long (Fig. 152) w h e n > , w h e r e lp is the e m b e d d e d


pile length.
In m o s t cases it is not possible to find a simple correlation b e t w e e n the coefficient of
subgrade reaction and the depth. However, k n o w i n g the correct values and distribution
of subgrade reaction w h i c h is the m o s t difficult part in the analysis, the p r o b l e m can be
easily solved by the aid of c o m p u t e r s .
T h e coefficient of subgrade reaction in cohesive soils can b e expressed as a function
of the u n d r a i n e d shear strength cu of the soil. T h e value of kls in l o n g - t e r m loading is
affected b y creep and consolidation p h e n o m e n a . In this case, empirical studies h a v e
s h o w n that a g o o d approximation is obtained by a s s u m i n g kls = 20cu. In an investigation
on m o d e l piles in clay (Bergfelt, 1964), this value of fc/5has b e e n found to yield theoretical
values of the b u c k l i n g load that are in good a g r e e m e n t with observations. T h e short-term
Deep foundations 211

value of kis can b e a s s u m e d equal to 8 0 c u . Of course, the kls value thus obtained can only
be applied w h e n the lateral soil r e a c t i o n p s = kls y is b e l o w the pressure leading to failure.
In cohesionless soil the coefficient of subgrade reaction according to Terzaghi (1955)
can be a s s u m e d to increase linearly with depth according to the relation kls = nhz, w h e r e
nh can b e estimated at 7 5 / t o 2 2 5 / f o r loose sand and at 7 5 0 / t o 1 5 0 0 / f o r dense sand
w h e r e / i s the effective unit weight of the soil.
In practice, a possible way of determining the m o d u l u s of lateral subgrade reaction is
by using the results of pressuremeter tests. According to the correlations between
settlement and pressuremeter m o d u l u s Epr previously given, Eq. (152), p. 146, w e find:

9F
k ls = (237)

2 ( 2 . 6 5 ) + 1.5
if Dp < 0.6 m,
9E
kls = ^ (238)
a
2(DQ/Dp)(2.65Dp/D0) + 15a

i f D p > 0 . 6 m,
w h e r e is the rheological coefficient at depth (Table 2 1 , p. 139),
D0 = 0.6 m.
T h e kls value determined accordingly can only be applied w h e n the lateral soil reaction
= s
Ps kisy * b e l o w the critical pressure p c r In case the lateral reaction exceeds the creep
pressure, the lateral subgrade reaction b e t w e e n creep pressure and failure can b e assumed
equal to half the values determined by Eqs. ( 2 3 7 - 2 3 8 ) .
N e a r the ground surface, the lateral subgrade reaction is less than at depth because of
ground h e a v e . T h u s Eqs. ( 2 3 7 - 2 3 8 ) are only valid below a critical depth zc w h i c h varies
from about 2Dp for cohesive soils to about 4Dp for granular soils. A b o v e the critical depth,
the kh value should b e replaced b y kls where:

. \+zlz c
= (239)

However, if soil h e a v e is prevented, for e x a m p l e by the pile cap, can b e a s s u m e d


equal to unity
Briaud (1992) suggests that the m o d u l u s of lateral s u b g r a d e reaction for non-
displacement piles b e estimated by the relation:

kls = Epri) + EprR (240)

where and EprR represent the first load and reload preboring pressuremeter moduli,
respectively.
212 Deep foundations

For full-displacement piles he suggests:

kls = 2EprR (241)

Example 32: Determine the horizontal displacement at the ground surface and at 1 m depth below the
ground surface of a steel pile with square cross-sectional area, 0,10 m in width, installed in clay with
constant shear strength cu - 30 kPa, if the pile is subjected to a horizontal load of 5 kN at a height of 0.5
m above the ground level.

4 4
Solu non : For a steel pile we have Ep=210 GPa. The moment of inertia islp=0.1 /12 = 8.3 l O ^ m . Assuming
6 6 1 /4 1
that kls = 20cu we find = [600/(4-210-10 8.31- )] = 0.542 m" .
The lateral movement becomes:
= 0 (ground level)
y = (250.542/600)(0.50.542 + 1) = 0.011 m
= 1 m
y = (2-5-0.542/600)-[0.5-0.542-(cos 0.542 + sin 0.542) + sin 0.542]exp(- 0.542) = 0.0047 m

6. BUCKLING OF PILES

T h e danger of buckling (elastic instability) of piles e m b e d d e d full length in soil is


normally negligible except for small-diameter piles in very soft cohesive soil. Therefore,
it is seldom required to check whether the permissible pile load is governed by buckling
or not. Buckling m a y be a problem in cases where the pile is not fully e m b e d d e d in soil,
for e x a m p l e in certain quay and bridge structures with the upper part of the piles in deep
water.

6.1 Fully embedded piles

(i) Straight pile. T h e buckling load of a straight pile e m b e d d e d in soil can be analysed
in the same w a y as in the case of a column e m b e d d e d in an elastic m e d i u m (Timoschenko,
1936). A s s u m i n g that the pile is hinged at both ends we have:

is
4 (242)
m , pip

w h e r e kls,lp,Ep and Ip as above


m = n u m b e r of half waves assuming sinusoidal buckling of the pile.
T h e n u m b e r of half waves (the m value) is determined so that ( p r e a c h e s its m i n i m u m
value.
Introducing the buckling length lp/m = into Eq. (242), w e find:
Deep foundations 213

2 1 \ 3

+ ( 2 4 3
* - * w * > >

T h e m i n i m u m value of Qb is determined by the relation:

= I (- + -) = 0 (244)
TCEpIp

W h e n buckling is taking place w e thus have:

X = X = K(EpIp/kls)U* (245)

from which:
(246)
Qb,min ~ k[sEpIp

a dn 1
Introducing = Qb/Qb,mm = into Eq.(243) w e find:

(247)
1 1 2

In practice, Qb m in and are determined from Eqs. ( 2 4 5 - 2 4 6 ) , w h e n c e m-lp I. Then,


if m is not an integer, the integer values mx and m2 are chosen that are closest to m (mx
<m< m2). This yields t w o values of i:
I] = { -m lmx

T h e value ix or i2 that gives the lowest value yields the buckling load (Mascardi,
1970): (248)
V
Qb ~ Qb, min

If the pile is free to m o v e laterally at either of its ends the buckling load will b e only
half the value obtained for the hinged pile.

Example 33: Determine the buckling load of a straight, 5 m long steel pipe pile (outer diameter = 50
mm; wall thickness = 4 mm) installed in clay with an undrained shear strength of 10 kPa.

4 4 7 4
Solution: We have kls = 200 kPa and Ip = ( 0 . 0 5 - 0.042 )/64 = 1.54 10" m . Furthermore, Ep = 210
GPa. Thus:
6 7 1/2
(2 /? = 2 ( 2 0 0 2 1 0 1 0 1 . 5 4 1 0 - ) = 161 kN
6 7 1 /4
A = (210 10 1.54-10~ /200) = 1.99
m = 5/1.99 = 2.51 and, consequently, m{ = 2 and m2 = 3 from which:
1 1 = 2.51/2 = 1 . 2 5 1 ^ = 1.10
1 2 = 2.51/3 = 0.84 v2= 1.06
214 Deep foundations

Fig. 153. Illustrative example of pile 'wandering'. A 37 m long steel -pile installed by means of the
pile driver shown to the left (top picture) is being curved during installation, then moving in a wide bow
below the ground surface, finally penetrating the ground surface 20 m away and hitting the parked
Honda which is being lifted 1.5 m before discovery. (By courtesy of Herkules Grundlggning AB).

The buckling load is equal to 1.06-161 = 171 kN


The steel stress in the pile under the buckling load is
= 1 7 1 4 / [ ( 0 . 0 5 2 - 0 . 0 4 2 2 ) ] = 2 9 6 1 0 3 kPa

(ii) Initially bent pile. B y experience w e k n o w that it is difficult to maintain the straightness
of driven piles (Fig. 1 5 3 ) . T h e danger of the pile being curved during installation is
particularly important in the case of small-diameter piles, so-called micro-piles. T h e
bearing capacity of such initially curved piles is governed by elasto-plastic buckling
(buckling in combination with plastic yield). A pile may, of course, b e curved along its
w h o l e length and fail in bending because the lateral soil resistance is insufficient. In the
case of buckling, however, the pile is a s s u m e d to b e initially curved between the
inflection points determined by the sinusoidal half w a v e s appearing in buckling (Fig. 1 5 4 ) .
A s s u m i n g that the initial deflection of a pile, hinged at both e n d s , is 0 and the total
deflection caused by the application of the load is ^ + y w e h a v e (cf. E q . 2 4 3 ) :

Qb = 21(- + -*-)-> (249)

T h e m i n i m u m value of Qh is given by the condition:

2 , 2Xk,s y
= 2(- (250)
9 A3 **Epl/y + 0
Deep foundations 215

- - - * Y -
l i n
line of thrust ^ ^ _ \{ I + I 2 I ^ e of thrust

Fig. 154. Buckling of initially bent pile. Radius of curvature = p. Initial deflection from line of thrust
= 0. For a pile curved along its whole length (right picture), the thrust line can be assumed to coincide
with a circle drawn through A, and C. In this case <50 can be taken as the average of the deflections
and b\ (Bernander & Svensk, 1970).

which yields an upper limit of Q due to buckling:

QbMm - 2 / k i s E p I Pp (251)
J + <So

T h e stresses in the pile are obtained from the relation (Bernander & Svensk, 1970):

Q .Mfn
=
E (252)

where Wp = flexural rigidity of the pile,

A M = y = =
p x 2 y y ^ p h QbMmj,

= Stress increment due to the initial deflection <50 of the pile.

The maximum stress in the pile is consequently obtained from the relation:

lim
-7?,max 0 + p
(253)
2 W

The limiting value of Q which leads to plastic yield is thus:


216 Deep foundations

~ ( p, yield ~ jp) An
G

(254)
A

a,iim = V1 + yv ++ Q P
g f t, A
2 W p

Solving y in Eq.(251), w e find:

y +<5Q _ S0y/kl5EpIp

2 2y/khEpIp-Qbyim
(255)

Inserting the expression for (y + <50)/2 into Eq. (254), w e finally have:

a,Hm = \[Qi + aod -^y:I + ftoo + ^ ) ] - 4 a


2
1 0 (256)

w h e r e Qx = ( " ^ 1 ( 1- ) ,
and

0 0.001 0.002 0.003 0.004 0.005


Relative deflection <50

Fig. 155. Examples of limiting load due to elasto-plastic buckling of initially bent piles. Modulus of
lateral subgrade reaction kls = 0.2 MPa. Yield stress of piles >yield = 260 MPa.
7
Deep foundations

T h e value of can b e found from the correlation ( p =radius of pile curvature):

1 8<50
Eplp 2

whence:

Aff =
'^ / W =
^ (257)

F r o m the results of Eqs. (246) and (256) w e find that the limiting load of initially bent
piles is generally governed by yield rather than by elastic instability (Fig. 155).

6.2 Partially embedded pile

For a pile that is not fully e m b e d d e d in soil, it can b e a s s u m e d that the pile is restrained
(rigidly fixed) at a depth below the ground surface equal to (Davidson et al, 1965):

ls=\A{Eplplklsy (258)

w h e r e EpIp = b e n d i n g rigidity of the pile,


kls = m o d u l u s of lateral subgrade reaction.

If the modulus of lateral subgrade reaction is increasing linearly with depth (kls - nhz), ls can
be approximated to:

ls=\WpIp/nh)^ (259)

T h e buckling length depends on the degree of restraint at the pile head. For a pile that
is rigidly fixedat the pile head it can b e a s s u m e d that = 0.5(ls + / 0 ) w h e r e l0 is the free
length of the pile. T h e length of the e m b e d d e d part of the pile should b e > 2.5ls.

Instability is reached for the Euler buckling load:

Qb = n*EpIpM (260)

Example 34: Determine the buckling load of a steel -pile, HEB 200, installed in sand to a depth of
10 m and with a free length of 5 m. The sand has a pressuremeter modulus Epr = 6 MPa.

Solution: The coefficient of subgrade reaction, determined on the basis of the pressuremeter tests (Eq.
238), becomes equal to:
9-6
kh = 7T,
1 / 3
= 16.5 MPa
22.65 +1.5/3 4
In the weak direction of the beam we have Ip = 28.43 1 ( H m . Moreover Ep = 210 GPa.
218 Deep foundations

This yields:
3 6 1 /4
/, = 1.4(21010 28.4310- /16.5) = 1.1 m
= ( 1 . 1 +5)/2 = 3.05m
The buckling load is equal to:
2 3 6 2
Qb = 21010 28.43 1 0 / 3 . 0 5 = 6.3 M N
3 2
Since the cross-sectional area of the -pile is 9.104-10 m this corresponds to a steel stress of
approximately 690 MPa (in most cases above the yield stress).

7. PILE GROUP BEHAVIOUR

7 . 1 Bearing capacity

(i) Friction pile groups. In a pile group, the pile forces will b e influenced by variations
in load/settlement behaviour between the individual piles in the group. This is particularly
the case with friction piles. If the soil were still h o m o g e n e o u s after pile installation, the
frictional forces along the pile shafts would cause a settlement bowl underneath the pile
footing with a tendency towards increasing load share a m o n g the outer piles and a
decreasing load share a m o n g the inner piles in the group. In granular soil, however, pile
installation, particularly in the case of displacement-type piles, generally causes changes
in the deformation characteristics of the soil inside the pile group in such a way that the
load distribution a m o n g the piles will change in another direction than that mentioned.
T h u s , the piles in the centre of the pile group will generally be subjected to a higher load
than the outer piles.
T h e ultimate bearing capacity of a pile group with friction piles is generally different
from the s u m of the ultimate loads of the individual piles in the group.
According to Kishida & Meyerhof (1965) the total bearing capacity of a piled
foundation can be estimated as the bearing capacity of the foundation and its surcharge
effect on the point resistance of the piles in the g r o u p e i t h e r by considering the bearing
capacity of the pile cap as a whole and the bearing capacity of the individual piles in the
group (individual pile failure) or by considering only the contribution to the bearing
capacity of the outer rim of the cap, outside the pile group (the pier), and the bearing
capacity of the pile group as a whole (pier failure), Fig. 156.
In an extensive test series on bored pile/cap/soil interaction effects in sand, comprising
51 pile groups and 23 single piles, Liu etal. (1985) investigated the group effect on both
pile groups with free space between cap and soil and pile groups with pile cap in direct
contact with soil. T h e pile groups consisted of 2 - 1 5 piles, 1 2 5 - 3 3 0 m m in diameter, and
with lengths of 8 - 2 3 times the pile diameter. Pile spacings were 2 - 6 times the pile
diameter. F r o m the results obtained they found no evidence of block failure and therefore
propose that the analysis of pile group failure be based upon the bearing capacity of the
cap as a w h o l e in combination with individual pile failure.
T h e ultimate load Qgrf of a pile group is generally expressed as:
Deep foundations 219

/ \ \
tr
1^--

Leap failure
N\ 1

\
base failure zone J
V

(a) (b)

Fig. 156. Assumed failure zones at piled foundations: (a) pier failure, (b) individual pile failure. (Kishida
& Meyerhof, 1965)

2^=77/7*0,/ (261)

w h e r e 77 = group efficiency factor,


m = n u m b e r of piles in the group,
Qsj = ultimate load of a single pile under the s a m e soil conditions as for the pile
group.
A large n u m b e r of investigations h a v e been carried out for the p u r p o s e of determining
the group efficiency factor 77 under various soil conditions and pile group geometries. T h e
investigations include both free-standing pile groups and piled footings with the pile cap
in direct contact with soil.
Large-scale and full-scale tests on free-standing pile groups in loose to m e d i u m dense
sand h a v e resulted in efficiency factors 77 >1 with a m a x i m u m (77 ~ 2) at a pile spacing
to pile diameter ratio of about 2 - 3 . For free-standing pile groups in dense sand, both 77
>1 and 77 <1 h a v e been found. However, in the latter case only small-scale tests h a v e been
carried out. In reality, a piled foundation on friction piles is usually in direct contact with
the soil. Consequently, the foundation itself m a y contribute considerably to the bearing
capacity of the piled foundation. In the case of cohesionless soils, the c o m b i n e d action
of the piles in the pile group and the piled footing itself has been s h o w n to h a v e a great
influence on the group efficiency. T h e m o s t influential factors, besides soil conditions,
are pile spacing S and the ratio of pile length lp to width of the pile cap.
Liu et al (1985) proposed that the bearing capacity Qj-pc of a a pile g r o u p with pile cap
in direct contact with soil be expressed by the relation:

Qfpc = m{r\s Qfs + 77, Qft) + Qfc (262)

w h e r e 77^ = Cs Gs = group efficiency factor with reference to pile shaft,


rj r = CtGt = group efficiency factor with reference to pile tip,
Gs and Gt = factors of influence of pile/soil interaction on shaft and tip bearing
capacities of the individual piles in the group,
220 Deep foundations

Cs and Ct = factors of influence of cap/pile/soil interaction on shaft and base


bearing capacities of the individual piles in the group,
Q and Q - shaft and tip bearing capacities of single piles,
Q bearing capacity of the cap alone.
In the case of a pile group with pile cap in direct contact with soil, the contribution to
the bearing capacity due to the cap has two main causes: the bearing capacity of the cap
itself and its surcharge effect on pile shaft friction. For pile lengths exceeding t w o times
the width of the pile cap, the surcharge effect on pile point resistance is negligible.
Doubtless, the m e t h o d of pile installation will affect the soil characteristics and m a y
consequently c h a n g e the load vs. settlement behaviour of the piled cap in comparison
with that of the unpiled cap. In order to separate this effect, P h u n g (1993) modified Eq.
(262) and proposed the relation:

Qpc = m(77, Qfi + 7] r Qft) + Cc Q (263)

w h e r e Cc = factor of influence of cap/pile/soil interaction on the bearing capacity of the


pile cap.
As regards driven piles an illustrative effect of pile group installation in a sand layer w a s
presented by E k s t r m (1989). A test series was performed on single piles on the o n e hand
and on individual piles in free-standing pile groups on the other with the n u m b e r of driven
piles varying from 5 to 2 5 . T h e piles which were of hollow steel with square cross-section
(width bp = 60 m m ; wall thickness = 5 m m ) were driven to a depth of 3.3 m. T h e influence
of pile installation on the lateral earth pressure coefficient is exemplified in Fig. 157 and
on the ultimate load of the centre and corner piles in the various pile groups in Table 26.
As can b e seen, there is a strong influence on the soil properties and the load distribution

Fig. 157. Average lateral earth pressure coefficient with reference to the centre pile in pile groups with
pile spacing equal to 4b, driven into sand with relative density ID ~ 4 7 % . The letter L represents results
after loading test. Arrows indicate change in lateral earth pressure with time, in days. (Ekstrm, 1989).
Deep foundations 221

TABLE 26.
Example of ultimate total pile loads and ultimate shaft loads observed by Ekstrm ( 1989). Free-standing
square pile groups. Individual pile tests. Pile loads in kN.

Pile spacing 3bp 4bp 6.5bp 3bp 4b 6.5bp


Initial ID % 47 47 60 47 47 60

Total Shaft

Single pile 8-10 11 24 3-6 6 10

Pile group, 5 piles:


Centre 18 21 33 13 14 20
Corner 11-18 25-26 5-8 8-11

Pile group, 9 piles


Centre 22 27 38 15 18 22
Corner 14 25 23 6 12 9

Pile group, 13 piles


Centre 35 36 19 21
Corner 25 35 10 17

Pile group, 25 piles


Centre 37 33 19 17
Corner 25 25 7

TABLE 27
Group efficiencies for bored piles according to Liu etal. (1985). Dp = 0.25 m. Loose silty sand. 77^ and
77, include pile/soil interaction effects while % also includes action of pile cap in contact with soil.

SJDp l ID Group Shaft Tip Total


PP
Tis %
3 8 3x3 0.36 1.44 1.64
3 13 3x3 1.09 1.51 1.69
3 18 3x3 1.16 1.49 1.51
3 18 3x3 1.42 0.91 1.36
3 23 3x3 1.16 1.12 1.15

2 18 3x3 0.98 0.70 1.21


4 18 3x3 1.11 0.93 1.46
6 18 3x3 0.82 1.06 2.23

3 18 1x4 1.11 1.10 1.49


3 18 2x4 0.88 1.51 1.40
3 18 4x4 1.03 1.45 1.19
3 18 2x2 1.20 1.22 1.60
222 Deep foundations

T A B L E 28.
Efficiency factors at pile failure obtained by Phung (1993). Pile cap action included.

ID(%) Slbn Tip ,


Shaft 5 Total
Free-standing group:
38 2.6 2.0 2.4
67 3.2 0.8 1.1
62 2.0 1.0 1.2

Cap in contact with soil:


38 3.2 3.0 3.1
67 4.4 0.7 1.3
62 4.4 1.4 2.0

0.5 m

-50 L
0 50 100 150 200 250
Cap load, kN
150

0.5 m
100

0.75 m
50

1.25 m
M rTi
1.75 m

-50
20 40 60
Settlement, mm

Fig. 158. Influence on lateral earth pressure against the shaft of the centre pile in the group due to the
surcharge induced by the pile cap (After Phung, 1993).
Deep foundations 223

a m o n g the piles in the group and, as w a s to be expected, the application of results of


loading tests on single piles for the determination of the pile group capacity, a s s u m i n g
= 1 brings about a conservative design. In his test series on the ultimate load of square
pile groups, c o m p r i s i n g 5 piles, Ekstrm found the pile group efficiency factor = 2.2
for pile spacing 4bp and = 1.4 for pile spacing 6.5bp.
F r o m their tests on group effects on bored piles with pile cap in direct contact with soil,
Liu et al. (1985) presented the group efficiencies given in Table 27.
Another illustrative example of cap/pile/soil interaction was presented by Phung (1993).
P h u n g performed a test series comprising cap without piles (shallow footing), single pile,
free-standing pile group (5 piles) and pile group (5 piles) with pile cap in direct contact
with the soil. T h e piles utilised were of the same type as in the test series performed by
Ekstrm, /. e. hollow sand-covered steel piles with square cross-section, 4 0 m m by 4 0 m m
and 5 m m wall thickness. However, the pile length e m b e d d e d in sand w a s in this case only
2 m. S o m e of the results obtained are presented in Figs. 1 5 8 - 1 6 0 .
T h e efficiency factors in respect of shaft resistance, base resistance and total resistance
obtained by P h u n g are presented in Table 28.
Concerning floating pile groups in clay soils, the ultimate load is governed by individual

Fig. 159. Comparison of load vs. settlement relationship of single pile to those of cap (shallow
foundation), free-standing pile group and pile group with cap in direct contact with the soil. Pile spacing
= 4b. Sand with In = 38%.
224 Deep foundations

pile failure at large pile spacing, or, w h e n the pile spacing is small, by the shear strength
along the perimeter and nearest below the b o t t o m of the pile group, so-called block
failure. A group efficiency factor of 0.7 is r e c o m m e n d e d for pile spacings in the range
of2.5Dp-4Dp (see B r o m s & Hansbo, 1981).

(ii) End bearing pile groups. T h e ultimate bearing capacity is generally calculated as the
sum of the bearing capacity of a single pile.

7.2 Settlement

Different proposals have been presented about how to define, under equal soil conditions,
the settlement of pile groups in relation to the settlement of an individual free-standing
pile the so-called settlement ratio . A m o n g these, it seems preferable to use either one
of the following definitions:

is the ratio of pile group settlement to single pile settlement at equal pile loads
Load, kN Load, kN

0 10 20 0 100 200 300 400

Fig. 160. Comparison of load vs. settlement relationship of single pile to that of cap (shallow foundation),
free-standing pile group and pile group with cap in direct contact with the soil. Pile spacing = 8/?. Sand
w i t h / D = 62%.
Deep foundations 225

i s the ratio of the initial slope of the average pile load vs. settlement curve of the pile
group to the initial slope of the load vs. settlement curve of the single pile.

(i) Pile groups in granular soil. A fairly large number of tests have been carried out to find
the value but the results obtained are often difficult to analyse both because of the
settlement ratios being related to different factors of safety against pile failure and the
settlement ratios not being defined. A s settlement is very much dependent on the factor
of safety applied, the results presented by different authors show great scattering and are
often contradictory. Roughly speaking, results of loading tests on free-standing pile
groups indicate that >\ in dense sand while <\ for driven piles in loose to medium
dense sand. Thus, in the latter case h a s been found to vary from about 0.2 at BgrIDp ~
3 (where Bgr = the width of the pile group) to about 0.7 at Bgr IDp ~ 10. A s for bored piles
in loose sand, h a s been found to vary from about 0.6 at BgfJDp ~ 3 to about 2 at Bgr/Dp
-5.
Based on the results of full-scale investigations, Vesic (1969) suggests that the Rvalue
be determined by the relation:

(264)

As in the case of the ultimate load of pile groups in sand, the influence on pile group
settlement of the surcharge exerted by the pile cap may be quite important (Figs. 158
160). The results presented by Phung show that the settlement ratio , defined as the ratio
of the pile group settlement to single pile settlement at equal loads, is fairly constant up
to an average pile load in the pile group of about 6 0 % of the single pile failure load and
from then on strongly decreasing. The results also show that a considerable decrease in
the Rvalue can be expected due to the pile cap being in contact with the soil. This has also
been found for bored piles at Bgr IDp between 2 and 5 (Garg, 1979).

0.15

a
CS
o.iof- 1
_ o Outer piles
1 'S

.05- Centre pile

0 0.25 0.50 0.75 1.00


Ratio of working load to ultimate load of pile group

Fig. 161. Load distribution between individual piles in a pile group in soft clay loaded to failure. Piles
of aluminium, 3 m in length and 0.037 m in diameter. Pile spacing 3.5Dp. Soft high-plasticity clay.
226 Deep foundations

(ii) Pile groups in cohesive soils. T h e settlement of friction pile groups in cohesive soils
depends on the preloading history of the soil.
T h e instantaneous settlement of the pile group can b e calculated on the assumption that
the soil behaves as an elastic m e d i u m with P o i s s o n ' s ratio v = 0.5. T h e settlement thus
obtained is also valid for the long-term settlement of pile groups in overconsolidated
clays w h e r e the in situ pressure induced by the pile group does not exceed the
preconsolidation pressure of the clay.
In clay soils, the m o d u l u s of elasticity Es to be applied in the analysis can b e chosen
on an empirical basis, for e x a m p l e in relation to the undrained shear strength cu of the clay.

TABLE 30
Rvalues for free-standing pile groups with lp/Dp= 25 in an elastic medium with v = 0.5. Rigid cap. Pile
spacing = S. Depth of the medium = d.

d/lp = CO 5 2.5 1.5 1.2


S/Dp

Group 2x2
2 2.91 2.80 2.76 2.46 2.01
3 2.59 2.46 2.41 2.10 1.70
5 2.19 2.08 2.00 1.69 1.39
10 1.70 1.63 1.54 1.29 1.16
Group 3x3
2 5.38 5.00 4.88 4.10 3.09
3 4.64 4.22 4.06 3.25 2.39
5 3.74 3.27 3.05 2.30 1.75
10 2.73 2.20 1.98 1.48 1.27

Group 4x4
2 8.34 7.56 7.29 6.02 4.18
3 6.96 6.12 5.77 4.01 3.05
5 5.34 4.43 4.00 2.82 2.05
10 4.43 2.66 2.29 1.60 1.33

With regard to the pile length, the values given in Table 30 can b e adjusted by the
factors given in Table 3 1 .

TABLE 31
Adjustment factors for lp/Dp.

S'Dp lptDp= 10 lp/Dp= 25 100

2.5 0.82 1.00 1.20


5 0.77 1.00 1.30
10 0.74 1.00 1.45
Deep foundations 227

I \
^TmiltkTIIIIIIIIIIIIIIIII
/ V Q \
(Bgr+zHLgr+z)

Fig. 162. Equivalent raft approach.

Thus, for normally consolidated clay Es ~ 150c w while for heavily overconsolidated clay
Es - 500c M.
T h e load distribution a m o n g the piles in working load condition is in good agreement
with the theory of elasticity (Fig. 161).
A c o m p r e h e n s i v e theoretical study of the relation between individual pile settlement
and pile group settlement in an elastic m e d i u m w a s presented by Poulus (1968). H e
assumes the soil to behave as an elastic medium and the pile cap to be either rigid or
perfectly flexible. In the rigid cap case, their analysis yields the settlement ratios given
in Table 30.
T h e Rvalues given in Tables 30 and 31 can b e considered representative for settlements
in undrained condition (immediate settlements). However, according to Poulos the
immediate settlements of pile groups represent the p r e d o m i n a n t part of the total, final
settlements.
T h e Rvalues obtained on a theoretical basis (Table 30) are supported by results report-
ed by, for example, Berezantzev et (1961), Sowers et (1961) and H a n n a (1963).

(iii) Equivalent raft method. According to the so-called equivalent raft m e t h o d , the load
applied on the pile group is assumed to act at the lower third of the pile length (Fig. 162).
T h e settlement of the pile group is calculated as the s u m of the equivalent raft settlement
and the compression of the upper t w o thirds of the piles. According to Poulos (1993), a
good correlation can b e expected between the results obtained b y this m e t h o d of analysis
and computer-based analysis of settlement, based on his o w n data p r o g r a m D E F P I G which
takes into account pile/soil interaction in an elastic m e d i u m (Poulos, 1990).
Regarding pile groups in low-permeable, cohesive soils, w h e r e the load induces in situ
stresses in excess of the preconsolidation pressure, the long-term consolidation settlement
228 Deep foundations

can be analysed in a similar way. For pile spacings presumably less than 8 times the pile
diameter Dp, and to a certain depth the relative m o v e m e n t between pile shaft and
surrounding soil seems m o r e or less negligible (Hansbo, 1973). T h e soil above the lower
third of the pile length behaves as an 'incompressible' layer (compression m o d u l u s M
tending to infinity). In consequence, the course of settlement can be analysed on the
assumption of full drainage (cv = kMlyw tending to large values) at the fictitious foundation
depth 2 / ^ / 3 .

7.3 Design

As shown in Section 2.8, the behaviour of individual piles in a pile group can b e quite
different from their behaviour as single piles. This fact is generally neglected in
conventional design.
By tradition, the pile group design is carried out on the basis of very simplified
assumptions:
T h e piles in the group are assumed to function as axially loaded c o l u m n s hinged at
pile head and pile tip. (Computer p r o g r a m m e s exist which take into consideration pos-
sible restraint of the pile head as well as lateral soil resistance).
Every pile is a s s u m e d to have the same axial stiffness, i.e. they are a s s u m e d to have
equal lengths and cross-sectional areas and to be supported by an unyielding m e d i u m .
T h e pile cap is considered rigid and is not assumed to contribute in carrying the load
applied.
Forces applied at the centre of gravity of the pile group are assumed to cause pure
translation. M o m e n t s applied at the centre of gravity of the pile group are a s s u m e d to
cause pure rotation.
In c o n s e q u e n c e of this model of pile group analysis, horizontal forces acting on the pile
group h a v e to be taken by raker piles. Moreover, in order to reduce the m o m e n t of
rotation, the pile group has to be arranged in such a w a y that its centre of gravity is as close
as possible to the line of action of the external force resultant.
In reality, the piles in a pile group m a y deviate considerably from the position given
in the design. Therefore, the real position of the piles has to b e checked and the pile forces
recalculated after pile installation is terminated.
According to conventional design, the load acting on the pile group, irrespective of
whether the piles are in cohesionless or cohesive material, is assumed to be carried by the
piles alone with a certain factor of safety against failure. This approach is rational w h e n
the piles are end bearing or mainly end bearing or w h e n w e h a v e to deal with footings on
normally consolidated clay. However, it is used even in cases w h e r e the bearing capacity
of the footing itself would b e satisfactory, the reason being that the settlements without
piles are felt to b e too large. In situations where the piles installed are friction piles for
w h i c h the load/settlement relationship does not show a m a r k e d decrease after peak, this
approach is quite conservative and unnecessarily expensive. A m o r e cost-effective
Deep foundations 229

approach is first to investigate h o w m u c h of the load can be carried by the pile cap without
causing excessive settlement and then design the pile group to carry the r e m a i n i n g part
of the load. T h e intricate p r o b l e m of analysing the influence of pile/soil/cap interaction
on the ultimate load of the piles in the pile g r o u p and on the settlement can b e totally
disregarded. T h u s , on the basis of the investigation carried out b y P h u n g (1993), the
settlement obtained under the load taken by contact pressure at the cap/soil interface in
a piled footing is very nearly equal to the settlement obtained under an equally large load
taken by a corresponding unpiled cap (spread footing). This fact simplifies the design
procedure. T h e n u m b e r of piles required to limit settlement can b e d e t e r m i n e d on the
basis of the ultimate load of the single pile and the settlement can b e calculated as if the
pile cap w e r e a shallow footing carrying the load not taken by the piles.
T h e principle of pile group design can thus b e s u m m a r i s e d as follows:
D e t e r m i n e the load (Qx) that can b e placed on the unpiled footing without causing
unacceptable settlements.
T h e r e m a i n d e r of the load (Q - Qx) should b e carried by settlement reducing piles.
As the permissible settlement will be large e n o u g h for shaft resistance to b e fully
mobilised, the piles can be designed as friction piles in a state of failure (in clay, in a state
of creep failure).
T h e settlement of the piled footing can b e estimated at about the s a m e value as the
settlement of the unpiled footing under load Qx. This leads to a conservative design. For
e x a m p l e , for footings on sand the design b e c o m e s m o r e conservative the looser the sand.
This approach has b e e n used with great success to b o r e d large-diameter piles in stiff
L o n d o n clay (Burland, 1986).

Example 35: Determine the load that can be carried by the piled cap shown in Fig. 160, following the
principle of design proposed above. The soil characteristics, determined by pressuremeter tests, are
given below (Phung, 1993):

Depth, m Epr MPa 'Pi MPa p c rM P a

0.5 4.74 0.33 0.33


1.0 2.55 0.22 0.15
1.5 5.89 0.47 0.32
2.0 8.23 0.53 0.38
2.5 6.33 0.52 0.37

Solution: The ultimate load of the single pile determined on the basis of the pressuremeter tests is given
by Eq. (211), p. 191. The mean value of the limit pressure along the pile shaft is 0.39 MPa and of the
creep pressure 0.30 MPa. According to Fig. 138, case (d), the shaft resistance can be estimated at 15 kPa.
This yields a total shaft resistance of 1 5 0 . 0 6 4 - 2 . 1 = 7.6 kN (observed value = 1.8 kN). For the
determination of the base resistance, Eq. (209), we have a value of 2.6. The net limit pressure is 0.49
2
MPa. This yields a tip resistance of 2.64900.06 /4 =3.6 kN (observed value = 8.2 kN). Thus, the total
bearing capacity is 7.6 + 3.6 ~ 11 kN (observed value = 10 kN). Choosing instead of *p{ the net creep
pressure *pcn the creep failure load can be estimated at 5.7 + 2 . 6 - 8 kN.
230 Deep foundations

Fig. 163. Piled raft with equally distributed wooden piles, 18 m in length. Soft highly plastic clay
reaching to a depth below foundation level of about 45 to 85 m.

The ultimate load / c f m e unpiled cap is determined according to Eq. (141), p. 136. We have ple
= (0.33-0.22) 1 72 = 0.27 MPa and bearing capacity factor k = 0.8. This yields Qfc = 0.8-270-0.8 2 = 138
kN (observed value 200 kN). The critical load (the 'creep' load) becomes 120 kN. Assuming a factor
of safety of minimum 1.5 we find Qx =90 kN (which is below the creep load), i.e. qx = 0.144 MPa.
The settlement is determined according to Eq. ( 151 ), p. 146. We have the shape coefficients (1=2
and := 1.10 and the rheological coefficients ad = a J; = 1/3. The pressuremeter moduli to be applied in
the analysis are determined according to Eqs. (152-153). We find = 4.7 MPa and Eprd = 4/[ 1/4.74
+ 1/(0.85-4.74) + 1/4.39 + 2/(2.5-7.15)] = 5.0 MPa. Thus, the settlement is given by the relation:

1.2. 0 . 1 4 4 . 2 0.6 1 . 1 3 - 0 . 8 1 3/ 1.10-0.8. . n A c/ :


.v = [ ( ),M+ ] = 0.0065 m
9 5.0 0.6 3-4.7
The total load that can be carried by the piled cap without exceeding a settlement of 6 - 7 mm is the
sum of the unpiled cap load and the pile loads at failure, i.e.:
Q = 1 + m - s /= 9 0 + 5 11 = 145 kN
The observed settlement under this load, according to Fig. 160, is 4 mm. Taking into account that
settlement, determined on the basis of pressuremeter tests, refers to settlement after a loading time of
10 years, the deviation between the calculated and observed values will most probably be strongly
reduced with loading time. However, in spite of the fact that the factor of safety against failure for the
unpiled footing is as low as 1.5, the design of the piled footing turns out to be conservative.

8. PILED RAFTS

Piled rafts (Fig. 163) are used instead of piled footings in poor soil conditions, particularly
w h e r e the subsoil consists of soft, normally consolidated clay. T h e traditional approach
in this case is the s a m e as for piled footings the piles h a v e been designed to carry the
Deep foundations 231

Fig. 164. Settlement contours in mm for two adjacent residential buildings in Gothenburg. Equal soil
characteristics (normally consolidated, highly plastic clay to great depth) and equal building loads (
2
60 kN/m ). Top: Traditional foundation design (factor of safety against pile failure equal to 3). Centre:
Raft foundation with settlement reducing creep piles. Bottom: Average settlement vs. time for the two
buildings, (cf. Hansbo, 1984, and Jendeby, 1986).

w h o l e load with a certain factor of safety against failure. A design principle, similar to
that applied to piled footings, has been used in S w e d e n since long ago with great success
(Hansbo & Kllstrm, 1983; H a n s b o , 1984; Jendeby, 1986; Svensson, 1 9 9 1 ; Randolph,
1993).
T h e approach, suitable for piled rafts on normally or lightly overconsolidated clays,
is as follows:
232 Deep foundations

determine the effective overburden pressure and the preconsolidation pressure a ' c
at different depths in the clay,
determine the decrease in overburden pressure due to excavation and the increase
caused by the construction of the building,
decide h o w m u c h of the building load can be carried by contact stresses at the soil/
raft interface without exceeding the preconsolidation pressure in the soil,
the remainder of the building load shall b e carried by piles in a state of creep failure
(pile load equal to creep load); the piles should be distributed in such a way that the
preconsolidation pressure is nowhere exceeded and so that the differential settlement is
minimised.
Besides the savings in foundation costs, this approach has the advantage that the pile
forces acting against the raft from below are k n o w n in size. C o m p u t e r p r o g r a m m e s have
been developed for the design of the raft, taking into consideration the stiffening effect
of internal walls in the b a s e m e n t (Svensson, 1991).
An e x a m p l e of the settlement distribution for two adjacent residential buildings, one
of which is designed according to the 'creep pile' approach, the other according to the
traditional approach, is given in Fig. 164.

9. DOWNDRAG

As previously mentioned, piling particularly pile driving entails a disturbance of


the original soil structure which in the case of l o w - p e r m e a b l e cohesive soils gives rise to
excess pore water pressure and reduced bearing capacity. Moreover, the architectural
layout of buildings often entails a raising of the ground level at the building site, or the
buildings are to b e constructed at sites covered by old fill material of considerable
thickness. Also, building and drainage activities often entail a lowering of the groundwater
level and, in consequence, an increase in the effective overburden pressure. Obviously,
there is a considerable risk of long-term settlement taking place in cases w h e r e the subsoil
consists of soft clay or organic material. T h e consolidation process thus induced gives
rise to so-called negative skin friction, or down-drag, which has to b e considered in the
design of end bearing piles. T h e piles will carry not only the applied load but also part
of the weight of the surrounding soil.
In old days, negative skin friction was seldom, if ever, considered and all the same the
buildings rarely suffered d a m a g e . However, in certain circumstances, the consequences
of ignoring negative skin friction can be serious. T h u s , in practice it has h a p p e n e d that
piles h a v e been pulled out of the foundation due to negative skin friction which has
entailed serious d a m a g e to the buildings.
T h e negative skin friction ca can b e estimated from the equation:

ca = tan0' fl G'V (265)


Deep foundations 233

w h e r e o\ is the effective overburden pressure,


is the lateral pressure against the pile shaft,
' is the angle of friction in the pile/soil interface.
According to B r o m s (1976), Ktancj)^ can be estimated at 0.4 in rock-fill, 0.35 in sand
and gravel, 0.3 in silt and normally consolidated clay of low to m e d i u m plasticity (IP <
5 0 % ) and 0.2 in normally consolidated clay of high plasticity.
For high-plasticity normally consolidated clay the negative skin friction can also be
estimated from the undrained shear strength cu by the relation:

ca - 0.6c t t (266)

Negative skin friction affects the individual piles only d o w n to the neutral point, w h e r e
the relative m o v e m e n t between pile and soil in the pile/soil interface is zero. B e l o w the
neutral point, the skin friction is positive. The position of the neutral point d e p e n d s on
the length of the piles and the bearing stratum at the pile point. For piles driven through
deep layers of soft clay to bedrock of high bearing capacity, it can be a s s u m e d that the
whole pile is subjected to negative skin friction. However, for piles driven into a sand or
gravel or to bedrock with an ultimate bearing capacity less than that of the pile section,
the position of the neutral point can be taken at the level w h e r e the settlement of the
surrounding soil is 5 m m (Norwegian Pile C o m m i s s i o n , 1973).
For pile groups with large pile spacing, each individual pile can be a s s u m e d to be
subjected to negative skin friction according to Eqs. ( 2 6 5 - 2 6 6 ) . At small pile spacing, the
d o w n - d r a g forces on the pile group can be assumed to comprise the weight of the fill
above the pile group plus the shear resistance of the soil on the perimeter area of the pile
group d o w n to the neutral point. In consequence, the d o w n - d r a g will be larger for the
peripheral piles than for the central piles.

10. PIERS, CAISSONS AND UNDERGROUND WALLS

10.1 Introductory remarks

T h e term pier has t w o different meanings: (a) it represents an u n d e r g r o u n d structural


m e m b e r serving the purpose of transmitting the load to a stratum capable of carrying the
load without danger of excessive settlement; (b) it represents a support for the superstructure
of a bridge, usually of concrete or masonry, rising above ground level or water level.
A c c o r d i n g to the definition given by Barker (1981) a pier in its first-mentioned usage
is an u n d e r g r o u n d structural member, having a m a x i m u m depth to width ratio of 4 : 1 .
However, according to the definition given by Peck et ai (1973) it is an u n d e r g r o u n d
structural m e m b e r having a depth to the width ratio that is usually greater than 4. Caissons
are used for excavation purposes and can be considered as o p e n - e n d e d piers.
Underground walls c o m p r i s e d i a p h r a g m walls, so-called secant walls constructed by
234 Deep foundations

Fig. 165. Secant pile wall prepared for intermediate floor slab. By courtesy of Bachy.

bored piles that intersect to form a solid wall, resembling the d i a p h r a g m wall, and walls
consisting of piles with interspace w h e r e the soil is retained by arching (Figs. 1 6 5 - 1 6 7 ) .
T h e construction of piers is carried out in very m u c h the s a m e w a y as of large-diameter
bored piles. A hole is excavated or drilled into the soil d o w n to the foundation level and
the pier is built inside. T h e sides of the hole usually have to be stabilised, for instance by
m e a n s of bentonite slurry or sheet pile walls. Another m e t h o d for construction of a pier
is the use of caissons, generally provided with a cutting edge to facilitate the caissons
being lowered to the depth of foundation.

10.2 Bearing capacity

T h e analysis of the bearing capacity of piers, caissons and underground walls can b e
carried out in the same w a y as for piles with due consideration to shape and depth/width
relations. T h e main contribution to the bearing capacity is obtained from base resistance
(cf. Brandl, 1993). In the case of granular soils, the most reliable m e t h o d of analysis seems
to b e the one based upon the results of pressuremeter tests. Calculations based on the
shear strength parameters c ' a n d (//of granular soils are carried out on the assumption of

Fig. 166. Secant wall with unreinforced 'female' piles and intersecting, reinforced 'male' piles.
Fig. 167. Assumed zone of failure for a deep wall foundation. Two cases may occur: the failure zone
does not reach the ground surface (left) and the failure zone intersects the ground surface (right).

'global' failure which does not usually agree with reality. For the sake of completeness,
this type of analysis will all the same be included.

(i) Geostatical method. In the case of cohesionless soil, it m a y be difficult to define the
failure load as the load/settlement curve is generally quite flat r e m i n d i n g of that obtained
for strain hardening soil. This has its explanation in local shear failure taking place before
total failure of the foundation. O n e w a y of taking this into consideration is to apply a
reduced value of the shear strength of the soil or to define failure in relation to a certain
relative settlement sib, w h e r e b is the width of the foundation. A c o m m o n assumption is
that failure takes place at sib = 0.10.
T h e analysis of the bearing capacity of piers and underground walls can be carried out
in principally the s a m e way as for shallow foundations, see p. 126. Let us first consider
the contribution to the bearing capacity represented by the bearing capacity factors Nq and
Nc. A s s u m i n g that w e h a v e a case of ' g l o b a l ' shear failure, comprising a combined
Prandtl-Rankine failure zone, t w o cases m a y occur: (1) the failure z o n e reaches the
ground surface, (2) the failure zone does not reach the ground surface (Fig. 167).
Assuming that the failure zone reaches the ground surface as shown to the right in Fig.
236 Deep foundations

167 and that w e h a v e a case of plain strain condition (strip foundation), the solution to
the p r o b l e m can b e derived as follows.
In the soil w e d g e A C D , the relative lengths A C / C D / D A , according to the law of sines,
can b e expressed as COS(t] + ^'. Equilibrium projections in the r 0 and ' 0
directions yield:
T5cos0'= [cos(r+</Ocosj - sin?] sin(rj+0O + (fx [COS(t] + ^ - sin?7 C O S ( t j + ")
/ = =
aoCos0 - Ti [cos(77+0Osinr] + sinrcos(7j+0O + <fx [cos(rj + 0Ocos77-sin7]sin(7] + ")

Introducing the M o h r - C o u l o m b failure condition = c' + \ t a n 0 ' w e find after


simplification:
^ c ' + c / . t a n ^ ^ ^ ) ( 2 6 ?

COS0

7
c + \ tan , , ,./m
= cfx - - - [sin(2i] 4- ') - sin '] (268)
cos '

By this equation system, and can be determined for any given value of cf0, r 0 and '.
Consider section A D as an equivalent free surface. A s s u m i n g further that r represents
the part of the shear strength m a d e use of, w e have:

, / 2 6 9
to = r ( c + CTotan0O ( )

T h e relation between and 2 is given by Ktter's equation:

2 = exp(2 tan^O = (c'+ o\ tan</Oexp(20 tan^O (270)

T h e equilibrium condition for the R a n k i n e failure zone below the base of the wall
(wedge A B E ) yields:

qf= (f2 + r 2 t a n ( 4 5 + 072) = ( r 2 - cOcot0' + r 2 tan(45+ 072) (271)

Substituting Eq. (268) for ' 0 and Eq. (270) for r 2 and expressing the bearing capacity
under the conventional form:

q^ibllWNy+NyG'v + c'N,

the bearing capacity factors Nq and Nc are given by the relations:

/ /
_ (1 + s i n 0 ) e x p ( 2 0 t a n 0 )
9 ) ( 2 7 2
" l - s i n 0 ' sin(2rj+0')
Deep foundations 237

1000
J3 = 90

/3 = 60

// /
V
/ /3 = 30
100 //
V

A //
/
Y // ////
VA j

10 /
//','///
//
<///.

V r -= 0
- r=1
1
10 20 30 40
Angle o f internal friction '

Fig. 168. Bearing capacity for deep wall foundation.

yVc = ( i V , - l ) c o t f (273)

where = 3/4 +-- 072,


= the angle of inclination to the horizontal of the upper boundary surface of the
failure zone (Fig. 167).
The value is a function of 0 ' a n d r and is governed by the relation:

cos(277
v 1
+ 0 ) tan<*
r = {1 [sin(27] + 0 - sin ']} (274)
COS0' COS0'
This yields 77 = / 4 + 0 7 2 for r = 0 and 77 = 0 for r = 1.
The ^ values according to Eq. (272) for r = 0 and 1 and for = 30, 60 and 90 are
given in Fig. 168.
In case the failure zone does not reach the ground surface, ' 0 is replaced by o's. The
earth pressure coefficient against the sides of the foundation, Ks = O'JG'V , can be assumed
to vary between active earth pressure and unity. The ' value should be chosen as an
average along the boundary between the failure zone and the sides of the foundation with
a maximum value governed by the critical depth according to Fig. 135 where the ratio dcr IDp
is replaced by dcrlb.
238 Deep foundations

= 90

0 10 20 30 40
Angle of internal friction </>'

Fig. 169. Bearing capacity deep wall foundation.

In the analysis of the bearing capacity factor 7Vr it is assumed that the failure zone has
a shape similar to the one prevailing in the former case treated above. T h e m i n i m u m
values of A^are found by trial and error. T h e R v a l u e s , according to Meyerhof (1951),
are given for = 30, 60 and 90 in Fig. 169.
For values of djb > 5, local shear failure seems to govern the b a s e bearing resistance.
M e y e r h o f (1951) r e c o m m e n d s in this case that the shear strength of the soil be reduced
to 8 5 % of the value determined.This is s o m e w h a t higher than 2/3 of the value which was
suggested by Terzaghi (1943).

TABLE 32
Shape factors based on Meyerhof s proposal.

Friction angle ' length to width ratio Shape factor

30 >1 1.0
35 1 1.2
>5 1.0
40 1 1.9
2 1.5
5 1.3
> 10 1.0
Deep foundations 239

S h a p e factors with regard to length-width relations of the foundations increase


according to M e y e r h o f with increasing angle of internal friction. T h e m a g n i t u d e of the
shape factors h a v e to b e determined on an empirical basis. For djlb b e t w e e n 4 and 10 the
shape factors p r o p o s e d in Table 32 can b e applied.
T h e contribution to the bearing capacity of skin friction should b e added to the base
bearing capacity. A m a i n difficulty arises in the appreciation of the horizontal pressure
along the shaft and the angle of skin friction <5. Its contribution to the bearing capacity can
normally b e neglected.
For cohesive soils the bearing capacity factors NqQ and are given b y the relations:

^ 0 = 1 (275)

2
Afo = 1 + 3 / 2 + 2/J + / - r - arccos r (276)

w h e r e r = ca lcu represents the ratio of wall adhesion to undrained shear strength


For values of djlb > 2, the angle of inclination can be a s s u m e d equal to / 2 (90). This
yields bearing capacity factors equal to = 8.28 w h e n r = 0 and = 8.85 when r =
1. In the case of soft clay, M e y e r h o f r e c o m m e n d s that the undrained shear strength be
reduced to 9 0 % of the measured value. Using instead the unreduced value cu in the analysis,
the values should b e reduced to NcQ = 7.45 w h e n r = 0 and Nc0 = 7.96 w h e n r = 1.
For a rectangular foundation, Meyerhof r e c o m m e n d s that the value of Nc0 b e multiplied
by the shape correction factor:

sc=\+0A5b/l (277)

Example 36: Determine the bearing capacity of a diaphragm wall with a thickness of 0.5 m, founded
at 10 m depth in a homogeneous sand layer with an internal angle of friction of 35. The groundwater
3 3
level is at 2 m depth. The effective density of the soil is 1.8 t/m above and 1.1 t/m below the groundwater
level.

Solution: First find out whether or not the zone of failure reaches the ground surface. With the notations
given in Fig. 167 we have AB = 0.25/cos(45 + 35/2) = 0.54 m. The radius vector of the logarithmic
spiral, representing the Prandtl failure zone, if extended to the sides of the diaphragm wall, is equal to
0.54-exp{7U-[(180 + 45 - 3572)/180]-tan35} - 6.8 m < 10 m. Thus the bearing capacity factors are
governed by = 90.
Since djb > 5, the angle of internal soil friction to be applied in the determination of Nq and ^ s h o u l d
be reduced to 0 ' = arctan(0.85-tan35) 31. According to Figs. 168-169 this yields Nq ~ 140 and
~ 160. Since = 90, we have cr^ = <fs. The magnitude of <fs is uncertain and has to be estimated. In
our case we assume Ks = 0.5. The & s value is taken as the average along the sides of the wall inside the
failure zone, i.e. from the base of the wall to a height of 6.8 m above the base.The critical depth ratio
for '= 31 according to Fig. 135 is about 10 which yields dcr = 5 m. The maximum value of ' to be
applied is thus 2-18 + 3-11 = 69 kPa. At 6.8 m above the base we have ' = 2 18 + 1.2 11 - 49 kPa.
The bearing capacity becomes:
3
^=0.25-11.160+ 140-0.5-[(49+ 69) 1.8/2 + 69-5)/6.8]-10- - 5.1 MPa
240 Deep foundations

TABLE 33.
Maximum values for strip foundations and corresponding, minimum embedment ratios. (Baguelin
et ai, 1978)

Category Net limit pressure * p l ( d / b ) mm



max

Rock 1 1 1.9
2 1 2.2
5 1 2.6
10 7 3.0

Sand and gravel 0.4 6 1.9


2 9 3.1
6 12 4.4
Silt 0.1 3 1.2
0.5 4 1.6
1 5 1.9
3 6 2.2

Clay 0.1 3 1.3


1 4.5 1.8
4 5 2.2

which, if u n r e d u c e d values of cu are applied, yields a m a x i m u m value of Nci) between 8.6


( w h e n r = 0) and 9.2 (when r = 1) for a circular foundation.
Janbu et al (1956) p r o p o s e d a s o m e w h a t m o r e conservative value of the bearing
capacity factor Nc0 (see Fig. 96, p. 130). Accordingly, reaches its m a x i m u m value at
djlb ~ 4. For values of djlb < 4, can be determined by the approximate relation:

7
Na = (1 + 0.2y ){5.14 + - A s - iV } (278)
/ 3 b b

(ii) The pressuremeter method. T h e bearing capacity of piers and u n d e r g r o u n d walls can
be estimated on the basis the net limit pressure determined by m e a n s of pressuremeter
tests in a c c o r d a n c e with Eq.(209). T h e values for strip foundations increase from 0.8
for the e m b e d m e n t ratio dlb = 0 to the m a x i m u m values presented in Table 3 3 .
For values of dlb less than the m i n i m u m values (d/b)min required to reach the m a x i m u m
values Kmax, according to Table 3 3 , the values can b e obtained b y the relation:

jf-O^ + i ^ - o W - ^ - (279)
Deep foundations 241

Example 37: Determine, by the pressuremeter method, the bearing capacity of a diaphragm wall, 0.8
m in thickness, with its base at 5 m depth in a sand deposit. The pressure limits pt in the sand, observed
at depth intervals of 1 m, are 2 . 5 , 2 . 1 , 1 . 9 , 2 . 2 , 1 . 8 , and 2.3 MPa, starting from 1 m depth below ground
surface downwards. The groundwater level is at 2 m depth and the porosity of the sand = 27%.

Solution: The density of the sand below and above groundwater is obtained by Eq. (5). Assuming Sr
3 3
= 100% below groundwater and pg = 2.65 t/m we obtain = 2.65(1 - 0.27) + 0.27-1.0 = 2.2 t/m ,
3
and assuming Sr = 0 above groundwater we find = 2.65(1 - 0.27) = 1.9 t/m . The effective overburden
pressure at the foundation level is then ' = 9.81(2-1.9 + 3-1.2) 73 kPa.
1 /3
The equivalent limit pressure with regard to end bearing capacity is ple = (2.2-1.8-2.3) = 2.1 MPa
from which the net limit pressure pte ~ 2.0 MPa.
1 /2
For dlb = 5/0.8 = 6.25 and (d/b)min = 9 we find = 0.8 + (3.1 - 0 . 8 ) ( 6 . 2 5 / 9 ) = 2.7.
The pressure limit with regard to shaft resistance can be taken as the average of the observations down
to 5 m depth, Le. pt pt = (2.5 +2.1 +1.9 +2.2 +1.8)/5 = 2.1 MPa
The shaft resistance, taken from Fig. 138, is t h e n / ^ ~ 40 kPa
The bearing capacity is now obtained from Eq. (211). We find:
qf= 0.073 + 2.7-2 + 2-5-0.04/0.8 = 6.0 MPa

10.3 S e t t l e m e n t

S e t t l e m e n t s of piers and u n d e r g r o u n d walls can b e analysed by g e o m e c h a n i c a l m e t h o d s


b a s e d on deformation p a r a m e t e r s of the soil or b y e m p i r i c a l m e t h o d s related to
penetration r e s i s t a n c e obtained by s o u n d i n g . T h e m o s t reliable m e t h o d for settlement
estimation is d o u b t l e s s the p r e s s u r e m e t e r m e t h o d .
T h e principles of settlement analysis are the s a m e as those p r e v i o u s l y p r e s e n t e d in
connection with shallow foundations with d u e consideration to depth effects.

Example 38: Determine the settlement of the diaphragm wall described in Example 37 for a wall load
of 1.5 MN/m. The pressuremeter moduli Epn at depth intervals of 1 m, are 20, 25, 24, 30, 35, 34 and
40 MPa, counted from the foundation level downwards.

Solution: The pressuremeter moduli in the different part layers (see p. 147) are Ex = 20 MPa, E2 = 22.2
MPa, 3 = 24.5 MPa, 4 = 26.7 MPa and E5 = 34.4 MPa. These values yield Epri = Ex = 20 MPa and
Eprd = 4 / [ l / 2 0 + 1/(0.85-22.2) + 1/24.5 + 1/(2-26.7) + 1/(2-34.4)] = 22.6 MPa. The shape factors
according to Table 22 are Xd = 2.65 and = 1.50, and the rheological coefficients according to Table
21 ad = ax =1/3. The net load at the foundation level is 1.5 - 0 . 0 7 3 0 . 8 = 1.44 MN/m.
The settlement according to Eq. (152), p. 146, is:

= IM 2_0_6 2 , 6 5 , 0 8 m + 1 5018 =

9-0.8 22.6 0.6 3-20

Thus, the settlement to be expected is about 20 mm.


242 Dynamically loaded foundations

DYNAMICALLY LOADED FOUNDATIONS

1. INTRODUCTION

D y n a m i c actions on foundations are a source of disturbance that has b e c o m e increasingly


important. Foundations carrying machines or instruments that are sensitive to vibrations
have to b e designed in a way to eliminate the risk of functional disturbances. Vibrations
are oscillatory motions of various character. T h e simplest w a y of describing their
character is by stating their amplitude and frequency. In each project the kind of vibration
sources m e t with h a v e to be estimated as well as their influence upon the project in
question. Generally speaking w e differ between the following disturbance sources due
to vibrations:
Periodic disturbances. T h e s e include disturbances caused by gang saws, reciprocating
compressors, internal combustion engines, generators, turbines, mills, electric motors,
etc.
Stochastic disturbances. T h e s e include disturbances due to blasting, pile driving,
traffic, earth quakes, wind, water w a v e s , etc.
Impulse disturbances. T h e s e include disturbances by impact (for e x a m p l e by forging
h a m m e r s , p u n c h presses, stamping machines, etc.) or detonation (for e x a m p l e by dust or
gas explosions, explosive substances, etc.)
When dealing with vibrations of foundations it is generally assumed that the subsoil
can be replaced by a lumped mass-spring-dashpot system whith the mass of the
foundation representing the effective mass, the spring representing the elastic response
of the soil and the dashpot representing the total damping of the vibrating system. Here,
only the p r o b l e m of foundations subjected to periodic and impact disturbances will be
dealt with. In order to create a basis for the understanding of these problems, a short
introduction to the vibration theory will b e given.
T h e theoretical background to the influence of d y n a m i c actions on foundations pre-
sented in the following is mainly based on the standard w o r k by Rausch (1959).

2. LINEAR VIBRATION THEORY

2.1 Free vibrations

(i) Undamped free vibrations. Consider a body with mass m e a r n e d by perfectly elastic
springs with a spring constant equal to k (Fig. 170). In a state of rest the shortening of the
springs required to balance the body is <50. We thus have:
Dynamically loaded foundations 243

Amplitude

Fig. 170. Undamped free vibrations with one degree of freedom can be illustrated by a body resting on
a perfectly elastic spring system. The vibration becomes a sinusoidal function of time.

0 = mglk (280)

By pushing the b o d y d o w n w a r d s a distance z 0 from its position of rest and then letting
it free, the b o d y will start swinging around the position of rest. T h e oscillatory motion of
the b o d y thus achieved can be demonstrated graphically as s h o w n in Fig. 170. T h e
velocity of the b o d y is given by the tangent to the zlt curve. Since the spring is perfectly
elastic w e h a v e according to N e w t o n ' s second law:

d\
mg - (mg +kz) = m (281)
that is:

m'z+kz = 0 (282)

Intoducing the b o u n d a r y condition = ZQ at time t = 0, this equation has the solution:

fk
z=Zo cos(\/ t) = zo cos(<*v) (283)
Ym

w h e r e is the circular (angular) frequency, usually expressed in radians per second. T h e


distance ZQ represents the vibration amplitude.
We find that at time t will repeat itself after time t[l + (2/)] w h e r e is an integer.
T h e time of a complete cycle , the period of vibration, is thus:

2 im
= = 2 \ - (284)
V k
244 Dynamically loaded foundations

mg + kz

mg

Fig. 171. Damped, free vibrations with one degree of freedom can be illustrated by a body resting on
a spring-daspot system.

As can b e seen, the period of vibration is independent of amplitude and time.


T h e natural frequency fn of the system, expressed in cycles per second (Hz), is the
inverse of the period of vibration:

(285)

Introducing k - mg/S0 (whith 0 in m) w e get an alternative expression f o r / n :

(ii) Free vibrations with viscous damping. In reality, u n d a m p e d vibrations do not exist.
T h e oscillation amplitude of the b o d y will decrease with time and finally the body will
turn into a position of rest. D a m p i n g counteracts the m o v e m e n t of the body. Assuming
that w e h a v e to do with viscous damping which is directly proportional to the vibration
velocity (Fig. 171), N e w t o n ' s second law of motion takes the form:

mg - c i - (mg + kz) = m (287)

w h e r e c is the coefficient of viscous d a m p i n g .


Eliminating mg w e have:
mz +cz+kz= 0 (288)

T h e solution to this differential equation is of type = exp(Ar). Substituting this into Eq.
(288) w e find:
Dynamically loaded foundations 245

2
2 - - {c yjc -4mk)
2m

1
Case 1: c > 4mk. ( O v e r d a m p i n g ) .

A 1 )2 are real quantities and the solution takes the form:

= i^expUjf) + 5 i e x p ( ^ i ) (289)

w h e r e Ax and are constants of integration.


T h e system is o v e r d a m p e d and vibrations are completely prevented (Fig. 172a).

2
C a s e 2: c = 4mk. (Critical damping).

This yields = = - c/m, w h e n c e :

= (A2t + B2)exp(- ctllm) (290)

w h e r e A2 and B2 are constants of integration.


Inserting the b o u n d a r y conditions = Zq and z = 0 for t = 0 w e find:

Fig. 172. Viscous damping can give rise to three different time dependencies: (a) Overdamping, (b)
Critical damping, (c) Underdamping.
246 Dynamically loaded foundations

z = z 0 ( ^ + l)exp(-^-) (291)
Zm Zm

In this case no continuing oscillations will take place, only a d a m p e d displacement


b a c k to the position of rest (Fig. 172b). T h e system is said to b e subjected to critical
damping. T h e critical d a m p i n g factor is thus:

=2y/mk
Ccr (292)

Introducing the d a m p i n g ratio D = clccr w e h a v e :

~ = conD (293)
Zm

2
Case 3: c < Amk. (Underdamping).

A] 2 are c o m p l e x quantities and the solution takes the form:

1 2 2 2
A 1 2 = -nD y/c -c = n{-D\y/\-D ) (294)
2m

This gives the solution:

2 2
= [A 3 sm(conty/1 - D ) + B3 cos(conty/1 -D )] exp(-CunDt) (295)

w h e r e A 3 and 2? 3 are constants of integration.


Inserting the boundary conditions z = Zq and z = 0 for t = 0 w e h a v e :

D
= z0[ , w(dt) + c o s ( ) d0 ] ( - ) (296)
y 1
2
-D

2
w h e r e = ny/l-D
Eq. (296) shows that oscillation is continuing with a gradually decreasing amplitude
2
and with a d a m p e d circular frequency (d - cnyl -D (Fig. 172c).

2.2 Forced vibrations with viscous damping

Let us n o w a s s u m e that a vibrating force <2 0sinG#, acting in t h e direction, is enforced


u p o n the system s h o w n in Fig. 173. T h e law of motion yields:
Dynamically loaded foundations 247

kz + mg

Fig. 173. Forced, damped vibrations.

mg-cz-{mg +kz) + 0 sin()f) = m f

whence:
mz + cz = <2 sin(&)0
0
(297)

After a certain period of time w h o s e length depends on the d a m p i n g ratio the vibrating
system will keep p a c e with the vibrations enforced upon it. T h e system is subjected to
continuous, h a r m o n i c vibrations with a phase shift between the oscillating force and
the oscillating system.
T h e simplest w a y of solving the resulting effect upon the system due to the enforced
vibrations is by vectorial studies (Fig. 174). During oscillation, the velocity vector zcois
perpendicular to to the displacement vector and the acceleration vector zo is per-
pendicular to the velocity vector. Furthermore, the m a x i m u m amplitude z 0 is reached for
sincot = 1. Consequently, the following relation can be established:


kzo
(

Fig. 174. Vectorial presentation of displacement, speed and acceleration (left) and of forces acting at
maximum amplitude.
248 Dynamically loaded foundations

mzoo + cz0co + kzo = Q0 (298)

T h e conditions of equilibrium in the vertical and horizontal directions (Fig. 174, right)
yield:

1
kz0+ Q0cos(p = mzoco (299)

0 sin<p + cz0co = 0 (300)

F r o m Eqs. ( 2 9 8 - 2 9 9 ) w e finally obtain:

Go
2 2 2
(301)
^(k-mco ) + (cco)

ceo
tan - 2
(302)
k - m

F r e q u e n c y ratio flfn

Fig. 175. Dynamic factor as a function of the frequency ratio f/fn and the damping ratio D.
Dynamically loaded foundations 249

///>///////////

Fig. 176. Vertical impact. A body with weight mx is dropped from height h against a foundation block
with mass m supported by a spring system.

Now, if o represented a static load (it has to b e realised that the spring constant k which
represents the soil refers to the d y n a m i c r e s p o n s e of the soil even w h e n determining the
(50 value) w e would h a v e the displacement <50 = Q$lk.
T h e ratio of the m a x i m u m amplitude z 0 to the 'static' displacement <50 is defined as the
dynamic factor . Substituting c/2m for and g/50 for con
2
the d y n a m i c factor and the
phase shift can b e expressed as:

= ? = (303)
So yj[ 1 - ( I ) ] + (2 I )

2 2

2

2Dco/
n
tan = (304)
\-( )
2

T h e ratio / is usually substituted forflfn. T h e variation of the d y n a m i c factor with


the frequency r a t i o / / / n is shown in Fig. 175.

2.3 Impact

Let us now consider a case w h e r e the b o d y (Fig. 176) is subjected to an i m p a c t in the


direction caused by the b l o w of a falling weight. If the falling height is a s s u m e d to be h

and the m a s s of the falling weight is m, the impact velocity is - y/2gh . On the as-
sumption that that the duration of the i m p u l s e is short in comparison with the period of
vibration of the system, the initial velocity of the b o d y due to the i m p a c t will b e c o m e
equal to (cf Eq. 175, p. 176):
m
v=v t (1 + g) (305)
m +m

where e is the coefficient of restitution.


250 Dynamically loaded foundations

T h e force of impact is counteracted by the spring force F = kz. T h e energy c o n s u m e d


by the compression of the springs is equal to the energy c o n s u m e d by the impact, i.e.:

2
1 2 1 2 IF
-mv = -kz = (306)
2 2 2 k

w h e n c e the spring force:


F- vJmk (307)

In the case of static loading w e have:


F
s t a t = mg = k0

T h e d y n a m i c factor can be expressed as F/Fsm which yields:

(308)

T h e total load transmitted to the subgrade is the sum of the d y n a m i c and static loads,
that is:

^tot^statO+) (309)

T h e m a x i m u m amplitude z m a x caused by the impact is obtained by the relation:

(310)

Example 39: A tilt hammer with a weight mx of 1 t impacts against a concrete block with a base area
2
of 2 by 3 m and a weight m2 of 26 t. The concrete block is resting via a bed of oak timber on a larger
2
concrete block with a base area of 2.5 by 3.5 m and a weight m 3 of 5 0 1 , founded on a 10 m thick layer
3
of sandy gravel with density = 1.9 t/m and, according to seismic investigations, a Rayleigh wave
velocity vR of 375 m/s.

D'i = 1 t

1.8 m

m a
2m3m
2.4 m
m3 = 50t

2.5 m 3.5 m
Dynamically loaded foundations 251

Determine the the maximum contact pressure and the maximum amplitudes for the upper and lower
concrete blocks due to the impact of the tilt hammer if the impact energy is 30 kNm. The oak bed has
2
a thickness of 0.38 m and a modulus of elasticity = 1000 MPa. The base area of the oak bed is 2.3 m
and its coefficient of restitution e = 0.6.

Solution: The calculation is carried out in two steps, firstly with respect to the oak bed and secondly with
respect to the subsoil.
Step 1. The contact pressure between the upper concrete block and the oak bed is:
qx = 9 . 8 1 - 2 6 / 2 . 3 = 111 kPa
The corresponding deformation of the oak planks is:
3 3
(5 01 = 0 . 3 8 - 0 . 1 1 1 / 1 0 = 0 . 0 4 0 - m
The velocity v 1 of the tilt hammer when it hits the concrete block is found by the relation:
2
mxvx /2 = 30 kNm
whence
2 3 0
' nn< ,
V l =
~
The initial velocity Vj of the upper concrete block thus becomes:

VI= 7 . 7 5 ( 1 + 0 . 6 ) = 0.46m/s
1+26
and hence the dynamic factor :
0 4 6
^ =
= - =22.7
9 8 1 0 0 4 1 0 3-
VgSi V - ' - '
The corresponding static load and contact pressure:
3
Px = (1 + Qx)m^ = 23.7-26-9.81 = 6.0-10 kN
ql = 6.0/2.3 = 2.6 MPa
The maximum amplitude of the upper concrete block becomes:
3 3
z m ax = 22.7-0.04-10- = 0.9-10- m
Step 2. The contact pressure due to weight ra3 of the lower concrete block is:
50-9.81
q\i
H
= = 56 kPa
2.5-3.5
From the result of the seismic investigation we find the dynamic shear modulus:
2 3
G 0 = 1.15-1.9-375 = 307.3-10 kPa
Assuming that Poisson's ratio of the soil is 0.3, we have
6
0 = 0.80-10 kPa
From Eq. (146) and Fig. 100c, yielding Te = 0.75, we find:

6 3
= 0.75-56-2.5/(0.8-10 ) = 0.13-10~ m
Moreover:
62
vu = vi (1 + 0 . 6 ) = 0 . 2 5 m / s
26 + 50
Thus, the dynamic factor becomes:
025
= , -7.0
V 9 . 8 - 0 . 1 3 - ICH
Equivalent static load on the subsoil is:
m m
* n = S ( 2 + 3 ) + % S " * 3 = 9.81-76 +7-9.81-50 =4179 kN
from which the corresponding equivalent contact pressure becomes 0.48 MPa.
The maximum amplitude of the lower concrete block becomes:
3 3
W = 7 . 0 0 . 1 3 - 1 0 - = 0.9-" m
252 Dynamically loaded foundations

Fig. 177. General modes of vibration with six degrees of freedom. Translation movement designated
as , etc., rotational movement with ^, ^ etc.

3. NON-LINEAR VIBRATION THEORY

3.1 Introduction

T h e linear vibration theory which is applicable for vibrating systems with one degree of
freedom is sufficient for the solution of m a n y vibration p r o b l e m s . However, vibrating
machines often give rise to non-linear vibrations. Considering the general case, the
vibrating b o d y has six degrees of freedom: displacements in the directions of the three
coordinate axes and rotation around each one of the three axes (Fig. 177). T h e vibrations
m a y include every kind of combination of these motions.
A so-called rocking and sliding m o d e of oscillation represents a type of non-linear
vibrations w h i c h is c o m m o n l y occurring and, therefore, quite important for dynamically
loaded m a c h i n e foundations. Torsional oscillations represent another type of non-linear
vibrations that is also c o m m o n .

3.2 Rocking and sliding mode of vibration

In the case of m a c h i n e foundations, the horizontal, principal elastic axis does not go
through the centre of gravity of the foundation block but normally through the contact
surface b e t w e e n the foundation and the subsoil. A horizontal, d y n a m i c force gives rise
to a horizontal translation of the foundation b l o c k only if it is applied at the elastic centre
(the intersection point between the horizontal and vertical principal axes) of the
foundation (Fig. 178) and has the s a m e direction as the horizontal principal axis. A
rotational m o m e n t acting in the plane formed b y the vertical and horizontal principal axes
only causes rotation around O. In practice, a horizontal force acting on the foundation
s e l d o m passes through the elastic centre and therefore the vibrations induced are
neither purely horizontal, nor purely rotational. Instead t w o rotational vibrations will take
place around t w o fixed points situated on the vertical principal axis. This t y p e of
vibrations is defined as rocking-sliding.
Dynamically loaded foundations 253

Fig. 178. Foundation block subjected to a horizontal, dynamic load outside of its elastic centre.

Let us a s s u m e that point (Fig. 178) represents one of these fixed points and that the
foundation block is turned at the angle y/from its position of rest. T h e centre of gravity
of the foundation C is then displaced a distance = by. In order to induce this m o v e m e n t
of the foundation, it has to b e subjected to a turning m o m e n t around and a horizontal
force through O. T h e m o m e n t and the horizontal force in combination can be replaced
by a horizontal force through point A at a distance from the centre of gravity.
Introducing = \/kx and y/y = l/k^ (Fig. 179) w e have:

=Px + P(a + s)y/}<> (311)

y/=/b = P(a + s)y/y (312)

whence:

2
+s+s = ob + bs (313)

N e w t o n ' s second law of motion yields:

Fig. 179. Definitions of the parameters and .


254 Dynamically loaded foundations

Fig. 180. Forces belonging to poles A and in connection with rocking movement. The radius of inertia
/ is the geometrical mean of the distances a{ and a2.

m'S+ko = 0 (314)

+=0 (315)

where lis the mass moment of inertia of the foundation with respect to the horizontal axis
of gravity perpendicular to the plane of the figure.
112
But = /b and = k. Hence, Ilm = ab. Introducing the radius of inertia / = (Ilm) , we
find:

2
ab = i (316)

Thus, the radius of inertia iis the geometric mean of the distances a and b. Consequently,
if the distance / is set off from the centre of gravity C and if its end point is connected
with the fixed points A and B, then the angle AEB will be 90 (Fig. 180).
Substituting for b the radius of inertia according to Eq.(316), a can be solved from the
2nd degree equation:

2 , 1 , ^ , 2 .2, .2 riM\
a + - ( + s -)- (J1/)
s

whence:

a n = -a0 y/afi + i
2
(318)
Dynamically loaded foundations 255

Fig. 181. Principle for dividing up an arbitrary force into components acting in the directions of the three
principal modes of vibration.

1
x 2 2
where 0 - ( + s - i ) .
2s y/y

As expected w e get two values of a, of which the positive value ax corresponds to a in


Fig. 178 and the negative value - a2 corresponds to b. (In order not to mistake the term
b for width of foundation, the terms ax and a2, with positive sign +, will be used in the
following). The force Px, causing rotation around pole goes through pole A, and the
force P2, causing rotation around pole A, goes through pole (Fig. 180). The foundation
body with mass m can be replaced by two bodies, one at pole A with mass mx and the other
at pole with mass ra2, where:

m i = - ^ (319)
ax +a2

m2 = (320)
ax +a2

These mass 'points' are dynamically equivalent to the total mass of the foundation, i.e.
the resulting horizontal movement can be obtained by studying the horizontal movements
of the two mass points. If, for example, mass point mx is subjected to a horizontal impact,
the mass point m2in its capacity as centre of rotation remains at rest, and vice versa.
The corresponding natural frequencies are:

(321)
256 Dynamically loaded foundations

fa = J = ( 3 2 2 )
2y/2

where and 2 represent the horizontal displacements (in m) of the mass points mx and
m2 due to horizontal loads of magnitude gml and g m 2 , acting at the respective mass point.
The two displacement values and 2 are calculated on the basis of and in the
following way: the horizontal force gm x is replaced by a horizontal force gm x, acti ng through
the centre of gravity C, and a turning moment gmx(a + s). The displacement of mass point
mx the becomes equal to:

\ =gmx[x + (ax + j ) V y (323)


By analogy w e have:

2
82^gm2[x + (a2-s) yy (324)

If the foundation is subjected to external, horizontal oscillating or impulse forces, the


following equivalent forces are obtained with regard to vibrations around poles A and B:

Oscillating force with amplitude Q0 and frequency/:

Px = QXQ0 and P2 = 2Q ( 3 2 5 )

2 2
f2 n
where , = / , and
1 l
r ) - f / I
\fn\ ' Vn2-f \

Impact:

--gmx - ! m j (326)

7
P2 = v2m2y/J7o 2 ( 3 2 7 )

Now, by the aid of equivalent static forces, w e can easily solve the problems caused by
dynamic forces acting in arbitrary directions on the foundation. For example, the
dynamic force R (Fig. 181) can be divided up into three components, one vertical V and
two horizontal Hx and H2. Each component is then studied separately. It has to be noticed
that the dynamic factor will differ between the three components.
Dynamically loaded foundations 257

3.3 Torsional vibrations

To be able to study study the influence of any type of d y n a m i c forces, there remains to
analyse the influence of a torsional m o m e n t acting around the vertical axis through the
centre of gravity. A s this axis represents o n e of the three principal axes, w e simply h a v e
to do with p u r e torsional vibrations. In this case, N e w t o n ' s law of motion yields:

'+^ = 0 (328)

w h e r e Iz = the m a s s m o m e n t of inertia with respect to the vertical principal axis,


= the torsional angle around the vertical principal axis,
kfo = the torsional resistance in the foundation/subsoil interface per torsional
angle unit.
Eq. (328) is s y n o m y n o u s with Eq. (282). T h u s , the natural frequency is obtained by the
relation (cf. Eq. 285):
1 77
(329)

If the foundation is subjected to enforced torsional vibrations with amplitude M and


f r e q u e n c y / t h e n the equivalent static force is equal to:

Mt = QtM (330)
2
f
fnt
w h e r e Qt =
5 - /

4. T H E SUBSOIL AS D A M P E D SPRING S Y S T E M

In the elementary vibration theory it w a s p r e s u m e d that the m a s s of the spring system was
negligible. In reality, the subsoil affected by the oscillating b o d y has a considerable mass
which seemingly w o u l d invalidate a direct application of the results of the mass-less
spring response. Test experience, however, gives full e v i d e n c e that the elementary
vibration theory can b e applied provided that the subsoil is treated both as a mass-less
spring system and a d a m p e r with the d a m p i n g ratio D.
In Fig. 182, typical rheological m o d e l s are presented w h i c h are utilised for the analysis
of foundations subjected to vertical, horizontal and torsional d y n a m i c loads

4.1 Vertical linear vibrations

T h e spring constants utilised in the vibration theory n o w h a v e to b e related to the real


behaviour of the subsoil.
T h e p r o b l e m of determining the kz value is reduced to determining the elastic settlement
258 Dynamically loaded foundations

Rigid block
with equivalent mass
Section

f// - ///

5
Rigid block with equivalent mass and
Section mass moment of inertia around horizontal axis

JAW

Rigid block with equivalent


mass moment of inertia around vertical axis
Plan

Real system Equivalent system

Fig. 182. Examples of rheological models used for the analysis of dynamically loaded foundations

<50 under the w e i g h t mg of the foundation with regard to the a m p l i t u d e in question. Since
the vibration amplitudes are generally small, <50 can b e calculated on the basis of the
theory of elasticity with a m o d u l u s of elasticity determined by d y n a m i c m e t h o d s , for
e x a m p l e from the results of Rayleigh w a v e or shear w a v e velocity m e a s u r e m e n t s . T h u s ,
in the case of a shallow foundation with rectangular b o t t o m area, <50 can b e determined
according to Eq. (147) by substituting for EQ = 2 G 0 ( 1 + v) and q for mg which yields:

8 0 )
0= (3313
2 G 0( 1 + ) /

w h e r e / = the length of the foundation,


(mg) = the net load of the foundation at the foundation level,
= settlement coefficient according to Fig. 105 a-c.
Dynamically loaded foundations 259

1 1 1 I 1 1
' M l '

/
/

1
X
/

Ui
C3 b
CU

1 1 1 1 _ I L_ __L
0.1 0.2 0.4 0.6 0.8 1 2 4 6 8 10
Length to width ratio lib

Fig. 183. Parameters x, z and as functions of the side ratio lib.

According to Eq. (280), the spring constant can n o w b e expressed by the relation:

kz = mg/<5b (332)

Alternatively, kz can b e determined according to Barkan (1962):

(333)
1-

w h e r e A = the foundation area (A - bl),


= settlement parameter according to Fig. 183.
In this case, the influence of the foundation depth on the spring constant can be
estimated from Fig. 184.
T h e natural frequency of the foundation is obtained as shown previously from Eq.
(285) by the relation:
260 Dynamically loaded foundations

I I I I I I I I I I I I I

Fig. 184. Influence of embedment upon the vertical spring constant for a circular foundation with radius
R (After Kaldjian, 1969). Top curve represents rigid foundation, bottom curve weak foundation.

4.2 Rocking mode of vibration

Horizontal forces in the y direction that are not acting through the elastic centre of the
foundation (which, therefore, give rise to a rocking m o d e of vibration around the y axis)
produce a triangular stress distribution under the bottom area of the foundation.
According to Rausch (1959), this stress distribution can be replaced by a uniform
pressure as shown in Fig. 185. For rocking m o d e of formation, the spring constant k^ can
be chosen according to Gorbunov-Possadov & Serebrajanyi (1961):

Fig. 185. Theoretical stress distribution due to rocking mode of vibration. However, in practice only one half
of the foundation will be loaded in turns. Replacement distribution suggested by Rausch (1959).
Dynamically loaded foundations 261

vyAl (335)
1-

where according to Fig. 183,


/ = side length of the base of the foundation in the direction,
A = b a s e area of the foundation.
Unlike the state of vertical vibrations, a lengthening of the base of the foundation in the
direction of the acting forces increases the spring constant considerably and, consequently,
also the natural frequency of the rocking foundation.
L o o k i n g n o w at the effect of horizontal forces acting through the elastic centre of the
foundation, these are resisted by shear stresses in the foundation/subsoil interface. T h e
horizontal displacements induced by the horizontal forces in this case are thus due to
shear deformations. For a shallow foundation with base area A acted u p o n b y horizontal
forces through the elastic centre in the direction, Barkan (1962) proposes a spring
constant equal to:

(336)

w h e r e according to Fig. 183.

4.3 Torsional vibrations

T h e spring constant for a shallow circular foundation can be chosen according to Reissner
& Sagoci (1944) from the relation:

(337)

w h e r e R is the radius of the base of the foundation.

4.4 Soil damping ratio

W h e n dealing with a real foundation/soil system, t w o kinds of d a m p i n g h a v e to b e


considered: Internal damping, caused by internal energy losses (hysteresis), and geo-
metrical damping, caused by loss of energy due to the w a v e propagation through the soil.
T h e sum of these t w o d a m p i n g ratios constitutes the total d a m p i n g ratio.

(i) Internal damping. T h e internal d a m p i n g ratio D is defined as the ratio of energy loss
to m a x i m u m stored energy in a stress/strain cycle according to the relation (Fig. 186):

1 A
D = (338)
4 As
262 Dynamically loaded foundations

where A/ is the area inside the hysteresis loop,


As is the dashed area in Fig. 186.
The damping ratio is strongly dependent on the amplitude of the shear deformation.
However, for shearing amplitudes of the magnitude prevailing in connection with
dynamically loaded foundations, the values of internal damping ratio are usually within
the limits given in Table 34.
The influence of internal damping is generally negligible in comparison with geometric
damping.

(ii) Geometrical damping. The geometrical damping ratio of a circular rigid foundation
placed on a semi-infinite elastic medium is presented in Table 35. Rectangular foundations
with side lengths / (in the plane of rocking) and b, can be converted to circular foundations
with an equivalent radius Re. For the different modes of vibration the equivalent radius
can be put equal to:
112
Translation: Re = (/)
2 1 /4
Rocking: Re = [ / / ( 3 ) ]
2 2 1/4
Torsional: Re = [A(l + b )/(6n)]

4.5 Instructions for practical analysis

The analysis of dynamically loaded foundations, subjected to rocking and sliding or


torsion, can be facilitated by following a certain calculation sequence. The following
procedure is recommended:
collect necessary information concerning the mass of the machine, the magnitude,
position and frequency of the dynamic forces, subsoil conditions, etc.,
determine, for the type of foundation chosen, its mass, centre of gravity and the
required mass moments of inertia as well as the corresponding radii of inertia,
determine the foundation area, the required surface moments of inertia and the
required spring constants,
determine the natural frequency of the vibrating system according to the relation
l
fn = (2y/~5)~~ by the aid of the displacements [m] in the direction concerned,
determine the equivalent static loads affecting the subsoil. This is done by determining
the dynamic factor for the modes of vibration in question. The equivalent static loads
thus obtained should be multiplied by a fatigue coefficient generally assumed equal to
3. The design loads are thus given by the relation:
= 0 0

The contact stresses under the foundation base caused by the design load must not
exceed the allowable stresses.
Determine the maximum amplitudes obtained under the influence of the static load
o-
263
Dynamically loaded foundations

TABLE 34.
Internal damping ratio for different soils

Soil type Damping ratio D

Sand and gravel 0.01 - 0.03


Sand 0.03-0.07
Silty sand 0.03-0.10
Clay 0.02 - 0.05

0.4


03

00
G

C3

0
4 3 2 1
- - - - 1 10

Shear strain amplitude y, %

Fig. 186. The damping ratio D can be obtained the basis of the r/yhysteresis curve (left) and increases
strongly with increasing /amplitude.

TABLE 35.
Geometrical damping ratio D for various modes of vibration.

Mode of vibration Mass or inertia ratio Damping ratio

Vertical * z - ^ ^ ^ ^

0.288
7-8v m =

Horizontal ~ 32{1-)PR* y[x

8 ~ ~ pR* d+B^y/B
h

_ 0 . 5 0
Torsion = ~ c
264 Dynamically loaded foundations

Example 40: A rotating mass oscillator is placed on a rectangular block foundation with a rectangular
base area 2 by 3 m. The total weight of the mass oscillator and the foundation is 251. The rotating masses,
0.5 t each, have a lever arm of 0.6 m and produce a vertical oscillation with a frequency of 10 Hz. The
foundation is placed with an embedment of 1 m on a 10 m thick sand layer on bedrock. Seismic refraction
measurements have given a Rayleigh wave velocity in the sand equal to 210 m/s. The total density of the
3 3
sand above the groundwater level (at 2 m depth) is 1.8 t/m and below the groundwater level 2.1 t/m .

500 kg / = 1 0 Hz

/ W^- ft = ///SW 3 / // //"/ HV/7

3 in 2 m J
3
Sand with = 1 . 8 t / m

Solution:
Subsoil condition:
The initial (dynamic) shear modulus G 0 of the sand is calculated from the R wave velocity. Assuming
that the average value of the density is 2 t/m3 we have (see Eq. 59):
2 3
G 0 = \5 = 1.15 2 . 0 2 1 0 = 101 1 0 kPa
whence, assuming Poisson's ratio = 0.3, E0 = 2(1 + v ) G 0 = 263 MPa
Natural frequency:
The effective foundation contact pressure is (mg)/A = (25/6 - 1.8)g = 23 kPa
The condition-of-rest displacement 0 is found by Eq. (146) and Fig. 100c:
2
5o = ^ = 0 . 8 0 4 ^ - = 0.14.10- m

AE0 263
whence
1
fn = = 42 Hz.

Enforced vertical oscillation:


2
g = mrosincut = 20.50.6(2 10) sin(207tf) kN
whence g 0 = 2369 kN
Total damping ratio:
Geometrical damping ratio:
0.7 25
= /= 1.38 m; Bz = 3 = 3 :0.83;
4 p/? 4-2.0-1.38

A = ^ = 0 . 4 7 .
B7
Internal damping ratio can be assumed equal to 0.03, whence total damping ratio:
D
z,tot = 0-47 + 0.03 = 0.50.
The dynamic factor:
1
= = 1.03
/ 2
[ 1 - ( 1 0 / 4 2 ) 2 ] 2 + (20.50 10/42)
Dynamically loaded foundations 265

The maximum amplitude becomes equal to:


r W n , o n1 . 0 3 - 2 3 6 9 - 2 _ 3i n
zm ax = =0.804 = 2 . 5 - 10 m.
3
AEo 6 - 2 6 3 - 10

Example 41: A machine foundation of concrete with a base area of 8 m by 3 m and a height of 3 m is
affected by two oscillating masses, 10 t each. The oscillations take place without phase shift in the
horizontal plane, 3.8 m above the base area, with a circular frequency of 60 cycles/s and a maximum
centrifugal force of 25 kN. The foundation is resting on deep layers of sand with a total density of 1.9
2
t/m and a R wave velocity of 250 m/s. Calculate the maximum foundation pressure and the amplitude
at point D in the xlz plane.

' 50sin)f k N
VU
25cos)/kN 25cosft)/kN

7
/// = /v-^///

S a n d w i t h = 1 . 9 t/iiv^

Solution:
Oscillating force:
F r e q u e n c y / = 2 = 60/2 = 9.55 Hz
Maximum force amplitude 2Q0 = 50 kN
Subsoil:
The dynamic shear modulus of the sand is obtained from the shear wave velocity:
2 3
G 0 = 1.151.9250 = 137 1(> kPa
whence, for = 0.3, E0 = 355 MPa
Foundation block:
Mass: m =8-3-3-2.4 = 173 t
Centre of gravity:
5(173 + 2 0 ) = 173-1.5 + 20-3.8
s = 1.74 m
Mass moment of inertia (r designates the polar distance of an infinitesimal mass dm from the centre
of gravity):
2
/ = jV dm
For rocking in the xlz plane we have to calculate the mass moment of inertia
2 2
Iy = l(x + z )dm = Iyx+Iyz
We have :
2 2
Iyx= 173-8 /12 = 923 t m
2 2 2 2
Iyz = 173(3 /12 + 0 . 2 4 ) +20(1.26 + 0 . 8 ) = 225 t m
whence:
2
Iy = 923 + 225 = 1148 t m
The radius of inertia / = / I l m for the total mass 193 t becomes:
iy = 2.44 m
266 Dynamically loaded foundations

Spring constants:
The maximum displacement amplitude is obtained when the maximum force amplitude is directed
perpendicular to the longer side of the foundation, i.e. in the y direction.
Remains to determine and .
For determination of we have to know kx. According to Eq. (336) we have kx = </ . For
the side ratio lib = 8/3 (Fig. 183), = 1.0 whence kx = 355/S^3 - 1740 M N / m . For determination
of y/y we have to know k^. According to Eq. (335) we have k^ = [GQ/(1 - v^yyAL For the side ratio
8/3 (Fig. 183), = 0.65, whence k^ = 136-0.65-24-8/0.7 - 24250 MNm/rad.
3 6
Thus, = Ukx = 0.575 1 0 - rn/MN and = Ukw = 41.2 1 0 - rad/MNm.
Rotational centres of rocking movement:
Now, ax 2 can be calculated according to Eq. (318). First we calculate a0.
2 2 2 2
a0 = +s -L ) = ^ ( + 1 . 7 4 - 2 . 4 4 ) = 3.17 m
2s 2- 1.74^41.2
whence
2 2
a h2 = -3.17/3.17 +2.44
Thus, the two rotational centres of the rocking movement (poles 1 and 2) are defined by ax = 0.83 m
and \a2 I = 7.17 m.
Pole displacements:
The pole displacements ( of the upper pole 1 and 2 of the lower pole 2) are given by Eqs. (323
324 ):
3 3 6 3
{ =9.81193(7.17/8.0)10- [0.57510- + (0.83 + 1.74)^41.2-10 ] = 1.44 10" m
3 3 2 6 3
2 = 9.81 193(0.83/8.0) 10- [0.574 -f (1.74 - 7.17) 41.2 1 0 ] = 0 . 3 5 - 1 0 m
Natural frequencies:
3 1
fi = ( 2 / 1 . 4 4 1 0 - ) " = 13 H z
3 -1
fn2 = ( 2 / 0 . 3 5 " ) = 27 H z
Total damping:
Geometrical damping:
3 7 1 4 8
Bw = 5
= - !2 45 /4= 0.272
V
8 pRe 8 1.9 (24 8 / 3 )
whence:
Q , 1
Vy = i =0.23
1.272/0.272
Internal damping can be assumed equal to = 0.03.
Total damping: D = 0.26
Dynamic factors:
1
i = = 1.7
2 2 2
/[1-(9.55/13) ] + (20.269.55/13)

1
2 = -^ 2 2
= =2 1 . 1
/ [ 1 - ( 9 . 5 5 / 2 7 ) ] + (2 0.26 9 . 5 5 / 2 7 )
Equivalent external force system:
The external force system = 2)0 and M of around the elastic centre is replaced by an
equivalent force system distributed between poles 1 and 2. We find:
P = P{+P2
M = Pl(al+s)-P2(a2-s)
whence:
Dynamically loaded foundations 267

D M + P{g2-s)
M = 7
\ + a2
P(ai+s)-M
r2 =
a\ +a2
We have = 1.7-50 = 85 kN,
M= 1.7-50-3.8 = 323 kNm,
ax = 0.83 m; a2 = 7.17 m and s = 1.74 m.
Introducing these values, w e find:
= 9 8 kN
P2 = - 13 kN
The moment around the elastic centre produced by Px: Mx = 98(0.83 + 1.74) = 252 kNm
Maximum contact stresses:
193# 3-252

a m ax = +
2
= 103 kPa
8-3 3 8 /6
W = 3 - 8 5 / 2 4 = 11 kPa
Maximum amplitude at point D:
In direction:
Mx I 252-l(h3. 8
zD = = = 0 . 0 4 - 10 m
kyyl 24250-2
In the direction:
Pi Mxh 85-10-3 252-10-3.3
3
xD = + = + = 0.08 10 m
kx = V D + 1740
kyy * D = - ' 24250= 0.09 mm.
z 0 9 1 0 3- m
Total amplitude
268 Retaining structures

RETAINING S T R U C T U R E S

1. I N T R O D U C T I O N

O n e of the first analysed problems in foundation engineering, also considered as one of


the major ones, is h o w to evaluate the earth pressure distribution on earth retaining
structures and structural elements e m b e d d e d in soil. A theoretical analysis of the earth
pressure distribution in the upper and lower limit states was published by C o u l o m b
already in the beginning of the 18th century. With increasing urbanisation and consequential
need of utilising expensive land property, buildings are provided with an increasing
number of underground floors. D a m a g e s to underdimensioned retaining structures often
lead to heavy additional foundation costs. Therefore, a correct evaluation of the earth
pressure is no doubt imperative in foundation engineering.
Earth pressure is defined as the force or the stress acting at the boundary between a
structural element and the soil. T h e magnitude, distribution and direction of the earth
pressure depend on the relative m o v e m e n t between the structure and the soil, on dynamic
actions, frost action, etc. If the soil displacements b e c o m e large e n o u g h for soil failure
to occur, certain limits will be reached depending upon the shearing resistance of the soil,
the roughness of the structure and, even m o r e important, the size and direction of the
relative m o v e m e n t s between structure and soil. Obviously, in order to find the most
probable earth pressure distribution the m o v e m e n t of the retaining structure has to be
clarified.

2. E A R T H P R E S S U R E A G A I N S T R E T A I N I N G W A L L S

2.1 Introductory remarks

An extensive analysis of the various failure patterns that can occur depending upon the
m o v e m e n t of a retaining wall has been presented by Brinch H a n s e n (1953), s o m e of
which are presented in Fig. 187. As can b e seen, t w o types of failure patterns exist: one
characterised by the development of two groups of failure surfaces intersecting each
other by the angle 9 0 - 0 ' ( z o n e failure) and another by the d e v e l o p m e n t of a single failure
surface (line failure). A combination of zone failure and line failure is c o m m o n .
In the design of retaining walls, the p r o b l e m of earth pressure evaluation is generally
simplified to three cases:
earth pressure at rest (no m o v e m e n t between soil and wall),
active earth pressure (the wall is m o v i n g a w a y from the soil until failure takes place),
Retaining structures 269

> 1.26
> 1.21

1.26 > > 0.52

1.21 > > 0.67

0.52 > > 0.49


I
0.67>>.33

0.49 > 7] > 0

n= z/h
= ZJh

0 > > - 0 > > -

Fig. 187. Examples of different types of failure depending upon the situation of the rotational centre of
the wall and the direction of rotation.

Passive earth pressure (the wall is m o v i n g against the soil until failure takes place).
T h e m e c h a n i c a l b a c k g r o u n d to these three cases will b e elucidated in brief.

(i) Earth pressure at rest. In a soil m a s s with horizontal ground surfaceunloaded or


subjected to an evenly distributed surface loadit is evident that the horizontal normal
stresses \ acting in an imaginary vertical plane through the soil is i n d e p e n d e n t of the
direction of the plane. Moreover, due to reasons of symmetry, shear stresses cannot exist
in the plane. Vertical and horizontal stresses are thus principal stresses.
Now, if the imaginary plane is e x c h a n g e d for a rigid, unyielding wall, the stress sit-
uation in the soil m a s s obviously does not change. On the assumption that no vertical
displacement b e t w e e n the wall and the soil is taking place, no shear stresses along the wall
will appear. Consequently, in such a case, the roughness of the wall has no influence on
the stress situation.
T h e horizontal stresses acting at the soil/wall interface u n d e r these p r e m i s e s represent
the earth pressure at rest.
270 Retaining structures

Dense state
Earth pressure

Critical density

Critical density

From the soil Against the soil


Wall movement

Fig. 188. Influence of wall movement on the development of earth pressure exemplified for dense and
loose cohesionless soil

(ii) Active earth pressure. Now, let us a s s u m e that the retaining wall is m o v i n g from its
position of rest, away from the soil mass. T h e earth pressure is then gradually decreasing
as shown in Fig. 188. For cohesionless soil, two borderline cases m a y occur depending
upon whether or not the soil is in a dense or in a loose state. In the case of dense soil, the
earth pressure, first of all, rapidly reaches a m i n i m u m value, but, by additional m o v e m e n t
of the wall, it increases again until it b e c o m e s constant, independent of further wall
m o v e m e n t . On the other hand, if the soil is in a loose state, the earth pressure gradually
decreases until it b e c o m e s constant, independent of further m o v e m e n t . In both cases,
when the earth pressure has b e c o m e independent of the wall m o v e m e n t , the soil has
turned into a state of critical void ratio.
T h e active earth pressure is defined as the m i n i m u m earth pressure value. Obviously,
the wall displacement required for reaching a state of active earth pressure depends on
the relative density of the soil.
In the case of cohesive soil, the earth pressure rapidly reaches a m i n i m u m value deter-
m i n e d by the undrained shear strength of the soil. Further wall m o v e m e n t s a w a y from the
wall lead to an increase of the earth pressure until it reaches an u p p e r limit determined
by the residual shear strength of the soil. If the wall is arrested w h e n the earth pressure
has reached its lower limit, an increase will occur with time d u e to creep effects.
W h e n the wall is m o v i n g outwards, the soil particles, due to gravitational forces, will
m o v e d o w n w a r d s in relation to the wall. T h e shear stresses thus generated at the soil/wall
interface will counteract the d e v e l o p m e n t of failure and, consequently, cause a decrease
in the active earth pressure. However, the opposite case m a y occur w h e r e the wall is
m o v i n g d o w n w a r d s in relation to the soil. T h e n the shear stresses will contribute to the
d e v e l o p m e n t of failure and, consequently, cause an increase in the active earth pressure.

(iii) Passive earth pressure. T h e passive earth pressure is the opposite to active earth
pressure. W h e n the wall is m o v i n g in the direction towards the soil the earth p r e s s u r e will
gradually increase until it reaches a m a x i m u m determined bv the shear resistance o f the
Retaining structures 271

soil. For dense, cohesionless soil the m a x i m u m value will b e reached at a fairly small wall
m o v e m e n t , m u c h larger though than in the case of active earth pressure. At increasing
wall m o v e m e n t , the earth pressure will decrease until it reaches a constant value
independent of further wall m o v e m e n t s . For loose, cohesionless soil the earth pressure
will gradually increase with increasing wall m o v e m e n t until it reaches a constant value.
In both cases the constant values represent the earth pressure at critical void ratio.
In the case of cohesive soil, the earth pressure reaches a m a x i m u m determined by the
undrained shear strength of the soil. Further wall m o v e m e n t will cause a decrease of the
earth pressure until it reaches a m i n i m u m value determined by the residual shear strength
of the soil. If the wall is arrested after the earth pressure has reached its m a x i m u m it will
decrease with time due to creep effects.
W h e n the wall is m o v i n g inwards, towards the soil, the latter will travel upwards
against the gravitational forces. T h e shear stresses thus generated at the soil/wall inter-
face will prevent soil failure and consequently cause an increase in the earth pressure. On
the other hand, if the wall is m o v i n g u p w a r d s in relation to the soil the shear stresses at
the wall/soil interface will contribute to soil failure and consequently cause a decrease
in the passive earth pressure.

2.2 Earth pressure at rest

(i) Influence of soil weight and uniformly distributed surface load. T h e earth pressure
at rest can b e expressed by the relation:

Po= Ko'v (339)

w h e r e K0 = the coefficient of earth pressure at rest,


' = the effective overburden pressure.
T h e earth pressure coefficient K0 can be calculated on the basis of the theory of elas-
ticity. Since, by definition, no horizontal movement of the wall is taking place, i.e. eh = 0, we
have for an isotropic elastic m e d i u m :

, /
\- ( + ) = 0 (340)

w h e r e = P o i s s o n ' s ratio,
\ = the earth pressure p0.
In other w o r d s :

K0 = T ^ - (341)
1-

T h e evaluation of Poisson's ratio for natural soil is quite a difficult matter. Therefore,
272 Retaining structures

empirical formulae h a v e been developed, a m o n g which the m o s t w e l l - k n o w n is the one


presented by Jaky:
K Q = l - sin0' (342)

w h e r e 0 ' i s the angle of effective internal friction.


J a k y ' s formula has been applied on both non-cohesive and cohesive soils and has
shown good correlation with observational data in the normally consolidated state (see,
for e x a m p l e , Holtz and Kovacs, 1981). In the case of cohesive soil, Larsson (1977)
suggests that K0 be a function of the plasticity index IP or the liquid limit wL. For normally
consolidated mineral clays the coefficient of earth pressure at rest can be approximated to:

K0NC = 0.15 + 0.7 wL (343)

For organic soils ^0 decreases with increasing organic content (incresing wL).
If the soil is overconsolidated Jaky's formula leads to an underestimation of the K0
value. Alphan (1967) suggests that in this case K0 b e evaluated b y the relation:

^OOC = ^ONC(OCR)^ (344)

/
where K0NC = 1 - sin0 = Jaky's earth pressure coefficient,
O C R = <J'C/G'V = the overconsolidation ratio,
- soil parameter.
T h e overconsolidation ratio O C R for non-cohesive soils is generally difficult to
estimate, which m a k e s the practical application of Eq.(344) tricky. For mineral clays,
w h e r e the O C R value can be determined with satisfactory accuracy, the value can be
estimated by the relation:
e x p ( 0 . 1 0 - 0 . 6 5 w L)
= ) ( 3 4 5
L8

If the ground surface is inclined at angle to the horizontal (Fig. 189), K0 can be found
by multiplying the K0 values according to the a b o v e by the factor (1 + ) w h e r e is in
radians and \ is determined on the assumption of horizontal ground surface.

(ii) Influence of line load. T h e analysis is generally based on the assumption of the soil
behaving as an elastic m e d i u m . By definition of the earth pressure at rest, p0 can b e found
from the resulting horizontal pressure on the plane of s y m m e t r y between two line loads,
and the earth pressure at rest of a line load can b e obtained by the relation (Fig. 190):

4P
z 9
p0 = c o s 0 s i n (346)

Retaining structures 273

Fig. 189. Ground surface inclination . The depth coordinate is measured from an assumed horizontal
ground surface.

w h e r e = the angle between the vertical and the line connecting the point of action of
the line load as shown in Fig. 190,
= the depth b e l o w the ground surface.

(iii) Influence of point load. T h e earth pressure can be calculated according to the same
principles as in the previous case, i.e. p0 is the resulting horizontal stress in the plane of
symmetry b e t w e e n two point loads Q (Fig. 190):

z
Q3z p l-2v
(347)
Po = - [ ~ - ~ D
:
a a(a + z)

w h e r e a = distance from the point of action of the point load to the wall at depth z,
= radius vector from the point load to the wall.
Again w e h a v e difficulties in evaluating the value of Poisson's coefficient v. For sand
can be assumed equal to 0.3, while for water-saturated clays in the undrained state
is equal to 0.5.

? P(Q)

'/< . o//a~ tit = _ / / / w


/
/
/
/

/
/
>/

Fig. 190. The earth pressure at rest due to the action of a line load or a point load Q is calculated on
the assumption of an identical action on the opposite side of the wall.
274 Retaining structures

Fig. 191. Directions of active and passive failure lines in a zone of failure and their angles of inclination
to the minor principal stress ' 3 and the major principal stress \.

2.3 Earth pressure in the state of failure

As previously mentioned, w e differ between line failure (failure along a single failure
surface), z o n e failure and combined failure (partly zone failure and partly line failure,
Fig. 187). Z o n e failure is characterised by two sets of failure surfaces, the intersection
angle between the two being 90 ' (Fig. 191). T h e intersection angle between the
failure surfaces and the major principal stress is 4 5 - 0 7 2 and b e t w e e n the failure surfaces
and the minor principal stress 45 + 072.

(i) Inclination to the horizontal of failure surface intersecting the ground surface. Let us
assume that the ground surface has an angle of inclination to the horizontal equal to .
A s s u m i n g further that the ground surface is subjected to a vertical overload q per unit area
of the inclined ground surface, the equilibrium condition for the triangular soil element
A B C , shown in Fig. 192, yields:

c' sin / 3 s i n ( 2 v 0 + ' - ) cos ' + (q sin ' +


+ c ' c o s c o s 0 ' ) c o s ( 2 v o + 0 ' - / ? ) + qsin= 0 (348)

from which v 0 can b e solved.


If c = 0, w e have:

2 v 0 = a r c c o s ( - sin/sin0O - '+ (349)


Retaining structures 275

Fig. 192. Intersection between two sets of failure surfaces and ground surface. Passive failure.

and if also = 0:

v0 = 4 5 - 072 (350)

If 0 ' = 0, i.e. if w e h a v e an undrained case (a clay with c ^ c j w e find:

v 0 = - arccos(- sin) + (351)


2 cu

and if also - 0:

v 0 = 45 (352)

Example 42: Determine the intersection angle between the failure surface in a non-cohesive soil and
the ground surface under passive failure condition if the inclination of the ground surface = 10 and
the angle of internal friction of the soil 0'= 36.

Solution: Eq.(351) yields:


2 v 0 = arccos(- sinl07sin36) - 36 + 10 = 81
Thus, the intersection angles become v 0 - = 30.5 and 90 - 0 ' - v 0 + = 23.5

Example 43: Determine the corresponding value under active failure condition.

/
Solution: In this case, the angle 0 shall be inserted with a negative sign. This yields:
2 v 0 = arccosf- s i n l 0 7 s i n ( - 36)] + 36 + 10 = 153
The intersection angles become v 0 - = 66.5 and 90 - '- v 0 + = 59.5

Example 44: Determine the intersection angle between the failure surface in a cohesive soil under
undrained passive failure condition, if the ground surface has an inclination of 20 and the undrained
2
shear strength of the soil cu = 20 kPa. (a) The surface load q = 10 k N / m . (b) q = 0.
Fig. 193. Intersection between two sets of failure surfaces and wall.

Solution: According to Eq. (351) we have:


(a) v 0 = 0.5-arccos[(- 10/20>sin20] + 20 = 69.9
(b) v 0 = 0.5-arccos 0 + 20 = 65
The intersection angles become: (a) v 0 - = 49.9 and 90 - v 0 + = 40. 1
(b) v 0 - = 45 and 90 - v 0 + = 45

Example 45: Determine the corresponding value under active failure condition.

Solution: In this case, cu should be inserted with negative sign. We have:


(a) v 0 = 0.5-arccos [(10/20)-sin20] + 20 = 60. 1
The intersection angles become v 0 - = 40. 1 and 90 - v 0 + = 49.9
(b) v 0 = 65

(ii) Inclination to the horizontal of failure surface intersecting the wall Let us a s s u m e that
the wall has a positive angle of inclination to the vertical equal to a as s h o w n in Fig. 193.
A s s u m i n g further that the angle of friction b e t w e e n wall and soil is <5 and the adhesion
cw, the equilibrium condition for the triangular soil element A B C (Fig. 193) yields:

/ , /
,/ ~ x s i n [ T ! - c c o s 0 + ( c v v/ 2 ) c o t s i n ( 2 0 ) ]
c o s ( 2 v ! + ' + -2a) = - - - - ( 3 5 3 )
sin

w h e r e = c'+ c r ^ t a n ^ '
Retaining structures 277

'
/
Fig. 194. Failure condition when cjc ' - tan5/tan0 .

A s s u m i n g , as shown in Fig. 194, that cjc'= tan<5/tan0' or, alternatively, that c '= cw =
0, w e find:

2vx = arccos(sin<5/sin0O + 2 - - 0 ' (354)

In the undrained case where c''= c w and both 0 ' = 0 and <5 = 0, w e have:

2vx = a r c c o s ^ / c j + 2 a (355)
Example 46: Determine the intersection angle between the wall and the failure surface in a non-
cohesive soil under passive failure condition if the inclination of the wall a = 20 and the internal angle
of friction in the soil '- 36 on the assumption that (a) - 24; (b) =().

Solution: According to Eq. (354) we have:


(a) 2vx = arccos(sin247sin36) + 2-20 - 24 - 36 = 26.2
The angles of intersection with the wall become 90 - v{ +cc = 96.9 and v{ + '- a - 29. 1
(b) 2vx = arccos 0 +2-20 - 36 = 94
The angles of intersection with the wall become 90 -vx+a = 63 and vx + '- a = 63

Example 47: Determine the intersection angles between the wall and the failure surface for the case
given in Example 46 under active failure condition.

Solution: In this case both <//and shall be inserted with negative signs. We find:
(a) 2vx = arccos[sin(-24)/sin(-36)l + 2-20 + 24 + 36 = 146.2
The angles of intersection with the wall become 90 -vx+a= 36.9 and vx - '- a - 17. 1
(b) 2vx = arccos 0 + 2-20 + 36 = 166
The angles of intersection with the wall become 90 - vx +a - 27 and vx - '- a - 21

Example 48: Determine the intersection angles between the wall and the failure surface for the case
given in Example 46, case (a), under passive failure condition, if the relative movement between wall
and soil is opposite to the general case, i.e. the wall is moving upwards in relation to the soil.

Solution: In this case shall be inserted with negative sign and 0 ' w i t h positive sign. We have:
(a) 2vx = arccos[sin(-24)/sin36] + 2-20 + 24 - 36 = 1 6 . 8
The angles of intersection with the wall become 90 -vx + cc = 29. 1 and vx + '- a = 96.9
278 Retaining structures

Fig. 195. Mhr circle at failure, showing the relation between effective overburden pressure <fz and passive
and active earth pressures p a and p p . Lines a and represent the inclination of the active and passive
failure surfaces.

Example 49: Determine the intersection angle between the wall and the failure surface under active
failure condition in a cohesive soil with an undrained shear strength cu = 25 kPa if the inclination of the
wall a- 20 and the adhesion between wall and soil cw = 20 kPa. The relative movement between wall
and soil is opposite to the general case, i.e. the wall is moving downwards in relation to the soil.

Solution: The shear strength cu shall be inserted with negative and the adhesion cw with positive value.
We have:
2vj = arccos[20/(-25)] + 20 = 163. 1
The angles of intersection with the wall become 90 - Vj + a = 38.4 and vx - a = 51.6

2.4 Earth pressure due to zone failure

(i) Horizontal ground surface: smooth vertical wall. In the case of a smooth vertical wall,
no shear stresses will appear at the wall/soil interface. Vertical and horizontal stresses in
such a condition are principal stresses. In the active case, when the wall is translated away
from the soil, the horizontal stress b e c o m e s the minor principal stress corresponding
to active earth pressure while in the passive case, when the wall is translated against
the soil, the horizontal stress b e c o m e s the major principal stress corresponding to
passive earth pressure. T h e magnitude of active and passive earth pressures can be
obtained geometrically from the M h r stress circles (Fig. 195).
By geometrical and trigonometrical considerations w e find:

' = ' ~*

;= ' ~ ':

; (356)
p'p + cfz + 2c' c o t 0 ' p'a + cfz + 2c' cot 0 '

Expressing the active and passive earth pressures p'a and p'p with the customary forms
p'a = ' andp'p = ' w e find {Mohr-Coulomb's failure criterion):
Retaining structures 279

1 2c'
= (357)


&JNA

= + (358)

1 + s i n < 2
w h e r e N* = ^ = tan (45 4 - 0 7 2 )
1 -sin0'

If the effective cohesion intercept c ' i s e x c h a n g e d for attraction (Fig. 195) w e find:

sin0
f

2c'J . = 2a- (359)


1 - sin 0 '
and
2c' sin0 '
= 2a- (360)
1 +sin0 '

As can b e seen, Ka b e c o m e s negative w h e n cfz < 2c'Jl^ (or cfz < 2 a ).


1 - sin 0 '
As soil cannot take long-term tensile stresses, this will eventually lead to fissuring.
T h e active and passive earth pressure h a v e been expressed in terms of effective stress
conditions. T h e total earth pressure is obtained by adding the g r o u n d w a t e r pressure uw
(which is positive b e l o w but negative a b o v e the g r o u n d w a t e r level in the capillary
zone). Consequently, the total active and passive earth pressures are obtained by the
respective relations:
pa = Kao'z+uw=Kaaz + uw(l-Ka) (361)


Pp = '+
u
w = pz
K u
- w( p
K
-
1
) (362)

w h e r e = total overburden pressure at depth z.

90-' 90 + </>'

Fig. 196. Failure surfaces and major and minor principle stresses in the active and passive states of
failure.
280 Retaining structures

In the case of water-saturated cohesive soil in undrained condition (c'- cu\ '- 0 ) , we find:

P a = az-2cu (363)

2 c
Pp = <*z + u (364)

In the case of active earth pressure presented above, the major principal stress is in the
vertical direction and is equal to the effective overburden pressure. Consequently, the
shape and inclination of the failure surfaces in the zone of failure are known since they
always form an angle with the major principal stress direction of 45 - 072. Then the
failure surfaces b e c o m e plane and have an inclination of 45 + 072 to the horizontal. In
the passive case, the major principal stress is in the horizontal direction and is equal to
p'p. T h e failure surfaces become plane and have an inclination of 45 - 072 to the horizontal
(Fig. 197).
T h e stress vectors acting against one group of parallel failure surfaces b e c o m e parallel
with the other group of intersecting failure surfaces.
In the deduction of the active and passive earth pressures it was presumed that the wall
is either subjected to a translation or a rotation around an axis situated below the foot of
the wall or at the foot level {cf. Fig. 187). Moreover, the wall m o v e m e n t w a s presumed
to be large enough. (Generally, the wall m o v e m e n t required for the development of active
earth pressure can be assumed to be 0.1 % to 0.5% and for passive earth pressure 1 % to 5%).

(ii) Horizontal ground surface: rough vertical wall. N o w let us a s s u m e that we have to
deal with a retaining wall with a rough surface which entails full or partial friction (and
adhesion) between the wall and the soil. Consider a case where the wall is translated in
parallel against the soil and where the angle of friction between the soil and the wall is
equal to the angle of internal friction of the soil (<5= 00- T h e angles of inclination v 0 and
vj (Figs. 192-193) according to Eqs. (349) and (354) become v 0 = 45 - 072 and vx = - '. Thus,
the angles of intersection between the failure surfaces and the wall become 90 + 0'and 0
respectively. B e t w e e n the wall and the failure zone, close to the ground surface with plane
failure surfaces forming an angle with the ground surface of ( 4 5 - 072) the "Rankine

Fig. 197. Prandtl and Rankine failure zones in cohesionless soil. The Prandtl zone forms a transition
zone between the wall and the Rankine zone. Arrows indicate the major principal stress directions.
Retaining structures 281

Fig. 198. Passive earth pressure resultant and failure zone in non-cohesive soil with an internal angle
of friction of 30. ='.

zone", a zone of transition is formed with one group of plane failure surfaces, starting
radially from the line of intersection between the wall and the ground surface and another
group of logarithmic failure surfaces, intersecting the radial failure surfaces by an angle
of 90 0 t h e "Prandtl zone"(Fig. 197).
In the normal passive case, the earth behind the wall m o v e s in the u p w a r d direction
relative to the wall. Consequently, the frictional forces acting along the wall h a v e a down-
ward direction (Fig. 198).
A s s u m i n g instead that the frictional forces h a v e the opposite direction, /.e. -- \ w e
find = 90 and the angles of intersection with the wall 0 and 90 - ', respectively. T h e
size of the failure z o n e in this case b e c o m e s strongly reduced (Fig. 199) and so is the case
with the passive earth pressure also.
T h e failure zones obtained for = 0.5 ' and for = 0 are shown in Fig. 200.
The earth pressure coefficients for various values of wall friction are presented in Table 36.
For cohesive soils in undrained condition, w e find v 0 = 45 and vx = 0 in the normal
passive case, when cw = cu (cw acting downwards on the soil), and = 90 when cw - - cu.
T h e failure zones obtained for cw = 0.5c w and cw = 0 are shown in Fig. 2 0 1 .

Fig. 199. Passive earth pressure resultant and failure zone in non-cohesive soil with an internal angle
of friction of 30. = -'.
282 Retaining structures

Fig. 200. Passive failure zones in cohesionless soil for (1) = - 0 . 5 0 ' (2) (5 = 0 and (3) =0.5'. Angle
of internal friction '- 30. Centre of logarithmic boundary of Prandtl failure zone for -0.50'shown.

Introducing r = cw/cu, the normal component of the p a s s i v e and active earth pressures
in c o h e s i v e soil can b e evaluated by the empirical e x p r e s s i o n s :

Ppn = z +
2 c
id
1
+2r/3 (365)

Pan = - 2 c M \ / 1 + 2 r / 3 (366)

Fig. 201. Passive failure zones in clay for (1) cw = - 0.5c M, (2) cw - 0 and (3) cw = 0.5c u. Centre of
circular boundary of Prandtl failure zone for cw = - 0.5cu shown.
Retaining structures 283

TABLE 36 (Caquot et , 1973).


Earth pressure coefficients, referring to the stress vector, for various values of /'

' -1 -2/3 0 2/3 1 1 1/2 0 -1/2 -2/3

'= 10 1.060 0.814 0.704 0.656 0.649 1.66 1.55 1.42 1.25 1.17
15 1.057 0.724 0.589 0.537 0.531 2.20 1.97 1.70 1.38 1.26
20 1.042 0.636 0.490 0.442 0.440 3.1 2.55 2.05 1.53 1.33
25 1.017 0.554 0.406 0.364 0.367 4.4 3.40 2.45 1.65 1.40
30 0.981 0.476 0.333 0.300 0.308 6.5 4.60 3.00 1.80 1.46
35 0.937 0.403 0.271 0.247 0.260 10.5 6.50 3.69 1.93 1.51
40 0.880 0.336 0.217 0.202 0.219 18 9.60 4.60 2.08 1.54
45 0.820 0.274 0.172 0.163 0.185 35 15.0 5.83 2.20 1.55

(iii) Leaning wall: inclined ground surface. T h e intersection angles between the failure
surfaces and the ground surface on one h a n d and the wall on the other are given by
Eqs.(349) and (354). This m a k e s possible a graphical determination of the earth pressure
as will b e shown in the following chapter. Quite complete tables have been worked out
by Caquot etal (1973) and some of the values given there are summarised in Figs. 202-205.
An analytical expression for the earth pressure in the case of weightless soil can be
obtained by K t t e r ' s equations and the equilibrium conditions governing the respective
intersection angles between ground surface/failure surface and wall/failure surface.
Thus, for a cohesionless soil, the normal earth pressure coefficient from a vertical load
q per unit area of the ground surface is obtained by the relation:

K
Pnq = nQ (367)
where:
2 2 /
cos (v1-a) + s i n ( 0 + v 1- a )
Kn
e x p [ 2 ( v 0 - vx) tan0 J
c o s v 0 s i n ( v 0 - )[tan v 0 + tan ( ' + v 0 ) ]

with a , /3, v 0 and vx defined as shown in Figs. 1 9 2 - 1 9 3 .


T h e pressure coefficient for the earth pressure vector acting against the wall is:

Kq = Kn/cosS (368)

For a soil with cohesion, the cohesion term can b e expressed as:

/
/ : c = ( / i : n- l ) c o t a n 0 (369)

A conservative value of the earth pressure coefficient due to the weight of the soil is
obtained b y the approximate relation:
284 Retaining structures

-30 -20 -10 0 10 20 30

-30 -20 -10 0 10 20 30


Fig. 202. Earth pressure coefficient Ka and Kp as functions of wall inclination and ground surface
inclination . Angle of internal friction ' = 35 and = 0. (In Figs. 202-205 = \ p p = Kp/Q

Ky~Kncos(-a) (370)
F o r ' - 0 w e h a v e :

2
Kq = cos (371)

Kc = 2 ( v 0 - ) + s i n [ 2 ( v 0 - j3)] + sin[2(v! - a)
V l (372)

^ - c o s a + sinjScosiv^aysinCvo-jS) (373)
Retaining structures 285

1 1
0.06 1 1 I
-30 -20 -10 0 10 20 30

Fig. 203. Earth pressure coefficient^ as function of wall inclination aand ground surface inclination Angle
of internal friction '= 30 (top) and '= 35. = 2073. Broken lines refer to negative values of .

2.5 Graphical determination of earth pressure

If the wall and the ground surface are irregularly shaped, the analytical solution to the
earth pressure problem becomes very intricate and graphical solutions are then a helpful
means of solving the problem. In the majority of cases, although it is true only when the
wall is smooth, it can be assumed that the failure surfaces are plane. It is true that the
resulting error is on the unsafe side but in the case of active earth pressure the error is
generally negligible. However, in most practical cases, the use of combined plane and
curved failure surfaces is a relatively simple procedure and ought to be applied, at least
in the case of passive earth pressure.
286 Retaining structures

-30 -20 -10 0 10 20 30

Fig. 204. Earth pressure coefficient Kp as function of wall inclination and ground surface inclination
. Angle of internal friction 0' = 30 and = 072. Broken lines refer to negative values of .

(i) Plane failure surfaces. A great variety of methods for determination of the earth
pressure based on plane failure surfaces h a v e been presented, generally to determine the
respective earth pressures obtained at a n u m b e r of a s s u m e d inclinations of the outer limit
of the failure zone. In this way, the inclination that leads to the m o s t critical earth pressure
value can b e established.
A m o n g the m e t h o d s introduced, Engesser's method seems quite simple and practical
to use.
T h e use of E n g e s s e r ' s method is demonstrated in Fig. 2 0 6 . T h r e e plane failure
surfaces, ( 1 ), (2) and (3), h a v e been selected to represent the outer limit of the failure zone.
For each one, the resultant of the soil weight inside the failure z o n e and possible overload
is k n o w n in size and line of action as well as the lines of action of the earth pressure
resultants against the wall and the plane failure surface in question. T h e earth p r e s s u r e
Retaining structures 287

Fig. 205. Earth pressure coefficient Kp as function of wall inclination and ground surface inclination
. Angle of internal friction '= 35 and = 072. Broken lines refer to negative values of 5.

vector reaching the envelope to the earth pressure resultants against the failure
surfaces in the force vector polygon, represents the true active earth pressure.

(ii) Curved failure surface. T h e graphical determination of the earth pressure on the
assumption of a c o m b i n e d R a n k i n e and Prandtl passive failure z o n e will b e exemplified
for a retaining wall with an inclination angle of a = 2 0 , backfilled with n o n - c o h e s i v e soil
with an internal friction angle 0'= 35. T h e surface of the fill has an inclination angle of
j3 = 10 (Fig. 207). T h e angle of friction in the wall/soil interface is a s s u m e d to be
negative and equal to half the ' value (the wall is m o v i n g u p w a r d s relative to the soil,
i.e. opposite to the n o r m a l case). N o w the p r o c e d u r e is as follows:
288 Retaining structures

Fig. 206. Graphical determination of active earth pressure by Engesser's method.

First determine the angles v 0 and vx according to Eqs. (349) and (354). T h e angle v 0
defines one of the two sets of failure surfaces in the R a n k i n e z o n e A B C while the other
set of failure surfaces is fixed by the intersection angles 90 '. T h e angle vx defines the
intersection angle b e t w e e n the failure surface and the horizontal nearest to the wall in the
Prandtl failure z o n e A C D .
T h e centre of the logarithmic spiral, forming the Prandtl failure zone, is situated
on the boundary line between the Rankine and the Prandtl zones (or on its extension). It
is also situated on the failure line extended from the b o t t o m corner D of the wall and
forming an angle with the logarithmic failure surface of 90 - '. T h e intersection point
between these two b o u n d a r y lines forms the centre of the logarithmic failure surfaces
with the centre angle determined by the relation:
- lv 0 v xl
T h e distance [OC] is determined in relation to the k n o w n distance [OD] by the
correlation:
[OC] = [OD]exp(->tan ')
w h e r e in radians and:

S i n ( V
o + a )
[OD] = [ A D ] ^ -
sin(| v 0 - v x \ )

T h e shape of the c o m b i n e d failure zone is n o w established.


D e t e r m i n e the magnitudes and lines of action of the gravitational forces Wx of the
R a n k i n e failure z o n e and W of the Prandtl zone. T h e resulting force vector R of the forces
Q i , Wi and W i s obtained from the force polygon and the intersection point b e t w e e n Px
(resultant to Qx and Wx) and W. T h e resulting frictional force Q acting against the outer
b o u n d a r y D C of the logarithmic spiral can b e determined from its line of action t h r o u g h
Retaining structures 289

Fig.207. Example of graphical determination of the passive earth pressure Pp on the assumption of a
combined Rankine/Prandtl failure zone. ('= 35; =- 0.50'; = 20; = 10).

the centre of the logarithmic spiral and by the intersection point b e t w e e n Pp (the earth
p r e s s u r e resultant acting against the wall) and the resultant R.
H a v i n g established the line of action of Q, the force p o l y g o n can finally b e closed
w h i c h yields the value of the earth p r e s s u r e Pp.

Example 50: Determine the passive earth pressure against a retaining wall with height 5 m and an angle
of inclination = arctan(2/5) = 21.8 if the ground behind the wall consists of sand with a surface
3
inclination of 1:1.5 [ = arctan (1/1.5) = 33.7]. The sand has a unit weight of 18 k N / m and an angle
of internal friction '= 35. The friction angle in the soil/wall interface is = 20. The groundwater level
is below the foot of the wall.

Solution: The intersection angles v 0 and v 1 are:


v 0 = 0.5[33.7 - 35 + arccos(- sin33.7/sin35)] = 82.0
V! = 0.5[2-21.8 - 35 - 20 + arccos(sin207sin35)] = 2 . 0
The intersecton angle between failure surface and ground surface is v 0 - = 82.0 - 33.7 = 48.3
The opening angle of the Prandtl zone = v 0 - v 1 = 6 1 (1.064 rad.)
The length of the radii vector [OD] and [OC] of the logarithmic spiral are:
[OD] = y/29 -sin(82 + 3 5 - 21.8)/sin(82 - 21) = 6.13 m
[OC] = 6.13exp(1.064-tan35) = 12.92 m
This gives us the outer boundary BC of the Rankine zone.
The weight of the soil belonging to the Rankine zone can now be calculated. We find Gx = 102 kN.
290 Retaining structures

The weight of the soil belonging to the Rankine zone can be calculated as the weight of the
logarithmic spiral sector area [OCD] minus the weight of the triangular area [ADO]. The area [OCD]
is found by the relation:

[OCD] = -i^^-[exp(2u)tan0 )- 1] /

4tan0
2
which yields [OCD] = 46.03 m
2
The triangular area [ADO] = O ^ r A D H O D l s i n ^ +'- ) = 0.5-V29-6.13-sin34.2 = 9.28 m
Thus the weight G of the soil belonging to the Rankine zone is G = (46.03 - 9.28) 18 = 662 kN
The line of action of G can be found in a similar way. Start by determining the centre of gravity of
the sector area [OCD]. Its horizontal distance from the centre of the logarithmic spiral can be
determined by the relation:
_ 4 r 0 tan0'[(3tan0'sin&>i - c o s c u i )exp(3&>i tan0O-(3tan0'sintoo-coswo)exp(3)otan0')]
2 / ,
3(1 +9tan 0 )[exp(2a>i t a n 0 ) - e x p ( 2 ) o t a n 0 ' ) ]

where r 0 = the length of the radius vector of the logarithmic failure surface if extended to the vertical
below the centre and { - 0 = and CQ represents the angle between OD and the vertical through
O, which gives 0)Q = /2 - (vx + ') with vx and ' radians.
We find = 6.32 m.
The horizontal distance from to the centre of gravity of the triangle O A D is 2.72 m.
Consequently, the line of action of G is situated at a horizontal distance from equal to:
(46.03-6.32 - 9.28-2.72)/(46.03 - 9.28) = 7.22 m.
Knowing the value and lines of action of Wx and Wand the lines of action of Q{ and P{, their resultant
R and its line of action can be found. The intersection point between the lines of action of R and Pp yields
Retaining structures 291

the line of action of Q which makes it possible to close the force polygon (which is the requirement of
equilibrium). W e find Pp = 3000 kN which yields the passive earth pressure coefficient = 3000/
(29-18/2) = 11.5, corresponding to = 11.5-cos5 = 10.8.
This value can be compared with the approximate value obtained by Eq. (367)which yields Kpn = 8.5
and, consequently, = 8 . 5 - c o s ( - ) = 8.3 (Eq. 370).

Example 5 1 : Determine graphically the passive earth pressure in the soil conditions presented in
example 50 if the ground surface has an inclination of 1:3 and =- 20 (the soil behind the wall is settling
in relation to the wall).

Solution: In this case v 0 and vx are found by the relations:


v 0 = 0.5[18.4 - 35 + arccos(- sinl8.4/sin35)] = 53.4
vx = 0.5{221.8 - 35 + 20 + arccos[sin(- 20)/sin35]} = 77.6
The intersection angles with the ground surface are respectively v 0 - = 35.0 at point and 180 -
(125 + 35) = 20 at point A, the latter forming the boundary between the Prandtl and the Rankine zones.
The centre of the logarithmic spiral is found on the extension of this boundary line and a line forming
an angle of 90 - ' with the failure surface at D. The centre angle of the Rankine zone = v 1 - v 0 =
24.2 = 0.422 radians. The lengths of radii vector [OD] and [OA] are respectively:

[OD] = / 2 9 -sin(53.4 + 35 - 21.8)/sin(24.2) = 12.05 m


[OC] = 13.14exp(- 0.422-tan35) = 8.97 m
The logarithmic spiral area [ODC] is found by the expression given in Example 50, exchanging [OD]
for [OC]. We find:
2 2
[ODC] = (8.97) exp[(2-0.422-tan35) - l]/(4tan35) = 23.1 m
In order to calculate the centre of gravity of the logarithmic spiral area [ODC] we have to find the
value of r 0. The angle C0Q = ( v 0 + ') =(53.4 + 35)/180 = 1.543 radians which yields:
r 0 = 8.97-exp(- 1.543-tan0O = 3.05 m
292 Retaining structures

whence, according to the expression given in Example 1, - 6.85 m.


2
The triangular area [ADO] can be calculated at 32.5 m and its centre of gravity 8.0 m at a horizontal
distance left of O. The horizontal distance from to the centre of gravity of the Prandtl zone area [ADC]
is thus found from the relation:
(32.5-8.0 - 23.l-6.85)/(32.5 - 23.1] = 10.9 m
Proceeding in the way described in Example 50, we find Pp = 430 kN which corresponds to the earth
pressure coefficient Kpy = 1.65.

Example 52: Determine the active earth pressure against a 5 m high wall supporting a non-cohesive soil
3
with an internal friction angle of 35 and a density of 1.8 t/m . The ground surface is inclined 1:4 and
the wall is leaning backwards by the angle arctan(l/5) = 11.3. The angle of friction in the wall/soil
interface is 10. The groundwater level is below the base of the wall.

Solution: We have - - 11.3 and = 14.0. Now (//and should be inserted with negative sign.
v 0 = {arccos[- sinl4.0/sin(- 35)] +35+ 14.0}/2 = 57.0
vx = {arccos[sin(- 10)/sin(- 35)] +22.6 +10 + 35}/2 = 47.6
The intersection angles between the ground surface and the failure surfaces in the Rankine zone are
thus v 0 - = 33.6 at point and 90 + 0 ' + -vQ = 82.0 at point A. The latter forms the boundary
AC between the Rankine and the Prandtl zones. The centre of the logarithmic spiral is situated at the
intersection point between the extended line AC and a line forming an angle of 90 + 0'with the failure
Retaining structures 293

surface at D. The centre angle of the Prandtl zone = v 0 - vx = 9.4 = 0.164 radians. The lengths of the
vector radii [OD] and [OC] are:
[OD] = / 2 6 -sin(57.0 - 35 + 1 l.3)lsin(9.4) = 17.0 m
[OC] = 17.0exp[0.164-tan(- 35)] = 15.2 m.
Furthermore, we have 0)Q = vx - '= 12.6.
The logarithmic spiral area[ODC] and the distance from its centre of gravity to the vertical through
can now be calculated by the relations given in Example 50. Inserting the values of r 0, 0 and , and
2
' = - 35, we find [ODC] = 21.27 m and = 3.14 m.
2
The triangular area [OAD] = 17.0 / 2 6 -sin(^ + )/2 = 17.43 m and the distance from the vertical
through to its centre of gravity is 2.8 m.
2
The area of the Rankine zone is [ODC] - [OAC] = 3.84 m and the distance from the vertical through
to its centre of gravity is (21.27-3.14 - 17.43-2.8)/3.84 = 4.7 m.
Proceeding along the lines given in Examples 5 0 - 5 1 , we find PA = 46 kN which corresponds to the
active earth pressure coefficient = 46/(18-26/2) = 0.20.
The approximate value obtained by Eq. (367) is Kn = 0.23 which yields Kay= 0.208/cos 11 .3 = 0.21.

Example 53. Determine the passive pressure against a 4 m high wall supporting a cohesive soil with
3
an undrained shear strength cu = 20 kPa and a unit weight of 16 k N / m . The ground surface is inclined
1:5 and the wall is leaning backwards by the angle arctan( 1/4) = 0.245. The adhesion between wall and
soil cw= 10 kPa.

Solution: We have cjcu = 0.5, = arctan(0.2) = 0.197 and a = - 0.245. The angles v 0 and vx thus become:
v 0 = 0.5-arccos 0 + 0.197 = 0.982 (56.3)
V! = 0.5-arccos 0.5 - 0.245 = 0.279 (16.0)
The boundary failure line between the Rankine and the Prandtl failure zones has an inclination to the

D
0 100 200 kN/m
294 Retaining structures

horizontal of /2 - 0.982 = 0.589 (33.7). The centre of the Prandtl zone is situated at the intersection
between this boundary failure line and a line perpendicular to the tangent of the Prandtl zone at D. As
the boundary of the Prandtl zone is a circle with ;ts centre at O, the outer boundary of the Rankine zone
is also given. The radius [OD] is found equal to 5.97 m and the opening angle co= 0.704.
The forces acting against the boundary of the Rankine zone represent the shear forces and normal
forces along BC and AC. The shear forces are given by the shear strength times the lengths of BC and
AC while the size of the normal forces are obtained from the force polygon. The resultant of the forces
acting at the interface between the Rankine and the Prandtl zone R and its line of action is obtained from
the force polygon. Regarding the forces acting against the Prandtl zone we only know the size and lines
of action of the shear forces C 2 and Cw, but not the size, nor the lines of action, of the normal forces Pp
and N2. The line of action of depends on the magnitude of Cw, and the line of action of N2 (passing
through O) depends on the line of action of Pp. However, it can be approximately assumed that N2
coincides with the bisector to the angle . By this assumption the force polygon can be closed and we
find Pp = 450 kN/m.
This can be compared with the result obtained by Eqs. (372-373). Accordingly, we have:
Kc = 2(0.982 - 0.279) + sin[2(0.982 - 0.197)] + sin[2(0.279 + 0.245)] = 3.27
Ky= cos (-0.425) + sin(0.197)-cos(0.279 + 0.245)/sin(0.982 - 0.197) = 1.21
This yields:

Pp = 4.12(1.21-4.12 16/2 + 3.27-20) = 434 kN/m

2.6 Earth pressure distribution at enforced wall rotation


As s h o w n in Fig. 180, the level of the rotational axis of the wall is vitally important to the
d e v e l o p m e n t of failure and earth pressure distribution. T h e zone failure normally used
as a basis for earth pressure evaluation tales place only w h e n the rotational axis of the wall
is b e l o w or at the level of the foot of the wall. A combination of z o n e failure and line
failure takes place w h e n the rotational axis is below the mid-height of the wall and line
failure w h e n the rotational axis is above the mid-height of the wall. In consideration of
the situation of the rotational axis, the normal component of the earth pressure can b e
expressed by the relation (Lundgren & Brinch Hansen, 1958):

/
p'n = K/z + Kqq + Kcc (374)

and its tangential component by the relation:

/= /< + (Kqq + Kcc')tenq + cw (375)

w h e r e cw represents the wall adhesion.


T h e earth pressure distribution due to the dead weight of the soil has been a s s u m e d to
b e triangular while the distribution caused by overload q and cohesion c 'has been a s s u m e d
rectangular. H o w e v e r , the real earth pressure distribution is not possible to establish
unless it is a case of zone failure. Therefore, it is necessary to m a k e an a s s u m p t i o n
regarding the earth pressure distribution w h i c h hopefully agrees with reality. B r i n c h
H a n s e n (1953) suggested that the earth pressure b e distributed as s h o w n in Fig. 2 0 8 . T h i s
Fig. 208. Earth pressure distribution according to Brinch Hansen (1953).

is characterised by a pressure j u m p at a height a b o v e the foot of the wall equal to Z y =


where h is the height of the wall and is a function of the level of the rotational axis.
T h e normal c o m p o n e n t pm and the tangential c o m p o n e n t / ^ a b o v e the pressure j u m p

7] = /

Fig. 209. Graph for determination of the value for smooth wall ( = 0; top figure) and rough wall (<5= ').
296 Retaining structures

0.05

Fig. 210. Kyx a n d Kyy for smooth w a l l (<5= 0; top figure) a n d rough w a l l ( = 0 0

and the normal c o m p o n e n t p^ and the tangential c o m p o n e n t / ^ below the pressure j u m p


depend on the direction of the wall rotation and are determined by the relations:

K + Kc
P'xn = + qx<! cx ' (376)


P'yn = /
K
+ qy <l + &cf' (377)

fx = ,^+
A
(Kqx q + Kcxc )ian5q + c w
(378)

fy = Kyyfztan5Y+ (Kqy q + KyCtanq + c w (379)

T h e tangential c o m p o n e n t s fx and fy are positive if acting in the u p w a r d direction.


T h e values of the parameters included in Eqs. ( 3 7 6 - 3 7 9 ) can be obtained from Figs.
210-215.
Retaining structures 297

Fig. 211. Kqx and for smooth wall ( = 0)

Example 54: Determine according to Brinch Hansen's method the earth pressure distribution against
a retaining wall, 5 m in height, if the wall rotates clockwise around a horizontal axis 1 m below the top
of the wall. The soil consists of sand with an internal friction angle '= 30. The friction angle in the
2
soil/wall interface = 30. The ground behind the wall is loaded with q = 20 k N / m .
Solution: W e have = ZJh = 4/5 = 0.8 and a positive rotation. According to Fig. 208 we obtain = 0.87
which yields Zj = 0.87-5 = 4.35 m. From Figs. 209 and 210 we find = 5.6, = 0.22, Kqx = 1.8 and
Kqy = 0 . 1 8 . The friction along the wall, according to Fig. 205, is t a n 5 7 = t a n ^ = - 0.58.
The earth pressure distribution follows the relation:
above a depth of 0.65 m ( < 0.65 m) / / n = 5.6-18-z + 1.8-20 = 101z + 36 kPa
below a depth of 0.65 m (z > 0.65 m) p'n = 0.22-18-z + 0.18-20 = 4.0z + 3.6 kPa
The earth pressure vector is inclined 30 downwards and is equal to:
2 2
/ / = / / n ( l + 0 . 5 8 ) " = 1.156/7;

Example 55: Compare the earth pressure distribution according to Brinch Hansen's theory with the
passive earth pressure distribution according to the classical Coulomb theory if the rotational axis in the
case given in Example 54 is situated at the foot of the wall.
298 Retaining structures

= ZJh

Fig. 212. Kqx and Kqy for rough wall (5 = )

Solution: We have = ZJh - 0, whence = 0 and = 5.8, Kqx = 5 . 0 and tan# r = tan<^ = 0.58. The
earth pressure vector is inclined 30 upwards and the earth pressure coefficient becomes Kp = Kx( 1 +
2 1 /2
0 . 5 8 ) = 1 . 1 5 6 ^ . We thus have = 5.8-1.156 = 6.7 and Kpq = 5.0-1.156 = 5.8.
The passive earth pressure coefficient according to Coulomb, Eq. (358), becomes = 6.5.

2.7 Design of gravity walls

In the design of gravity walls three m o d e s of failure have to be considered:


Insufficient bearing capacity
Rotational failure
Sliding failure
A s regards the bearing capacity this can be analysed according to the m e t h o d s given
for spread footings.
T h e possibility of rotational failure is strongly coupled with the bearing capacity
problem. However, rotational failure m a y occur even in cases w h e r e the bearing capacity
Retaining structures 299

40-^

30

-20^

- u -

-40-
J
- 20

I I 1 1 LI

-
1
.40--"
T T
-


30^
\
\
-20
\\
OP-H

;40_

. W 0
_ 1 ., -L

-4 -1 -0.1 0 0.5 1.4 2 5


r\ = ZJh

Fig. 213. Kxc and Kyc for smooth wall (cw - 0)

is not e x c e e d e d , for e x a m p l e if the gravity wall is f o u n d e d on h a r d rock. T h u s , in such


a case it is n e c e s s a r y to c h e c k that the m o s t u n f a v o u r a b l e line of action of t h e resulting
force against t h e wall falls i n s i d e the toe of the wall.
Sliding failure m a y o c c u r if t h e horizontal c o m p o n e n t of the resulting force e x c e e d s
the frictional r e s i s t a n c e in the interface b e t w e e n the b a s e of the wall a n d t h e subsoil.

Example 56: A 6 m high cantilever wall of concrete with a base width of 5 m is backfilled with sand
3
with an internal angle of friction of 35 and a unit weight of 19 kN/m . The depth of embedment of the
3
front of the wall is 1.5 m. The unit weight of the concrete is 24 k N / m . The sand fill, the surface of which
2
is horizontal, is subjected to an evenly distributed load of 20 k N / m . The thickness of the concrete wall
and the bottom plate is 0.8 m. Check whether or not the stability conditions are satisfied for at rest earth
pressure condition.

Solution: The earth pressure coefficient K0 = 1 - sin35 = 0.43. The horizontal component of the resul-
ting force against the retaining wall from the load is thus equal to 0.43-6-20 52 kN/m, acting at mid-
2
height of the wall, and from the self-weight of the soil 0.43-19-6 /2 147 kN/m, acting at the lower third
of the wall. The total horizontal force is thus / / = 199 kN/m, acting at a height above the base of the wall
equal to (52-3 + 1472)/199 = 2.25 m.
300 Retaining structures

7\ = ZJh

Fig. 214. Kxc and Kyc for rough wall (cw = cO

The vertical component V= (4.35-19 + 0.65-24)-5.2 + 5 0 . 8 - 2 4 = 607 kN/m, acting 2.5 m inside the
toe of the wall.
The inclination of the resultant is arctan( 199/607) 18 which, ignoring the restraint from the earth
pressure at the passive side of the wall, is also the friction angle required to prevent sliding.
The line of action of the resultant intersects the base of the wall at a distance form the toe of b'= 2.5
- 2 . 2 5 - 1 9 9 / 6 0 7 - 1.76 m.
Retaining structures 301

"- - < l 7 --
40^
30- %
\
20 40^
>\ J

>
0

\
20
30 - 1

40 - I
\ ""
40- i 1 t 1

" T T "
20
30
40 ^ ( ^ \
-20'
ono
-40

40^
/
ft
\ ^
I 1 1
- 4 -1 -0.1 0 0.5 1 1.4 2 5

= ZJh

Fig. 215. Diagram for determination of tanc)^ tan8 q and cjc'

For a friction angle of 35 we have Nq = 33 (Eq. 112) and = 45 (Eq. 114). According to Eqs. ( 127)
and (129) we find:
3
iq = (\ - 0.7 199/607) = 0.46
3
i y = ( l - 199/607) = 0.30
The bearing capacity of the retaining wall is thus:
Vf= 1.5 19-330.46 + 1.76-19-45 0.3 - 884 kN/m (> 607 kN/m)
This result is unsatisfactory in practice. Thus the design value of the friction angle, for example
according to Eurocode 7, is chosen as arctan (tan07l .25) which in our case would lead to a design value
of '= 29. This gives K0 = 0.52, Nq ~ 17 and ~ 18. The increase in the K0 value yields a horizontal
force = (0.52/0.43) 199 = 241 kN/m. Now the line of action of the resultant intersects the base of the
wall at a distance from the toe of b'= 2.5 - 2.25-241/607 = 1.61 m. The inclination factors will be changed
to ^ = 0.38 and i r = 0.22.
To satisfy the demands of Eurocode, the bottom plate must be corbelled out along the front of the
wall. Assuming a corbelling width of x m we find the correlation:
302 Retaining structures


Ph I I I Iii I I I , + dz
A ydz \
Phtan6

Fig. 216. Stresses acting on a slice element of the silo.

607 = 1.5 19-170.38 + (x + 1.61)-19-180.22


whence = 4.0 m

3. EARTH PRESSURE AGAINST OTHER STRUCTURES

3.1 Silo pressure

T h e first theoretical solution to the p r o b l e m of silo pressure was published by Janssen


(1895). Janssen's solution was later modified by Knen (1896).
T h e silo pressure according to Janssen-Knen is deduced in the following way.
Consider a horizontal slice, d z in thickness, of the silo (Fig. 216) at depth below the
surface of the in-filled material (assumed horizontal). T h e silo pressure acting in the
normal direction against the walls of the silo is termed p h and the frictional forces are
p / ?tan<5 w h e r e 5 i s the angle of friction between the silo walls and the material inside. T h e
unit weight of the material is .
Equilibrium in the vertical direction is satisfied by the relation:

-) + Cph tan = 0 (380)


dz.

w h e r e = internal cross-sectional area of the silo,


C = internal circumference.
Introducing AIC = R (the so-called hydraulic radius) and the silo pressure coefficient
= pjJOy, the relation takes the form:

(381)

w h e r e = t a n i s the wall friction coefficient.


Integrating and introducing the boundary conditions, the horizontal c o m p o n e n t of the
silo pressure ( p h ) b e c o m e s equal to:
Retaining structures 303

Fig. 217. Idealisation of a silo with eccentric discharge gate.

Ph = [1 - e x p ( ) ] (382)
ji R

Originally, the value was determined with regard to the earth pressure obtained
during filling of the silo or in at rest condition. Later investigations (Bergau, 1959;
Turitzin, 1963), initiated by a number of silo damages, showed that the most critical
condition occurred during emptying of the silo where its pressure was found to be
approximately twice as high as it had been during filling. The reason behind this fact is
that of arching phenomena. Special arrangements can be made in order to create an even
outflow of the fill material and thus prevent arching. The discharge gates should be placed
centrically. In the case of eccentric discharge gates, the silo pressure according to
Janssen-Knen will have to be corrected. The German code D I N 1055:6 recommends to
consider the eccentricity by assuming an idealised silo, with its centre coinciding with the
eccentric discharge gate (Fig. 217).
In a supplement of 1977 to D I N 1055:6 it is recommended to use the correction factor:

c = 1 +0.20t + e/1.5) (383)

where k = material parameter (1 for organic and 0 for inorganic material),


e = eccentricity of the discharge gate (m),
R = hydraulic radius (m).
The value of depends on the internal friction of the 'filling-in' material and on the
wall friction. Assuming that w e have to deal with an active earth pressure condition
during filling and a passive earth pressure condition during emptying (caused by
arching), the earth pressure coefficients will be governed by a negative value of the
friction angle (the movement between earth and wall is opposite to the usual case). This
would result in the silo pressure coefficients shown in Fig. 218.
It is doubtful, however, whether real active and passive states can occur since this
requires a relatively large earth<r->wall movement. Therefore, from a practical viewpoint,
304 Retaining structures

-1 -2/3 -1/3 0 0 1/3 2/3 1


' /'

Fig. 218. Theoretical silo pressure coefficients in the active and passive state.

the silo pressure coefficient during filling, should not be b e l o w the earth pressure
coefficient at rest:

Xf=K0=l -sin0' (384)

T h e value of to b e used in the design according to D I N 1055:6 is 0 . 7 5 0 ' f o r filling


and 0 . 6 0 ' f o r emptying.
On the basis of experimental evidence, the silo pressure coefficient during emptying,
, can b e chosen as 2Xj. This principle has been followed in the G e r m a n code D I N 1055:6
w h e r e is put equal to 0.5 (corresponding to 0'=3O) in the case of filling and 1.0 in the
case of emptying.
Of course, arching can only take p l a c e if the surface of the silo walls is rough enough
to carry momentarily the granular d o m e s that h a v e a tendency to form. T h e silo pressure
during e m p t y i n g is d y n a m i c in character with relatively large oscillations around the
mean. T h e application of Xe yields the m a x i m u m values of the silo pressure.

Example 57: A concrete silo with an inner diameter of 8.5 m and a height of 20 m shall be designed
for the storage of iron ore. The silo has two rectangular discharge gates (2.8m by 5.0 m) located 2.1 m
from the centre of the silo. The iron ore to be stored has an iron content of 55% by volume and shall be
filled into the silo from the top by means of a conveyor. The porosity of the ore material filled into the
silo is estimated at 45% and the water content w = 10%. The angle of internal friction is '= 40 and
the angle of wall friction 5 = 0.750'. Determine the design silo pressure.
Retaining structures 305

Solution: The grain density pg of the ore material is obtained by the relation:
P8 = PftPj\Pk - 0.55-(p fe - p s t) ]
3
where p f e = specific weight of pure iron = 7.85 t/m
3
p s t = specific weight of stone material = 2.75 t/m
3
This yields pg = 4.2 t/m
The bulk density of the material becomes equal to (Eq. 12):
3
= pg{\ - n){\ +w) = 4.2-0.55-1.1 = 2 . 5 4 t/m
Due to the eccentricity of the discharge gate, the hydraulic radius becomes equal to:
2
R = (8.5 /4 + 22.18.5)/(8.5 + 2-2.1) = 3.0 m
The correction factor with regard to eccentric discharge gates:
c = 1+0.20-2.1-3.0/1.5 = 1.84
With the values of 0'and 8given, we find (Fig. 211) a discharge pressure coefficient Xe - 1.4 while
= tan(0.75-40) = 0.577. Thus the design silo pressure varies with depth below the top surface of the
material filled into the silo according to the relation:

OA 2 . 5 4 - 9 . 8 1 - 3 . 0 ri 0.577 1.4 1
Ph = 1 . 8 4 [l-exp(- )] = 2 3 8 [ l - e x p ( - 0 . 2 6 9 z ) ]

The pressure during filling becomes:

2.54-9.81-3.0M , 0.577 0.5 - _1 , _ n Q. ^


ph = [l-exp( )] = 1 3 0 [ l - e x p ( - 0 . 0 9 6 z ) J
y
0.577 3.0
The maximum lateral design pressure just above the bottom of the silo (z = 20 m) is obtained during
emptying of the silo (completely filled) and becomes 238 kPa.
The maximum pressure during filling to full height becomes 111 kPa, i.e. 0.47 times the pressure
during emptying.
Using instead ^= (1 - sin^O = 0.36 and = 2^= 0.72, the maximum values of the silo pressure
become 223 kPa during emptying and 97 kPa during filling.

3.2 Earth pressure against underground pipelines.

In the design of buried pipes, one m u s t distinguish b e t w e e n rigid pipes and flexible pipes.
A pipe can be considered as rigid if the pipe stiffness Sp fulfils the condition:

3
Sp = -Ep(b) > 5000 kPa (385)
3 A + tp

w h e r e Ep = m o d u l u s of elasticity of the pipe material, in kPa,


tp = thickness of the pipe wall,
Dt = inner diameter of the pipe.
Thin-walled steel pipes and plastic pipes can be considered to b e completely flexible.

(i) Rigid pipes. T h e earth pressure against buried pipes d e p e n d on the conditions of
installation. Generally speaking w e differ b e t w e e n trench pipelines and embankment
pipelines. T h e first is defined as a pipe installed in a narrow trench e x c a v a t e d in the soil
and covered with back-fill while the second is defined as being covered with an
e m b a n k m e n t fill, the width of w h i c h is large in c o m p a r i s o n with the width of the pipe.
Thus, a pipe installed in a w i d e trench can b e considered as an e m b a n k m e n t pipe. T h e
306 Retaining structures

Fig. 219. Trench condition.

m i n i m u m width of such a trench is referred to as the transition width. T h e transition width,


measured at the crown level of the pipe, can b e assumed equal to 4 times the outer diameter
of the pipe.
A great n u m b e r of design m e t h o d s h a v e been developed with reference to different
types of pipe installation. Extensive information regarding types of pipe installation and
relevant design m e t h o d s is presented by Liedberg (1991). In this connection, only the
most c o m m o n semi-empirical design principle, the so-called M a r s t o n / S p a n g l e r theory,
will b e presented.

Fig. 220. Embankment condition.


7
Retaining structures

In the case of trench pipelines, the back-fill is a s s u m e d to settle in relation to the natural
soil at the sides of the trench. Consequently, shear stresses at the interface between back-
fill and natural soil are acting on the back-fill sides in an u p w a r d direction (Fig. 219).
Now, the vertical earth pressure at the crown level, exerted by the weight of the soil,
can b e deduced according to the s a m e principles as applied in the silo case. With the
symbols used in Fig. 2 1 9 , the condition of equilibrium for the soil element inside the
trench yields:

^b-Yb+2Kcfzt<m(l) ' = 0 (386)


dz.

w h e r e b = the trench width,


= the earth pressure coefficient,
G'Z = the effective overburden pressure,
'- the angle of internal friction of the backfill,
Y - effective unit weight of the soil.
Integrating and introducing the boundary condition ' - 0 for = 0, the vertical stress
at depth h (the crown level) is obtained by the relation:

^ Yb 2Kph
oi = ^ r [ l - e x p ( ) ] (387)
2Ku b

where = tan '.


T h e vertical load per unit length of the pipe at the crown is thus equal to:

2
Yb 2Kph
= ^[l-exp(-^-)] (388)
2Kji b

T h e value of the earth pressure coefficient K, according to Marston, can b e assumed


equal to the active earth pressure coefficient. However, it is obvious that its value depends
on whether or n o t the backfill in the trench has been compacted. T h e r e is a possibility,
due to compaction of the backfill, that tends to the passive earth pressure value given
in Fig. 2 1 2 for = - ', i.e. K^\. According to Wetzworke (1960), the l v a l u e can be
chosen equal to 0.5 for loose backfill and equal to 1.0 for c o m p a c t e d backfill.
In the case of embankment pipe lines (Fig. 220) placed on a less compressible soil than
that of the e m b a n k m e n t , the settlement at a certain depth below the e m b a n k m e n t surface
will b e larger on either side of the pipe than directly above it. T h e condition of equlibrium
for the soil element with height dz is now obtained by the relation:
308 Retaining structures

d&7
--2<\' = 0 (389)
dz

w h e r e D = the outer diameter of the pipe.


A s s u m i n g hs to b e t h e height above the c r o w n of the pipe w h e r e the settlement is equal
both a b o v e and outside the pipe, w e find, after integration and insertion of the boundary
condition ' - Y(h - hs) for - 0, the following overburden pressure at depth z\

YD IKiiz IKiiz
&z = [ e x p ( ) - 1] + / ( - , ) ( ^ ) (390)
2Kji D D

T h e crown load QCT per unit length of the pipe is given by the relation Da\ at depth
- hs, w h e n c e :
2
YD 2Kphs , 2Kphs
Q* = ^ [ e x p ( - ^ ) - 1] 4 - / D ( A - ^ e x p ( p ) (391)
2Kji D D

T h e depth h-hsto the level of equal settlement is an important parameter w h o s e value


depends on the rigidity of the pipe and the compressibility characteristics of the natural
ground on the one h a n d and the e m b a n k m e n t on the other. For the determination of hs,
a reduction factor rsd is introduced with reference to the relative vertical settlements at
the crown level of the pipe, above and outside of the pipe, according to the relation:

j . + - f r + M ) ( 3 9 2

w h e r e sg = settlement of the natural ground surface outside of the pipe,


sj- = settlement due to compression of fill with height aD alongside the pipe,
sp = settlement of the invert of the pipe,
A D = vertical deflection of the pipe itself.
T h e hs value can be calculated by considering the settlement s of the level of equal
settlement due to the weight of the fill above.
A b o v e the p i p e w e have:

rhS/i , exp(2Kpz/D)
V P J
, K (h-hsWD
s = L(h-h
0 s)Y ; -dz= [exp(2^/D)-l] (393)
2

Beside the pipe w e have:


Retaining structures 309

s =
(h-h)Y
J (hs + rsdoD) (394)
M

Equating these two expressions w e find:

2Kuhs 2Kuhs
exp(^) - = 1 + 28 (395)

If rsd has a n e g a t i v e value, the frictional forces along the vertical section lines h a v e the
opposite direction to that s h o w n in Fig. 2 1 3 . This implies that will h a v e an u p w a r d
instead of a d o w n w a r d direction, /. e. will h a v e to b e inserted into E q s . (391 ) and (395)
with a negative sign. W e thus find:

2
YD 25 , 28
= -exp(p)] +/D(A-AJ)exp(^) (396)
2 D D

e x p ( ^ )+ = 1 - 23(1 (397)

Spangler ( 1948) proposed an alternative method for determination of hs. T h e interaction


effects b e t w e e n the fill a b o v e and beside the conduit is considered by a s s u m i n g that the
load difference a b o v e and beside, across breadth jD, is a s s u m e d to b e acting in the u p w a r d
direction on the fill beside. T h e settlement of the fill outside the conduit is thus obtained
by the correlation:

rhv . D dz D t-aoD
s= f '[ q - (' - q ) ] + [ / f t - ( < 4 - Yh) (398)
J z z
o 2jD M 2jD M

w h e r e \ = the stress at depth according to Fig. 2 1 8 ,


or z = at
'hs - ^ hs ( the crown level),
qz = YQi - hs +z).
Equating this expression and Eq. (393), w e find:

(2/)-1 1 h hs rsda1 ^(2^/) h hs


( H H ) + . / ( - ) =
2 2 D D 1+2/ 1+2/ D D
h h, h, h,1
= -(- +rsda)-
d s
(- - ) (399)
DD 2D D Ku

T h e hs values according to M a r s t o n (Eq. 397) and Spangler (Eq. 399) for rsd a = 0.5
and 0.8, respectively, and - 0.3 are given in Fig. 2 2 1 .
310 Retaining structures

1.5
Marston , rsda = 0.8
1 0.5
Marston , rsda=
Spangle r, rsda = 0.%

il
Spangler, rt/Ja = 0.5
hJD
1
0.5 I
1
1
1
1
4 6 10
HID

Fig. 221. The level of equal settlement according to Marston and Spangler for rS(fi - 0.5 and 0.8,
respectively, and - 0.3. In Spangler's expression j is assumed equal to 1.

T h e design values of the settlement ratio rsd according to the A m e r i c a n Concrete Pipe
Association (1988) for embankment pipes of the type shown in Fig. 220 are given in Table 37.

T h e M a r s t o n - S p a n g l e r theory only considers the vertical pressure exerted on the


crown of the pipe while the pressure distribution around the periphery of the pipe is not
treated. A s a matter of fact, the p r o b l e m of solving the pressure distribution is very
complicated and is very m u c h depending on the bedding conditions. A n empirical
solution w a s presented by Olander (1950) and further developed by Smith (1978). The
distribution of the design earth pressure a ' n against the top part of the pipe and a \ b against
the b o t t o m part according to Smith is given b y the relation (Fig. 2 2 2 ) :


n = <7cr [ cos
(400)
2(l-ab/2n)

, -
= 4 i n Vc o s ( ) (401)
a b in

TABLE 37.
1
Design values of the settlement ratio rsd according to ACPA (1988)

Foundation condition Usual range Design value

Rock or unyielding soil 1.0 1.0


Ordinary soil 0.5 - 0.8 0.7
Yielding soil 0.0 - 0.5 0.3

1
For ordinary soil the rsd value depends on the degree of compaction of the fill adjacent to the sides of
the pipe. With construction methods resulting in proper compaction of bedding and side-fill materials,
it is recommended to use the design value rsd = 0.5.
Retaining structures 311

Fig. 222. Theoretical stress distribution around rigid pipe according to Smith (1978).

w h e r e = angle from vertical axis,


ccb = b e d d i n g angle,
qCT = effective design pressure at the crown,
q i m = effective design pressure at the invert.
Generally speaking, the Marston-Spangler theory gives conservative results. Several
numerical m e t h o d s h a v e been developed, a m o n g which the m o s t a d v a n c e d ones are finite
element p r o g r a m m e s assuming the soil to b e h a v e as a non-linear elastic m e d i u m . T h e
reliability of the results obtained b y the various existing m e t h o d s has b e e n investigated
for e m b a n k m e n t pipes u n d e r various bedding conditions (Liedberg, 1991). T h e best
agreement with the test results w a s obtained by the S P I D A finite e l e m e n t p r o g r a m m e .
According to the results obtained by Liedberg, the most favourable earth pressure
distribution and internal stress condition in an e m b a n k m e n t pipe will b e obtained if the
pipe is laid on a soft cushion (Fig. 223).

(ii) Flexible pipes. In the case of flexible pipes, for w h i c h the bending rigidity of the
cylindrical walls can b e neglected, the pipe will deform under the earth pressure until the
line of thrust b e c o m e s m o r e or less coincident with the centre line of the pipe walls. T h u s ,
the ring force and the consequential risk of buckling will generally determine the pipe
design. However, traffic load or other sources of load variation m a y lead to pipe
deformations w h i c h do not recover.
T h e buckling stress ob in the pipe is determined by the relation ( T i m o s c h e n k o & Gere,
1961):
312 Retaining structures

Fig. 223. Observed stress distribution around two reinforced concrete pipes, 0.6 m in diameter, one
placed directly on the trench bottom (open circles) and the other on a soft inclusion, 0.10 m in thickness
(filled circles). (Liedberg, 1991)

b = (402)
t + 2yklsD/Gf

w h e r e kls = m o d u l u s of s u b g r a d e reaction,
t = t h i c k n e s s of the p i p e walls,
Of- the failure stress of the p i p e m a t e r i a l in c o m p r e s s i o n .
Embankments 313

E M B A N K M E N T S AND FILL ON SOFT SOIL

1. INTRODUCTION

T h e p r o b l e m s encountered in the construction of e m b a n k m e n t s (or, for that matter, in


connection with the placement of fill, whatever the purpose) on ground consisting of soft
clay or organic soils, such as dy, gyttja or peat, concern safeguarding stability and keeping
settlement within allowable limits.
Stability p r o b l e m s can b e solved in various w a y s , for instance by stabilising loading
berms, e m b a n k m e n t piles, replacement of soft soil by granular fill or soil improvement.
As regards settlement, the analysis is mainly related to consolidation p h e n o m e n a . T h e
interrelation b e t w e e n stability and settlement has to be taken into account. T h u s , shear
strains at high shear stress level may h a v e a strong influence on settlement.
For high-speed roads, it is very important that settlement should not jeopardise the
roadholding ability at bends or cause dangerous j u m p s . As regards the construction of
buildings, the ground at the building site is often brought up to a higher level. Although
buildings m a y b e founded safely on deep-lying strata with high bearing capacity,
consequential ground settlement m a y entail great p r o b l e m s on the connections between
the buildings and the surroundings (Fig. 224). In order to avoid breakage of connecting

Fig. 224. Subsidence due to placement of fill around a building founded on piles.
314 Embankments

water pipes and sewers it m a y b e necessary to apply preloading techniques or other


methods of soil improvement.

2 LOADING BERMS

T h e stability of e m b a n k m e n t s on soft clay and organic soils, such as dy and gyttja, is


generally analysed on the assumption of undrained soil conditions. T h e shear strength
variations with depth and the possible existence of w e a k layers are of imperative
importance for the analysis of the stability conditions. Possible effects of strength
anisotropy and differences in stress-strain characteristics b e t w e e n different soil types, for
e x a m p l e between mineral clays and organic material, have to be taken into account. Time
may have a negative influence on the shear resistance (cf. pp. 1 0 0 - 1 0 1 ) , although
consolidation p h e n o m e n a generally lead to long-term stremgth increase.
T h e analysis of loading berms as a means of improving stability w a s e a r n e d out in
detail by Jacobson and Odenstad (1940) and by Odenstad (1956, 1960).

2.1 Constant undrained shear strength

As previously shown (p. 120), the bearing capacity qyof cohesive soil with an undrained
shear strength equal to cu can be determined by the relation:

9/=< (403)
where = 2 + .
In other words, the height h of an e m b a n k m e n t , leading to failure, is equal to:
h = (2 + n)cu I y
where y- unit weight of the fill material.
T h e value = 2 + is deduced on the assumption of a Rankine-Prandtl failure con-
dition. Investigations regarding the shape of the failure surface envelope in real cases
indicate that failure in connection with e m b a n k m e n t s generally takes place along a
circular-cylindrical failure surface. For e m b a n k m e n t s on h o m o g e n e o u s soft soil, the
assumption of circular-cylindrical failure surfaces is therefore generally accepted as a
basis of analysis.
A s s u m i n g that the ground is subjected to an arbitrary load (Fig. 225) leading to failure
along a circular-cylindrical failure surface, and ignoring any horizontal, external load, the
condition of m o m e n t equilibrium around yields:

2
Q(a -e) = cu'2cxa lxfia (404)

or:

cu - Q(a -
2
e)sm oc/2oca
2 (405)
Embankments
315
316 Embankments

Fig. 226. Centre of critical failure surface with regard to loading condition.

i.e. ^ = 5.52

T h e situation of the critical failure surface is determined by the condition that the
driving m o m e n t of external load around its centre shall be m a x i m u m . Considering only
vertical, external load w e h a v e (Fig. 226):

= * (-)<*

+
(412)
T h e condition of m a x i m u m m o m e n t yields:

| = ^ f l - f V = 0 (413)

Consequently, the m a x i m u m driving m o m e n t is obtained for:

= + /? + (
4 1 4
>

Geometrically, the m a x i m u m driving m o m e n t is thus obtained w h e n the dashed areas


A and (Fig. 226) are equally large.
For the determination of the width of the loading berm, Jacobson and Odenstad (1940)
a s s u m e that the load of the e m b a n k m e n t fill has only vertical c o m p o n e n t s . Horizontal
earth pressure in the fill is thus neglected.
A s s u m i n g that qx represents the load exerted by the e m b a n k m e n t and q2 the load exerted
by the loading b e r m (Fig. 227), w e h a v e the stability condition:

-q2<5.52cul
qi (415)

q2<5.52cu2 (416)

w h e r e cul represents the undrained shear strength nearest b e l o w the b o a r d e r of the


Embankments 317

e m b a n k m e n t and cu2 the undrained shear strength nearest below the outer boarder of the
embankment.
T h e centre of the failure surface is given b y the area relation = (Fig. 226), i.e.:

(qx-q2)c = q2(b2-c) (417)

whence:

c = b2q2/qx (418)

(i) The failure surface touches firm bottom and intersects the ground surface inside or at
the opposite rim of the embankment. T h e load resultant (Fig.227) b e c o m e s equal to Q =
q2b2 + qxx and the line of action of the resultant is obtained by the relation:

2
(q2b2 + qxx)e = q2b2(b2 /2 + x) + qxx /2 (419)

Introducing the value of c, w e find:

X 1 1 Q\ <7i
^= M - + - +- - ( - ) ] / + < )
42

b2 2 2q2 b2 q2b2
Furthermore w e have:

a = c+x = b2 (1+ ) (421)


<7l <72^2

2
2cua a
- V z" = (92^2 +qxx){a-e) (422)
sin a

Fig. 227. Failure surface intersecting ground surface inside embankment. Depth restricted.
318 Embankments

w h e r e = 2arctan(D/a)
and e according to Eqs. ( 4 2 0 - 4 2 1 )
Introducing , and e into Eq. (422), x/b2 can be solved by the relation:


n > 2 2 V v 2 f , <7l ,
2 c w( X ) arctan( ) =
q\ b2q2X

<?i b2 2 2q2b$ b2q2X

x
Q\
where X = 1 +
qib2

Eq. (423) is valid only \fx<bx. For a given width of the loading b e r m b2 and load ratio
q\lq2 the value of that yields the m a x i m u m shear strength has to be found. T h e width
has to b e chosen in order that the safety requirements are satisfied.
In order to facilitate the design of the loading berm, J a c o b s o n and Odenstad have
presented the diagram shown in Fig. 228.

Fig. 228. Diagram for determination of width of loading berm when the failure surface touches firm
bottom and intersects ground surface inside or at the opposite side of the embankment.
Embankments 319

Example 58: Determine the loading berm required to achieve a factor of safety of 1.5 for a road em-
bankment, 12 m in width and 3 m in height, placed on a 10 m thick clay deposit with an average undrained
shear strength cu = 10 kPa. The clay is underlain by bedrock. The fill material in the embankment and
3
in the loading berm has a unit weight 7 = 18 kN/m .

2
Solution: The load qx - 3-18 = 54 kN/m . The hight h of the loading berm is obtained by the relation:
5 4 - 18/ = 5.5240/1.5
2
whence h = 1 m and q2 = 18 kN/m .
The design value of the undrained shear strength is cud = 10/1.5 = 6.7 kPa.
We have qx/q2 - 3 and (cu/qx)/Fs = 0.123. From Fig. 228 we find b2/D 2.5 which yields b2 = 25 m.
Moreover, we find xlD 1 . 7 . Since > bx the critical failure surface will intersect the ground surface
at = 12 m. Inserting the value = bx = 12 m and the given values of qx and q2 into Eq. (423) the solution
yields b2 = 22.2 m which is very nearly the same value as that obtained from Fig. 228.

(ii) The failure surface does not reach firm bottom. In this case the resultant (Fig. 229)
b e c o m e s equal to:

4 2 4
Q = q\b\ + 9 2 * 2 ( )

Furthermore:

qxb\l2
L LJ
+ q2(bx+b2l2)
e = - - (425)
qxbx+q2b2

A c c o r d i n g to Eq. (410) w e h a v e Q - \ \.04cue, w h i c h yields:

2
(qxbx +q2b2)
2
UM[qlb L/2 + q2b2(b1+b2/2)]

Example 59: Determine the loading berm required to achieve a factor of safety of 1.5 for a road
embankment, 12 m in width and 3 m in height, placed on a deep clay deposit with an average undrained
shear strength cu - 10 kPa. The fill material in the embankment and in the loading berm has a unit weight
3
7 = 18 kN/m .

Fig. 229. Failure surface does not reach firm bottom.


320 Embankments
2
Solution: As in example 58, the embankment load qx = 54 k N / m and the hight of the loading berm h
= 1m
The width of the berm is determined according to Eq. (426):
2
JLO (54 1 2 + 1 8 - f r 2 )
2
1.5 " 11.04[54 1 2 / 2 + 5 4 b2(l2 + b2/2)]
whence b2 = 33.7 m ~ 34 m.

2.2 Shear strength increasing linearly with depth

Assuming that the undrained shear strength increases with depth according to the relation
cu = cu0 + kz and that the horizontal projection 2a (Fig. 229) is given, the stabilising
m o m e n t is obtained by the relation:

CO0 + 2a 2 r)0
+2a + 2a 2
Ms = \ (cll +kz)R dco = \ [cu + kR (sin -sin 0)]R dco (427)
n m
C)

whence:
2a cu0[a + (ka/cu0)(l - acota)]
2
(428)
sin

T h e m i n i m u m value of Ms is determined by the condition dMJda - 0, w h e n c e :

2
ka a sin 2a - s i n a
(429)
CHQ a + sin2a(acota- 1.5)

T h e opening angle of the failure surface is governed by the ratio of D/a according to
Eq.(409), i.e. cx = 2arctan(D/a).
2
T h e values of Ms/(a cu0) and D/a for various values of ka/cu0 is presented in Fig. 2 3 1 .
In this case the loading b e r m can be inclined as shown in Fig. 232 with a m a x i m u m load
step between embankment and berm at failure of 5.52c w 0. T h e m a x i m u m rotational m o m e n t
is obtained when the centre of the failure surface is placed so that the area A is equal to

c K
u0 + Z

Fig. 230. Failure surface when the undrained shear strength increases linearly with depth.
Embankments 321

0.5

0.4

0.3

Dia

0.2

0.1

0
0 20 40 60 80 100 120 140
k a / c
u0

Fig. 231. Stabilising moment and depth of failure surface as functions of ka/cu0

area B. T h e m i n i m u m width of the b e r m is sought w h i c h satisfies the equilibrium


condition.
For the design of inclined loading b e r m s , Odenstad has presented the d i a g r a m shown
in Fig. 232.

Example 60: Determine the loading berm required to achieve a factor of safety of 1.5 for a road
embankment, 12 m in width and 3 m in height, placed on a 10 m thick clay deposit with an average
undrained shear strength cu = 10 + kz kPa, where k = 2 kPa/m and is the depth, in m, below ground
3
surface. The fill material in the embankment and in the loading berm has a unit weight 7 = 18 kN/m .

Solution: As in examples 57 and 58 the height of the loading berm next to the embankment should be
h = 1 m. The value oqxFJcu0 = 54 1.5/10 = 8.1. From Fig. 220 we find k/a= 0.355 kPa/m which gives
= 5.6. The required width of the loading berm is b2 = 18-1.5/5.6 5 m.

2.3 Influence of weak layers

T h e shapeof the failure surface m a y b e governed by the existence of layers in the soil with
lower shear strength than than that of the soil as a w h o l e (cf Slope stability, p. 345). T h e
m e t h o d of analysis in the case of a w e a k horizontal layer will b e demonstrated for a case
w h e r e a w e a k layer exists at a depth equal to D.
Now, let us a s s u m e that the undrained shear strength of the soil a b o v e the w e a k layer
is cui and in the w e a k layer cu2. Let us further a s s u m e that the width of the loading b e r m
is b2 and that the load of the e m b a n k m e n t is qx and of the loading b e r m q2 (Fig. 233). Then
stability requires that the resultant to the active earth pressure on the vertical section at
322 Embankments

Fig. 232. Diagram for determination of width of loading berm in the case of shearing strength increasing
linearly with depth. The width of the loading berm b = {q- 5.52c )/a.
u0

the r i m of the e m b a n k m e n t and the passive pressures on the vertical section just outside
the b e r m can b e counteracted b y the shear resistance exerted in the w e a k layer along the
loading berm.
T h e stability condition can thus b e written:

qxD - DNcQcul/2 = DN^cJl + b2cu2 (430)

from which:

b2 = D(q1-Nd)cul)/cu2 (431)

w h e r e Nc0 = 2 + n = 5.14

A s in the previous case qx-q2< Nc0cul.


In the analysis it has been p r e s u m e d that the width of the e m b a n k m e n t bx ~ Dyfl. If
Embankments 323

I I IM
<?2 yKJIl

sDNc0cuJnj j j T i D - D N c 0c M /l 2
2 2
yp l2 4 y / > / 2
~-Wcak layer

Fig. 233. Assumed failure condition in the case of a weak horizontal layer.

bx > Dyfl the horizontal earth pressure in the e m b a n k m e n t itself will h a v e to b e added.
A s s u m i n g that the height of the e m b a n k m e n t is h and its unit w e i g h t / w e find:

b2 = [yh(D + Kah/2)-Nc0cul]/cu2 (432)

w h e r e Ka can b e taken as the active earth pressure coefficient for /' = 1 (see Table 36,
p. 283).

Example 6 1 : Determine the loading berm required to achieve a factor of safety of 1.5 for a road
embankment, 12 m in width and 3 m in height, placed on a clay deposit with an average undrained shear
strength cu = 10 kPa. At a depth of 3 m there is a continuous, weak horizontal clay layer with an undrained
shear strength of 7 kPa. The fill material in the embankment and in the loading berm has a unit weight
3
7 = 18 k N / m . The angle of friction in the fill material can be assumed equal to 35.

Solution: According to Table 36, the active earth pressure coefficient in the embankment material is Ka
= 0.26. Utilising the relation given by Eq. (432), we find:
b2 = [18-3(3 + 0.26-1.5) - 5.14 10/1.5] 1.5/7 = 33.6 m - 34 m.

3. SETTLEMENT

3.1 Vertical settlement

Provided that the stability of an e m b a n k m e n t is satisfactory, it can b e stated that the


settlement distribution generally agrees closely with the vertical stress distribution
induced b y the load as determined by the theory of elasticity. A s fill material is completely
flexible, the analysis of the stress distribution is simplified and can b e carried out
according to Eqs. ( 9 0 - 9 3 ) for a strip load and according to Fig. 77 for a circular load
distribution. For square or rectangular load distributions S t e i n b r e n n e r ' s method, based
on Eq. (94), can b e applied.
For the analysis of the m e a n settlement caused by an e m b a n k m e n t , the stress
distribution according to the empirical 2:1 m e t h o d (p. 109) gives satisfactory accuracy.
T h e analysis of settlement due to e m b a n k m e n t s and fill material is m a i n l y restricted
to the determination of the consolidation process and the l o n g - t e r m settlement under the
imposed load. A s previously mentioned, settlement consists of t w o c o m p o n e n t s : v o l u m e
324 Embankments

decrease due to a decrease in void ratio and v o l u m e change due to shear. T h e instantaneous
settlement taking place when the e m b a n k m e n t is placed on water saturated, low-
p e r m e a b l e soil is completely due to shear while, in the case of non-saturated soil, a
v o l u m e decrease takes place by compression of gas bubbles.
Instantaneous settlement taking place in connection with filling operations is generally
insignificant. However, if i m m e d i a t e settlement should b e of s o m e concern, it can be
calculated on the basis of the theory of elasticity on the assumption that the elastic
m o d u l u s Eis in the range of 150c M for water saturated, normally consolidated clay and
5 0 0 c w for overconsolidated clay.
Doubtless, the dominating part of settlement, taking place under the load of embankments
and fill on soft clay and organic soils, is due to primary and secondary consolidation. T h e
analysis of primary consolidation settlement is generally carried out according to
Terzaghi's consolidation theory. Refined computer p r o g r a m s exist which take into
account c o m b i n e d primary and secondary consolidation, self-induced excess pore water
pressure and c h a n g e s in the consolidation characteristics (primarily c h a n g e s in
permeability) during consolidation. However, from a practical viewpoint the information
available regarding soil parameters does not ordinarily justify advanced calculations by
the aid of computer p r o g r a m s .

(i) One-dimensional primary consolidation. Conventional analysis of the consolidation


process generally includes the determination of the final primary settlement, based on the
results of o e d o m e t e r tests, and the average degree of consolidation utilising diagrams of
the kind shown in Fig. 234.
T h e average degree of consolidation, however, is not a satisfactory m e a s u r e of the
consolidation process. It is quite important to get an idea of h o w the consolidation process
is advancing throughout the soil layer. For the determination of the degree of consolidation
Uv, reached at a certain depth z, a numerical m e t h o d suggested by Helenelund (1951)
gives a quick and quite accurate answer. A s s u m i n g that the soil profile is divided into
layers of equal thickness Az, Terzaghi's consolidation equation (Eq. 29) can be written:

'+1~ Ut_ ]
(433)
At Az
Az Az Az

where ut_ l9 ut and ui+ represent the excess pore water pressure at time t in the middle
of layers i - 1, / and / + 1, respectively,
Au = change in pore water pressure in the middle of layer / during a time increment At.
2
Introducing ATV = cvAt/Az , Eq. (433) can be rewritten:

Aui = ATv(ui_l + ui+l-2u$ (434)


Embankments 325

Sand

0.01 0.05 0.1 0.5 1 5 10


Time factor Tv

2
Fig. 234. Correlation between time factor Tv = cvtlh and the average degree of consolidation according
to Terzaghi.

2
Now, if CyAt/Az is constant, the c h a n g e in excess pore water pressure with time can be
calculated stepwise. A simple graphical solution is made possible by assuming ATV = 1/4,
whence:
1 II; + U: U: + Uj+

u, + , = "(^ + ~ -J )
! 1
(435)
i.e. the excess p o r e water pressure in the middle of layer i at time t + A n s the arithmetic
m e a n of the excess pore water pressures existing at time t in the interfaces between layer
i and the t w o surrounding layers (Fig. 235).
T h e assumption of ATV = 1/4 entails that the clay layer is divided into part layers of
thickness:
Az = lyfc^At (436)

T h e time increment At can now be chosen in order that at least 3-4 isochronal lines of
pore pressure distribution are obtained u p to the consolidation time in question. T h e
practical application of the m e t h o d is illustrated in Fig. 236.
326 Embankments

' / 2 | _
"_

4f
/2
r

/2

Fig. 235. Graphical determination of the change in excess pore water pressure from time t to time
t + At.

If the soil is not h o m o g e n e o u s the thickness of the respective layers will vary in
accordance with the variation of the coefficient of consolidation. Moreover, the condition
of continuity of flow between the different layers has to b e satisfied. This can also be done
numerically in a fairly simple way. T h u s , assuming validity of D a r c y ' s law the condition
of continuity of flow yields:
du du
ki()i = ki + l () i +l (437)
dz. dz

T h e graphical construction of the isochronal excess pore pressure gradients on both


sides of the interface between two neighbouring layers requires auxiliary lines as
indicated in Fig. 2 3 7 . For e x a m p l e , in this figure, at time 3, w e h a v e (du/dz)\ - tana
and (duz)i - tan/3. With the aid of the auxiliary lines and the construction shown w e have
kx t a n a = k 2 tan/3. H e n c e it follows that the boundary flow condition between layers 1 and
2 is satisfied. T h e graphical m e t h o d indicated for layers 1 and 2 is applied to all the layers
of the soil profile.

(ii) Two-dimensional consolidation. In the case of preloading in order to avoid long-term


settlement of road e m b a n k m e n t s , the width of the load is often limited in comparison with
the thickness of the soft soil deposit. In such a case the consolidation process is shortened
by two-dimensional consolidation taking place. This can also b e solved numerically by
H e l e n e l u n d ' s m e t h o d in combination with Carillo's equation (Carillo, 1942):

U=Uv+Uh- UvUh (438)

w h e r e U = total degree of consolidation,


Uv = degree of consolidation due to pore water flow in the vertical direction,
Uh = degree of consolidation due to pore water flow in the horizontal direction.
Embankments 321

Sand . - . - Sand

V * Sand ' / ' . - Impermeable base

Fig. 236. Graphical determination of the excess pore pressure distribution at times At, 2At and 3At after
the initial stage (t = 0). Homogeneous clay deposit. (In reality, the isochronal pore pressure lines form
a curve through the values obtaining in the middle of the respective layer). In the case of impermeable
base, the graphical construction is determined by the fact that the hydraulic gradient AuIAz must be zero
in the interface between the clay deposit and the impermeable base.

f/ \ . . . .Sand . /, ;: v

:
'."* / I"- : Sand

Fig. 237. Graphical construction of the pore pressure distribution at times At, 2At and 3 At after the initial
stage (t = 0). Inhomogeneous deposit. (In reality, the isochronal pore pressure lines are curves passing
through the pore pressure values obtaining in the middle of the respective layer and fulfilling the
condition of flow continuity in the interface between the layers).
328 Embankments

Alternatively, the equation can be expressed in excess pore pressure u:

UyUfr
U (439)
u0

w h e r e u0 = initial excess pore water pressure,


uv = remaining excess pore pressure due to vertical pore water flow,
uh = remaining excess pore pressure due to horizontal pore water flow.

Example 62: Determine the degree of consolidation attained after 4 years at 2.5 m depth below the
centre of a 4 m wide embankment, placed on a 6 m thick, water saturated, normally consolidated clay
underlain by sand. Determine also the excess pore water pressure after 4 years at 2.5 m depth, 5 m from
the centre of the embankment. The clay deposit is fully drained at top and bottom. The coefficient of
2
consolidation is cv = 0.25 m /year. The permeability in the horizontal direction is 4 times that in the
vertical direction. Skempton's pore pressure coefficients are = 1.3 and = \ .

Solution: Utilise the Carillo equation, i.e. divide the consolidation process into two phases: one with
pore water escape only in the vertical direction and one with pore water escape only in the horizontal
direction. The stress increase induced by loading is assumed to obey the theory of elasticity, Eqs. ( 9 0 -
91). The initial pore pressure distribution with depth under the centre of the strip loading is found by
Eq. ( 18). Its distribution at 2.5 m depth is symmetrical around the plane of symmetry of the strip loading
and is also given by Eqs. (90-91) and Eq. (18). Thus, for example, below the centre of the strip loading
we find Au/q = 1.11 at 1 m depth, Au/q = 0.93 at 2.5 m and Aujq - 0.51 at 6 m depth. At 2.5 m depth,
4 m away from the plane of symmetry, we find Au^Jq = 0.41, and 6 m away from the plane of symmetry,
Auo/q = 0A0.
1 /2 l /2
Choosing the time interval At = 1 year we find Az = 2(1 - 0 . 2 5 ) = 1 m and Ax = 2 ( M ) = 2 m. The
construction of the pore pressure decrease with time can now be carried out as shown below.
We find Uv = 0.11/0.93 = 0 . 1 2 and Uh = 0.25/0.93 = 0.27.
According to Carillo's equation, the total degree of consolidation at 2.5 m depth is:
U = 0.12 + 0.27 - 0.12-0.27 = 0.35 (35%)
At a distance of 5 m from the centre line we find uv = 024q and uh = 0.37q, whence ulq = 0.24-0.37/0.29
= 0.31.
Embankments 329

Example 63: In the test field at Sk-Edeby, established by the Swedish Geotechnical Institute in 1957,
3
a gravel fill, 1.5 m in thickness and with a unit weight of 18 k N / m , was placed on an undrained test area
with a diameter of 35 m. The soil at the site consists of water saturated clay underlain of sand at 12 m
depth. The clay is normally consolidated and has a dry crust of 1 m thickness. The coefficient of
2
consolidation can be estimated at 0.4 m /year. Determine the excess pore pressure at 5 m depth in the
soil under the centre and the periphery of the test area, and also at the distances 10 m and 20 m from the
centre. Skempton's pore pressure coefficients were estimated a t = 1 and = 0.75. Determine also the
excess pore pressure distribution below the centre of the area after 30 years.

Solution: Taking into account the slope at the boundary of the fill, the nominal diameter is 35 - 2.3 =
32.7 m. The pIR values in question are thus 0, 0.61, 1.0 and 1.22. The pore pressure distribution is
determined on the basis of Eq.(18) and Fig. 77. The stresses induced by the load for the different values
of are:
p = 0 10 m 16.35m 20 m
Aojq 0.98 0.90 0.45 0.12
0.58 0.48 0.37 0.29
0 0.14 0.29 0.16
Aoxlq 0.98 0.94 0.70 0.39
Aa3/q 0.58 0.44 0.12 0.02

where 13 is obtained from the relation:

2 2
, 3 = ^ ( + ) ^ - ) ] +

Introducing these values into Eq. ( 18) and q = 27 kPa, the excess pore pressure becomes equal to AuQ
= 24 kPa for = 0, Au0 = 22 kPa for = 10 m, Au0 = 15 kPa for = 16.35 m and Au0 = 8 kPa for =
20 m. These values are almost identical with those observed: Au0 = 23 kPa for = 0, Au0 = 22 kPa (max.
25 kPa; min. 20 kPa) for = 10 m and Au0 = 6 kPa for = 20 m.

27 kPa
330 Embankments

The excess pore pressure distribution with depth below the centre of the area becomes equal to AuQ
= 27 kPa for = 0, Au0 = 25 kPa for = 2.5 m, Au0 = 24 kPa for = 5 m, Aw0 = 22 kPa for = 1.5 m and
Au0 - 18 kPa for = 12 m. In this case, the influence of horizontal pore water flow can be disregarded.
The pore pressure dissipation can now be determined according to Helenelund's method. Assuming the
top dry crust to be drained because of the existence of frequent fissures and dividing the remaining 11
2
m of clay into 4 part layers, we have = 2.75 m. This yields = (2.75/2) /0.4 = 4.73 years. Thus 30
years correspond to 30/4.73 = 6.3.
The construction of the pore pressure dissipation is shown at bottom of p. 328.
The maximum excess pore water pressure after 30 years, about 20 kPa, is found at 6-7 m depth.

(iii) Secondary consolidation. According to Eq. (35) the relative compression due to
secondary consolidation can be expressed by the relation = asA(\ogt), which can also
be formulated as:
= as\og(t/tp) (440)

where tp = time required to reach 100% primary consolidation.


The secondary settlement may be important, especially in the case of highly organic
soils and extremely soft clays. For example, in the test field at Sk-Edeby outside
Stockholm, where the undrained shear strength of the clay cu ~ 8-10 kPa and the void
ratio e0 ~ 1.8-2.2 and a compression index CJ(\ + e0) ~ 0.5, a value of as, based on 30
years of observation, of between 0.02 and 0.05 has been observed.
Using instead the correlation given by Eq. (34) we have:

Ae = Ca\og(tltp) = as(l + e0)\og(t/tp) (441)

According to Mesri and Godlewski (see p. 79),most typical values of CJCC for inorganic
soft clay and for organic soft clays are 0.04 0.01 and 0.05 0 . 0 1 , respectively. The
correlation between as and Ca is given by the expression as = CJ(\ + e0). Thus, in the Sk-
Edeby case, CJCC would reach as high as 0.1.
The as value increases with increasing natural water content and reaches its maximum
when, in the consolidation process, the m a x i m u m past pressure is being passed by the
effective overburden pressure. Furthermore, its value decreases almost linearly with
increasing relative compression. Secondary consolidation can be reduced, and even
eliminated (cf. Fig. 325), by preloading the soil to a higher effective stress level than that
existing under working load condition.

3.2 Horizontal displacements

Long-term settlement due to shear is very much dependent on the shear stress level in
relation to the shear strength. Often, the factor of safety against failure is between 1.3 and
1.5. With a factor of safety of 1.3 the horizontal displacement in the soil along the rim of
the embankment can be quite important (Fig. 238).
Embankments 331

4. EMBANKMENT PILES

In places w h e r e the stability of the e m b a n k m e n t is insufficient it is often economically


advantageous to u s e e m b a n k m e n t piles in stead of loading b e r m s . T h e e m b a n k m e n t piles
also serve the p u r p o s e of reducing (floating piles) or eliminating settlement (end bearing
piles).
T h e design of e m b a n k m e n t piles can b e m a d e in accordance with the principles shown
in Fig. 239. T h e design diagram in this figure is based on the assumption that the whole
of the e m b a n k m e n t load, including traffic load, is carried by the piles. T h e relation
between pile spacing S and width of the pile caps is chosen on the basis of empirical
experience in order that arching in the soil prevents the piles from penetrating through
the e m b a n k m e n t .
It should b e noticed that batter piles h a v e to b e installed in order to resist the earth
pressure exerted b y the e m b a n k m e n t fill. Alternatively, the e m b a n k m e n t can be rein-
forced by m e a n s of horizontal layers of geotextiles. In very soft soil condition, it m a y be
necessary to p l a c e a mattress, or geotextiles, on top of the pile caps which serve the double
purpose of improving arching and preventing piles from being displaced sidewards
during the filling operation.
T h e pile caps are of reinforced concrete and prefabricated. Their width varies usually
between 0.8 and 1.5 m.
Floating piles without pile caps are also used in order to i m p r o v e the stability condition
of the e m b a n k m e n t (Fig. 240). T h e stabilising effect of the piles is obtained by the

Horizontal displacement, mm

100 50 0 A 0 50 100 150

15

Fig. 238. Horizontal displacement underneath a road embankment on soft clay deposits at Ska- Edeby.
Shear strength cu at the time of construction around 7 kPa. Height of embankment 1.5 m (load q - 27
2
kN/m ). After Larsson, 1986.
332 Embankments

0 2 4 6 8 10

5 5

Fig. 239. Design diagram for embankment piles

resisting m o m e n t around the centre of the potential failure surface induced by shaft
friction. If the piles m o v e with the sliding b o d y (as, for e x a m p l e , pile 1 in Fig. 240), the
frictional forces along the part of the piles that are outside of the sliding b o d y should be
taken into account. If, on the other hand, the piles do not m o v e with the sliding body (as,

Fig. 240. Analysis of stability of embankment placed on floating embankment piles. The part of the piles
contributing to stability are indicated by arrows. The lever arms for the piles with respect to the centre
of the potential failure surface exemplified for piles 1 and 2.
Embankments 333

for e x a m p l e , pile 2 in Fig. 240), the frictional forces along the part of the piles that are
inside the sliding b o d y should b e considered.

Example 66: A 4m high embankment shall be built on a deep clay layer with an undrained shear strength
increasing with depth below the ground surface according to the relation cu = 10 + 1.5 kPa, where
3
in m. The density of the embankment material is 1.8 t/m . Determine the maximum pile spacing if the
pile caps are 0.8 m by 0.8 m and the required pile length. The safety against pile failure is taken as
minimum 1.3.
Determine also the number of batter piles using a pile inclination of 5:1
2 2
Solution: W e have hi = 4/0.8 = 5 whence S /a = 4.8. This yields 5 = 1.75 m.
The pile working load is = 4.8-0.644 18 - 220 kN
Using spliced timber piles with a total length of 25 m, the lower pile segment consisting of a 17 m
long pile with a tip diameter of 0.13 m and a head diameter of 0.30 m and the upper of an 8 m long pile
with a diameter varying from 0.30 m at the splice to 0.22 m at the top, the bearing capacity of the piles
according to Eq. (207) becomes:

Q = 0.3 5 [- (0.22 + 0.30) 14.5 + - (0.22 + 2 0.30) 1.5 5]


2 6

+ 0 . 7 1 7 [ - ( 0 . 3 0 + 0 . 1 3 ) 2 2 + - ( 0 . 3 0 + 2 0 . 1 3 ) 1.5 17] = 288 kN


2 6
This corresponds to a factor of safety of 288/220 = 1.31
The number of batter piles is determined by the earth pressure exerted by the embankment. Assuming
an 'at rest' earth pressure condition with K0 = 0.5, the earth pressure becomes:
2
P= 0.5-18-4 /2 = 7 2 k N / m
The horizontal component of the pile resistance for a pile inclined 5:1 is 288 / / 2 6 = 56 kN. With
a factor of safety of 1.3, the allowable horizontal resistance per pile is 43 kN, i.e. 43/1.75 = 25 kN/m.
Consequently, 3 rows of batter piles are required along each border of the embankment.

Example 67: Determine the factor of safety against failure for the hypothetical slide surface shown
below. The 4 m high embankment is stabilised with 25 m long concrete piles with square cross-section,
3
0.25 m in width. The unit weight of the embankment material is 19 kN/m . The undrained shear strength
of the clay is 15 kPa. The inclination of the batter piles is 5:1 and the pile spacing is 4 m across the
embankment and 2 m in the length direction.

h=4m
334 Embankments

Solution: The value (Fig. 220) of the slide surface is 29-sin(2.33/2) = 26.6 m. The driving moment of
the embankment is thus:
Md = 41926.6 2/2 =26887 kNm/m
The stabilising moment is the sum of the shear resistance along the failure surface and the resistance
exerted by the piles. The pile lengths contributing to the stability are l{ = 8.0 m, l2 - 7.8 m, / 3 = 8.5 m,
/ 4 = 11.2 m, / 5 = 11.5 m, l6 = 6.6 m and / 7 = 2.0 m. The lever arms are rx = 3.2 m, r2 = 7.3 m, r3= 11.4
m, r 4 = 13.0 m, r5 - 17.0 m, r 6 = 18.3 m and r 7 = 22.3 m. This yields:
Ms = 15[2.33-29 2 + (8.0-3.2 + 7.8-7.3 + 8.5-11.4 +11.2-13.0 + 11.5-17.0 +6.6-18.3 + 2.0-22.3)/2]
= 34537 kNm/m
The factor of safety Fs = 34537/26877 = 1.3

5. E N F O R C E D D I S P L A C E M E N T OR EXCAVATION O F SOFT SOIL

In cases w h e r e settlement requirements are less strict, the soft soil can b e displaced by
overloading the soil to p r o d u c e ground failure. In this way the load of the fill will cause
an outward m u d w a v e and fill will replace the soft soil thus displaced. If the e m b a n k m e n t
fill is not displacing the soft soil in a satisfactory way, the overload m a y either be
increased or masses in front of the fill b e excavated.
Alternatively, a combination of overloading and blasting can be used in order to obtain
a m o r e complete removal of the soil underneath the fill (Fig. 241). T h e explosives are
usually placed in plastic tubes, 7 5 - 1 0 0 m m in diameter, which are inserted in the soil to
the required depth in a square grid with 3 - 4 m spacing. T h e best result is obtained when
the explosives are evenly distributed along the lower third of the tubes and a delayed
blasting technique is used. A b o u t 100 g of explosives per m 3 is generally chosen.

Fig. 241. Embankment sinking into underground removed by blasting.


Embankments 335

T h e e m b a n k m e n t masses should preferably consist of blasted rock as this improves


soft soil removal. Blasted rock has also a high internal angle of friction which reduces the
risk of fill sliding sidewards.
If the depth of soft soil is small, excavation and refilling is usually technically and
economically preferable to enforced displacement. Regarding the analysis of the stability
conditions of the slopes of the excavation, see ' S l o p e Stability'.

6. OTHER MEANS OF STABILISATION

In m o d e r n road construction it has b e c o m e economically advantageous to apply soil


i m p r o v e m e n t m e t h o d s for the purpose of eliminating detrimental settlements and
improving the stability of road e m b a n k m e n t s . A m o n g the m o s t c o m m o n soil i m p r o v e m e n t
m e t h o d s w e find vertical drains and lime/cement c o l u m n installations (see 'Soil
Improvement').
In the last few years, the use of light fill material to avoid an increase in the in situ stress
condition in the subsoil has b e c o m e quite popular. For e x a m p l e , e x p a n d e d polystyrene
3
with a b u l k density of 1 5 - 3 0 k g / m , is n o w used extensively in road e m b a n k m e n t s . In
3
the case of e x p a n d e d polystyrene with a density of 15 k g / m , the permissible normal
stress can b e put eqaul to 12 kPa for p e r m a n e n t load and 30 kPa for p e r m a n e n t and traffic
3
load in combination. For e x p a n d e d polystyrene with a density of 30 k g / m the corre-
sponding values are 35 k P a and 75 kPa, respectively. Blocks of polystyrene should be
placed so that the joints in the different layers are displaced in relation to eachother.
Geotextiles of different kinds are used as reinforcement of the soil to i m p r o v e the
stability condition but has no effect on settlement.
336 Slope stability

S L O P E STABILITY

1. INTRODUCTION

Before building activities close to river valleys or other natural slopes are begun, the
overall stability condition at the building site has to b e checked. By e x p e r i e n c e w e know
that m a n y disastrous and unforeseen slides with consequential fatal casualties and heavy
e c o n o m i c losses h a v e occurred due to unsatisfactory stability conditions (Fig. 242).
In the following an outline of the analytical m e t h o d s utilised to solve the slope stability
problems will b e presented. For m o r e detailed studies, reference is m a d e to text books and
papers specially devoted to the problem.

2. CIRCULAR-CYLINDRICAL FAILURE SURFACE

2.1 Undrained analysis

In the case of cohesive soil, the traditional analytical approach is based on the assumption'
of circular-cylindrical failure surfaces acted u p o n by the undrained shear strength cu of

Fig. 242. Air view of the slide at Tuve north of Gothenburg that took place on Nov. 3 0 , 1 9 7 7 . In the slide
65 dwelling-houses were carried away and 9 people were killed. The slide, which was initially of limited
extent but developed into a retrogressive type of slide, took place after a long period of raining.
Slope stability 337

the soil (Fellenius, 1 9 1 8 , 1 9 2 6 , 1 9 3 6 ) . Strength anisotropy (see pp. 3 4 and 1 0 1 ) with regard
to the inclination of the failure surface and time effects are generally taken into account. T h e
critical failure surface with regard to the loading condition is found by trial and error.
T h e rotational driving m o m e n t of the dead weight of the soil comprised by the sliding
surface, external load included, is easily established. T h e counteracting stabilising
m o m e n t is obtained by s u m m a t i o n of the shear strength contributions exerted along the
a s s u m e d failure surface times their lever a r m / ? . Failure occurs w h e n the driving m o m e n t
exceeds, or equals, the stabilising m o m e n t thus obtained.
In the case of a simple, elementary slope w h e r e the total unit w e i g h t of the soil is
constant, the driving m o m e n t Md due to the weight of the soil enclosed by the circular-
cylindrical surface can b e obtained by the relation (Fig. 2 4 3 ) :

2 2 2
Md = ^ [ R - x - y + H(y-") + B(x- ")] (442)

If the slope is subjected to water pressure the influence of the water pressure against
the slope has to b e added.
In h o m o g e n e o u s soil conditions the critical failure surface can b e determined once and
for all according to J a n b u ' s direct m e t h o d (Janbu, 1 9 5 4 ) . In this case, the safety against
failure in undrained condition can be expressed as:

f
' = ^ ^ ^ <) 4 4 3

w h e r e the stability factor jV 0 is given in Fig. 2 4 4 and the correction coefficients pq with

Fig. 243. Notations used for the analysis of rotational moment due to weight of soil comprised by
circular-cylindrical slide surface along slope.
338 Slope stability

Example 67: The risk of failure of a slope with 5 m height and an inclination of 1:1.5 shall be
investigated. The water depth outside the slope is 3 m. Calculate the rotational moment around the centre
of a circular-cylindrical failure surface assuming x = 2m,y = 1 0 m and R = 12 m. The unit weight
3
of the clay = 18 kN/m .

7.5 m

Solution: We have H = 5 m and = 1.5-5 = 7.5 m. The water pressure exerted on the slope is "equal to
5.4-3-9.81/2 = 79.6 kN/m and its lever arm with reference to is 5.41 m.
The rotational moment becomes:
2 2 2
Mr = (18/2)-5[12 - 2 - 1 0 + 5(10 - 5/3) + 7.5(2 - 7.5/3)] - 79.6-5.41 = 3076 kNm/m

regard to external load q, with regard to water depth Hw and w i t h regard to depth
Hjof superficial fissures in the dry crust in Fig. 2 4 5 .
T h e stabilising m o m e n t in an arbitrary case with a m o r e c o m p l i c a t e d slope g e o m e t r y
and varying u n d r a i n e d shear strength can b e obtained b y slicing up the b o d y into slices

90 80 60 40 20 0
Slope angle

Fig. 244. Stability factor N0 under homogeneous soil and undrained failure conditions
Slope stability 339

/3=0 DIH =oo


n i l // 1.0

0.9 .60

90

> ) / / / ) > > r ;


0 0.5
HJH

D/H = o o

^7
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5

Fig. 245. Correction coefficients ^ and ^

w h o s e width is determined by the variation in shear strength. A s s u m i n g that the width of


a slice along the failure surface is A/, the driving m o m e n t can b e written:

Md= 0 + /?(^ + W 2) s i n a (444)

w h e r e P0a is the rotational m o m e n t due to an external horizontal force (e. g. water pressure
or earth pressure in fissures),
Wi denotes full weight of the soil in the lamella a b o v e the water level outside the
slope,
W2 denotes s u b m e r g e d weight of the soil in the lamella b e l o w the water level
outside the slope,
u s i n a i s the lever a r m with respect to the centre of the circular-cylindrical failure
surface.
T h e stabilising m o m e n t Ms is obtained by the relation:

Ms = / ? I c MA / (445)

w h e r e R is the radius of the failure surface and cu is the undrained shear strength along
the part Al of the failure surface belonging to the lamella in question..

Example 68: Express in general terms the stabilising moment for a circular-cylindrical failure surface
in a clay slope if the undrained shear strength of the clay increases linearly with depth according to the
relation: cu = cu0(l+kz) kPa, where is the depth, in m, below the top of the slope.

Solution: With the notations given in the figure we have:


340 Slope stability

o + (Q+a
MS=R J cuRda>= R J cu[l+kR(s'm-s'mo)]d
CQ CQ

whenceM s = Rcu0{Bs + a[l - k(y - H)]} kNm/m

2.2 Effective stress analysis


T h e first attempt to determine the slope stability for a soil with an internal angle of friction
was publicised in connection with the investigation of the cause for the slide (Fig. 246)
which t o o k p l a c e on M a r c h 5 , 1 9 1 6 , at the Stigberg quay (Pettersson, 1916). T h e analysis,
which was carried out by the Gothenburg Harbour engineer S. Hultin (Hultin, 1916), was
based on the assumption of a circular-cylindrical failure surface and is generally referred
to as the friction circle m e t h o d (Fig. 247). T h e m e t h o d is very laborious and requires trial
and error to find the friction angle which satisfies the equilibrium condition of the sliding
b o d y investigated.

Fig. 246. The slide at the Stigberg quay in the harbour of Gothenburg on March 5, 1916,' initiated the
development of the slope stability analysis based on the assumption of circular-cylindrical failure
surfaces. Notice that the quai front wall is fully intact and hardly affected by the slide while the ground
behind with the railway tracks has sunk below water. This indicates that the centre of the circular-
cylindrical slip surface is situated nearly straight above the quai front wall.
Slope stability 341

Fig. 247. The analysis of the slide at the Stigberg quai in 1916 was based on the assumption of clay
behaving as a friction material. The method of analysis, named 'the friction circle method', based on
the assumption of a circular-cylindrical failure surface, was applied for the first time in history by Sven
Hultin (1916).
342 Slope stability

Pi0

Fig. 248. The slice method applied in effective stress analysis of a circular-cylindrical failure surface.

In m o d e r n analysis, the slope stability conditions are generally studied by a method


similar to that used in the undrained analysis with varying u n d r a i n e d shear strength.
According to Bishop (1955), the body is cut into slices where, at at the b o t t o m of each
slice, the values of c ' a n d 0'are constant (Fig. 248). S u m m i n g up over all slices with their
respective widths, the driving m o m e n t Md b e c o m e s equal to:

M,r = Pfl + 1(W + W2)x = 0 + RliW^ + W 2) s i n a (446)


d X

w h e r e Wx denotes full weight of the soil a b o v e the water level outside the slope,
W2 submerged weight of the soil below the water level outside the slope.
T h e stabilising m o m e n t is now given b y the expression:

(447)

T h e & value is obtained from the equilibrium condition of the slice. A vertical projection
equation yields:
Wx + W2 +AT-l(uecosa+ c smoc/F^)
/
(448)
c o s a + sinatan0 //^0

w h e r e ue = ywz = the pore water pressure at the b o t t o m of the slice expressed as an excess
over the hydrostatic pressure corresponding to the water level outside
the slope. (If no part of the slice is submerged, ue = u).
Slope stability 343

Substituting / for b/cosoc and inserting ' into Eq. (447) w e have:

/
R X [ c + ( W 1 + W2 + AT-ueb) tan ']
Ms = (449)
Fc<t> c o s a ( l + tan tan 0 7F^)

w h e r e the s u m is taken over all slices with their respective width b and soil characteristics.
T h e factor of safety can thus b e expressed by the relation:

Rl[c'b + (Wl + W2+AT-ueb)ton<l> ']lmc


F
cd> (450)
0 + Rl{Wx+W2) sin a

w h e r e ma = cosoc (1 + t a n a tan '/F^).


Eq. (450) is referred to as Bishop's rigorous method.
In the case of a circular-cylindrical slide surface, the influence of AT on the factor of
safety can be neglected (see E x a m p l e 71) and the factor of safety consequently be ex-
pressed as:

Rl[c'b + (WX + W 2 - n g f r ) t a n 0 ']lma


F
c (451)
P0a-^Rl(Wl + W2)s'ma

Eq. (451) is referred to as Bishop's simplified method.

Example 69: Determine the factor of safety for the potential circular-cylindrical failure surface across
the river bank shown below). The soil consists of sand with an angle of friction '= 37 underlain by
sulphide silt with an angle of friction ' - 35. The density of the sand, which is situated above the

20 m
344 Slope stability

3 3
groundwater level, is 1.8 t/m . The density of the silt is 1.7 t/m on the average above groundwater and
3
1.75 t/m below groundwater. The pore pressure increase with depth is hydrostatic.

Solution: The body comprised by the potential failure surface is sliced up, as shown, into 10 lamellae.
For each slice, the values and the total and effective vertical pressures, acting on the potential failure
surface, are calculated. As a first attempt, put F^= 1.05. Then the result, ignoring the A7term, and
denoting Wx + W2 = W, is as follows:

Slice b m WkN/m ueb kN/m ' AMJR kN/m

1 3.5 57.5 173 0 37 108.7


2 3 51.5 390 25 221.2
3 3.5 45.0 588 123 35 265.4
4 3.5 39.7 627 158 261.6
5 5 33.0 825 210 341.2
6 26.3 665 125 302.1
7 " 18.8 585 60 301.4
8 12.8 543 30 304.7
9 6 5.0 380 0 " 240.4
10 8 -4.0 180 0 126.2
2472.9

Thus assuming Fc(p - 1.05, the stabilising moment around the centre of the circular-cylindrical surface
becomes 111280 kNm/m.
The driving moment around the centre of the circular-cylindrical surface becomes 105335 kNm/m.
The factor of safety becomes Fc(p = 1.06.

3. LOGARITHMIC FAILURE SURFACE

R a d i u s vector to the l o g a r i t h m i c s p i r a l / ? = / ? 1e x p ( ( t a n 00 has the virtue of forming at each


point of the c u r v e an angle 0 ' w i t h its normal. T h u s , a s s u m i n g the shape of the failure
surface to a g r e e with that of a l o g a r i t h m i c spiral all the u n k n o w n frictional forces along
the failure surface pass through the centre of the spiral. T h e calculation of the risk of

Fig. 249. Logarithmic failure surface in non-cohesive soil.


Slope stability 345

failure in a cohesionless soil is then limited to the problem of determining the friction
angle ' for w h i c h the body, under the weight of the soil and the influence of possible
external forces, is in equilibrium. T h e factor of safety is generally expressed as the ratio
of the tangent of the existing value of ' to the tangent of the ' value required for equi-
librium. T h e logarithmic spiral for which the lowest factor of safety is obtained is found
by trial and error.
If the cohesion intercept c' 0 the stabilising m o m e n t due to c' is obtained b y the
relation (Fig. 249):

COS0' 2 ,
Ms = -^(R-R)c' (452)
2tan0

4. COMPOSITE FAILURE SURFACE

In m o r e c o m p l i c a t e d situations with layers of varying shear strength the failure surface


is often following the weakest layers. T h e depth of the failure surface m a y also be
restricted d u e to limited depth to firm bottom. There is obviously a need of analytical
m e t h o d s taking into account an arbitrary shape of the failure surface.
As in the case of a circular-cylindrical failure surface, the b o d y is cut into slices with
width b (Fig. 2 5 0 ) .
The value of 7 ( a n d thereby AT) can be found from the condition of m o m e n t equilibrium.
For a slice with infinitesimal width dx w e have (Janbu, 1973):

Pht - (P + dP)(ht - dx t a n a , ) + (dWx + dW2)dx/2 = (T+dT+ T)dx/2 (453)

Dividing by dx and omitting terms of m i n o r significance, the shear force Tii+l at the
interface b e t w e e n slices / and i+1 is obtained by the relation:

Fig. 250. The slice method applied for the analysis of a composite failure surface.
346 Slope stability

dP
Ti,i+ = ( + i(tana,)M+ 1 (454)
dx

where at is the angle of inclination of the internal pressure line at the interface between
slices i and i+l.
The exact position ht of the earth pressure resultant, acting in the interface between two
slices, cannot be determined. However, the influence of its position on the safety factor
is relatively small and, therefore, from a practical viewpoint it can be assumed that it is
acting at the lower third of the sides of the slices.
The value of AP is obtained from a tangential projection equation along the base of the
slice:
(Wx + W2 + AT) una - AP c o s a = (455)

Inserting the value of ' (Eq. 448) into = ( c ' + a'timc^y/F^ and substituting / for
b/cosa w e find:
1 c'fe + (Wi + W2-ueb)ton<b '
= - - (456)
FC0 c o s a ( l + tan tan 0 F^)

2
Introducing na - c o s a (1 + tan a tan (p'/FC(p), w e obtain:

AP = B-A/FC(I) (457)

where A = [bc'+ (Wx + W2 + AT-bue) tanO"


= (Wx + W2 + AT) tana
The value of dPIdx in the interface between two slices i and i+l can be obtained by
the approximate relation (Fig. 251):

Fig. 2 5 1 . Approximate determination of dPIdx.


Slope stability 347

AP + APi+ 1
(458)
k +h

S u m m i n g u p over all slices, the condition of force equilibrium yields:

1AP = -P 0 = 1B-(FCO)-^1A (459)

=0 (460)

Introducing the values of and into Eq. (459) w e find:

l[bc ' + (W! +W2 + AT - bue) tan ']


(461)
[0 + 1(WX + W2 + AT) tan a]

2
w h e r e na = c o s a (1 + t a n a t a n ^ ' / F ^ ) .
Eq. (461) is referred to as J a n b u ' s general procedure of slices (GPS).

T h e iteration procedure takes the following course:


a s s u m e a potential slide surface and cut the sliding body into slices w h o s e widths are
governed by the strength properties of the soil,
calculate for each slice its weight (W=Wl + W2) and the value of ueb,
7
determine for each slice the value of t a n a and tan0 (and cO,
a s s u m e a value of Fc(p (for instance F^ = 1.5),
calculate the values of A (= A0) and (= B0) for AT= 0 and s u m up the values thus
obtained to get a corrected factor of safety Fc(p = 0/(0 + 0) ,
calculate AP = B0- AJF^ for the respective slices and determine the values at each
interface b e t w e e n the slices by summation of ,
determine the values of dPIdx at the interface between the slices according to Eq.
(458),
draw the thrust line through the lower third of the height of the elements at the
respective interfaces and determine its inclination (tanc^) at each interface between the
slices,
calculate the value of T a t each interface (which in turn yields A T o v e r the respective
slice),
check that = - P0 and that = 0,
the values of A T t h u s obtained yield new values of A(=A{) and (= ) and a new
factor of safety F^ = ^ I( + \),
repeat the procedure until Fc(p tends to a final constant value.
Fig. 252. Correction coefficient/^ with regard to the effect of internal forces acting between lamellae.

T h e safety factor is obtained from E q s . ( 4 5 5 - 4 6 1 ) by successive approximation. As


previously mentioned, the exact position ht of the earth pressure resultant acting against
the sides of a slice cannot b e determined but since the choice of position has little
influence on the result, it can be assumed that the earth pressure is acting at the lower third
of the sides of the slices.
T h e influence on the factor of safety of the internal forces acting at the interfaces
between the slices has been investigated for various types of failure surface. As a result
of the investigation, a correction factor fQ has been established (Janbu et al, 1956) by
which the safety factor, calculated without consideration of the internal forces, should be
multiplied (Fig. 252). Accordingly:

_^ ?Xbc' + {Wx+W2-bue)\suL$']
~ J o 7^ ^7777 777 ~ (462)

Example 70: The river embankment in example 69 is to be built upon with terrace houses and the factor
of safety required is F C (^> 1.5. For the construction of the buildings, the embankment is excavated and
2
the buildings placed as shown below. The weight of the buildings is 20 kN/ m , except for the uppermost
2
building whose weight is 26 k N / m . In order to guarantee against erosion by the river, a rock fill is placed
at the toe of the slope with an inclination of 1:3. The soil characteristics are those given in example 2.
3
The rock fill has = 2.0 t / m and '= 45. Determine the factor of safety for the composite failure surface
indicated.
Slope stability 349

Solution: The body comprised by the slide surface is sliced up, as shown, into lamellae 1 to 12. For each
of the lamellae the value and the total and effective normal stresses along the slide surface are
calculated. The calculation is carried out step by step by successive approximation. In the expression
for a it is assumed, as a first assumption, that FCQ = 1.5. It is further assumed, as a first approximation,
that = 0. The values of ( = B 0 ) and (= A 0 ) thus obtained yield a corrected factor of safety and are
applied in a second iteration for the determination of and which in turn yield new values of A ( = A { )
and (= j), etc. The analysis is carried out as shown in the arrangement shown below.

Slice b o WkN/m bue kN/m <t>' 0 ( k N / m ) A (kN/m)

1 8 42 540 0 37 486.2 507.3


2 5.5 37 855 129 35 644.3 589.6
3 34 1190 358 802.7 644.6
4 26 1185 410 578.0 547.2
5 " 20 1000 275 364.0 491.4
6 7 12 1030 224 218.9 536.6
7 6.5 8 995 163 139.8 557.5
8 5 6 640 100 67.3 364.4
9 3 525 60 27.5 318.7
10 " 0 370 0 " 0 259.1
11 7.5 -2 480 0 -16.8 342.1
12 8 -13 310 0 45 -71.6 385.9
3240.3 5544.5

The factor of safety F C (^ = Aq/ B 0 = 1.71, i.e. higher than the assumed value 1.5.

The values of A 0 and B 0 are now applied for calculating with a factor of safety of Fc(p = 1.7.
Introducing dP/dx - (, + AP t + ) / ( + bi+l) and the ht and tana, values at the interface between the
respective slices, the and values can be calculated as shown below. Inserting the AT values thus
obtained, new values of A ( = A { ) and ( = B{) are found as shown below. These values yield a revised
factor of safety.
350 Slope stability

Slice B0 ^0 dPIdx tana. ht AT *1 *1

- 0 0
1 486.2 507.3 189.7 -28 461.0 500.2
- 189.7 36.3 0.625 2.5 -28
2 644.3 589.6 299.7 -57 601.3 561.2
489.4 66.0 0.578 3.0 -85
3 802.7 644.6 426.0 -184 678.6 517.3
- 915.4 62.2 0.532 3.5 -269
4 578.0 547.2 258.2 -171 494.6 436.4
1173.6 30.5 0.466 3.5 -440
5 364.0 491.4 76.8 29 374.5 520.3
- 1250,4 -1.4 0.325 3.5 -411
6 218.9 536.6 -94.7 135 247.6 633.5
1155.8 -20.8 0.176 3.5 -276
7 139.8 557.5 -186.0 65 149.0 605.6
- 969.7 -28.8 0.105 3.8 -211
8 67.3 364.4 -145.7 -17 65.5 355.0
- 824.1 -30.5 0.140 3.7 -228
9 27.5 318.7 -158.8 18 28.5 332.0
- 665.3 -31.0 0.176 3.0 -210
10 0 259.1 -151.4 70 0 308.1
- 513.9 -29.4 0.140 2.3 -140
11 - 1 6 . 8 342.1 -216.7 90 -19.9 405.4
- 297.2 -33.2 0 1.5 -50
12 - 7 1 . 6 385.9 -297.2 50 -83.1 438.4
- 0 0 2997.5 5613.4

The revised factor of safety F c <p = 1/ 1 = 5613/2997 = 1.87

In a third iteration process the factor of safety found in the second iteration process is applied. The
result is as follows.
Slope stability 351

Slice B X dPIdx tana, ht AT B2 A2

_ 0 0
1 461.0 500.2 193.9 -29 460.1 511.6
- 193.9 36.7 0.625 2.5 -29
2 601.3 561.2 301.6 -65 595.3 666.1
- 495.5 64.0 0.578 3.0 -94
3 678.6 517.3 402.4 -172 686.6 536.8
- 897.9 60.3 0.532 3.5 -266
4 494.6 436.4 261.5 -160 499.9 450.7
- 1159.4 32.6 0.466 3.5 -426
5 374.5 520.3 96.7 20 371.2 519.9
- 1256.1 0.5 0.325 3.5 -406
6 247.6 633.5 -90.7 132 247.0 635.9
- 1165.4 -19.6 0.176 3.5 -274
7 149.0 605.6 -174.4 71 149.8 612.5
- 991.0 -26.0 0.105 3.8 -203
8 65.5 355.0 -124.1 -19 65.3 354.8
- 866.9 -27.3 0.140 3.7 -222
9 28.5 332.0 -148.8 2 27.6 321.6
- 718.2 -31.3 0.176 3.0 -220
10 0 308.1 -164.5 69 0 307.4
- 553.6 -32.1 0.140 2.3 -151
11 -19.9 405.4 -236.4 98 -20.2 410.5
- 317.2 -35.7 0 1.5 -53
12 - 8 3 . 1 438.4 -317.2 53 -83.8 436.2
- 0 0 2998.9 5764.3

The safety factor becomes F c <p = 5764/2999 = 1.92.

A fourth iteration yields 3 = 5678 kN/m and A 3 = 2999 kN/m whence F c 0 = 1.89.
Using instead the correction factor/ 0 to be applied on the safety factor calculated without consideration
to the internal forces between the slices we find the ratio Lid = 11/78 = 0.14, whence f0 = 1.08.
In this case the factor of safety becomes Fc<p = 1.08-1.71 = 1.85 which is about 2% below the value
obtained by the iteration procedure.

Example 71 : Investigate how the internal forces acting in the interfaces of the slices will affect the factor
of safety with regard to moment equilibrium of the circular-cylindrical slide surface analysed in
Example 69.

Solution: For this purpose we have to establish the equilibrium condition for each slice and for the sum
of slices according to Eqs. (454-460). Starting the iteration procedure on the assumption that 7 = 0
and that F c 0 = 1.05, as was previously found in Example 69, we get the following result (forces in kN/
m and lengths in m):
352 Slope stability

Slice b o W ueb
/0
^0

1 3.5 57.5 173 0 37 212.3 271.6


2 3 51.5 390 25 490.3 373.1
3 3.5 45.0 588 123 35 588.0 390.7
4 3.5 39.7 627 158 520.5 357.1
5 5 33.0 825 210 535.8 427.2
6 " 26.3 665 125 328.7 353.8
7 " 18.8 585 60 " 199.2 334.3
8 " 12.8 543 30 123.4 328.0
9 6 5.0 380 0 33.2 253.3
10 8 -4.0 180 0 -12.6 132.8
3077 3163

The values of A 0 and B0 are now applied for calculating AP with a factor of safety of Fc<^ = /^-
Introducing dPIdx = (, + APt + x)l(bi + bi + )l and the ht and tana, values at the interface between the
respective slices, the and AT values can be calculated as shown below. Inserting the AT values thus
obtained, new values of (= A t ) and (= ) are found.

Slice B0 AP dPIck tan<7, ht AT *1

_
0 0
1 212.3 271.6 65.0 -4 265.3 205.3
- 65.0 29.6 0.93 1.9 -4
2 490.3 373.1 127.2 -3 486.5 362.6
- 192.2 51.5 0.84 3.0 -7
3 588.0 390.7 207.8 -120 468.0 287.6
- 400.0 54.4 0.78 3.4 -127
4 520.5 357.1 173.0 -130 412.6 256.3
- 573.0 34.5 0.66 3.5 -257
5 535.8 427.2 120.0 -122 456.5 340.5
- 693.0 10.4 0.60 3.5 -379
6 328.7 353.8 -15.6 56 356.3 388.7
- 677.4 -14.2 0.41 3.2 -323
7 199.2 334.3 -126.1 37 211.7 356.4
551.3 -32.1 0.31 3.6 -286
8 123.4 328.0 -195.8 79 141.3 377.6
- 355.5 -37.2 0.25 3.2 -207
9 33.2 253.3 -213.3 133 44.9 341.6
142.2 -25.4 0.19 1.9 -75
10-12.6 132.8 -142.2 75 -17.8 188.6
3077 3163 0 0 2825 3105

The new ratio of , \ yields = 1 . 1 and, therefore, this value will be applied in a second iteration
process.
Slope stability 353

Slice Bx * 1
AP dPIdx tana, AT B2 A2

_
0 0
1 265.3 205.3 78.5 -4 265.3 212.6
- 78.5 36.2 0.93 1.9 -4
2 486.5 362.6 156.6 -26 457.6 354.2
- 235.1 55.8 0.84 3.0 -30
3 468.0 287.6 206.3 -127 461.0 289.2
- 441.5 55.1 0.78 3.4 -157
4 412.6 256.3 179.4 -119 422.6 271.6
- 620.9 38.4 0.66 3.5 -275
5 456.5 340.5 146.7 -133 449.4 339.5
- 767.6 14.9 0.60 3.5 -408
6 356.3 388.7 2.7 57 356.8 395.7
- 770.3 -11.0 0.41 3.2 -351
7 211.7 356.4 -112.6 34 211.7 356.4
657.7 -31.5 0.31 3.6 -317
8 141.3 377.6 -202.2 67 138.6 373.1
- 455.5 -42.6 0.25 3.2 -250
9 44.9 341.6 -265.9 152 44.9 341.6
- 189.5 -32.5 0.19 1.9 -98
10-17.8 188.6 -189.5 98 -19.4 204.7
2825 3105 0 0 2789 3098

We find Fc(p = 3098/2789 = 1.11 which is in good agreement with the value obtained in the first
iteration. The values of AT obtained in the second iteration can thus be used to find the influence of the
internal forces on the factor of safety according to Eq. (450). Inserting these values of 7 , we find:

Slice bm o WkN/m A7kN/m ueb kN/m ' AMJR kN/m

1 3.5 57.5 173 -4 0 37 106.2


2 3 51.5 390 -26 25 205.5
3 3.5 45.0 588 -127 123 35 191.2
4 3.5 39.7 627 -118 158 " 195.8
5 5 33.0 825 -133 210 " 261.4
6 " 26.3 665 57 125 334.1
7 18.8 585 34 60 " 314.0
8 12.8 543 67 30 " 344.5
9 6 5.0 380 152 0 " 336.5
10 8 -4.0 180 98 0 " 194.9
2490

This yields a stabilising moment around of 45-2490 = 111280 kNm/m and a safety factor equal to
Fc<p = 111280/105335 = 1.06.
The result shows that the influence of the internal forces can be neglected and that, consequently, the
simplified solution given by Eq. (451) can be applied. It also indicates that the factor of safety deter-
mined on the basis of equilibrium in the horizontal direction gives a slightly higher factor of safety than
moment equilibrium around the centre of the circular-cylindrical failure surface.
354 Slope stability

5. GENERAL ASPECTS

(i) Use of computer programs. The examples given in Sections 2 and 3 clearly demonstrate
the strenuous w o r k required in normal hand calculations, especially in consideration of
the need of finding the m o s t dangerous potential slip surface. Therefore, the analysis of
slope stability is generally carried out by the aid of advanced c o m p u t e r p r o g r a m s . T h e
results obtained by computer analysis, however, sometimes show peculiarities which
indicates a need for checking by hand analysis the m o s t critical slip surface obtained.
A m o n g the computer programs existing on the m a r k e t at present, those which are
based upon slope analysis methods satisfying both force and m o m e n t equilibrium
conditions [for e x a m p l e the method of analysis presented by Morgenstern and Price
(1965)] are most advanced.

(ii) Stress distribution in slopes. T h e in situ stress condition prevailing at various depths
across a slope is mainly affected by the situation between the toe and the crest of the slope.
T h e ratio of horizontal to vertical stresses is generally m u c h higer close to the toe than
close to the crest. T h e real stress variation may be quite different from the values assumed
in the analysis, but its effect on the safety factor does not seem decisive.
T h e most important influence on the result of effective stress analysis of slope stability
is of course the variation of pore water pressure with location and depth across the slope.
Therefore, the seasonal variations and their extremes h a v e to b e carefully examined in
order to h a v e a reliable basis for the analysis. T h e m o s t critical time of the year therefore
concurs with periods with high rain falls.

(iii) Retrogressive slides. In regions with clays of extremely high sensitivity (quick clays,
Fig. 21), the occurrence of a slope failure often leads to retrogressive slides which, in
consequence, m a y finally e m b r a c e large areas with catastrophic consequences. A typical
e x a m p l e of this p h e n o m e n o n is the earth slide at Rissa, north of Trondheim, Norway, on
April 2 9 , 1978 (a slide that has been d o c u m e n t e d in a film taken by a bystander). Here,
3
a small excavation for an extension of a farm house, comprising 7 0 0 m of soil that was
placed on the waterfront, initiated a slide with a width of around 20 m and a length of
2
around 80 m. A sequence of retrogressive slides finally e m b r a c e d an area of 330000 m .
Seven farmsteads and five dwelling houses disappeared with the slide.

(iv) Reinforcement of slopes. Slope stability conditions are generally quite improved by
the reinforcing effect of root systems of bushes and trees. N a k e d slopes ought to b e
planted with bushes in order to improve the stability of the superficial parts of the slopes.
This is particularly important in silty soil regions subjected to ground frost. Superficial
slides in such regions usually occur during spring when the ground is thawing.
Slope stability can also be improved by soil nailing.
Excavation 355

EXCAVATION

1. INTRODUCTION

T h e lack of s uitable ground for building purposes and an increasing need of parking space
has entailed a general tendency to provide n e w buildings with an ever increasing n u m b e r
of b a s e m e n t floors. T h e cost of these b a s e m e n t floors will h a v e to b e related to the cost
of excavation in order to reach a cost-effective design. Quite often deep excavations
cause serious p r o b l e m s , giving rise to economical disputes on the contract work. O n c e
the construction of the building has reached ground level, unforeseen p r o b l e m s are con-
siderably lessened. Therefore, an optimal and reliable solution to the excavation prob-
lems is an important part in foundation design.
T h e m e a n s and type of machinery anticipated by the contractor in an excavation
contract are very important for the tendering price. Therefore, deviations from reality in
the d o c u m e n t s provided for the tenderers m a y lead to unforeseen claims for compensation

Fig. 253. Example of weathered rock below a seemingly un weathered rock surface.
356 Excavation

with possible arbitrations or court trials in consequence. T h e m o s t c o m m o n disputes


concern the soil and rock characteristics (classification with regard to diggability,
strength properties, depth to bedrock, groundwater conditions, etc.). T h e properties of
covered bedrock material, in particular, are difficult to predict from the results of
conventional ground investigations (cf. Fig. 253).
T h e m o s t urgent p r o b l e m encountered in connection with deep excavations is h o w to
ensure the stability of the sides and the bottom of excavation with regard to the shear
strength of the soil and groundwater condition. Sides of deep excavations in densely
populated areas, for the most part, have to be supported by anchored sheet pile walls. In
order to safeguard stability, additional measures m a y h a v e to b e taken, above all, by
dewatering or lowering the groundwater pressure. Injection of weathered rock or
fractured zones m a y be required to avoid leakage and/or internal erosion.
T h e time during which the excavation is left open is another factor of great importance.
T h e longer the time, the higher the stability requirements.
T h e design of the excavation is doubtless just as an important part of a buildi ng contract
as the design of the building itself.

2. DEWATERING

2.1 Hydraulic ground failure

There are mainly two causes of hydraulic ground failure: instability d u e to liquefaction
p h e n o m e n a , or instability due to the hydraulic uplift of impervious soil, underlain by
pervious soil in which the pore water pressure exceeds the weight of the overlying
impervious soil.

(i) Liquefaction. As mentioned on p. 2 1 , liquefaction due to u p w a r d groundwater seepage


can b e expected w h e n the hydraulic gradient / in the vertical u p w a r d direction is equal
to p'lpw In a physical sense this m e a n s that the soil behaves as a heavy liquid with a unit
weight equal to the unit weight of the soil. Expressed in the terms of porosity of soil or
void ratio e, liquefaction will occur when i - > 1.65(1 - n) = 1.65/(1 + e).
T h e liquefaction condition / = p 7 p w is based on the assumption of n o interlocking
forces between the soil particles and, therefore, in realityparticularly in dense s o i l
the hydraulic gradient m a y exceed p'lpw without liquefaction taking place.
T h e most c o m m o n type of liquefaction p h e n o m e n a appearing in deep excavations
below the groundwater table is the formation of erosion channels (piping) in the b o t t o m
of excavation. T h e u p w a r d seepage pressure carries away the finer particles u p to the
bottom surface where they settle in a cone-shaped form (Fig. 254).
T h e risk of liquefaction should be checked by pore pressure observations during
excavation. If the pore water pressure at the point of observation tends to the total
overburden pressure, the stability of the excavation is in jeopardy. In the design of a deep
Excavation 357

Fig. 254. Hydraulic bottom failure piping due to upward seepage of groundwater. Fine particles
are carried to the bottom of excavation by the seepage pressure and settle in cone-shaped forms around
the seepage channels.

excavation b e l o w groundwater level w h e r e there is a risk of liquefaction, detailed


instructions regarding h o w to control the stability conditions should b e included. T h e
pore water pressure should b e lowered to its acceptable value already before the ex-
cavation is started and then b e subjected to continuous observations in case an unforeseen
increase in the pore water pressure should take place. For e x a m p l e , the total pressure
relief caused by excavation in l o w - p e r m a b l e fine-grained soil will lead to a simultaneous
reduction of the p o r e water pressure b e l o w the b o t t o m of excavation. This m a y lead to
misjudging the risks involved. It has to b e realised that with time the p o r e water pressure
will b e restored to its original value due to inflow of water from the surroundings.

(ii) Hydraulic uplift. In connection with deep excavations in clay, the pressur exerted by
the pore water in sand or silt layers, or in pervious b o t t o m layers, often cause hydraulic
uplift of the b o t t o m of excavation. Alternatively, the reduction of effective stresses due
to excavation entails a reduction in average shear strength b e l o w the b o t t o m of
excavation which m a y lead to failure. It is often difficult to distinguish which of the two
p h e n o m e n a triggers a failure occurrence.
T h e existence of of silt or sand layers in clay deposits always has to b e m a p p e d and
358 Excavation

considered in stability analyses. Their existence is particularly risky at sites located in


deep valleys. In such cases the silt and sand layers are usually fed with water under
artesian pressure from the surroundings.
Excavations in clay m a y be carried out to great depth b e l o w the g r o u n d w a t e r table
without any visible sign of water seepage. Unless pore pressure observations are carried
out, the risks involved m a y not b e realised until it is too late and hydraulic uplift of the
b o t t o m of excavation has suddenly occurred (Fig. 255).
T h e only possible w a y of eliminating the risk of hydraulic uplift is to lower the pore
water pressure in the pervious soil to an acceptable level before excavation is started.
Hydraulic uplift m a y also b e avoided b y e n c o m p a s s i n g the excavation by a cut-off sheet
pile wall that prevents the inflow of water from the surroundings. In this way, the decrease
in pore water pressure that is obtained due to the pressure relief caused b y excavation will
be maintained throughout the excavation period. However, one has to m a k e sure that
water inflow is completely cut off, otherwise excavation m a y result in hydraulic uplift
in spite of all the precautions taken.
Even in cases w h e r e the clay is underlain by liquefied silt, hydraulic uplift will take
place because of inflow from the surroundings of soil in liquid state (Fig. 256).
In the analysis of hydraulic uplift, the stabilising effect of the vertical shear planes along
the sides of the excavation is often taken into account. This m a y b e justified for narrow

Fig. 255. Excavation in clay for a waste water pump station immediately after hydraulic uplift had taken
place. Depth of excavation about 11 m; diameter about 23 m. The uplift was caused by too high a pore
water pressure in thin silt layers, wedging into the excavation just below the tip of the encompassing
sheet pile wall.
Excavation 359

Fig. 256. Liquefied silt giving rise to hydraulic uplift of the bottom of excavation is pumped away by
means of an Archimedian screw. Liquefaction caused by stress release during excavation for shopping centre.

excavations w h e r e the 'soil b e a m ' (or 'soil p l a t e ' ) can resist the bending m o m e n t caused
by the uplift pressure. In the case of wide excavations it is not justified.

Example 72: In order to create a dry dock for the construction of oil rig platforms, a circular excavation
with a bottom diameter of 120 m was carried out to a depth of 12.7 m in sand and silt underlain by clay.
The excavation was enclosed by a sheet pile wall driven to a depth of minimum 2m into the clay. Below
the bottom of excavation the clay contained seams of sand and silt. The density of the clay was 1.9 t/
m 3 and its undrained shear strength 40kPa. In order to avoid hydraulic uplift, relief wells (vertical sand
drains) were installed to a depth of 10 m below the bottom of excavation. The area around the site was
mountainous and the decrease in pore pressure in the silt and sand layers below the relief wells, due to
unloading by excavation, could be expected to vanish with time because of inflow of water from the
surroundings. In truth, piping and local hydraulic failure, forming the lake that is visible in the figure
below, occurred when the dock was emptied from water for the construction of the second platform.
Determine the maximum pore water pressure, leading to hydraulic failure of the bottom of
excavation, with regard to the 'plate' stiffness of the 10 m thick clay layer drained by the relief wells
if the short-term tensile (or compressive) strength in the soil is (a) 40 kPa, (b) 80 kPa.
360 Excavation

Solution: Assuming that the clay layer behaves as an elastic material, rigidly fixed along its boundary,
the maximum stresses in the horizontal direction are obtained by the relation:
2
, 3(1 + v)(u-\.9gh)R
G h =
W
at the centre of the excavation, and:

(/, = r
2
4h
at the outer boundary of the excavation.
Introducing v = 0.5 (undrained condition), R = 60 m and h = 10 m, we find oh = 20.25w - 3774 kPa
at the centre and ah = 27.Ou - 5032 kPa at the outer boundary.
In case (a) we have um.dX = 188 kPa and in case (b) w m ax = 1 8 9 kPa. In both cases the critical stress
value is almost the same at the centre and the boundary of the excavation. If the tensile strength is
assumed zero we have w m ax = 186 kPa. The contribution by the plate effect is negligible.

Example 73: A vertical cut in clay, 1.5 m in width, shall be carried out in clay with a density of 1.6 t/
3
m and an undrained shear strength of 20 kPa. The clay is underlain by sand at 5 m depth. The hydraulic
head in the sand layer is 1 m below the ground surface. Determine the allowable depth of excavation
d.dX if the partial safety factor on undrained shear strength is 1.5.

Solution: If the contribution of the shear planes along the sides of the cut are taken into consideration
we have:
( 5 - 4 1 > l ^ + 2 { 5 - J a l) - 2 0 / 1 . 5 = 4^
which yields dal = 4.1 m.
Ignoring the side effects we have d.dl = 5 - 4/1.6 = 2.5 m
If the clay layer, left below the bottom of a cut, carried out to a depth of 4.1 m, is considered as an
elastic beam, rigidly fixed along the sides of the cut, the tensile and compressive horizontal stresses in
the clay along the sides of the cut become equal to:
2
Ju-\.6gh)-l
Oh -
2
2h
Inserting the values h = 0 . 9 m, / = 1.5 m and u-Ag kPa, we find oh=32.1 kPa. It is doubtful iflong-
term tensile stresses of this magnitude can be allowed.

2.2 Damages caused by groundwater lowering

In cases w h e r e it is decided to stabilise the excavation by reducing the groundwater


pressure one has to keep in m i n d what effect this might h a v e on existing buildings in the
neighbourhood. T h u s , groundwater lowering in regions with compressible soils, such as
normally consolidated clays and organic soils, m a y cause serious d a m a g e to other
buildings. In order to minimise possible damage, the hydro-geological conditions ought
to be m a p p e d very carefully, not only at the site itself but also in the surroundings where
a negative effect on on the existing groundwater condition can b e suspected of causing
d a m a g e (Figs. 2 5 7 - 2 5 8 ) . For the assessment of possible d a m a g e , the proprietor of the
building project w h o is responsible ought to m a k e a continuous check on possible
changes in the groundwater pressure during the time of construction. T h e costs of such
a check are insignificant in comparison with the cost of repairs that m a y be b r o u g h t down
on the proprietor. Claims on cost of repairs m a y also include old d a m a g e s which have
Excavation 361

Fig. 257. Leakage of groundwater into rock tunnels, here visible by the existence of stalactites, may
cause serious problems.

WHm> groundwater l o w e r i n g > 10 m


5 - 10 m
crush-zones

Fig. 258. Pore pressure decrease in sand beneath a 10 m deep clay deposit due to leakage of groundwater
into a raw water tunnel. N o leak observed in the tail race tunnel. Maximum decrease in pore pressure
observed about 300 m away from the leaking raw water tunnel. A large number of dwelling houses
kilometres away were damaged due to differential settlements caused by the pore pressure decrease.
362 Excavation

nothing to do with the building project in question. By m a k i n g continuous pore water


pressure observations during the time of construction such claims can b e rejected by
submission of evidence that no changes in the groundwater situation has occurred.

Example 74: A normally consolidated, homogeneous clay layer is underlain by sand at 20 m depth. Due
to water outtake in the sand aquifer the hydraulic head in the sand drops 10 m. Determine the pore
pressure distribution in the clay layer and the subsidence after 40 years if the clay has a coefficient of
2
consolidation ch = 0.5 m /year and a compression modulus M = 200 kPa. The groundwater level is at
2 m depth and the original pore pressure distribution is hydrostatic. Due to the influence of superficial
fissures and root systems and precipitation, the pore water pressure distribution remains hydrostatic to
a depth of 6 m.

Solution: The consolidation process is most easily determined by the graphic method suggested by
1 /2
Helenelund (p. 325). Choosing t = 8 years we have Az = 2 ( 0 . 5 8 ) = 4 m. The graphical determination
requires five isochronal lines and is carried out as shown below.

6 m

2m

4 m

Au= 10-9.81 kPa

The decrease in hydraulic head after 40 years is 7.6 m at 18 m depth, 3.5 m at 14 m depth and 1.1m
at 10 m depth.
Since in this case - Au = ', the settlement becomes:
s HI.6 + 3.5 + 1.1)9.81/200 = 2.4 m

2.3 Methods of groundwater lowering

F r o m what is said above, groundwater lowering is often required in order to m a k e


possible excavation to full depth. G r o u n d w a t e r lowering generally entails also other
advantages for the contractor. A m o n g the m o s t important achievements w e find:
hydraulic ground failure is prevented,
seepage of water through the sides and the b o t t o m of excavation is prevented,
the stability of the sides of excavation is improved,
Excavation 363

loss of material by internal erosion is prevented,


the earth pressure against e n c o m p a s s i n g sheet pile walls is reduced,
excavation and loading of sand and silt are facilitated.
T h e m e t h o d applied for groundwater lowering has to b e selected with due regard to the
soil characteristics and to the type of excavation. A m o n g the m o s t c o m m o n m e t h o d s w e
find:
draining of s u m p s in the b o t t o m of excavation,
p u m p i n g of water from well points or deep wells,
dewatering by the use of electro-osmosis,
pneumatic methods.
T h e field of application of the different m e t h o d s of groundwater lowering is presented
in Fig. 259.
T h e first d o c u m e n t e d case of groundwater lowering w a s carried out in 1838 in
connection with the construction of the Kilsby tunnel for the L o n d o n B i r m i n g h a m
railway. T h e well point m e t h o d was first applied in connection with tunnel constructions
in Berlin in 1896 while electro-osmosis was first utilised in 1939 for the stabilisation of
a railway cut in Salzgitter.

(i) Draining of sumps. In case the hydraulic conductivity of the soil is high, as in gravel
and cobbles, a very high flow rate is required before the critical hydraulic gradient is
reached to lead to liquefaction. Dewatering can then b e carried out from s u m p s placed

Fig. 259. Field of application of the most common methods of groundwater lowering.
364 Excavation

in the b o t t o m of excavation without risk of liquefaction. T h e m a i n p r o b l e m in this case


is to ensure e n o u g h p u m p i n g capacity. If the p u m p i n g capacity turns out to be insufficient,
either injection of cement grout into the voids of the soil m a y b e required to r e d u c e the
hydraulic conductivity or excavation has to be carried out under water. Installation of
sheet pile walls around the excavation is another alternative.
Draining of s u m p s m a y also b e possible in dense, fine-grained soils, such as till. A n y
tendency of grain displacements in these soils gives rise to an increase in the contact
pressure between the individual grains (increased effective stress) due to interlocking
effects and dilatancy which counteracts liquefaction. However, in this case great care has
to b e taken to avoid disturbance of the b o t t o m of excavation. Disturbance will strongly
reduce the favourable influence exerted by dilatancy and induce superficial liquefaction.
If there are signs of liquefaction tending to p e r m e a t e deeper d o w n into the soil, a
protecting, inverted filter should be placed on the b o t t o m of excavation. T h e filter must
be coarse e n o u g h to permit free outflow of seepage water, but fine e n o u g h to prevent
escape of soil particles through the voids. For e x a m p l e , the filter can b e m a d e up of 5 cm
sand at the bottom, covered by 10 c m of gravel and, on the top, 15 c m of shingle or
m a c a d a m . T h e filter ought to cover also the sides of the excavation and the sumps.
Excavation m a y h a v e to b e carried out in small part areas, one at a time, starting at the
location of a s u m p , immediately followed by placing of the filter.
By installation of cut-off sheet pile walls around the excavation, the flow path can be
prolonged and consequently the hydraulic gradient reduced (Fig. 260). T h e depth of the
sheet pile wall required to eliminate the p r o b l e m of water inflow can b e determined by
the use of so-called flow nets. T h e result of the flow net analysis is very m u c h dependent
on the h o m o g e n e i t y of the soil. Unfortunately, in m o s t cases soil contains layers of
varying permeability which limits the value of the flow net m e t h o d . Therefore, the

Original Sheet pile walls


groundwater level 1/ \

100% 100%

Fig. 260. Flow net and potential lines around excavation with sheet pile walls installed for the purpose
of reducing the hydraulic gradient. Figures indicate the hydraulic head in % of the original hydraulic
head above the bottom of excavation (before groundwater lowering).
Excavation 365

Fig. 2 6 1 . Dewatering of an open excavation by means of two well point systems placed at different
levels.

necessity of p o r e pressure observations during excavation is equally important in this


case as in.a case w h e r e cut-off sheet pile walls are not utilised.

(ii) Well points. For excavations in fine-grained soils, drainage of s u m p s would entail too
large a risk of liquefaction. T h e pore water pressure has to be reduced at greater depth
before excavation below the groundwater table is started, otherwise difficulties are likely
to turn up.
O n e of the m o s t c o m m o n methods used for this purpose is to install filter wells, so-
called well points, that e n c o m p a s s the excavation. T h e well points are usually placed in
a row with 1-4 m spacing just outside and above the intersection line between the sides
of the excavation and the groundwater table. T h e well points are connected to a v a c u u m
p u m p by a joint header pipe. T h e depth of installation varies from 5 to 9 m below the
suction pipe.
T h e well points can b e utilised in t w o w a y s , either as gravity wells or as v a c u u m wells.
In the former case, the filter wells are under prevailing air pressure (in connection with
open air) and in the latter case under reduced air pressure (shut off from open air). In
gravity wells the water flows towards the wells by the influence of gravity alone, while
in v a c u u m wells water is sucked out of the soil. Vacuum is achieved by sealing the wells
with clay or bentonite slurry. T h e v a c u u m m e t h o d will have to be used w h e r e the soil
consists of fine sand or silty sand in which the water is bound by capillary forces.
T h e lifting height of the water up the return pipe by m e a n s of a v a c u u m p u m p is limited
to about 7 - 7 . 5 m. In practice, the decrease in hydraulic head at the centre of ordinary
excavations, achieved by the gravity method, is limited to about 3.5 m. By using the
vacuum method, a m o r e deep-going and considerably flatter depression of the groundwater
366 Excavation

level is obtained. If the groundwater level has to b e lowered to greater depth this has to
be carried out in stages b y m e a n s of two or several well point systems placed at different
levels (Fig. 261).
A rough estimate of the groundwater lowering sm achieved by the gravity m e t h o d in the
middle of a circular area with radius R0 surrounded by wells can b e m a d e by the relation:

5, = / ^ / / / - % 1
2
(463)

w h e r e H = the original hydraulic head above the impervious b o t t o m layer,


R = radius of influence,
Qw = total a m o u n t of p u m p e d water per time unit,
k = permeability of soil subjected to groundwater lowering.
If the wells are placed in rows, forming arectangle with sides and b, / ? 0 c a n be calculated
by the relation:

fob
Ro = V-

(464)

If and b differ considerably (a > b), RQ can be obtained from:

R0 = (a + 2b)/5 (465)

A rough estimate of the R value can be m a d e by the relation:

/ 0 9kFt

i0
R= R
+~ (466)
w h e r e t - time of p u m p i n g ,
= porosity of the soil.

Example 75: An excavation with an area of 15 m by 10 m shall be earned out to a depth of 3 m below
the groundwater table in silt underlain by bedrock at 20 m depth. In order to avoid problems during
excavation the groundwater table is lowered by installation of well-points immediately outside the
boundaries of the excavation. The aim is to achieve a depression on the groundwater level of 4 m in the
centre of the excavation one week after pumping has started. Determine the required pumping capacity
if the silt has a porosity of 40% and a permeability of 300 m/ year.

1 /2
Solution: We have R0 = (15 1 0 / ) = 6.9 m. The R value after one week of pumping becomes equal
to:
2 1 /2
R = [ 6 . 9 + 930020/(520.4)] = 51 m

The pumping capacity required is obtained from the relation:


Excavation 367

4 = 20-J(202 ^ l n
V 3000 6.9
3 3 3
whence Qw > 67.8 1 0 m / y e a r = 0.13 m / m i n .

(iii) Ejector well points. In ejector well points, also called deep wells, v a c u u m is created
at the tip of the wells by m e a n s of a high-speed flow of water through a j e t nozzle. T h e
water sucked out of the soil is then forced by the water j e t stream u p the return pipe. By
the u s e of ejector well points the lifting height of the water u p the return p i p e can b e u p
to 6 0 - 1 0 0 m. This m a k e s possible a very deep-going lowering of the g r o u n d w a t e r level
and the n u m b e r of wells required can b e reduced accordingly (Fig. 262).

(iv) Electro-osmosis. In soils consisting of fine to m e d i u m silt there is no possibility to


accomplish dewatering by m e a n s of well points because of the water-binding forces
being too strong. However, by applying an electric gradient to the soil, hydrated cations
will b e transported b y the current from the anod towards the cathod. T h e water transport
thus taking place is n a m e d electro-osmosis.
In practical application of electro-osmosis, the anodes generally consist of iron bars or
iron tubes while the cathodes m a y consist of perforated tubes enclosed by inverted filters.
T h e water transported to the cathodes is p u m p e d u p , for instance by m e a n s of ejector
p u m p s . W h e n using electro-osmosis in large scale, special cathodes designed as wells are
generally utilised from which the water is p u m p e d up.
T h e a m o u n t of water transported by electro-osmosis is governed by the electric
potential drop between anodes and cathodes and the electro-osmotic permeability
coefficient. D u e to the electro-osmotic water transport, an excess p o r e water pressure will

Three-stage set up of well points Ejector well points

Fig. 262. Comparison between stagewise lowering of the groundwater level by ordinary well points and
lowering in one step by means of ejector well points.
368 Excavation

be built up in the direction of the cathodes with a consequential hydraulic gradient


counteracting the electro-osmotic flow. Dewatering at the cathodes is a means of
reducing the build-up of the counteracting hydraulic gradient.
The combined effect of electro-osmotic flow and hydraulic flow can be expressed by
the relation:

vx = - k e ~ - k j x (467)
OX

where ke = electro-osmotic permeability coefficient,


dVxJdx = electric potential gradient,
kx = hydraulic permeability coefficient in the direction,
ix = hydraulic gradient in the direction.
By dewatering of the cathode, the influence of hydraulic flow can be neglected and the
amount of water Qw transported to the cathode during time t can be expressed by the
relation:

Qw ~ ke(Vlx)At (468)

where A - cross-sectional flow area.


Substituting (V/x)A for J, where pe = electric resistivity of the soil and current, we
finally have:
Qw~kepjt = pjt (469)

The parameter pe = kepe yields the amount of water transported per electric charge unit.
The consumption of electric charge required to reach a certain desired result depends to
a great extent on the design of the electro-osmotic field (the choice of electrodes and their
location). The total resistance in the electro-osmotic field which governs the interrelation
between amperage and voltage (according to Ohm's and Kirchhoff's laws) is dominated
by the electrode to soil resistance.
In the case of a staff-shaped electrode, the electrode to soil resistance R is given by the
relation (Kupfmller, 1959):

Pe 4L
R =ln() (470)
2nle De

in which le = length of the electrode below groundwater table,


De = diameter of the electrode.
The total electrode to soil resistance for electrodes coupled in parallel can be obtained
from the relation (Mourn, 1967):
Excavation 369

15
3
\ p = 1.9 t/m

I 10

i . p = 1.8 t/m
3

3
6 t/m

3
= 1.4 t / m ^ " ^ -

5 10 15 20
Sait content (equivalent NaCl), %o

Fig. 263. Correlation between electric resistivity of mineral soil and salt content. (Larsson, 1975).

D M l n [ ( c o t ( f l 2 / 2 ) - c o t ( 0 3 / 2 ) - cot(012)} R
K, = { 1 ) (471)
ll nn((8b4V
/ D^, )) -- l
1 n

w h e r e = axctan(nSJle),
Se = distance b e t w e e n the electrodes.

40 60 100

Water content,

Fig. 264. Electro-osmotic permeability coefficient of mineral soils as a function of water content.
370 Excavation

Both ke and pe can be determined in the laboratory. Measurements have shown that, for
a given density of mineral soil, there is a close relationship between the electric resistivity
and the salt content of the soil (Fig. 263) while ke is mainly a function of the water content
and the salt content (Fig. 264)
In order to maintain full efficiency of electro-osmotic dewatering, the consumption of
the anodes must be considered (Fig. 265). This can be calculated on the basis of Faraday's
law. Accordingly, the weight AGa (in grammes) set free by the passage of Qe coulombs
is given by the relation:

Qe atomic weight
AGa = (472)
96494 valence

For iron anodes, having an atomic weight = 55.9 and a valence = 2, the consumption
AGa (in milligrammes) is 0.29 Qe.
A classical example of the use of electro-osmosis for dewatering purposes is the
successful application designed by Leo Casagrande to achieve a 12 m deep excavation
in very soft, water-saturated silt for a submarine pen in Trondheim during the second
world war (Fig. 266). Previous attempts to reach the required 12 m excavation depth by
the use of a 17 m deep sheet pile cut off led to liquefaction and sheet pile failure already
at 8 m depth of excavation (Casagrande, 1947).

Fig. 265. Iron anode after termination of electro-osmosis. Most of the material is consumed and the
anode is steaming.
Excavation 371

Fig. 266. Failure of sheet pile wall during excavation for submarine pen in Trondheim and solving the
problem by the use of electro-osmosis. The electro-osmotic installation can be seen in the background.
(After Casagrande, 1947).

Example 76: In order to stabilise the slope of an excavation in silty clay, two rows of electrodes are
installed along the slope, of which the anodes are rails with a mass of 25 kg/m and the cathodes, which
are constructed as pumping wells, are made of steel with an outer diameter of 300 m. The distance
between the rows is 4 m, the spacing between the electrodes in the rows is 2 m, and the embedded length
of the electrodes is 10 m. The number of anodes is 10 and the number of cathodes is 9.
Determine the required pumping capacity and how long time it will take before the anodes are
consumed if the electro-osmotic permeability coefficient of the soil is ke = 2-1 ( ) 8 m 2/ s V and the resisitivity
is pe = 6 . Only the electrode to soil resistance needs to be considered. The electric potential anode
to cathode is 30 V.

Solution: The equivalent diameter of the anodes (the specific density of iron is 7.85 t/m 3) is equal to:
d = 2(0.025/7.85) 1 /2 = 0.064 m
The resistance anode/soil becomes:

tfa/s = l n ( ^ ^ ) = 0.61
210 0.064
372 Excavation

and the resistance cathode/soil:

Rc/S = ln(^^) = 0.47


210 0.30

The anode to soil resistance for the whole system is determined by Eq. (471). We have:
= arctan(rc-2/10) which yields:
R& =( 0.61/10){ 1 + ln[cot(0.190)-cot(0.270)-cot(0.337)-cot(0.393)-cot(0.438)-cot(0.475)-
cot(0.506)cot(0.532)-cot(0.554)]/[ln(8-10/0.064)- 1]} = 0 . 1 4 0
Regarding the cathode to soil resistance for the whole system we have:
Rc = (0.47/9){1 + ln[2662-tan(0.554)]/[ln(8-10/0.3) - 1]} = 0.137 Qm
The total resistance becomes R.d + Rc = 0.277
The electric currency / = 30/0.277 = 108 A and the water carried from the anodes to the cathodes is
equal to:
8 3
Qw = 2-10 -6-108-24-3600/9 = 0.124 m per day and cathode well.
6
The anode consumption is AG = 0 . 2 9 - 1 0 108-24-3600/10 = 0.27 kg per day and and anode. Each
anode has a mass of 250 kg. Thus the anodes will last maximum 2.5 years.

(v) Pneumatic method. B y the p n e u m a t i c m e t h o d , the air p r e s s u r e in air-tight caissons,


inside w h i c h the excavation is b e i n g carried out, is kept h i g h e r than the g r o u n d w a t e r
pressure. W h e n the air p r e s s u r e , required to p r e v e n t g r o u n d w a t e r s e e p a g e into the
caisson, tends to 350 kPa,.the w o r k i n g condition will w o r s e n at the risk of causing caisson
decease, and then m a n y p e r s o n s are not able to w o r k efficiently m o r e than 1 - 1 . 5 hours
at a time.
B e c a u s e of the e x c e s s air p r e s s u r e acting u p o n the ceiling of the caisson, it m a y be
necessary to e m p l o y a n c h o r a g e against uplift.
T h e m e t h o d requires very e x t e n s i v e safety a r r a n g e m e n t s .

Example 77: A heat pump in operation gives rise to a groundwater lowering followed by consolidation
of a clay deposit. The clay deposit, which is normally consolidated and 10 m in thickness, is underlain
by sand. The groundwater level is at 1 m depth with hydrostatic groundwater condition. Determine the
settlement taking place in 10 years if the groundwater lowering during 4 months per year is 0.3 m on
the average. The consolidation characteristics of the clay are as follows:
2
cv = 0.25 m /year; M = 100 kPa from the groundwater level to a depth of 7 m
2
cv = 0.55 m /year; M = 300 kPa from 7 to 10 m depth.

Solution: Choose the time interval = 1 year. Then, the layer thickness according to Eq.(436) becomes
equal to:
Az - 2"/0.25 = 1.0 m from 1 to 7 m depth
Az - 2TJ0.55 = 1.73 m from 7 to 10 m depth (which can be approximated to Az = 1.75 m)
The permeability k = cvyJM = 0.025 m throughout the clay deposit.
The groundwater level, at 1 m depth, depends on the amount of precipitation and evaporation and is
assumed not to influence the pore water pressure in the sand below the clay deposit. The total time of
groundwater lowering during a 10 years period is 4 years.
The graphical determination of the pore pressure decrease is carried out as shown on next page.
We find, since -Au = ' that ' is about 2.2 kPa in the middle of the layer nearest to the sand, and
0.8 kPa, 0.2 kPa and 0.05 kPa in the middle of the consecutive layers above. This yields:
s = 1.5(2.2 + 0.8)/220 + (0.2 + 0.05)/100 = 0.023 m
Excavation 373

2.4 Reinfiltration of water

In order not to d a m a g e neighbouring buildings by groundwater lowering it m a y be


necessary to reinfiltrate water into the soil to maintain, or to reestablish, the original
groundwater balance in the surrounding of the building under construction. T h e m e t h o d
of infiltration will depend on the soil properties. Mostly, infiltration is carried out from
boreholes or recharge wells in contact with pervious soil layers. This technique, however,
has certain d r a w b a c k s . Difficulties arise in finding proper installation points w h e r e the
hydraulic connection with the total area under consideration can b e guaranteed. Moreover,
clogging p r o b l e m s (siltation, bacteriological activity, etc.) are often encountered.
A very promising infiltration m e t h o d has been tested in s o m e infiltration plants in
S w e d e n w h e r e the original groundwater balance has been restored in large areas sub-
jected to groundwater lowering. T h e infiltration has taken place from infiltration tunnels
blasted in the bed-rock and put under water pressure corresponding to the original
g r o u n d w a t e r level. T h e water pressure in the tunnel has been transmitted to the w h o l e area
under consideration by boreholes driven from the tunnel in all directions to ascertain
contact with crush-zones in the bed-rock. T h e hydraulic head of the filled-in water is kept
constant by m e a n s of a simple charging device.
374 Excavation

3. U N B R A C E D EXCAVATIONS

3.1 Introduction

After appropriate measures have been taken to avoid stability problems due to unfavourable
groundwater conditions, other causes for instability or excessive ground displacements
in connection with excavation have to be taken into account. T h e excavation has to be
carried out in such a m a n n e r that the shear stresses induced do not approach the shear
strength of the soil. T h e measures required depend on type of soil, depth of excavation
and neighbouring conditions.

3.2 Sloping sides

In m a n y cases the excavation can be carried out without the need of side supports.
Generally, the only measure to fulfil side stability requirements is to carry out the
excavation with sloping sides.
In cohesionless soil the slope angle is determined by the angle of repose of the soil
(approximately equal to the internal critical angle of friction). If there is a load acting in
the near vicinity of the slope the stability condition will h a v e to be checked by a detailed
analysis searching for the most dangerous failure surface. In cohesionless soil the shape
of the failure surface can b e assumed to be a logarithmic spiral.
In cohesive soil the short-term stability of the slopes is determined by the slope angle
and the undrained shear strength of the soil. Since excavation leads to a long-term
reduction of the effective stresses in the soil, long-term stability has to b e determined on
the basis of effective strength parameters. T h e sliding surface can generally be assumed
to h a v e the shape of a circular cylinder. T h e most dangerous sliding surface can be found
by trial and error. In the case of simple geometry and h o m o g e n e o u s soil conditions the
slope stability conditions can be determined by the aid of diagrams.
T h e analysis of the stability of sloping sides of an excavation in simple and more
complicated situations was treated in the section dealing with slope stability, p. 336.

3.3 Cantilevered walls

Sheet pile walls represent the most c o m m o n type of retaining structure for the support of
the sides of deep excavations. In cases w h e r e the depth of excavation is limited, canti-
levered sheet pile walls can be utilised. T h e earth pressure distribution in such a case
depends on the bending rigidity of the wall and the depth of e m b e d m e n t (Rowe, 1951).
Obviously, in the design of a cantilever wall both the conditions of equilibrium and the
limits set to wall m o v e m e n t s will have to be satisfied. F r o m a theoretical viewpoint, the
latter requires good k n o w l e d g e of the soil properties and finite e l e m e n t analysis.
T h e stability against failure can b e treated in a fairly simple way. Accordingly, the wall
is a s s u m e d to b e subjected to classical Mohr-Coulomb type of earth pressure (pp. 278-279).
Excavation 375

P3/(d-zr)

Fig. 267. Earth pressure distribution along a cantilevered wall.

Below the centre of rotation (Fig. 267) the resulting earth pressure is replaced by a force
acting at the centre of rotation. With the symbols given in Fig. 260, the requirements of
equilibrium yield:

Pizl-P2z2 =0
P3 = P2-PX (473)
P3/(D-zr) = PpiJF-Par

w h e r e Ppr = passive earth pressure at the centre of rotation (on the active side),
Par = active earth pressure at the centre of rotation (on the passive side),
F = factor of safety.
Normally, the depth of e m b e d m e n t d is taken as m i n i m u m 1 2 z r.
For the d e v e l o p m e n t of the earth pressure needed to balance the rotational m o m e n t of
the wall, caused b y earth pressure a b o v e the b o t t o m of excavation (cf. D e e p foundations,
Section 2.6), quite large lateral and rotational m o v e m e n t s of the sheet pile wall generally
take place. T h e chances of failure or unacceptable time-dependent m o v e m e n t s are
considerable (Fig. 268).

Example 78: Determine the depth of embedment of a cantilever sheet pile wall installed for the purpose
of stabilising an excavation, 3 m in depth, in a homogeneous clay deposit with a unit weight of 16 kN/
3
m and an undrained shear strength of cu = ( 15 + 2z) kPa, where z, in m, is the depth below ground surface.
The possibility of water pressure against the wall shall be considered on the assumption that the
groundwater level coincides with the ground surface. The adhesion ca = 0.5c M. Outside the excavation,
2
a load of 10 k N / m has to be considered.
376 Excavation

1 / 2
5o/iii//i: The active earth p r e s s u r e ^ = 1 6 z - 2 ( 1 5 + 2>(1 + 0 . 5 ) + 1 0 = 1 L l z - 2 6 . 7 kPa. The active
earth pressure is nil to a depth of = 26.7/1 LI = 2.4 m
The water pressure uw increases from zero at the ground surface to 30 kPa at the bottom of excavation.
Below the bottom of excavation it dominates over the active earth pressure to a depth obtained by the
relation 11. lz - 26.7 = 30, whence = 5.11 m.
1/2
The passive earth pressurep p = 16(z - 3) + 2(15 + 2z>(l + 0 . 5 ) ] = (20.9z - 11.3) kPa.
The resulting earth and water pressure at, and below, the depth of excavation is:
pp-uw = 20.9z - 41.3 kPa to a depth of = 5.11 m
P p - ~ (9.8z + 15.4) kPa from = 5.11 m downwards.
2
We find Px = 103 /2 = 45.0 kN/m
Assuming the depth below the bottom of excavation to the centre of rotation to be equal to zr we find:
2 2 2
P 2 = 21.4z r +20.9-2.11 /2 + 44.1(z r - 2.11) + 9.8(z r - 2.11) /2 = 4 . 9 z r + 44.8z r - 2 4 . 6
Moment equilibrium around the rotation centre yields:
2 2 3
45(z r + l ) = 21.4-z, /2 + 44.L[(2.11/2)-(z r - 2.11-2/3) + (z, - 2.11) /2 + 9.8(z r - 2 . 1 1 ) / 6
or, in rewritten form:
3 2
zr + 13.8 z r - 42.7 z r = 16.8
whence zr - 2.90 m.
2
2 = 4.9-2.09 + 44.8-2.9 - 24.6 = 146.6 kN/m
P3 = 146.6 - 45.0 = 101.6 kN/m
Choosing F = 1.5 we find:
101.6/(D-2.09) = [16-5.90+10 + 2(15 + 5 ^
whence D = 4.2 m (= 1.45z r)
The sheet pile wall should be installed to a depth of 3.0 + 4.2 = 7.2 m

2
q = 10 k N / m

4. B R A C E D EXCAVATIONS

4.1 Introductory remarks

Sheet pile walls r e p r e s e n t the m o s t c o m m o n t y p e of retaining structure for t h e support of


the sides of d e e p e x c a v a t i o n s . O t h e r types of retaining structures, often f o r m i n g an
integral part of the structure, are d i a p h r a g m walls and secant pile w a l l s .
D i a p h r a g m walls are carried out in sections leaving intervals w h i c h are later filled in
(Figs. 2 7 0 - 2 7 1 ) a n d j o i n t e d with the p r e v i o u s l y built sections.
S e c a n t p i l e w a l l s are built in a similar w a y w i t h piles first installed w i t h a s p a c i n g less
than t h e pile d i a m e t e r a n d then c o m p l e t e d w i t h i n t e r m e d i a t e piles i n c i s e d in t h e piles
Excavation 377

Fig. 268. Failure of cantilever sheet pile wall one day after excavation to full depth had been terminated.

previously installed (cf. Fig. 1 6 5 - 1 6 6 , p. 234). Alternatively, a small e n o u g h space is left


between the piles for the soil to be retained due to arching. To safeguard against long-term
effects, the open space is generally covered by shotcrete (Fig. 271).
According to the so-called Berlinoise System, -piles driven into the soil serve as a
support for timber lagging placed during excavation.
A most important d e v e l o p m e n t in respect of braced excavations is the gradual change
in construction technique, from the use of internal propping which places obstacles to the
construction w o r k inside the excavation, to the use of tiebacks which leave the excavation
completely open and do not interfere with the construction work.
In order to limit time-dependent m o v e m e n t s the top row of braces should b e placed at
a high level, preferably not deeper than 2 m below the ground surface. Prestressing of the
braces is also an effective m e a n s of reducing lateral m o v e m e n t s of the sheet piles.
T h e p r o b l e m s connected with propped walls are s o m e w h a t different from those
connected with tied-back walls. Props are placed horizontally (in narrow excavations) or
with an u p w a r d inclination (in wide excavations). T h e prop force resultant therefore acts
on the wall either in a horizontal direction or has an u p w a r d compressive force
component. Tiebacks are usually inclined d o w n w a r d s and pretensioned, giving rise to a
downward compressive force c o m p o n e n t on the wall.
T h e m a g n i t u d e and distribution of earth pressure are very m u c h d e p e n d e nt on the
execution of the propping or anchoring system and of the excavation. It is, therefore, of
utmost importance that the execution does not deviate from what has been p r e s u m e d in
378 Excavation

Fig. 269. Schematic picture of the construction of diaphragm wall. In a second step of construction, the
diaphragm wall will be completed and the joints between the panels sealed.

Fig. 270. Diapragm wall constructed as part of the basement walls in a shopping centre. The wall, which
is made water-tight, is carried down to bed-rock at about 10m depth below groundwater and is anchored
by tiebacks.
Excavation 379

Fig. 271. Pile wall, anchored by tiebacks. Open space between the piles covered by shotcrete.

the design. M o s t of the cases w h e r e things go wrong depend on lack of contact between
the designer and the contractor or lack of efficient control.
A wealth of information regarding the analysis and design of anchored structures is
given in the textbook by H a n n a (1982). T h e problems encountered in connection with
deep excavations with side support are related to the earth pressure distribution, as
mentioned above, and to the risk of b o t t o m h e a v e in soft cohesive soils and the risk of
piping in cohesionless soils.
For the sake of simplicity, "anchored w a l l s " will b e used in the following as a term for
both anchored and propped (strutted) walls.

4.2 Walls braced at one level

T h e earth pressure distribution is generally calculated on the assumption that there is a


Rankine-type z o n e failure and no friction at the interface b e t w e e n the wall and soil (see
Retaining walls, Section 2.4). This implies that the wall is subjected to active earth
pressure a b o v e the b o t t o m of excavation and to the c o m b i n e d action of active and passive
earth pressures below the b o t t o m of excavation. T h e required depth of the wall is
determined by the conditions of equilibrium.

(i) Cohesionless soil. For a cohesionless soil the depth of the wall is determined in order that
the net passive earth pressure is at least 1.3 times the earth pressure required for rotational
380 Excavation

Fig. 272. Earth pressure distribution applied in the design of walls with one row of anchors in
cohesionless soil (left) and cohesive soil. (Sahlstrm and Stille, 1979).

equilibrium. T h e net passive earth pressure applied in the analyst s of rotational equilibrium
is assumed to be distributed as shown in Fig. 272 (Terzaghi, 1943) which yields:

zp = 4 , ( l - / r r ^ ) (474)

Pmax = {Kp -KQ)Ydp{\ - yJ\-F~ ) (475)


x

where zp, dp and pmax according to Fig. 272,


Fs = factor of safety (> 1.3).
Choosing Fs = 1.3, w e have zp = 0.52dp and / ? m a x = 0.52(A^ - Ka)Ydp.
T h e required depth D - b + dp can now be found from the rotational equilibrium
condition around point A i n Fig. 272. Finally, the force equilibrium condition yields the
value of Qe.
D u e to arching effects in non-cohesive soil, the m a x i m u m m o m e n t calculated for a
flexible wall (such as sheet pile walls) on the basis of the design earth pressure will be
too large. T h e m o m e n t reduction is dependent on the rigidity of the wall and the soil
characteristics.
In Fig. 2 7 3 , the reduction factor is given as a function of the density of the soil and
the flexibility p m of a sheet pile wall (Rowe, 1952; Janbu et ai, 1956), defined by the
correlation:
^50 100 200 500 1000 2000 5000

Flexibility of sheet pile wall pm (m/MPa)

Fig. 273. Diagram for determination of the moment reduction factor to be applied in the design of
sheet pile walls with one row of props/anchors.

where Lshp = H + Dn = m i n i m u m length of sheet piles for equilibrium,


Eshp - m o d u l u s of elasticity of the sheet piles,
Ishp = m o m e n t of inertia per length unit of the sheet pile wall.
T h e design m o m e n t M d e s is then given b y the relation:

A 4 7 7
Ancles = A * m * m ax ( )

Example 79: To realise a 5 m deep excavation in sand with an internal angle of friction of 30 a sheet
3 3
pile wall has to be installed. The unit weight of the sand is 7= 18 k N / m above, and / = 11 k N / m below,
the groundwater table. Determine the depth to which the sheet pile wall has to be installed if the
groundwater level in the sand is at 5 m depth and the wall is propped at one level, 1.5 m below the ground
2
surface. A load of 20 k N / m is placed on the ground outside the wall. Possible soil/wall friction can be
neglected.

Solution: For a soil with '= 30, we have the earth pressure coefficients Ka = 1/3 and Kp = 3.0. Thus,
the active earth pressure above the bottom of excavation follows the relation:
pa = (20+ 18z)/3 kPa
and the net pressure (active minus passive pressure) below the bottom of excavation:
pnei = 1(20 + 185)/3 + ( - 5X1/3 - 3.0)11 = 36.7 - 29.3(z - 5) kPa
The net pressure is zero for = 6.25 m, i.e. 1.25 m below the bottom of excavation. Choosing a factor
of safety Fs= 1.3 we have:
zp = 0.52dp
a nd 5 2 3
Pmax = - ( - 1 / 3 ) 1 1 ^ = 1 5 . 3 ^ kPa
The earth pressure distribution thus obtained is shown in the figure below.
Moment equilibrium around A requires:
2 2
(20/3>5(5/2 - 1.5) + (185 /6)(52/3 - 1.5) +(1.25 36.7/2)(3.5 + 1.25/3) =
= 0.52d?p.15.3dfp(3.5 + 1.25 + 2-0.52^3) + 1 5 . 3 ^ 0 . 4 8 ^ ( 3 . 5 +1.25 + 0 . 5 2 ^ + 0 . 4 8 ^ / 2 )
382 Excavation

2
q = 20 k N / m

which rewritten becomes:


^ + 8 . 7 1 ^ 2 = 33.9
whence
dp = 1.80 m
D = 1.80+ 1.25 = 3.05 m
The required depth of the sheet pile wall is equal to 5.0 + 3.05 = 8.1 m.
From the condition of force equilibrium we have:
Qe = (6.7 + 36.7)5/2 + 36.7-1.25/2 - 0.94-27.5/2 - 27.5(1.8 - 0.94) - 95 kN/m

(ii) Cohesive soil In the case of cohesive soil, disturbance effects due to pile installation
and/or installation of sheet pile walls m a y cause a considerable reducti on of the undrained
shear strength of the soil (Fig. 274). This entails an increase in the active earth pressure
and a decrease in the passive earth pressure as well as an increased risk of bottom heave.
T h e p h e n o m e n o n of b o t t o m h e a v e resembles base failure of a footing or a deep
foundation in clay and can b e treated correspondingly by considering the pressure relief
at the b o t t o m of excavation as a negative load (Fig. 275). B o t t o m h e a v e will occur when:

+ Ncbcu + 2cwD/(B + L) (478)

T h e factor of safety against b o t t o m h e a v e can thus b e expressed by the relation:

Ncbcu+2cw/(b + L)
(479)
+q

Normally, the term 2cJ(B + L) cannot be taken into account (see Example 73), whence:

Net
(480)
+q

The bottom heave factor Ncb for trench excavations can be determined in a way similar
to that of the bearing capacity factor for deep foundations. Assuming D to represent the
Excavation 383

Fig. 274. Example of disturbance effects caused by pile driving and installation of sheet pile wall.

r
_JII I I I I, I I I

\
AM = t/f = /ft -./// ~ rm
V \

\
I

Fig. 275. Bottom heave of deep excavations in cohesive soils is analysed in a similar way as bottom
failure of deep foundations.
384 Excavation

1
BIL = 1

BIL = 0.5

/L = 0

0 1 2 3 4 5

Depth to width ratio (D + H)IB

Fig. 276. Diagram for the determination of Ncb {cf. Eq. 480).

depth below the b o t t o m of excavation to the base of the supporting wall, the stability
factor Ncb for an excavation with length L, width and depth H, and for (H + D)IB < 4,
can be expressed by the approximate relation (Fig. 276):

1 H+D H+D 707


^ / 7= ( 1 + 0 . 2 - ) { 5 . 1 4 + - [ - ( 8 - ^ )
0
} (481)
L j

For values of (H + D)/B > 4, the value obtained for (H + D)/B = 4 is used, i.e. Ncb - 7.5
for BIL = 0 and Nch = 9.0 for BIL = 1.
T h e shear strength cu in Eqs. ( 4 8 0 - 4 8 1 ) should be taken as the m e a n of the shear
strength values observed to a depth below the b o t t o m of excavation of 0.7. Disturbance
effects due to excavation, installation of the sheet pile wall and/or pile driving should b e
considered (cf Fig. 274).
For w i d e excavations in clay the overall stability should be checked, for e x a m p l e by
circular-cylindrical slide surface analysis.
T h e net passive earth pressure below the b o t t o m of excavation that can be allowed in
the design will b e restricted by the risk of b o t t o m heave. A s s u m i n g a factor of safety of
Fs against wall rotation around the anchor (strut) level, the m a x i m u m net passive earth
pressure is determined by the relation:

Pmax = ^ T T - ( 9 + ? ) (482)
r
Excavation 385

w h e r e Ncb = the stability factor (In the case of inclined anchors, Ncb is governed b y
the vertical stability of the sheet pile wall, see p . 3 8 9 ) ,
cu = undrained shear strength.
A n allowance should b e m a d e for water pressure against the wall on the assumption that
the g r o u n d w a t e r level coincides with t h e ground surface (as s h o w n in Fig. 2 7 2 ) .
Again the depth D is determined by the rotational equilibrium around point A in Fig. 272.
If the anchors (struts) are prestressed, the total earth pressure against t h e wall and the
anchor (strut) force are determined b y the relation:

Qiot = 0.8Qe + 0AQpr (483)

w h e r e Qe = force required to b a l a n c e the earth pressure,


Qpr = total prestressing force,
or b y the relation:
tot = p r (484)

The former relation is used when 0.4 < QprIPa < 1.35 and the latter relation when QprJPa >
1.35.
T h e additional earth pressure resultant Qioi - Qe due to prestressing can b e a s s u m e d to
act at the anchor (prop) level.

Example 80: A trench, supported by a sheet pile wall, shall be made to a depth of 3.5 m in clay with
3
a unit weight of 16 k N / m and an undrained shear strength cu = 15 kPa. The width of the excavation is
2
2 m and the length 10 m. The load on the ground surface outside the excavation is 10 k N / m . The wall
is propped at a depth of 1.5 m. Determine the required depth D below the bottom of excavation to which
the sheet pile wall must be installed if the required factor of safety is 1.5. Determine also the strut force.

Solution: The active earth pressurep a = \6z - 2-15 + 10 = \6z - 20 kPa. Water pressure uw = lOz kPa
dominates over the active earth pressure to a depth of 20/6 = 3.3 m. The stability factor Ncb = 1.04-7.5.
With a factor of safety against wall rotation of 1.5, the maximum allowable, stabilising earth pressure
below the bottom of excavation becomes equal to:
Pmax = 7.8-15/1.5 - (16-3.5 + 10) = 12 kPa
Moment equilibrium around the anchor level yields:
2
10-(3.5 /2)(3.5-2/3 - 1.5) = 12D(D/2 + 2.0)
2
10 k N / m

1.5 m
U
\ u

^ =^
D

Pmax
386 Excavation

or, in rewritten form:


2
D = 4.5
whence D = 2.1 m and H + D = 5.2 m
The heave stability factor becomes equal to:
7
Ncb = (1 + 0.22/10){5.14 + [(5.6/2)(8 - 5.6/2)]- /3} = 7.6
The safety factor against bottom heave becomes Fs = 7.6 15/(16-3.5 + 10) = 1.73
2
The strut force Qe = 10-3.5 /2 - 12-2.1 = 36 kN/m

4.3 Walls braced at several levels

W h e n the wall is anchored (propped) at several levels a redistribution of the earth


pressure will take place depending on the execution of the anchors and excavation and
on the soil characteristics (see Tschebotarioff, 1951). T h e redistribution suggested by
Terzaghi & Peck (1967), or minor modifications of it, is frequently used in the design.
In the design method proposed by Sahlstrm & Stille (1979), the earth pressure is
calculated according to Rankine, assuming no friction in the contact surface between
sheet piles and soil. As regards cohesionless soil, the depth to which the sheet pile must
be driven is governed by the depth of excavation below the lowest anchor level and by
the possible risk of piping (cf. p. 357). In the case of cohesive soil, the corresponding
depth is governed by the risk of bottom h e a v e and by possible hydraulic uplift. T h e
stability of the base of anchored walls is acceptable w h e n the factor of safety against
bottom h e a v e is a b o v e 1.5. If this is not the case the sheet piles ought to be driven into
firm b o t t o m layers. T h e risk of hydraulic uplift in layered clay m u s t always b e considered,
i.e. the total overburden pressure at the tip level of the cut-off wall must exceed the uplift
exerted by the pore water pressure.
For the calculation of anchor (prop) loads, the earth pressures are redistributed as
shown in Figs. 2 7 7 - 2 7 8 , increasing linearly from 0 at the ground level to a constant value
Pi at depth 0.2H and below. T h e p{ value is determined from the relation:

Pi

Fig. 277. Design earth pressure and its redistribution in cohesionless soils. The value of Pa included in
the redistributed earth pressure is chosen as the Rankine earth pressure above point D. (Sahlstrm and
Stille, 1979).
Excavation 387

tot
0.9H+d

Fig. 278. Design earth pressure and its redistribution in cohesive soils. The value of included in the
redistributed earth pressure is chosen as the Rankine earth pressure above point D. (Sahlstrm and Stille,
1979).

tot
(485)
0.9//+ d

w h e r e H and d according to Figs. 277 and 2 7 8 ,


Pa = total earth pressure (including water pressure),
Qtot = 0.SPa + 0AQpr>Pa,
or: tot = QPr
T h e former value of Qlot is used w h e n 0.4 < Qpr/Pa < 1.35 and the latter w h e n QprJPa
> 1.35 ( c / E q s . 4 8 3 ^ 8 4 ) .
A s m e n t i o n e d before, Sahlstrm and Stille (1979) p r o p o s e that an allowance b e m a d e
for the possibility of water pressure in the top part of a clay deposit with a groundwater
level coinciding with the ground surface (as shown in Fig. 272). This m a y s e e m odd since
the total weight of the clay is included in the calculated active earth pressure. However,
the pressure acting against the wall can never b e less than the water pressure.
T h e earth pressure b e l o w the b o t t o m of excavation represents the difference b e t w e e n
active and passive earth pressures. For excavations in clay, w h e r e Ncb < ( + q)/cu, the
base of the wall should b e carried d o w n to firm soil layers (Fig. 270).

Example 81: A sheet pile wall shall be installed for the purpose of stabilising a 5 m deep excavation
3
in a 9 m thick clay layer underlain of sand. The clay has a unit weight of 17 k N / m and an undrained
3
shear strength of 20 kPa. The sand has a saturated unit weight of 21 k N / m and an internal angle of
friction of 38. The groundwater level in the sand is kept 3.5 m below the ground surface. The width
of the excavation is 4 m and the length 20 m. The wall shall be supported with horizontal struts at depths
388 Excavation

1 m, 3.0 m and 5 m below ground surface. A prestress load of Q =0JPa shall be applied. External load
2
outside the excavation q = 10 k N / m . Determine the required length of the sheet pile wall and the strut
loads. A partial factor of safety of 1.15 should be applied on the angle of friction in the sand.

Solution: The active earth pressure is equal to Pa = llz + 10 - 2-20 = 17z - 3 0 kPa, where denotes the
depth below ground surface. The water pressure uw 10z kPa dominates to a depth of 30/7 = 4.3 m.
The total pressure above the bottom of excavation is equal to:
2
Pa = 104.3 /2 + 17(4.3 + 5.0)/2 - 30 = 141.5 kN/m
whence g t ot = (0.8 + 0.4-0.7)141.5 = 152.8 kN/m
Assuming that the sheet pile wall is installed to a depth of D = 4 m, the safety factor against bottom
heave Fs becomes equal to 1.17 and against hydraulic uplift 17-4/55 = 1.24.
The sheet pile wall has to be driven down into the sand.
The earth pressure in the clay below the bottom of excavation is equal to
pa-pp= 1 7 z - 3 0 - [ 1 7 ( z - 5 ) + 40] = 15 kPa
The design angle of friction in the sand is '= arctan(tan3871.15) = 34
2
The earth pressure coefficients in the sand are Ka = tan (45 - 3 4 7 2 ) = 0.28 and = 1/0.28 = 3.57.
In the sand we thus have:
/ 7 - / 7 = 3 . 5 7 [ 4 1 7 - 5 . 5 1 ( ) + 1 1 ( - 9 ) ] - 0 . 2 8 [ 9 1 7 + 1 0 - 5 . 5 - 1 0 + l l ( z - 9 ) ] = 16.2 + 3.29(z-9)kPa
Thus in our case 2c = 4 m and d = c/2 = 2 m (see Fig. 278)
The redistributed earth pressure pt = 152.8/(0.9-5 + 2) = 23.5 kPa
whence the prop loads: Qx = 23.5(1.0 + 0.5) = 35.3 kN/m
Q2 = 23.5(1.0 + 1.0) = 47.0 kN/m
Q3 = 23.5(2.0 + 1.0) = 70.5 kN/m
The length of the wall is determined by the relation:
2
2-15 = 16.2(D - 4) + 3.29(D - 4 ) / 2
whence D = 5.6 m
The total length required of the sheet pile wall is 10.6 m.
Excavation 389

4.4 Influence of inclined tiebacks

W h e n the sheet pile wall is anchored by m e a n s of inclined tiebacks the vertical stability
of the sheet piles m a y b e adventured by the vertical c o m p o n e n t of the anchor loads. This
p r o b l e m has b e e n dealt with by Stille (1976).
Let us a s s u m e that the anchor inclination is (Fig. 279). T h e vertical c o m p o n e n t of the
anchor loads will b e counteracted by wall adhesion. Taking into account the influence on
the earth pressure exerted by wall adhesion, the normal c o m p o n e n t s of the active and
passive earth pressures are given by the a p p r o x i m a t e relations, given by Eqs. 3 6 3 - 3 6 4 :

Pa = + 4 - NaCu

Pp = JZ+q + Npcu

w h e r e (Janbu et al, 1956)

JV = 2 v / l + 2 r / 3
Np = 2y/l+2rp/3
r = cjcu

T h e wall adhesion parameter r can b e positive or negative d e p e n d i n g on the direction


of the relative displacement between soil and sheet pile wall. T h u s , w h e n the soil is
m o v i n g d o w n w a r d s in relation to the wall, r b e c o m e s positive on the active side of the
wall and negative on the passive side, and vice versa.
A s s u m i n g that the forces are acting on the sheet pile wall as s h o w n in Fig. 2 7 9 , the
following equilibrium conditions are obtained:

(Pp-Pa)une-(Cp-Ca)cose =0 (486)

Fig. 279. Assumptions behind the analysis of the stability of sheet pile walls with inclined anchors.
Cohesive soil.
390 Excavation

Rsme-(Cp- Ca) = 0 (487)

where Pa = 0.5[y(H + D) - Nacu](H + D- //,),


Pp = 0.5yD+NpCuD,
Ca = racu(H + D-H^,
C rc D
p = pu >
Hi = NacJy.

Introducing the stability n u m b e r Ncb = yH/cu and the values of Pp, Pa, Cp,Ca and / / , into
Eq.(486), Ncb is obtained by solving the following equation of the second degree:

- = (488)

where
Na{\+HID)+Np + [rp-ra{\+HID)]cote
1+0.5H/D

Na(2racote-Na)
B =
1+2D/H

T h e anchor load R can b e calculated from Eq.(487):

Hc rnD D Na
R = ra(l+-- )] (489)

// V2/3

-i 0.5
// / / /I72/3
/1/3
I Vo.5
^Vl/3
N* 4 r y I S ) J .
/ - <P
= 0

Is*
D D 2
~ =1 - = 0.5
H~3

0 1
RcO
Hcu

Fig. 280. Values of stability number Ncb as a function of anchor load and wall adhesion for anchor
inclination = 45.
Excavation 391

B y a s s u m i n g t h e v a l u e of rp, Ncb and ra can b e solved b y an iterative calculation


p r o c e d u r e . T h e stability of the sheet pile w a l l is j e o p a r d i s e d w h e n ra > - 1.
T h e influence of a n c h o r inclination on t h e stability n u m b e r is s h o w n in Fig. 2 8 0 . [The
values p r e s e n t e d in Fig. 2 8 0 differ from t h o s e p r e s e n t e d b y Stille for the r e a s o n that in
l/2
his analysis the Na a n d Np v a l u e s w e r e b a s e d on the relations Na = 2(l + ra) and Np =
2 1 / 2
2(1 + rp)M instead of Na = 2(1 + 2ra/3)^ and Np = 2(1 + 2 r / / 3 ) ].

Example 82: For the side support of a 4 m deep excavation in a clay layer with a unit volume weight
3
of 16 k N / m and an undrained shear strengh of 20 kPa a sheet pile wall is installed to a depth of 7.3 m.
The wall is anchored at a depth of 2 m by tiebacks with an inclination of 45. The anchor force is 110
2
kN/m. External load outside the excavation q = 10 kN/m . Determine the factor of safety on the assumption
that we can allow a value of rp = 0.5. Determine also the earth pressure distribution and the factor of
safety against rotation around the anchor level.

Solution: W e have Rcos6l(cuH) = 110-cos457(20-4) = 0.97 and HID = 4/3.3 = 1.2.


For a value of HID = 1, rp = 0.5 and Rcos6l(cuH) = 1, we have Ncb ~ 4.3.
Introducing this value into Eq. (489) we find:
1/2
110 = (4-20/sin45)[0.5-3.3/4 - ra(\ + 3.3/4 - 2 ( l + 2 r t / 3 ) / 4 . 3 ]
whence ra = - 0.39
Inserting this value of ra into Eq. (489) we find Ncb = 4.28.
Another iteration procedure yields the same values of Ncb and ra.
The safety factor Fs = 4.28-20/(16-4 + 10) = 1.16
1 72
The active earth pressurep a = I6z + 10- 2-20(1 - 2-0.39/3) = I6z - 24.4 kPa. Water pressure uw
= 9.8z kPa dominates over active earth pressure to a depth of 24.4/6.2 = 3.9 m.
1 /2
The passive earth pressure pp = 16(z - 4) + 2-20(1 + 2 - 0 . 5 / 3 ) = 16(z - 4) + 46.2 kPa.
Net earth pressure at the bottom of excavation:
p p - P a = 46.2 - 16-4 + 24.4 = 4.9 kPa.
Rotational stability yields:
2
4.9-3.3(3.3/2 + 2) = 9.8(4 /2)(2-4/3 - 2)FS
whence Fs= 1.13.

2
10 k N / m
392 Excavation

4.5 Design of wale beams and anchors

(i) Design of wale beams. In the design of the wale b e a m s , consideration m u s t b e paid to
the possibility of anchor failure. Such anchor failure will cause a redistribution of the
earth pressure against the sheet pile wall and induce a load increase in the anchors nearby.
T h e consequence of anchor failures has been investigated by Stille (1976). In 13 cases
of simulated anchor failures, the load increase in adjacent anchors was less than 2 5 % . In
these cases the earth pressure, due to prestressing of the anchors, exceeded the Rankine
earth pressure. Generally, the load increase was very small, particularly w h e n the anchor
failure took place in the lower anchor rows.
Obviously, considering the insignificant effect of anchor failure and normal safety
requirements, the possibility of a d o m i n o effect, leading to a total collapse of the sheet
pile wall, seems negligible. However, the possibility of anchor failure cannot be
neglected and ought to be considered in the design of the w a l e b e a m . Sahlstrm and Stille
(1979) r e c o m m e n d that the design of the wale b e a m b e based on a modified version of
the limiting creep stress method. Assuming that the load acting on the w a l e b e a m before
anchor failure is q, the m a x i m u m allowable bending m o m e n t , M m ax can be determined
by the relation:

Mm ax = ^-<1.5aa ll Wx (490)
lo

w h e r e / = double the anchor spacing,


a a l l = m a x i m u m allowable wale b e a m stress,
Wx = bending resistance of the w a l e b e a m in the load direction.

At the outer ends of the wale beam, the cantilever length can be a s s u m e d equal to lyj 2
w h e r e / as above.
In the case of very rigid wale b e a m s and w e a k anchors, special analyses of the m o m e n t
distribution m a y be required.

(ii) Design of ground anchors. Very complete information regarding ground anchors is
provided in the text b o o k by H a n n a (1982). Further details about different anchoring
systems and installation m e t h o d s used throughout the world, as well as detailed aspects
of their design, can also b e found there.
T h e ground anchors can either b e fixed in bed-rock or in soil. R o c k anchors are most
c o m m o n l y used w h e n the depth to bedrock is within reasonable limits. In this case, a
cylindrical hole is generally bored into the rock and filled with grout or s o m e other fixing
agent in which the anchor tendon can b e fixed.
T h e load that can b e carried by rock anchors will depend on the grouted length and the
b o n d resistance between tendon and grout or rock and grout. T h e distribution of b o n d
along the anchor before yielding can b e assumed to follow the relation:
Excavation 393

/// a . //c ^ 'V = s/r ^ //| ^ 7 ^ m j

/yan0'|jjj| / > l f t al n 0 '

\W\

2&

Fig. 281. Anchoring of retaining wall by means of a horizontal tie rod fixed to a vertical plate.

= T0exp(-Ax/d) (491)

where = b o n d stress at distance from the top of the fixed anchor,


0 = b o n d stress at the top of the fixed anchor,
d = tendon diameter (or, alternatively, rock hole diameter),
A = constant relating the b o n d stress to axial stress in the tendon.
A s s u m i n g that the bond stress can be expressed as a function of the relative displacement
ux between the tendon and the bond and the relative displacement m o d u l u s K, according
to the relation = Kux, then:

A = 2y/KdlE (492)

w h e r e = the elastic m o d u l u s of the tendon.


Integrating over the total length L of the fixed anchor w e find:

2
nd T(\
P= - [ 1 - e x p ( - A L Id)} (493)
A

We thus realise that the higher the displacement m o d u l u s the shorter the length taking
part in carrying the load and the higher the bond stresses. Obviously, yielding will take
place in the b o n d in the upper part of the anchor already at an early stage of the loading
and progress deeper d o w n with increasing load.
Soil anchors h a v e to b e formed with due regard to the soil characteristics with the aim
to producing a reliable and cost-effective solution.
T h e m o s t simple type of soil anchors consist of vertical plates buried in the soil and
fixed to the sheet pile wall by horizontal tie rods. T h e anchoring force is determined as
the difference b e t w e e n passive and active earth pressures acting against the plate. For
superficial anchor plates the vertical stability has to checked b y the relation (Fig. 281):
394 Excavation

W + 2 tm5a + 1 tan0'> P2p tmp + Plp tan0' (494)

where W = weight of the plate and the overlying soil,


P2a = normal component of active earth pressure against the plate,
Pla = normal component of active earth pressure against the overlying soil section,
P2p = normal component of passive earth pressure against the plate,
Plp = normal component of passive earth pressure against the overlying soil section,
and are the angles of friction at the active and passive plate/soil interfaces.
The upward force on the passive side can be reduced by providing the passive plate
surface with a friction-reducing coating.
The distance between the sheet pile wall and the anchor plate is determined by the
condition that the passive failure zone in front of the anchor plate and the active failure
zone behind the sheet pile wall should not intersect. The overall stability also has to be
checked.
Nowadays, these types of anchor are more or less completely replaced by inclined
tension anchors. In cohesionless soil, the anchoring zone is generally enlarged by
injection of grout that permeates into the granular material, but the form and homogeneity
of the grouted zone thus achieved is difficult to control. In order to produce a well and
easily controlled anchor body in any type of soil, so-called expander anchors (Fig. 282)
have been developed and used with great success under various soil conditions. T h e
expander body consists of a steel container which is folded into the shape of a cylinder
with square cross section. The expander body is attached to a steel tube (of the same kind
as those used as micro-piles) and is then inserted into the soil to the required depth, either
by hammering it into the soil or by preboring. After installation, the body is expanded by
grout injection through the steel tube. By recording the volume increase vs. grouting
pressure, information is obtained about the soil conditions and the bearing capacity of the
anchor can be estimated according to the principles given for the pressuremeter test.
The expander bodies now on the market, in their folded shape before installation, have
dimensions varying from 1 m in length and 60 m m in width (EB 300) to 3 m in length
and 110 m m in width (EB 800). The expanded diameter of E B 300 is around 300 m m and
of E B 800 around 800 m m .
The first type of inclined anchors in clay, used in practice, consisted of H E A steel beams
driven with 45 inclination deep into the clay and fixed to the sheet pile wall by tendons
(Fig. 283). Nowadays, expander anchors form a cost-effective solution also in the case
of cohesive soils.
The grout injection of the expander anchor brings to mind the pressuremeter test.
Actually, results of pressuremeter tests form an excellent basis for determination of the
bearing capacity and the deformation behaviour in loading. As shown by Sellgren (1991),
the displacement s for the tension load can be calculated by the relation:
Excavation 395

Fig. 282. Expander bodies EB 800 (folded and expanded) and EB 300 (expanded).

s= - (495)
i-Q/Qf

w h e r e is the initial slope of the load vs. displacement curve,


Qf is the failure load in tension.
T h e initial slope depends on one hand on the extension of the tendon and on the other
on the displacement of the anchor b o d y itself. Sellgren (1991) has s h o w n that there is a
good correlation b e t w e e n calculated and measured oad/displacement behaviour if be
determined by the relation:

Fig. 283. HEA steel beam, installed with an inclination of 45 to the horizontal, to serve as an anchor
for a sheet pile wall in soft, high-plasticity clay.
396 Excavation

= ^ + ^ (496)
EtAt 9DaEs

w h e r e Lt = the free length of the tendon,


Et = the elastic m o d u l u s of the tendon,
At = the cross-sectional area of the tendon,
Da = the diameter of the anchor body,
Es = the elastic m o d u l u s of the soil.
T h e most reliable m e t h o d of determining the bearing capacity of the e x p a n d e r anchor
is based of the injection pressure used during installation. T h e bearing capacity is
s o m e w h a t higher in compression than in tension (Fig. 284).

Fig. 284. Ultimate load for various types of expander bodies determined by the results obtained in
expansion of the bodies (Sellgren, 1992).
So/7 improvement 397

SOIL I M P R O V E M E N T

1. I N T R O D U C T I O N

T h e a i m of soil i m p r o v e m e n t is to bring about a condition w h e r e the g e o m e c h a n i c a l


properties of the subsoil b e c o m e good e n o u g h to instigate cost-effective solutions to the
foundation p r o b l e m s and m i n i m i s e m a i n t e n a n c e costs. Soil i m p r o v e m e n t can be a m e a n s
to avoid expensive piling and to reduce the risk of d a m a g e to adjacent buildings during
foundation w o r k s .
In the case of cohesive soils, soil i m p r o v e m e n t is most c o m m o n l y used to reduce, or
to eliminate, l o n g - t e r m settlement which otherwise can create serious p r o b l e m s . For
e x a m p l e , in the architectural planning of buildings, the ground level is often raised and
the groundwater level often lowered by drainage through s e w e r a g e and water pipes. In
cases w h e r e the subsoil consists of compressible soil this m a y cause serious p r o b l e m s in
the connection b e t w e e n building and environs (Fig. 285) unless the deformation prop-
erties of the soil h a v e been i m p r o v e d in advance.
In urbanised areas, land with g o o d foundation properties has been m o r e or less used
up for building purposes. Therefore, soil that w a s originally considered too bad has
started being utilised. T h e r e is a tendency to place, first of all, industrial activities to such

Fig. 285. Large settlement of the ground encompassing buildings on piles brings about access problems
and may cause breakage of connecting sewerage and water pipes.
398 So/7 improvement

bad areas whereas dwellings are mostly concentrated to areas with better foundation
properties. T h e problems encountered will certainly increase with increasing need of
ground for building purposes. Obviously, soil i m p r o v e m e n t techniques will b e c o m e
increasingly important in the future.

(i) Improvement by compaction. T h e most important compaction w o r k s are carried out


in connection with the construction of earth dams which due to their i m m e n s e heights (up
to 300 m ) and the catastrophic consequences of failure require extremely careful
w o r k m a n s h i p and control. T h e compaction technique is well established and detailed
instructions for the compaction process exist in most countries.
N o w a d a y s , compaction of fill and natural soil has also b e c o m e quite c o m m o n in
connection with industrial building and house-building activities in order to find cost-
effective foundation solutions. Recently, the need of land reclamation for industrial and
harbour development has urged the development of deep-compaction technique onward
and compaction can now be carried out to depths of 3 0 - 4 0 m below the compaction
surface.
Compaction has two aims: Firstly it should bring about a densification of the soil and
secondly it should crush sharp edges and sharp contact points between the soil grains.
Crushing leads to an increase in the size of the contact areas and, consequently, to a
decrease in contact stresses between the grains which in turn yields a lower compressibility.
T h e crushed material also fills the voids in the soil and results, as does the reorientation
of the grains, in a v o l u m e decrease.
Densification by reorientation of grains is best achieved by vibrations while crushing
is best achieved by impact. Reorientation of grains i s the foremost reason for densification
in fine-grained soils. D u e to small grain size, the n u m b e r of contact points b e t w e e n the
grains per unit v o l u m e is very large (for example, in a case w h e r e the grain size is 2 m m ,
2 6
the n u m b e r of contact points per m cross-sectional area is about 1 0 ) , and the contact
stresses thus b e c o m e insignificant. In a soil with large grain size, such as boulder and
cobble soil, crushing of grains is the foremost reason for densification. This is particularly
the case for blasted rock fill. On the one hand, the n u m b e r of contact points per unit
v o l u m e is quite small (in a case w h e r e the grain size is 6 0 0 m m , the n u m b e r of contact
2
points per m cross-sectional area is about 10) and on the other hand the grains have not
been subjected to the wear of nature and time.
In water saturated soil the v o l u m e decrease achieved by compaction is always
accompanied by escape of an equally large volume of water. In soils with low permeability,
the water escape m a y take place very slowly and compaction of such soils therefore
involves a risk of liquefaction p h e n o m e n a occurring. Even if the soil is not originally
water saturated there m a y b e a risk of liquefaction provided that the v o l u m e decrease
during compaction is large e n o u g h for the soil to pass into a state of full water saturation.
T h e choice of compaction m e t h o d and e q u i p m e n t obviously depends on the grain size
So/7 improvement 399

distribution and the content of fines and cobbles and boulders. W h e n the content of fines
is large then the degree of water saturation is of significant importance.

(ii) Improvement by preloading. In the case of water saturated l o w - p e r m e a b l e soils, such


as clay and silt, densification b y m e a n s of compaction m e t h o d s is not possible.
Densification is a c c o m p a n i e d by water being squeezed out of the soil which m a y take
c o n s i d e r a b l e time. Therefore, i m p r o v e m e n t of l o w - p e r m e a b l e soils can only b e
accomplished by long-term stabilisation methods: preconsolidation by m e a n s of temporary
overloading, chemical treatment or electrochemical treatment. Temporary overloading
is c o m m o n l y c o m b i n e d with the installation of vertical drains in order to speed u p the
consolidation process.

(iii) Improvement by chemical and electrochemical treatment. Chemical and


electrochemical treatments a i m at creating a m o r e stable and stronger soil structure by
ion e x c h a n g e . C h e m i c a l stabilisation is generally achieved by m i x i n g soil with lime or
cement. Electrochemical stabilisation is accomplished by the aid of direct electric
current, causing electro-osmotic processes to take place. T h e water that is assembled at
the cathodes by electro-osmosis is generally p u m p e d out.

2. SHALLOW COMPACTION

T h e most c o m m o n m e t h o d of soil i m p r o v e m e n t is related to the p l a c e m e n t of fill material


in lifts which are successively compacted. T h e thickness of the layers is rather limited
(generally less than 1.5-2 m ) and the compaction technique applied in this connection is
therefore referred to as shallow compaction.
Shallow compaction m e t h o d s h a v e mainly been developed for earth d a m and subgrade
construction. Characteristic for earth d a m construction is the d e m a n d for stability, water
tightness and safety against erosion (both external and internal). T h e d e m a n d for water
tightness requires the use of fine-grained soils in the d a m core. T h e necessity of
compacting these soils in a satisfactory w a y has led to detailed n o r m s regarding the
determination of the compaction characteristics of the soil and the choice of compaction
equipment.

2.1 Optimum water content

O n e of the m o s t important compaction characteristics of fine-grained soil is the o p t i m u m


water content, i.e. the water content at w h i c h the best c o m p a c t i o n result can b e expected.
T h e o p t i m u m water content is c o m m o n l y determined in the laboratory b y m e a n s of s o m e
standardised c o m p a c t i o n procedure, the m o s t c o m m o n of w h i c h is in a g r e e m e n t with
either A A S H O ( A m e r i c a n Association of State H i g h w a y Officials) or the Standard
Proctor Test. T h e a i m of this test is to find out h o w the density of the soil achieved by a
400 Soil improvement

0 5 10 15 20 25
Water content w (%)

Fig. 286. Dry density as a function of water content at laboratory compaction. Compaction is carried
out at varying water contents in order to enable a full presentation of pd vs. w. Lines represent water
saturation for different values of particle density.

certain compaction w o r k is influenced by the water content of the soil. T h e result is


presented in a diagram w h e r e the dry density is given as a function of the water content
at compaction (Figs. 2 8 6 - 2 8 8 ) . T h e reason for choosing dry density pd instead of bulk
density p i s that it gives us an indirect m e a s u r e of the porosity (void ratio) of the soil (see
p. 14). According t o E q . (11), p. 14, the bulk density at the water content w in question
is obtained by:
p = Pd(\+w) (497)

T h e correlation b e t w e e n dry density and water content m u s t theoretically fall inside


the b o u n d a r y of water saturation ws, defined by:

w 5 = pJMPd - l/p8) (498)

Increasing flocculation Increasing dispersion

Water content w

Fig. 287. Influence of compaction energy on optimum water content and dry density. Compaction wet
of optimum gives a more parallel particle orientation than compaction dry of optimum. Large
compaction energy also leads to a more parallel particle orientation.
So/7 improvement 401

Fig. 288. Example of the influence on the optimum water content of compaction energy. Notice that
loose infilling of material at optimum water content according to laboratory compaction may result in
minimum dry density.

However, from a statistical viewpoint this is not always the case due to deviations from
the a s s u m e d value of pg (cf. Fig. 288).

Example 83: As a result of laboratory compaction, the optimum water content of the soil was found
3
equal to w o pt = 7% and the dry density equal to pd = 2.18 t/m . Determine the porosity and the degree
of saturation of the compacted soil.

3
Solution: Assuming that the particle density of the soil is pg = 2.7 t/m , we find by Eq. (11):
2.18 = 2.7(1 - n), whence =0.19 (19%)
The water content at full saturation, according to Eq. (498), is:
ws = 1/2.18 - 1/2.7 = 0.088, whence Sr = 0.07/0.088 = 0.79 (79%)

T h e o p t i m u m water content determined in the laboratory differs generally from the


o p t i m u m water content prevailing during compaction under field conditions as it is
dependent on the compaction energy. T h e heavier the compaction equipment, the lower
the o p t i m u m water content and vice versa. L o o s e infilling of soil at the o p t i m u m water
content m a y lead to the lowest possible density (Fig. 288).
402 So/7 improvement

2.2 Dry of optimum and wet of optimum

In order to achieve the results intended, the compaction m a y h a v e to b e carried out at a


water content below o p t i m u m (dry of o p t i m u m ) or a b o v e o p t i m u m (wet of o p t i m u m ) . In
the case of clay, the answer to the question whether compaction should b e carried out dry
of o p t i m u m or wet of o p t i m u m depends on the intended effect on clay structure,
permeability, compressibility and shear strength . T h e following general rules can b e
applied.
Soil structure. Flocculation tendencies increases dry of o p t i m u m while dispersion
tendencies increases wet of optimum. Material compacted dry of o p t i m u m has a low
degree of water saturation and can consequently suck m o r e water than material compacted
wet of o p t i m u m (Fig. 289). Water suction m a y cause considerable swelling if the material
contains swelling clay mineral (smectite).
Permeability. T h e permeability reaches a m i n i m u m value w h e n compaction is
carried out slightly wet of optimum. T h e permeability will u n d e r g o a long-term decrease
in soil compacted dry of optimum.
Compressibility. T h e compressibility of soil compacted dry of o p t i m u m is lower than
for soil compacted wet of o p t i m u m except at very high pressure. T h e compression curve
dry of o p t i m u m resembles that of undisturbed clay and wet of o p t i m u m that of disturbed
clay. T h e consolidation process is quicker dry of o p t i m u m .
Shear strength. T h e undrained shear strength is higher dry of o p t i m u m but m a y be
lower than wet of o p t i m u m if swelling takes place, resulting in water saturation. T h e
drained shear strength will be approximately the s a m e irrespective of whether compaction
is carried out dry or wet of optimum. Pore pressure at failure is higher wet of optimum.
Sensitivity generally b e c o m e s higher dry of optimum.
In conclusion, compaction of clay fill ought to be carried out dry of o p t i m u m to reach

Water content at compaction

Fig. 289. Material compacted dry of optimum has a low degree of water saturation and, therefore,
due to suction, is prone to swelling unless prevented.
So/Y improvement 403

high bearing capacity and small settlement while it ought to b e performed wet of o p t i m u m
to achieve low hydraulic conductivity.

2.3 Compaction equipments

T h e various e q u i p m e n t s used for shallow compaction are designed so as to achieve the


best possible effectiveness and are founded on either of the principles of static pressure,
vibration or impact. In the case of large filling w o r k s very h e a v y self-propelled or tractor-
drawn vibratory rollers are mostly used. S o m e of the m o s t c o m m o n e q u i p m e n t s in use
will b e described briefly.

(i) 'Static'compaction equipments. A m o n g the static c o m p a c t i o n e q u i p m e n t s w e h a v e


heavy-weight caterpillars, bulldozers, scrapers and trucks etc. T h e s e are not primarily
meant for c o m p a c t i o n and, therefore, no instructions are given for their u s e as compaction
tools. All the same, the u s e of bulldozers especially can give quite satisfactory compaction
results depending upon the fact that they induce high horizontal in situ stresses in the fill.
W h e n speaking about static compaction e q u i p m e n t s w e normally allude to rollers of
various t y p e s s m o o t h rollers, sheepsfoot rollers, padfoot rollers, grid rollers and
rubber-tired or pneumatic-tired rollers (Figs. 2 9 0 - 2 9 2 ) . T h e rollers can either b e self-
propelled or tractor-drawn or coupled together to form so-called t a n d e m rollers as well
as coupled together sideways.
In order to p r o d u c e as deep a compaction as possible the weight of the rollers has gradu-
ally been increased and n o w a d a y s weights of u p to 15 t for smooth steel rollers, 3 0 1 for
sheepsfoot and padfoot rollers and 50 t for rubber-tired rollers are c o m m o n . In special

Fig. 290. Padfoot rollers are recognised by feet protruding from the cylindrical steel shell of the roller.
The feet of the padfoot roller are shorter and larger than those of sheepsfoot rollers.
404 Soil improvement

Fig. 291. Rubber-tired rollers are utilised for the compaction of fine-grained soils especially clay with
low shear strength.

cases even heavier rollers are utilised and their weight can be increased by the use of
ballast.
The compaction effect of smooth rollers is achieved by high static contact pressure
induced by the line load of the roller. Sheepsfoot and padfoot rollers, and to a certain
extent also grid rollers, exert high contact pressure and kneading together. The contact
pressure under the feet is, of course, considerably higher than under the steel cylinder
itself.
Rubber-tired rollers also bring about compaction by the combined effect of static
pressure and kneading. High wheel pressure should be used to achieve deepest possible
compaction.
The depth of compaction accomplished by the use of static rollers is fairly small
(generally limited to 0.2-0.3 m) and their compaction capacity is therefore low.

(ii) Vibratory rollers. All the static rollers mentioned above also exist as vibratory
compaction equipments.
Vibratory rollers generally work in a frequency range of 2 0 - 8 0 Hz with nominal
amplitudes varying from 0.3 to 2.5 mm. The vibrations are accomplished by means of one
or several rotating eccentrics. (The nominal amplitude is obtained theoretically as the
ratio of moment of eccentric weight to mass of roller. The real amplitude at certain
frequencies can be as much as 5 0 - 1 0 0 % higher than the nominal one, depending on
resonance phenomena).
The vibratory action of the roller induces a long sequence of compression waves into
the subsoil. The result of compaction by means of vibratory rollers is dependent on
(Forssblad, 1981):
So/7 improvement 405

Fig. 292. Tractor-drawn, smooth vibratory roller.

static weight of the roller,


ratio of d r u m m a s s to frame mass,
d r u m diameter,
frequency and amplitude,
roller speed.
It seems reasonable to believe that the effectiveness of c o m p a c t i o n w o u l d reach its
m a x i m u m at r e s o n a n c e frequency. However, both centrifugal force and m o m e n t u m
increase with increasing frequency of the rotating masses. Practical experience has
shown that the best compaction effect is obtained at a higher frequency than the resonance
frequency.

20

40

"5 60

80

100
16
Number of passes

Fig. 293. Average settlement of the surface of 1.5 m thick rock fill vs. number of passes of a 5 1 smooth
vibratory roller (Lindblom,1973).
406 So/7 improvement

T h e d y n a m i c action of vibratory rollers induces stress w a v e s into the soil (mainly


longitudinal compression waves). T h e intensity of the stress w a v e s m u s t b e sufficiently
high to cause a reorientation of the soil particles into a denser state. T h e densification
obtained is dependent on the n u m b e r of passes of the roller. B a s e d on experience, the
induced settlement of the soil surface can b e a s s u m e d to increase linearly with the
logarithm of the n u m b e r of passes of the roller (Fig. 293).
T h e depth effect of vibratory rollers is considerably larger than in the case of static
rollers. A higher relative density is also obtained. Effective c o m p a c t i o n can b e achieved
to depths of about 3 m b y using the heaviest rollers n o w in existance. However, the density
of the u p p e r m o s t part of the c o m p a c t e d layer m a y b e relatively low due to the uplift
caused by reflected compression w a v e s .

(iii) Oscillatory roller. A conventional vibratory roller gives rise to longitudinal


compression and tension w a v e s in the soil. T h e principle behind the oscillatory roller
(Fig. 294) is to induce transversal shear w a v e s into the subsoil w h i c h is very favourable
from the point of view of compaction. T h e soil particles are sheared into a denser state
and loosening of the soil surface due to reflected tension w a v e s does not occur.
C o m p a c t i o n is faster and the wear of the e q u i p m e n t is reduced (Thurner, 1992). Accord-
ing to Thurner, the layer thickness should not exceed 0.4 m.

(i v) Vibratory plate compactors. Vibratory plate compactors are mostly used for compaction
of fill in trenches or in places w h e r e the space is too limited to allow the use of rollers and
for complementary compaction of loosened surface layers after vibratory roller compaction
has been terminated.
M o d e r n vibratory plate compactors are provided with two counteracting, adjustable
eccentrics. In this w a y the direction of the centrifugal force can b e varied and the
compactor can b e m a d e to m o v e forward or b a c k w a r d with a speed of u p to 24 m/min.

Fig. 294. The oscillatory roller is provided with eccentrics arranged in a way to induce transversal shear
waves into the soil.
Soil improvement 407

The adjustment of the eccentrics is carried out by hydraulic action by the use of a finger
tip controlled back-and-forth lever on the handle. The amplitude is high enough to give
good compaction results to fairly great depth.

2.4 Practical application

In the following a brief summary will be given regarding the applicability of the various
equipments for compaction of different types of fill (see also Hausmann, 1990). The
limits of the layer thickness that can be accepted to achieve a good compaction result is
governed by the type of equipment and geotechnical characteristics of the fill. The type
of equipment and the limiting height of the lifts for different fill types can be chosen
according to the following general outlines.
Blasted (crushed) rock fill. Heavy vibratory smooth rollers using a layer thickness of
2 - 3 m. The m a x i m u m width of the rock fragments should not exceed 2/3 times the layer
thickness. The vibratory roller should be allowed to act directly on the rock fill surface
without an intermediate layer of finer material. Spreading of rock fill material by the aid
of heavy bulldozers generally generates good compaction.
Sand and gravel. Vibratory plates using a layer thickness of 0.5-1.5 m. If static rollers
are used the layer thickness must be reduced to a maximum of about 0.25 m. The best
result is obtained if the fill is water saturated or nearly dry. Compaction of fine sand,
however, should be performed at a water content close to optimum.
Silt and well-graded soils, for example till. Heavy vibratory, rubber-tired rollers or
padfoot rollers. The water content should be close to optimum. The layer thickness should
not exceed 0.5 m. If static rollers are used the layer thickness should be limited to a
m a x i m u m of 0.2 m. Rainy weather obstructs or makes impossible compaction work.
Clay. Rubber-tired rollers, padfoot rollers or sheepsfoot rollers. If rubber-tired rollers
are applied the wheel pressure should be adjusted with regard to water content and shear
strength of the fill. Sheepsfoot rollers ought to be used if the undrained shear strength of
the clay fill exceeds 200 kPa.

2.5 Compaction in winter time

Compaction in winter entails special problems that have to be taken into account. Thus,
pore water in the soil will freeze to ice and form an obstacle to void ratio reduction. Even
apparently dry soil contains enough condensed humidity to give a similar effect. The soil
particles will be more or less locked in their positions or will be covered by a frozen ice
film that prevents direct contact between the particles. The problems increase with
increasing content of fines (Fig. 295).
In order to accomplish a good compaction result it is very important to supply the soil
with enough heat during the compaction process to prevent freezing of the soil. To reduce
408 So/7 improvement

+20C

-0.5C

V
- -2C

-5C
-10C

0 5 10 15
Water content w (%)

Fig. 295. Compaction of frozen soil becomes increasingly difficult with decreasing temperature. The
best result is obtained if the soil is completely dry. (After Forssblad, 1981)

the negative consequences of possible freezing, flushing or watering should not be


allowed. If flushing is considered necessary the water must be heated.
The freezing process is very rapid, especially in cold and windy weather. Therefore,
even when unfrozen fill material is collected from borrow pits in winter time, there is
always a considerable risk of water freezing in the soil during transport. Ice lenses and
frozen pieces of loose soil that are included in the fill are strongly bonded and cannot be
compacted. This may lead to considerable settlement in the process of thawing. For the
same reason the ground on which the fill is being placed has to be freed from frost and
ice.
If frozen fill has to be used vibratory sheepsfoot rollers are the best means to reduce
the chances of future detrimental settlement. The best is to recompact the fill in summer
after the fill has thawed completely. If the fill is placed inside an unheated building, the
thawing process can require considerable time.
In case compaction has to be carried out in winter, fine-grained material should be
avoided and only dry material should be allowed. Moreover, the compaction energy
should be increased in relation to what is generally considered necessary and close
surveillance of the work is required.

2.6 Control methods

Generally, the purpose of compaction is to achieve as high a density (as small a void ratio)
of the compacted soil as possible. High density is considered to be coupled with good
geotechnical properties so the most common control method is based on the results of
density observations. Samples which are taken at the surface of the compacted fill, and
whose volumes are determined by some standardised procedure (for example by means
So/7 improvement 409

of water or sand volumeters), are dried to d e t e r m i n e their dry densities and the values
obtained are c o m p a r e d with the m a x i m u m dry density d e t e r m i n e d beforehand by the
Proctor m e t h o d , for e x a m p l e .
As an alternative to the use of volumeters the use of nuclear density m e a s u r e m e n t s
h a v e b e c o m e quite popular. This m e t h o d gives a quick a n s w e r (and takes only o n e tenth
of the time required by the use of a sand volumeter) and requires less m a n u a l input.
Moreover, the dry density is obtained directly w h i c h is very a d v a n t a g e o u s in a field
control w h e r e an i m m e d i a t e answer is required.
T h e requirements to b e placed on the c o m p a c t i o n result naturally h a v e to b e adjusted
with regard to the p u r p o s e of compaction. Forssblad (1967) r e c o m m e n d s the following
1
guiding values of lowest allowable degree of c o m p a c t i o n (Rd = p ^ / p ^ m a x )
subgrade for streets, roads and airports 95%
reinforcing layer for streets, roads and airports 90%
earth d a m s 90-95%
fill in trenches 90%
fill under floors and concreted areas 90%
fill under building foundations 90-95%
T h e s e values are based on long experience and are considered as a guarantee that the
result is satisfactory.
However, in m a n y cases better control criteria can be established to c h e c k that the
result satisfies the conditions set up. In situ m e t h o d s for determination of settlement and
strength characteristics of the fill after compaction has been terminated are m a n y times
preferable ( H a n s b o & P r a m b o r g , 1980).
N o w a d a y s , d y n a m i c compaction m a c h i n e s , such as vibratory or oscillatory rollers,
can b e e q u i p p e d with a c o m p a c t i o n d o c u m e n t a t i o n system (Thurner & S a n d s t r m , 1991 ;
Thurner, 1992). Such a s y s t e m consists mainly of a so-called c o m p a c t i o n meter, a
speedometer and a c o m p u t e r unit with an L C D display. T h e c o m p a c t i o n meter provides
continuous information about vibration frequency and roller speed and presents a
compaction meter value, representing the ratio of the amplitude of the first h a r m o n i c to
the amplitude of the fundamental tone of vertical vibrations. T h e c o m p a c t i o n m e t e r value
gives a m e a s u r e of the resistance to compaction and consequently of the result obtained.

3. DEEP COMPACTION

T h e m e t h o d of c o m p a c t i n g fill material in successively placed, relatively thin layers is


often impossible to apply in practice; this is particularly true in the case of creating
reclaimed areas. T h e n all the fill has to b e placed before c o m p a c t i o n can b e started. T h e
same is true w h e n there is a need for c o m p a c t i n g natural loose soil at great depth to
improve the natural soil characteristics. In the following, the m o s t c o m m o n m e t h o d s of
deep compaction and their applicability will be described. A m o n g the m e t h o d s , compaction
410 Soil improvement

piling will be included, although this m e t h o d can be considered as something between


friction piles and soil compaction.

3.1 Heavy tamping

(i) Basic principle. H e a v y tamping, or, as it is also n a m e d , d y n a m i c consolidation (or


d y n a m i c compaction), is the most generally applicable m e t h o d of compaction, in the
sense that it can b e used for densification of any type of soil except for low-pervious,
water-saturated, fine-grained soils, such as clay and highly organic soils( dy, gyttja,
d e c o m p o s e d peat). T h e basic principle behind heavy tamping is to accomplish a
densification of the soil by m e a n s of the impact of a falling w e i g h t a pounder. As a spin-
off effect, the ratio of horizontal in situ stresses to vertical in situ stresses increases which
is favourable from a settlement point of view.

(ii) Equipment. T h e principle of using impact as a m e a n s of compaction has been utilised


for a long time on a small scale. Thus, the Proctor m e t h o d used on a laboratory scale is
based on the s a m e principle as well as manual tampers and soil r a m m e r s used in the field.
However, the depth effect of the latter e q u i p m e n t s is quite small because of insignificant
mass (generally less than 100 kg). Heavy tamping, as a full-scale m o d e r n compaction
technique, was developed and introduced by the French engineer Louis M n a r d and his
c o m p a n y T L M (Technique Louis M n a r d ) . F r o m a modest start with a m a x i m u m mass

Fig. 296. Heavy tamping applied at Nice. The mass of the falling weight is 190 t and the drop height
(free fall) 25 m. As can be seen, the weight is made by steel plates bolted together. By courtesy of TLM.
So/7 improvement 411

of the pounder (the falling weight) of 10 t and a drop height of m a x i m u m 10 m the


dimensions of the e q u i p m e n t s h a v e increased considerably. For e x a m p l e , during
compaction w o r k for a new airfield at N i c e in France, the m a s s of the p o u n d e r was 190
t and the drop height 25 m (Fig. 2 9 6 - 2 9 7 ) . A tripod, especially m a d e for the purpose with
a m a s s of 4 0 t and a drop height of 4 0 m, was developed by T L M .
To begin with, the p o u n d e r was generally a steel container filled with concrete or soil
but with increasing mass requirements the pounders are now generally m a d e of steel
plates bolted together. T h e b o t t o m area of the pounders is usually 4 m 2 .

(iii) Design aspects. According to experience based on case records, the depth dcp of
compaction caused b y the falling weight can b e estimated by the relation:

dcp = njmh (499)

w h e r e m = m a s s of the p o u n d e r in t,
h = drop height in m,
= empirical constant varying b e t w e e n 0.3 and 1.0.
T h e choice of value depends upon the grain size distribution of the soil and the degree
of saturation.

Fig. 297. Heavy tamping can be performed as shown here, dropping the weight in a direct sequence
successively at each intended point, or, alternatively, first in every second point of design and then in
the points in between not yet compacted. By courtesy of TLM.
412 Soil improvement

For cohesionless soil has been found to vary b e t w e e n 0.5 and 1. T h e lower b o u n d a r y
value concerns cases w h e r e the loss of energy nearest to the soil surface is considerable.
This is for instance true for very coarse-grained soil (boulder size) w h e r e , at the beginning
of compaction, a large part of the compaction energy is c o n s u m e d in fragmentation. As
previously mentioned, the n u m b e r of contact points per unit v o l u m e of coarse-grained
soil is smaller the larger the particle size (p. 398). T h e stress shock i n d u c e d b y the impact
cannot therefore b e resisted by the material and is the cause of fragmentation. It is
advisable, also in the case of fine-grained material to a s s u m e the lower b o u n d a r y value.
In the case of silt and clayey soils, values in the range of 0.3 to 0.5 are r e c o m m e n d e d .
T h e view is s o m e t i m e s expressed that heavy t a m p i n g can b e u s e d also for consolidation
(compaction) of clay. However, it is very doubtful if h e a v y t a m p i n g on clay has any
favourable effect at all. T h u s , n o remaining v o l u m e decrease of the soil can b e obtained
unless water and/or gas are expelled from the voids in the soil. Since the permeability of
clay soils is extremely low and, consequently, the time required for the water and/or gas
to escape is very long, ' c o m p a c t i o n ' only results in shear deformations and not in a
v o l u m e decrease. (In unsaturated clay, gas bubbles will certainly b e c o m p r e s s e d under
the impact of the b l o w but will quickly retain their original size). Moreover, the
disturbance effects of h e a v y tamping will induce high excess p o r e pressure in the clay
which starts u p a consolidation process a c c o m p a n i e d by a long-term v o l u m e decrease

Fig. 298. The depth and the outlook of the craters formed by the impact of the pounder give a good
indication of the effect of compaction. A good sign of effectiveness is that there is no heave of soil around
the craters.
So/7 improvement 413

(Hansbo, 1978). Overconsolidated clay m a y thus turn into underconsolidated clay that
after consolidation b e c o m e s normally consolidated with a lower void ratio than in its
original state. Consequently, h e a v y t a m p i n g should not b e used on clay with a high degree
of saturation and the water content of the clay should not b e a b o v e the plastic limit.

(iv) Application. T h e compaction procedure is n o w a d a y s generally following a standard


in which the p o u n d e r is dropped successively in the points settled in the design, usually
in a square pattern. T h e spacing between the points of compaction is chosen in order to
get as even a compaction result as possible. B y increasing the compaction energy at points
which will b e subjected to heavy loads (beneath columns and walls), the risk of
detrimental differential settlements can b e eliminated. C o m p a c t i o n is generally carried
out in several passes. After each pass the craters formed in the ground at each point of
compaction are filled with soil before the next pass of compaction is started. S o m e t i m e s
it m a y b e suitable to start compaction in a m o r e open pattern than according to the design
and then finalise compaction in agreement with the design. To begin with, when
compaction is started, the imprint of the p o u n d e r is generally considerable but will
successively b e c o m e reduced with the increasing n u m b e r of b l o w s . C o m p a c t i o n is
continued until the imprint of the impact is reduced to an acceptable upper limit or until
the compaction results m e e t the requirements for the subsoil based on in situ control
m e t h o d s . A h e a v e of soil around the imprints of the p o u n d e r is an indication that the
kinetic energy of the blow is partly or, if the worst c o m e s to the worst, completely
consumed by shear deformations and, therefore, thatthecompactionresultisunsatisfactory.
T h e appearence of the imprints is a good indication of the possibilities of achieving a good
compaction result (Fig. 298).

W h e n the compaction requirements at depth are fulfilled, heavy tamping is usually


terminated by a so-called ironing pass in which the pounder is dropped from about 1-2
m height in a very dense pattern to compact the superficial part of the soil (Fig. 299).
T h e densification of the subsoil in the course of heavy tamping is often accompanied
by a sudden increase in pore pressure which in the case of fine-grained soil, such as silt
and silty sand, m a y cause liquefaction and create vertical flow channels. T h e flow
channels in combination with vertical fissures, caused by horizontal tension stresses,
accelerate the dissipation of excess pore pressure. In addition, the excess pressure will
h a v e a p e a k b e l o w the point of compaction and therefore three-dimensional outflow takes
place. This explains w h y heavy tamping can be used as an effective m e a n s of compaction
also in the case of saturated silt.
414 So/7 improvement

Fig. 299. Ironing of the surface of compacted silt and sand fill at the Vnern terminal (Hansbo, 1974
and 1978)

3.2 Vibrocompaction

(i) Basic principle. T h e basic principle behind vibrocompaction is to cau.se a reorintation


of the soil particles into a denser state by m e a n s of induced vibrations in the horizontal
or the verticafdirection.

(ii) Equipment. There are two methods utilised for vibrocompaction, namely vibroflotation
and the vibro-wing or vibro-rod method. Vibroflotation is the oldest m e t h o d and has been
used since the 1930s while the vibro-rod has been developed fairly recently in Japan and
the vibro-wing even m o r e recently in Sweden.
T h e vibrating unit used for vibroflotation consists of a steel tube, 3 - 8 m in length and
0 . 3 0 - 0 . 5 0 m in diameter, provided with a vibrator and a water jetting device. An electric
motor at the top of the vibrator operates an eccentric m o u n t e d on a vertical axis to produce
horizontal vibrations at the lower part of the unit. T h e vibrator is connected to an
extension rod w h o s e length can be adjusted with regard to the depth of compaction.
Normally, the frequency of the vibrator is 3 0 - 5 0 Hz. T h e m a x i m u m centrifugal force is
about 2 2 0 k N and the m a x i m u m amplitude 25 m m . In order to facilitate penetration and
withdrawal of the equipment water jetting is utilised with a flushing pressure of up to 0.8
M P a and a water flow of u p to 3 m 3 / m i n . During compaction water jetting brings fines
to the ground surface. By adding coarse material, which replaces the fines, well com-
pacted columns of coarser material is obtained (Fig. 300).
So/7 improvement 415

Fig. 300. Sequence of operations in compaction of sand by means of vibro-flotation.

Another type of vibro-compaction is represented by the so-called vibro-rod and vibro-


wing m e t h o d s . In this case the vibrations take place in the vertical direction and not in the
horizontal direction as w a s the case with vibroflotation. T h e vibro-wing is a further
d e v e l o p m e n t of the vibro-rod w h e r e the proturberances attached to the vibro-rod are
e x c h a n g e d for steel fins with a length of 0.8 m at 0.5 m spacing (Fig.301). T h e vibro-rod
and the vibro-wing are operated by conventional piling rigs with a h e a v y vibratory
h a m m e r w h o s e frequency is in the range of 5 - 2 0 H z for the Franki tristar p r o b e and
typically about 2 0 H z for the vibro-wing.
T h e Franki tristar p r o b e (Fig. 302) is another vibro-compaction tool. It has three long
steel plates, 5 0 0 m m in width and 2 0 m m in thickness, attached to a 1 5 - 2 0 m long steel
rod at 120 angles to eachother. Additional 300 m m long steel ribs, 5 0 m m in width and
10 m m in thickness, are fixed on top of the steel plates at 2 m intervals in order to i m p r o v e
the compaction efficiency.
A new technique has recently been developed according to w h i c h c o m p a c t i o n is
carried out at the resonant frequency of the subsoil (Massarsch, 1991 ). A d y n a m i c probe
416 So/7 improvement

Fig. 301. Compaction of sand fill by means of the Vibro-wing. By courtesy of NCC.

with low i m p e d a n c e is operated by m e a n s of a heavy vibrator with a centrifugal force of


up to 4 M N and with variable operating frequency. After the probe has been inserted into
the soil the frequency of the probe is adjusted to the resonance frequency of the soil. T h e
intention is to h a v e the w h o l e soil oscillating simultaneously in the vertical direction
during compaction and, thus, to h a v e an optimal transfer of the vibration energy to the
soil. T h e result is a m o r e effective compaction.

Fig. 302. Compaction of sand by means of the Franki tristar probe. By courtesy of Mller Geosystem.
So/7 improvement 417

Fig. 303. Upper and lower boundaries of grain size distribution curves for soils suitable for compaction
by means of vibration and blasting.

() Application. Vibratory deep compaction is successful in sand soil w h o s e grain size


distribution curve falls within the limits given in Fig. 3 0 3 . In the case of vibro-wing or
vibro-rod, the soil should preferably b e water saturated. This is not required in the case
of vibroflotation w h e r e water jetting can b e used. Vibrocompaction is generally carried
out in an equilateral triangular pattern with a spacing b e t w e e n the points of compaction
varying from about 5 m in fine sand to about 1.5 m in coarse sand.
T h e u p w a r d flow of flush water in vibroflotation carries fine material in the soil u p to
the ground surface w h e r e it settles. This material is generally taken a w a y and replaced
by coarse material w h i c h sinks d o w n into the soil around the vibrator and forms well
c o m p a c t e d c o l u m n s with high bearing capacity. A similar principle is used in so-called
vibro-replacement w h e r e cohesive soil around the vibrating unit is r e m o v e d by m e a n s of
jetting and replaced b y a c o l u m n of coarse material c o m p a c t e d b y vibration.

3.3 Blasting

(i) Basic principle. T h e shock w a v e caused by detonation of explosives can b e used as


another m e a n s of achieving deep compaction. T h e m e t h o d can b e applied for compaction
of water-saturated n o n c o h e s i v e soils with a grain size distribution falling within the limits
given in Fig. 3 0 3 , i.e. for soils suitable for c o m p a c t i o n b y m e a n s of vibroflotation or by
the vibro-wing.
418 So/7 improvement

Fig. 304. Liquefaction phenomenon taking place after blasting in loose sand. By courtesy of L. Kok.

T h e basic principle behind compaction by blasting is that the detonation of explosives


induces in the soil a shock w a v e of such a m a g n i t u d e that it causes a state of liquefaction
(Fig. 304) followed by water escape and a reorientation of the grains into a denser state.
T h e success of the m e t h o d thus depends on the ability of the s h o c k w a v e to break down
the existing soil structure.

(ii) Design. T h e r e is as yet no theoretical basis for predicting with satisfactory degree of
accuracy the effective z o n e of influence of blasting. T h e m a x i m u m depth to which
blasting can cause compaction is not k n o w n , but the technique has b e e n used successfully
d o w n to depths of m o r e than 30 m below ground surface. T h e radius of influence in the
horizontal direction will b e dependent on the overburden pressure. M o r e energy is
required to m o v e the grains and cause liquefaction at high overburden pressure than at
low overburden pressure. Therefore, the effective radius of c o m p a c t i o n will decrease
with increasing depth.
According to Ivanov (1980), the stresses (in kPa) in the soil skeleton due to a shock
w a v e can b e expressed as:

G=K(Owy (500)

w h e r e R = distance (in m ) from the centre of the charge,


C = size of charge (in kg T N T ) ,
and are empirical coefficients w h o s e values d e p e n d on the gas content. (In
fine-grained sand, o v a r i e s from 6 0 0 at full saturation to 45 at 4 % gas content and
So/7 improvement 419

correspondingly from 1.05 to 2.5. At a gas content of 1%, ~ 250). The influence
of gas content on and shows that even a small amount of gas will strongly
reduce the shock wave stress level, and consequently reduce the effectiveness of
blasting.
The size C of the explosives (in kg T N T ) can be taken as (Ivanov, 1978):

3
C = 0.055 dch (501)

where dch = the depth (in m) at which the explosive is detonated.


The propensity for liquefaction as a result of blasting, which governs the potential for
densification, can be expressed as a function of the liquefaction coefficient w/' 0 , in which
Aw = excess pore water pressure due to blasting and '0 = effective overburden pressure
prior to blasting. If the depth effect is disregarded then the expected value of the lique-
faction coefficient can be expressed by the relation (Kok & Trense, 1979):

1/3
/'0 = 1.65 + 0.641n(C /#) (502)

1/3
with a maximum value of Aula'0 equal to 2.15 + 0.741n(C //?) and a minimum value equal
1/3
t o l . 5 3 + 0.771n(C /#)
Liquefaction occurs when Aula ' 0 tends to unity. This gives a mean value for the radius
of liquefaction:

1 /3
Rliq = 2.8C (503)

1 /3 1/3
with a m a x i m u m of 4.7 C and a minimum of 2 . 0 C .
Based on practical experience, the effective radius (in m) of compaction (defined as
the radius, within which fairly uniform compaction occurs and approximates the base of
the settlement saucer) can be expressed as:

l/3
ReJf=kC (504)

where k = 2 - 5 , i.e. close to the range of radius of liquefaction.


The depth of compaction dcp (in m) can be estimated by the relation:

rfc/,= 1.5</cft = 3.9Ci'3 (505)

Compaction by blasting is most frequently achieved by detonation of buried explosives


but can also be achieved when the explosives are placed in free water, e.g. in the sea at
a certain water depth. The shock wave is then directed towards the sea bottom and the soil
below bottom will react partly in the same way as if the explosives had been buried in the
420 So/7 improvement

soil itself. In this case, the charge is placed at a height h (in m ) a b o v e b o t t o m according
to the relation:

h = 0.35C (506)

and the size of the charge (in kg) according to the relation:

2 46
C = 0.1// (507)

w h e r e H = depth of water (in m ) .


Design diagrams are given in Figs. 305-306.

(iii) Application. T h e blasting technique has been developed and rationalised, for
e x a m p l e in the Netherlands and Russia, since about 4 0 years b a c k (Fig. 3 0 7 ) . As a great
deal of uncertainty about the effect of blasting still exists, a test field is r e c o m m e n d e d for
the study of the effectiveness and the z o n e of influence of blasting.
T h e following practical guidelines can b e given:
T h e charges should be placed at approximately 2/3 (not less than 1/4; 1 Z 2 - 3 / 4 common)
of the desired depth of compaction.
T h e spacing between the detonation holes is usually b e t w e e n 5 and 1 5 m (not less than
3 m ) . T h e spacing should be based on experience gained from the test field.
Delayed blasting is preferable to simultaneous blasting.
T h e n u m b e r of coverages is usually 2 - 3 . Each coverage consists of a n u m b e r of
individual charges. Successive coverages are separated by hours or days. For each new
coverage the effectiveness is gradually reduced.


0 4 I 2 I 6
C kg T N T

Fig. 305. Design diagram for blasting in water saturated sand. Legend: Reff= effective radius; dcp = depth
of compaction; dch = depth of charge.
So/7 improvement 421

"0 10 20 30 40 50
C, kg TNT

Fig.306. Design diagram for blasting in water. Legend: Re^= effective radius; dcp - depth of compaction;
H = required depth of water; h = height of charge above sand bottom.

T h e individual charge size varies generally b e t w e e n 1 and 12 kg. T h e total explosive


use a m o u n t s to 8 - 1 5 0 g / m 3 of soil, usually 1 0 - 3 0 g / m 3 .
T h e soil closest to the ground surface will b e poorly c o m p a c t e d and will therefore h a v e
to be c o m p a c t e d in s o m e other way, or b e taken away.
T h e compaction achieved by blasting is a function of the relative density of the soil.

Fig.307. In modern blasting technique, special rigs for installation of explosives are utilised (left). The
explosives are enclosed in thermo-shrinked plastics for direct placement in the detonation holes. By
courtesy of L. Kok.
422 So/7 improvement

T h e following typical behaviour of the soil can b e expected:


A l m o s t i m m e d i a t e settlement of the ground surface with little further settlement with
time. T h e surface settlement m a y be 2 - 1 0 % of the thickness of the layer subjected to
compaction.
Initially loose zones show little immediate c h a n g e in penetration resistance. Penetration
resistance increases slowly with time, until after several w e e k s the soil indicates a marked
i m p r o v e m e n t in properties as compared to its initial condition.
Zones which are initially very dense may be permanently loosened or w e a k e n e d by the
blast. However, the resultant overall condition is still likely to b e satisfactory.

(iv) Case record: As previously mentioned, gas in the soil obstructs the effectiveness of
compaction by blasting. A case record described by D o n c h e v (1980), w h e r e the use of
blasting for densification of unsaturated loess soil to i m p r o v e the foundation base for
three high-rise buildings, is therefore of particular interest. Before blasting, the soil was
saturated by feeding water to drainage boreholes with a spacing of 2 - 2 . 5 m. Each second
drainage borehole was loaded with explosives. T h e explosions were realised in a series
of 2-4 boreholes at a time interval of 1-5 s.
T h e blasting w a s immediately followed by a settlement of 1.0-1.4 m. Settlement
increased considerably for two weeks and was finished within t w o to three months.
According to D o n c h e v the settlement of the buildings, built on shallow foundations on
the compacted soil, h a v e b e c o m e even and a m o u n t to 0 . 1 4 - 0 . 1 8 m at a foundation
pressure of 200 kPa.

3.4 Compaction piling

(i) Basic principle. According to experience, driving of piles in the neighbourhood of


buildings often give rise to detrimental building settlements. T h e settlements are caused
by both soil vibrations induced by pile driving and soil displacements corresponding to
the pile v o l u m e inserted into the soil. This p h e n o m e n o n is taken advantage of in
compaction piling, w h e r e the main object of pile driving is to achieve compaction. T h e
piles also serve as reinforcement of the soil.

(ii) Application. C o m p a c t i o n piling is generally carried out with timber piles, less usual
with concrete piles. W h e n timber piles are used the pile cut-off should b e satisfactorily
below the lowest groundwater table in order to avoid rotting. T h e pile spacing is generally
chosen at 1.5-2.5 m. T h e compaction effect, according to experience, extends to a
distance from the pile periphery of about 3 times the pile diameter ( s o m e w h a t m o r e
nearest to and b e l o w the pile tip). T h e increase of average density due to compaction
piling can b e estimated by m e a s u r e m e n t of the settlement of the soil surface and the pile
v o l u m e driven into the ground.
Soil improvement 423

Compaction piling is effective only in water saturated or completely dry non-cohesive


soils.

3.5 Compaction control

In the case of deep compaction, the control methods used to check the results of shallow
compaction are no longer applicable. A rough indication of the result is obtained by
studying the settlement of the soil surface during compaction, and this should always be
done since it gives us information of the relative compression of the subsoil due to
compaction. However, other characteristics of the soil after compaction will have to be
determined by more sophisticated methods. In this case, it is necessary to use in situ
methods, such as sounding (for example C P T or SPT), preferably in combination with
pressuremeter or dilatometer tests (Fig. 308). Sounding results give a good picture of the
homogeneity of the soil and may be used as an indirect measure of the deformation and
bearing characteristics of the compacted soil (p. 103). A safer measurement is obtained
by the use of the pressuremeter (Fig. 53) or, where this is possible, the dilatometer (Fig.
56).
In case neither sounding nor pressuremeter or dilatometer tests can be performed, for
instance in rock fill, the results obtained can be checked by measuring the surface wave
velocity. The surface wave velocity serves as a basis for determination of the average
dynamic shear modulus of a surface layer, with a thickness equal to the wave length of
the surface wave (cf. p. 87). By varying the wave length a good picture can be obtained
regarding the variation with depth of the shear modulus.

qc, MPa Epr MPa pl, MPa

Fig. 308. Influence of heavy tamping (12 t pounder dropped from 12 m height) on the characteristics
of silty sand at Vnerterminalen, Karlstad (Hansbo et al, 1974). Cone penetration test carried out before
compaction, after 2 passes and after 4 passes of heavy tamping. Pressuremeter values obtained before
and after termination of compaction.
424 So/7 improvement

Ground pressure, MPa Ground pressure, MPa


0 1 2 3 4 0 1 2 3 4
0 0

0.20'
0.25

Fig. 309. Ground pressure vs. settlement curves obtained by measurements of the deceleration of a 40
2
t pounder (bottom area 4 m ) as compared to the results of corresponding plate loading test (bottom area
2
of plate 4.9 m ) . 2nd test (right) carried out after complementary compaction on the test spot (four blows
from 20 m height and one the last from 30 m height).

In the case of h e a v y tamping, a new m e t h o d of checking the characteristics of the soil


nearest b e l o w the b o t t o m of the crater, caused b y the i m p a c t of the pounder, w a s utilised
in connection with h e a v y tamping of r o c k fill at Uddevalla, S w e d e n (Andrasson and
H a n s b o , 1977; H a n s b o , 1977). T h e deceleration of the pounder, w h e n it hits the ground,
w a s m e a s u r e d by m e a n s of an accelerometer attached to the pounder. T h e imprint of the
p o u n d e r in each b l o w was also measured. Since the m a s s and the b o t t o m area of the
p o u n d e r are k n o w n , w e can easily find the stress/strain relation in each blow by
integrating twice the w e l l - k n o w n N e w t o n ' s 2nd law. T h e results of the integration are
checked against the measured imprint. Although this is a d y n a m i c method, the deformation
m o d u l u s determined is approximately equal to the static deformation m o d u l u s (Fig. 309)
b e c a u s e of the fact that w e h a v e to deal with deformations of equal size (cf. Fig. 59).

4. PRELOADING

4.1 The preloading technique

(i) Application. Preloading is applied in order to reduce, or eliminate, settlements that


w o u l d otherwise occur under the load of, for e x a m p l e , a road e m b a n k m e n t , a fill or a
building to b e constructed. T h e preloading m e t h o d has b e e n u s e d occasionally long since
but has not b e c o m e general until the u s e of vertical drains c a m e into practice for the
p u r p o s e of speeding u p consolidation. In reality, it has b e e n so intimately c o n n e c t e d with
vertical drains that its application in other connections is exceptional.
T h e best result of preloading is obtained if it is d o n e in a w a y similar to the final loading
condition. T h u s , it is important that the relation b e t w e e n shear stresses and n o r m a l
So/7 improvement 425

Fig. 310. By taking advantage of occasional overloading the consolidation settlement for the design load
can be attained in relatively short time. The broken curve indicates the consolidation process under the
design load in case overloading is not utilised.

stresses induced by preloading on the one hand and final loading on the other is as equal
as possible. In order to avoid misinterpretation of the results of preloading and to m a k e
possible an effective check-up of the consolidation process (for instance the required time
of preloading) it is necessary to k n o w in detail the geological and geotechnical conditions
at the site. T h e planning and execution of preloading m u s t b e carried out by specialists
and a careful follow-up of the consolidation process is a must.
Preloading is mostly used to eliminate settlement in highly compressible cohesive
soils but is s o m e t i m e s used also in the case of coarse-grained soils.
In the former case, the time of preloading should be long e n o u g h to eliminate primary
consolidation settlement and, at least partially, secondary settlement for the final load
(Fig. 310). T h e real process of consolidation ought to be carefully monitored with both
settlement gauges and pore pressure meters and the results obtained should be checked
against theoretical prediction. If there is good agreement preloading can be terminated
w h e n the consolidation settlement predicted for the design load has been reached (or
preferably e x c e e d e d ) .
W h e n preloading is applied on coarse-grained soils, it is suggested that the loading
time b e chosen so as to reduce creep to a m a x i m u m of 0.02 m m / m i n .

4.2 Design for preloading of low-permeable soils

(i) Homogeneous soil conditions. T h e time required for preloading of low-permeable


soils is determined on the basis of the consolidation theory (p. 324). Generally, earth fill
is applied as a m e a n s of preloading but alternative m e t h o d s , such as occasional
groundwater lowering, h a v e also been utilised. T h e width of the preloaded area is usually
426 So/7 improvement

large in comparison with the thickness of the compressible soil layer. Therefore, one-
dimensional consolidation can be assumed to take place (cf. E x a m p l e s 6 2 - 6 3 , p. 328).
It is important to check that the required degree of consolidation is reached throughout
the soil layer. T h u s , the consolidation time cannot be based on the average degree of
consolidation determined according to Terzaghi 's consolidation theory. T h e degree of
consolidation Uv reached at a certain depth can easily b e obtained by the numerical
m e t h o d presented on pp. 3 2 5 - 3 2 7 , or, in the case of equally large excess p o r e pressure
throughout a h o m o g e n e o u s soil deposit, by the relation:

2
1 - Uv = - s i n ( W - ) e x p ( - / V 7 v ) (508)
h

w h e r e - / 2 , 3/2, , (2m + 1)/2, , (ra is an integer),


2
Tv = Cytlh - the time factor.

4.3 Speeding up consolidation by vertical drains

(i) General aspects. In cases where the subsoil consists of fine-grained soils with low
permeability, the time of preloading required before consolidation settlements have
terminated can be very long, particularly w h e r e the thickness of the fine-grained soil layer
is large. Settlement will continue until the excess pore water pressure induced by
preloading has dissipated and the soil skeleton has b e c o m e strong e n o u g h to carry the new

sand gravel
100

0.06 2 60
Grain size d, mm

Fig. 311. The grain size distribution boundaries of ideal sand suitable for sand drains.
So/Y improvement 427

load on its o w n . T h e philosophy behind the u s e of vertical drainage is to shorten the


drainage paths b y installing drains vertically into the soil in s o m e regular pattern. In this
way the consolidation process can b e speeded up considerably. Both drains of natural
sand and drains m a d e artificially are utilised.
T h e u s e of sand drains is connected with certain disadvantages. T h e sand to b e used
for the drains m u s t b e chosen with extreme care in order to h a v e good drainage properties
(Fig. 311) and can therefore s e l d o m b e found in the vicinity of the site of drain installation.
Moreover, there is an obvious risk that the drains b e c o m e discontinuous (necking) and
thus ineffective d u e to careless installation or excessive lateral soil displacements during
consolidation. T h e diameter of sand drains, generally varying from about 160 to 5 0 0 m m ,
represents another problem. In order to reduce disturbance of soil during installation,
which w o u l d r e d u c e the efficiency of the drains, preboring m a y b e required for the drains
(non-displacement type drains). In order to avoid the need of preboring without causing
unacceptable disturbance effects, sand drains with a diameter of only about 5 0 m m , so-
called sand wicks (Fig. 312), are n o w a d a y s prefabricated on the site b y enclosing the
drain sand in a "stocking". In this w a y the risk of necking is also eliminated.
T h e artificial drains are generally so-called "band drains". Generally speaking, these
are built u p b y a grooved core enclosed in a filter sleeve w h i c h prevents soil particles
m o v i n g with the p o r e water from penetrating into the grooves, w h i c h m i g h t lead to
clogging. Prefabricated band drains (Fig. 313) h a v e an obvious a d v a n t a g e over sand
drains in that their quality and characteristics can b e well defined and the risk of necking
eliminated.

Fig. 312. The sand wick represents a type of prefabricated small-diameter sand drain.
428 So/7 improvement

T h e cardboard wick, invented by Kjellman in the late 30s, is the prototype of all the
m o d e r n band drains existing on the market. It is an interesting fact that sand drains were
first installed for stabilization purposes only 5 years earlier than the cardboard wicks
(Johnson, 1970). B a n d drains and sand drains are thus of very nearly the same age.
T h e overall features of the different proprietary drains are very nearly the same. A
successive development is taking place of both core and filter design, mainly for the
purpose of cost reduction. N o w a d a y s , all types of filter are m a d e of synthetic material.
S o m e filters are in the shape of a loose sleeve enclosing the core while others are fixed
to the core. Besides various types of filter, one and the s a m e trade m a r k m a y represent
a n u m b e r of different drain cores.
N o w a d a y s the various drain m a k e s differ only in respect of the channel system in the

Fig. 313. A considerable number of band drains are marketed. Shown here from top to bottom and from
left to right are: Cardboard wick (the prototype of all band drains), Geodrain, Castle Board, Alidrain,
PVC, Desol, Mebradrain and Colbond.
So/7 improvement 429

drain core and the filter characteristics, whereas their geometrical properties are m o r e or
less the s a m e (see H a n s b o , 1992).
Sites in need of vertical drainage quite often h a v e a low bearing capacity. Consequently
there m a y b e a need of reinforcing the ground surface, for e x a m p l e by m e a n s of
geotextile, in order to enable the machinery for drain installation to enter the site without
danger.
T h e drainage layer should b e at least 0 . 3 - 0 . 5 m in thickness. T h e material in the drain-
age blanket shall h a v e good drainage properties; otherwise it is necessary to i m p r o v e
horizontal drainage by connecting the vertical drains with a horizontal drainage system
in order to avoid the build-up of back-pressure in the drains which w o u l d delay the
consolidation process.

(ii) Drain installation techniques. T h e next step of construction includes drain installation
in accordance with the directions given in the design. M o s t l y the drains are installed in
an equilateral triangular pattern or in rows with a drain spacing equal to the spacing
between the drain rows. Irrespective of which of the two drain patterns is chosen, the
n u m b e r of drains required to achieve a certain rat of consolidation will b e the same. It is
important that the drains after installation do not deviate appreciably from the vertical
direction.
For sand drains, several installation procedures are used. T h e m o s t c o m m o n types of
sand drains with reference to the installation m e t h o d are driven closed-end mandrel sand
drains, continuous flight hollow stem augered sand drains, internally jetted sand drains,
rotary jet sand drains and dutch jet-bailer sand drains. A m o n g these, the driven closed-
end mandrel causes causes m o r e disturbance of the soil than the others. A c c o r d i n g to
L a d d (1976), the continuous flight hollow stem auger and the j e t bailer auger m e t h o d s
h a v e advantages over the internally jetted and the rotary j e t m e t h o d s of drain installation
in that they cause a lesser disturbance and do not require as close a supervision as the other
installation m e t h o d s .
Contractors w o r k i n g in the field of b a n d drain installations h a v e generally developed
their o w n type of e q u i p m e n t for drain installation. A c o m m o n feature is that the drains
are held inside a steel mandrel which protects the drain from being d a m a g e d during
installation. D u e to the small cross-sectional area of the b a n d d r a i n s 3 - 7 m m in
thickness and 100 m m in w i d t h t h e m a n d r e l for their installation can b e m a d e very
slender and, therefore, the disturbance effects during installation can b e greatly reduced.
Two principles of installation can b e distinguishedthe so-called static and d y n a m i c
installation m e t h o d s . In the first case, the m a n d r e l with the drain inside is p u s h e d into the
soil by static pressure, while in the second case it is driven into the soil b y m e a n s of a
gravity h a m m e r or a vibratory driver. Both static and d y n a m i c installation rigs of various
types exist, from conventional piling rigs to quite advanced rigs such as those shown in
Fig. 314. Floating rigs exist by w h i c h drains can b e installed from o p e n water, Fig. 3 1 5 .
430 So/7 improvement

Fig. 314. Rigs for installation of band drains: static, to a depth of more than 30 m (left); dynamic to a
depth of more than 40 m.

Fig. 315. Rig for installation of band drains from open water.
So/7 improvement 431

T h e m e t h o d of installation static or d y n a m i c does not s e e m to affect the


efficiency of the drainage system unless it is carried out in a w a y leading to excessive
smear. D y n a m i c installation should, however, b e avoided wherever disturbance effects,
usually evidenced by excess pore pressures being built up during installation, m a y
adventure stability.
Before the b a n d drains are inserted into the soil they must b e provided with an anchor
that keeps the drains in their position w h e n the mandrel is w i t h d r a w n (Fig. 316). T h e
anchor also prevents soil intruding into the mandrel during installation (which might lock
the drain to the mandrel by friction). Different types of anchor are used by different
contractors. T h e design of the anchor is important for the extent of the z o n e of smear.
After the mandrel is withdrawn the drain should be cut in such a w a y that a good
connection with the drainage blanket is ascertained. Rigs exist that are provided with a
cutting device by w h i c h f o r e x a m p l e in a drain installation from open w a t e r t h e drains

Fig. 316. Cutting the drain after installation (top). An anchor is the fixed to the drain. The anchor
prevents soil intruding into the mandrel during drain installation and keeps the drain in place when the
mandrel is withdrawn.
432 So/7 improvement

can be cut just above the drainage blanket placed on the bottom. In deep water this
technique can save a considerable length of drains.
T h e present types of mandrel are often slender. Therefore, w h e n they are inserted into
the soil they m a y deviate considerably from the vertical, particularly in the case of deep
installations. Since it is essential for a well-functioning drain system that the prescribed
drain spacing is maintained throughout the drained soil layer such a deviation from the
vertical can h a v e a negative and unpredictable effect on the consolidation process at great
depth. Therefore, either the mandrel should be equipped with an inclinometer that gives
information about the horizontal position of the drain at various depths or the mandrel
should be stiff enough to ensure verticality of the drains.

4.4 Design of drain installations

(i) Theoretical assumptions. In the theoretical analysis of the effect of vertical drainage,
each drain is assumed to dewater a circular-cylindrical soil column w h o s e v o l u m e equals
the parallelepiped formed by four neighbouring drains in two adjacent r o w s , Fig. 317. In
order to m a k e this parallelepiped coincide in the best possible way with a circular cylinder
the drains are often installed in equilateral triangular pattern. If the cross-sectional area
of the parallelepiped is A, obviously the diameter of the circular cylinder is:

D = 2</7. (509)

Fig. 317. Terms used in the analysis of vertical drains. D = diameter of soil cylinder dewatered by a drain,
d= drain diameter, ds = diameter of zone of smear, / = length of drain when closed at bottom (21 = length
of drain when open at bottom), qw = discharge capacity of drain.
So/7 improvement 433

T h o s e w h o h a v e a profound interest in different a p p r o a c h e s to the vertical drain


analysis are referred to M a g n a n (1983) and L o (1991). T h e solution to the p r o b l e m is
greatly simplified if o n e a s s u m e s that horizontal sections of the soil r e m a i n horizontal
throughout the consolidation process ( the so-called equal strain theory). O t h e r w i s e a
rather c o m p l i c a t e d solution is obtained. H o w e v e r , B a r r o n ( 1 9 4 4 ) s h o w e d that the
numerical value of the average degree of consolidation obtained by a s s u m i n g equal
strains is a l m o s t equal to the value obtained b y a s s u m i n g that the strains in the soil
develop freely in full a c c o r d a n c e with the rate of p o r e pressure dissipation (the so-called
free strain theory).
Originally, in the theoretical analysis, the permeability of the drains w a s a s s u m e d to
be high e n o u g h not to cause any well resistance ('ideal d r a i n s ' ) . It w a s also a s s u m e d that
the installation of the drains could be m a d e without causing any disturbance of the soil,
an a s s u m p t i o n still considered acceptable in case preboring techniques for drain
installation are utilised. As the installation of drains generally entails a z o n e of smear with
reduced permeability around the drain, and as s o m e drains m a y h a v e quite a low
discharge capacity (considerable well resistance), it seems necessary to include in the
theoretical analysis the influence on the consolidation p r o c e s s of disturbance and in
certain cases also of well resistance.
T h e analysis of the influence of vertical drainage can b e based on the following
assumptions:
equal strains irrespective of the radial distance from the drain centre,
the soil is water-saturated,
the drain is circular-cylindrical with diameter d,
the drain has limited discharge capacity,
the installation of a drain causes a circular-cylindrical z o n e of s m e a r around the drain
with diameter ds and l o w e r permeability than in the undisturbed soil,
D a r c y ' s law is valid, /. e. the rate of flow is directly proportional to the hydraulic
gradient,
the effect of p o r e water flow in the vertical direction b e t w e e n the drains is neglected,
the total stress is the s u m of the effective stress and the p o r e w a t e r pressure.
T h e s e a s s u m p t i o n s lead to the following a p p r o x i m a t e solution of the consolidation
process (for details, see H a n s b o , 1981), expressed in the terms of a v e r a g e d e g r e e of
consolidation at depth z:

u
hz=
1
- hAo
u
z
1
= -
e x
P(-
8 7
)/) (510)

where uhz = a v e r a g e excess p o r e pressure at depth and t i m e t,


u0z = initial a v e r a g e excess p o r e pressure at depth z,
2
Th - cht ID - time factor,
ch = coefficient of consolidation in horizontal p o r e water flow,
434 So/7 improvement

t = time of consolidation,
D = diameter of drained circular cylinder,
= 1() + (kf/ks)lns - 0.75 + (2 - z)kf/qW9
n=D/d,
s =ds Id,
I = full (half) length of drain if closed (open) at bottom,
kh = permeability (in the horizontal direction) of soil,
ks= permeability (in the horizontal direction) of zone of smear,
qw = specific discharge capacity of drain (qw = k^A^, where kw = permeability in
the vertical direction of the drain and Aw = cross-sectional area of the drain).
Since band drains have a rectangular cross-section it is necessary to transform them
into circular-cylindrical drains with equivalent draining capacity. Kjellman (1948) stated
that "the draining effect of a drain depends to a great extent upon the circumference of
its cross-section, but very little upon its cross-sectional a r e a " and that "certain
considerations show that the cardboard wick is as effective as a circular drain with a 1-in.
radius". This statement was later confirmed by the aid of finite element analysis (Hansbo,
1979, 1981). Thus, band drains and circular drains can be assumed to be equally efficient
provided that their circumference is equal, in other words:

deq = 2(b+t)ln (511)

w h e r e b = breadth and t = thickness of the band drain.


As previously mentioned, the assumption of equal strain taking place irrespective of
the distance from the drain yields almost exactly the same average degree of consolidation
as the assumption of free strain taking place. Since Barron first presented his free strain
and equal strain solutions a great number of theoretical studies have been presented
among which the most complete one in respect of the influence of well resistance is that
presented by Yoshikuni in his doctoral thesis (1979).
A study of the influence of well resistance performed at the University of Illinois by
the aid of a finite element programme named ILLICON has shown acceptable agreement
between the simple solution according to Eq. (512) and the results obtained according to
ILLICON (Lo, 1991) and other more sophisticated solutions (Fig. 318).

(ii) Alternative design based on non-validity of Darcy's law. The vertical drain theory
presented above is based on the assumption that Darcy's law is valid. This may not always
be the case. We know from soil physics that the pore water in claywhich in reality
consists of water molecules and ionsis more or less strongly bound to the mineral
surface of the clay particles. Moreover, in the pore water there are most probably a great
number of microscopic mobile mineral particles and/or organic matter which are bound
only by sorption and hydrodynamic forces and which do not belong to the load-carrying
So/7 improvement 435

Fig. 318. Comparison between approximate and rigorous solutions of consolidation of clay provided
with vertical drains. Drain characteristics: = 5; s = 1; HD = 16; kjkh = 1730. (Mesri, 1990).

clay skeleton. Since the binding forces increase in the direction of the mineral surfaces
it is but logical to a s s u m e that h y d r o d y n a m i c forces will h a v e an influence upon the
apparent porosity of the clay. T h e higher the h y d r o d y n a m i c forces the larger the flow
channels, but only u p to a certain limit. This indicates non-validity of D a r c y ' s law at small
hydraulic gradients.
An alternative solution to consolidation by vertical drains, b a s e d on non-validity of
Darcy's law w a s derived b y H a n s b o (1960). In this solution the flow of p o r e water is
assumed to follow the relation given by Eq. (17), p . 22, a s s u m i n g = 1.5:

= (512)
436 So/7 improvement

w h e r e kx = coefficient of permeability in horizontal pore water flow,


/ = hydraulic gradient in radial direction towards the drain.
Full-scale tests have indicated that one can assume -1.5. Introducing a new coefficient
of consolidation = kxMlyw the consolidation equation can b e written under the form:

1) (513)
* 1 "o ]/l-Uh

w h e r e a according to Fig. 319,


t, D and y w as before,
w 0 = initial excess pore water pressure.
Eq. (513) does not take into account the influence of smear or well resistance. In
practice well resistance can usually b e neglected as indicated before. T h e effect of smear
can b e taken into account by a reduction of the value. Generally it can b e a s s u m e d that
lcv - 0 . 5 - 1 , w h e r e the lower value should b e applied w h e n the drain spacing is small
(0.9-1.2 m ) and the higher value w h e n the drain spacing is larger. T h e initial excess pore
pressure w 0 can be put equal to the consolidation load.

(iii) Influence of vertical pore water outflow between the drain. Vertical drains are generally
installed in places w h e r e the thickness of the clay layer is so large that the effect on the
consolidation process of vertical outflow of water b e t w e e n the drains can b e neglected.
This is also justified with regard to all the uncertainties involved in the choice of
consolidation characteristics of the soil. If, for s o m e reason, it is considered necessary to
take into account the effect of vertical pore water flow in the clay itself, this can be done
according to Carillo (Eq. 438).

Fig. 319. The a value as a function of Did.


So/7 improvement 437

(i ) Partially penetrating drains. In the case of partially penetrating drains, obviously the
rate of consolidation of the clay beneath the drain tips is slower than in the clay penetrated
by the drains. This will result in a hydraulic gradient between the undrained and the
drained parts of the clay w h i c h will delay the consolidation process to a certain height
above the drain tips but, on the other hand, speed u p the consolidation process to a certain
depth below the drain tips. As shown by Runesson et al. (1985) the delay in the consolidation
process within the z o n e penetrated by the drains is noticeable only to a very limited height
above the drain tips. F r o m a practical viewpoint, it can be neglected w h e n the height
above the drain tips e x c e e d around 2 0 % of the drain length.

(v) Multilayer systems. T h e solution to the consolidation problem, given by Eq. (510) for
drains w h o s e well resistance has to be considered, cannot be applied directly on soils in
which the permeability varies strongly with depth. However, for a multilayer system with
different consolidation characteristics, O n o u e (1988) developed a p r o c e d u r e which takes
into account the influence of well resistance on the consolidation process as given by Eq.
(510) for h o m o g e n e o u s soil. T h e consolidation process is calculated under the assumption
that the w h o l e soil profile is h o m o g e n e o u s and has the consolidation properties of each
of the respective layers. T h e distribution of excess pore pressure in the respective layer
is extracted and plotted as s h o w n in Fig. 320.

\ L1

f \
L1 L2
L2 \
\
I
1 I
Single-layered LI Two-layered Single-layered L2

Fig. 320. Consolidation of multi-layered anisotropic soil by vertical drains with well resistance.
Approximate solution according to Eq. (510).
438 So/7 improvement

(vi) Zone of smear. T h e method applied for the installation of the drains is of p a r a m o u n t
importance for the extent and characteristics of the zone of smear. Concerning sand
drains, installation by careful jetting or by the Dutch jet bailer auger m e t h o d seems to
cause a m i n i m u m of disturbance, while installation by m e a n s of a closed-end mandrel,
particularly if driven into the soil by drop h a m m e r s or vibratory h a m m e r s , seems to cause
a m a x i m u m of disturbance. In the former case, smear can be ignored and h e n c e ds - d.
In the latter case, the effect of installation is very similar to that of conventional pile
driving. Investigations including both drains installed by m e a n s of a closed-end mandrel
and driven piles (Holtz & H o l m , 1973; Akagi, 1976) indicate that the zone of smear can
be assumed to have a diameter of ds ~ 2d. There are undoubtedly cases where the remoulding
effect m a y reach further out in the surrounding of the drains which has been indicated by
an overall reduction in shear strength. However, as a rule of t h u m b , for driven closed-end
mandrel sand drains w e can put s = 2.
T h e extent of the zone of disturbance caused by installation of band drains will be
dependent of the type of mandrel used for installation and the size of the anchor fixed to
the tip of the drain during installation. As a rule of t h u m b , ds can be chosen on the same
premises as in the case of driven sand drains, i.e. the cross-sectional area of the mandrel
(also taking into account the area of the folded anchor) is replaced by an equally large
circular area w h o s e diameter is then doubled. T h e s value is the ratio of this ds value to
the equivalent diameter of the drain.
F r o m a practical viewpoint, the question regarding which diameter to choose in the
design is also coupled with the question regarding which permeability to choose for the
zone of smear. A s s u m i n g a value of s equal to 2, the permeability ratio k}/ks can be assumed
equal to k}/kv. On this assumption, a design diagram of the type shown in Fig. 321 can
be utilised.

(vii) Well resistance. As shown analytically, the efficiency of a drain in field conditions
depends mainly upon its long-term discharge capacity. Too low a discharge capacity can
seriously delay the consolidation process, especially in a case where the drains are
installed to great depth. It is specially important that the discharge capacity is not too low
in the beginning of the consolidation process while it is of lesser importance at a later
stage of the consolidation process.
To ascertain a high enough discharge capacity of sand drains, the grain size distribution
curve of the sand should fall within the limits given in Fig. 3 1 1 .
T h e discharge capacity of a certain band drain is generally given by the proprietor of
the drain. However, before a certain drain m a k e is selected for a j o b , evidence of the drain
fulfilling the discharge capacity requirements placed upon the drain in the design
assumptions ought to be given. Generally, appropriate laboratory testing can give suffi-
cient evidence of the discharge capacity to be expected under field conditions (Hansbo,
1983), but in case of a drain never before m a d e use of in practice, full-scale field tests are
So/7 improvement 439

f=-t ( y e a r s ) / l n ( l - Uh) fch = D^S

Fig. 321. Graphs for design of band drain installations. Negligible well resistance. Zone of smear ds - Id.
Permeability ratio kh lks = kh lkv. From the design value of average degree of consolidation Uh we find
the/value. Then t h e / c A value gives the drain spacing ( the D value).

E f f e c t i v e lateral p r e s s u r e , kPa

Fig. 322. Results of discharge capacity tests for different band drains carried out on a laboratory scale.
Drains enclosed in soil. In the ENEL tests (Jamiolkowski et al., 1983) the drains were tested in full scale,
in the CTH tests (Hansbo, 1983b) and in the KU tests (Kamon, 1984) with reduced width (40 and 30
mm, respectively). Legend: A = Alidrain, BC = Bando Chemical, CB = Castle Board, C = Colbond, G
= Geodrain and M = Mebradrain. (p) indicates filter sleeve of paper.
440 So/7 improvement

r e c o m m e n d a b l e . Full-scale field tests also serve the purpose of showing that the drains
are strong e n o u g h to resist the strains subjected to t h e m during installation. Moreover,
compression of the soil in the course of consolidation settlement entails folding of the
drain which m a y bring about clogging of the channel system. T h e latter effect is difficult
to discern in a laboratory test.
As a general rule, investigations on a laboratory scale of the discharge capacity of a
drain ought to be carried out in a way so as to simulate field conditions as closely as
possible. Such tests, performed in several laboratories (see, for e x a m p l e , H a n s b o , 1983),
h a v e revealed a great influence on the discharge capacity of the lateral consolidation
pressure (Fig. 322). T h e reason is that the filter sleeve is squeezed into the channel system
which entails a reduction of the channel area, or that the channels themselves are
squeezed together.
T h e influence of well resistance for drains with a discharge capacity above 100 mVyear
can b e neglected unless the drains are very long (/ > 20 m ) .

(viii) Filter sleeve. T h e filter permeability required can be estimated by considering the
filter as a zone of smear with a diameter ds equal lod+2t w h e r e t represents the thickness
of the filter. By studying the influence of the filter thickness on the course of consolidation
w e find that the permeability need not b e higher than the permeability of the soil in which
the drains are installed. It is the discharge capacity of the channel system that is decisive,
not too low a permeability of the filter enclosing the channel system. On the contrary, too

Fig. 323. Influence on discharge capacity of filter deterioration (Koda et al., 1986). Tests on Geodrains
with filter sleeves of synthetic material (broken lines) and paper (full lines) which were pulled out of
the soil after different lengths of time after installation (number of days given in figure).
So/7 improvement 441

high a permeability m a y entail a risk of clogging of the drains. Fine soil particles m a y
penetrate through the filter m e s h and finally lead to clogging (siltation).
With time a certain deterioration of the filter will h a v e to b e expected due to
bacteriological activity or fungi attacks. A n interesting investigation of the influence of
filter deterioration on the discharge capacity of Geodrains w a s publicized by K o d a et al.
(1986). As can b e seen from Fig. 3 2 3 , both paper and synthetic material are subjected to
deterioration. A previous investigation on paper filter after two years in organic soil at
Porto Tolle in Italy, carried out by the R o d i o Company, s h o w e d serious deterioration
s y m p t o m s (Hansbo, 1986). Fortunately, w h e n deterioration is beginning to affect the
discharge capacity of the drains the consolidation process is generally near its termination
and then a decrease in discharge capacity has a very small influence on the consolidation
process.

4.5 Practical aspects of vertical drainage

(i) Preloading. In order to reduce the time of preloading, a temporary surcharge (load in
excess of the design load) is often applied. T h e temporary surcharge can b e r e m o v e d
when the settlement exceeds the predicted final settlement preferably not before the
remaining excess pore water pressure is below the the stress increase caused by the
temporary surcharge. O n e has to keep in m i n d that the preloading time is governed by the
soil layer w h i c h shows the lowest rate of consolidation. B y increasing the time of tem-
porary overloading, or the size of the overload, secondary settlement can b e reduced or
even eliminated (Fig. 310).
T h e p l a c e m e n t of the fill material, representing the fill and surcharge, takes s o m e time,
which in the case of vertical drains, m a y h a v e a considerable influence on the early stage
of the consolidation process. By a simple graphical method, suggested by Terzaghi, the
consolidation curve can be adjusted accordingly (Fig. 324).
T h e stipulated time of consolidation is decisive of the drain spacing and the surcharge
to b e applied. F r o m a practical viewpoint the solution to b e chosen depends upon the
relative costs of drains and surcharge and is a matter of e c o n o m i c a l optimisation.
In places with low undrained shear strength, stability along the e d g e of a fill e m b a n k -
ment m a y b e at stake (see E m b a n k m e n t s and fill on soft soil). Stabilising loading berms
m a y b e required to avoid failure or excessive lateral m o v e m e n t s in the soil. W h e n con-
sidering the need of loading b e r m s one has to take into account that drain installation m a y
entail an overall decrease in the undrained shear strength of the subsoil.
T h e cost of loading b e r m s w h e n e v e r needed m a y add considerably to the total cost of
the vertical drainage project. B y using the so-called v a c u u m m e t h o d which is an alter-
native to preloading by the u s e of a fill e m b a n k m e n t , loading b e r m s are no longer
required. A c c o r d i n g to the v a c u u m m e t h o d an underpressure is created in the drainage
blanket and consequently also in the vertical drains by m e a n s of a v a c u u m p u m p .
442 Soil improvement

Fig. 324. Correction of the consolidation curvewith regard to gradual application of the load. Broken
curve indicates the consolidation process caused by instantaneous loading. Full line curve represents
consolidation process by gradual application of the load.

T h e main difficulty with the v a c u u m m e t h o d consists in creating an air-tight sealalong


the boundaries of the drainage blanket.
T h e v a c u u m m e t h o d is of special interest w h e r e soft soil under deep water has to b e
consolidated. Full v a c u u m w o u l d then have the s a m e effect as a surcharge represented
by atmospheric pressure plus the water pressure at the seabed. For e x a m p l e , at 10 m water
2
depth 8 0 % v a c u u m would represent sa surcharge of 180 k N / m (about 16 m of fill placed
on the seabed).

(ii) Application and limitations. According to Eq. (510), the parameters required for the
design are related to the drain characteristics and the m e t h o d of installation (disturbance
effects) on the one hand and the consolidation characteristics of the soil on the other. T h e
most important drain characteristic is the discharge capacity u n d e r long-term conditions.
T h e discharge capacity of the drain should b e high e n o u g h and the drain m u s t not b e
clogged d u e to the channel system being silted u p in the course of consolidation which
places certain requirements on the filtration of water squeezed into the channels of the
core. Finally, the strength of the drain must b e sufficient so that the drains do not b r e a k
during installation.
M a y b e even m o r e important for the prediction of the consolidation process are the soil
parameters governing Eq. (510), in particular the coefficient of consolidation in horizontal
pore water flow b u t also the hydraulic conductivity in the z o n e of smear. A correct choice
of the design parameters n o doubt represents the m o s t important part of the design of a
vertical drain installation, considerably m o r e important than the question regarding
So/7 improvement 443

whether the theoretical solution to the consolidation process is rigorous or an approximation


suitable for practical purpose.
In particular, the relation b e t w e e n prevailing effective overburden and preconsolidation
pressures represents a question of utmost i m p o r t a n c e for the standpoint w h e t h e r or not
vertical drains will b e effective in the circumstances. Should w e deal with a case where
the soil is overconsolidated to such an extent that the preconsolidation pressure is higher
than the effective vertical stress to be expected after loading, then an installation of
vertical drains can b e counterproductive. T h e disturbance caused by drain installation
can entail an increase in final settlement while the consolidation process will b e m o r e or
less unaffected by the drains.
As is w e l l - k n o w n the preconsolidation pressure is determined through oedometer
tests. T h e interpretation of its value is very sensitive to whether the effective stress is
plotted in logarithmic or linear scale. O e d o m e t e r curves of heavily disturbed samples
w h e r e a c o m p l e t e b r e a k - d o w n of the structure has occurred can give the impression
if plotted in a semi-logarithmic d i a g r a m that they are preconsolidated (Hansbo,
1979a). Therefore, a presentation in linear scale should always b e m a d e as a control of
sample disturbance. S a m p l e disturbance will give incorrect information about the
consolidation characteristics as a whole, particularly in the case of highly sensitive clays

(iii) Coefficient of consolidation. T h e coefficient of consolidation ch in horizontal pore


water flow is usually larger than the coefficient cv in vertical p o r e water flow determined
by conventional o e d o m e t e r tests. As the M value is equal in both cases, then obviously
the ratio Cj/cv = k/kv T h e ch value is thus very dependent of the p r e s e n c e of pervious
horizontal layers in the soil. K n o w i n g the M value from the results of o e d o m e t e r tests,
the ch value can be determined by measuring the permeability in the horizontal direction
of the soil sample. A n even better w a y of determining the ch value is to use the piezocone.
As is w e l l - k n o w n , during penetration of clay by the p i e z o c o n e a high excess p o r e water
pressure is induced. A fairly reliable value of the coefficient of consolidation can b e
obtained by interrupting the penetration and studying the rate of dissipation of excess
pore water pressure.
In practice, normally only the value of cv is given and therefore ch has to be estimated
by simple j u d g e m e n t based on studies of the soil structure, in particular the frequency and
thickness of pervious layers. If the spacing between such layers is small the clay between
will consolidate mainly by outflow of water in the vertical direction to the pervious
layers. Consequently, in such a case the ch value is governed by the properties of the per-
vious layers and not by those of the m o r e or less impervious clay.
For seemingly h o m o g e n e o u s soil it can generally be a s s u m e d that c}/cv = 2.

(iv) Checking the rate of consolidation. T h e m o s t logical m e t h o d of checking the degree


of consolidation achieved at a certain time w o u l d be to m e a s u r e the rate of pore pressure
444 So/7 improvement

dissipation. M a n y case histories indicate, however, that this m e t h o d m a y not b e fully


reliable. T h u s , excess pore water pressure m a y s e e m to remain although settlement has
c o m e to a complete stop. O n e difficulty met with, w h e n faced with the p r o b l e m of j u d g i n g
the degree of consolidation on the basis of pore pressure m e a s u r e m e n t s , is the lack of
k n o w l e d g e about the final groundwater situation after that consolidation has seized. In
f a c t , the groundwater level is generally following very closely the ground surface level.
Therefore, if the ground level has been subjected to changes by, for e x a m p l e , the
placement of a load e m b a n k m e n t the reference value of the pore water pressure will also
be subjected to changes.
There are certainly also other reasons w h y the result of pore pressure measurements
are s o m e w h a t doubtful as a check of the degree of consolidation attained. This does not
m e a n that pore pressure observations are meaningless as control method. However,
doubtless the m o s t feasible w a y of checking the rate of consolidation is to measure the
rate of settlement and, in addition, the increase in undrained shear strength and
preconsolidation pressure. Pore pressure observations should serve as a c o m p l e m e n t to
these control m e t h o d s .

Example 84: A clay deposit, underlain by sand at 20 m depth, is provided with vertical band drains with
3
across-sectional area of 4 mm by ,100 mm and a discharge capacity of 20 m /year. The drains are installed
in a square pattern with a drain spacing of 1.3 m. The coefficient of consolidation of the clay to a depth
2 2
of 8 m is ch = 0.5 m /year and the permeability kh = 0.03 m/year and underneath 8 m depth ch = 2.0 m /year
and kh = 0.05 m/year. Determine the excess pore pressure variation with depth 3 months after the
2
application of a a sand fill of 20 kN/m . The installation of the drains is assumed to cause a zone of smear
with a diameter of ds ~ 2.2d and a permeability ks = 0.5kh.

2
Solution: First calculate the consolidation process on the assumption that ch = 0.5 m /year throughout
2
the clay layer, and then on the assumption that ch - 0.2 m /year throughout the clay layer.
The equivalent drain diameter is:
d = 2 ( 1 0 0 + 4)/ = 0.066 m
The diameter of the zone of disturbance ds 0.15 m
The diameter of the cylinder dewatered by a drain is:
D=2/hVdK= 1.47 m
We have in the first case:
-3
Pi = ln( 1.47/0.15) + 21(0.15/0.066) - 0 . 7 5 +(20 - )0.03/20 = 3.174 + 4 . 7 U ( 2 0 - ) 1 0
and in the second case:
3
2 = 3.174 +(20 - )0.05/20 = 3.174 + 7.85z(20 - )10-
Eq. (512) and Fig. 318 yield the following values of Uh with depth:
z, m: 0 1 2 3 4 5 6 7 8
U%: 13.5 13.2 12.9 12.6 12.4 12.3 12.1 12.0 12.0
z, m: 8 9 10 11 12 13 14 15 16 17 18 19 20
Uh%: 37.5 37.4 37.3 37.4 37.5 37.8 38.2 38.8 39.5 40.4 41.4 42.9 44.1
At 8 m depth Uh can be taken as the average ofg 12.0 and 37.5, i.e. Uh = 24.8%
So/7 improvement 445

Example 85: A 30 m thick clay layer, underlain by sand, shall be consolidated for a design load of 100
2
k N / m in one year. For the purpose band drains of good quality, 100 mm in width and 4 mm in thickness,
shall be installed with the drain tips penetrating into the sand layer blow the clay. For preloading, gravel
3
fill is available with a density of 1.8 t/m . The cost of drain installation, including the drain material,
has been offered at a price of l/m drain and for placement and removal of temporary fill at a price of
3
8 /m of fill material. The drainage layer, 0.5 m in thickness, to be placed on top of the clay surface
2
is offered at 5/m . Determine the optimal drain spacing for the project, if the clay has the following
2
characteristics: coefficient of consolidation ch = 1.5 m /year; permeability kh = 0.05 m/year. The discharge
3
capacity of the drains can be estimated at qw =500 m /year. The installation is assumed to cause a zone
of smear with a diameter of ds - 22d and a permeability ks = k/3.

Solution: The equivalent drain diameter is 0.208/ = 0.066m. Assuming that the drains are installed in
an equilateral triangular pattern, x m in spacing, the diameter of the dewatered cylinder becomes equal
to D = 1.05JC. Assuming further that the settlement during consolidation is directly proportional to the
size of the load and the average degree of consolidation U, we have:

J_ <i + i
=
A<
u
2
where spq = primary consolidation settlement under the load q = 100 kN/m
Aq = temporary overload
Obviously, for the condition U = 100% to be fulfilled it is required that Umin = q/(q + Aq). Since the
lowest consolidation degree is obtained in the middle of the clay layer we have to choose in our analysis
of the effect of well resistance = / = 15 m as given by Eq . (510).
The analysis gives the following result:

Drain spacing, m 1.1 1.2 1.3 1.4 1.5

uh% 91 86 81 75 69
2
Temporary surcharge Aq, kNm 10 16 24 33 45
Temporary fill, m 0.6 0.9 1.3 1.8 2.5
2
Drain costs, / m 29.6 24.8 21.2 18.3 15.9
2
Cost of fill, / m 4.8 7.2 10.4 14.4 20.0
2
Cost of drainage layer, / m 5.0 5.0 5.0 5.0 5.0
2
Total cost, / m 39.4 37.0 36.6 37.7 40.9
The most cost-effective drain spacing is found to be 1.3 m.

Example 86: Determine the average degree of consolidation, 9 months after load application, by Eq.
(510), on the one hand, and by Eq. (512), on the other, for a thick clay deposit provided with vertical
band drains, 100 mm by 4 mm, with a spacing of 1.0 m in equilateral triangular pattern. The drains have
a high discharge capacity and, therefore, well resistance can be ignored. The coefficient of consolidation,
2
as determined by oedometer tests, is cv = 0.5 m /year. A fill with large extension has been placed on the
2
ground representing a load q = 30 kN/m .

Solution: In the analysis, based on Eq. (510), the value of ch can be assumed equal to 2 c v. The diameter
of the zone of disturbance ds can be assumed equal to 22d and the ratio khlks = 2. In the analysis, based
on Eq. (513), can beassumed equal to 0.75c v.

The diameter D=2^JO.5/3/K = 1.05 m


(1) Eq. (510) yields: = ln(1.05/0.15) + 21n(0.15/0.066) - 0.75 = 2.838
2
Moreover, the time factor TH = 1.0-0.75/1.05 = 0.680, and hence:
Uh = 1 - exp(-80.680/2.838) = 0.85 (85%)
446 So/7 improvement

(2) The value of a, for Did = 1.05/0.066 =16, according to Fig. 317, is 0.24. Eq. (513) yields:

, = - + A JILy* = _( 1 + ' ' = ,.85 (85%)


8 0 7 5 0 5
f ^ ^ y

aD\DyJ 1 2 - 0 . 2 4 - 1.05 V 1.05-9.81

4.6 Case histories

A large n u m b e r of case histories h a v e been publicised [see, for e x a m p l e , the special issue
of Gotechnique of M a r c h 1981 on vertical drains and the publications by Lo ( 1991 ) and
H a n s b o (1992)]. Two case histories of special interest will be presented here: the test field
at Sk-Edeby, Sweden, and the test field at Tianjin harbour, China.
(i) Test field at Sk-Edeby, Sweden. T h e Sk-Edeby test field, located about 25 k m west
of Stockholm, was established in 1957 and is one of the oldest monitored test fields for
the study of vertical sand drain and band drain behaviour. It contains five circular test
areas, three of which provided with sand drains , 0.18 m in diameter, with drain spacings
varying from 0.9 to 2.2 m, and one with band drains, type Geodrain, with drain spacing
equal to 0.9 m. In order to distinguish the effect of vertical drains from the effect of one-
dimensional consolidation, one of the areas was left undrained. T h e test areas were
2
loaded with 1.5 m gravel, corresponding to a load of 27 k N / m except for test area III
2
which was loaded up to 2.2 m gravel (36 k N / m ) T h e geology at the site and the geotechnical
properties of the different test areas w e r e described in detail by H a n s b o (1960). T h e
results obtained h a v e been presented in several publications (Holtz & B r o m s , 1972;
Larsson, 1986; H a n s b o , 1987).
In short, for the clay b e t w e e n 2.5 and 7.5 m depth, selected for study in order to avoid
the influence of the dry crust and of irregularities in soil characteristics near firm bottom
(at 8 - 1 5 m depth), the coefficient of consolidation c v determined by oedometer tests

C o n s o l i d a t i o n time
months years

Fig. 325. Consolidation process for the 5 m thick clay layer between 2.5 and 7.5 m depth.
So/7 improvement 447

2
varied from 0.13 to 0.22 m / y e a r . T h e compression ratio CR [= Ccl(\ + e0)] varied from
0.43 to 0.56. T h e coefficient of consolidation in horizontal pore water flow ch, w h i c h was
determined in an o e d o m e t e r test w h e r e drainage was allowed only through a central drain,
2
was 0.7 m / y e a r . T h e corresponding permeability value kh ~ 0.03 m/year.
T h e settlement vs. time relationship for the test areas with sand drains and n o drains
is presented in Fig. 325.
T h e degree of consolidation in the undrained Area IV and in Area III which was
unloaded is not as well-defined as for Areas I and II w h e r e w e h a v e entered into a stage
of secondary consolidation. All the same, a good estimate can b e m a d e on the basis of the
shape of the settlement curve. N o w the discharge capacity qw of sand drains, 0.18 m in
3
diameter, can be estimated on the basis of sand permeability at about 100 m /year. Assuming
a zone of smear ds - Id = 0.36 m and a ratio kh lks = 4 (equal to ch lcv a v ) an acceptable
agreement is obtained if a successively decreasing value of ch is applied as s h o w n in Table
38. T h e Uh values given in Table 38 are determined from the m e a s u r e d values of Uioi and
Uv according to Carillo's equation. T h e values represent the corresponding coefficients
of consolidation assuming non-validity of D a r c y ' s law, Eq.(513).

TABLE 38.
2
Best fit of ch and values in Areas I, II (load 27 k N / m ) .

Drain spacing 0.9 m 1.5 m 2.2 m


Time, years 1/6 1 2 1/2 1 4 1 4 9
flot. % 34 80 94 36 60 95 33 69 93
uv,% 1 10 19 5 10 34 10 34 51
uh.% 33 78 92 33 56 92 25 53 85
2
ch m /year 0.85 0.53 0.45 0.89 0.91 0.70 0.75 0.50 0.55
2
A, m /year 0.07 0.06 0.07 0.14 0.16 0.20 0.16 0.12 0.19

According to Eq. (513) the consolidation process depends on the m a g n i t u d e of the load
placed on the ground. This is not the case with Eq. (510). In test area the load was 36
2
k N / m . F r o m the test results in area III w e find the best agreement b y choosing the values
given in Table 39.

TABLE 39.
2
Best fit of ch and values in Area III (load 36 kN/m ).

Drain spacing 1.5 m


Time, years 1/6 1/2 1 2
ftot.* 20 42 65 87
Uy.% 1 5 10 19
uh,% 19 39 61 84
2
ch m /year 1.4 1.1 1.0 1.0
2
A, m /year 0.18 0.16 0.17 0.21
448 So/7 improvement

J
-1 ti-

I i
li
Loading time, years
0 1 2 3 4 5 6

\
\

Points of >hserv Hion

p H ~~ 0 6 tn ^

t;OP settiei lent = 0.9 m

Fig. 326. Observed compression of clay layer situated between depths of 2.5 and 7.5 m in Test area V
at Sk-Edeby. Band drains, type Geodrain, with 0.9 m spacing. Broken line represents theoretical
2
correlation based on Eq. (515) with = 0.15 m / year.

T h e values are m o r e consistent than the ch values throughout the consolidation process.
T h e values w h i c h include the influence of disturbance are very close to the cv values
2
obtained in the oedometer test ( 0 . 1 3 - 0 . 2 2 m / y e a r ) , except at a drain spacing of 0.9 m
w h e r e they are considerably lower. T h e ch values are higher at the beginning of the
consolidation process than at the end but are fairly close to the coefficient of consolidation
2
ch determined in the laboratory (0.7 m / y e a r ) .
T h e result obtained in the test area provided with b a n d drains, 100 m m in width and
4 m m in thickness, is presented in Fig. 326.
T h e consolidation process observed in the undrained area on the basis of settlement
indicates that the coefficient of consolidation c v in field condition is m u c h higher than
the o e d o m e t e r value of cv in the beginning of the consolidation process. T h e rate of excess
pore pressure dissipation, however, observed after 14 and 2 5 years of loading time
(Larsson, 1986), is in good agreement with the cv value determined by o e d o m e t e r tests.

(ii) Pilot test at Tianjin harbour, China. T h e pilot test at Tianjin harbour is interesting from
the point of view that it m a k e s possible a direct comparison of the v a c u u m m e t h o d to the
conventional m e t h o d of using an overload to p r o d u c e consolidation. T h e test is described
in detail by C h o a (1989) and here only a short s u m m a r y will b e presented.
T h e v a c u u m m e t h o d w a s first successfully applied in a full-scale test b y the S w e d i s h
Geotechnical Institute (SGI, 1949). T h e idea behind the m e t h o d is to p r o d u c e a p o r e water
underpressure in the drains instead of an excess pore water pressure in the surrounding
soil. In b o t h cases a hydraulic gradient in the radial direction t o w a r d s the drains is created
followed by consolidation.
So/7 improvement 449

Fig. 327. Observed settlement vs. time relationship in the Tianjin pilot test.

According to the v a c u u m m e t h o d the surface drainage blanket in contact with the


drains is sealed by m e a n s of an air-tight m e m b r a n e and connected to a v a c u u m p u m p . T h e
consolidation effect depends upon h o w effectively the v a c u u m procedure can b e realised.
In the Tianjin pilot test, band drains manufactured in C h i n a w e r e utilised. T h e s e have
the conventional dimensions, 100 m m by 4 m m , and consist of a plastic core enclosed in
a filter sleeve of geotextile material.
T h e pilot test comprised five test areas, 5 0 m by 50 m each. Out of these, four areas
w e r e provided with drains while o n e was left without drains in order that the effect of the
vertical drains could be discerned. Three of the areas w e r e loaded conventionally by an
2
earth-fill e m b a n k m e n t representing a load of 97 k N / m . T h e drain spacings in these areas
were 1.3 m , 1.5 m and 1.8 m , respectively. T h e area placed under v a c u u m was provided
with drains, 1.3 m in spacing.
T h e soil to b e consolidated at the site consists of about 4 m of hydraulic fill underlain
by about 20 m of soft, silty clay with intermediate layers of clayey, sandy loam. T h e
undrained shear strength was found to increase almost linearly from about 5 kPa at 3 m
depth to about 30 kPa at 20 m depth. According to the results of o e d o m e t e r tests on
samples taken in the silty clay the compression index Cc was estimated at 0 . 3 5 - 0 . 5 (CR
estimated by the Author at about 0.2) and the coefficient of consolidation cv at 1.3-2.5
2
m /year.
T h e settlement vs. time of consolidation obtained in the test areas with 1.3 m drain
spacing, using a surcharge on the one hand and v a c u u m on the other, is shown in Fig. 327.
As in the previous case record a good agreement is obtained between measured
2
consolidation process and that calculated by Eq. (515) a s s u m i n g = 1.5 m / y e a r .
In conclusion w e find that the v a c u u m m e t h o d has w o r k e d very satisfactorily. It has
the advantage over conventional preloading of being able to be used in areas with very
low bearing capacity without causing stability problems.
450 So/7 improvement

5. GRANULAR COLUMNS

5.1 Aim and application

Large-diameter sand drains are often considered to act as reinforcing and stiffening
elements in the soil which contribute to reducing settlement. However, in the case of
ordinary sand drains, which are usually in a loose state, this effect can b e disregarded. T h e
purpose of granular columns is both to increase stability and to r e d u c e settlement. They
also serve as drains and, consequently, contribute to speeding u p excess p o r e pressure
dissipation. D u e to dilation characteristics and drainage capacity, granular columns are
also used as a m e a n s to reduce liquefaction potential.
T h e most c o m m o n types of granular columns are so-called stone c o l u m n s and sand
compaction piles. A m o n g the installation methods utilised, the cased borehole method,
the J a p a n e s e vibro-composer method and vibro-replacement m e t h o d are the most
c o m m o n ones.
According to the cased borehole method, gravel and/or sand is filled into the casings
and compacted during the withdrawal of the casing by m e a n s of a 1 . 5 - 2 1 gravity h a m m e r
dropped from a height of 1 to 1.5 m. T h e granular material can also b e c o m p a c t e d by re-
driving the casing several times during withdrawal.
By using the vibro-composer method, large-diameter sand c o m p a c t i o n piles, up to 2
m in diameter, can b e produced. T h e sand c o m p a c t i o n piles are constructed by driving
the casing to the required depth by m e a n s of a heavy vibratory h a m m e r , located at the top
of the casing. During the driving operation, the casing is closed at the b o t t o m by means
of a sand plug. T h e casing is filled with a specified v o l u m e of sand and the casing is then
repeatedly extracted and partially redriven by m e a n s of the vibratory hammer, starting
from the bottom, in order to c o m p a c t the sand. T h e operation is repeated until a well
c o m p a c t e d sand column is obtained.
T h e vibro-replacement method is mainly applied in silt soils. T h e granular columns are
created in a w a y similar to vibroflotation (see p . 415). T h e process can b e either wet or
dry. In the wet process, a hole is formed in the ground by jetting the vibroflot cylinder
down the the required depth with water. W h e n the cylinder is w i t h d r a w n a hole is formed
in the ground with a larger diameter than that of the cylinder. T h e uncased hole is filled
with gravel and densified by vibroflotation.
In the dry process no water jetting is used when the hole is formed by the cylinder. T h e
dry process can only b e used where the uncased borehole can stand open without the
support of water.

5.2 Design aspects

T h e design principles regarding granular columns installed in clay can be s u m m a r i s e d as


follows.
So/7 improvement 451

(i) Load share between columns and surrounding soil. A s in the case of vertical drains
the piles are generally installed in a triangular or square pattern. T h e contributory area of
the soil surrounding the granular pile can b e approximated to a circular cylinder with a
cross-sectional area equal to the area A enclosed b y four neighbouring granular piles (see
p. 4 3 2 ) , i.e. the diameter of the cylinder D - lyjA In .
A s s u m i n g that the diameter of the granular c o l u m n is d, the area r e p l a c e m e n t ratio can
2
b e expressed as as = (d ID) . T h e load share between the granular c o l u m n s and the
surrounding clay is generally calculated on the assumption that horizontal sections in the
ground remain horizontal in the course of settlement ('equal strain' theory). A s s u m i n g
that the normal vertical stress in the granular c o l u m n is 1 and in the surrounding soil
a s o i l, t h e strain in the vertical direction is then obtained by the relation:

= 01 / M c ol = a s o il / M s o il (514)

w h e r e Mcol = the compression (oedometer) m o d u l u s of the granular c o l u m n ,


M s o il = the compression m o d u l u s of the clay.
A s s u m i n g further that the load placed on the ground is q w e have:

a ) ( 5 1 5
^ t f c o l ^ + ^ s o i r O ~ s)

Introducing m = McoXIM^OX w e find:

m
q
a ml = (516)
1 + (m - \)as

)
tfS0il =
1 + (m - \)as

(ii) Bearing capacity. T h e ultimate load of granular c o l u m n s in clay depends u p o n the


internal angle of friction of the column material and the shear strength of the surrounding
clay. In principle, the failure condition can b e c o m p a r e d with that occurring in the triaxial
test w h e r e the radial, confining pressure ar is governed b y the horizontal overburden
pressure and the undrained shear strength cu of the clay. A s the top part of the ground
is generally weathered (dry crust formations) with high shear strength, failure will
normally take p l a c e at depth in the form of bulging within the z o n e w h e r e the shear
strength of the soil is at its m i n i m u m .
In the case of cylindrical expansion the radial stress at failure orj, according to the
theory of plasticity, can b e expressed by the relation:
452 So/7 improvement

crf = al0 + cu[\ + ln (518)


2cu(\+vc)

w h e r e c = elastic m o d u l u s of the clay,


vc = Poisson's ratio of the clay.
By experience w e k n o w that the elastic m o d u l u s of clay is in the r a n g e of 1 5 0 c u - 5 0 0 c u
which, in undrained condition (vc = 0.5), yields Grjvalues between + 5cu and +
6cu. In m o s t cases in practice it can be assumed that the governing, effective confining
pressure at failure is equal to "' r /= c r ^ ) + 5cu. Introducing this value into the Mohr-
C o u l o m b failure criterion (pp. 278-279), we find the ultimate, effective column stress <fjQ0\
from the relation:

1 + sin0'
(519)
1 -sin0
/

w h e r e '= the internal angle of friction of the granular column.


T h e thickness of the w e a k layer governing the occurrence of failure can b e estimated
at 2 - 3 times the diameter of the granular column.
If the granular columns are not penetrating the clay layers (floating c o l u m n s ) , their
bearing capacity can b e determined in a way similar to that of floating piles (p. 189).
A normal zone failure of the top part of the granular column and surrounding soil may
occur in cases w h e r e the soil surface is in a loose state. This p r o b l e m can be treated on
a theoretical basis by the application of the bearing capacity formulae given for shallow
foundations.

Example 86: In a full-scale test on a granular column-improved clay soil at the campus near
Bangkok (Bergado et ai, 1992), the bearing capacity of individual columns, 0.3 m in diameter, was
determined on five groups of columns, comprising three columns each. The clay characteristics are
given in the figure below.

Undrained shear
Consistency limits, % U n i t weight]
3 s t r e n g t h cu, kPa
Soil l a y e r s Natural water content, % [kN/m
(field v a n e test)
20 40 60 80 100 15 16 10 20 30 40
B r o w n to r e d d i s h
brown weathered
clay

D a r k g r e y soft c l a y
often w i t h d e c o m -
posed w o o d and
sandy seams
Soil improvement 453

The following results were obtained.

Group: 2 3 4 5
Friction angle (average) 38.2 36.9 35.6 37.7 43.3
Ultimate loads, kN:
Measured (average): 32.7 30.2 21.3 30.7 35.6

Determine theoretically the bearing capacity of the column groups.

Solution: The bearing capacity is governed by the clay layer with the lower shear strength values. The
mean value of cu based on the shear strength values obtaining from 2.5 to 3.5 m depth (length 3d) is
cu av = 16.5 kPa. The effective overburden pressure at 3 m depth, at an average groundwater level of 1.2
m below ground surface, is about 30 kPa. Assuming K0 = 0.5 we have 15 kPa.
For column group 1, Eq. (519) yields:
e e , 1+sin38.2 10
eyed = (15 + 5 16.5)- = 413.5 kPa
l-sin38.2
from which the ultimate load:
2
u lt = 4l3.5-7t-0.3 /4 =29.2 kN
The result of the analysis for all the column groups is summarised below.

Group: l 2 3 4 5
Ultimate loads: 29.2 27.6 26.1 28.6 37.0

The measured mean value of all the tested columns is 30.1 kN while the calculated mean value is 29.7 kN.

(iii) Slope stability improvement. Granular columns are also installed for the purpose of
increasing slope stability conditions. In this case failure is assumed to take place along
a curved (normally circular-cylindrical) failure surface in which the shear strength of the
columns and the surrounding soil is fully mobilised. With regard to the fact that the stress-
strain behaviour of the columns and the natural soil up to failure may be completely
different from eachother this assumption is very questionable.
According to a method of analysis of a granular column improved slope, suggested by
Chambosse & Dobson (1982), named the lumped parameter method, the driving and
stabilising forces are first determined for the natural, unimproved soil (see Section on
slope stability). The additional effects of the granular columns on driving and stabilising
forces are then calculated and added to the respective values of unimproved soil. T h e
safety factor is the sum of the stabilising forces divided by the driving ones.

(iii) Settlement. As mentioned above, granular columns, like vertical drains, speed up the
primary consolidation process induced by loading. For fully penetrating granular
columns the final primary consolidation settlement sp is simply given by the relation:

h 2 soil ,
(520)
M, soil A f i e o i [l l + ( m - l ) f l j]
454 So/7 improvement

where = the overall vertical stress increase at depth due to the load q,
h = depth of the clay layer (length of the columns),
M s o i lz = the compression modulus of the soil at depth z.
Alternatively, the settlement is calculated as:

h A a
sp = l ^ d z = (521)
J J
" 0 M z c ol 0 Mzwl[l + {m-l)as]

If the extension of the loaded area is large in comparison with the thickness of the clay
layer (spread load), can be assumed equal to q.
In the case of partially penetrating columns, the compression of the unstabilised soil
layer underneath the tips of the columns will, of course, have to be added to the settlement
given by the above relations.
The course of settlement can be determined in a similar way as was done for vertical
drains. However, in this case the diameter of the columns is generally too large for using
the simplified solution given for small-diameter vertical drains. Assuming that the
installation of the granular columns causes a zone of disturbance with reduced permeability
ks and a diameter ds (as in the case of vertical drain installation) a more correct solution
of the consolidation process will take the form (cf. Eq. 510):

Uh = l - e x p ( - ) (522)

2
where Th = ch t/D ,

kh 1 1 kh 1
= - T - r O n - + f \ns-0.75 + - - ) + (21 - ) (I - - J ,
-l s ks An qw
= Did,
s= ds/d.

2
Example 87: Determine the primary consolidation settlement due to a spread fill of 25 kN/m on a 10
m thick, granular column-improved clay layer if the compression modulus of the clay M c l ay = 300 kPa.
The granular columns, which are 0.5 m in diameter and have a compression modulus M c ol = 2500 kPa,
are placed in an equilateral triangular pattern with a spacing of 1.3 m. Determine also the time required
to reach half the final settlement if the coefficient of consolidation of the clay in horizontal pore water
2
flow is 0.5 m /year. The well resistance of the granular columns and the zone of smear can be neglected.

2 1 / 2
Solution: The diameter of the zone of influence of a granular pile is D = 2- ( 1 . 3 / 3 / 2 ) / JK - 1.365
2
m, whence as = (0.5/1.365) = 0.134. Furthermore we have m = 2500/300 = 8.33.
The primary settlement thus becomes equal to:
25-10
So/7 improvement 455

The time required to reach half this settlement (U = 50%) can be estimated by Eq. (522). For a value
of = Did = 1.365/0.5 =2.73 we have:
= [2.73 2/(2.73 2 - l)][ln(2.73) - 0.75 + 2.73~ 2 - (42.73 4)- 1]= 0.443
whence:
Th = - ln(0.5)0.443/8 = 0.0384
which yields:
t = 0.0384 1.365 2/0.5 = 0.143 years - 50 days

6. STABILISATION BY A D M I X T U R E S

6.1 Superficial stabilisation

Admixtures are often used for the purpose of increasing the bearing capacity of the
ground surface or of the bottom of excavations in soft soil conditions. The most common
additives utilised in clay soils are lime and cement which bring about a strength increase
by chemical reactions. Other common additives are bitumen and, for preventing leakage,
bentonite.
The stabilisation caused by cement and lime is achieved by ion exchange. The effect
of ion exchange can be easily observed, for example by mixing common salt (sodium
chloride) with a remoulded quick clay. The consistency of the clay will convert from a
liquid state into a plastic state. Thus the salt has drastically changed the shear strength as
well as the sensitivity of the clay. On the other hand, mixing sodium pyrophosphate with
clay of normal sensitivity will convert the clay from a plastic or viscous state into a liquid
state (the clay.has turned into a quick clay).

Fig. 328. Water saturated soil with a high silt content easily turns into a liquid state. The shear strength
of the soil becomes strongly reduced.
456 Soil improvement

T h e influence of ion e x c h a n g e on the undrained shear strength of clay varies with the
type of ions involved. T h e best effect is reached by a l u m i n i u m ions, while the effect of,
for e x a m p l e , calcium ions is quite limited. T h e i m p r o v e m e n t obtained by the use of
admixtures is very m u c h dependent on the efficiency of the mixing process.
Briefly, lime is utilised for stabilisation of clay and silt (Figs. 3 2 8 - 3 2 9 ) . T h e favourable
effect of adding lime is caused by ion e x c h a n g e in the first place and, in the second place,
by cementation. T h e shear strength is improved and the plasticity index of the soil and
the o p t i m u m water content are increased. T h e best effect is obtained by m e a n s of
quicklime.
Cement is used as a m e a n s of stabilising practically any type of inorganic soils. T h e best
result is obtained in well graded soils with less than 5 0 % fines and a plasticity index IP
< 2 0 % . Sand and cement should be mixed dry of o p t i m u m , clay and cement s o m e w h a t
wet of optimum. T h e soil-cement mixture should be kept moist during curing. T h e
increase in shear strength is speeded up by high temperature. Certain additives, such as
lime, calcium chloride and, even better, sodium sulphate i m p r o v e the result and are cost-
effective. T h e a m o u n t of cement needed is about 5 - 1 0 % (by weight) in gravel, 7 - 1 2 %
in sand, 1 2 - 1 5 % in silt and 1 2 - 2 0 % in clay.
T h e resulting shear strength varies from about 350 kPa in silty clay to about 10 M P a
in well grade gravel.
Bitumen (for e x a m p l e asphalt and tar) is used mainly for stabilisation of sand and
gravel. Asphalt, which is most c o m m o n l y used, binds the particles together and prevents
water absorption (important in the case of swelling material). T h e u s e of proper additives,
such as phosphorus pentoxide, will increase the shear strength (an admixture of 2 %
phosphorus pentoxide + 5 % asphalt was found to p r o d u c e a threefold increase of the shear
strength of silt).

Fig. 329. The superficial soil shown in Fig. 328 is treated with a lime admixture. Due to ion exchange,
and to the capacity of binding surplus water exerted by lime, the ground surface becomes firm and
obtaines a high bearing capacity.
So/7 improvement 457

6.2 Lime and/or cement columns

A stabilising effect on soft clay and silt to great depth cah b e attained by forming soil
columns m i x e d with lime or c e m e n t or armixture of lime and cement. G y p s u m is also used
in certain conditions. In the following, the c o l u m n s w i l l b e referred to as lime c o l u m n s ,
irrespective of the admixture utilised.
L i m e c o l u m n s are utilised for stabilisation purposes (for instance of trench walls) and
for the p u r p o s e of reducing and speeding u p consolidation settlement. T h e pattern of
installation will h a v e to b e chosen with regard to the a i m of stabilisation.

(i) Installation techniques. In E u r o p e , m a c h i n e s for lime c o l u m n installations exist which


can p r o d u c e c o l u m n s to a depth of normally about 15 m and with a c o l u m n diameter of
0.5 m. T h e design and efficiency of the mixing tool is very important for the result
achieved.
T h e e q u i p m e n t developed by Fundator, Finland, can b e taken as an e x a m p l e . This is
a caterpillar with a weight of 3 6 1 and a contact pressure of about 30 kPa (Fig. 330). It is
provided with t w o containers, each with a v o l u m e of 3 m 3 , for storage of lime and/or
cement. T h e installation of the columns can b e performed to a depth of 25 m with a
column diameter varying from 0.12 to 1.0 m. T h e rotational speed of the m i x i n g tool (Fig.
331) can be varied from 25 to 2 1 0 revolutions/min and the speed with which it is
withdrawn can b e varied ffom nil to 25 mm/revolution. T h e c o l u m n s can b e raked at an
angle of m a x i m u m 4 5 . T h e installation is carried out by the aid of a data processing
system based on the results of test c o l u m n installations. F r o m the results, the pressure in

Fig. 330. Equipment for installation of lime columns provided with two containers for storage of lime
and/or cement.
458 So/7 improvement

the container, the feed pressure, impeller delivery, type of mixing tool, a m o u n t of
admixture to be used, etc. are settled. For each column that has been installed a check tape
is obtained from a data printer. This m a k e s it possible to j u d g e whether or not a
readjustment of the installation technique is required and if extra c o l u m n s will have to
be installed.
In Japan a so-called deep-mixing method has been developed, comprising up to eight
mixing units, by which lime or cement columns can b e created to a depth of m a x i m u m
4 0 m. T h e mixing units overlap and in this w a y a cross-sectional area of the columns can
be obtained of u p to about 9.5 m 2 , depending upon the n u m b e r and diameter of mixing
units. Under-water installations can also be m a d e .

(ii) Choice of admixture. T h e type of admixture utilised should serve the special purpose
settled for the project. T h e choice is generally based on laboratory investigations in which
various amounts of lime, cement or g y p s u m are m i x e d with the soil to be stabilised.
However, it has to be recognised that uncritical j u d g m e n t of laboratory results generally
entail an overestimation of the results to be expected from field installations. A main
difficulty met with in field conditions is to check the in situ characteristics of the columns
installed, such as shear strength, homogeneity, deformation m o d u l u s and permeability.
At present the control of the results achieved is generally performed with the aid of a
special sounding tool w h o s e penetration resistance is m e a s u r e d w h e n pressed down in
the centre of the column. However, by this m e t h o d only a general view of the shear
strength and of the h o m o g e n e i t y of the columns can be obtanied with reasonable degree
of accuracy. Deformation moduli and permeability characteristics are generally appreciated
only on the basis of laboratory tests.
T h e admixture to be used depends on the soil type. Q u i c k l i m e has a better effect than
slaked lime but is problematic to handle. A mixture of cement and lime results in a higher

Fig. 331. Mixing tool for the production of lime or cement columns
So/7 improvement 459

strength and lower compressibility than is achieved with lime alone. G y p s u m improves
the growth of strength in organic soils. In silt with a clay content below 2 0 % , it is
r e c o m m e n d e d to add fly-ash. T h e o p t i m u m lime or lime/cement content increases with
increasing plasticity of the soil. Normally, a content corresponding to 6-8 wt.% of the soil
is sufficient. T h e strength increase of the stabilised soil takes p l a c e very rapidly in the
beginning but slows d o w n with time. T h e increase m a y continue for years.
T h e texture of lime stabilised soil b e c o m e s grainy and reminds of dry crust clay. T h e
columns are often quite i n h o m o g e n e o u s depending upon insufficient mixing. At the tip
of the c o l u m n s a r e m o u l d e d zone of soil will b e formed by the mixing tool, m a y b e without
i m p r o v e m e n t effects being achieved.

(iii) Design with regard to failure. T h e bearing capacity of lime or c e m e n t columns


depends on the shear strength in planes of w e a k n e s s that are m o s t likely to exist in the
columns. T h e ultimate load is generally calculated on the assumption that failure takes
place in drained condition. T h e angle of internal friction in lime c o l u m n s , determined on
a laboratory scale, has been found to vary between 30 and 40. A s s u m i n g that the column
has the strength characteristics </>'= 30 and c - 0 and that the creep strength (p. 34) of
the surrounding clay is c c r c l a ,y the bearing capacity of a lime c o l u m n , according to Eq.
(519), b e c o m e s equal to:
tf/col = ( / +
3 c r clay)
5 c ( 5 2 )3

If c o l u m n failure takes place in undrained condition w e h a v e , Eq.(364):

/\ = r0a + $ccr clay + ^ccr col (524)

Fig. 332. Lime columns installed for stabilisation of trench walls. Lime/cement columns used for
stabilisation of the sides of deep excavations should be installed in several parallel rows.
460 So/7 improvement

where c c r c ol = the creep strength of the column.


When columns are placed in groups, local shear failure or block failure may take place
in a similar way as was described for floating pile groups.

(iv) Stabilisation of slopes and trench walls. Often the purpose of the columns is to
stabilise slopes or trench walls (Fig. 332). In the latter case, they are then generally
installed in a number of rows on both sides of the trench. The stability of the trench is
analysed on the basis of the strength properties of the untreated soil and those of the
columns. Failure is assumed to take place along a plane, or curved, failure surface in
which the shear strength of the columns and the surrounding soil is fully mobilised. With
regard to the fact that the stress-strain behaviour of the columns and the natural soil up
to failure may be completely different from eachother this assumption is very questionable.

(v) Design with regard to settlement. The settlement due to loading of a lime column
stabilised soil is analysed in the same way as previously described for granular column
stabilised soil, Eqs. (520-522). In the case of lime columns, however, it is necessary to
check that the stresses in the column is below its failure load determined by its creep
strength. The part of the load that induces stresses in the columns exceeding its creep
strength will have to be carried by the untreated soil.
The most difficult problems of settlement analysis are mainly related to difficulties of
determining the true value of the creep strength the compression modulus and the
permeability of the columns. By experience, the compression modulus of lime columns
is often assumed equal to (75 25) times the shear strength ^ the column. The permea-
bility is generally assumed equal to 100-1000 times the permeability of clay.
An illustrative example of the scatter of values to be expected regarding the properties
of lime/cement columns was reported by Pramborg & Albertsson (1992). In this case the

TABLE 40.
Characteristics of stabilised soil determined by different methods.

Measured value minimum maximum average

Permeability, m/year:
1
(Pressure-permeameter) 1.6 ll 4.4
Shear strength rf, kPa (scatter 300%):
Pressuremeter test 218
Field vane test 104
Lime column penetrometer 305
Unconfined compression test (after 28 days) 204
Compression modulus, MPa:

Pressuremeter (M 3 Epr) 21.9


1
Two extreme values out of sixteen omitted (170 and 30 m/year)
So/7 improvement 461

columns w e r e a mixture of 30 g quicklime and 36 g cement per kg of the soil [soft,


normally consolidated clay and with the characteristics: cu ~ 5 - 1 5 k P a (av. 9 kPa); w ~
6 0 - 9 5 % (av. 6 5 % ) ; w L - 4 0 - 7 5 % (av. 4 8 % ) ; St - 1 5 - 3 5 (av. 26); Af 1 4 0 - 4 0 0 kPa (av.
2 2
260 kPa); c v 0 . 3 - 1 m / y e a r (av. 0.6 m / y e a r ) ] . L i m e / c e m e n t c o l u m n s w e r e installed
with diameters 0.5 m in o n e test field (spacing = 1.3 m ) and 1.0 m in another (spacing 1.8
m). Two months after installation the columns were tested in situ with regard to permeability
(by m e a n s of the pressure-permeameter), shear strength (by pressuremeter tests, field
vane test and lime c o l u m n penetrometer) and the deformation characteristics (by
pressuremeter tests). T h e shear strength of soil, stabilised with an admixture of 25 g lime
and 35 g c e m e n t per kg of natural clay, w a s determined at certain intervals after mixing
in the laboratory by unconfined compression tests. T h e test results obtained are summarised
in Table 40.
A s s u m i n g that the compression m o d u l u s determined on the basis of the pressuremeter
test is representative of the true compression characteristics of the column w e find the
best agreement with the empirical relations:
M c o l ~ l O O ^ f o r the unconfined compression test
M c o l ~ 7 0 i ^ f o r the lime column penetrometer
Mcol ~ l O O ^ f o r the pressuremeter
T h e results of the field v a n e tests which w e r e carried out in the centre of the columns
and are m o s t probably not representative because of insufficient mixing of clay and
additives.
F r o m the coefficient of consolidation and the compression m o d u l u s of the clay w e find
an average value of the permeability c l ay ~ 0.023 m/year. T h u s the ratio of column
permeability to clay permeability is:
2 00
^col^clay^
In this case, the specific discharge capacity qw of a column, 0.5 m in diameter, is found
3
to b e only about 0.9 m / y e a r , and of a column, 1 m in diameter, about 3.6 mVyear. Thus,
in the analysis of the course of settlement of a soil, stabilised with lime/cement columns,
the influence on the consolidation process of well resistance qw cannot be neglected. On
the other hand, the installation of the columns is m a d e in a way which does not normally
give rise to disturbance (ds - d).

2
Example 88: In the test field just described, a fill of 1.5 m sand and gravel, corresponding to 27 kN/m was
placed on the ground surface after the installation of the lime/cement columns. Determine the settlement
in the test field in the two cases, ( 1 ): 0.5 m columns with 1.3 m spacing and (2): 1.0 m columns with 1.8
m spacing. The groundwater level is at 1 m depth. The lime/cement columns are assumed to have the
characteristics given by the pressuremeter tests. Average values can be applied except in the determination
of column stress leading to creep failure of the column. This is governed by the minimum shear strength
of the clay, 5 kPa, at 2 m depth.

Solution: First calculate the final primary settlement.


2 2
Case 1: We have = 21.9/0.26 = 84 and as = (0.5 /4)/ ( 1 . 3 / 3 / 2 ) = 0.134. This yields:
462 Soil improvement

1 = 84-27/(1 + 83-0.134) = 187 kPa


Before we carry out the settlement calculation we have to check whether the column stress 189 kPa
exceeds the stress leading to creep failure of the columns. The creep strength of the column material can
be estimated at 0.6Xj 130 kPa and of the clay at 3.5 kPa. The radial stress a r c a n be taken as half the
effective overburden pressure at 2 m depth (about 23 kPa). The stress leading to creep failure is thus
0.5-23 + 5-3.5 + 2 130 = 289 kPa which is less than o c ol (187 kPa).
The final, primary settlement is thus:
sp = 10 187/21900 = 0,085 m
2 2
Case 2: We have as = (TT-1.0 /4)/(1.8 V3/2) = 0.280 and as before. This yields:
o c ol = 84-27/(1 + 83 0.280) = 94 kPa
The final, primary consolidation settlement becomes:
sp = 10-94/21900 = 0.043 m

The settlement after a loading time of 13 months is analysed by the equation for consolidation by
granular columns, Eq. (522). We have:
Uh = \ - exp(-87VM)
2
Case 1: Assuming c A = 2c vand*: A = 2c vy w/Af = we have Th = 20.6(13/12)/( 1.05 -1.3) = 0.698. Furthermore,
2 2 2 4 1
we find = [ 2 . 7 3 / ( 2 . 7 3 - l)][ln (2.73) - 0.75 + 2.73~ - (4-2.73 )- ] + (2/-)20.69.81/(2600.9)(1
2
- 2.73~ ) = 0.384 + 0 . 1 3 8 z ( 2 / - z) . For 10 m columns, drained at top and bottom, we have / = 5 m.
Choosing equal to 0, 1 , 2 ,, 5 to find the mean consolidation degree we find - 0.384, 1.626, 2.592,
3.282, 3.696 and 3.834, respectively.
This yields Uh = 86% whence the settlement after 13 months s = 0 . 8 6 ^ 0.073 m
2 2
Case 2: Proceeding in the same way we find Th = 0.364 and = [1.89 /(1.89 - 1)] [In (1.89) - 0.75 +
2 4 1 2
1.89- + (4-1.89 )- ] + (2/-)20.69.81/(2603.6) = 0.166 + 0 . 1 5 8 z ( 2 / - z ) ( l - 1 . 8 9 ) . the average
consolidation after 13 months equal to Uh = 98%. Hence the settlement after 13 months s = 0.042 m.
The measured average settlement after 9 months in case 1 is 0.092 m and in case 2 0.039 m. Thus
the agreement between calculated values and observations is quite satisfactory.

7. ELECTROCHEMICAL STABILISATION

7.1 Electro-osmosis

Electro-osmotic dewatering can also be used for the purpose of consolidating silt and
clay. In the case of one-dimensional consolidation in the direction in an electric field
with constant voltage V over two rows of electrodes with large extension, the consolidation
equation can be written (Esrig, 1968):

2 2
dV kh d u 1 du
L z
dx Yw dx M dt

where ke = electro-osmotic permeability coefficient,


k = hydraulic permeability coefficient,
u = excess pore water pressure,
M - compression modulus,
So/7 improvement 463

Yw - unit w e i g h t of water.
Introducing the auxiliary varaiable = (ke/k)ywV+ w, the consolidation equation takes
the form:


3 3

* - ( 5 2 6 )

w h e r e ch = k^Aly^ = consolidation coefficient in horizontal pore water flow (in the


direction).
A s s u m i n g that the anodes are arranged circularly around the cathode w e have:

C ( + ) = )
* " T T 2 i :
1
p dp dp at

w h e r e denotes the radial distance from the cathode (radius vector).


However, these consolidation equations are difficult to solve b e c a u s e of difficulties in
determining the b o u n d a r y conditions.
T h e consolidation process will b e finalised w h e n the steady-state electrokinetic water
flow and the steady-state hydraulic water flow, acting in opposite directions, are equally
large. T h e c h a n g e in average effective stress, assuming one-dimensional horizontal
consolidation (Fig. 3 3 3 ), then b e c o m e s equal to:

dx kh dx

Fig. 333. Variation of electric potential and pore pressure gradient between anodes and cathodes at
equilibrium.
464 So/7 improvement

A s s u m i n g that the potential gradient between the row of anodes and the row of
cathodes is linear the average effective stress increase can b e expressed as:

/
A^ av = ( f t e y w / * A ) W 2 - A i i c a t h. (529)

If p u m p i n g is taking place at the cathodes, A w c a th b e c o m e s zero (pore water pressure


u n c h a n g e d in relation to the original value) or negative.
With regard to the decrease in water content and ion e x c h a n g e effects, the increase in
shear strength b e c o m e s largest nearest to the anodes. Therefore, the best average
consolidation effect is obtained by switching from time to time the direction of the
potential between the electrodes (anodes are m a d e cathodes and vice versa).
2
T h e current density at the anodes should be restricted to a m a x i m u m of about 30 A / m .
Too high a current density leads to rapid desiccation of the soil around the anodes which
strongly increases the electric anode to soil resistance.

7.2 Ion diffusion

D u e to the effect of ion diffusion the characteristics of the soil can b e thoroughly changed.
Thus, depending on the type of ions diffused the strength and character of the soil may
change drastically for the w o r s e or the better. T h e rate of ion diffusion is normally very
slow and can, for example, be calculated in the case of potassium chloride. T h u s , if KCl
is poured into a vertical hole in the soil, the rate of diffusion can be calculated by the
relation:

de \ d dc
- - - ( D p ) (530)
dt dp dp

where c = concentration of the electrolyte,


D = diffusion coefficient,
= radial distance from the hole (radius vector),
t = time of diffusion.
A solution to this equation was presented by Schneider (1963) in the form of a series
2
of curves for various values of p/R, Dt/R andc7c 0 , w h e r e R denotes the radius of the hole
and c 0 the concentration of the electrolyte in the hole.
T h e use of electric current is an effective m e a n s of speeding u p ion diffusion into the
soil. T h u s , electro-osmosis does not only produce a water flow from anodes to cathodes
but also brings about ion e x c h a n g e which contributes to i m p r o v e the soil properties. T h e
effect of ion e x c h a n g e can be strongly i m p r o v e d by choosing a favourable anode
material, such as aluminium.
Another w a y of improving the electrochemical stabilisation effects is by installing
So/7 improvement 465

anodes in wells or drains filled with electrolytes containing suitable ions and utilising the
electro-osmotic potential of enforcing and accelerating the ion escape into the surroundings.
A l u m i n i u m and p o t a s s i u m ions are well fitted for the purpose.

8. SOIL INJECTION

Soil injection is often a cost-effective m e t h o d of improving the strength and compression


characteristics of non-cohesive soils. Another purpose of injection is to reduce soil
permeability. T h e material suitable for injection depends on the grain size distribution of
the soil and the purpose of injection. T h e most c o m m o n types of injection are cement
grouting and chemical grouting. It is important that the course of injection can be guided
and properly controlled. For this purpose, special injection devices h a v e been developed.
T h e radius of influence of injection depends on the viscosity and the speed of
sedimatation of the particles suspended in the agent utilised. W h e n the rate of flow of the
suspension gets too low the particles will settle leading to a separation which m a k e s the
injection agent ineffective. T h e radius of influence of chemical agents depend on the time
of jelling.
T h e injection pressure m u s t b e limited in order not to cause an u p h e a v e of overlying
soil masses or buildings or a destruction of b a s e m e n t walls.
T h e selection of injection material shall b e m a d e with due consideration to the grain

Fig. 334. T u b e manchette' for guided injection of soil. A steel pipe, provided with peripheral
holes,.generally spaced 330 mm vertically apart, is inserted into a prebored hole in the soil. The annular
space between the pipe and the borehole is filled with semi-plastic material. The injection agent is
introduced through an injection tube, the ends of which has packers on both sides of the orifice in order
to make possible injection through any selected sleeve in the tube manchette.
466 So/7 improvement

size distribution of the soil. Roughly speaking, injection material utilised in coarse sand
should h a v e an upper grain size limit of 0.04 m m , while injection of m e d i u m or fine sand
should b e carried out with colloidal material and in coarse silt with material having a
viscosity close to that of water.
T h e injection is generally carried out by the aid of compressed air or injection p u m p s
through tubes provided with peripheral holes. According to a m e t h o d developed by the
French contractor Soltanche, a sleeve device, k n o w n as 'tube m a n c h e t t e ' (Fig. 334)
is utilised. A casing, 75 m m in diameter, is first installed in the soil. Then a tube
manchette, 38 m m in diameter and provided with peripheral holes, spaced 330 m m
vertically apart and closed by means of rubber sleeves, is inserted into the casing. T h e
annular space between the tube manchette and the casing is filled with semi-plastic
material after which the casing is withdrawn from the soil. T h e injection tube has a
diameter of 25 m m and is provided with packers on both sides of the orifice. Before
injection the orifice of the tube is placed just inside the sleeve w h e r e injection is to be
m a d e . B y increasing the injection pressure the sleeve opens as does also the enclosing
plastic material and the injection material penetrates into the soil at the selected level.
Injection generally starts from bottom. T h e m e t h o d of injection j u s t described has the
advantage of m a k i n g possible injection at any selected level as well as re-injection
without the need of new boreholes.

8.1 Cement and clay-cement grouting

T h e u s e of c e m e n t grouting is governed by the grain size of the soil and is mainly


restricted to coarse-grained soils, chiefly gravel and coarse sand. Special, finely ground
cement has to b e used for injection of sand and bentonite suspension m a y also have to be
added (clay/cement injection) to widen the radius of influence of the injection material.
T h e grain size of the main part of the soil to be injected should always exceed 1 m m .
A m e r i c a n investigations h a v e indicated that cement grouting is possible w h e n the ratio
dl5 of the soil to dS5 of the cement grout is a b o v e 25 (Fig. 335). As a rule of t h u m b , cement
grouting is only possible in soils w h e r e the groundwater level can b e lowered by gravity
wells (see p. 365). Obviously, the permeability k of the soil is of p a r a m o u n t importance
for the choice of grout.
For injection of fissures in rock, the m a x i m u m grain size of the c e m e n t grout should
not exceed on third of the width of the fissures.
T h e radius of influence of injection can b e estimated from the relation (Maag, 1938):

3kr^ 3vl/3
Ht+r ) (531)
ni]

w h e r e k = permeability of the soil,


So/7 improvement 467

Fig. 335. Limiting grain size distribution curve of grout for soil injection.

= porosity of the soil,


r = radius of spherical injection source,
77 = viscosity of grout,
77^ = viscosity of pore water,
= hydraulic driving head,
t = time.
4
For a cylindrical injection source, such as t h e tube m a n c h e t t e ' , of length / and diameter
d, w e can a s s u m e that the surface area of the sphere approximates the e n v e l o p e surface
2 1 / 2
area of the cylinder, Le. 4nr ~ izdl, which yields r ~ 0 . 5 ( J / ) .
In order to i m p r o v e its efficiency, the grout is often activated by adding bentonite, fly-
ash or water glass to the c e m e n t grout. T h e grout can also b e activated by heating it to 2 5 -
35 C or b y m e c h a n i c a l treatment in special mixers. In this w a y it is possible to apply a
lower injection pressure or to enlarge the radius of influence.

8.2 Chemical grouting with hardener

Injection of soil w h o s e grain size varies from m e d i u m sand to coarse silt requires the use
of diluted chemical grout, such as silicates (water glass) and p o l y m e r s ( c h r o m e lignins,
synthetic resins, plastics), which sets by m e a n s of catalysts (hardeners). T h e lower limit
of soil permeability for chemical grouting to b e successful according to Littlejohn (1992)
is 1 /s (30 m/year).
T h e radius of influence depends on the viscosity of the c h e m i c a l grout, the a m o u n t of
chemical grout injected per time unit and the jelling time of the grout. T h e jelling time
ought to b e short in comparison with the total injection time at the injection level in
question, particularly if the soil contains coarse-grained layers or if there is a groundwater
current. A rough approximation of the radius of influence is given by the relation (Karol,
1960):
468 So/7 improvement

1 /3
R - 0.6(^^) (532)

w h e r e Q = v o l u m e of injected grout per time unit,


t = jelling time,
= porosity of the soil,
= viscosity of grout,
r j w = viscosity of p o r e water.

(i) Water glass. A c c o r d i n g to the original m e t h o d of water glass injection, the 'two-shot'
solution introduced by Joosten, water glass is first injected into the soil and afterwards
a salt solution, usually C a C l 2 . T h e n a silica gel is formed and the soil b e c o m e s 'petrified'.
Using C a C l 2 the following chemical reaction takes place:
N a 20 ( S i O ) 2 + C a C l 2 + ( H 2 0 ) m - ( S i O ^ O ^ O ) ^ + C a ( O H ) 2 + 2NaCl
T h e chemical reaction thus results in the formation of silica gel plus calcium hydroxide
plus c o m m o n salt. Ion e x c h a n g e processes and other consequential reactions m a k e the
soil increasingly firmer.
T h e field of application of this m e t h o d is restricted to sand b e c a u s e of relatively high
viscosity of the water glass. T h e compressive strength of the grouted soil is generally in
the order of m a g n i t u d e of 3 - 5 M P a , higher in fine sand than in coarser sand.
According to the 'one-shot' solution later introduced, a catalyst is added to the water
glass before injection. In this w a y the chemical reaction can b e delayed and the time of
setting controlled b y adding either acid or alcalic catalysts. T h e following factors govern
the time of setting:
neutralisation of the water glass solution gives short setting time,
very active catalysts, such as sulphuric acid, increases the jelling sensitivity of the
solution with regard to changes in the degree of neutralisation,

TABLE 4 1 .

Radius of influence of water glass injection according to experience (Szchy, 1965)

Soil type Permeability, mm/s Radius of influence, m


Medium to coarse sand 0.2-1.2 0.3-0.4
(two-shot solution) 1.2-2.3 0.4-0.6
2.3-5.8 0.6-0.8
5.8-9.0 0.8-1.0

Fine to medium sand 0.03-0.06 0.3-0.4


(one-shot solution) 0.06-0.12 0.4-0.6
0.12-0.23 0.6-0.8
0.23-O.58 0.8-1.0
So/7 improvement 469

low concentration of the solution results in higher strength,


a temperature increase (in an interval of 10C to 4 0 C ) the setting time of an acid
solution will decrease while that of an alcalic solution will increase.
Injection of water glass by the one-shot procedure can b e used in m o r e fine-grained soil
than the two-shot procedure. Based on experience, Szchy (1965) presents the radii of
influence given in Table 4 1 .

(ii) Polymers. P o l y m e r s h a v e a wider r a n g e of applicability but are m o r e e x p e n s i v e to use


than cement or water glass. D u e to very low viscosity (tending to the viscosity of water)
they can b e injected into very low-pervious, granular soils, such as coarse silt. A m o n g the
polymers used can b e m e n t i o n e d chrome-lignin and acrylamide.
T h e jelling time of chrome-lignin, which is a byproduct in the cellulose industry, can
be controlled b y m e a n s of various catalysts and the final product is lignin-sulphonate. T h e
strength of the grouted soil is low, about 0.5 M P a .
A c r y l a m i d e is a m o n o m e r of calcium acrylate with either s o d i u m sulphate or
a m m o n i u m persulphate catalyst. Its polymerisation depends on temperature and the
concentration of solution and catalyst. T h e strength of the grouted soil is of the order of
magnitude of 100 kPa.
T h e lower limits of various injection grouts with reference to the soil permeability is
shown in Fig. 336.

1000

Fig. 336. Injectability limits for various grouts with reference to soil permeability (Cambefort, 1967)
470 So/7 improvement

8.3 Bitumen and bentonite injections

(i) Bitumen. Diluted b i t u m e n emulsions with catalysts are characterised by very small
particle size ( 1 - 2 ) and low viscosity and can, therefore, b e injected easily in fine-
grained soils, such as coarse silt. T h e grout drives a w a y the p o r e water without mixing
with it. T h e bitumen solution sets by the effect of the hardener and sticks to the grains in
the soil.

(ii) Bentonite. Bentonite is a volcanic clay characterised by its p o w e r to take u p water with
swelling in consequence. T h u s , for e x a m p l e , 'Volclay ' bentonite swells 12 times its orig-
inal v o l u m e from dry to water saturated condition. Bentonite is injected in sand and
gravel for the purpose of creating barriers against water leakage.
At injection of coarse gravel, coarsely-ground bentonite is m i x e d with bentonite slurry.
This prevents the coarse bentonite grains from forming agglomerates which w o u l d m a k e
p u m p i n g of the grout impossible. T h e coarser grains in the bentonite slurry will gradually
swell and cause clogging.
Injection of fine to m e d i u m gravel can also b e carried out with a s o m e w h a t lumpy
bentonite slurry, though with smaller grain size than in the previous case. L u m p s cannot
be allowed at injection of sand. T h e m o r e fine-grained the soil, the m o r e dilute the
bentonite slurry.
Bentonite slurry is sometimes mixed with gravel, sand, sawdust, metal filings, or the
like in order to i m p r o v e the possibilities of creating water-tight barriers in current water.
A bentonite slurry with a bentonite content of only 8 % quite simply keeps sand in
suspension.

9. THERMAL IMPROVEMENT

9.1 Burning

As is well-known, clay which undergoes burning in the brick industry b e c o m e s an


excellent building material with considerable tensile and c o m p r e s s i v e strengths and high
compression m o d u l u s . This fact is sometimes utilised to i m p r o v e the properties of fine-
grained soil in situ, especially in the case of loess and unsaturated clays. T h e plasticity
characteristics of the soil c h a n g e essentially w h e n the soil is heated to a temperature of
4 0 0 C and its capability of taking up water b e c o m e s strongly reduced at a temperature
of 6 0 0 C .
T h e practical application of burning varies. M o s t experience of the m e t h o d is gained
in Eastern Europe. As an e x a m p l e of the application m e t h o d s , holes are prebored and
filled with gasohol which is kept burning until the expected result has been achieved. In
such an application in the former Soviet Union, 1 0 - 1 2 days of burning resulted in burned
c o l u m n s , 2 - 3 m in diameter, to a depth of 6 - 1 5 m.
So/7 improvement 471

Fig. 337. Freezing in connection with tunnelling.

9.2 Freezing

Freezing in order to i m p r o v e the strength of soil seems to h a v e b e e n first applied 1862


in Wales and is thus a very old soil i m p r o v e m e n t method. It is e x p e n s i v e and is therefore
used only in cases w h e r e alternative m e t h o d s cannot b e applied. T h e freezing technique
is for instance utilised in deep excavations b e l o w the g r o u n d w a t e r table to stop internal
erosion and leakage and in tunnel construction, particularly in occasional r o c k to soil
crossings (Fig. 337).
T h e freezing operation has to b e very fast in order to prevent frost h e a v e . In free-
draining granular soils, freezing will not p r o d u c e any v o l u m e increase (instead the water
content will decrease). It is s e l d o m possible to create full water-tightness of the frozen
soil w h i c h can b e s o m e w h a t of a p r o b l e m w h e r e construction is going on b e l o w water.
T h e strength of frozen soil is strongly dependent on freezing temperature and time of
loading. It is considerably larger than the strength of ice and increases with increasing
grain size of the soil (Fig. 338). Organic content reduces the the strength.
Frozen soil has visco-elastic properties. T h e tensile strength is considerably lower than
the c o m p r e s s i v e strength.
T h a w i n g of frozen fine-grained soil, such as silt and clay, gives rise to strong reduction
in the shear strength and increasing compressibility. T h e c o n s e q u e n c e s of t h a w i n g can
be particularly serious in silt and clayey silt w h e r e ice lenses m a y h a v e b e e n formed
during freezing.
472 So/7 improvement

" - 10 -20 - 30 -40 - 50


Temperature of sample, C

Fig. 338. Influence on compressive strength (unconfined compression tests) of ice and soil samples due
to change in temperature.

Freezing can b e carried out by the aid of a m m o n i a gas, carbon dioxide or liquid
nitrogen.
Freezing by the add of carbon dioxide is usually carried out in the following way. T h e
carbon dioxide is first compressed to a gas pressure of about 6 - 7 M P a . This causes a
temperature increase to about + 4 5 C . T h e gas is then conveyed to a condensator where
it turns into a liquid at a temperature of about + 2 0 C . N o w the liquid carbon dioxide is
allowed to pass through a narrow orifice into the freezing pipes. This causes a fall in
pressure to about 0.5 M P a and this, in turn, entails gasification a c c o m p a n i e d by strong
absorption of heat from the surroundings w h i c h freeze. T h e gas is then returned to the
compressor for a new cycle to begin.

10. REINFORCED SOIL

10.1 General aspects

Reinforcement of soil to m a k e possible the u s e of very steep slopes along road cuts or to
reduce earth pressure against retaining structures is an old technique(Figs. 3 3 9 - 3 4 0 ) but
has gained n e w and interesting applications in soil m e c h a n i c s . T h e reinforcement is
generally carried out b y m e a n s of various geosynthetic materials or, m o r e seldom, b y the
u s e of steel. T h e behaviour of reinforced soil depends on the strength and stress/strain
characteristics and the long-term resistance of the reinforcement materials, their design
(strips, grids, m a t s , m e s h e s , strands, etc.) and positioning and the friction along the
contact area b e t w e e n reinforcement and soil.
So/7 improvement 473

Fig. 339. Reinforcement with grass in an ancient monument in Iraq.

T h e o p t i m u m result is achieved if the reinforcement is positioned in the minor


principal stress direction and is designed so that the strains induced in the soil and
reinforcement are compatible. T h e reason is that the minor principal stress direction
coincides with the direction of pure tension. If the reinforcement is positioned in a
direction of compression or in a direction of pure shear, it b e c o m e s inefficient or m a y
even create planes of w e a k n e s s leading to failure.
T h e effectiveness of the reinforcement depends also on the relative density of the soil.
A s s u m i n g that w e h a v e to deal with plain strain condition (which is c o m m o n l y the case
in practice), the v o l u m e c h a n g e in shear can b e expressed as = + 3 (compression
and tension given positive and negative signs, respectively). In other words, the higher

Fig. 340. Fascines represent another old type of soil reinforcement, here in a modern version.
474 So/7 improvement

yll

Fig. 341. Mhr's strain circle utilised for positioning of soil reinforcement. The reinforcement should
preferably be placed in the direction of pure tension 3.

the relative density (i.e. the larger the dilatency - ) , the larger the pure tension 3 (the
centre of the M h r strain circle shown in Fig. 341 will b e displaced to the left). This
tendency towards increasing tension due to dilation is counteracted by the reinforcement.
Dilation also increases the earth pressure against the reinforcement bodies which
contributes to improved co-operation between soil and reinforcement. T h e latter obviously
b e c o m e s m o r e effective in dense than in loose or m e d i u m dense soils.
T h e reinforcement can be installed in two different w a y s either in connection with
filling operations or directly in natural soil, for e x a m p l e in cut slopes. T h e latter technique
is referred to as soil nailing.

10.2 Design

For c o m p l e t e information on the design of reinforced soil, reference is m a d e to a textbook


by H a u s m a n n (1990). A complete description of the properties of geosynthetics and of
design m e t h o d s belonging to these materials is also presented by L a w s o n (1993). In this
context, only a brief summation will be given of the principles of design used at present.
T h e most important aspects of design concern the mechanical properties of the
reinforcement material. T h e requirements to b e placed on the reinforcement are very
m u c h dependent on the p r o b l e m dealt with. As previously stated, its in situ mechanical
properties h a v e to b e compatible with those of the soil under the conditions prevailing in
the reinforcement zone.
Other aspects of importance in the design are the durability and long-term capacity of
the reinforcement. This is of particular interest since reinforced soil is used in p e r m a n e n t
structures and there is no material used for reinforcement purposes that can b e considered
permanent. As regards geosynthetics these are mostly very sensitive to ultra-violet light.
So/7 improvement 475

Thus, gesynthetic materials cannot be exposed to sunlight without being subjected to


long-term degradation (polyester is the most resistant and polypropylene the least
resistant material). The resistance to ultra-violet light can be improved by U V stabilisers.
Polyethylene and polypropylene are also to a certain extent subjected to oxidation and
polyester to hydrolysis (Lawson, 1993).
From a mechanical viewpoint the analysis of reinforced soil walls is carried out in
accordance with the analysis of retaining structures.

Example 89: For stabilisation purposes, a 5 m high vertical cut shall be anchored by means of geo-
synthetics placed in horizontal layers. Determine the total anchoring force required if the soil has an
3
internal angle of friktion of 35 and a unit weight of 18 kN/m . The top side of the cut is subjected to
2
an evenly distributed load of 20 kN/m .

Solution: Since there is no friction on the frontside of the cut we have to deal with an active Rankine
type zone failure. The anchoring force required will be equal to the active earth pressure:
2 2
Pa = [ 1 8 ( 5 / 2 ) + 20-5]tan (45 - 072) = 88.1 kN/m
Assuming a factor of safety of 2, the design anchoring force should be 166 kN/m. The distribution
of the geosynthetic reinforcement with depth should be compatible with the earth pressure distribution.
476 Design aspects

ASPECTS OF GEOTECHNICAL DESIGN

1. GEOMECHANICAL VIEWS

1.1 Accuracy in methods ofcalcufation

T h e m e t h o d s of calculation presented in this textbook are mostly b a s e d on the assumption


of elastic-plastic behaviour of soil. As s h o w n in chapter ' F u n d a m e n t a l s ' , however, this
is a very simplified picture of reality. Hence, in the selection of soil parameters to b e used
in the various m e t h o d s of calculation, there is a great need for sound j u d g m e n t and
profound k n o w l e d g e of the real behaviour of soil under the conditions prevailing in the
design. Without any doubt, the results obtained even by fairly simple m e t h o d s of analysis
m a y be j u s t as reliable as those obtained by the use of advanced soil m o d e l s and computer
analysis. T h e apparently good correlation b e t w e e n computer-based analysis and case
records can mostly be related to the fact that the soil parameters used in the analysis h a v e
been adjusted by back-analysis to tally with the real behaviour. In reality, the soil
parameters forming an integral part of the computer m o d e l represent a great d i l e m m a in
computer-based analysis. By experience w e k n o w that the results obtained, although
seemingly trustworthy, can b e seriously deceptive. With continued research the capacity
of predicting real behaviour in foundation engineering will certainly be greatly improved,
but as yet m u c h remains.

(i) Bearing capacity. B y tradition, the determination of the bearing capacity has b e c o m e
one of the corner-stones in foundation engineering. This is understandable from the point
of view that ground failure would entail very serious c o n s e q u e n c e s for the building and
the people working or living in the building. However, in the case of building foundations
the main p r o b l e m in the design is not to ensure a certain factor of safety against ground
failure (except, of course, for foundations adjacent to slopes) but to m a k e sure that the
total and differential settlements will not cause d a m a g e to the building. By tradition, a
high e n o u g h factor of safety against failure was considered as the only criterion necessary
to avoid d a m a g e due to settlement. Only in the case of soft, highly compressible soils,
settlement analysis was normal part of the design.
T h e determination of the bearing capacity, however, is important from quite another
aspect. T h u s , a high-enough safety against ground failure is needed for the soil to b e h a v e
as a pseudo-elastic material applicable to the computational m e t h o d s used in settlement
analysis.
Design aspects All

(ii) Stability. T h e m o s t important aspect of stability and bearing capacity problems


concerns the choice of strength parameters.
T h e undrained (total) shear strength, of interest for fine-grained soils, varies with test
type and testing routine. T i m e to failure, especially in the case of organic soils, and the
direction of shear in the test m a y have a great influence on the result obtained. Hence, it
is important that the tests b e carried out under similar stress and strain conditions as can
be expected in reality. T h e strength depends on the pore pressure induced at shear and can
thus not be considered as an intrinsic soil parameter.
T h e effective strength parameters c ' a n d 0'are dependent on the dilation characteristics
of the soil. Subjected to shear deformations, dense soils tend to dilate while loose soils
tend to contract. T h u s , the shear strength of dense soils will pass a p e a k value before
falling d o w n to the value obtained at critical void ratio. On the other hand the shear
strength of loose soil will gradually increase from below up to the value obtaining at the
critical void ratio. Again, the determination of the strength values to be applied in the
design have to b e properly adjusted to reality. In effective stress analysis, the actual pore
pressure and effective stress distribution are of p a r a m o u n t importance.

(iii) Stress distribution. F r o m experience w e k n o w that the stresses induced in the soil by
the working load can be determined with acceptable accuracy on the basis of the theory
of elasticity. In granular soils, the stress distribution is influenced b y the fact that the soil
cannot take tensile stresses. However, in practice the difference in result obtained for a
particulate m e d i u m and for an elastic m e d i u m is negligible.

(iv) Settlement. As pointed out above, it is presumed in the settlement computations that
the loading condition does not result in essential plastic deformations of the soil. F r o m
a practical viewpoint, settlement analysis for foundations on n o n - c o h e s i v e soils cause
problems mainly for the reason of difficulties in determining relevant elastic properties.
N o n - c o h e s i v e soil is a particulate material and, hence, far from b e h a v i n g as an elastic
material; undisturbed sampling is practically impossible and the deformation parameters
determined on a laboratory scale therefore quite misleading. However, m o d e r n in-situ
methods h a v e greatly i m p r o v e d the accuracy of prediction.
Settlement analysis for cohesive soils is facilitated by the fact that undisturbed sam-
pling is possible. This in turn facilitates the evaluation of requisite parameters for set-
tlement analysis. Calculation m e t h o d s may often seem too simplified to b e trusted, but the
use of more sophisticated analytical solutions may not be worthwhile with regard to the
complexity of nature. Moreover, the deviations between the results obtained with simplified
and more sophisticated approaches, using the same input parameters, are generally negligible.
An illustrative example of this is the simplified Terzaghi consolidation equation (Eq. 29):
2
du du
478 Design aspects

Let u s replace this equation b y the m o r e accurate o n e (Eq. 3 2 ) :

2
du M d du du M dk du ?
~dt Ywdz dz dz Ywdu dz

Let u s further a s s u m e that w e have to deal with a 1 0 m thick clay layer drained at top
and b o t t o m with an induced excess pore water pressure u0 = 1 0 k P a (constant with depth).
T h e compression m o d u l u s of the clay M=200 k P a (constant value) and the permeability
k can b e expressed as (k is a linear function in a log ' / - d i a g r a m ) :

UQ-U
k = k0exp(-k) (533)
M

dk
whence
du M M

with k0 - 0 . 0 3 m/year and k - 8 .


2
T h e coefficient of c o n s o l i d a t i o n cv = ^ Q M / ) ^ ~ 0 . 0 3 - 2 0 0 / 1 0 = 0 . 6 m / y e a r .
T h e excess pore water pressure variation throughout the clay layer according to the
simplified and m o r e accurate relation, solved by c o m p u t e r analysis, is s h o w n in Fig. 3 4 2 .
A s can b e seen, t h e difference in results is negligible from a practical viewpoint.

E x c e s s pore water pressure, kPa

0 0.5 1.0 1.5 2.0


o

\
\
\
\
\
1
r 6 1
/
/
/

10

Fig. 342. Remaining excess pore water pressure after 50 years of consolidation according to Terzaghi's
simplified solution (broken line; average degree of consolidation 95%) and the more accurate solution
taking into consideration a decrease in permeability due to a decrease in void ratio during the
consolidation process (average degree of consolidation 90%). Initial excess pore water pressure u0 = 10
kPa. Clay layer drained at top and bottom. Terzaghi's simplified solution would yield the same result
2 2
as the more accurate one provided that c v is 0.435 m /year (J: r ed = 0.0217 m/year) instead of 0.6 m /year.
Design aspects 479

1.2 Probabilistic approach

T h e i m m e d i a t e and long-term behaviour of the foundation of a prospective building are


matters of greatest concern to the o w n e r and, therefore, has to b e predicted by the
geotechnical engineer to the o w n e r ' s full satisfaction. T h e two m o s t important problems
to b e solved concern the stability and deformation of the ground u n d e r the weight of the
building. T h e safety against ground failure m u s t not fall b e l o w a given limit and the
deformation of the structural elements m u s t not cause inconveniences or structural
damage.
T h e probabilistic approach is generally based on the assumption of normal distribution.
Accordingly, the distribution function of a stochastic variable X is given by the relation
(Fig. 343):

1 l - ?
fx M = 7 = exp[--( -) (534)
2
Gyj2%

w h e r e = E(X) - m e a n value of X
a - D(X) = standard deviation of X
A useful m e a s u r e of the scatter of a r a n d o m variable is its coefficient of variation V(X)
= al.
T h e standardised form of normal distribution,

1
1
() = j = e x p ( - ),
2
OyJlK

is obtained if = 0 and = 1. In this case, the cumulative-distribution function (),


which is the area b e l o w the normal distribution curve from t = -<> to t = JC, is obtained by
(Fig. 344):

Fig. 343. Normal d i s t r i b u t i o n / ^ ) .


480 Design aspects

Fig. 344. Cumulative distribution ()

2
1 . t
&x(x) = = exp(--)dt 2
(535)
/ 2 "~

Since the cumulative-distribution function is equal to 0.5 for = 0, () for > 0 is


obtained by the relation:

2
1 t
() = 0.5 + - = P e x p ( - - ) A (536)
0 2
/2

Introducing () = 0.5 + (), w e have:

= - 7 = | * ( - ^ - ) (537)
2
V2k

which generally forms the basis for tabulated values of the cumulative-distribution curve.

(i) Safety margin. T h e classical measure of fitness of a structural element to perform an


i m p o s e d d e m a n d is the factor of safety F, defined as the quotient of capacity and demand
( F = CID). Failure is indicated where F < 1. Lately, a new additional approach has c o m e
into use based u p o n probability theory. Accordingly, the fitness is measured by the safety
margin SM, defined as capacity minus demand (SM-C-D). T h e safety margin is generally
analysed statistically on the assumption of normal distribution. This is fundamental to the
reliability index approach. Failure is indicated when SM = 0. In this case, the probability
of the safety margin b e c o m i n g less than zero [ p ( 5 M < 0 ) ] is a m e a s u r e of fitness (Fig. 345).
Reliability is a c o m p l e m e n t to failure, thus, for a probability of failure of 5 % the
reliability is 9 5 % . T h e reliability is an important measure w h e n a system of several
Design aspects 481

Fig. 345. Capacity and demand probability distribution.

similar elements is considered. T h e reliability index is defined as the ratio of the m e a n


value to the standard deviation:

(538)

Hence, the reliability index tells us h o w m a n y standard deviations pSM are a b o v e zero
(Fig. 346). Obviously, increases in imply increases in the reliability and the safety
margin. T h e reliability index tells us that the probability of failure is:

(539)
aSM Gsm

1
where {) = = \ tx^(--)dt.
2
If w e h a v e to deal with a population of parameters, the so-called general point estimate
m e t h o d ( G P E M ) developed by Rosenblueth ( 1 9 7 5 ; 1981) simplifies the statistical
approach (Harr, 1987). T h e output function is formed by taking all the input samples in
combination and producing the usual statistical measures (mean, standard deviation,

Safety margin, SM

Fig. 346. Safety margin probability distribution and reliability index.


482 Design aspects

n
etc.). T h e n u m b e r of possible combinations b e c o m e equal to 2 , where is the n u m b e r
of input variables.
M
If t w o r a n d o m variables are correlated, say y =f(xhx2), the expected value E(y ) for
any n u m b e r of k n o w n m o m e n t s M of x, is obtained by the relations:

++ M+ + M +
E(yM) =p (y+ ) + p+ ~(y -) + pr (y~ +)A* + p - -(y- -)M (540)

where y =/(1 1, 2 2),


p+ + = pr- = (i + p)/4,
p+- = / r + = ( l - p ) / 4 .
In these equations, is the correlation coefficient between xx and x2. T h e weighting
factors/? can vary from - 1 t o + 1 . Should there be no correlation, then = 0 and consequently
p= 1/4.
R o s e n b l u e t h ' s generalised version of the point estimate method (see Harr, 1987) is
valid for any n u m b e r of r a n d o m variables. For example, for a function of three variables
y =fijcl,x2y)c3) w e have:

j =f(pxl ],2 2, 3 3\
p+ + + =r
p = (1 + p 1 2 + p 2 3 + p 3 , ) / 2 3 ,
p+ + - = pr-+ = (1 + p , 2 - p 2 3 - p 3 1) / 2 3 ,
+ +
p - = / r + - = (l - p 1 2- p 2 3 + p 3 1) / 2 3 ,
+
p ~ = /r ++= ( l - p 1 2 + p 2 3 - p 3 1) / 2 3 .

Then M t h expectation is obtained from the relation:

M + ++ + + + M + + + + M + + M M
E(y ) =p (y ) +p -(y ~) + + / ? - - ( y - - ) + p ~(y~ ~ ~) (541 )

T h e sign of the correlation coefficients is determined by the sign of the product of the
corresponding + and - of the weighting factors, for example (+)(-) = (-) and (-)(-) = (+).
T h e expected value E(y) = py of the function is obtained for M - 1 while the variance
2
GY is obtained for M - 2.

Example 90: A strip footing with width b = 1 m is subjected to a vertical load Q = 500 kN/m. The footing
is founded at a depth dj- 1 m in sand with an internal angle of friction ' = 30 and a cohesion intercept
/ 3
c = 10 kPa. The groundwater level is at 1 m depth. Above the water table we have = 18 k N / m and
3
below Y =11 kN/m . Determine the reliability index and the probability of failure according to GPEM

if the coefficients of variation are V(Q) = V(f)=0, ( ) = 20% and V(c 0=40%. The correlation coefficient
is p i c ^ O = - 0.5, i.e. if (//increases, ^decreases, and vice versa.

Solution: The safety margin is defined by the relation (p. 126):


SM = 0.5 YbNY+ yjjNq + c'Nc - Qlb
The probability is determined by the safety margin SM < 0. We have to consider two random variables
0'and (/with standard deviations = 6 and oc = 4 kPa. The bearing capacity factors become N + = 23.8,
Design aspects 483
+ + /+
Nq = 37.7, Nc = 50.5, c = 14 kPa and yVy- = 7.65, N~ = 9.60, Wcr = 17.1, c'- = 6 kPa. Furthermore,
+
we have p + =p-- = ( 1 + p c 0) / 4 = 1/8 and p+ - = / ? - + = ( 1 - p c 0) / 4 = 3/8. Hence,
++ ++ + 2
p SM = (0.5-11-23.8 + 18-37.7 + 14-50.5 - 500)/8 = 127.1 p+ +(SM +) = 129159.0
+ + + 2
p SM+- = 3(0.5-11-23.8 + 18-37.7 + 6-50.5 - 500)/8 = 229.7 p ~(SM -) = 140683.6
+ + + + 2
p- SM~ = 3(0.5-11-7.65 + 18-9.6 + 14-50.5 - 500)/8 = 158.2 p- (SM~ ) = 66741.9
2
p--SM = (0.5-11-7.65 + 18-9.6 + 6-50.5 - 500)/8 = 2.2 p(SM) = 39.9
2
The mean value pSM = lp(SM) = 517.2 and () = 336624.5
2 2
The variance (oSM) = 336624.5 - (517.2) = 69128.7 which yields oSM=262.9 kPa and the reliability
index = pSMloSM = 517.2/262.9 = 1.97.
Hence, the probability of failure (SM < 0) = 2.4%

Example 9 1 : A square footing with width b = 2 m is founded on the ground surface of a sand deposit
of great depth. The deformation characteristics of the sand are characterised by Poisson's ratio v = 0.3
and the pseudo-elastic modulus = 15 MPa. Determine the probability of of the settlement exceeding
2
5 = 1 5 mm for an average foundation pressure q = 0.1 M N / m if the coefficients of variation are V(v)
= 30%, V(E) = 20% and V(q) = 10%.

Solution: The safety margin can be expressed by the relation (pp. 141-142):
SM = 0.015 -reqblE
where 0.015 is the settlement limit (in m),
Te is a function of v, depth of soil deposit and ratio of foundation length to foundation width.
The probability is determined by SM < 0. The standard deviations become - 0.09, = 3 MPa and
2 + +
q = 0.01 M N / m . In our case we find / = 0.728 and ~ = 0.820. Moreover, q = 0.11, q~ = 0.09, E
= 18 MPa and E~ = 12 MPa. We have:
+ + + + + +2 3
SM = 0.015 -0.728-0.11-2/18 = 0.00610 (SM ) = 0.0372-10-
+ + + + 2 3
SM - = 0.015 -0.728-0.09-2/18 = 0.00772 (SM - ) = 0.0596 1 0 -
3
SM+-- = 0.015 -0.728-0.09-2/12 = 0.00408 (SM+- " ) 2 = 0.0166-10-
SM = 0.015 -0.820-0.09-2/12 = 0.00270 (SM-- " ) = 0.0073-10-3
2

+
(SM~ " ) 2 = 0.0000-10-3
+
SM- - = 0.015 -0.820-0.11-2/12 = - 0.00003
+ + + 3
SM- = 0.015 -0.820-0.11-2/18 = 0.00498 (SM- + ) 2 = 0.0248-"
(SM + ) 2 = 0.0462-10-3
+
SM-~ = 0.015 -0.820-0.09-2/18 = 0.00680
+ +
- ) = 0.0027-10-3
2
SM+ + - = 0.015 -0.728-0.11-2/12 = 0.00165 (SM
2 3
(5 )/8 = 0.00425 = pSM and ( ) /8 = 0.02432-10"
2 3 2 3
The variance (oSM) = 0.02432-10" _ (0.00425) = 0.00626-10~ which yields oSM = 0.00250 and
the reliability index j3= pSMloSM = 0.00425/0.00250 = 1.70.
Thus the probability of s exceeding 15 mm is 4.5%.

2. FOUNDATION REQUIREMENTS

2,1 Design criteria for buildings

(i) Total and differential settlement. O n e of the m o s t important questions in the geotechnical
design of f o u n d a t i o n s is to d e c i d e w h a t settlement, total and differential, c a n b e a l l o w e d .
T h e total settlement m u s t b e restricted, primarily in o r d e r that c o n n e c t i n g s e w e r s are n o t
being b r o k e n or t e n d i n g to slope b a c k w a r d s . T h e differential settlements m u s t b e restricted
in o r d e r n o t to entail structural d a m a g e to the b u i l d i n g or restrict its serviceability.
484 Design aspects

Fig. 347. Settlement of a block of old buildings founded on deep clay deposits. Original situation of the
buildings indicated with black lines. (By courtesy of BAAB).

E x a m p l e s of buildings subjected to large total and differential settlements are given in


Figs 3 4 7 - 3 4 8 .

(i) Empirical damage limits. Probably the most w e l l - k n o w n inventory of buildings


subjected to d a m a g e was carried out by S k e m p t o n and M c D o n a l d (1956). T h e inventory
included 98 buildings, a m o n g which 48 had suffered d a m a g e by differential settlement.
S k e m p t o n and M c D o n a l d found that cracks in main walls and supporting frameworks
had been formed when the angle of rotation b e t w e e n nearby supports (Fig. 349) exceeded
6.7%o (1:150) and in secondary structural elements w h e n exceeded 3.3%o (1:300).
A similar investigation carried out by Rethaty (1964) s h o w e d that the d a m a g e s caused
by excessive differential settlements also depend on the n u m b e r of storeys (Fig. 350). T h e
m o r e the n u m b e r of storeys, the higher the angular c h a n g e a giving rise to structural
d a m a g e . This can b e explained by the length of the construction time: the longer the time
of construction, the better the possibilities for the building to a c c o m m o d a t e itself to
differential settlements, for e x a m p l e by formation of plastic hinges and by creep. T h e
d a m a g e occurring in this case is mostly concentrated to the b o t t o m floors which are first
erected.
Design aspects 485

Fig. 348. Differential seulement of a small family house due to influence of road embankment causing
consolidation settlement.

According to Bjerrum (1966), the empirical limiting values of the angle of inclination
can b e estimated from Fig. 3 5 1 . Bjerrum also presented an empirical correlation
between the m a x i m u m differential and the m a x i m u m total settlements of buildings
founded on sand and on clay (Fig. 352).

(ii) Design settlement limits. Detailed regulations regarding permissible total and
differential settlements exist in several countries, for instance in C a n a d a and in some

A D

Fig. 349. Definitions used in differential settlement analysis.


486 Design aspects

11

<

> 4> ]

_=H
<
'
ML j
j
j

2 4
Number of storeys

Fig. 350. Inventory of damages due to differential settlements (Rethaty, 1964). Filled circles represent
damage to frameworks; open circles damage to partition walls and Wallings. Line ( 1 ) denotes lower limit
for damage to frameworks; line (2) lower limit for damage to partition walls and Wallings.

Eastern European countries. In the Canadian M a n u a l on Foundation Engineering (1985),


the following limiting values are presented:

high continuous brick walls 0.001-0.005


residential brick h o u s e s 0.003
filling walls of bricks between columns 0.001
reinforced concrete frameworks 0.0025-0.004
reinforced concrete walls 0.003
continuous steel frameworks 0.002
freely supported steel frames 0.005

.5 *'3I
.S
13

4 6 10
Rotation, %o

Fig. 351. Empirical limiting values of of angle of rotation according to Bjerrum (1966).
Design aspects 487

0 10 20 30 40 0 5 10 15
Maximum settlement, cm

Fig. 352. Maximum differential settlements Asmax and maximumangle of rotation or building founded
on clay (left) and sand according to Bjerrum (1966).

T h e limiting angular changes a, given as a function of the span width L, are:


masonry, glass or other brittle material L/360
metal covering or similar non-brittle material L/240
frameworks of steel or reinforced concrete L/150-L/180
frameworks of w o o d L/100

T h e limiting total settlement according to the Canadian M a n u a l on Foundation


Engineering is 0.15 m for buildings founded on clay and 0.05 m for buildings founded
on sand.
A s indicated in Fig. 3 5 1 , the d e m a n d s on settlement limitation for m a c h i n e foundations
are very strict. For e x a m p l e , in order to avoid d a m a g e to the bearings in paper and
calendering m a c h i n e s , the m a x i m u m allowable inclination of a b o t t o m rail section must
not e x c e e d 0 . 2 5 - 1 m m / m in the length direction and 0 . 3 5 - 1 m m / m in the transverse
direction of the m a c h i n e . Moreover, the difference in inclination b e t w e e n the points of
observation m u s t not e x c e e d 0 . 2 - 0 . 4 m m / m .

2.2 Design criteria for tanks

Tanks are large cylindrical steel containers built for storage and distribution of liquids
(petroleum products, molasses, etc.). T h e y are mostly grouped together in tank farms in
close proximity to w a t e r w a y s used for the transportation and delivery of the products to
b e stored in the tanks. In c o n s e q u e n c e , they are located in places w h e r e subsoil conditions
are often very poor.
488 Design aspects

T h e cost of foundation of tanks represents the major part of the building cost. In order
to m i n i m i s e these costs, foundation methods may be accepted which, according to
prediction, will entail settlement that verge on adopted criteria for permissible settlement.
If in reality settlements should tend to exceed the limits given in the design criteria, there
is always a possibility to re-level the shell or the b o t t o m plate of the tank.

(i) Damage problems. T h e problems normally encountered are summarised below.


Radial deformation of shell (or shell out-of-roundness). In the case of floating roof,
shell out-of-roundness m a y lead to binding of the floating roof during operation, or to a
gap between the tank shell and the floating roof. In the case of fixed cone roof tanks, shell
out-of-roundness m a y result in upper shell course buckling.
Rupture of shell, or bottom plate, or bottom plate-shell connection

(ii) Design settlement limits. Empirical design criteria, proposed by M a r r etal.(\ 982) and
by Malik et al. (1977), represent the most recent and well-founded design criteria. T h e
criteria are divided into tilt, differential settlement of shell and differential settlement of
bottom plate.
Tilt. To avoid spilling oil from the floating roof or stressing the roof of cone roof tanks,
the following condition should b e satisfied (Fig. 352):

5m!a<2Ahd (542)

w h e r e Ahd is the design freeboard.


Moreover, the change in diameter caused by rotation of the shell relative to the roof
must not exceed the tolerance of the seal A/? t o l, w h e n c e :

5 n i < 2 / D A / ? 1 0, (543)

Differential settlement of shell. T h e criterion given by Malik et al. ( 1977) is considered


as being the best in respect of ovalness resulting from out-of-plane distortion of floating
roof tanks. Accordingly (Fig. 353):

Sl
" 2
=
~^~ ) ( 5 4 4

w h e r e N= total n u m b e r of survey points spaced at equal angles along the shell perimeter,
Si = out-of-plane settlement in the /:th survey point.
However, it is suggested that large tanks can stand m o r e differential settlement than
is indicated by this relation.
Design aspects 489

Aspects of performance Mode of failure

Planar tilt

Overtopping of shell
L o s s of roof seal

Non-planar tilt

Binding of roof seal


Overstress of shell
point

Rupture from d i s h - s h a p e d settlement

Rupture from l o c a l i s e d depressions

a. R e m o t e from shell
b. Adjacent to shell

Non-planar settlement

deformed tank wall Rupture of c o n n e c t i o n as shell


bridges over soft spots
gap.

d e f o r m e d annular ring

Fig. 353. Aspects of the performance and modes of failure for oil tanks.

T h e out-of-plane settlements, i.e. the settlements relative to planar tilt, are obviously
an important part of the design criteria for tanks.
T h e planar tilt is given by the first h a r m o n i c of the Fourier series:

Z ^ A o + A ^ o s C ^ + jS) (545)

w h e r e { = 2ni/N (i = 1, 2, 3, - , N),
A0 = (1/0;, = m e a s u r e d average settlement of shell,

Fig. 354. Definition of symbols used in Eq. (546).


490 Design aspects

2 2
A ! = - y/; (Pi cos q>i ) + , (p, sin { ) ,

= arctan[(, , sin { ) / (, pt- cos >-f )]
T h e angle of tilt is equal to 2AXID.
T h e out-of-plane settlement the results from:
s z
i = Pi - i
T h e criterion for verstress of the shell, based on experience of 90 case records, is given
by the relation:
2 2
1 1 )
A s
' - ^ r < 546)

w h e r e oy is the yield stress of the steel in the shell,


is the elastic m o d u l u s of the steel in the shell,
D and according to Fig. 352.

(iii) Differential settlement of bottom plate. In the case of dish-type settlement, it is


r e c o m m e n d e d that:

2
wl + 031afD
vv < \
1
(547)
V E(FS)
} K

w h e r e w and w 0 according to Fig. 3 5 3 ,


^= rupture steel stress limit of the steel in the b o t t o m plate,
(FS) = safety factor of tank material (< 4 if localised yielding is possible, and < 2 if
severe overstressing is possible).
If the case of localised depressions (Fig. 353) remote from the shell, w e have the criterion
(d in m):

f ,0.28,
s<d\l^
E(FS) ) ( 5 4 8

and in the case of localised depression adjacent to the shell:

3 4/
/ 2.25 afDld
S ) ( 5 4 9
" V EH(FS)

T h e value of the safety factor (FS) is chosen in the s a m e way as in the case of dish-
shaped settlement.
Design aspects 491

T h e criteria concerning localised depressions are established on the assumption that


the b o t t o m plate s h o w s dish-shaped settlement of local depression (Fig. 353). A
combination of the t w o m o d e s m a y lead to higher stresses than those experienced in either
of the two m o d e s .

2.3 Important aspects of foundation design

(i) Soil-structure interaction. T h e most obvious cause of settlement is the load placed on
the ground in connection with the building activity. T h e analysis of the settlement due to
the weight of the building is generally carried out by the geotechnician on the basis of the
building loads presented to h i m by the structural engineer. Now, the general procedure
adopted by the structural engineer in his analysis is to a s s u m e that the building is resting
on unyielding ground. However, differential settlements will give rise to changes in the
loading condition and consequently the settlements calculated by the geotechnician on
the premises given by the structural engineer will not b e correct. L o a d s on c o l u m n s that
settle m o r e will partly be carried over to columns that settle less. H e n c e , the resulting error
is generally on the safe side as regards the a m o u n t of calculated differential settlements.
T h e errors in calculated internal forces, however, m a y b e serious. Calculated bending
m o m e n t s m a y h a v e an opposite sign to that occurring in reality which m e a n s that
reinforcement bars m a y b e placed on the wrong side of the beam. A d v a n c e d computer
p r o g r a m s will probably take care of these problems in the future by m e a n s of a successive
iteration process. Of course, several difficulties will be encountered with regard to the
gradual build-up of the structural stiffness during the construction period. Fortunately,
the structure has an admirable capacity to adjust itself to its design. For e x a m p l e , in a
concrete structure fine cracks will appear at w e a k sections, plastic hinges will be formed,
and so forth. M o s t important of all, a considerable a m o u n t of a c c o m m o d a t i o n will have
taken place during the construction period.

(ii) Other causes of settlements. In regions with compressible fine-grained soils, settlements
often take place d u e to goundwater lowering or to changes in ground level. This fact is
often ignored and attention is mainly paid to the foundation of the buildings, in particular
whether piling is required or not. T h e differential settlements often observed between
piled buildings and the surrounding area (Fig. 224) m a y cause serious trouble, for
instance, b r e a k a g e of connecting sewers.
T h e influence of building activities on the groundwater condition can b e quite hard to
predict. For e x a m p l e , groundwater leakage into tunnels and caverns in b e d r o c k often
entail serious g r o u n d w a t e r lowering (Fig. 257). Therefore, if tunnels or caverns are
situated in areas w h e r e the overlying soil consists of clay or other compressible soil
deposits, leakage m u s t be prevented in o n e w a y or the other or water infiltration be
arranged in order to maintain the existing state of pore pressure distribution in the area.
Observations h a v e shown that it is extremely difficult to predict w h e r e the groundwater
492 Design aspects

will b e mostly affected by tunnels in bedrock. In the case shown in Fig. 258 (p. 361) with
a raw water tunnel in b e d r o c k passing a crush-zone, water l e a k a g e entailed a pore
pressure decrease of over 100 kPa in a sand aquifer underneath clay within a fairly limited
area, with its centre about 300 m away from the leaking tunnel line. Obviously, theoretical
m o d e l s for calculation of the depression in groudwater level around a tunnel or a rock
cavern, based on condition of symmetry, are useless.
L e a k a g e can b e prevented by grout hole drilling and cement injection (p. 466).
Infiltration is often connected with difficulties of various kinds (see p. 373).
In semi-arid regions, desiccation during lengthy dry periods can lower the groundwater
table beneath u n c o v e r e d areas to considerable depth. D u e to desiccation, cracks are
broken u p in the soil and are filled with water during the rainy season. If the soil contains
swelling minerals (smectite), a consequential swelling and h e a v e will take place. Beneath
buildings and other covered areas the loss of water is very m u c h smaller than in adjacent,
uncovered areas and, hence, a long-term increase in the water content in the soil below
central parts of the covered areas will take place. T h e consequential h e a v e is m o r e or less
independent of the external loading condition. T h e differential settlements caused by
h e a v e inside buildings and by settlement outside can entail serious d a m a g e to buildings
(Fig. 355).
Retaining structures supporting frost-susceptible soil are subjected to a considerable
increase in earth pressure in wintertime (Fig. 356). H e n c e , it is very important that frost-
susceptible soil adjacent to the retaining structure should be replaced with frost-
insusceptible soil. Frost h e a v e is a c o m m o n cause of differential settlement in buildings
which are founded at smaller depths than the depth of frost penetration or subjected to
frost penetration which is insufficiently protected against.

Fig. 355. Example of differential settlement due to desiccation effects.


Design aspects 493

Fig. 356. Damage to retaining wall due to frost action.

(iii) Safety against ground failure.The possible negative influence of building activities
on the stability of of the building area is another important matter to consider. T h e bearing
capacity of the of the foundation itself is generally guaranteed by the application of safety
factors in the design and is therefore seldom in jeopardy. M a n y times, however, the
overall stability of the building site or its neighbourhood can be j e o p a r d i s e d due to
unforeseen influence on the overall long-term stability condition. It is thus extremely

Fig. 357. Slide in built-up area in Norway. The slide scar indicates a thick and strong dry crust formation
which may have been misleading in the judgment of the stability condition.
494 Design aspects

Fig. 358. Piping caused by an earthquake in Japan. (By courtesy of Y. Yoshimi).

important that the stability conditions are carefully investigated wherever the building
activities take place on sloping terrain or in immediate vicinity of slopes with clay and
silt deposits (Fig. 357).
In seismic regions with loose sand or silt, complete liquefaction of large soil masses
have taken place. In 1920, in the Chinese province Shansun, an earthquake was reported
which induced an earth slide in loess soil comprising an area, 480 k m in length and 160
k m in width, resulting in about 100,000 casualties. Piping phenomena in connection with
strong earthquakes are not as disastrous but quite spectacular (Fig. 358).
The vibrations induced by blasting can also entail liquefaction, although in more rare
cases. However, blasting in the vicinity of slopes where the stability is at stake should
certainly be avoided (Fig. 359).

Fig. 359. Slide triggered by rock blasting on the west coast of Sweden.
Design aspects 495

In certain areas, formation of sinkholes in soluble rocks and sensitive soils are
experienced. D r a m a t i c e x a m p l e s of the catastrophic consequences that m a y follow upon
sinkhole formations h a v e been presented by Wagener (1989).
Besides long-term settlement caused by mining operations there m a y also b e a risk of
ground failure. A n e x a m p l e of such a failure due to the collapse of the roof of an old iron
m i n e w a s given in Fig. 8 (p. 17).

3. INTERNATIONAL CODES

At present, concurrent w o r k is going on in the International Standardisation Organistation


(ISO) and the E u r o p e a n C o m m i t t e e for Standardisation ( C E N ) with the purpose of
drawing u p standards for geotechnical design. Several of the specialists e n g a g e d in the
standardisation w o r k s h a v e their c o m m i s s i o n from both I S O and C E N and, therefore, the
b a c k b o n e s of the international and the E u r o p e a n standards will m o s t probably b e m o r e
or less the same.
T h e d o c u m e n t s p r o d u c e d by the I S O and the C E N committees are still in a preliminary
stage and will m o s t probably b e subjected to i m p r o v e m e n t s after having b e e n tested in
practice. It w o u l d lead too far to go into details which m a y b e altered in the final version
of the codes. Therefore, only the philosophy behind the c o d e will b e dealt with.

3.1 Scope of the Eurocode

T h e E u r o c o d e is an essential part of the establishment of the E u r o p e a n C o m m o n Market.


T h e intention is to agree on a c o m m o n code that can b e applied within the w h o l e of E u r o p e
and which facilitates competition between all contractors from the various countries.
In all, the c o d e includes eight different titlesBasis of design and actions on structures
(Eurocode 1), Design of concrete structures (Eurocode 2), D e s i g n of steel structures
(Eurocode 3), Design of c o m p o s i t e steel and concrete structures ( E u r o c o d e 4), Design of
timber structures (Eurocode 5), Design of m a s o n r y structures (Eurocode 6), Geotechnical
design (Eurocode 7) and Design of structures in seismic regions (Eurocode 8).
E u r o c o d e 7 aims at creating standards for laboratory testing and for field testing and
sampling and at establishing c o m m o n rules for geotechnical design, including also
specific geotechnical structures, such as ground anchors, soil i m p r o v e m e n t , dewatering,
bridges, off-shore structures, etc. E u r o c o d e 7 is expected to b e finalised and approved of
in 1997.

3.2 Design philosophy of Eurocode

(i) Limit states. T h e design of structures should b e carried out with the view to satisfying
the requirements connected with states, b e y o n d w h i c h the structure no longer satisfies the
design performance criteria. T h e r e are two limit states that h a v e to b e considered in the
design, n a m e l y the ultimate limit state associated with any form of failure or instability
496 Design aspects

that m a y endanger the safety of people, and the serviceability limit state associated with
malfunctioning of the structure due to settlement, causing discomfort to people, or
malfunctioning of m a c h i n e installations due to vibrations.
T h e design values to b e used in the two limit states are taken as the characteristic values
of the parameters involved modified b y certain partial safety factors. Moreover, a partial
safety factor with reference to model uncertainty is also introduced. This is another
approach than using l u m p safety factors as w a s previously the case. T h e partial safety
factors h a v e been m o r e or less adjusted with regard to the results of previous design. It
has been considered m o r e appropriate from a statistical point of view to introduce this
new safety philosophy.

in) Actions. In the c o d e a distinction is m a d e between forces taking part in action and soil
parameters taking part in reaction. In geotechnical design, not only the external forces are
considered as action but also the weight of the soil and rock, the internal stress situation
(effective stresses, pore pressure, etc.), changes in stress situation (for instance due to
excavation), influence of vegetation or climatic variations, m o v e m e n t s (for instance due
to consolidation, decomposition, solution, creep), temperature effects, etc.
For the ultimate limit state, the design values of actions should either b e assessed
directly or b e based on characteristic values modified by a certain partial safety factor.
T h e characteristic values depend u p o n the type of supported structure. For soil and rock
density, the characteristic value is defined as for soil properties in general. T h e values of
the partial safety factor are chosen with due regard to the uncertainty in the values of
action.
For the serviceability limit states, the design values can generally b e taken equal to the
characteristic values.

(iii) Soil parameters. As in the case of actions, the design values of the soil properties
should either b e assessed directly or b e based on characteristic values modified by a
certain partial safety factor for the soil properties.
T h e characteristic values of soil properties is a cautious estimate of the m e a n value
observed over a certain surface or v o l u m e of ground, relevant to the design problem.
Statistical m e t h o d s (such as Baysian statistical m e t h o d s ) m a y b e e m p l o y e d in the
selection of characteristic values. In this case, the characteristic value should b e derived
such that the calculated probability of a w o r s e value governing the occurrence of a limit
state is not greater than 5 %. A s s u m i n g that the material property \ normally distributed
with a m e a n value E(X) and a coefficient of variation V(X), then the 5:th percentile
corresponds to i/<1.64). T h e characteristic value in this case is:

Xk = [l-l.64V(X)]E(X) (550)
Design aspects 497

Fig. 360. Failure of earth fill cofferdam caused by insufficient bearing capacity. Subsoil consisting of
clay with silt layers. An earth fill cofferdam is a typical example of a structure belonging to category 3.

(iv) Geotechnical categories. Naturally, the extent of the soil investigation required as a
basis for geotechnical design depends upon the type of structure and risks involved. In
the E u r o c o d e , three geotechnical categories are defined:

category 1 , including small and relatively simple structures 'for which it is possible to
ensure that the fundamental requirements will b e satisfied on the basis of experience and
qualitative geotechnical investigations' and w h e r e there is negligible risk for property
and life,
category 2, including 'conventional types of structures and foundations with no
abnormal risks or unusual or exceptionally difficult ground or loading c o n d i t i o n s ' ,

Fig. 361. Cellular cofferdam, a typical example of a structure belonging to category 3.


498 Design aspects

category 3, including 'very large or unusual structures, structures involving abnormal


risks, or unusual or exceptionally difficult ground or loading conditions and structures
in highly seismic a r e a s ' .
It m a y b e difficult, from a legal point of view, to d r a w the border lines b e t w e e n the
different categories. This j u d g e m e n t has to be left to the geotechnical expertise. If
something should go w r o n g , however, there will always b e a judicial procedure, and then,
certainly, the question regarding w h i c h of the three categories the project belongs to will
be one of the m a i n issues.
T w o projects illustrating the need for a sound j u d g e m e n t of design categories are
presented in Figs. 3 6 0 - 3 6 1 . In both cases w e h a v e to deal with cofferdams w h e r e a failure
would entail a considerable risk of catastrophic consequences. All the same, the
investigation used as a basis for the design of the earth fill cofferdam s h o w n in Fig. 360
w a s of a very limited extent which, as a consequence, resulted in d a m failure due to
insufficient bearing capacity of the subsoil (circular slide surface). Fortunately, in this
case the failure took place after working-hours and, therefore, brought about no
casualties.
As regards the cellular cofferdam, shown in Fig. 3 6 1 , a very detailed investigation and
follow-up during the construction was carried out, in full a g r e e m e n t with the content of
category 3.

3.3 ISO standards

T h e I S O w o r k on international standards in geotechnics has been in process for several


years in parallel with the w o r k on the Eurocode. T h e people e n g a g e d in the I S O com-
mittee on geotechnical standards are to a great extent also e n g a g e d in the w o r k on Euro-
code, and until the latter w o r k has been completed, an international standard in geotech-
nical design approved of is not to b e expected. It has been a very laborious task to reach
an agreement regarding the extent of detailed information and obligatory principles of
design to b e included in the E u r o c o d e and whether or not calculation m e t h o d s should be
included. Considering the great variety of design rules and geotechnical experience in the
different parts of the world, the time required to reach an international acceptance of a
universally prevailing standard, whether it concerns merely geotechnical investigation
routines or geotechnical design as a whole (an international code), seems indefinite.
References 499

REFERENCES

Akagi, T., 1976. Effect of displacement type sand drains o n strength and compressibility
of soft clays. Ph. D . Thesis, University of Tokyo.
Andrasson, . & H a n s b o , S., 1977. C o m p a c t i o n control by d y n a m i c m e t h o d s . V g - och
Vattenbyggaren, N o . 8 - 9 .
Andrasson, , 1979. Deformation characteristics of soft, high-plastic clays under
d y n a m i c loading conditions. Ph. D. Thesis, C h a l m e r s University of Technology,
Gothenburg, D e p . of Geotechnical Engineering.
Andrasson, L., 1974. Frslag till ndrade reduktionsfaktorer vid reduktion av vingborr-
bestmd skjuvhllfasthet m e d ledning av flytgrnsvrdet. (Proposal for altered correction
factors w h e n reducing field vane shear strength on the basis of the liquid limit value).
Internal report, C h a l m e r s University of Technology, Gothenburg.
Baguelin, F., Jzequel, J. F. & Shields, D. H., 1978. T h e pressuremeter in foundation
engineering, Trans. Tech. Publications, Paris.
Barden, L. & Sides, G., 1971. Sample disturbance in the investigation of clay structure.
G o t e c h n i q u e 2 1 , N o . 3, 2 2 1 - 2 2 2 .
Barkan, D . D., 1962. D y n a m i c s of basis and and foundations. M c G r a w - H i l l B o o k
C o m p a n y , N e w York.
Barker, J. ., 1 9 8 1 . Dictionary of Soil m e c h a n i c s and F o u n d a t i o n E n g i n e e r i n g .
Construction Press, L o n d o n and N e w York.
Barron, R. ., 1944. T h e influence of drain wells on the consolidation of finegrained
soils. (R. I.) Ph. D. Thesis, Providence, U. S. Engineering Office.
Barron, R. ., 1948.Consolidation of fine-grained soils by drain wells. Proc. A S C E ,
Paper N o . 2 3 4 6 (with discussions).
Barton, J O, 1982. Laterally loaded m o d e l piles in sand: Centrifuge tests and finite
element analysis. Dissertation, Univ. of C a m b r i d g e .
Beigler, S.-., 1976. Soil-structure interaction under static loading. Ph. D . Thesis,
C h a l m e r s University of Technology, Gothenburg.
Berezantsev, V. G., Khristoforov, V. S. & Golubkov, V. N., 1961. L o a d bearing capacity
and deformation of piled foundations. Proc. 5th Int. Conf. Soil M e c h . F o u n d . Eng.,
Paris, Vol. 2, 1 1 - 1 5 .
Bergau, W., 1959. M e a s u r e m e n t s in grain silos. Swedish Geotech. Inst., Proc. N o . 17.
Bergdahl, U. & Eriksson, U., 1983. B e s t m n i n g av j o r d e g e n s k a p e r m e d s o n d e r i n g e n
litteraturstudie. (Determination of soil characteristics by s o u n d i n g a literature survay).
S w e d i s h Geotech. Inst., Report N o . 22.
500 References

Bergfeit, ., 1964. Armeringsjrn som plar. (Reinforcement bars as piles). Halmstad


Jrnverk, Tekn. m e d d e l a n d e No. 17.
Berggren, B . , H a n s b o , S. & Lundahl, B., 1977. Prediction and i m p r o v e m e n t of static
behaviour of large bored piles by dynamic preloading. Vg- och Vattenbyggaren, No. 8-9.
Berggren, B. & Bengtsson, P.-E., 1985. Grvpalar i friktionsjord. Anvisningar for
d y n a m i s k frbelastning. (Bored piles in cohesionless soil. Instructions for d y n a m i c
preloading). R. Swedish Academy of Eng. Sc., Commission on Pile Research, Report 77.
Bernander, S. & Svensk, I., 1970. Plars brfrmaga i elastiskt m e d i u m under hnsyns-
tagande till initialkrkning och egenspnningar i plmaterial. (The bearing capacity of
piles in an elastic m e d i u m with regard to initial curvature and self-induced stresses in
pile material). R. Swedish A c a d e m y of Eng. Sc., C o m m i s s i o n on Pile Research,
Offprints and Preliminary Reports 2 3 .
Berntson, J. ., 1983. Portrycksvariationer i leror i Gteborgsregionen. (Pore pressure
variations in clays in the Gothenburg region). Licentiate's Dissertation, Chalmers
University of Technology, Gothenburg.
Bishop, A. W., 1955. T h e use of the slip circle in the stability analysis of slopes.
Gotechnique, Vol V, N o . 1, 1-17.
Bjerrum, L., 1966. Generelle krav till fundamentering av forskjellige byggverk. (General
requirements on foundations of various buildings). Norwegian Association of Engineers,
NIF-course, Bergen.
Bjerrum, L., 1973. Problems of soil mechanics and construction on soft clays and
structurally unstable soils (collapsible, expansive and others). Proc. 8th Int. Conf. Soil
M e c h . Found. Eng., Moscow, Vol. 3. (State-of-the-art Report).
B o m a n , P., 1973. N y metod att drnera lera. (A n e w method of draining clay.) V g - och
Vattenbyggaren, N o . 2, 192.
B o w l e s , J. E., 1988. Foundation analysis and design. M c G r a w Hill, N e w York (4th
Edition).
Brandl, H., 1993. Installation, monitoring and design of caissons. Proc. 2nd Int. Geotech.
Seminar on on D e e p Foundations on Bored and A u g e r Piles, Ghent, 3 - 2 0 .
Briaud, J.-L., 1992. T h e pressuremeter. A. A. Balkema, Rotterdam.
Brinch H a n s e n , J., 1953. Earth pressure calculations. Teknisk Forlag, C o p e n h a g e n .
Brinch Hansen, J., 1967. St0ttemurers baereevne. (The bearing capacity of retaining
walls). B-undervisning og forskning 67, D a n m a r k s Ingenj0rakademi, D I A B .
B r o m s , . , 1964. Lateral resistance of piles in cohesive soils. Proc. A S C E , J. S M F D , N o .
S M 2.
B r o m s , . , 1964. Lateral resistance of piles in cohesionless soils. Proc. A S C E , J. S M F D ,
N o . S M 3.
B r o m s , . , 1976. Pile foundationspile groups. Proc. 6th E u r o p e a n Conf. Soil M e c h .
Found. E n g . , Vienna, 2 . 1 , 1 0 3 - 1 3 2 .
References 501

B r o m s , . & H a n s b o , S., 1981. Foundations on soft clay. In Brand, E. W. & Brenner, R.


R (Editors), Soft Clay Engineering. Elsevier Sc. Publ. Co., Chapter 6.
Burland, J. B . , B r o m s , . & D e Mello, V., 1977. Behaviour of foundations and structures.
Proc. 9th Int. Conf. Soil M e c h . Found. Eng., Tokyo, Vol. 2, 4 9 5 - 5 4 6 .
Burland, J., 1986. T h e value of field measurements in the design and construction of deep
foundations. Proc. Int. Conf. on D e e p Foundations, Beijing, Vol. 2, 1 7 7 - 1 8 7 .
Burland, J. B . , 1989. Ninth Laurits Bjerrum M e m o r i a l Lecture: "Small is beatiful" the
stiffness of soils at small strains. Can. Geotech. J. 26, 4 9 9 - 5 1 6 .
Bustamante, M . & Gianeselli, L., 1982. Pile bearing capacity prediction b y m e a n s of
static penetrometer CPT. Proc. Int. S y m p o s i u m on Penetration Testing, A m s t e r d a m ,
Vol. 2, 4 9 3 - 5 0 0 .
Cadling, L. & Odenstad, S., 1950. T h e vane borer. A n apparatus for determining the shear
strength of clay soils directly in the ground. Swedish Geotech. Inst., Proc. N o . 2.
Carillo, N., 1942. Simple t w o and three dimensional cases in the theory of consolidation
of soils. J. M a t h . Phys., Vol. 2 1 , N o . 1.
Casagrande, ., 1947. Classification and identification of soils. Proc. A S C E , 7 3 : 6 , 7 8 3 -
810.
Casagrande, L., 1947. T h e application of electro-osmosis to practical p r o b l e m s in
foundations and earthworks. D e p . of Sc. and Ind. Res., Building Res., Technical Paper
30.
Cassan, M., 1966. L e tassement des pieux. Synthse des recherches rcentes et essais
comparatifs. Sols-Soils, Vol. 5, N o . 1 8 - 1 9 .
Chellis, R. D., 1 9 5 1 . Pile foundations. T h e o r y D e s i g n P r a c t i c e . M c G r a w - H i l l , N e w
York.
C h a n g , Y C. E., 1981. L o n g - t e r m consolidation beneath the test fills at Vsby. S w e d i s h
Geot. Inst., Report N o . 18.
Choa, V , 1989. Drains and v a c u u m preloading pilot test. Proc. 12th Int. Conf. Soil M e c h .
Found. Eng., R i o deJaneiro, Vol. 2, Paper 18/6, 1 3 4 7 - 1 3 5 0 .
D ' A p p o l o n i a , D . J., D ' A p p o l o n i a , . & Brissette, R. F., 1970. Discussion on settlement
of spread footings on sand. A S C E , J. S M F D 9 6 , N o . S M 2 , 7 5 4 - 7 6 2 .
Darcy, H., 1856. L e s fontaines publiques de la ville de Dijon. Paris
Davisson, M . T. & Robinson, K. E., 1965. Bending and buckling of partially embeddedpiles.
Proc. 6th Int. Conf. Soil M e c h . Found. Eng., Montreal, Vol. 2.
Davisson, M . T., 1970. B R D vibratory driving formula. Foundation Facts, Vol. 6, N o . 1.
Dcourt, L., 1982. Prediction of the bearing capacity of piles b a s e d exclusivly on
values of the SPT. Proc. Int. S y m p o s i u m on Penetration Testing, A m s t e r d a m , Vol 1 , 2 9 -
34.
D e Josselin de Jong, G., 1957. Application of stress function to consolidation p r o b l e m s .
Proc. 4th Int. Conf. Soil M e c h . Found. Eng., L o n d o n , Vol. 1.
502 References

Dubin, . & Molin, G., 1986. Influence of a critical gradient on the consolidation of clays.
In R. . Yong & F. C. Townsend (Editors), Consolidation of soils: testing and
evaluation. A S T M Special Tech. Publ. 892, 3 5 4 - 3 7 7 .
D u n c a n , J. & C h a n g , C . - Y , 1970. Nonlinear analysis of stress and strain in soil. A S C E ,
J. Soil M e c h . Found. Eng., Vol. 96, N o . S M 5.
Du Thinh, ., 1984. T h e built-in stiffness of footings in sand. Ph. D. Thesis, C h a l m e r s
University of Technology, Gothenburg.
Ekstrm, J., 1989. A field study of model pile group behaviour in non-cohesive soils.
Influence of compaction due to pile driving. Ph. D. Thesis, C h a l m e r s University of
Technology, Gothenburg.
Eriksson, L. & Ekstrm, ., 1983. T h e efficiency of three different types of vertical
drainsresults from full-scale tests. Proc. 8th European Conf. Soil M e c h . Found.
Eng., Helsinki, Vol. 2, Paper
Faber, O., 1933. Pressure distribution under basis and stability of foundations. Struct.
Engineer.
Fellenius, W., 1918. Kaj- och jordrasen i Gteborg. (The quay and earth slides in
Gothenburg). Teknisk Tidskrift, Hafte 2, 1 7 - 1 9 .
Fellenius, W., 1926. Erdstatische Berechnungen. Wilhelm Ernst & Sohn.
Fellenius, W., 1936. Calculations of the stability of earth d a m s . Proc. 2nd Int. Conf. on
Large D a m s , Washington D.C., Vol. 4, 4 4 5 ^ 4 4 9 .
Flodin, N . , 1973. S v e n s k a pappdrner i Canada. (Swedish cardboard wicks in Canada.)
Vg- o c h Vattenbyggaren, N o . 3, 320.
Forssblad, L., 1 9 8 1 . Vibratory soil and rock fill compaction. D y n a p a c M a s k i n A B ,
Stockholm.
Garg, K. G., 1979. Bored pile groups under vertical loads in sand. J. Geot. E n g . Div.,
A S C E , Vol. 105, N o . G T 8.
Gereben, L. & P r a m b o r g , B . , 1990. Rayleighvagsmtningar i friktionsjord och fast lera.
(Rayleigh w a v e m e a s u r e m e n t s in frictional soil and stiff clay). Swedish Council for
Building Res., Report R 1 0 2 : 1 9 9 0
Gilbert, P. ., 1991. Rapid Water content by computor-controlled m i c r o - w a v e drying.
A S C E , J. Geot. Eng., Vol. 117, N o . 1, 1 1 8 - 1 3 8 .
Goble, G. G., Garland Likings, J. & Rausche, F., 1975. Bearing capacity of piles from
d y n a m i c m e a s u r e m e n t s . Final Report, Dept. of Civil Eng., C a p e Western Univ.,
Cleveland, O h i o .
G o r b u n o v - P o s s a d o v , M . I. & Serebrajanyi, R. V , 1961. Design of structures u p o n elastic
foundations. Proc. 5th Int. Conf.Soil M e c h . F o u n d . Eng., Paris, Vol. 1, 6 4 3 - 6 4 8 .
H a n n a , ., 1963. M o d e l studies of foundation groups in sand. G o t e c h n i q u e , Vol. 13.
H a n n a , ., 1982. Foundations in t e n s i o n G r o u n d anchors. Trans Tech Publ., Series
on R o c k and Soil Mech., Vol. 6.
References 503

Hansen, . , 1 9 6 1 . Shear b o x tests on sand. Proc. 5th Int. Conf. Soil M e c h . Found. Eng.,
L o n d o n , Vol. I, 1 2 7 - 1 3 1 .
H a n s b o , S., 1957. A n e w approach to the determination of the shear strength of clay by
the fall-cone test. Proc. Swedish Geotech. Inst., N o 14.
H a n s b o , S., 1960. Consolidation of clay, with special reference to influence of vertical
drains. S w e d i s h Geotech. Inst., Proc. N o . 18. Ph. D. Thesis, C h a l m e r s University of
Technology.
H a n s b o , S., 1973. Influence of mobile particles in soft clay on permeability. Proc. Int.
S y m p . on Soil Structure, Gothenburg, 1 3 1 - 1 3 5 .
H a n s b o , S., 1975. Jordmateriallra. (Soil material science). A W E / G e b e r s , Stockholm.
H a n s b o , S., 1977. D y n a m i c consolidation of rock fill at Uddevalla ship yard. Proc. 9th
Int. Conf. Soil M e c h . Found. Eng., Tokyo, Vol. 2, 2 4 1 - 2 4 6 .
H a n s b o , S., 1978. D y n a m i c consolidation of soil by a falling weight. G r o u n d Engineering,
N o . 5, July.
H a n s b o , S., 1979a. Discussion on design parameters in geotechnical engineering, Proc.
7th E u r o p e a n Conf. Soil M e c h . Found. Eng., Brighton, Vol. 4, 1.22, 4 5 .
H a n s b o , S., 1979b. Design parameters for soft clays. Proc. 7th European Conf. Soil
M e c h . Found. Eng., Brighton, Vol. 4, 1.22, 141.
H a n s b o , S. & P r a m b o r g , B., 1980. Compaction control. Int. Conf. on C o m p a c t i o n . Paris,
Vol. II, 5 5 9 - 5 6 4 .
H a n s b o , S., 1 9 8 1 . Consolidation of fine-grained soils by prefabricated drains. Proc. 10th
Int. Conf. Soil M e c h . Found. Eng., Stockholm, Vol. 3, Paper 12/22.
H a n s b o , S., 1983. H o w to evaluate the properties of prefabricateddrains. Proc. 8th
E u r o p e a n Conf. Soil M e c h . Found. Eng., Helsinki, Vol. 2, Paper 6.13.
H a n s b o , S., 1984. Foundations on friction creep piles in soft clays. Proc. Int. Conf. on
C a s e Histories in Geot. Eng., St Louis, Vol. 2, 9 1 3 - 9 2 2 .
H a n s b o , S. 1986. Preconsolidation of soft compressible subsoil b y the use of prefabricated
vertical drains. Tijdschrift der openbare werken van Belgi, N o . 6
H a n s b o , S., 1987. Fact and fiction in the field of vertical drainage. In: R. C. Joshi & F.
J. Griffiths (Editors), Prediction and Performance in Geotechnical Engineering. A. A.
Balkema, pp. 6 1 - 7 2 .
H a n s b o , S., 1987. State-of-the-art report on groundwater problems in S w e d e n . Proc. 9th
E u r o p e a n Conf. Soil M e c h . Found. Engng., Dublin, Vol. 1, 3.4, 2 9 3 - 2 9 8 .
H a n s b o , S. & P r a m b o r g , B., 1990. Experience of the M n a r d pressuremeter in foundation
design. Proc. 3rd Int. S y m p . on Pressuremeters, Oxford.
Hansbo, S., 1992. B a n d drains. In: M . P. Moseley (Editor), G r o u n d I m p r o v e m e n t .
Blackie A c a d e m i c & Professional, C R C Press, Inc., 4 0 - 6 4 .
Hardin, B . O. & Drnevich, V P., 1972. Shear m o d u l u s and d a m p i n g of soils: Design
equations and curves. Proc. A S C E , J. S M F D , Vol 9 8 , N o S M 7.
504 References

Hardin, 1978. T h e nature of stress-strain behaviour of soils. Proc. A S C E Specialty Conf.


on Earthquake E n g . and Soil D y n a m i c s , Pasadena, Vol. 1, 3 - 9 0 .
Harr, M., 1977. M e c h a n i c s of particulate media. A probabilistic approach. M c G r a w - H i l l ,
N e w York.
Harr, M . , 1987. Reliability based design in civil engineering. M c G r a w - H i l l , N e w York.
H a r d e n , J., 1974. S k n s k a mornlerors hllfasthets- och deformationsegenskaper.
(Strength and deformation characteristics of Scanian boulder clays). Ph. D . Thesis,
C h a l m e r s University of Technology, Gothenburg, D e p . of Geotech. E n g .
H a u s m a n n , . ., 1990. Engineering principles of ground modification. M c G r a w - H i l l ,
N e w York.
Helenelund, . V , 1 9 5 1 . O m konsolidering och sttning av belastade marklager. (On
conolidation and settlement of loaded soil layers). Ph. D. Thesis, Jord och vattentekniska
forskningar 6, Helsinki.
Helenelund, . V , 1965. Investigations on the bearing capacity and engineering
properties of silt. V T T : N , Tiedotus, Sarja 3 , R a k e n n u s 8 5 .
Holtz, R. D . & B r o m s , . , 1972. Long-term loading tests at Sk-Edeby, Sweden. Proc.
Spec. Conf. on Performance of Earth-Supported Structures, Vol. 1, Purdue University,
Lafayette, Indiana.
Holtz, R. D . & H o l m , G., 1973. Excavation and sampling around s o m e drains at Sk-
Edeby, S w e d e n . Proc. Nordic Geot. Meeting in Trondheim 1972, N o r w e g i a n Geotech.
Inst.
Holtz, R. D . & K o v a c s , W . D., 1981. A n introduction to geotechnical engineering. Text
b o o k (733 p.), Prentice-Hall, Inc., Englewood Cliffs, N e w Jersey.
Hultin, S., 1916. Grusfyllningar for kajbyggnader. (Gravel fill for quay constructions).
Teknisk Tidskrift, Hafte 3 1 , 2 9 2 - 2 9 4 .
Hutchinson, J. N . , 1 9 6 1 . A landslide on a thin layer of quick clay at Furre, Central
Norway. N o r w e g i a n Geotech. Inst., Publ. N o . 4 4 .
Imai, T. & Tonouchi, K., 1982. Correlation of TV value with S-wave velocity and shear
m o d u l u s . Proc. 2nd European S y m p . on Penetration Testing, A m s t e r d a m , V o l . 1, 6 7 -
72.
Ivanov, P. L., 1978. Liquefaction and consolidation of noncohesive soils b y d y n a m i c
action. N . J. Kalinin Polytechn. Inst. (In Russian)
Ivanov, P. L., 1980. Consolidation of saturated soils by explosions. Proc. Int. Conf. on
C o m p a c t i o n , Paris, Vol. 1.
Jacobson, . & Odenstad, S., 1940. Ett geotekniskt stabilitetsproblem m e d speciell till-
lmpning p tryckbankar. (A geotechnical stability problem with special application to
loading berms). Teknisk Tidskrift, Vg- och Vattenbyggnadskonst, Husbyggnadsteknik,
Hafte 2, 1 7 - 2 3 .
J a m i o l k o w s k i , M . , Lancelotta, R. & Wolski, W., 1983. S u m m a r y of discussion. Proc. 8th
European Conf. Soil M e c h . Found. Eng., Helsinki, Vol. 3, Spec. Session 6 , 1 2 4 2 - 1 2 4 5 .
References 505

J a m i o l k o w s k i , M . & Lancelotta, R., 1984. E m b a n k m e n t s on vertical drains Pore


pressures during construction. Proc. Int. Conf. on C a s e Histories in Geot. Eng., St
Louis, Vol. I, 2 7 5 - 2 7 8 .
J a m i o l k o w s k i , M . & Robertson, P. K., 1988. Future trends for penetration testing. Proc.
Conf. on Penetration Testing in the U K , B i r m i n g h a m .
Janbu, N., 1954. Stability analysis of slopes with dimensionless parameters. Ph. D.
Thesis, H a r v a r d University, C a m b r i d g e , M a s s .
Janbu, N., Bjerrum, L. & Kjaernsli, B., 1956. Veiledning v e d l 0 s n i n g av fundamenterings-
opgaver. ( G u i d a n c e on the solution of foundation problems). N o r w e g i a n Geotech.
Inst., Publ. N o . 18.
Janbu, N., 1973. T h e generalized procedure of slices. E m b a n k m e n t D a m Engineering,
C a s a g r a n d e Volume, 4 7 - 8 6 .
Janbu, N . & Senneset, K., 1973. Field compressometerprinciples and applications.
Proc.8th Int. Conf. Soil M e c h . Found. Eng., Moscow, Vol. 1.1.
Janssen, . ., 1895. Versuche ber Getreidedruck in Silozellen. Zeitschrift vereinigter
deutschen Ingenieure, B a n d 39.
Jelinek, R., 1966. Eigenschaften des B o d e n s . Grundbau-Taschenbuch, B a n d 1, 2:e
Auflage, Verlag v. Wilhelm Ernst & Sohn, Berlin-Mnchen, Chapter 1.2, 11.
Jendeby, L., 1986. Friction piled foundations in soft clay. A study of load transfer and
settlements. Ph. D . Thesis, C h a l m e r s University of Technology, Gothenburg.
Johnson, S. J., 1970. Foundation precompression with vertical sand drains. J. Soil M e c h .
Found. Div., A S C E , Vol. 96, S M I.
Karlsson, R., 1977. Consistency limits. M a n u a l for the Performance and Interpretation
of Laboratory Investigations, Part 6, Swed. Council for Building Research, D 6 : 1 9 7 7 .
Karlsson, R. & H a n s b o , S., 1984. Jordarternas indelning och b e n m n i n g . (Soil classifica-
tion and identification). M a n u a l for the Performance and Interpretation of Laboratory
Investigations, Part 2, Swed. Council for Building Research, T 2 L 1 9 8 2 .
K a m o n , M., 1984. Function of band-shaped prefabricated plastic board drain. Proc. 19th
Jap. Natl. Conf. on Soil M e c h . Found. Eng.
Kishida, H. & Meyerhof, G. G., 1965. Bearing capacity of pile groups u n d e r eccentric
loads in sand. Proc. 6th Int. Conf. Soil M e c h . Found. Eng., Toronto, Vol. 2
Kjellman, W., 1948. Consolidation of fine-grained soils by drain wells. Trans. A S C E ,
Vol.113. (Contribution to the discussion).
Koda, E., S z y m a n s k i , A. & Wolski, W., 1986. Laboratory tests on G e o d r a i n s D u r a b i l i t y
in organic soils. S e m i n a r on Laboratory Testing of Prefabricated B a n d - S h a p e d Drains,
Milano, April 2 2 - 2 3 .
Kok, L. & Trense, R. W., 1979. Blast-induced soil liquefaction. State-of-the-art Report,
6th I M A B S , C a h o r s , France.
Koskinen, M., 1 9 9 1 . Horizontal capacity of steel pipe pile. Proc. 10th E u r o p e a n Conf.
Soil M e c h . F o u n d . Eng., Florence, Vol 2, 4 5 7 ^ 1 6 2 .
506 References

Kupfmller, ., 1959. Einfhrung in die theoretische Elektrotechnik. Springer, Berlin,


6th Edition.
Knen, M., 1896. Berechnung des Seiten- und Bodendruckes in Silozellen. Zentralblatt,
Zeitschrift der Bauverwaltung.
Ladd, C. C , 1976. U s e of precompression and vertical sand drains for stabilization of
foundation soils. Presented at Soil and Site Improvement, Continuing Education in
Engineering University of California, Berkely.
Larsson, R., 1986. Consolidation of soft soils. Swedish Geotech. Institute, Report N o . 29.
Larsson, R. & Sllfors, G., 1986. Automatic continuous consolidation testing in Sweden.
In R. N . Yong & F. C. Townsend (Editors), Consolidation of soils: testing and
evaluation. A S T M Special Tech. Publ. 892.
Larsson, R. & M u l a b d i c ' , ML, 1991. Shear moduli in Scandinavian clays. Swedish
Geotech. Inst., Report N o . 4 0 .
L a w s o n , C. R., 1992. Soil reinforcement with geosynthetics. Workshop on Applied
ground i m p r o v e m e n t techniques, S E A G S and AIT, Bangkok.
Liu, J. L., Yan, Z. L. & Shang, K. P., 1985. Cap-pile-soil interaction of bored pile groups.
Proc. 11th Int. Conf. Soil Mech. and Found. Eng., San Francisco, Vol. 3, 1433-1466.
Lo, D., ., K., 1 9 9 1 . Soil improvement by vertical drains. Ph. D . Thesis, University of
Illinois at U r b a n a - C h a m p a i g n .
Lo Presti, D . & Lai, C , 1989. Shear w a v e velocity from penetration test. Politecnico di
Torino, D e p . of Structural Eng., N o . 2 1 .
L o v e , . . H., 1929. T h e stress produced in a semi-infinite solid by pressure on part of
the boundary. R. Soc. Phil. Trans., Vol. 2 2 8 .
Lundgren, H. & Brinch Hansen, J., 1958. Geoteknik. Teknisk Forlag, C o p e n h a g e n .
M a g n a n , J. P., 1983. Thorie et pratique des drains verticaux. Tec & D o c - L a voisier,
Paris.
Malik, Z., M o r t o n , J. & Ruiz, C , 1977. Ovalization of cylindrical tanks as a result of
foundation settlement. J. of Strain Analysis, Vol. 12, N o . 4, 3 3 9 - 3 4 8 .
Marchetti, S., 1975. A n e w in situ test for m e a s u r e m e n t of horizontal soil deformability.
Proc. Conf. on In Situ M e a s u r e m e n t of Soil Prop., Raleygh, N . C , Vol. 2, A S C E .
Marr, W. ., R a m o s , J. A. & L a m b e , T. W., 1982. Criteria for settlement of tanks. Proc.
A S C E , J. Geotech. E n g . Div., Vol. 108, N o . G T 8, 1 0 1 7 - 1 0 3 9 .
Mascardi, C. 1970. Il c o m p o r t a m e n t o dei micropali sottoposti a sforzo assiale, m o m e n t o
flettente e taglio. R o d i o , Publ. 12.
Massarsch, R., 1991. D e e p soil compaction using vibratory probes. A S T M S y m p o s i u m
on Design, Construction and Testing of D e e p Foundation Improvement: Stone C o l u m n s
and Related Techniques. A S T M Special Technical Publ., Philadelphia, 2 9 7 - 3 1 9 .
(Editor: Robert C. Bachus).
Mesri, G., 1990. Personal communication.
References 507

Mesri, G., Feng, T. W. & Benak, J. M., 1990. Postdensification penetration resistance of
clean sands. Proc. A S C E , J. Geotech. Eng., Vol. 116, N o 7, 1 0 9 5 - 1 1 1 5 .
Meyerhof, G. G., 1951. T h e ultimate bearing capacity of foundations. G o t e c h n i q u e , Vol.
II, N o . 4, 3 0 1 - 3 3 2 .
Meyerhof, G. G., 1 9 6 1 . Discussions. 5th Int. Conf. Soil M e c h . Found. Eng., Paris, Vol.
3, 193.
Meyerhof, G. G., 1963. S o m e recent research on the bearing capacity of foundations.
Can. Geotech. J., Vol. 1, N o . 1.
Meyerhof, G. G., 1976. Bearing capacity and settlement of pile foundations. Proc. A S C E ,
J. Geotech. Eng. Div., N o . G T 3.
Morgenstern, N . R. & Price, V. ., 1965. T h e analysis of the stability of general slip
surfaces. G o t e c h n i q u e , Vol. XV, N o . 1.
Mourn, J., 1967. Elektro-osmosevirkning og anvendelse innen geoteknikken. (Electro-
osmosisfunctioning and utilisation in geotechnical engineering), T h e N G F Lecture,
N o r w e g i a n Geotech. Soc.
N e w m a r k , N . M., 1942. Influence charts for computation of stresses in elastic foundations.
University of Illinois, Bull. N o . 4 5 .
Nguyen, T. Tien, 1987. D y n a m i c and static behaviour of driven piles. Ph. D. Thesis.
C h a l m e r s University of Technology. Swedish Geotech. Inst., Report N o . 3 3 .
N o r w e g i a n Pile C o m m i t t e e , 1973. Veiledning ved pelefundamentering. (Guidelines for
pile foundations). N o r w e g i a n Geotech. Inst.
Odenstad, S., 1956. Markbrighet och tryckbankar vidkohesionsjord. (Bearing capacity
and loading b e r m s in the case of cohesive soils). Teknisk Tidskrift, N o v e m b e r , 1 0 3 3 -
1035.
Odenstad, S., 1960. Markbrighet och tryckbankar vid kohesionsjord. (Bearing capacity
and loading berms in the case of cohesive soils). Vg- och Vattenbyggaren, No. 2, 6 0 - 6 2 .
O n o u e , ., 1988. Consolidation of multilayered anisotropic soils by vertical drains with
well resistance. Soils and Foundations, Vol. 2 8 , N o . 3 , 7 5 - 9 0 , Japanese Soc. Soil M e c h .
Found. E n g .
Parry, R., 1 9 7 1 . A direct m e t h o d of estimating settlement in sands from SPT-values.
A S C E , G T 9, 2 9 - 3 7 .
Peck, R. B . , H a n s o n , W. E. & Thornburn, T. H., 1973. Foundation Engineering. John
Wiley & Sons, N e w York, Second Edition.
Pettersson, K., 1916. Kajraseti Gteborg 5 mars 1916. (The quay collapse in Gothenburg,
M a r c h 5, 1916). Teknisk Tidskrift, Hafte 3 1 , 2 8 9 - 2 9 1 .
Phung, D . L o n g , 1993. Footings with settlement reducing piles in non-cohesive soils. Ph.
D . Thesis. C h a l m e r s University of Technology, Gothenburg,
von Post, L , 1 9 2 1 . U p p l y sningar rrande S veriges Geologiska Undersknings torvmarks-
rekognocering. (Information regarding the peatland exploration of the Geological
Survay of S w e d e n ) . Sv. Geol. U n d e r s k n i n g , Ser. D , Stockholm.
508 References

Poulos, H. G., 1968. Analysis of the settlement of pile groups. Gotechnique, Vol. 108,
449-471.
Poulos, H. G., 1993. Settlement prediction for bored pile groups. Proc. 2nd Int. Geotech.
Seminar on D e e p Foundations on Bored and A u g e r Piles, 1 0 3 - 1 1 8 .
Pusch, R., 1967. A technique for investigation of clay microstructure. J. de Microscopie,
Vol. 6, 9 6 3 - 9 8 6 .
Randolph, M . F. & Clancy, P., 1993. Efficient design of piled rafts. Proc. 2nd Int. Geotech.
Seminar on D e e p Foundations on Bored and A u g e r Piles, 1 1 9 - 1 3 0 .
Rausch, ., 1959. Maschinenfundamente und andere dynamisch beanspruchte B a u k o n -
struktionen. VDI-Verlag G m b H . , Dsseldorf.
Rausche, F., M o s e s , F. & Goble, G. G., 1972. Soil resistance predictions from soil
d y n a m i c s . Proc. A S C E , J. S M F D , Vol. 89, N o S M 9.
Reissner, E. & Sagoci, H. F., 1944. Forced torsional oscillations of an elastic half-space.
J. Applied Physics, Vol. 15, 6 5 2 - 6 6 2 .
Rethaty, L., T h e investigation of d a m a g e to buildings. Osnovanya, F u n d a m e n t y i M e k h a -
nika Gruntov, N o . 6.
Robertson, P. K., Campanella, R. G., Brown, P. T. & Robinson, . E., 1988. Prediction
of wick drain performance using piezometer cone data. Canadian Geotech. J. 2 5 , 5 6 -
61.
R o s e n b l u e t h , E . , 1981. Two-point estimates in probabilities. Appl. Math. M o d e l l i n g , Vol.
5, 3 2 9 - 3 3 5 .
Rosenblueth, E., 1975. Point estimates for probability m o m e n t s . Proc. Natl. Acad. Sc.
U S A , Vol. 72, N o . 10.
Rosenblueth, ., 1981. Two-point estimates in probabilities. Appl. Math. Modelling, Vol.
5, 3 2 9 - 3 3 5 .
R o w e , P., 1951. Cantilever sheet piling in cohesionless soil. Engineering, Vol. 172, N o .
4467,316-319.
R o w e , P., 1952. A n c h o r e d sheet pile walls. Proc. Institution of Civil Engineers, Pt. 1, Vol.
l , N o . 1,27-70.
Runesson, K., H a n s b o , S. & Wiberg, . E., 1985. T h e efficiency of partially penetrating
drains. G o t e c h n i q u e 3 5 , N o . 4, 5 1 1 - 5 1 6 .
Sahlstrm, P.-O. & Stille, H., 1979. Frankrade sponter. (Anchored sheet pile walls).
S w e d i s h Council for Building Research, T 3 0 : 1 9 7 9 .
Sandven, R., Senneset, K. & Janbu, N., 1988. Interpretation of piezocone tests in cohesive
soils. International Symposium on Penetration Testing, I S O P T - 1 , Orlando, USA, 939 - 953.
Seed, H. B . & Idriss, I. M., 1971. Simplified procedure for evaluating soil liquefaction
potential. J. S M F D , A S C E , Vol. 9 7 , S M 9, 1 2 4 9 - 1 2 7 3 .
S c h m e r t m a n n , J. H., 1955. T h e undisturbed consolidation behaviour of clay. A S C E , S o i l
M e c h . and F o u n d . Div., Vol. 9 6 , N o . S M 3.
References 509

S c h m e r t m a n n , J. H. & Osterberg, J. O., 1961. A n experimental study of the d e v e l o p m e n t


of cohesion and friction with axial strain in saturated cohesive soils. A S C E , R e s . Conf.
on Shear Strength of C o h e s i v e Soils, Boulder, Colorado.
S c h m e r t m a n n , J. H., 1970. Static cone to c o m p u t e static settlement over sand. A S C E , Soil
M e c h . and F o u n d . Div., Vol. 96, N o . S M 3.
Schultze, E. & Sherif, G., 1973. Prediction of settlement from evaluated settlement
observations for sand. Proc. 8th Int. Conf. Soil M e c h . Found. Eng., M o s c o w , Vol. 1.3,
225-230.
Sellgren, E., 1 9 8 1 . Friction piles in non-cohesive soils. Evaluation from pressuremeter
tests. Ph. D. Thesis, C h a l m e r s University of Technology, Gothenburg.
Sellgren, E. 1985. Prediction of the behaviour of friction piles in non-cohesive soils. Proc.
11th Int. Conf. Soil Mech. Found. Eng., San Francisco, Vol. 4, Paper A 3 1 , 1463-1468.
Sellgren, E., 1 9 9 1 . A soil anchoring technique in expansion. Proc. 10th E u r o p e a n Conf.
Soil M e c h . Found. Eng., Florence, Vol. 2, 7 4 5 - 7 4 8 .
Sellgren, E., 1992. Expanderplar och expanderstag. (Expander piles and e x p a n d e r
anchors). Statens rd for byggnadsforskning, Stockholm, T 3 : 1 9 9 2 .
Senneset, K., Sandven, R. & Janbu, N., 1989. T h e evaluation of soil parameters from
piezocone tests. S y m p o s i u m on Soil Properties for Transportation Facilities, 1235
Transportation Research Board, Natl. Res. Council, Washington D. C , 2 4 - 3 7 .
SGI, 1949. R e d o g r e l s e for Statens geotekniska instituts verksamhet under ren 1 9 4 4 -
1948. (Report on the activities of the Swedish Geotechnical Institute in the years 1944
- 1 9 4 8 ) . Swedish Geotech. Inst., M e d d e l a n d e N o 2.
Skempton, A. W., 1954. The pore pressure coefficients A and B. Gotechnique, Vol. 4, No. 4.
S k e m p t o n , A. W., 1954. Discussion of the structure of inorganic soil. Proc. A S C E , Soil
M e c h . Found. Div., Vol. 80, Separate N o . 4 7 8 .
Skempton, A. W., 1959. Cast in-situ bored piles in London clay. Gotechnique, No. 9 , 1 5 3 - 1 7 3 .
S k e m p t o n , A. W. & M c Donald, D. H., 1956. T h e allowable settlements of buildings.
Proc. Inst, of Civil Engineers, Part 3, Vol. 5.
S J: G e o t e k n i s k a K o m m i s s i o n e n , 1 9 1 4 - 1 9 2 2 . Slutbetnkande.(Swedish State Railways:
Geotechnical C o m m i s s i o n , 1 9 1 4 - 1 9 2 2 . Final report). Geotekniskt M e d d e l a n d e N o . 2,
Stockholm.
Sliwinski, Z. J. & Fleming, W. G. K., 1984. T h e integrity and performance of bored piles.
Piling and G r o u n d Treatment, T h o m a s Telford Ltd., L o n d o n , Paper 15, 2 1 1 - 2 2 3 .
Sowers, G. F , 1961. T h e bearing capacity of friction pile groups in h o m o g e n e o u s clay
from model studies. Proc. 5th Int. Conf. Soil Mech. Found. Eng., Paris, Vol. 2, 155 - 1 6 1 .
Steinbrenner, W., 1936. A rational method for determination of the vertical normal
stresses under foundations. Proc. Int. Conf. Soil M e c h . Found. Eng., C a m b r i d g e ,
Massachusetts, Vol. 2.
Stille, H., 1976. Behaviour of anchored sheet pile walls. Dissertation. R. Inst, of
Technology, S t o c k h o m
510 References

Sllfors, G. & Larsson, R., 1981. Berkning av sttningar i lera. (Calculation of


settlement in clay). Vg- och Vattenbyggaren N o 3, 3 9 ^ - 2 .
Taylor, D. W. & Merchant, W., 1940. A theory of clay consolidation accounting for
secondary compressions. J. Mathematics and Physics, N o . 1.
Terzaghi, K., 1955. Evaluation of coefficients of subgrade reaction. Gotechnique, Vol. 5.
Terzaghi, K. & Peck, R. B., 1967. Soil mechanics in engineering practice. John Wiley and
Sons Inc.
Thurner, H. & Sandstrm, ., 1991. Quality assurance in soil compaction. Proc. X l X t h
World R o a d C o n g r e s s , Marrakesh, Questions II, Construction and maintenance,
General questions 1.4 Quality Assurance Systems, 4 6 9 - 4 7 7 .
Thurner, H., 1992. Verdichtungstechnik und Verdichtungskontrolle der 90-er Jahre. 1st
Int. S y m p . on Technique and Technology in Road Construction, B a u m a 9 2 , M n c h e n .
T i m o s c h e n k o , S., 1936. Theory of elastic stability. M c G r a w - H i l l , N e w York.
T i m o s c h e n k o , S. & Gere J. M., 1961. Theory of elastic stability. M c G r a w - H i l l , N e w
York.
Torstensson, B.-A., 1973. Kohesionsplar i ls lera. En fltstudie i modellskala. (Cohesion
piles in soft clay. A field study in model scale). Ph. D. Thesis, C h a l m e r s University of
Technology, Gothenburg.
Tschebotarioff, G. P., 1 9 5 1 . Soil mechanics, foundations and earth structures. M c G r a w
Hill Civil Engineering Series, N e w York.
Turitzin, A. M., 1963. D y n a m i c pressure of granular material in deep bins. Proc. A S C E ,
J. Structural Div., St. 2.
Van I m p e , W. F., 1991. Deformations of deep foundations. Proc. 10th E u r o p e a n Conf.
Soil M e c h . Found. Eng., Florence, Vol.III, 1 0 3 1 - 1 0 6 2 .
Vesic, A. S., 1969. Experiments with instrumented pile groups in sand. A m e r i c a n Society
for Testing Materials, Performance of deep foundations, A S T M S T P 4 4 4 , 1 7 2 - 2 2 2 .
Walkley & Black, 1934. An examination of the Dergtjareff method for determining soil
organic matter and a proposed modification of the chromic and titration method, Soil
Science 37.
Wetzworke, ., 1960. ber die Bruchsicherheit von Rohrleitungen in parallellwndigen
Grben. Technische Hochschule Hannover, D e p . Public Water Supply, Heft 5.
Wiegers, H., 1974. T h e interaction between classification and terminology in engi-
neering geology and associated disciplines, particularly with relation to unconsolidated
sediments. Proc. 2nd Int. Congress I A E G , Sao Paulo, Vol. I, Paper IV. 12.
Yoshikuni, H. & N a k a n a d o , H., 1974. Consolidation of soils by vertical drain wells with
finite permeability. Japanese Soc. Soil Mech. Found. Eng., Vol. 14, N o . 2.
Yoshikoni, H., 1979. Design and construction control of vertical drain m e t h o d s . Ph. D.
Thesis, Found. Eng. Series, Gihodo, Tokyo. (In Japanese)
References 511

Zeevaert, L., 1986. Consolidation in the intergranular viscosity of highly compressible


soils. In R. N . Yong & F. C. T o w n s e n d (Editors), Consolidation of soils: testing and
evaluation. A S T M Special Technical Publ. 892.
Zeng, G. X. & X i e , . H, 1989. N e w d e v e l o p m e n t of the vertical drain theories. Proc. 12th
Int. Conf. Soil M e c h . Found. Eng., Rio de Janeiro, Vol. 2, Paper 18/28, 1 4 3 5 - 1 4 3 8 .
Astedt, B., Weiner, L. & H o l m , G., 1990. Friktionsplar. Brfrmagans tillvxt m e d tid-
512 Subject index

SUBJECT INDEX

Actions, 496 Bearing capacity of piles:


Active earth pressure, 270, 279 bearing capacity factors, 186
Admixtures for soil stabilisation purpose: creep load, 175
bitumen, 456 critical depth, 187
cement, 456 empirical method based on CPT, 193
cement columns, 457 empirical methods based on SPT, 192
lime, 456 floating piles, 187, 189
lime columns, 457 Hiley's pile formula, 177
Anisotropy Case analysis, 179
strength, 34 CAPWAP analysis, 181
structural, 4 pressuremeter method, 190
Angle of intersection: stress wave analysis, 162, 179
between failure surface and ground surface, 274 Svidyn analysis, 182
between failure surface and wall, 276 underreamed piles, 187
Angular frequency, 243 Bearing capacity of pile groups:
Atterberg limits, see Consistency limits friction pile groups, 217
group efficiency factor, 218
influence of cap/pile/soil interaction, 218
Bearing capacity of footings based on strength influence of pile/soil interaction, 219
parameters: influence of pile installation, 220
bearing capacity factors, 126, 128 pile cap action, 221
depth coefficients, 129 Bedding, 54
footings next to a slope, 134 Begeman sampler, 72
inclined base of the footing, 135 Berlinoise system, 377
influence of eccentric load, 137 Blasting, compaction by, 417
influence of inclined load, 131 Bored piles, 171, 193, 199
shape coefficients, 129 Bottom heave, 382
B e a r i n g c a p a c i t y of f o o t i n g s based on Braced excavations, 376
pressuremeter results: Buckling of piles:
eccentric load, 137 initially bent pile, fully embedded, 213
eccentric, inclined load, 138 sraight pile, fully embedded, 210
centric load, 135 straight, partially embedded pile, 216
centric, inclined load, 138 Bulk density, 14
footings next to a slope, 138 Bulk modulus, 25
Bearing capacity of piers, caissons, diaphragm
and secant pile walls: Caissons:
bearing capacity factors, 236, 239 bearing capacity, 234
pressuremeter method, 234 settlement, 240
shape factor, 237 Camcometer, 82
Subject index 513

Cantilevered walls, 374 plastic limit, 48


Capillary zone, 17 shrinkage limit, 48
CAPWAP analysis, 181 Contact pressure:
Casagrande shear box, 94 approximation for rafts, 120
Casagrande's plasticity chart, 51 influence of superstructure, 124
Case analysis, 179 rigid footing on elastic medium, 107
Chemical stabilisation: raft on elastic medium, 119
cement and clay-cement grouting, 466 subgrade coefficient theory, 118
grouting with hardener, 467 Continuous sampling, 71
polymers, 469 Coring, 169
water glass, 468 CPT, see cone penetration test
Clay mineral, 4 CPTU, see piezocone test
Christalline materials, 8 Creep, 9, 32
Circular frequency, 243 Creep limit, 32
Coefficient of consolidation, 78 Creep strength, 34, 101
Coefficient of curvature, 41 Critical damping, 245
Coefficient of theoretical compressibility, 25 CRS oedometer tests, 78
Coefficient of volume compressibility, 25
Coefficient of subgrade reaction, 119, 210 Damping:
Collorimetric test, 44 critical damping, 245
Compaction: geometric damping, 263
blasting, 417 hysteretic damping, 183
compaction control, 408, 423 internal damping, 261, 263
compaction in winter time, 407 overdamping, 245
compaction piling, 422 radiation damping, 183
dry of optimum, 402 underdamping, 246
dynamic compaction equipment, 404 viscous damping, 184
frozen fill, 408 Darcy's law, 21
heavy tamping, 4 1 0 Deformation properties of soil:
oscillatory rollers, 406 correlation with sounding resistance, 90
optimum water content, 399 dilatometer test, 82
padfoot rollers, 403 direct shear test, 74
practical application, 407 empirical strength correlations, 89
rubber-tired rollers, 404 empirical stress correlations, 87
sheepsfoot rollers, 403 oedometer test, 76
static compaction equipment, 403 pressuremeter test, 80
tristar probe, 415 strain dependence, 26
vibratory plate compactors, 406 stress dependence, 28
vibratory rollers, 404 triaxial test, 75
vibroflotation, 414 time dependence, 28
vibro-wing, 415 Degree of water saturation, 12
wet of optimum, 402 Design crieria:
Compensated foundation, 106 buildings, 483
Compression modulus, see Oedometer modulus calendering machines, 487
Compression wave, 66, 86 tanks, 487
Cone penetration tests, 64 Design of vertical drain installations, 4 3 2 , 4 3 6 , 4 3 9
Consistency index, 49 Design settlement limits, 485, 488
Consistency limits: Deviation from Darcy's law, 22
fall-cone liquid limit, 48 Diameter of influence of vertical drains, 4 3 2
percussion liquid limit, 47 Diaphragm walls, 233, 378
514 Subject index

Diaphragm walls: electric resistivity of soil, 369


bearing capacity, 234 electrode consumption, 370
settlement, 240 electrode to soil resistance, 368
Dilatometer, 82 electro-osmotic permeability coefficient, 369
Dilatometer modulus, 84 current density, 464
Direct shear test, 74 Embankments:
Discharge capacity of drains, 432, 434 consolidation settlement, 323
Discontinuities, 54 critical failure surface, 315
Displacement piles, 160 embankment piles, 330
Dissolved matter, 11 enforced displacement of soft soil, 334
Drain installation techniques, 429 excavation, 334
Drained strength parameters, 32 failure surface touching firm bottom, 317
Driven piles, 160, 175, 199 horizontal displacement, 330
Dry density, 14, 400 light-weight fill, 335
Dry of optimum, 402 loading berms, 314
Dynamic compression modulus, 86 shear strength increasing with depth, 320
Dynamic consolidation, see Heavy tamping weak layers, 321
Dynamic factor, 248 Embankment pipes, see Underground pipe lines
Dynamic preloading of piles, 196 Empirical damage limits, 484
Dynamic shear modulus, 86 Equivalent drain diameter, 434
Erratic strata, 10
Earth pressure: Eurocode, 38, 495
active, 270, 279 Excavation:
at rest, 271 braced, 376
distribution, 279, 295 dewatering, 356
enforced wall rotation, 294 ejector wells, 367
graphical determination, 285 electro-osmosis, 367
horizontal ground: smooth vertical wall, 278 flow net, 364
horizontal ground: rough vertical wall, 280 gravity wells, 365
leaning wall: inclined ground surface, 283 hydraulic ground failure, 356
passive, 270, 279 hydraulic uplift, 357
Earth pressure at rest: liquefaction, 356
line load, 272 sumps, 364
point load, 273 unbraced, 374
Earthquake, liquefaction due to, 34, 494 vacuum wells, 365
Echo sounding, 67 well points, 365
Effective strength parameters, 32 Excess pore pressure, 23, 30
Effective stress, 20 Expander body, 394
Ejector wells, 367
Elastic modulus:
definition, 23 Failure surface:
determination, 74 intersection angle with ground surface, 274
Electric resistivity method, 67 intersection angle with wall, 276
Electro-chemical stabilisation: Ktter's equation, 125
electro-osmosis, 4 6 2 Fall-cone test, 96
ion diffusion, 464 False refusal, 167
Electro-osmosis: Fascines, 473
consolidation, 462 Field vane test, 99
dewatering, 367 Filter requirements for band drains, 4 4 0
electric potential gradient, 368 Flow net, 364
Subject index 515

Footings: Ion diffusion, 464


bearing capacity, see Bearing capacity of footings ISO standards, 38, 498
contact pressure, see Contact pressure
settlement, see Settlement of footings Karst formations, 16
Forced, damped vibrations, 246 Kotter's equations, 125
Foundation requirements, buildings:
design settlement limits, 485 Laminar flow, 22
empirical damage limits, 484 Laterally loaded piles:
Foundation requirements, tanks: ultimate resistance in non-cohesive soil, 202
design settlement limits, 488 ultimate resistance in cohesive soil, 206
Freezing, 471 deflection, 209
Frequency: Lime/cement columns:
angular, 243 bearing capacity, 459
circular, 243 choice of admixture, 456
natural, 244 installation techniques, 457
Frost activity, 52 settlement, 460
Frost susceptibility, 53 stabilisation of trenches/slopes, 460
Lime stabilisation, 456
Geometrical damping, 263 Limit state design, 495
Geotechnical categories, 496 Linear vibration theory, 242
Grain density, 13 Liquefaction, 22, 356
Grain size: Liquidity index, 49
distribution, 39
fractional groups, 39 Macro-structure of soil:
fractional limits, 39 anisotropy, 9
Granular columns: non-homogeneities, 10
bearing capacity, 451 Macro-structure of rock:
design, 451 appertures, 16
load share columns/surrounding soil, 451 argillaceous zones, 16
settlement, 453 discontinuities, 16
slope stability improvement, 453 fabrics, 15
Gravity walls, 298 joints, 16
Gravity wells, 365 karst formations, 16
Ground anchors, 392 shear zones, 16
Mass points:
Heavy tamping, 410 definition, 255
Helenelund's graphical method, 324 displasement, 256
Hydraulic ground failure, 356 Massive rock, 16
Hydraulic radius, 303 Mnard pressuremeter, 80
Hydraulic time lag, 73 Metal foil sampler, 72
Hydraulic uplift, 357 Micro-structure of soil:
Hydrodynamic condition, 21 aggregates, 7
Hydrodynamic time lag, see Primary consolidation bridges, 7
Hydrostatic condition, 21 clay minerals, 4
Ignition loss method, 44 crystalline, rock-forming materials, 8
Illite, 7 domains, 7
Impact, 249 organic material, 8
Inclined anchors, 378, 389, 394 Microstructure of rock, 15
Internal damping, 263 Mohr-Coulomb failure criterion, 27, 32, 278
International codes, 495 Moment reduction, sheet pile walls, 380
516 Subject index

Montmorillonite, 6 Pore water pressure, 20


Non-displacement piles, 169 Porosity, 11
Non-homogeneity, 10 Prandtl failure zone, 127, 280
Non-linear vibrations, 252 Preconsolidation pressure, 28, 88
Preconsolidation pressure:
Oedometer modulus: correlation with sounding resistance, 90
definition, 26 strength correlations, 89
determination, 76 Preloading, 424, 441
Oedometer test, 76 Pressure distribution:
Optimum water content, 399 empirical, 116
Organic material, 8 Newmark's influence diagrams, 113, 114
Oscillatory rollers, 406 probabilistic approach, 115
Overconsolidation ratio, 88 Steinbrenner's method, 111
Overdamping, 245 theory of elasticity, 109
Pressuremeter modulus, 82
Padfoot rollers, 403 Pressuremeter test, 80
Particle density, see Grain density Prestressed anchors, 385, 387
Particle size, see Grain size Primary compression index, 25, 77
Peat: Primary compression modulus, see oedometer
amorphous, 44 modulus
fibrous, 44 Primary compression ratio, 77
pseudo-fibrous, 44 Primary consolidation:
Phase shift, 247, 249 Carillo's equation, 326
Piers: Helenelund's graphical solution, 324
bearing capacity, 234 Terzaghi's consolidation theory, 30, 478
settlement, 240 Probabilistic approach, 115, 479
Piezocone tests, 65
Piezometers, 68 Quake value, 183
Piles: Quick clay, 5, 33, 354
bearing capacity, see Bearing capacity of piles Quicksand, 33
buckling, see Buckling of piles
laterally loaded, see Laterally loaded piles Raft foundation, 117
settlement, see Settlement of piles Rankine failure zone, 127, 280
Pile group design, 227 Rayleigh wave, 87
Pile loading test, 174 Reference shear strain, 27
Pile loading test, interpretation: References, 499
Brinch Hansen's 80% criterion, 175 Reinfiltration of water, 373
creep load, 175 Reinforced soil, 474
Polish method, 175 Relative density, 45
Piled raft design, 229 Reliability, 480
Piston sampler: Reliability index, 481
inside clearance ratio, 71 Rock anchors, 392
edge taper angle, 71 Rock classification:
Plasticity index, 49 bedding and foliation spacing, 57
Plate loading test, 85 compressive strength, 57
Poisson's ratio: description of mass structure, 56
definition, 25 discontinuity aperture, 58
determination, 74 discontinuity spacing, 58
Pore gas pressure, 20 identification, 55
Pore pressure measurements, 67 weathering grades, 56
Subject index 517

Rocking-sliding mode of vibrations, 252 Silo pressure:


Rubber-tired rollers, 404 eccentric discharge gate, 303
Janssen-Knen equation, 302
Skempton's pore pressure equation, 23
Safety margin, 480 Slope stability:
Sampling, 89 Bishop's rigorous method of analysis, 343
Seasonal pore pressure variations, 19 Bishop's simplified method of analysis, 343
Secant pile walls, 233 circular-cylindrical failure surface, 336
Secant pile walls: composite failure surface, 345
bearing capacity, 234 effective stress analysis, 340
settlement, 240 Janbu's correction factor, 348
Secondary compression index, 31, 330 Janbu's general procedure of slices, 347
Secondary consolidation, 31, 330 logarithmic failure surface, 344
Sedentary soils, 38 undrained analysis, 336
Sedimentary rock, 16 Soil anchor plates, 393
Sedimentary soils, 38 Soil classification:
Seepage pressure, 22 grain size distribution, 42
Seepage zone, 17 lime content, 43
Seismic refraction method: organic content, 43
longitudinal wave (P-wave), 66 Soil injection:
surface (Rayleigh) wave (R-wave), 66 bentonite, 470
shear wave (S-wave), 66 bitumen, 470
Sensitivity, 33 cement and cement-clay grouting, 466
Serviceability limit state, 495 chemical grouting with hardener, 467
Settlement of footings: injectability limits, 469
characteristic point, 142 polymers, 469
pressuremeter method, 146 water glass, 468
Steinbrenner's method, 141 Soil investigations:
theory of elasticity, 141 cone penetration test (CPT), 64
Settlement of single piles: dilatometer test, 82
empirical method, 200 direct shear test, 74, 93
pressuremeter method, 198 dynamic probing, 61
Settlement of pile groups: echo sounding, 67
equivalent raft method, 226 electric resistivity method, 67
cohesive soil, 225 fall-cone test, 96
granular soil, 223 field vane test, 99
settlement ratio, 223 geophysical methods, 65
Sheepsfoot rollers, 403 oedometer test, 76
Sheet pile walls: piezocone test (CPTU), 65
anchor plates, 393 plate loading test, 85
braced at one level, 379 pore pressure measurements, 67
braced at several levels, 386 pressuremeter test, 80
cantilevered, 374 sampling, 69
expander body anchors, 394 seismic method, 65
inclined tiebacks, 378, 389 standard penetration test (SPT), 61
moment reduction, 380 triaxial test, 75, 95
prestressed anchors, 385, 387 unconfined compression test, 95
Shear modulus: weight sounding, 63
definition, 27 Soil/structure interaction, 124, 219, 491
determination, 74 Specific density, 13
518 Subject index

Spread foundation: design settlement limits, 488


bearing capacity, 125 differential settlement of bottom plate, 490
contact stresses, 107 differential settlement of shell, 488
settlement, 139 tilt, 488
Spring constants: Temperature effects on compressive strength, 472
horizontal vibrations, 261 Terzaghi's consolidation theory, 30, 478
rocking-sliding mode of vibration, 260 Threshold gradient, 22
torsional vibrations, 261 Tie backs, 379
vertical vibrations, 258 Time factor, 78, 325
SPT, see Standard penetration test Torsional vibrations, 257
Standard penetration test, 61 Total strength parameters, 32
Static rollers, 403 Total stress, 20
Strain hardening, 73 Trench pipe lines, see Underground pipe lines
Strain softening, 73 Triaxial test, 75
Strength correlations: Tristar probe, compaction by means of, 415
penetration resistance, 102 True strength parameters, 33, 36
preconsolidation pressure, 101
Strength parameters, definitions: Ultimate limit state, 495
effective, 32 Unbraced excavations, 374
drained, 32 Undamped, free vibrations, 242
total, 32 Underground pipelines:
undrained, 33 buckling of flexible pipes, 312
Strength properties of soil, determination: embankment pipe line, 305, 307
direct shear test, 93 flexible pipes, 311
fall-cone test, 96 rigid pipes, 305
field vane test, 98 trench pipe line, 305, 307
pressuremeter test, 80 Undisturbed sampling, 69
triaxial test, 75 Undrained shear strength, correction with regard
unconfined compression test, 96 to:
Strength properties: anisotropy, 36, 101
non-cohesive soils, 46 consistency limits, 100
cohesive soils, 46 time, 35, 100
intermediary soils, 47 volume effects, 36
Stress distribution: Unified soil classification system, 52
circular flexible load (e.g. oil tanks), 112 Uniformity coefficient, 41
embankments, 110-112, 323
raft foundations, 117-124 Vacuum wells, 365
rocking mode of vibration, 252 Vadose zone, 17
spread foundations, 107-117 Vertical drains:
Stress wave analysis, 162, 179 band drains, 428
Structural discontinuities, 16 case records, 446
Subgrade coefficient theory, 118 checking the rate of consolidation, 443
Superstructure, influence of, 124 coefficient of consolidation, choice of, 443
Suspended matter, 11 design graphs, 439
Svidyn analysis, 182 diameter of influence, 432
Swedish Standard Piston Sampler, 70 discharge capacity, 432, 434, 438
drain anchor, 431
Tanks: dynamic installation, 429
damage problems, 488 drain installation techniques, 429
design criteria, 487 equivalent diameter of band drains, 434
Subject index 519

filter requirements, 440


multi-layer systems, 437
preloading, 441
sand drains, 427
sand wicks, 427
static installation, 429
theoretical approach, 432, 436
well resistance, 438
zone of smear, 438
Vibration amplitude, 243
Vibration period, 243
Vibratory plate compactors, 406
Vibratory rollers, 404
Vibroflotation, 414
Vibro-wing, compaction by, 415
Void ratio, 11

Wale beams, 392


Water content, 12
Water glass, 468
Water seepage in rock, 59
Weight sounding, 63
Well points, 365
Well resistance, 438
Wet of optimum, 402
Yield surface, 37
Young's modulus, see Elastic modulus

Zone failure at spread foundations, 126


Zone failure behind sheet pile walls:
curved failure surface, graphical determ., 287
graphical determination, 285
horizontal ground surface, 278
inclined ground surface, 283
leaning wall, 283
plane failure surface, graphical determ., 286
smooth vertical wall, 278
Zone of smear, 432

Vous aimerez peut-être aussi