Vous êtes sur la page 1sur 10

ACI STRUCTURAL JOURNAL TECHNICAL PAPER

Title no. 107-S09

Investigation of Ultimate Strength of Deck Slabs


in Steel-Concrete Bridges
by Yu Zheng, Su Taylor, Des Robinson, and David Cleland

In the last 50 years, many bridges have been built as composite steel bridge decks, the influence of varying the sized steel
structures with decks of reinforced concrete that are supported by supporting beams has not been considered.9-11 Therefore,
longitudinal steel beams. The presence of the longitudinal steel the effect of varying the size of steel supporting beams on the
beams and the unloaded area of concrete slab cause the loaded loading capacity of the slab was one of the primary areas of
deck slabs to be restrained against lateral expansion. As a result, a
focus in this research. Using an improved QUB model,7 this
compressive membrane thrust is developed. In experimental tests,
the authors built a series of one-third scale steel-concrete paper details a practical design approach to account for the
composite bridge models with several varying structural parameters, enhancing effects of arching action.
including concrete compressive strength, reinforcement percentage,
and the size of steel supporting beams. After comparing the results EXPERIMENTAL PROGRAM
of different models, the influence of these structural parameters on Test models
the amount of compressive membrane action in the deck slab was eval- The one-third scaled models were designed to represent
uated. Furthermore, the improvement of an existing theoretical model
the lowest level of external restraint in the real bridge
provided accurate predictions for the loading-carrying capacities.
(external bay) and to verify the theory under realistic loading
conditions. Another aim was to provide information on the
Keywords: beam; bridge; compressive strength; membrane.
most effective restraint system, thereby enabling the most
efficient enhancement in the load-carrying capacity.
INTRODUCTION
Based on an elastic finite element analysis result, an
Bridge deck slabs in typical beam-and-slab-type bridges
overall length of 2 m (6.7 ft) was chosen for the test models,
have inherent strength due to in-plane forces set up as a result
and the cross-sectional dimensions were typical of a
of the restraint provided by the slab panel boundary
composite steel-concrete bridge deck slab at one-third scale
conditions, including beams, diaphragms, and slab
(Fig. 1). As shown in Table 1, the variables were concrete
continuity. This is known as compressive membrane action
compressive strength, reinforcement percentage, and
(CMA) or arching action.
external restraint stiffness (by varying the minor axis I value—
Although the effect of arching action in concrete bridge Iyy value of edge beam). The typical model consisted of a
deck slabs has been recognized for some time, it is only one-way spanning concrete slab of 50 mm (1.96 in.) thickness
recently that there has been acceptance of a rational treatment of supported by two steel I-beams connected at the end by
arching action in concrete slabs. Some design and assessment channel section diaphragms. Shear studs were simulated as
codes now acknowledge the benefits of CMA. These include 25 x 25 x 50 mm (0.98 x 0.98 x 1.96 in.) steel equal angles
the Department of Regional Development (NI), Design with a spacing of 150 mm (6 in.), which provided the equivalent
Specification for Bridge Decks1; The Canadian Bridge shear area as the full- scale shear studs. The composite effect
Design Code 2 ; and the UK Highways Agency Standard between the steel beam and concrete bridge deck was
BD81/02.3 The latter came about as a direct result of achieved by the steel stud weld on the top flange of the steel
research at Queen’s University Belfast (QUB).4-7 beam. The steel beams were connected by an end diaphragm
Composite steel-concrete bridges are one of the most consisting of parallel flange channel sections of size 125 x
common types of bridge form. The presence of longitudinal 65 x 15 kg (4.92 x 2.55 x 33.1 lb) connected with four 6 mm
steel beams, together with shear stud connectors, provide (0.24 in.) diameter bolts (refer to Fig. 1).
restraint against expansion of the deck slab. As a result, The models were reinforced with bars at one-third scale. In all
compressive membrane forces are developed (refer to of the test models, the steel reinforcement was positioned at the
Fig. 1(b)), which caused an enhancement in loading capacities in middepth of the concrete slab. The 8 mm (0.32 in.) reinforcing
concrete girder-slab-type bridge decks.8 bar with a yield strength of 597 N/mm2 (86,588 psi) was used
In this paper, the effects of several structural variables on in the model with 0.5% reinforcement percentage and 10 mm
the load capacities of the deck slabs of the composite steel- (0.39 in.) reinforcing bar with a yield strength of 560 N/mm2
concrete bridges were studied by experimental tests. A (81,221 psi) in those with a 1.0% reinforcement percentage.
design method is presented for assessing the restraint stiffness that
To understand the effect on the external restraint stiffness
exists in this typical steel composite bridge deck and thereby by the steel supporting beams, two types of steel I-beam were
the strength enhancement due to CMA.

RESEARCH SIGNIFICANCE ACI Structural Journal, V. 107, No. 1, January-February 2010.


MS No. S-2008-395.R3 received May 13, 2009, and reviewed under Institute
The existence of CMA in bridge decks has been neglected publication policies. Copyright © 2010, American Concrete Institute. All rights reserved,
in flexural and shear design methods. Furthermore, in the including the making of copies unless permission is obtained from the copyright proprietors.
Pertinent discussion including author’s closure, if any, will be published in the November-
previous research into arching action in composite concrete- December 2010 ACI Structural Journal if the discussion is received by July 1, 2010.

82 ACI Structural Journal/January-February 2010


Yu Zheng is a Lecturer in structural engineering at the Dongguan University of
measured along the center line parallel to the steel support
Technology, Dongguan, China. He received his MSC from Shenzhen University, beams. Some transducers were placed directly under the slab
Shenzhen, Guangdong, China, and his PhD from Queen’s University Belfast, midspan (refer to Fig. 2). To monitor the effects of the
Northern Ireland, UK. His research interests include the durability of structures and
application of fiber-reinforced polymer. arching thrust on the steel beam, three transducers were set,
as illustrated in Fig. 2 (average of T7&T10, T8&T11, and
Su Taylor is a Senior Lecturer in structural engineering at Queen’s University Belfast. She
graduated from the University of Bath, Bath, UK, and received her PhD in structural analysis T9&T12) to measure any rotations. Concrete strains were
from Queen’s University Belfast. measured using vibrating wire strain gauges. In an attempt to
identify the in-plane or membrane strain, the arrangement of
Des Robinson is a Senior Lecturer at Queen’s University Belfast. He received his
PhD from the University of Edinburgh, Edinburgh, UK. gauges is summarized in Fig. 2. To observe the deformation
and stress in the reinforcement, epoxy-bonded electrical
David Cleland is a Professor in structural materials and Head of the School of
Planning, Architecture and Civil Engineering (SPACE) at Queen’s University Belfast.
resistance strain gauges were used to measure axial strains in
the reinforcement. This type of gauge was also used on the
outer surface of the steel beam and diaphragm to investigate
trialed in the experimental tests, namely, 305 mm x 102 mm x the strain in these structural components.
25 kg (12 in. x 4 in. x 55.1 lb) and 305 mm x 165 mm x 54 kg
(12 in. x 6.5 in. x 119.1 lb),12 respectively.
Test procedure
Test arrangement and instrumentation Each model was loaded under monotonically increasing
A 500 x 25 mm (20 x 1 in.) knife-edge load was applied at load up to the service load to ensure bedding in of the test
the midspan of the slab and parallel to the longitudinal model and remove residual deflection. After the service load
supporting beams. The loading pattern simulated two wheel test, the test model was then loaded in 5 kN (1.12 kips)
loads at a spacing of 1.5 m (4.9 ft) (assumed to be 0.5 m [1.6 ft] at increments until cracking occurred on the bottom surface at
one-third scale). It was applied via an accurately calibrated midspan. After cracking, the load increment was decreased
hydraulic actuator. Deflection of concrete slabs was to 2 kN (0.44 kips) up until the failure load.

Table 1—Nominal variables in experimental model


Ultimate deflection/
Model h, mm (in.) ρ, % fcu ,* N/mm2 (psi) ft , N/mm2 (psi) Support beam size† Iy, cm4 (in.4) Pt , kN (kips) thickness
M36SB05 50 (1.96) 0.5 35.8 (5192) 3.83 (556) 305 mm x 102 mm x 25 kg 123 (96) 58 (13) 0.4
M77SB05 50 (1.96) 0.5 77 (11,168) 4.8 (696) 305 mm x 102 mm x 25 kg 123 (96) 78 (17.5) 0.4
M38BB05 50 (1.96) 0.5 37.7 (5468) 3.51 (509) 305 mm x 165 mm x 54 kg 1063 (830) 79 (17.8) 0.2
M69BB05 50 (1.96) 0.5 68.8 (9979) 4.4 (638) 305 mm x 165 mm x 54 kg 1063 (830) 99 (22.3) 0.3
M33SB10 50 (1.96) 1 32.8 (4757) 3.77 (638) 305 mm x 102 mm x 25 kg 123 (96) 63 (14.2) 0.2
M34BB10 50 (1.96) 1 33.8 (4902) 3.43 (497) 305 mm x 165 mm x 54 kg 1063 (830) 95 (21.4) 0.2
*
Concrete strength was based on cube test.
†Based on BS4 Part 1: 1993 and BS EN10056 (1999).
Note: 1 in. = 25.4 mm; 1 lb = 0.45359 kg.

Fig. 1—Typical test model (NTS). (Note: 1 in. = 25.4 mm; 1 kip = 4.448 kN.)

ACI Structural Journal/January-February 2010 83


Fig. 2—Measurement arrangement. (Note: 1 in. = 25.4 mm; 1 kip = 4.448 kN.)

EXPERIMENTAL RESULTS
Observed behavior
Representative crack patterns at failure load for the
different loading levels are presented in Fig. 3. In all the
models—due to the line load—a longitudinal crack occurred
under the load, resulting in an obvious one-way spanning
effect. An increase in restraint stiffness or concrete compressive
strength reduced the ductility of the slab. As a result, the
punching effects became more obvious, and the failure mode
was more localized, brittle, and sudden.
In Models M77SB05, M38BB05, and M69BB05, a brittle
behavior was observed with punching the primary failure
mode. Crushing of the concrete top surface was a significant
aspect of failure for all of the experimental models. The other
Fig. 3—Crack patterns in Models M36SB05, M77SB05, and models, however, did not exhibit brittle failure modes,
M69BB05. particularly Model M36SB05, which exhibited a more
ductile behavior compared to the other slabs, as shown by its
load-deflection responses (refer to Fig. 4).

Model deflections
From the load versus deflection response at the slab
midspan (T3) in Fig. 4, it can be seen that, for the slabs with
higher external restraint stiffness or concrete compressive
strength, a delay in the stiffness degradation was evident.
This characteristic was associated with the deformation
response of a laterally restrained slab.
Figure 5 shows the deflected shape at the midsection at the
same level of applied load (50 kN [11.2 kips]) in the transverse
direction for Models M36SB05 and M38BB05. The deflections
have been magnified by 10 times with respect to the cross-
sectional dimensions. The deflected shapes clearly show the
horizontal deflections in the supporting beams, which is the
result of membrane action. Furthermore, it can be seen that
Fig. 4—Load-deflection response in validated experimental the vertical deflection in the midspan was lower as the size
test. (Note: 1 in. = 25.4 mm; 1 kip = 4.448 kN.) of the steel supporting beams increased.

84 ACI Structural Journal/January-February 2010


Table 2(a)—Comparison between experimental
results and calculated capacity using current
design methods: British standard
Pt , PfBS , PvBS,
Model no. kN (kips) kN (kips) kN (kips) Pt /PpBS
M36SB05 58 (13) 25 (5.6) 45 (10) 2.3
M77SB05 78 (18) 27 (6.1) 58 (13) 2.87
M38BB05 79 (18) 25 (5.6) 46 (10) 3.1
M69BB05 99 (22) 27 (6.1) 56 (13) 3.67
M33SB10 64 (14) 44 (10) 56 (13) 1.47
M34BB10 95 (21) 45 (10) 57 (13) 2.1
Average 2.59
Standard deviation 0.79
Coefficient of variation 0.30

Table 2(b)—Comparison between experimental


results and calculated capacity using current
design methods: American standard
Pt , PfACI , PvACI,
Fig. 5—Idealized deflected shapes at 11.2 kips (50 kN) Model no. kN (kips) kN (kips) kN (kips) Pt /PpACI
applied load (based on deflection measurements). M36SB05 58 (13) 27 (6.1) 51 (12) 2.12
M77SB05 78 (18) 29 (6.5) 75 (17) 2.65
Ultimate loads M38BB05 79 (18) 28 (6.3) 53 (12) 2.86
The ultimate loads for all of the test models are presented
M69BB05 99 (22) 29 (6.5) 71 (16) 3.39
in Table 2. As expected, the ultimate strength of the concrete
bridge deck slab with the larger steel beam and higher strength M33SB10 64 (14) 47 (11) 50 (11) 1.35
concrete had the highest ultimate capacity. This enhancement was M34BB10 95 (21) 47 (11) 50 (11) 2
partly due to increased compressive membrane action. In Models Average 2.40
M36SB05 and M38BB05, where the Iyy (second moment of area) Standard deviation 0.72
of the steel beams varied by a factor of 10 times and the cross- Coefficient of variation 0.30
sectional area varied by a factor of 2, the loading capacity was
improved by more than 40%. This implies that an increase Table 2(c)—Comparison between experimental
in the stiffness of the restraint system is an efficient way results and calculated capacity using current
to improve the loading capacity of the deck slab. Furthermore, design methods: European standard
it was found that reinforcement percentages did not influence
the ultimate strengths significantly in the comparison Pt , PfEuro , PvEuro ,
Model no. kN (kips) kN (kips) kN (kips) Pt /PpEuro
between Model M36SB05 and M33SB10.
M36SB05 58 (13) 25 (5.6) 38 (8.5) 2.29
PREDICTION METHODS OF ULTIMATE M77SB05 78 (18) 27 (6.1) 48 (11) 2.9
STRENGTHS IN BRIDGE DECKS M38BB05 79 (18) 25 (5.6) 38 (8.5) 3.1
In this study, three current standards, including the British M69BB05 99 (22) 27 (6.1) 47 (11) 3.69
Standard,13 American Codes,14,15 and the European Standard,16 M33SB10 64 (14) 42 (9.4) 48 (11) 1.54
were used to predict the ultimate strengths of the bridge M34BB10 95 (21) 44 (9.9) 49 (11) 2.17
decks. As shown in Table 2, it can be concluded that the Average 2.62
current design codes provided highly conservative Standard deviation 0.77
predictions in load capacities of a concrete bridge deck slab
Coefficient of variation 0.29
because they do not consider compressive membrane action.
The test results clearly show an increasing ultimate strength
with increasing restraint stiffness from the change in the the torsional stiffness form the supporting beam was omitted.
steel supporting beams but this was not recognized by any of Furthermore, the determination method of the effective width
the design codes. of the slab subjected to arching forces was empirical.

ANALYTICAL METHODS BASED Degree of lateral restraint


ON ARCHING THEORY The lateral restraint stiffness was assumed as shown as
Background of analytical method Fig. 6. The components that contribute to the external
To determine the ultimate capacities of bridge deck slabs, restraint stiffness were assumed as: a) horizontal bending
a proposed theoretical method based on the QUB model4,7 stiffness of the supporting beams about the minor axis (Iyy);
was used in the study. It can be known that two types of b) axial stiffness of the unloaded concrete slab and end
punching failure—flexural punching and shear punching diaphragm; and c) torsional stiffness of the supporting beams.
failure—were considered in the analysis procedure and an Horizontal bending stiffness of the supporting beam—By
accurate estimation of the effective width and lateral assuming that the horizontal arching thrust applied a
restraint stiffness was the crucial basis for accurate load capacity uniformly distributed load Par, which causes a deflection,
prediction. In the current QUB model, however, the influence of the following equation is valid

ACI Structural Journal/January-February 2010 85


Torsional stiffness of the steel supporting beams—In a
bridge deck, the concrete slab is located above the
supporting beams, and arching thrusts from the compressive
membrane effects do not act at the middepth of the
supporting beams. For concrete beams, twisting can be
neglected due to the high torsional stiffness; but for steel
beams, it is more significant due to the lower torsional
stiffness of the I-section.
Similarly to the calculation for the minor axis bending
stiffness, the applied load Par produces torsional effects,
which are distributed along the slab edge and beam, causing
a deflection δ given by the following equation

δ = Par /Ktor (4)

The torsional effect caused by the arching force in the edge


beams can be simplified as a beam subjected an eccentric
uniform load. Therefore, the torque in the section with length
Fig. 6—External lateral restraint stiffness system for bridge
dx can be expressed by
deck slab.
h
δ = Par /Kb (1) dT = q ⋅ dx ⋅ --- (5)
2

A uniformly distributed load applied to a beam produced where T is torque, q is uniform load, and h is height of the beam.
a variation in the value of deflection along the length of the The rotational angle in the unit can be given by
beam, but this can be equated to an average value of deflection by
considering the deflected shape. Integrating the expression
h
for the deflected shape and dividing by the base length gives q ⋅ dx ⋅ --- ⋅ dx
the average deflection, and by equating the average dθ = dT ⋅ dx- = -------------------------------
---------------- 2 (6)
deflection in lateral bending to the stiffness of a spring of an GJ GJ
equivalent area, the lateral restraint can be expressed by7
Therefore, the rotational angle in the edge beam can be
obtained from the integration of Eq. (6) between 0 and L (or
EI yb
K b = ζ ---------
- (2) over beff) with the boundary conditions of θ = 0, whereas x =
3
b eff 0 and x = L. Therefore,

where ζ is supporting condition constant [ζ = 114.6 (simply qh 2 qhL


θ = ----------x – ----------x (7)
supported)7 or 985 (fixed end)].7 In this analysis, the fixity 4GJ 4GJ
at the edge of beams was assumed to be simply supported for
conservatism. The effective width of the slab beff , subjected The maximum rotation angle should occur in the midspan
to arching forces, is discussed in a following section. The of the edge beams—θmax , which can be expressed as
lateral restraint stiffness due to the out-of-plane bending stiffness
of the edge supporting beam, Iyb (Eq. (2)), was a critical 2
θ max = ---------- ⎛ -----⎞ – ---------- --- = ------------
parameter. In the analysis of composite steel-concrete qh L 2 qhL L qSL
(8)
bridge decks, it is assumed that the transformed second 4GJ ⎝ 2 ⎠ 4GJ 2 8GJ
moment of the area of the support beam about the vertical
axis Iyb (Eq. (2)) could be considered as the combination of By integrating this expression to give the total area and
the edge slabs and the steel supporting beam. dividing by the base length, the average value of the deflected
Axial stiffness of the unloaded concrete slab and end shape is 0.667 θmax. Therefore, the average horizontal
diaphragms—This stiffness is influenced by the cross- displacement can be expressed by the equation as
sectional area of the concrete slab and the end diaphragm, the
span of the deck slab, and the elastic modulus of the material. 2 2
0.667qS L
The area of the unloaded slab outside the effective width acts δ avg = θ avg ⋅ S = 0.667θ max ⋅ S = ---------------------------- (9)
8GJ
in conjunction with the diaphragms in resisting the outward
arching thrust; the restraint stiffnesses are cumulative and where S is the distance from arching forces locations to the shear
can be added to give total effective restraint stiffness of
center of the edge beams. In Eq. (9), G is the shear modulus, and J
is the torsional constant according to the cross section. From
Eslab A 1 E diaphragms A 2 Eq. (4), the restraint stiffness from the torsional effects can be
⇒ K d = -----------------
- + --------------------------------- (3) given by
Le Le

12GJ
where A1 is the cross-sectional area of slab outside effective K tor = ---------------- (10)
2 3
width, and A2 is all the areas of diaphragms. S b eff

86 ACI Structural Journal/January-February 2010


Fig. 7—Distributions of transverse membrane forces
(NLFEA) in bridge deck slabs.

Fig. 9—Influence from ratio of cx to slab span: (a) length of


effective length versus ratio of cx and span; and (b) angle Φ
versus ratio of cx to slab span. (Note: 1 in. = 25.4 mm.)

“unloaded” area of slabs and diaphragms. As shown in Fig. 6, the


combined flexibility of the system was found by adding the
individual flexibilities given by the following

1- = -----
1- + -----
1- + ---------
1 - or K = -----------------------------------------------------------------------
1
----- (11)
K r K b K d K tor r
( 1 ⁄ K b ) + ( 1 ⁄ K d ) + ( 1 ⁄ K tor )

Effective width
The restraint model for the bridge deck was assumed to be
the predominately one-way spanning of the deck slab and the
assumption that the diaphragms and surrounding area of
slab, in combination with edge beams, were resisting arching
thrust in the direction parallel to the slab span direction,
which is different from the approach by Rankin for slab and
Column.17 This assumption was validated by the NLFEA
model proposed by Zheng18 (refer to Fig. 7). In the model
proposed by Taylor et al.,7 it was estimated that the influence
of the arching force was sufficiently low at a distance equal
to the effective span plus the depth of the slab (Le + h) from
the face of the support. The effective width of slab subjected
to arching forces can be given by

beff = cy + 2Le + 2h (where Le = L/2 – cx /2) (12)

where cx and cy were the transverse and longitudinal length


of the patch loads, as shown in Fig. 7.
This equation, however, was developed based on empirical
Fig. 8—Influence of cx and cy of loaded area on effective width assumptions. To verify this assumption, the results from
and Φ (refer to Fig. 7). (Notes: cx = 25 mm [1 in.] for (a) and (b); NLFEA were used to determine a more accurate effective
cy = 100 mm [3.94 in.] for (c) and (d); and 1 in. = 25.4 mm.) width. The possible influencing factors on this parameter
included concrete compressive strength, effective depth,
where L is length of the effective width of the slab subjected depth of deck and load area, and sizes and/or shape.
to arching force, which is discussed in the following. It was found that effective depth (position of reinforcement),
The combined external restraint stiffness Kr is the sum of depth of bridge deck, the concrete compressive strength, and
the contribution of the bending stiffness and torsional stiffness of loading styles did not influence the effective width of the
the edge beams and the axial stiffness of the surrounding bridge deck slab subjected to arching forces. The size of the

ACI Structural Journal/January-February 2010 87


loaded area influenced the length of the effective width, Ma = 0.168 · b · fcu · d12 · Mr(Le /Lr) (14)
however, as shown in Fig. 8. From Fig. 8(a), it can be seen
that the increase of cy increased the effective widths
approximately proportionally. That is, the value of Φ is where b is the width of the slab; d1 is half of the arching
fairly constant at 300 degrees (refer to Fig. 7 and 8(b)). In depth; and Lr is half the span of the equivalent rigidly
contrast, increasing cx reduces the effective width up to a restrained arch, which was provided by Rankin 17 with
value of cx equal to 200 mm (7.8 in.) (refer to Fig. 8(c)). idealization of the slab as a three-hinged arch.22 The arch
Thereafter, the effective width is constant. Therefore, the could be used to describe the load-deformation response of a
angle Φ increased with an increasing value of cx (refer to Fig. 8(d)) shorter finitely restrained slab as follows
but was not influenced by the value cy (refer to Fig. 8(b)).
In Fig. 9(a), it was also noted that the length of the effective
Ec A ⎞
width was constant when the ratio of cx to the slab span was L r = L e 3 ⎛ --------
-+1 (15)
⎝ KL e ⎠
above 0.4. Therefore, it is possible to establish the relationship
between Φ and ratio of cx to the slab span (<0.4) as shown in
Fig. 9(b). This relationship was based on the relative stiffness per
Based on the previous analysis, the effective width of slab unit width along the perimeter of the restrained slab. In other
subjected to arching action can be given by words, the stiffness of the bridge deck representing the arch
leg was calculated using an effective width (refer to Eq. (13)).
b eff = c y + L ( 1 – r cp ) × tan ( Φ ); Φ = 23.3 × r cp + 35.1 (13) The flexural capacity of the bridge deck is then expressed
by combination of bending capacity and arching capacity as
when rcp > 0.4; take rcp = 0.4
where rcp is the ratio of cx and the span of the bridge deck. Ppf = kd(Ma + Mb) (16)
From Table 3, it can be concluded that the proposed equation
for effective width of arching action at the loaded area gave Based on the simplified theoretical model to a strip
a better correlation with those from the NLFEA compared analogy by Taylor,23 the moment factor for the bending
with the original effective width determination. capacity was regarded as fixed at the end between beam and
concrete slabs and the arching capacity was assessed using a
Prediction approach three hinge arch, which was equated to an equivalent longer
The theoretical assumptions of the effective width and rigidly restrained arch. The simplified strip method showed
lateral restraint stiffness have been verified against NLFEA a good correlation between a more rigorous method,23 which
and provide a reasonable means of assessing the bending and analyzed the full deck slab and the experimental test results.
arching capacity of the two-way reinforced concrete slabs Therefore, Eq. (16) can be expressed as
based on the theoretical model provided by Taylor et al.4 for
one-way spanning slab strips. The bending capacity Mb can Ppf = ktaMa + Ktb Mb (17)
be derived from flexural theory using a stress block valid for
high-strength concrete. The method for estimating the
arching capacity Ma was based on the previous research on where kta = 8/L and ktb = 4/L, and L is the clear span of the
arching action in masonry walls by McDowell et al.21 Based bridge deck.
on the research by Rankin17 and the material model for high- The shear punching equation of Kirkpatrick24 for the punching
strength concrete,4 the equation for arching moment is capacity of slabs under a concentrated load was adapted for a less
provided4 as than rigid restraint system and can be described by

Table 3—Comparison of effective width of NLFEA and theoretical model23


beffNLFEA, Clear span, beff1 —original, beff2—proposed, beff NLFEA/ beff NLFEA/
Test Model in. cy , in. cx , in. in. Φ in. in. beff1 beff2
M77SB05 29 20 1 498 35.7 42 33 0.69 0.89
Author M69BB05 30 20 1 433 35.9 40 31 0.75 0.95
Taylor et al.7 D5 29 20 1 350 36.3 36 29 0.80 1.01
6 A1 17 5 5 516 41.3 25 18 0.67 0.90
Kirkpatrick et al.
A11 20 6 3 608 37.8 32 22 0.61 0.89
A12 22 8 4 608 38.9 33 24 0.65 0.90
Azad et al.19
A13 25 16 8 608 43.6 37 31 0.69 0.82
A14 29 20 8 608 43.6 41 35 0.72 0.84
20 A/23 24 3 3 828 37.3 39 26 0.60 0.92
Azad et al.
Khanna et al.8 A 62 10 20 1971 41.5 69 64 0.89 1.03
Average 0.71 0.91
Standard deviation 0.09 0.06
Coefficient of variation 0.13 0.07
Note: beff1 is effective width predicted by Eq. (12)—original effective width; beff2 is effective width predicted by Eq. (13)—proposed effective width; 1 in. = 25.4 mm.

88 ACI Structural Journal/January-February 2010


Table 4(a)—Comparison of results of theoretical models and experimental tests—authors’ tests
Model fcu , psi Beam size %As Pf1, kips Pv1, kips Pf 2, kips Pv2, kips Pt , kips Pt /Pp1 Pt /Pp2
M36SB05 5221 305 mm x 102 mm x 25 kg 0.5 14.8 18.2 13.1 19.1 13.0 0.88 1
M77SB05 11,168 305 mm x 102 mm x 25 kg 0.5 17.3 27.5 17.3 30.2 17.5 1.02 1.02
M38BB05 5511 305 mm x 165 mm x 54 kg 0.5 20.1 19.6 16.5 20.7 17.8 0.91 1.08
M69BB05 10,008 305 mm x 165 mm x 54 kg 0.5 25.9 28.0 22.7 30.6 22.3 0.86 0.98
M33SB10 4786 305 mm x 102 mm x 25 kg 1 18.1 18.4 15.1 18.9 14.4 0.79 0.95
M34BB10 4931 305 mm x 165 mm x 54 kg 1 24.8 20.0 19.9 20.4 21.4 1.07 1.07
Average 0.92 1.02
Standard deviation 0.1 0.05
Coefficient of variation 0.11 0.05
Note: Pf1, Pv1, and Pp1 are flexural punching capacity, shear punching capacity, and loading capacity predicted by original theoretical models, respectively; Pf2,Pv2,and Pp2 are flexural
punching capacity, shear punching capacity, and loading capacity predicted by proposed theoretical models with modified determination methods for effective width; 1 in. = 25.4 mm; 1 kip =
4.448 kN; 1 lbf2 = 47.88 N/m2; and 1 lb = 0.45359 kg2.

Table 4(b)—Comparison of results of theoretical models and experimental tests—Taylor et al.’s 23 tests
Beam
Model fcu , psi width, in. %As d, in. Pf1, kips Pv1, kips Pf2, kips Pv2, kips Pt , kips Pt /Pp1 Pt /Pp2
D1 15,983 5.9 0.53 1.3 43.1 48.2 42.6 51.4 41.6 0.97 0.98
D2 14,533 5.9 0.53 1.3 41.1 45.3 42.3 48.1 45.0 1.09 1.06
D5 13,663 5.9 0.53 1.0 35.8 36.9 34.6 39.2 33.7 0.94 0.97
D6 14,214 7.9 0.53 1.0 44.5 39.0 44.6 43.1 40.9 1.05 0.95
D7 14,547 3.9 0.53 1.0 24.5 33.6 26.9 36.4 30.3 1.24 1.13
Average 1.06 1.02
Standard deviation 0.12 0.08
Coefficient of variation 0.11 0.07
Note: Pf1, Pv1, and Pp1 are flexural punching capacity, shear punching capacity, and loading capacity predicted by original theoretical models, respectively; Pf2,Pv2,and Pp2 are flex-
ural punching capacity, shear punching capacity, and loading capacity predicted by proposed theoretical models with modified determination methods for effective width; 1 in. =
25.4 mm; 1 kip = 4.448 kN; 1 lbf2 = 47.88 N/m2; and 1 lb = 0.45359 kg2.

Table 5—Comparison of results of theoretical models with other theoretical models in authors’ tests
Model Ppf , kips Ppv , kips PCanada , kips Pkirk , kips Ppark , kips Pt , kips Pt /Pp Pt /PCanada Pt /Pkirk Pt /Ppark
M36SB05 13.1 19.1 11.2 22.2 28.2 13 1 1.16 0.59 0.46
M77SB05 17.3 30.2 16.9 38.2 34.7 17.5 1.02 1.04 0.46 0.51
M38BB05 16.5 20.7 11.2 23.3 33.2 17.8 1.08 1.58 0.76 0.53
M69BB05 22.7 30.6 15.7 35.7 40.3 22.3 0.98 1.41 0.62 0.55
M33SB10 15.1 18.9 13.5 21 35.8 14.4 0.95 1.07 0.69 0.4
M34BB10 19.9 20.4 13.5 21.6 50.8 21.4 1.07 1.58 0.99 0.42
Average 1.02 1.31 0.68 0.48
Standard deviation 0.05 0.25 0.18 0.06
Note: 1 kip = 4.448 kN.

0.43 0.25 f y ⎞ ⎛ M a + M b⎞ ⎛ f y ⎞
P pv = ---------- f cu × ( critical perimeter ) × d ( 100ρ e ) (18) ρ e = ( ρ a + ρ ) ⋅ ⎛ --------
- = --------------------- ⋅ --------- ρ (20)
rf ⎝ 320⎠ ⎝ M b ⎠ ⎝ 320⎠

The term rf is a reduction factor that accounts for a where ρe is the equivalent arching reinforcement percentage,
variation in the shape of the column. A value of 1.15 was ρ is the actual slab reinforcement percentage, and the critical
used for square columns compared to circular columns due perimeter is as follows: b0 = 2(cx + cy + 6d) for a rectangular
to the stress concentrations in the corners. This formula load; or b0 = 4(c + 3d) for a square load; or b0 = π(φ + 6d) for
quantifies the shear punching strength in terms of the a circular load.
“equivalent” area of reinforcement due to the combined According to the flexural and shear punching mode
effects of bending and arching. This quantified the arching predictions, the ultimate strengths of the bridge deck slabs
moment in terms of an equivalent bending resistance, that is can be obtained as

M
ρ e = ⎛ ------a-⎞ ⋅ ρ a (19) Pp = lowest of Ppf and Ppv (21)
⎝ M b⎠
Modification of original model
A correction for the variation in the yield strength was Compared to the predicted capacities from Taylor’s
necessary, however. Therefore, the total “equivalent” percentage approach,23 the proposed analysis of the authors, which
of reinforcement can be described by incorporates the torsional stiffness of supporting beams and

ACI Structural Journal/January-February 2010 89


composite bridge deck slabs and the correlation coefficients
for the composite-only test were similar to the aforementioned.

CONCLUSIONS
Based on the aforementioned analysis, the following
conclusions have been drawn:
1. Experimental results indicate that CMA are influenced by
concrete compressive strength and lateral restraint stiffness.
2. The ultimate strengths of the experimental models are
directly influenced by the concrete compressive strength and
lateral restraint stiffness. The reinforcement percentage has
a less significant effect on the capacity of the slab in the
experimental study.
Fig. 10—Correlation of predicted with test results. 3. Current design standards are highly conservative in
(Note: 1 in. = 25.4 mm.) predicting the strength of laterally restrained bridge deck slabs.
4. In the proposed lateral restraint stiffness model, the
an improved effective width prediction, showed a better influence of the torsional stiffness in the supporting beams is
collection with the results of experimental tests (refer to incorporated. It was found, however, that for typical dimension,
Table 4). It was found, however, that the proposed restraint the torsional stiffness was not significant in the influencing the
model, with the incorporation of the torsional stiffness of the overall external restraint stiffness.
edge supporting beams, did not improve the prediction of 5. From studying the results from NLFEA, a more
load capacities significantly. This is due to Ktor being much accurate determination of the effective width (arching zone)
higher in value than Kb. Therefore, Kr was more significantly was achieved. The use of an improved effective width
influenced by the value of Kb. The horizontal displacement provided improved ultimate strength predictions.
due to minor-axis bending of the supporting beams 6. The proposed method, developed from the predictions
(concrete/steel beams) was far larger than the rotational approach of Rankin17 and Taylor,23 combined with the
displacement due to the torsional stiffness of the beams. The restraint model and concept of an “equivalent” area of
modification in determining the effective width, however, arching reinforcement, gave an accurate prediction for
provided better prediction of the load capacities compared to the strength of a wide range of laterally restrained bridge
those using the previous estimation. deck slabs. The method provided consistent but slightly
conservative predictions.
Comparison of proposed method with other
methods in authors’ tests REFERENCES
A summary of five predicted failure loads is given in Table 5. 1. Department of Regional Development for Northern Ireland (formerly
These are based on Kirkpatrick et al.’s approach,6 Park’s Department of the Environment or DOE), “Design of M-Beam Bridge
Decks—Amendment No. 3 to Bridge Design Code,” N.I. Roads Service
approach,25 the prediction method provided by Canadian Headquarters, 1986, pp. 11.1-11.5.
researchers (Desai et al.),11 and the authors’ proposed 2. CAN/CSA-S6-00(R2005), “Canadian Highway Bridge Design Code,”
method. Because the approach from Kirkpatrick et al.6 and Canadian Standards Association, Mississauga, ON, Canada, 706 pp.
Park25 can just provide one punching failure mode, some 3. BD 81/02, “Use of Compressive Membrane Action in Bridge Decks,”
Design Manual for Roads and Bridges, V. 3, Section 4, Part 20, UK Highways
unsafe capacities were predicted using these two methods. Agency, Aug. 2002.
Desai et al.11 provided a rational model to predict the ultimate 4. Taylor, S. E.; Rankin, G. I. B.; and Cleland, D. J., “Arching Action in
loads of the bridge deck with consideration of compressive High Strength Concrete Slabs,” ICE Proceedings—Structures and Buildings,
membrane action and showed a good correlation at low levels of No. 146, Nov. 2001, pp. 353-362.
lateral restraint stiffness but were slightly conservative at the 5. Rankin, G. I. B., and Long, A. E., “Arching Action Strength Enhancement in
Laterally Restrained Slab Strips,” ICE Proceedings—Structures and Buildings,
higher levels of external restraint stiffness. This is due to an No. 122, Nov. 1997, pp. 461-467.
assumption of constant restraint stiffness from the supporting 6. Kirkpatrick, J.; Rankin, G. I. B.; and Long, A. E., “Strength of Evaluation of
beams. In the comparisons with other predictions, the proposed M-Beam Bridge Deck Slabs,” Structural Engineer, V. 62b, No. 3, Sept.
method provided more accurate, reliable, and safe predictions 1984, pp. 60-68.
with an average ratio of test-to-predicted failure load of 1.02. 7. Taylor, S. E.; Rankin, G. I. B.; and Cleland, D. J., “Guide to Compressive
Membrane Action in Bridge Deck Slabs,” Technical Paper 3, UK Concrete
Bridge Development Group/British Cement Association, June 2002.
Comparison with other test data 8. Khanna, O. S.; Mufit, A. A.; and Bakht, B., “Experimental Investigation of
Because the punching capacity is highly influenced by the Role of Reinforcement in Strength of Concrete Deck Slabs,” Canadian
Journal of Civil Engineering, V. 27, No. 3, 2000, pp. 475-480.
loaded area, to fully confirm the accuracy of the proposed 9. Barrington, deV., B.; Hewitt, B. E.; Casgoly, P.; and Holowka, M.,
model, a larger number of test results6,8,9,19,20,23with “An Investigation of the Ultimate Strength of Deck Slabs of Composite
different loaded areas, particularly composite bridge deck Steel/Concrete Bridges,” Transportation Research Record No. 664,
tests in both laboratory and field tests, were used for Transportation Research Board, 1978, pp. 162-170.
comparison. In total, 67 bridge deck slabs were analyzed 10. Mufti, A. A., and Newhook, J. P., “Punching Shear Strength of
Restrained Concrete Bridge Deck Slabs,” ACI Structural Journal, V. 95,
with the following results: the average value of Pt /Pp = 1.09; No. 4, July-Aug. 1998, pp. 375-381.
the sample standard deviation is 0.10; and the coefficient of 11. Desai, Y. M.; Mufti, A. A.; and Tadros, G., “Finite Element Analysis of
variation is 0.09. Steel-Free Decks,” User Manual for FEM Punch Version (2.0), July 2002, pp. 3-8.
The sample was considered to have satisfactorily covered a 12. BS EN10056, “Specification for Structural Steel Equal and Unequal
Angles: Tolerances on Shape and Dimensions,” 1993, pp. 20-56.
wide range of variables and the good correlation obtained has 13. BS 5400, “Steel, Concrete and Composite Bridges,” Parts 2 and 4,
validated the proposed method, as shown in Fig. 10. In the British Standards Institute, London, UK, 1978 and 1990, pp. 20-43.
correlation models, more than 70% of the specimens were 14. ACI Committee 318, “Building Code Requirement for Reinforced

90 ACI Structural Journal/January-February 2010


Concrete (ACI 318-05) and Commentary (318R-05),” American Concrete Kareem, K., “Loss of Punching Capacity of Bridge Deck Slabs from Crack
Institute, Farmington Hills, MI, 2005, 430 pp. Damage,” ACI Structural Journal, V. 90, No. 1, Jan.-Feb. 1993, pp. 37-41.
15. AASHTO, “Standard Specification for Highway Bridges,” third 21. McDowell, E. L.; McKee, K. E.; and Sevin, E., “Arching Action
edition, American Association of State Highway and Transport Officials, Theory of Masonry Walls,” Journal of Structural Engineering, ASCE,
Washington, DC, 1996, pp. 5.1-5.208. V. 82, No. ST2, 1956, pp. 915-1 to 915-18.
16. BS EN 1992-1-1, “Eurocode 2,” 2004., pp. 27-40. 22. Lind, N. C., and Puranik, B., “The Stability Analysis of Reticulated
17. Rankin, G. I. B., “Punching Failure and Compressive Membrane Domes with Grid Discontinuities,” Space Structures, Chapter 39, R. M.
Action in Reinforced Concrete Slabs,” PhD thesis, Queen’s University Davies, ed., Blackwell Scientific Publications, 1966, pp. 439-443.
Belfast, Northern Ireland, UK, 1982, pp. 56-106. 23. Taylor, S. E., “Compressive Membrane Action in High-Strength
Concrete Bridge Deck Slabs,” PhD thesis, Queen’s University Belfast,
18. Zheng, Y., “Modelling of Compressive Membrane Action in
Northern Ireland, UK, 2000, pp. 223-450.
Concrete Bridge Decks,” PhD thesis, Queen’s University Belfast, Northern
24. Kirkpatrick, J., “An Analytical Field and Model study of M-Beam Bridge
Ireland, UK, 2007, pp. 298-324.
Decks,” PhD thesis, Queen’s University Belfast, Northern Ireland, UK,
19. Azad, A. K.; Baluch, M. H.; Ababasi, M. S. A.; and Kareem, K., 1982, pp. 61-158.
“Punching Capacity of Deck Slabs in Girder Slab Bridges,” ACI Structural 25. Park, R., “Ultimate Strength of Rectangular Concrete Slabs under
Journal, V. 91, No. 6, Nov.-Dec. 1994, pp. 656-662. Ultimate Load with Edge Restraint,” Proceedings of the ICE, V. 28, June
20. Azad, A. K.; Baluch, M. H.; Al-Mandil, M.; Sharif, A. M.; and 1964, pp. 125-150.

ACI Structural Journal/January-February 2010 91

Vous aimerez peut-être aussi