Vous êtes sur la page 1sur 9

Clay Minerals (1987) 22, 329-337

T H E R E D U C T I V E D I S S O L U T I O N OF S Y N T H E T I C
GOETHITE AND HEMATITE IN DITHIONITE
J. T O R R E N T ,

U. S C H W E R T M A N N *

AND V. B A R R O N

Departamento de Cienciasy Recursos Agricolas, Escuela T~cnica Superior de Ingenieros Agr6nomos, Apdo.
3048, 14080 C6rdoba, Spain, and *InstitutJ~r Bodenkunde, Technische Universitiit Mi2nchen,8050 FreisingVVeihenstephan, Federal Republic of Germany.
(Received 30 July 1987)

ABSTRACT: The reductive dissolution by Na-dithionite of 28 synthetic goethites and 26


hematites having widely different crystal morphologies, specific surfaces and aluminium
substitution levelshas been investigated. For both minerals the initial dissolution rate per unit of
surface area decreased with aluminium substitution. At similar aluminium substitution and
specific surface, goethites and hematites showed similar dissolution rates. These results suggest
that preferential, reductivedissolution of hematite in some natural environments, such as soilsor
sediments, might be due to the generallylower aluminium substitution of this mineral compared
to goethite.

Although thermodynamically extremely stable, goethite and hematite, the most frequent
Fe(III)-oxides in soils, may be completely redissolved under an anoxic environment if the
redox potential drops below about 0.15 V (at pH 7). This reductive dissolution is usually
caused by the anaerobic respiration of microorganisms which transfer an electron from the
metabolized organic compounds to the Fe(III)-oxides, as exemplified by the following
reaction:
FeOOH + 2H + + C H 2 0 - ~ Fe 2+ + CO2 + H20.
The dissolution of Fe(III)-oxides can be easily recognized in soils as the typical red or
yellow colours give way to the grey colours of the matrix minerals.
The ease with which the reduction takes place under fixed conditions with respect to Eh
and pH should depend on the type of Fe(III)-oxide, and its crystal properties such as crystal
size, morphology, disorder and chemistry (e.g. Al-for-Fe substitution).
Observations from red tropical soils containing hematite and goethite have led to the
conclusion that under anoxic conditions a yellowing, i.e. an apparent preferential dissolution
of hematite as opposed to goethite, occurs. Fey (1983) predicted this from thermodynamic
considerations based on an increasing stability of goethite if its AI substitution increases (see
also Yapp, 1983; Tardy & Nahon, 1985; Trolard & Tardy, 1987). However, because of the
variability of thermodynamic data for fine-grained minerals, which has resulted in a certain
arbitrariness of the selected solubility products (Trolard & Tardy, 1987), an experimental
investigation into the dissolution kinetics of fine-grained goethites and hematites appeared
desirable. This paper reports the results of a study in which the dissolution in Na-dithionite of
28 synthetic goethites and 24 hematites of widely varying crystallinity and Al-for-Fe
substitution was investigated. Dithionite was chosen because, under the conditions used, the
9 1987 The Mineralogical Society

330

J. Torrent et al.

rate of dissolution was high but still easy to follow analytically. In addition, dithionite is
widely used to extract Fe(III)-oxides from soils (Mehra & Jackson, 1960).

MATERIALS AND METHODS


Several series of synthetic goethites and hematites were used. Series 34 and 35 were A1
goethites synthesised in 0.3 M KOH at 70~ and 25~ respectively (Schulze & Schwertmann,
1984, 1987). Series 39 were pure goethites synthesised in 0.3 M KOH between 4~ and 80~
(Schwertmann et al., 1985). Goethites GV1, GV2 and GV3 were produced by adding 2 u
KOH to a solution of Fe(NO3)3 (60 mmol in 200 ml) until the final pHs were 11.5, 12 and 13,
respectively, and storing the suspensions at 35~ for 30 days. Goethite GV4 was prepared as
for GV1 but Al(NO3)3 (6 mmol) was added to the initial solution. For goethite GT2, 50 mmol
of Fe and 25 mmol of A1 (as nitrates) were dissolved in 200 ml of water, and 100 ml of 3 M
KOH were added (final OH concentration: 0.5 M); the resulting suspension was stored for 16
days at 28~ Goethite GT4 was similar to GT2 but no A1 was added.
All hematites of series A, B, C, D, E, M and MH were synthesised from Fe(NO3)3 solutions
(60 mmol in 300 ml) having different amounts of AI(NO3)B and to which 2 M KOH was
added. For series A the final pH was 8 and the suspensions were stored at 98~ for 25 days.
For series B the same procedure was followed except that oxalate was added at a
concentration of 10 -2 M. For series C different concentrations of oxalate were added: 2 x
l0 -2 for C1, 2.5 x 10-4 for C2, 5 x 10-3 for C3 and C4 and 2 x 10-2 for C5. In series D the
final pH was 9 and citrate (10 -5 M) was added. In series E the final pH was 10 and citrate
concentration was 10-4 M. In series M, maltose (10 -4 M) was added and final pH was 10 (M3)
or 9.5 (M4). Samples MH22 and MH27 were prepared as for M3 but storage temperature was
120~ Samples MH41 and MH47 were prepared by adding oxalic acid (10 -2 M) to the
solutions, precipitating to a final pH of 5, and storing the suspensions at 98~ Hematites S1,
$2 and $3 were prepared by hydrolysing Fe(C104)3 at different concentrations at 90-100~
All the hematites were washed once with acid NH4-oxalate to remove noncrystalline Feoxides, several times with 1 M (NH4)2CO3, once with water and then finally dried at 70~
Tables 1 and 2 summarize the properties of the goethites and hematites, respectively.
All samples were X-rayed using a Philips instrument with Co-Kct radiation. The crystal
thickness perpendicular to a given hkl plane (MCDhkl) was calculated using the widths at half
weight (WHH) of the most important lines and the Scherrer formula, after correcting for
instrumental broadening as described by Schulze & Schwertmann (1984) for the goethites
and by a folding procedure (H. Stanjek, unpublished) for the hematites. From MCDhk~ the
thickness perpendicular to any plane can be calculated by multiplying MCDhk~ by the cosine
of the angle between hkl and the plane. For the goethites MCD was calculated along the a and
b axes (i.e. MCD, and MCDb). For the hematites MCD a and MCDc were calculated. These
MCD values were used to obtain 'calculated' specific surfaces as follows. It was assumed that
the goethite particles were endless prisms along the e direction. In this case, if A is the mole
percent A1 substitution, it can be shown that, for a specific gravity of the pure goethite equal
to 4.37 g cm -3 and that of diaspore equal to 3.44g cm -3, the specific surface (m 2 g - l ) is:
1

where MCDa and MCDb are given in nm.

Reduction of goethite and hematite

331

TABLE 1. General characteristics of the goethites.

Goethite

Specific surface
calculated
(m2/g)
(m2/g)

MCDa MCD b measured

Feo/F%

(nmol F e m -2 min -1)

30
29
23
20
19
20
19
22

0'0
0.9
1'5
3.0
3.4
5.8
7"9
10.9

0'000
0.002
0"002
0.001
0.001
0-001
0.006
0-017

31
27
31
23
28
17
17
10

94
82
73
70
60
42
30
22
19
18

63
56
50
52
52
37
24
22
19
18

0'0
0-0
0"0
0.0
0.0
0-0
0.0
0"0
0.0
0"0

0'020
0-014
0.013
0.008
0.007
0.008
0.005
0.001
0.001
0"001

22
23
25
22
23
35
31
32
34
31

58
61
62
34

34
31
48
28

30
27
34
37

0.0
0.0
0.0
7.6

0-006
0.005
0.005
0.005

40
36
25
19

32
10

56
42

38
63

23
57

18.7
0.0

0.006
0-010

8
25

17
24
33
47

52
104
70
95

40
32
27
26

36
24
21
15

0"0
4.9
7'9
11'6

0"000
0-000
0"000
0'005

30
19
13
9

(nm)

34/0A
34/1
34/2
34/3
34/4
34/5
34/6
34/7

21
24
27
31
33
36
40
38

60
63
77
83
94
66
59
48

32
29
23
22
20
17
15
21

39'4
39'10
39r15
39'25
39'30
39'40
39'50
3~70
3980

9
10
11
11
11
16
26
29
36
38

38
43
53
43
45
54
71
80
80
80

GV1
GV2
GV3
GV4

21
24
17
20

GT2
GT4
35/0
35/3
35/4
35/5

39'60

hISS*

A1 substitution
(mol ~)

(nm)

no.

* Dissolution rate per unit of surface.


F o r hematite, if we assume that the specific gravity of the pure m i n e r a l is 5.26 g c m -3 and
t h a t of c o r u n d u m is 4.05 g c m -3 and the particles are h e x a g o n a l prisms w h o s e w i d t h is equal
to M C D a and whose height is equal to M C D c , it can be s h o w n t h a t the specific surface is:
(760/MCD~) + (380/MCDc)
SShm =

1 - 0-00228A

Specific surface was m e a s u r e d by w a t e r sorption at 2 0 ~ relative humidity. T h i s m e t h o d


gives usually surfaces about 3 0 ~ lower t h a n those m e a s u r e d with E G M E ( C a r t e r et al., 1965).
T h e oxalate-extractable F e (Feo) was d e t e r m i n e d after S c h w e r t m a n n (1964) and the
dithionite-soluble F e (F%) after M e h r a & J a c k s o n (1960) as modified by T o r r e n t & G 6 m e z M a r t i n (1985). T h e dissolved F e and A1 were d e t e r m i n e d by a t o m i c a b s o r p t i o n
s p e c t r o p h o t o m e t r y (AAS).

J. Torrent et al.

332

TABLE 2. General characteristics of the hematites.

Hematite
no.

Specific surface
measured calculated
(m2/g)
(mZ/g)

MCD a
(nm)

MCDc
(nm)

A2
A3
A4
A5

67
75
80
44

40
54
37
16

28
29
95
138

B1
B2

67
55

56
33

C1
C2
C3
C4
C5

44
69
47
68
55

D1
D2
D3

b/SS*

AI substitution
(mol %)

Feo/Fed

(nmol F e m -2 min -1)

21
17
20
42

2.0
4-5
8.5
11-5

0.007
0-016
0.051
0.310

25
15
10
12

23
96

18
26

0.0
14.3

0.002
0.025

37
6

47
35
23
48
41

37
91
101
73
98

26
22
34
20
24

0.0
7.5
14.2
10-2
13.4

0.003
0.064
0.047
0-032
0.061

29
14
8
10
7

80
76
39

67
61
17

21
20
57

15
17
44

0-0
7.5
14.0

0-008
0.012
0.052

13
15
9

E1
E2
E3

94
88
77

88
85
56

11
6
23

13
13
17

4.1
7-8
12.6

0.005
0.002
0.022

17
21
11

M3
M4

40
40

13
12

93
100

49
53

12.6
12.8

0-194
0.189

9
9

MH22
MH27
MH41
MH47

90
80
37
28

81
42
42
27

11
12
43
67

13
19
29
42

1.7
12.5
0.0
8-2

0.019
0.006
0-012
0.005

36
32
50
14

S1
$2
$3

55
55
38

44
38
23

44
36
27

23
24
37

0.0
0.0
0.0

0.001
0.001
0-001

40
29
48

* As for Table 1.

F o r investigation o f the kinetics of reductive dissolution, 60 ml polyethylene bottles were


filled with 50 ml o f a solution of s o d i u m citrate (0.25 M)/sodium b i c a r b o n a t e (0.1 M). T h i s
solution was continuously stirred with a m a g n e t i c bar while b u b b l i n g N2 at a rate of
~ 1 ml s-1 and k e p t at 25~ A f t e r N 2 had bubbled for at least 15 min, 250 m g o f solid s o d i u m
d i t h i o n i t e was added. In the m e a n t i m e , 10 m g of the sample o f goethite or h e m a t i t e were
suspended in 1 ml o f w a t e r and treated ultrasonically for 5 min. This suspension was added to
the stirred solution 5 m i n after all the solid d i t h i o n i t e had dissolved. T h e n 1 ml portions of the
final stirred suspension were t a k e n at selected times and r e d u c t i o n was i m m e d i a t e l y stopped
by a d d i n g 0.050 ml o f 30% h y d r o g e n peroxide. Finally, the dissolved F e was analysed by A A S
in the clear s u p e r n a t a n t after centrifuging the suspensions. All d e t e r m i n a t i o n s were carried
out in duplicate or triplicate and the coefficient of v a r i a t i o n was < 15%.

Reduction of goethite and hematite


RESULTS

AND

333

DISCUSSION

Goethites
For the goethites the plots of dissolved Fe against time were essentially linear for the first
30 min of reaction, when 9 ~ (for GT2) to 55~ (for 39/4) of the total Fe had dissolved.
Departure from linearity, when it occurred, was due to a small initial curvature of either
decreasing or increasing slope (Fig. 1), which is typical of first order and S-shaped curves,
respectively. Fitting straight lines to the 0-30 rain plots gave small positive or negative
intercepts ( < 2~ of total Fe). In several cases, positive intercepts corresponded to samples
having significant amounts of Feo, which would indicate some preferential, quick dissolution
of poorly-crystalline material. The slope, b, of these fitted lines is, obviously, a measure of the
initial dissolution rate which, in turn, can be taken as a measure of the 'reducibility' of the
goethites. The initial dissolution rate per unit of surface area can be obtained by dividing that
slope by the specific surfaces (SS). The resulting b/SS values (in nmol F e m -2 min-1), shown
in Table 1, are negatively correlated with A1 substitution (Fig. 2). Hence A1 appears to be a
major factor determining the dissolution rate.
For the unsubstituted goethites the initial rate of dissolution per surface area varied less if
referred to the calculated surface area than if referred to the surface area measured by H20 or
EGME adsorption. In fact, there was a negative correlation between the dissolution rate and
the difference SS~2o SScalc..For the series 39 this difference increased from zero for goethite
synthesised at 60~ to 31 m 2 g-1 for the goethite synthesised at 4~ As shown earlier by
electron microscopy (Schwertmann et al., 1985) this extra surface may be attributed to an
'internal' surface due to fissures and cracks between single domains. In contrast to the
-

5~

r~
bJ

/
/

_J
Q
or)
Or)
r-~

/,

4~

39/18

//

"~

w
b-r-

39/50

Z'

20
~

bJ
C~

/
f

10

.~J

,*

,-'~

- - -

39170

9 -"

t"

,.~"

_.

34/6

--m-

o ~/,P"
'

10

20

50

40

TIME

CMIN]

FIG. I. Dissolution-time curves for several goethites showing, in some cases, small initial
curvatures.

334

J. Torrent et al.
i

50J

"-7t..-

E
r

4~

E
LL

3g

o
y = 29-2x

E
t-

+ 0.05x 2

20

R=0.84

U3
m

12

16

2~

AI SUBSTITUTION
FIG.

2.

(MOLE

*/I,)

Relationship between dissolution rate per unit of surface area (b/SS) and Al substitution

in goethites.

dissolution by proton attack, it may be speculated that for reductive dissolution the electron
carrier, i.e. the dithionite anion, may not easily penetrate these fissures so that the main
dissolution process is more or less limited to the 'external' surface.
Hematites

Many of the hematites used in this study had significant amounts of oxalate-soluble Fe
(Table 2). In three cases (A5, M3, M4) the Feo/Fed ratio was > 0.1 (Table 2) although the
products had been washed for 2 h with acid NH4-oxalate in amounts higher than those
required for the stoichiometric removal of noncrystalline Fe-oxides. The presence of this
noncrystalline Fe can explain, in part, why the measured specific surfaces were often much
higher than the calculated ones, since noncrystalline Fe-oxides show high specific surfaces of
about 400 mZ/g (Schwertmann & Fischer, 1973; Borggaard, 1984). However, for several
samples having a Feo/F% ratio of < 0.05 (and, consequently, a specific surface attributable to
noncrystalline oxides of < 400 x 0.05 = 20 m: g-l) the differences between calculated and
measured specific surfaces were much higher than 20 m: g-1 (see e.g. A4, B2 and C3). This
fact suggests that either the morphology of the particle is markedly different from the ideal
plate assumed here, or that the particle is platy but the surface is irregular, or that the crystals
have micropores to produce a marked increase in the specific surface (Schwertmann &
K/~mpf, 1985). Irregular, grainy hematite crystals have been described specifically for
hematites prepared in oxalate (Fischer & Schwertmann, 1975 ; Schwertmann, 1987) and this,
indeed, is the case for some of the hematites of the present investigation, as for instance B2
and C3.

Reduction of goethite and hematite

335

50.
.",4

173
Lid

Jl"

41~

A5

--- .,,"
/,

,/

J
0"3
0'3

,J

30

t")

20

.,~/
H-

~ --C2

CI2
fl

T"

1.'/

.-~-

-..1,_..,r

~4

--"-

~-

,
10

MH 27

,
20

TIME

,
$0

~
4~

CMINI

FIG. 3. Dissolution-time curves for hematites A5, M3, C2 and MH-27. The first two are
curvilinear and the last two essentially rectilinear.

Most of the plots of dissolved Fe against time for the hematites were, after a few (0-5) rain
close to linearity (Fig. 3) until, at least, the first 30 min, and similar, therefore, to those of the
goethites. A few samples with high Feo/F% ratios (C3, B2, M3, M4 and A5) showed curved
dissolution curves with a decreasing slope and no sharp breaks (Fig. 3), suggesting that
noncrystalline oxides and hematite have not dissimilar dissolution rates (because in the
opposite case sharp breaks close to the Feo values would have been observed).
As for the goethites, the initial rates of dissolution per unit surface area of the hematites
(hISS) were also negatively correlated with the A1 substitution (Fig. 4). Comparison of Fig. 2
with Fig. 4 shows that the fitted quadratic curves are similar, i.e. the rate of dissolution per
unit of surface of the two minerals is similar at similar A1 substitution.
Unsubstituted hematites showed markedly different values of b/SS. None of the
mineralogical characteristics studied here were able to explain these differences. Particle
morphology might affect dissolution. This would be in line with the anisotropy shown by
hematites being etched by acids (Warren et al., 1969) although no proofs can be offered here.

CONCLUSIONS
Under the conditions of the present investigation, synthetic goethites and hematites having
similar specific surfaces and A1 substitution showed similar dissolution rates in dithionite.
Although this observation may not be valid in toto for natural goethites and hematites present
in soils or sediments it suggests, nevertheless, that the apparent preferential dissolution of
hematite in natural environments is due either to its smaller size (higher specific surface) or to

336

J. Torrent et al.

"7"( -

50,

0__

[]

E
o

LL

'%.
%

38

Y = 3 5 - 3.5 x+ 0.14 x 2

,% /

[]

,%

,%

R =0.79

,%
%.

U3
(J3

[]

,%

[]

"t ~..
[]

18-

12

16

AL SUBSTITUTION (MOLE %)
FIG. 4. Relationship between b/SS and A1 substitution in hematites.

its lower level of A1 substitution, or both. Although it has been frequently reported that
hematite has larger crystal size than goethite (Schwertmann, 1987) this is not always true
(Pefia & Torrent, 1984). As a consequence, particle size cannot be invoked as the main cause
for preferential dissolution of hematite. In contrast, in natural environments A1 substitution
is lower in hematite than in the coexisting goethite (Torrent et al., 1980; Pefia & Torrent,
1984; Schwertmann, 1985). This could be then the main reason for the preferential reductive
dissolution of hematite in soil materials or sediments which have changed from red to yellow,
and when only the pigmenting effect of the goethite remains.
ACKNOWLEDGMENT
This work was supported, in part, by the Comisi6n Asesora de Investigaci6n Cientifica y T6cnica (Spain)
under Project No. 2010/83.

REFERENCES
BORGGAARDO.K. (1984) Influence of iron oxides on the non-specific anion (chloride) adsorption by soil. J. Soil
Sei. 35, 71-78.
CARTER D E , HEILMAN M.D. & GONZALESC.L. (1965) The ethylene glycol monoethyl ether (EGME)
technique for determining soil-surface area. Soil Sei. I00, 409-413.
FEY M.V. (1983)Hypothesis for the pedogenic yellowing of red soil materials. Tech. Commun. Dept. of ARt. and
Fisheries, Republic of South Africa 18, 130-136.
FISCHER W. & SCHWERTMANNU. (1975) The formation of hematite from amorphous iron (III)-hydroxide.
Clays Clay Miner. 23, 33-37.
MEHRA O.P. & JACKSONM.L. (1960) Iron oxide removal from soils and clays by dithionite-citrate systems
buffered with sodium bicarbonate. Clays Clay Miner. 7, 317-327.

Reduction o f goethite and hematite

337

PE~A F. & TORRENTJ. (1984) Relationships between phosphate sorption and iron oxides in Alfisols from a
river terrace sequence of Mediterranean Spain. Geoderma 33, 265-282.
SCnULZED.G. & SCh'WERT~Nr~ U. (1984) The influence of aluminum on iron oxides: X. The properties of A1substituted goethites. Clay Miner. 19, 521-529.
SCI-IULZED.G. & SCHWERTMANNU, (1987) The influence of aluminum on iron oxides. XIII. Properties of
goethites synthesized in 0.3 M KOH at 25~ Clays Clay Miner. (in press).
ScnwEarraAr,~N U. (1964) Differenzierung der Eisenoxide des Bodens dutch photochemische Extraktion mit
saurer Ammonium-oxalat-L6sung. Z. Pflanzenerniihr. Bodenk. 105, 194-202.
SCnWERrMAr~N U. (1985) The effect of pedogenic environments on iron oxide minerals. Pp. 171-200 in:
Advances in Soil Science, 1, Springer-Verlag, New York.
SCl-tW~.RrMANNU. (1987) Some properties of soil and synthetic iron oxides. In: 1ton in Soils and Clay Minerals,
Nato Advanced Institute (J. W. Stucki, B. A. Goodman & U. Schwertmann, editors). Reidel, Bad
Windsheim, Germany.
SCnWERTMANNU., CAMBmRPH. & Mtr~,D E. (1985) Properties of goethites of varying crystallinity. Clays Clay
Miner. 33, 369-378.
SCh'WER'rMAr,rNU. & FiSChER W. (1973) Natural 'amorphous' ferric hydroxide. Geoderma 10, 237-247.
SCnWERTr~tANNU. & KL~u'F N. (1985) Properties of goethite and hematite in kaolinitic soils of Southern and
Central Brazil. Soil Sci. 139, 344-350.
TARDY Y. & NAHOND. (1985) Geochemistry of laterites, stability of Al-goethite, Al-hematite, and Fe 3+kaolinite in bauxites and ferricretes: an approach to the mechanism of concretion formation. Am. J. Sci.
285, 865-903.
TORRENTJ. & GOMEZ MARTIN F. (1985) Incipient podzolization processes in Humic Acrisols of Southern
Spain. J. Soil Sei. 36, 389-399.
TORRENTJ., SCHWERTMA~U. & SCHUnZED.G. (1980) Iron oxide mineralogy of some soils of two river
terrace sequences in Spain. Geoderma 23, 191-208.
TROLARDF. & TARDYY. (1987) The stabilities of gibbsite, boehmite, aluminous goethites and aluminous
hematites in bauxites, ferricretes and laterites as function of water activity, temperature and particle size.
Geoehim. Cosmochim. Acta (in press).
WARREN I.H., BATH M.D., PROSSERA.P. & ARMSTRONGJ.T. (1969) Anisotropic dissolution of hematite.
Trans. Inst. Mining. Met. C78, 21-27.
YAPP C.L (1983) Effects of A1OOH-FeOOH solid solution on goethite-hematite equilibrium. Clays Clay
Miner. 31, 239-240.

Vous aimerez peut-être aussi