Sei sulla pagina 1di 9

THE MONO-OXALATO COMPLEXES OF IRON(II1)

PART 11. KINETICS O F FORMATION

R. F. BAUERAND W. RIIAcF. SMITH


Departnzent of Chemistry, Queen's University, Kingston, Ontario
Received October 1, 1964

ABSTRACT
The kinetics of the formation of the mono-oxalato complexes of iron(II1) have been examined
spectrophotometrically over the range of temperatures 5 to 25 "C in a n aqueous medium of
ionic strength 0.50 and the range of hydrogen ion concentrations 0.03 to 0.45 M . The kinetic-
ally significant paths under the conditions studied involve reactions first order in iron(II1)
and in bioxalate but there appears to be some decrease in the second order rate constant with
increase in hydrogen ion concentration a t the highest acidities and a t the highest temperatures.
Although there is no significant contribution to the rate by a n acid-independent path first
order in free oxalate under the experimental conditions, the possibility of the rate constant for
such a path being greater than t h a t first order in bioxalate is not precluded.

INTRODUCTION
The kinetics of the anation reaction of iron (111) ion with fluoride (I), chloride ( 2 ) ,
bromide (3), thiocyanate (4), and sulfate (5) have been examined by techniques ranging
from relaxation methods to stopwatch timing and show a fairly inarlted dependence of
the second order rate constant on ligand type. Such behavior is in contrast to that exhibited
by certain bivalent metal ions such as nicltel (11) whose rate constants for corresponding
anation reactions show little dependence on ligand type (6) and are presumably controlled
by the rate of water loss from the metal ion.
This paper presents more information bearing on the influence of ligand type on the
anation reactions of iron(II1). The reaction studied, that of iron(II1) with oxalate ion
and its protonated forms, is convenient to examine because of its relative slowness in acid
solution. I t differs from the anation reaction of iron(II1) with singly charged ligands which
have been previously examined in that the principal complex formed is probably bidentate
rather than monodentate.
EXPERIMENTAL
Materials
Solutions were prepared from triply distilled water, the second distillation being from alkaline perman-
ganate. Ferric perchlorate (non-yellow), perchloric acid (double vacuum distilled 70%), sodium perchlorate
(anhydrous), and oxalic acid were all reagent grade. The sodium perchlorate was dried a t 110' for 2 h before
weighing. Stock solutions ferric perchlorate and perchloric acid were analyzed for acid and iron content by
the method previously discussed (1).
Method and Apparatus
The reaction was followed spectrophotometrically with a Beckman DIG-1 spectrophotometer a t X 310 m p
using thermostated standard 1 cm silica cells as reaction vessels. Two solutions, one containing iron(II1)
perchlorate, the other oxalic acid, and both adjusted for ionic strength and acidity, were simultaneously
injected into the reaction cell from thermostated 2 ml syringes fitted with polythene tips to ensure mixing.
The solution in the cell in the reference beam had a composition identical with the reacting solution except
t h a t it contained no oxalic acid. A t X 310 m p the uncomplexed oxalic acid and its ions d o not contribute
significantly to optical density and the difference in optical densities ( A D ) of reference and sample solutions is

where x is the total concentration of the complexes, r is the fraction of the coinplexes which is FeHC204++,
and E is the extinction coefficient in mole-' 1 cm-I of the species indicated by the subscript. If it is assumed
t h a t a t a given acidity the ratio [FeC204+]/[FeHC204++] remains a t the equilibrium value, the change of
Canadian Journal of Chemistry. Volume 43 (1965)
2764 CANADIAN JOURNAL O F CHEMISTRY. VOL. 43, 1965

optical density is proportional t o x, the concentration of complexed iron. Since as indicated in the preceding
paper the extinction coefficients of the two complexed species differ by less than 2075, AD would remain
nearly proportioned t o x even if there were moderate deviations from the equilibrium value.

I N T E R P R E T A T I O N O F T H E K I N E T I C DATA
In interpreting the experimental rate data we have used the integrated form of the
relation

where a is the total concentration of iron(II1) in all species, b is the total concentration
of oxalate, and x is the total concentration of iron(II1) complexed as FeC204+ and
FeHC204++.
kt" is a composite rate constant for the formation of the complexes and will depend on
the rate constants of the various paths. I t will involve factors for the fractions of uncom-
plexed oxalate in the protonated and unprotonated forms and will consequently depend
on the concentration of hydrogen ions. k b is a corresponding composite constant for the
reactions involving breakup or aquation of the conlplexes and like ktr' will be acid
dependent.
The relation [2] will be valid for a variety of possible mechanisms, e.g. three parallel
forward reactions opposed by back reactions if the ratio of the concentrations of the pro-
tonated and unprotonated complexes is maintained close to the equilibrium value.

kz
Fe+++ + C?OaP= FeC?04+
k-z

In this case eq. [2] takes the form

which may be written

*TBe transition state f or tluis reaction w ~ ~ of~course


l d be different from that for reaction [I].
BAUER A N D SMITH: MONO-OSALATO COMPLEXES O F IRON(II1).PART I1

where Qn2is the acid ionization quotient for HC204-,

Equation [2]would of course also apply if the protonated complex were formed solely b y
protonation of FeCZO4+if the rates of protonation and deprotonation of the complexes
were rapid compared to the rate of complex formation. I n this event k twould equal

instead of k1 + k3
Qn 2
[H'] + k2.

There is the marked possibility that the protonated and unprotonated complexes are
formed from a common intermediate. The rate expression [2]would then apply only if the
rate-controlling steps involved solely the transformation of the initial reactants to the
intermediate or if the intermediate occurred in the relatively small concentrations con-
sistent with applicability of the stationary state approximations.
If x , represents the total concentration of complexes a t equilibrium

[51
&-
- (a - xc)s(b - xe)
kt xe

Substitution of [5]into [3]with rearrangement yields

which on integration and introduction of the condition that x = 0 when t = 0 yields

log t*) + log( a6 -ab x,x ) = - k is (xe2- ab)t .


2.303~~

The second term on the left of [GI is negligible relative t o the first. a is much greater
than b under our experimental conditions and even when x approaches x , the product x,x
is rarely as great as 1% of ab.
I t follows from [ I ]that

where x , and AD, are corresponding values for equilibrium.


Substituting [7]into [GI and dropping the second term on the left of [GI yields

( ti~)
log 1 - - - -k-ts- (xe2- ab)t.
- 2.303~.
2766 CANADIAS JOURXAL O F CI-IEMISTRY. VOL. 43, 1965

The slope S1of the plot of log (1 - (ADJAD,)) against t should then be related to k
by the relation
2.303x,S1
kf =
s (xo" ab) '

x, and s can be calculated from the values of the association quotients of the mono-
oxalate conlplexes and the ionization quotients for oxalic acid given in the previous paper.

RESULTS OF T H E KINETIC MEASUREMENTS AND DISCUSSION


The data from kinetic runs for all conditions investigated yielded linear plots for
log (1 - (ADJAD,)) against time. Rate constants k p were calculated from the slopes of
these plots using eq. [9] and showed no significant dependence on total concentration of
iron(II1) or oxalate a t a given acidity over the range of concentrations investigated, i.e.
2X to 8 X lop4 for iron(III), 2 X to 12 X for oxalate. Mean values for
k p for each acidity and temperature along with the standard deviations are given in
Table I and plotted in Fig. 1against the concentration of hydrogen ions.

250 QIO 0.20 030


GiE
[H4 (mole
0 4 0 I'

FIG.1. Plots of observed forward rate constant kr against hydrogen ion concentration.

As indicated above

if there are three parallel paths, one first order in oxalate and the other two first order in
bioxalate.
Figure 1indicates a linear relationship between kp and [I-I+] (i.e. proportionality between
rate and bioxalate ion concentration) over the range of acidities investigated for the two
lowest temperatures and a significant falloff from first order dependence on hydrogen ion
concentration for the highest acidities a t the two highest temperatures. Lines of regression
through the points a t each temperature, ignoring those for concentrations of hydrogen
ion in excess of 0.2 iM a t 20 and 25 "C, have slopes implying, through eq. [4], values for
+
k1 k3 listed in Table 11. Cursory exainination of Fig. 1 suggests that k f approaches zero
a t zero hydrogen ion concentration and consequently that k: equals zero. However,
BAUER AXD SMITH: MONO-OSALATO COMPLEXES OF IRON(II1).PART I1

TABLE I
Average values for the forward rate constant
(kr) for reaction between iron(II1) and oxalate
at various temperatures and acidities. Ionic
strength 0.50

Temp. H+ kt X
("c) (mole 1 - I ) (mole-' 1 s-1)

statistical treatment of the data, implies values for kz of (19 f 20*) X lo3 a t 15 O C and
(30 f 32) X lo3 mole-l 1 s-I a t 25 "C. I t is apparent that there is the possibility that k2
is larger than kl + k3. However, under all conditions investigated k2[C204=] must be
negligible compared with (kl + k3) [HCz04-] because of the very large value of the ratio
[HC204-]/[C204=]and consequently there will be negligible contributions to the overall
rate by the path with first order dependance of rate on free oxalate ion concentration.
The plot of log (kl + k3) against 1 / T implies effective values for AH* and A F of
18 A 4 kcal molep1 and 16 f 10 cal deg-l mole-l respectively. In the subsequent argu-
ment we tentatively identify the experinlental (kl +
k3) with kl and consequently imply
that the values for AH* and AS* apply to the path with rate constant kl. I t is of interest
that AH* for other anation reactions of iron(II1) range from 13 ltcal (thiocyanate) t o
22.8 kcal (fluoride) and A F from -5 cal deg-I (thiocyanate) t o +35 cal deg-l (fluoride).
However, the points in the plot of log k against 1 / T for oxalate show considerable dis-
persion and do not rule out curvature. We do not therefore consider our values for AH*
and AS* suitable for critical comparison with those for other anation reactions. I t is possible
that an extensive group of measurements obtained with a technique such a s "stop flow"
which would reinove coinplications due to back reaction and perinit a n extension of the
*95%con$dence limit.
CANADIAN JOURNAL O F CHEMISTRY. VOL. 43. 1965

TABLE I1
Values for effective rate constant
(kl+ ks) for reactions forming
mono-complexes with first order
dependence on bioxalate ion con-
centration

Temp. +
( k ~ kt) X
("c) (mole-' 1 s-l)

temperature range might help elucidate through the shape of the log k against 1/T plot
whether (kl + k3) is indeed composite.
I t is apparent that an acceptable mechanism for the reaction between iron(II1) and
oxalate must be in accord with a second order rate law, first order in iron(II1) ion and
first order in bioxalate for all conditions except the highest acidities a t the two highest
temperatures where there is some decrease in the magnitude of the second order rate
constant with increase in acidity.
k1 (and also k3) could conceivably be composite and involve rate constants for steps in a
series reaction leading to the formation of the final complex. For example, if the protonated
complex could be ignored the reaction path might be

where FeC204+.represents a monodentate intermediate. T h a t the formation of a mono-


dentate intermediate is rate controlling in the formation of the multidentate coinplexes of
nickel(I1) is suggested by the findings of Wilkins and co-workers (7) and of Margerum and
co-workers (8).
If the stationary state approach is applicable, the following relation should apply for the
forward rate of formation of the bidentate complex.

At any one hydrogen ion concentration the rate should be first order in iron(II1) and in
bioxalate with an effective rate constant kl equal to k,jka/(kb[H+] +kd. If kb[H+] is
much less than kd then the rate constant kl equals k, (i.e. the first step is rate controlling)
and the rate constant is independent of acidity. If [H+] is such that kb[H+] becomes sig-
nificant relative to k, there will be a decrease in kl. A corresponding mechanism involving
a protonated monodentate intermediate would imply that kl should be independent of
hydrogen ion concentration. In fact, in the range of higher acidities where there appears to
be a falloff in the observed rate constant, the protonated complex is a significant species a t
equilibrium and should not be ignored. Identification of a protonated intermediate with
the final protonated complex would imply a series reaction which in general would show a
dependence of rate on concentration more involved than the second order dependence
which is actually observed. The possibility of there being a reaction path parallel to that
BAUER A N D SMITH: MONO-OXALATO COMPLEXES O F IRON(II1).PART I1 2769

summarized by eq. [lo] and involving a protonated intermediate different from the final
product cannot be definitely ruled out b u t we shall ignore this possibility and assume that
the essential reaction path is that summarized b y eq. [lo] and that the final protonated
and unprotonated species are in labile equilibrium. Such an assumption is in accord with
our data and our method of interpreting them and implies values quoted for k1 $ k3above
are those for kl.
The work of Eigen and co-workers (6,9,10) and of Wendt and Strehlow (11) has led t o
general acceptance of a mechanism of ion association which emphasizes t h a t there is a
rate-determining exchange between ligand and the water of the first coordination shell of
the metal ion within ion pairs where the partners are separated b y one water molecule.
T h e rate of water loss may be independent of or dependent on the nature of the potential
ligand of the ion pair. I n the latter case the reaction could approach in character t h a t often
designated by SN2. Ion pair formation should be a labile process and a n equilibrium con-
centration of ion pairs relative to reactant species may be expected. On this basis the
concentration of ion pairs between iron(II1) and bioxalate would be Klo[Fe+++J[HC204-]
where Klo is a n equilibrium quotient whose magnitude may be roughly estimated by the
method of Hammes and Steinfeld (12, 13). For ions of opposite charge and valences three
and one the value of the association quotient is about 14 mole-' 1if the separation of centers
is assumed to be 5 A. The values for ion pairs between iron(II1) and various singly charged
anions may be expected to vary somewhat but not markedly and the value 14 may be
considered to represent the order of magnitude of association quotients.
The rate constant k1 for complex formation should on this picture be Klokl* where kl*
is a first order rate constant. Also, if our value for kl($ k3) is indeed t h a t for formation of
the monodentate intermediate, kl($ k3)/K10 will represent the effective first order rate
constant for transformation of the ion pairs into the intermediate and a t 25 OC will be of
the order of 1440 mole-' 1 s-'/I4 mole-I 1 = 100 s-'. T h e rate constants a t 25 OC for the
anation reactions of iron (111) in aqueous solution are listed in Table 111 along with values
for the reciprocal of the ionization constants (K,) for the acids conjugate to the associating
ligand. The ionic strengths to which the rate constants apply vary from 0.1 to 1.0 but the
effect of such variation on the constants should be small. On the basis of the theory out-
lined above, the first order rate constants kl* for transformation of ion pair to complex
should be roughly parallel to the rate constants listed for all singly charged ligands. The
values listed for the ionization constants apply to zero ionic strengths b u t their reciprocals
should give fair measure of the ability of the ligand in the ion pair to accept protons from
water in the inner hydration sphere of the metal ion. I t is apparent t h a t there is a fairly
marked increase in rate constant with increase in ligand basicity and that the value which
we have determined for bioxalate is roughly in accord with there being a correlation
between rate constant and ligand basicity for singly charged anions and monodentate
complex formation.
Wendt and Strehlow (11) have suggested that the dependence of rate on ligand type is
related to the ability of the potential ligand in an ion pair to accept a proton from one of
the waters coordinated with the metal ion. I n this picture such an inner hydrolysis leads
to a n increase in the rate of loss of one of the other water molecules coordinated with the
metal ion. If it is assumed t h a t the partially desolvated metal ion is the precursor to the
complex, such an increase in the rate of water loss could lead to an increase in the first
order rate constant for complex formation from the ion pair and in the second order rate
constant with regard to the primary reactants.
Superficial assessment of the data of Table I11 suggests that they are in accord with such
2770 CANADIAN JOURNAL OF CHEMISTRY. VOL.43, 19~5

TABLE I11
Rate constants of anation reactions yielding mono-complexes of iron(II1) a t 25 "C and
estimates of the basicities of the anions involved in the reactions

Anion involved Rate constant Ref. l/k, Ref.


- -

B r- 20 3 14
C1- 9.4 2 lo-& 15
SCN- 127 4 7 16
HCa0.i- 1 440 This paper 17.5 17,18
HF 12* 1
F- 4 000* 1 1.5X103 19
Sod- 6 370 5 1O2 20
"Obtained by extrapolation.

a view. However, we doubt the direct applicability of Wendt and Strehlow's hypothesis
to the reactions of iron(II1) mainly because the rate constant for water exchange from
iron(II1) is markedly in excess of the value of the rate constant for complex formation.
The rate constant for the exchange of a water nlolecule from the first coordination sphere
of hexahydrated iron(I11) ion has been estimated by Connick and co-workers on the basis
of nuclear magnetic resonance (n.1n.r.) measurements to be in the range 2.4 X lo4 to
1.1 X lo6s-I (21, 22). The presence of a ligand in the ion pair could conceivably raise the
value for water exchange of the hydrated metal ion through inner hydrolysis in the ion
pair above this. We have implied t h a t the first order rate constant for formation of the
oxalate complex from the ion pair is approximately 100 s-l, a value substantially less than
t h a t for water exchange. If it is assumed that the rate of water exchange measures the
rate of formation of the pentahydrated iron(II1) and t h a t this is the precursor to complex
formation i t is apparent t h a t very few of the partially desolvated ion pairs complete
reaction t o form the complex and that the reaction

is essentially in equilibrium. If the equilibrium constant for this reaction is represented


by IP and the rate constant for transformation of the partially desolvated ion pair into
mono-complex (presumably with proton transfer from the bioxalate ion to the water of
the solvent) by k1 then

The associating ligand mill then be influential via its effect on K 1 and k'. Presumably
there will be partial proton transfer from water coordinated with the metal t o the associ-
ating ligand in both types of ion pairs and although such inner hydrolysis may well affect
the lability of loss and gain of water i t is difficult t o see how there could be much influence
on the equilibrium constant K1. On the other hand the rate constant k1 for the relatively
slow nucleophilic attack b y the ligand would doubtlessly be dependent on the nucleo-
philicity of the ligand. We suggest that the influence of ligand is through its nuc'eophilicity
towards the metal ion rather than its basicity with regards to proton acceptance, and that
the apparent parallelism of rate constant on the basicity is a reflection of the parallelism
between basicity and nucleophilicity.
The behavior of iron(II1) and nickel(I1) in anation reactions seems t o be in marked
contrast. For nickel the rate of reaction is close to the rate of water loss and in terms of the
above picture this implies t h a t the probability of nucleophilic attack b y ligand within the
BAUER AND SMITI-I: MONO-OXALATO COMPLEXES OF IROX(1II). P A R T I1 2771

p a r t i a l l y d e s o l v a t e d i o n p a i r i s m u c h g r e a t e r t h a n t h a t for attack b y water, w h e r e a s


f o r i r o n ( I I 1 ) t h e converse applies. As t h e r e l a t i v e nucleophilicities of w a t e r and a g i v e n
l i g a n d (e.g. o x a l a t e ) s h o u l d not be g r e a t l y different for the r e a c t i o n s i n v o l v i n g nicltel(I1)
and i r o n ( I I 1 ) p e r h a p s t h e differences i n b e h a v i o r a r e to be e x p l a i n e d i n t e r m s of a v a i l -
a b i l i t y of w a t e r for nucleophilic attack w i t h i n the i o n pairs.

REFERENCES
1. D. POULIand \\I. NIAcF. S ~ I I T H .Can. J . Chem. 38, 567 (1960).
2. R. E. CON NICK^^^ C. P. COPPELL. J. Am. Chem. Soc. 81,6389 (1959).
3. P. MATHIES and H. WENDT. Z. Physik. Chem. Frankfurt, 30, 137 (1961).
4. J . F. BELOW,JR., R. E. CONNICIC, and C. P. COPPEL. J. Am. Chem. Soc. 80,2961 (1958).
5. G G. DAVISand W. MAcF. SMITH. Can. J. Chem. 40, 1836 (1962).
6. M. EIGENand I<. T~arhr. 2. Electrochem 66, 107 (1962).
7. G. A. MELSON and R. G. ~ ~ ~ I L K I X SJ.. Chem. Soc. 4208 (1962).
8. D. W. MARGERUM, D. B. RORABACHER, and J . F. G. CLARICE,
Jr. Inorg. Chem. 2,667 (1963).
9. M. EIGEN. Proc. 7th Intern. Congr. Co-ord. Chem. Stockholm. 1962.
10. M. EIGEN.Z. Elektrochem. 64,119 (1960).
11. H. ~ V E N Dand
T H. STREHLOW.Z. Elecktrochem. 66,228 (196'2).
12. G. G. HA\r\ras and J . I. STEINFELD.J . Am. Chem. Soc. 84,4639 (1962).
13. G. H. NANCOLLAS and N. SUTIN. I~iorg.Chem. 3,360 (1964).
14. J . C. NICCOUBREY.Trans. Faraday Soc. 51,743 (1955).
15. A. NI. POSNER. Nature, 171,519 (1953).
16. T. SUZUKI and H. HAGISAWA.Bull. Inst. Phys. Chem. Res. Tokyo, 21, 601 (1942).
17. H. M. DAWSON. C. R. HOSICINS. and 1. E. SMITH. 1. Chem. Soc. 1884 (1929).
- > - -
19. H. H. BROENE and T. DEVRIES. j. An]. Chem. Soc. 69,1644 (1947).
20. C. \\I. DAVIES,
H. \Ii.JONES,and C. B. MONK. Trans. Faraday Soc. 48,921 (1952).
21. R. E. CONNICICand R. E. POULSON.1. Chem. Phvs. 30,759 (1959).
2'2. R. E. COXNICK and E. D. STOVER. JUphys.Chem. 65,2075 (1961).

Potrebbero piacerti anche