You are on page 1of 9

Indian Journal of Chemistry

Vol. 54A, October 2015, pp. 1183-1191

Nanoparticle mediated copper(II) catalyzed oxidation of mercaptosuccinic acid


by methylene blue in aqueous acetone medium: Non-Arrhenius behaviour
Mahender Pal∗, Maya Shukla, Ranjana Sharma & K K Mishra∗
Department of Post Graduate Studies and Research in Chemistry, Rani Durgavati University,
Jabalpur 482 001, India
Email: kkmishra.chem@gmail.com (KKM)/palmahendra100@gmail.com (MP)

Received 10 March 2015; revised and accepted 14 September 2015


The oxidation of mercaptosuccinic acid by methylene blue catalyzed by Cu(II) (ca. 2.0×10-5 M) in acidic medium leads
to the formation of the corresponding disulphide and leuco dye. A half order kinetics is followed in methylene blue while
the order in the substrate varies between half and unity. The rate constant increases linearly on increasing [Cu(II)] while
remaining unaffected on varying [H+]. At lower concentrations of hydrogen ion, however, the reaction exhibits zero order
kinetics in methylene blue. The reaction shows a non-Arrhenius behaviour between 20 °C and 30 °C but subsequently the
rate constant increases at 35 °C and attains a limiting value. The rate of reaction increases on increasing the time of
incubation of the catalyst. TEM, SEM, XRD and FTIR data indicate the participation of copper nanoparticles and XRD
analysis suggests the population of (100), (111), (210) and (211) planes at 35 °C. A new active facet, (200), is created in the
face centered cubic lattice of Cu on increasing the time of incubation. The non-Arrhenius behaviour is attributed to the
temperature dependent changes in size and morphology of the nanoparticles and to a variation in the Fermi band gap energy.

Keywords: Kinetics, Nanocatalysis, Nanoparticles, Oxidation, Non-Arrhenius behaviour, Copper, Mercaptosuccinic acid, Methylene blue

Metal ions play a major role in regulation of DNA16 and in treatment of Alzheimer’s disease and
enzyme action and are vital for many biological hepatitis C17-20. Due to such wide and varied
applications1. Largely, the metal ion activity has applications, methylene blue has been the subject of
been attributed to the participation of nanoparticles2-5. several physicochemical and kinetic investigations21-25,
The nanoarchitectonics for mesoporous materials and including our investigations on metal ion catalyzed
nanowires of semiconductors have found enormous oxidation of various sulphydryl substrates by this
applications in nanotechnology especially in the oxidant26-30. In a relatively recent study, we have
generation of electronics, photonics and sensing found that the catalytic activity of Cu(II) in the
devices6,7. Recently, TiO2 thin films doped with oxidation of cysteine hydrochloride (cys) and
Au/Ag nanoclusters have been synthesized and have 2-mercaptoethylamine hydrochloride by MB in acidic
been exploited in device technology8. It is found that medium, is considerably enhanced in presence of
the size and shape of a nanocatalyst considerably redox inactive Zn(II)31. The reaction involves the
influences the course of a chemical reaction9,10. participation of nanoparticles and the morphology of
Mercaptosuccinic acid (thiomalic acid, TMA), the Cu nanoparticles changes from nanorods to
substrate used presently, has been widely employed in nanogranules on adding Zn(II) to cys-MB system.
the synthesis of various biologically active Keeping in view these novel and fascinating features
compounds such as nonleukemic spiro [indoline-3,2’- of the reaction and the biochemical and
thiazolidine]-2,4’-diones11 and in the formation of pharmaceutical significance of these reaction systems,
highly efficient non-cadmium based quantum dots the present investigation has been taken up. These
(QDs) recognized as nontoxic optical probes for studies highlight the temperature and time dependent
imaging live pancreatic cancer cells12. The oxidant changes in size and surface atomic configurations, i.e.,
methylene blue (MB) has been found to be an morphology of the participating nanoparticles leading
inhibitor of plasmodium falciparum malaria13 and has to non-Arrhenius behaviour. The motivating force for
been applied as a key compound in a wide range of this study is to correlate the catalytic activity of metal
optical oxygen sensors14,15, as a model electron ions with the size and morphology of the
receptor for the degradation of surface-confined nanoparticles involved in nanocatalysis.
1184 INDIAN J CHEM, SEC A, OCTOBER 2015

Materials and Methods washed and extracted with ether and the solid mass
The reaction of mercaptosuccinic acid (thiomalic was subsequently dissolved in methanol (Merck, GR).
acid, TMA, RSH) with methylene blue (MB) The participation of nanoparticles was ascertained by
catalyzed by Cu(II) was investigated in aqueous using scanning electron microscope Quanta 250
acetone medium (34% v/v) in presence of HCl (model D9393), Singapore and transmission electron
(ca. 2.0×10-2 M) at a fixed ionic strength (ca. 0.1201 M). microscope (Philips Technai 20). These conclusions
TMA (Evans Chemetics Inc. USA; assay 99.3%), were further verified by XRD analysis of the reaction
MB (E Merck, Germany) and CuSO4.5H2O system by employing a Goniometer PW 3050/60
(Qualigens, ExcelaR) were used as supplied. Their (theta/theta diffractometer system XPERT-PRO under
solutions were prepared by dissolving an accurately 40 mA, 45kV X-ray and minimum step size 2θ: 0.01°)
weighed quantity of TMA in acetone (Merck GR) and at different temperatures and by varying the time of
those of MB and Cu(II) in doubly distilled water incubation. Data were collected for the 2θ range of
respectively. The stock solution of CuSO4.5H2O 10-80 deg. with the step size of 0.06 deg. The
(ca. 0.1 M) was standardized with a standard solution diffraction pattern was indexed and Miller indices
of sodium thiosulphate inter alia a standard potassium (hkl) to each peak were assigned. The IR spectra were
iodate solution32. The solutions were prepared afresh recorded on a Shimadzu FTIR spectrometer (model
for each kinetic run. Dithiodimalic acid, the 8400 S) using KBr pellets.
corresponding disulphide,was prepared by oxidizing
TMA with hydrogen peroxide (Merck, 30% v/v)33 and Results and Discussion
extracting the product with ether. Dihydromethylene Kinetic studies
blue (leuco base, H2M) was prepared by reducing MB Analytical and spectrophotometric methods suggest
solution with Sn-HCl couple as described earlier34. that two mols of TMA react with one mol of MB to
give the corresponding disulphide and the leuco base.
Kinetic studies
The kinetics of the reaction were followed by The formation of disulphide as one of the end products
taking TMA solution and other ingredients of the was confirmed by the presence of characteristic
reaction system excepting MB in a conical flask and frequency at 550-510 cm-1 in the IR spectra35. The
after equilibrating the ingredients for 10 minutes, the reaction follows a half order kinetics in MB at varying
reaction was started by adding the requisite amount of concentrations of TMA (Fig. 1 and Supplementary
MB to the reaction mixture. The runs were made in Data, Table S5) as well as in TMA but the order in
reaction vessels (Pyrex, England), coated black from the substrate shows a transition from 1/2 to unity
outside to avoid photochemical interference.
The progress of the reaction was followed at
664 nm on a Thermospectronic Unicam UV
spectrophotometer 530 in terms of the depletion in
concentration of methylene blue inter alia its
absorbance (εmax 664 for MB is 6.76×104 M-1cm-1)
with the help of Beer-Lambert law for MB at a
constant temperature employing Julabo water bath
(Germany, model F–34; temperature stability ±0.03 °C).
The reaction mixture was purged with purified
nitrogen gas throughout the course of reaction.
Methylene blue and acetone do not react under the
prevailing conditions. The kinetic runs remained
homogeneous and reproducible results were obtained
for 90-95% completion of the reaction
(Supplementary Data, Table S1-S4) and thus,
the possibility of light scattering was ruled out. Fig. 1—Half order kinetics in MB. {[TMA] = 0.5×10-3M
(1); 1.0×10-3M (2); 1.5×10-3M (3); 2.0×10-3M (4); 2.5×10-3M
The reaction products were isolated by evaporating (5); 3.0×10-3M (6); [HCl] = 2.0×10-2 M; [Cu(II)] = 2.0×10-5 M;
the reaction mixture at room temperature after [MB] = 2.0×10-5 M; Acetone = 34% (v/v); [KCl] = 0.1 M;
completion of the reaction. The residue was then µ = 0.12 M; Temp. = 35 °C}.
PAL et al.: NANOPARTICLE MEDIATED Cu(II) CATALYZED OXIDATION OF MERCAPTOSUCCINIC ACID 1185

at higher concentrations of TMA (ca. >2.0×10-3 M) Table 1—Rate coefficient at different [H+]a
(see Supplementary Data, Fig. S1). These conclusions
[HCl] ×10 2 (M) k0×108 (M s-1) k1/2×106(M1/2 s-1)
were also verified by van’t Hoff differential and initial
rate measurement methods. The statistical treatment of 1.0 1.0 -
kinetic data also confirms the half order kinetic 2.0 1.1 -
behaviour in methylene blue. The rate constants were 2.5 - 2.4
reproducible within 5%. The rate increases on 3.0 - 2.3
increasing the initial concentration of MB while it 4.0 - 2.8
decreases on adding KCl and NaCl to the reaction a
[TMA] = 1.0×10 M, [Cu(II)] = 2.0×10 M, [MB] = 2.0×10-5 M,
-3 -5
system. The rate constant also decreases on increasing Acetone = 34% (v/v), [KCl] = 0.1 M, Temp. = 35 oC, µ = 0.20 M.
acetone content of the system (Supplementary Data,
Table S6). The addition of the reaction products, viz., Table 2—Rate coefficients at different time of incubationa
the disulphide and dihydromethylene blue (leuco dye, Time (min.) At T = 35 oC At T = 20 oC
H2M) to the system did not influence the rate. The k1/2×106 (M1/2 s-1)
k1/2×106 (M1/2 s-1) k1×103 (s-1)
half order rate constant increased from 1.6×10-6
5 3.5 4.6 -
to 10.1×10-6 M1/2 s-1 on increasing [Cu(II)] from 10 4.0 5.2 -
1.0×10-5 to 6.0×10-5 M, while it remains almost 20 4.5 - 1.6
unaffected on varying [H+] (Table 1). It can also be 30 4.7 - 1.9
seen from Table 1 that the order in MB shows a 30
transition from half to zero at lower concentrations of a
[TMA] = 1.0×10-3 M, [HCl] = 2.0×10-2 M, [Cu(II)] = 2.0×10-5 M,
HCl (ca. ≤ 2.0×10-3 M). Such a complex behaviour of [MB] = 2.0×10-5 M, Acetone = 34% (v/v), [KCl] = 0.1 M,
the reaction is in consonance with the earlier µ = 0.12].
observations on pH dependence of these reaction
systems36. were evaluated from the Arrhenius plots and for
the non-Arrhenius reaction, the values are found
In order to investigate the mode of interaction of to be -19.9 kJ mol-1, -418 J K-1 mol-1 and 122 kJ mol-1
Cu(II) ions with the substrate, the time of incubation of respectively. For the Arrhenius reaction, these were
the metal ion was varied in successive runs. It was calculated by employing van’t Hoff isochore and the
found that the rate constant increases on increasing the values are found to be 147.8 kJ mol-1, -315.3 J K-1 mol-1
time of incubation at 35 °C as well as at 20 °C and 108.4 kJ mol-1 respectively.
(Table 2). The rate constant decreases on increasing
The kinetic features of the reaction suggest that
the temperature from 20 °C to 30 °C showing Cu(II) coordinates with thiol acid TMA (RSH) to give
non-Arrhenius behaviour but the reaction reverts to
a labile complex C as has been frequently reported in
Arrhenius behaviour at 35 °C and the rate constant the literature35,39. The in situ formation of the complex
levels off at 40 °C. Admittedly, the temperature range C is indicated by the fact that Cu-RSH complex
employed in these studies is relatively small but it was
prepared externally when added to the system did not
chosen primarily keeping in view the kinetic catalyze the reaction. Hence,
smoothness of the runs and the reproducibility of
the results. The presence of acetone did not permit
the variations above 40 °C while erratic results
were obtained at temperatures below 20 °C.
Incidentally, a similar temperature dependence of
the rate has been noticed by Banerjee and coworkers37
in the gold nanoparticle catalyzed oxidation of … (1)
hydrazine by a metallo-superoxide complex in the where R' represents the remaining part of the
temperature range 15-35 °C. Similarly, Sheiko and molecule.
cowrkers38 have also investigated the anti-Arrhenius As shown in Scheme 1, the species C may interact
cleavage of covalent bonds in bottlebrush with the protonated methylene blue (II, MBH+;
macromolecules on aqueous substrates in a narrow equilibrium constant K)40 to give the transient species C*
temperature range (21–50 °C). The activation (III) which in turn may dissociate to produce RS radical
parameters of the reaction, viz., ∆H*, ∆S*and ∆G* and half reduced MB radical (protonated, IV)14.
1186 INDIAN J CHEM, SEC A, OCTOBER 2015

The HM· radical may react with the substrate On presuming step (2) as the rate limiting step, the
molecule to give the leucobase (V)41 and the thiyl rate expression would be given as Eq. (6).
d [MB ]
radical which will dimerise to give the disulphide.
The participation of free radicals is quite prevalent - = k ' [C][MBH + ] − k ' [C*][Cu(II )] … (6)
dt 2 −2
in these reaction systems14, 33 and in the present
case, this was qualitatively confirmed by the On applying steady state treatment for C* and C
capability of the system to initiate the polymerization and substituting [MBH+] = K[MB][H+], the rate of
of acrylonitrile42. reaction is given as Eq. (7).
H
k1' k2' k3' K[Cu(II)][RSH][MB]
N d [ MB] +k1' k−' 1k−' 2[Cu(II)]2[RSH]
C + - = … (7)
dt k2' k3' K[MB]
(CH3)2N S N(CH3)2 +k−' 1k−' 2[Cu(II)] + k−' 1k−' 3
(II, MBH) Since [Cu(II)]2 is expected to be very small, the
rate expression is given as Eq. (8).
k2' k-2'
d [ MB] k ' k 'k ' K [Cu(II)][RSH][MB]
- = 1 2 3 … (8)
dt k2' k3' K [MB]
H
+ k−1 ' k−2 '[Cu(II)] + k-1 ' k3 '
R

N Again, [Cu(II)] is comparatively smaller in the


S denominator in Eq. 8 and thus, the rate will be given
as

d [MB ] k1 ' k 2 K [Cu(II)][R SH][MB]


(CH3)2N S
N(CH 3)2 +Cu(II) '

- = … (9)
(III, C*) dt k -1 '+ k 2 ' K [MB]
… (2)
The rate expression (Eq. 9, see Supplementary
k3 ' Data SI for the complete derivation) explains a
H
C*
RS + fractional order in methylene blue and a first order
N kinetic behaviour in TMA and Cu(II) as has been
actually observed. The expression also predicts a zero
order kinetics in [H+] in consonance with the
(CH3)2N S experimental observations.
N(CH3)2
It may be mentioned here that although at lower
(IV, HM)
… (3) concentrations of Cu(II) (ca. 10-6 M), copper mainly
exists as Cu(II) as reported by Ehrenberg et al.43 in
fast
H Cu(II) catalyzed oxidation of cysteine by dioxygen,
HM + RSH however, traces of the metal ion may be reduced to
N
give Cu nanoparticles as observed by Zhang and
+ RS + H
coworkers44. It thus, appears that the kinetics of the
(CH3)2N S
reaction becomes appreciably complicated due to the
N(CH3)2
participation of nanoparticles and this obscures the
(V, Leuco base, H2M)
simple relationship between the rate constant and
… (4) [Cu(II)]. Further, Cu(II) and thiol acids interact in
different mole ratios, forming complexes such as
fast − − −
2RS RSSR
… (5) [Cu(II)R S ] + , [Cu(II)(R S) 2 ], [Cu(I)(R S )] , etc.39,
which may result in kinetic deviations, preventing us
Scheme 1 from establishing quantitatively a half order kinetic
PAL et al.: NANOPARTICLE MEDIATED Cu(II) CATALYZED OXIDATION OF MERCAPTOSUCCINIC ACID 1187

behaviour in TMA at lower concentrations. At lower aggregation under these conditions (Fig. 2 (a&b)). It
[H+], the equilibrium as shown in Eq. (1) should lie has already been mentioned that the reaction shows a
far to the right giving larger values of k1’ and k2’ and at non-Arrhenius behaviour giving a negative energy of
the same time, k-1’ is expected to decrease. Under such activation in the temperature range of 20-30 °C. The
conditions, k2’ K[MB>> k-1’ (Eq. 9) which will lead to three dimensional representations45 of the distribution
the rate expression Eq. (10). of nanoparticles at 20 °C and 35 °C with 10 minutes
d [MB]
incubation while that of the system at 35 °C with the
− = k 1 ' [Cu(II)][R SH] … (10) incubation time of 10 and 30 minutes have been given
dt in Fig. 3(a-c). The negative value of energy of
Thus, at lower [H+], the reaction should exhibit a activation indicates that the free energy barrier arises
zero order kinetics in MB which has been from a significant loss of entropy and the deviating
experimentally observed (Table 1). Eyring plots may be attributed to temperature
From Eq. (10), the zero order rate constant, dependence of either the energy of activation or the
k0 = k1’ [Cu(II)][RSH] … (11) entropy of activation46; these highlight the participation
’ of two transition states during the course of reaction as
Using the value of k0 from Table 1, k1 is found to
reported by Gypser et al.47A smaller value of the rate
be 0.5 Ms-1.
constant at 40 °C in comparison to that at 20 °C
Effect of nanoparticles highlights the prominence of non-Arrhenius behaviour
The TEM, SEM images and XRD analysis of the
reaction system indicate that Cu(II)-substrate
interaction resulting in the formation of species C, is
accompanied by the formation of nanoparticles in the
homogeneous domain. This contention seems justified
on the grounds that sulphydryl substrates are known
reducing agents and at the same time, they stabilize
the nanoparticles4. It appears that Cu(II) is, in situ,
reduced by the thiol acid to give copper nanoparticles
which in turn, may be oxidized back to Cu(II).
Incidentally, homogeneity of solution has also been
reported by El- Syed in the case of transition metal
nanocatalysis mainly due to the fact that the particles
are over an order of magnitude smaller than visible
light10.
TEM and SEM studies
The TEM images show that the size of nanoparticles
decreases on increasing the temperature and is found to
lie in the range of 119.9-198.3 nm at 30 °C, while it
ranges between 24.5 and 40.5 nm at 40 °C
(Supplementary Data, Fig. S2(a) & S2(b) respectively).
SEM images of the reaction system indicate that the
surface morphology of the nanoparticles also
appreciably depends on temperature as is evident from
the formation of nanogranules and a narrow
distribution of particles at 20 °C (Supplementary Data,
Fig. S3) and the formation of nanorods at 35 °C
(Fig. 2a). The rate increases on increasing the time of
incubation of the ingredients with the metal ion as
Fig. 2—SEM image of the reaction system at 35 °C.
already described (Table 2). The SEM images of the [(a) incubation time 10 min; (b) incubation time 30 min. [TMA]
reaction system (after completion of the reaction) show = 1.0×10-3 M; [HCl] = 2.0×10-2 M; [Cu(II)] = 2.0×10-5 M; [MB]
a change in morphology of nanoparticles leading to = 2.0×10-5 M; Acetone = 34% (v/v); [KCl] = 0.1 M; µ = 0.12 M].
1188 INDIAN J CHEM, SEC A, OCTOBER 2015

Fig. 3—Three dimensional representation of distribution of nanoparticles. [(a) at 20 °C; (b) at 35 °C with incubation time of 10 min;
(c) at 35 °C with incubation time of 10 and 30 min].

Table 3—XRD parameters of the reaction at different temperatures and time of incubation
Temp. (°C); 2θ (deg.) d-spacing (Ao) D (nm) hkl
Incubation time (min)
Expt. Calc. Expt. Calc.
35 °C; 10 min 28.6743 28.6500 3.11329 3.1101 3.28 100
40.8369 40.4009 2.20979 2.2079 14.14 111
58.8056 58.9125 1.57031 1.5737 8.37 210
66.6714 66.8580 1.40171 1.4044 6.81 211
20 °C; 10 min 28.6255 28.6012 3.11849 3.1159 3.28 100
58.7354 58.8420 1.57202 1.57463 6.09` 210
66.6886 66.8720 1.40139 1.40488 19.1 211
35 °C; 30 min 28.3037 28.2797 3.15322 3.1506 4.10 100
40.6414 40.7118 2.21996 2.2236 16.99 111
50.5517 50.6400 1.80557 1.8085 8.81 200
58.6399 58.7400 1.57435 1.57696 18.28 210
66.5000 66.6244 1.40607 1.40830 9.53 211

of the reaction. Further, a larger negative ∆S* at lower nanoparticles (D) varies between 3.28 and 14.14 nm,
temperatures suggests that adsorption is more favored as calculated by Scherrer equation48. It also shows the
resulting in a decrease in entropy under these participation of planes with Miller indices (hkl) (100),
conditions while at higher temperatures, desorption of (111), (210) and (211). On increasing the time of
MBH+ takes place37.This aspect, however, could not be incubation to 30 minutes (Fig. 4b) the size of particles
probed in detail due to presence of acetone as already increases and ranges between 4.10 and 18.28 nm
explained. without altering the d-spacing of the referred planes
XRD studies (Table 3). In addition, a new plane (200) is generated
XRD analysis of the system with 10 minutes with nanoparticles of size 8.81 nm on increasing the
incubation (Fig. 4a) reveals that the size of the time of incubation. This indicates that the larger size
PAL et al.: NANOPARTICLE MEDIATED Cu(II) CATALYZED OXIDATION OF MERCAPTOSUCCINIC ACID 1189

in size should lead to a decrease in rate of reaction on


similar grounds. In our case, however, the opposite
observation has been recorded i.e. an increase in
particle size at lower temperature leads to an increase
in rate exhibiting non-Arrhenius behaviour. In this
context, reference may be made to the size- dependent
catalytic behaviour of Pt nanoparticles in
hexacyanoferrate(III)/thiosulphate redox reaction
where kinetic study of the reaction has shown that for
a fixed mass of catalyst, the catalytic rate did not
increase proportionately with the decrease in particle
size over the entire range from 10-80 nm49. The
maximum rate of reaction has been observed for
average particle diameter of 38 nm. Particles with
diameter below 38 nm exhibit a retarding influence on
rate on decreasing the particle size. This is explained
on the grounds that with particles having a diameter
less than 38 nm, a downward shift of Fermi level with
a consequent increase of band gap energy takes place.
Consequently, the particles require more energy to
pump electrons to the adsorbed ions for the electron
transfer reaction. This results in a decrease in rate of
reaction on decreasing the particle size. On the other
hand, for nanoparticles above diameter 38 nm, the
change of Fermi level is not appreciable and due to a
decrease in surface area of the particles, the rate of
reaction decreases. It is, therefore, seen that
nanoparticles exhibit a diverse behaviour vis-a vis
their size, depending on factors such as surface-to-
volume ratio and Fermi energy level etc. This may
explain the retarding influence of nanoparticles of
smaller size obtained in the present study at higher
temperatures.
Calculation of the size of particles in the present
case was made by using Scherrer equation and the
Miller indices, for a particular angle θ, were calculated
by employing Bragg’s equation. Since the crystal of Cu
has a face centered cubic structure, the edge of the unit
cell a is equal to (4 / 2).r ) (r = 128 pm for Cu)48.
Fig. 4—XRD pattern of the reaction system. [(a) at 35 °C,
(b) at 20 °C, incubation time 10 min., and, (c) at 35 °C, incubation FTIR studies
time 30 min. [TMA] = 1.0×10-3 M, [HCl] = 2.0×10-2 M, A non-Arrhenius behaviour has been reported by
[Cu(II)] = 2.0×10-5 M, [MB] = 2.0×10-5 M, Acetone = 34% (v/v), Islam & Mukherjee50 while studying the effect of
[KCl] = 0.1 M, µ = 0.12 M].
temperature on the synthesis of silver nanoparticles in
of nanoparticles favours the reaction. This is also a copolymer micellar solution where this behaviour
verified by the XRD analysis of the system at 20 °C has been attributed to probable chemical changes in
(Fig. 4c) where the particle size is found to lie the nanoparticles on increasing the temperature. This
between 3.28 and 19.10 nm (Table 3). was also noticed by recording FTIR spectra of
It is expected that a decrease in particle size will the present reaction system at 30 °C and 40 °C
result in an increase in surface-to-volume ratio and (see Supplementary Data, Figs S4 and S5
thus, tend to increase the rate. Conversely, an increase respectively). The spectra show that the transmittance
1190 INDIAN J CHEM, SEC A, OCTOBER 2015

of CH2 (rock) increases on increasing the temperature spectra of the reaction system at 30 °C and 40 °C
and a new band appears in the fingerprint region at respectively (Figs S1-S5) as well as the complete
1545 cm-1 probably due to the coupling of C-O, derivation of the rate expression (SI) are available in
C-C or C-H stretching vibrations51. There was the electronic form at http://www.niscair.res.in/jinfo/
no appreciable change in C=O stretching vibration ijca/IJCA_54A(10)1183-1191_SupplData.pdf.
(1719 cm-1), stretching vibration Acknowledgement
(1215 cm-1) and the fundamental –CH3 asymmetric The authors are thankful to University Grants
deformation vibration (1402 cm-1). However, the Commission, India New Delhi for financial support
stretching vibration of –SH (2623 cm-1), asymmetric to work on this project (No. F 37- 522/2009 SR)
stretch of -CH3 (3090 cm-1), stretching vibrations of and to Prof. V S Rangra, Department of
≡CH (3358 cm-1) and –OH (3620 cm-1) seem to Physics, Himachal Pradesh University, Shimla,
participate in the structural and/or chemical India, for extending SEM and XRD facilities. The
modifications in the Cu–TMA-MB reaction system on authors also wish to thank Sophisticated
increasing the temperature. This indicates that the Instrumentation Centre for Applied Research and
catalysis by Cu(II) in this system is size as well as Testing (SICART), Vallabh Vidyanagar, Gujarat,
morphology dependent, which is perhaps due to the India for TEM imaging.
selective exposure of reaction site to the anisotropic
catalyst particle 9. References
1 Xiao Q, Huang S, Su W, Chan W H & Liu Y,
Conclusions Nanotechnology, 23 (2012) 495717.
Cu(II) catalyzed oxidation of mercaptosuccinic 2 Liao C, Wang H P, Chen F L, Huang C H & Fukushima Y,
J Nano, (2009) 1.
acid by methylene blue in aqueous acetone medium in
3 Datta K K R, Srinivasan B, Balaram H & Eswaramoorthy M,
presence of HCl exhibits complex kinetic features and J Chem Sci, 120 (2008) 579.
shows a non-Arrhenius behaviour between 20 and 4 Wei Y, Chen S, Kowalczyk B, Huda S, Gray T P &
30 °C. Subsequently, the rate constant increases at Grzybowski B A, J Phys Chem C, 114 (2010) 15612.
35 °C and attains a limiting value. TEM, SEM, XRD 5 Shimizu M, Takeda Y & Hiyama T, Bull Chem Soc Japan,
and FTIR data indicate the participation of copper 84 (2011) 1339.
nanoparticles in the reaction system which may be 6 Chiaramonte T, Tizei L H G, Ugarte D & Cotta M A, Nano
Letters, 111 (2011) 1934.
produced in situ by the partial reduction of Cu(II). 7 Ariga K,Vinu A, Yamauchi Y, Ji Q & Hill J P, Bull Chem
The rate increases on increasing the time of Soc Japan, 85 (2012) 1.
incubation of the metal ion with other components of 8 Gultekin A & Sonmezoglu S, Z Phys Chem, 228 (2014) 649.
the reaction system. Also, the XRD studies indicate 9 Yong L, Qiying L & Shen W, Dalton Trans, 40 (2011) 5811.
the creation of a new active facet (200) besides 10 El-Sayed M A, Acco Chem Res, 34 (2001) 257.
the population of (100), (111), (210) and (211) planes 11 Zekarias M T & Rao G N J, Chil Chem Soc, 57 (2012) 1054.
in the face centred cubic lattice of Cu. The 12 Yong K T, Ding H, Roy I, Law W C, Bergey E J , Maitra A
& Prasad PN, ACS Nano, 3 (2009) 502.
non-Arrhenius behaviour of the reaction is attributed
13 Buchholz K, Schirer R H, Eubel J K, Akoachere M B,
to the counter balancing of surface-to-volume ratio Dandekar T, Becker K & Gromer S, Antimicrob Agents
and Fermi band gap energy with regard to the Chemother, 52 (2008) 183.
temperature dependent changes in the size of 14 Impert O, Katafias A, Kita P, Mills A, Pietkiewicz –Graczyk
nanoparticles. A & Wrzeszcz G, Dalton Trans, 348 (2003).
15 Katafias A, Lipinska M & Strutynski K, React Kinet Mech
Cat, 101 (2010) 251.
Supplementary Data
16 Pan D, Zuo X, Wan Y, Wang L, Zhang J, Song S & Fan C,
Supplementary data associated with this article, Sensors, 7 (2007) 2671.
i.e., representative runs made at different 17 Tardivo J P, Gilio A D, de Oliver C S, Gabrielli D S,
concentrations of TMA and different percentage of Junqueria H C, Tada DB, Severino D, de Fatima,
acetone are described in Tables S1-S6, giving the Turchiello R & Baptista M S, Photodiagn Photodyn Ther, 2
(2005) 175.
composition of the reaction mixture, the figures
18 Tang W, Xu, H, Kopelman R & Philbert M A, Photochem
showing dependence of half order rate constant on Photobiol, 81 (2005) 242.
[TMA], TEM and SEM images of the reaction system 19 Riha P D, Riha A K, Echevarria D J & Gonzalez–Lima F,
at 30 °C, 40 °C and 20 °C respectively and FTIR Eur J Pharm, 511 (2005) 151.
PAL et al.: NANOPARTICLE MEDIATED Cu(II) CATALYZED OXIDATION OF MERCAPTOSUCCINIC ACID 1191

20 Kang X & Chen S, J Mater Sci, 45 (2010) 2696. 37 Das R S, Singh B, Mukhopadhyay S & Banerjee R,
21 Katafias A & Fenska J, Trans Met Chem, 36 (2011) 801. Dalton Trans, 41 (2012) 464132.
22 Agarwal R & Mishra K K, Z Phys Chem, 211 (1999) 91. 38 Lebedeva N V, Nese A, Sun F C, Matyjaszewski K &
23 Mishra K K & Sylvester J, J Chem Res, 10 (2006) 678. Sheiko S S, Proc Natl Acad Sci U S A, 109 (2012) 9276.
24 Mishra K K, Dwivedi U, Chaturvedi R, Pal M & Sharma R, 39 Mishra R, Banerjee R & Mukhopadhyay S, J Phy Org Chem,
Proc Nat Acad Sci India, 84 (2014) 37. 25 (2012) 1193.
25 Mishra K K, Sylvester J, Dwivedi U & Chaturvedi R, Oxid 40 Mukherjee P & Ghosh A K, J Am Chem Soc, 92 (1970)
Comm,37 (2014) 41. 6403.
26 Gupta R, Chaturvedi R & Mishra K K, Indian J Chem, 45A 41 Lehninger A L, Biochemistry, 2nd Edn, (Kalyani Publishers,
(2006) 2431. New Delhi) 1978.
27 Chaturvedi R & Mishra K K, Int J Chem Kinet, 40 (2008) 42 Seturam B, Some Aspects of Electron Transfer Reaction
145. Involving Organic Molecules, (Allied Publishers, Hydrabad)
28 Mishra K K & Chaturvedi R, Prog React Kinet Mech, 33 2003.
(2008) 253. 43 Ehrenberg M, Harms- Ringdahl M, Fedoresak I & Granath F,
29 Mishra K K, Chaturvedi R & Shukla M, Indian J Chem, 49A Acta Chem Scand, 43 (1989) 177.
(2010) 185. 44 Shen J, Chen Y, Wang Q, Yu T, Huang X, Yang Y &
30 Sharma R, Pal M & Mishra K K, Phosph Sulf Silicon Rel Zhang H, J Mater Chem C, 1 (2013) 2092.
Elements, (2015) (In press). 45 Femto Scan Online Software, (Advanced Technologies
31 Mishra K K, Gupta Ritu, Shukla Maya, Chaturvedi Ranu, Center, Moscow, Russia).
Pal M & Sharma R, Indian J Chem, 52A (2013) 724. 46 Sorgenfrei S, Chiu C, Ruben L G Jr, Yu Y, Kim P, Nuckolls C
32 Vogel A I, Quantitative Inorganic Analysis, 3rd Edn, (ELBS & Shepard K L, Nature Nanotechnol, 1 (2011).
& Longmans, London) 1968. 47 Gypser A & Norrby P O, Perkin Trans, 2 (1997) 939.
33 Patai S, The Chemistry of Thiol Group, Part II (John Wiley 48 Theivasanthi T & Alagar M, Arch Phys Res, 1 (2010) 112.
and Sons, London) 1974. 49 Sharma R K, Sharma P & Maitra A, J Colloid Interf, 265
34 Mishra K K & Chansoria K, Oxid Commun, 11 (1988) 285. (2003) 134.
35 Nakamoto K, Infrared and Raman Spectra of Inorganic and 50 Maidul Islam A K M & Mukherjee M, J Exp Nanosci,
Coordination Compounds, 4th Edn, (John Wiley and Sons, 6 (2011) 596.
New York) 1986. 51 Silverstein R M & Webster F X, Spectrometric Identification
36 Kharash N, Org Sulf Comp, Vol. I (Pergamon Press London) of Organic Compounds, 7th Edn, (John Wiley and Sons
1961. New York) 2005.

You might also like