You are on page 1of 13

Article

Progress in Reaction Kinetics


Kinetics and mechanism and Mechanism
2019, Vol. 44(3) 244–256
of formation of the complex Ó The Author(s) 2019
Article reuse guidelines:
[Ru(CN)5INH]32 through sagepub.com/journals-permissions
DOI: 10.1177/1468678319825737
the ligand substitution journals.sagepub.com/home/prk

reaction between the


aquapentacyanoruthenate(II)
anion and isoniazid

Rupal Yadav and Radhey Mohan Naik

Abstract
The formation kinetics of the complex, [Ru(CN)5INH]32, formed through the ligand substitution
reaction between isoniazid (INH) and aquapentacyanoruthenate(II) ([Ru(CN)5H2O]32), have
been investigated, under pseudo first-order conditions, as a function of concentrations of [INH]
and [Ru(CN)5H2O]32, ionic strength and temperature at pH = 4.0 6 0.02 in 0.2 M NaClO4
spectrophotometrically at 502 nm (lmax of intense yellow colour product [Ru(CN)5INH]32)
corresponding to metal-to-ligand charge-transfer transitions, in aqueous medium. The pseudo
first-order condition was maintained by taking at least 10% excess of [INH] over
[Ru(CN)5H2O]32. The stoichiometry of the reaction product was found to be 1:1 which was
further supported and characterized using elemental analysis, infrared, nuclear magnetic
resonance and mass spectrometric techniques. Thermodynamic and kinetic parameters have
also been computed, using the Eyring equation, and the values of DH6¼, Ea, DG6¼ and DS6¼ were
found to be 47.3 kJ mol21, 49.8 kJ mol21, 28.62 kJ mol21 and 187.6 J K21mol21, respectively.
The reaction was found to obey first-order kinetics with respect to [INH]. It exhibited a
negative salt effect on the rate upon variation of ionic strength of the medium. A tentative
mechanistic scheme was proposed on the basis of experimental findings.

Keywords
Ligand substitution, aquapentacyanoruthenate(II) ion, isoniazid, thermodynamic parameters, spec-
troscopic techniques

Department of Chemistry, University of Lucknow, Lucknow, India

Corresponding author:
Radhey Mohan Naik, Department of Chemistry, University of Lucknow, Lucknow, Uttar Pradesh 226007, India.
Email: radheynaik@gmail.com
Yadav and Naik 245

Introduction
There has been a considerable advancement in the past few years concerning the kinetics and
mechanism of ligand substitution reactions of catalysed and uncatalysed transition metal
complexes in aqueous medium.1–15 Previous studies revealed that the transition metals, exhi-
biting complexes like [M(CN)5L]n2 (where M(II/III) = Fe, Ru and L = H2O, CN2, imi-
des) show wide applications towards physiological as well as industrial approaches.16–23
Numerous transition metal ions, alone or as binary mixtures, such as osmium(III) oxide,
palladium(II) chloride, ruthenium(III) chloride, platinum(IV) chloride and iridium(III)
chloride, have been used extensively in redox reactions as a homogeneous catalysts, and
some systems among these have been proved quite suitable for the purpose of kinetic analy-
sis.24–26 The present developing field of transition metal–mediated organic methodologies
has incorporated several new and desirable properties, in combination with ruthenium metal,
to attain the present requirement of organic synthesis.25 Various oxidation processes of
organic and inorganic substrates, involving ruthenium metal as a catalyst,26,27 have been
reported. Several ruthenium compounds possess the ability to damage genetic materials,28–31
so they act as bacterial mutagens. Research concerning the activity of ruthenium complexes
shows that many have antitumor properties; moreover, they are found to be less toxic in
comparison with cisplatin.32,33 The bioactivity of ruthenium(II/III) complexes still requires
much attention.34
At present, isoniazid (isonicotinoylhydrazide; INH) is a first-choice medical drug, known
for the prevention and treatment of tuberculosis (TB).35 INH, a pro-drug, activated by
Mycobacterium tuberculosis catalase-peroxidase KatG, generates an isonicotinoyl acyl radi-
cal, which in subsequent steps checks the synthesis of mycolic acid which is a necessary com-
ponent in cell wall synthesis of M. tuberculosis.36 Treatment of TB currently involves several
issues such as its latency, co-infection with HIV and resistance developed against drugs due
to the emergence of multidrug-resistant tuberculosis (MDR-TB) and extensively drug-
resistant tuberculosis (XDR-TB).37 MDR-TB emerges due to resistance against most potent
first-line antitubercular drugs such as INH and rifampicin (RIF), while XDR-TB is resistant
not only to INH and RIF but also to fluoroquinolones and at least one injectable second-
line drug.
A wide variety of complexes containing nitrogen, oxygen, sulphur and phosphorous
donor atoms have been synthesized from the labile [Fe(CN)5H2O]32 ion, forming the
pentacyanoferrate(II) complex, [Fe(CN)5L]32.38,39 The ligand substitution reactions of low-
spin pentacyano(ligand)ferrate(II) complexes with various ligands, containing heterocyclic
atoms, have attracted much attention.40,41 The ligand substitution behaviour of the low-spin
[Ru(CN)5H2O]32 ion with some nitrogen-containing aromatic heterocyclic ligands, with ref-
erence to its kinetics and mechanism, was first studied by Hoddenbagh and Macartney,42
Baran and Ulger39 and recently by Naik et al.9 The low-spin aquapentacyanoruthenate(II)
ion, [Ru(CN5)H2O]32, has the capability of binding easily with one incoming or substituting
ligand, it easily loses a coordinated water molecule and subsequently forms a highly stable
nitrogen heterocyclic substituted complex.
In the recent past, we have been interested in investigating the ligand substitution kinetics
of aquapentacyanoruthenate(II) with some nitrogen heterocyclic ligands6 and some
naphthalene-substituted ligands.9 In order to strengthen our previous proposed mechanistic
scheme, we considered it worthwhile to investigate further the kinetics and mechanism of
the aquapentacyanoruthenate(II) anion with the antitubercular drug isoniazid [INH] as a
246 Progress in Reaction Kinetics and Mechanism 44(3)

H
O N
C NH2

N
NC CN
Ru
NC CN
CN

Figure 1. Structure of the product, [Ru(CN)5(INH)]32.

ligand, leading to the formation of [Ru(CN)5INH]32, as a function of pH, concentration of


[Ru(CN)5H2O]32, [INH], ionic strength and temperature at 502 nm (lmax of
[Ru(CN)5INH]32 complex). The product composition has been found to be 1:1 by the mole
ratio and slope ratio methods. The structure of the complex is shown in Figure 1.

Experimental
Materials and methods
All chemicals used were of analytical reagent grade and used as received. All solutions were
prepared by weighing them accurately in doubly distilled deionized water. K4Ru(CN)63H2O
(Alfa Aesar), INH (S D Fine Chem Ltd), NaClO4 (Aldrich) and KBr (S D Fine Chem Ltd)
were used. Using bromine from an ampoule, a 1022 M stock solution of bromine was pre-
pared and standardized regularly against a standard solution of sodium thiosulphate
(Sarabhai M Chemicals), using starch and KI solution as an indicator (as reported in the lit-
erature).43 A 1023 M stock solution of INH and a 0.2 M stock solution of NaClO4 were pre-
pared by directly weighing the compound. All the prepared stock solutions were wrapped in
aluminium foil to avoid any photodecomposition.
A solution of [Ru(CN)5H2O]32 was prepared by rapid aquation of the [Ru(CN)6]42 ion
by mixing equimolar concentrations of K4[Ru(CN)6] and bromine, in the presence of a 10-
fold excess of potassium bromide by the method described in Johnson and Shepherd.44
Standard BDH buffers were used to standardize the pH meter before use. Adjustment of the
pH up to the desired value was attained, using KCl/HCl or potassium hydrogen phthalate
and HCl/NaOH buffers. NaClO4 was used to maintain the ionic strength of the reaction
mixture.

Kinetic measurements
All kinetic runs were performed, under pseudo first-order conditions, by taking [INH] in at
least a 10-fold excess over [Ru(CN)5H2O]32 in aqueous medium at 25 °C. Solutions of the
desired concentrations were obtained by accurate dilution of their stock solutions whenever
Yadav and Naik 247

Figure 2. Absorption spectrum of the product [Ru(CN)5INH]32 under the conditions:


[Ru(CN)5H2O]32 = 1 3 1024 M, [INH] = 4 3 1023 M, ionic strength = 0.2 M (NaClO4),
temperature = 25.0 6 0.01°C and pH = 4.0 6 0.02.

required. A double-beam UV-Vis spectrophotometer (Systronics-2203) was used to record


the spectra. A single-beam visible spectrophotometer model DIGI-110 (SISCO) equipped
with a circulatory arrangement of water for thermostating the cell compartment was used to
observe the kinetics of the reaction between [Ru(CN)5H2O]32 and INH spectrophotometri-
cally. All kinetic measurements were performed at 502 nm (lmax of light yellow
[Ru(CN)5INH]32 formed). All pH measurements were made on a Systronics m pH System
(Model-361). No absorbance correction was made since at this wavelength the only strongly
absorbing species is [Ru(CN)5INH]32 and absorbances due to other reactants, namely,
[Ru(CN)5H2O]32, NaClO4 and INH, have negligible values at 502 nm (Figure 2).
In order to maintain thermal equilibrium, all reactant solutions were thermally equili-
brated at 25.0 6 0.01°C, placing them in a self-designed thermostat prior to the start of
reaction for about 30 min. All the thermally pre-equilibrated reagent solutions were mixed
(2 cm3) in the order of [Ru(CN)5H2O]32, NaClO4 and the ligand INH into a 50 mL Borosil
flask. The mixture was shaken thoroughly and quickly transferred into a 10 mm quartz cuv-
ette cell in a temperature-controlled cell compartment of the spectrophotometer.
The absorbance at time t (At) was measured at 502 nm and that at infinite time (AN) was
measured after completion of the reaction by heating the reaction mixture at 60 °C or by
keeping it overnight at room temperature. The first-order rate constants (kobs) were obtained
by least-squares fits of the data by plotting ln(At – AN) versus time t as given in equation
(1)

lnðAt  A‘ Þ = kobs t ð1Þ


248 Progress in Reaction Kinetics and Mechanism 44(3)

where At is the absorbance at time t and AN is the final absorbance.


The standard error was found to be within 62% during the determination of the kobs
values.

Results and discussion


For the preparation of the above-mentioned complex, [Ru(CN)5INH]32, to a solution of
0.1 mmol K4[Ru(CN)6]3H2O in 5 cm3 of water, 1 mmol INH was added and further,
1.7 cm3 of aqueous bromine solution (0.0117 M in Br2 and 0.0117 M in KBr) was added
slowly dropwise. Then, at room temperature, the mixture was left for 30 min and more INH
was added (1 mmol) to this solution. After 24 h, the solution was made to evaporate in a
rotary evaporator at 25 °C, until the volume reduced to 1 cm3. Upon the addition of 5 cm3
of acetone to the reaction mixture at low temperature (0 °C), a precipitate was obtained. The
synthesized solid exhibited hygroscopic properties and so it was stored under CaCl2.45 The
product was further characterized using elemental analysis along with other techniques such
as infrared (IR) spectroscopy, mass spectrometry and nuclear magnetic resonance (NMR)
spectroscopy. Finally, the formed product was identified as the [Ru(CN)5INH]32 complex
(as given in Scheme S1 in the supplemental information).
Elemental composition of the complex [Ru(CN)5INH]32. Yield: 68%–72% (0.268 g).
Anal. calcd for C11H7N8ORu: C, 35.06; H, 1.88; N, 30.16; found: C, 35.87; H, 1.92; N,
30.42%.
[Ru(CN)6]42 reacts with Br2 very rapidly and produces a pale yellow coloured solution of
the aquapentacyanoruthenate(II) ion [Ru(CN)5H2O]32, as given in equation (2)44
 4 H2 O  3
RuðCNÞ6 + Br2 ! RuðCNÞ5 H2 O + Br + BrCN ð2Þ

This pale yellow solution of the aquapentacyanoruthenate(II) ion, formed during the reac-
tion, showed lmax at 310 nm. The [Ru(CN)5H2O]32 generated above reacts with INH and
forms the [Ru(CN)5INH]32 complex which follows the Beer–Lambert law at 502 nm (molar
extinction coefficient (e) at 1640 6 50 M21 cm21) over a wide range of concentrations. The
[Ru(CN)5H2O]32 ion undergoes a slow dimerization reaction at higher concentrations
(.1024 M) (Kd ’ 1022 M21 s21 at pH = 7, 25.0 °C) to yield most probably a cyanide-
bridged Ru2(CN)1062 ion,46 as reported in the case of the pentacyanoiron(II) system.47
The dependence of reaction rate on various reaction variables is discussed below.

The IR spectrum
The IR spectral study can be achieved in vapour, liquid, solution and in the solid phase.
Since water shows absorption near 3710 and 1630 cm21, the sample used should be com-
pletely dried. To record the spectrum of a solid sample, 1–3 mg of compound and 100–
200 mg of alkali halide (KBr or KCl) were mixed together in powder form, dried further to
remove moisture and pressed under pressure to produce a small transparent disc (pellet),
which produces a clean spectrum. The IR spectrum of the product (KBr disc) showed strong
bands at 3302.13, 3111.18, 2054.19 and 1668.43 cm21 together with other prominent bands
at 1635.64, 1556.55, 844.82 cm21 , etc., (see supplemental material in Figure S1 and Table
1). The presence of strong absorption bands at 3049.46 and 3014.74 cm21 for asymmetric
and symmetric stretching vibrations, respectively, shows the presence of a free amino (NH2)
Yadav and Naik 249

Table 1. IR absorption bands (cm21) and assignments for the [Ru(CN)5INH]32 complex.

Vibrational wavenumbers (cm21) Vibrational assignments

3302.13 m n(NH) secondary amine stretching


3111.18 m n(=C–H) stretching
3049.46 m n(NH2) asymmetric stretching
3014.74 m n(NH2) symmetric stretching
2054.19 s n(CN) stretching
1668.43 vs n(CO) stretching
1635.64 m d(NH2) bending
1556.55 s n(C=C) stretching
844.82 s v(NH2)
675.09 s d(=C–H) out-of-plane bending
530.00 m n(Ru–N) stretching

IR: infrared.

group. The very sharp band present at 1668.43 cm21 indicates the presence of a carbonyl
group in the complex. The small band observed at 530 cm21 corresponds to the formation
of a Ru–N bond in the product.48 Therefore, the spectrum suggests that ruthenium(II) is
likely to be coordinated through the nitrogen atom of the pyridine ring due to the strong
coordinating ability of the nitrogen atom.

NMR spectrum
The NMR spectrum of the complex, (see Figure S2 in the supplemental material), gave
chemical shifts at 8.62 and 7.64 ppm, respectively. The first value is attributed to the pro-
tons, attached directly to the electronegative nitrogen atoms. This high value is due to the
deshielding effect of the electronegative nitrogen atom of the pyridine ring as compared to
the second value, that is, 7.64 ppm, for subsequent protons attached at vicinal carbon atoms.
Since D2O was the solvent used, which leads to the rapid exchange of NH and NH2 protons
with the deuterium atoms, this suppresses the peaks of NH and NH2 protons in the spec-
trum. All values are tabulated in Table S1 in the supplemental material.

Mass spectrum
The electrospray ionization-mass spectrum of the formed complex is shown in Figure S3 in
the supplemental material. The major peak, [M + H]+, recorded at m/z = 365.00, confirms
the stoichiometry of the complex formed. Some deviations, observed in the first-order beha-
viour of reactions of the [Ru(CN)5L]n2 complex, were attributed to possible side reactions
such as cyanide substitution and dimer formation.49 The peak at m/z = 684 is due to the
substitution of cyanide ion by excess INH and water, and the other peaks at m/z = 536 and
610 are due to the formation of dimers as well as substitution of cyanide.

Effect of pH on initial rate of reaction


In order to select a suitable pH value, corresponding to the optimum rate of reaction, the
pH of the reaction mixture was varied from 2.0 to 8.0, keeping all the other reaction
250 Progress in Reaction Kinetics and Mechanism 44(3)

(
(

Figure 3. Effect of pH for the reaction of [Ru(CN)5H2O]32 with isoniazid ligand in aqueous medium
under the optimum reaction experimental conditions: [INH] = 4 3 1023 M, [Ru(CN)5H2O]32
= 1 3 1024 M, ionic strength = 0.2 M and temperature = 25.0 6 0.01°C.

variables constant. The pH up to 6 was varied using potassium hydrogen phthalate/NaOH


or HCl buffer. A higher pH of all the working solutions was maintained using 5 M NaOH.
The effect of pH on the rate of reaction was studied, using the fixed time procedure method
as a measure of the initial rate. For each pH, the values of kobs and kf were determined and
kf (kf refers to formation rate constant) was plotted against pH as shown in Figure 3 which
indicates that the rate of reaction was initially low at low pH and increased with increasing
pH value up to 4.0 6 0.02, became maximum and then decreased slowly.
The equilibrium constant values for protonation of the hexacyanoruthenate(II) ion have
still not been reported in the literature, but they are assumed to follow the same pattern as
in the case of H + [Fe(CN)6]42 equilibria.50 At low pH, the reduction observed in rate and
ultimately in the kf values is attributed to the formation of various much less reactive proto-
nated species of [Ru(CN)6]42,51 in comparison to [Fe(CN)6]42. Two protonated species of
[Ru(CN)6]42, namely, monoprotonated H[Ru(CN)6]32 and diprotonated H2[Ru(CN)6]22,
have their protonation constants known,51 while two other protonated species of
[Ru(CN)6]42, namely, H3[Ru(CN)6]12 and H4[Ru(CN)6], have their protonation constants
still unknown in the literature.
This is the reason that distributions of all the species of the hexacyanoruthenate(II) ion
(for protonated and deprotonated forms) as a function of pH have not been obtained.
However, as provided in the literature, on the basis of various species of [Fe(CN)6]42 and
dissociation constants of INH, it is quite reasonable to say that deprotonated forms of
[Ru(CN)6]42 and INH are found to be the active species at a pH value of 4.00. Hence, it
supports the justified criteria for the constancy of the initial rate in the pH range between
2.0 and 8.0.
Yadav and Naik 251

(
(

(M)

Figure 4. Effect of ligand [INH] for the reaction of [Ru(CN)5H2O]32 with isoniazid in aqueous medium
under the optimum reaction experimental conditions: [Ru(CN)5H2O]32 = 4 3 1024 M, pH = 4.0 6 0.02,
ionic strength = 0.2 M and temperature = 25.0 6 0.01°C.

Effect of ligand [INH]


The effect of variation of INH on the initial rate was determined by varying its concentra-
tion from 4 3 1023 to 15 3 1024 M by keeping the fixed optimum conditions for other
reaction variables. The plot of kobs versus [INH] was a straight line (R2 = 0.9891), as shown
in Figure 4.
The reaction is considered to proceed through equations (3) and (4) as reported in our ear-
lier investigation9
 4 k1  3
RuðCNÞ6 þ H2 O
RuðCNÞ5 OH2 þ CN ðSlowÞ ð3Þ
k1

 3 k2  3
RuðCNÞ5 OH2 þ INH
RuðCNÞ5 ðINHÞ þ H2 O ð4Þ
k2

The lack of intercept in Figure 4 suggests the minimum role of the reverse reaction in
equation (4). A similar route for substitution reactions of [Ru(CN)5L]32 has also been
observed with some nitrogen-containing heterocyclic ligands.52

Effect of complex concentration


To study the effect of variation of concentration of [Ru(CN)5H2O]32, the rate of reaction
was measured as a function of [Ru(CN)5H2O]32, by changing its concentration in the range
4 3 1024–2 3 1024 M, keeping all other experimental variables fixed at optimum values.
The rate of reaction was found to be invariant with the concentration of [Ru(CN)5H2O]32.
252 Progress in Reaction Kinetics and Mechanism 44(3)

Figure 5. Effect of ionic strength for the reaction of [Ru(CN)5H2O]32 with isoniazid in aqueous medium
under the optimum reaction experimental conditions: [INH] = 4 3 1023 M,
[Ru(CN)5H2O]32 = 4 3 1024 M, pH = 4.0 6 0.02 and temperature = 25.0 6 0.01°C.

The variation of kobs with [Ru(CN)5H2O]32 is shown in Table S2 in the supplemental


information.

Effect of ionic strength (m)


The effect of ionic strength variation on the rate of reaction was studied by varying the con-
centration of NaClO4 within the experimental range 0.025–0.2 M, keeping all other experi-
mental variables fixed at their optimum values. The plot of ln(kf) versus I1/2/(1 + I1/2) was
found to be linear (Figure 5) (R2 = 0.9988). The graph was in accordance with the
Brønsted–Bjerrium–Christiansen relationship53 and found to exhibit a negative salt effect,
that is, as the ionic strength of the medium increases, the rate of reaction decreases.

Effect of temperature
The effect of variation of temperature on the reaction rate was analysed in the range 303–
318 K, keeping other reaction variables fixed. A higher temperature was avoided to reduce
the possibility of decomposition of the product. The values of the forward rate constant (kf)
were determined at different temperatures. The reaction followed Eyring’s equation. The for-
ward rate constant (kf) is related to the entropy of activation (DS6¼) and enthalpy of activa-
tion (DH6¼) through Eyring’s equation, given in equation (5)
     6¼ 
kf kB DS 6¼ =R DH
ln = ln e  ð5Þ
T h RT
Yadav and Naik 253

( (

Figure 6. Effect of temperature for the reaction of [Ru(CN)5H2O]32 with isoniazid in aqueous medium
under the optimum reaction experimental conditions: [INH] = 4 3 1023 M,
[Ru(CN)5H2O]32 = 4 3 1024 M, pH = 4.0 6 0.02 and ionic strength = 0.2 M (NaClO4).

The plot of ln (kf/T) versus 1/T exhibited a straight line as shown in Figure 6
(R2 = 0.9985). The values of activation parameters (DH6¼ and DS6¼) were estimated from
the slope and intercept of the above plot and are tabulated in Table 2. The high positive
DS6¼ value represents a greater degree of bond breaking (SN1 character), it follows therefore
that the given reaction operates through a dissociative pathway.

Mechanism
On the basis of kinetic data and activation parameters, the reaction under study preferen-
tially takes place through an ion-pair dissociation mechanism. The most plausible mechan-
ism is considered to take place through equations (6) and (7)
 3 k1  3
RuðCNÞ5 H2 O
RuðCNÞ5 + H2 O ð6Þ
k1

 3 k2  n3
RuðCNÞ5 + Ln ! RuðCNÞ5 L ð7Þ

where n is the charge on the incoming ligand.


After applying the steady-state approximation for [Ru(CN)5]32, one can deduce equation
(8)
 3  3
k1 RuðCNÞ5 H2 O = RuðCNÞ5 fk2 ½L + k1 ½H2 Og ð8Þ

After simplifying equation (8), we obtain the concentration of ½RuðCNÞ5 3 , given
through equation (9)
254 Progress in Reaction Kinetics and Mechanism 44(3)

Table 2. Thermodynamic activation parameters: enthalpy (DH6¼, kJ mol21), energy of activation


(Ea, kJ mol21), Gibbs free energy (DG6¼, kJ mol21) and entropy (DS6¼, J K21 mol21), for the cyanide
ion–substituted reaction between [Ru(CN)5H2O]32 and isoniazid (INH) ligand in aqueous medium
under the optimum experimental conditions and their change for reaction at 25.0 6 0.01°C.

Reaction condition – [Ru(CN)5H2O]32 = 4 3 1024 M, [INH] = 4 3 1023 M, pH = 4.0 6 0.02, ionic


strength (I) = 0.2 M (NaClO4) and temperature = 25.0 6 0.01°C
Thermodynamic activation parameters DH6¼ Ea DG6¼ DS6¼

Values 47.3 49.8 28.62 187.6

 
 3 k1 RuðCNÞ5 H2 O 3
RuðCNÞ5 = ð9Þ
k2 ½L + k1 ½H2 O

Since, according to equation (7), the rate of reaction depends on the concentration of
intermediate and ligand, it can be expressed through equation (10)
 3
Rate = k2 RuðCNÞ5 ½L ð10Þ

After solving equations (9) and (10), equation (11) is deduced


 3
k2 k1 ½L RuðCNÞ5 H2 O
Rate = ð11Þ
k2 ½L + k1 ½H2 O

Since the rate of reaction is directly proportional to the concentration of complex and
ligand, we can easily represent the rate through equations (12) or (13)
 3
Rate = kf ½L RuðCNÞ5 H2 O ð12Þ
 3
Rate = kobs RuðCNÞ5 H2 O ð13Þ

After substituting equation (13) into equation (11), the value of kobs can be represented
through equation (14)
k2 k1 ½L
kobs = ð14Þ
k2 ½L + k1 ½H2 O

Under the present experimental conditions


k1 ½H2 O..k2 ½L

Therefore, equation (14) reduces to equation (15)


k2 k1 ½L
kobs = ð15Þ
k1 ½H2 O

By considering
Yadav and Naik 255

k2 k1
= k ap
k1 ½H2 O

equation (15) can be transformed into equation (16)


kobs = k ap ½L ð16Þ

This equation proves that the observed rate constant follows the first-order rate law. The
plot of variation of the concentration of [INH] with kobs (kobs vs [INH]; Figure 4) shows a
straight line having no intercept, which is in good agreement with equation (16). The values
of kf = kap are obtained from the slope of the plot.

Conclusion
This work describes a successful method for the preparation of the [Ru(CN)5INH]32 com-
plex which could be a better medication than INH itself for the treatment of M. tuberculosis
like its [Fe(CN)5INH]32 counterpart. Further studies will be undertaken in our laboratory in
this regard. A tentative mechanistic scheme for the substitution of one of the six coordinated
cyanides in the hexacyanoruthenate(II) ion by the INH ligand has been proposed. The struc-
tural integrity of the synthesized complex has been made using IR, NMR, mass spectrometry
and elemental analysis.

Acknowledgements
The authors thank the Head, Department of Chemistry, Lucknow University, Lucknow, for providing
required departmental facilities to carry out the research work. The authors are grateful to the Director
of CDRI, Lucknow, for spectral analysis.

Declaration of conflicting interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship and/or
publication of this article.

Funding
The author(s) received no financial support for the research, authorship and/or publication of this
article.

Supplemental material
Supplemental material for this article is available online.

References
1. Kislenko VN and Olijnyk LP. Int J Chem Kinet 2000; 32: 184.
2. Baran Y and Ulgen A. Int J Chem Kinet 1998; 30: 415.
3. Prasad S. Anal Lett 2004; 37: 2851.
4. Naik RM and Sarkar J. Indian J Chem Technol 2005; 12: 567.
5. Naik RM, Agarwal A, Verma AK, et al. Int J Chem Kinet 2009; 41: 215–226.
6. Naik RM, Verma AK, Agarwal A, et al. Trans Met Chem 2009; 34: 209–215.
256 Progress in Reaction Kinetics and Mechanism 44(3)

7. Prasad S and Naik RM. Inorganic Reaction Mechanisms 2008; 6: 337–344.


8. Naik RM, Tiwari RK, Yadav SBS, et al. Prog React Kinet Mech 2009; 34: 211–226.
9. Naik RM, Singh R and Asthana A. Int J Chem Kinet 2011; 43: 21–30.
10. Naik RM, Tiwari RK, Yadav SBS, et al. Int J Chem Kinet 2012; 44: 1–9.
11. Naik RM and Kumar B. Prog React Kinet Mech 2012; 37: 147–160.
12. Naik RM and Kumar B. J Dispersion Sc Tech 2012.
13. Naik RM and Kumar B. Prog React Kinet Mech 2012; 37: 291-300.
14. Naik RM, Kumar B, Prasad S, et al. Microchem J 2015; 122: 82–88.
15. Agarwal A, Prasad S and Naik RM. Microchem J 2016; 128: 181–186.
16. Baraldo LM, Forlano P, Parise AR, et al. Coor Chem Rev 2001; 219: 881–921.
17. Tejera I, Jimenez R, Rodriguez A, et al. React Kinet Catal Lett 1992; 46: 427–434.
18. Moya ML, Burgess J and Sanchez F. Int J Chem Kinet 1993; 25: 469–477.
19. Lopez P, Sanchez F, Jimenez R, et al. Int J Chem Kinet 1996; 28: 57–60.
20. dela Vega R, Perez P, Pradogotor R, et al. Chem Phys 2004; 297: 163–169.
21. Allardyce CS and Dyson PJ. Plat Met Rev 2001; 45: 62–69.
22. Szulbinoki WS and Malato S. Pol J Chem 2001; 75: 1543–1551.
23. Alessio E, Iengo E, Serli B, et al. J Inorg Biochem 2001; 86: 21–30.
24. Puttaswamy Sukhdev A and Shubha JP. J Mol Catal A: Chem 2009; 310: 24.
25. Jagadeesh RV and Puttaswamy. J Phys Org Chem 2008; 21: 844.
26. Vinod KN, Puttaswamy and Gowda KNN. Ind Eng Chem Res 2010; 49: 3137.
27. Shetti NP, Malode SJ and Nandibewoor ST. J Mol Catal A: Chem 2011; 30: 1785.
28. Zhao M and Clarke MJ. J Biol Inorg Chem 1999; 4: 325.
29. Galardon E, Lc Maux P, Bondon A, et al. Tetrahedron: Asymmetry 1999; 10: 4203.
30. Frasca DR and Clarke MJ. J Am Chem Soc 1999; 121: 8523.
31. Povsc VG and Olabc JA. Transition Met Chem 1998; 23: 657.
32. Clarke MJ. Met Ions Biol Syst 1980; 11: 231.
33. Yasbin RE, Matthews GR and Clarke MJ. Chem BiolInteract 1980; 31: 355.
34. Mukherjee A and De K. Transition Met Chem 2005; 30: 677–683.
35. Korokovals A. Essentials of medicinal chemistry, 2nd ed. New York: Wiley-Interscience, 1998.
36. Argyrou A, Vetting MW, Aladegbami B, et al. Nat Struct Mol Biol 2006; 13: 408–413.
37. World Health Organization (WHO). World Health Organization Report Global tuberculosis
controls. Geneva: WHO, 2003.
38. Coelho AL, Moriea IS, Miguel AB, et al. Polyhedron 1994; 3: 1015.
39. Baran Y and Ulger A. Int J Chem Kinet 1998; 30: 415.
40. Fernando M, Francisco S and Burgess J. Transition Met Chem 2000; 25: 537.
41. Foretic B, Lovric J and Burger N. J Coord Chem 2006; 59: 1537.
42. Hoddenbagh JMA and Macartney DH. Inorg Chem 1986; 25: 380–2099.
43. Vogel’s text book of quantitative inorganic analysis. 4th ed. New York: Longman, 1978, p. 374.
44. Johnson CR and Shepherd RE. Inorg Chem 1983; 22: 2439–2444.
45. Tokman AL, Gentil LA and Olabe JA. Polyhedron 1989; 8: 2091–2097.
46. Schug K and Crean FM. Inorg Chem 1984; 23: 853.
47. Garafalo AR and Davies G. Inorg Chem 1976; 15: 1787.
48. Karmakar P, Bera BK, Barik KL, et al. J Coor Chem 2010; 63: 2158–2171.
49. Crean FM and Schug K. Inorg Chem 1984; 23: 853.
50. Jordan J and Ewing G. J Inorg Chem 1962; 1: 587.
51. Hicks KW and Chappelle GA. Inorg Chem 1980; 19: 1623.
52. Hoddenbagh JMA and Macartney DH. J Am Chem Soc 1982; 104: 7509.
53. Basolo F and Pearson RG. Mechanism of inorganic reactions: a study of metal complexes in
solutions, vol. 34. 2nd ed. New York: Wiley, p. 1967.

You might also like