Vous êtes sur la page 1sur 8

Article

pubs.acs.org/JPCC

Kinetic Analysis of the Catalytic Reduction of 4Nitrophenol by


Metallic Nanoparticles
Sasa Gu, Stefanie Wunder, Yan Lu, and Matthias Ballau*
Soft Matter and Functional Materials, Helmholtz-Zentrum Berlin fur Materialien und Energie, Hahn-Meitner-Platz 1,
14109 Berlin, Germany

Robert Fenger and Klaus Rademann


Department of Chemistry, Humboldt-Universitat zu Berlin, Brook-Taylor-Strasse 2, 12489 Berlin, Germany

Baptiste Jaquet
Institute for Chemical and Bioengineering, ETH Zurich, Wolfgang-Pauli-Strasse 10/HCI F123 ETH Zurich,
CH-8093 Zurich, Switzerland

Alessio Zaccone
Physik-Department and Institute of Advanced Study, Technische Universitat Munchen, 85748 Garching, Germany
S Supporting Information
*

ABSTRACT: We present a study on the catalytic reduction of 4-nitrophenol (Nip)


to 4-aminophenol (Amp) by sodium borohydride (BH4) in the presence of metal
nanoparticles in aqueous solution. This reaction which proceeds via the intermediate
4-hydroxylaminophenol has been used abundantly as a model reaction to check the
catalytic activity of metallic nanoparticles. Here we present a full kinetic scheme that
includes the intermediate 4-hydroxylaminophenol. All steps of the reaction are assumed to proceed solely on the surface of metal nanoparticles (LangmuirHinshelwood
model). The discussion of the resulting kinetic equations shows that there is a
stationary state in which the concentration of the intermediate 4-hydroxylaminophenol stays approximately constant. The resulting kinetic expression had been used
previously to evaluate the kinetic constants for this reaction. In this stationary state
there are isosbestic points in the UV/vis-spectra which are in full agreement with
most published data. We compare the full kinetic equations to experimental data
given by the temporal decay of the concentration of Nip. Good agreement is found underlining the general validity of the
scheme. The kinetic constants derived from this analysis demonstrate that the second step, namely the reduction of the
4-hydroxylaminophenol is the rate-determining step.

INTRODUCTION

nanoparticles or for a given type of nanoparticle immobilized in


dierent carrier systems.
In recent years, the reduction of 4-nitrophenol (Nip) to
4-aminophenol (Amp) by borohydride (BH4) in aqueous
solution has become such a model reaction that meets all
criteria of a model reaction.10 It can be monitored easily with
high precision by UVvis spectroscopy.1013 This is due to the
fact that Nip has a strong absorption at 400 nm and the decay
of this peak can be measured precisely as the function of time.
Moreover, the reaction rate is small enough so that the conversion

Metallic nanoparticles (NP) have been the subject of intense


research during the recent years because of their potential use
in catalysis.16 It is now well-established that even inert metals
such as gold may become active catalysts when divided down to
nanoscale.7,8 Very often, nanoparticles are attached to a suitable
colloidal carrier for easier handling and in order to avoid potential hazards.9 However, these carrier systems may impede the
activity of the nanoparticles for a given reaction. Comparing the
catalytic activity of nanoparticles bound in various systems
hence requires a model reaction, that is, a well-controlled reaction without side reactions.10 Kinetic data and rate constants
obtained from such a reaction can be compared for dierent
2014 American Chemical Society

Received: June 18, 2014


Revised: July 22, 2014
Published: July 24, 2014
18618

dx.doi.org/10.1021/jp5060606 | J. Phys. Chem. C 2014, 118, 1861818625

The Journal of Physical Chemistry C

Article

we have modeled this stationary state in terms of an apparent


reaction rate kapp (see Figure 1). We20 and others25 demonstrated that this rate constant can be fully evaluated in terms of
a LangmuirHinshelwood kinetics: Both reactants, namely
Nip and BH4 must be adsorbed on the surface to react. This
kinetic model has met with gratifying success when compared
to experimental data.19,20 However, the theory presented in
refs 19 and 20 only models the decay rate of Nip; no follow-up
products are considered.
Here we present a full kinetic analysis of the reduction of Nip
in the presence of metal nanoparticles. The present analysis
aims at a quantitative understanding of the entire kinetics
starting just after the delay time and the subsequent transition
to the stationary state. We consider the direct route15 that takes
place on the surface. The goal of this work is to develop a
kinetic model for the entire dependence of the concentration of
Nip on time and the comparison of this model with the experimental data given in ref 20. The model is general, however, and
applies to reductions catalyzed by other nanoparticles as well.
The only prerequisite is that all steps take place at the surface.

can be conveniently monitored over several minutes. The


presence of isosbestic points in the UVvis spectra measured at
dierent time gives clear evidence that Nip is fully reduced to
the nal product Amp, no byproducts can be detected. The
analogous reaction, namely the reduction of nitrobenzene is a
well-studied reaction. Since the classical work by Haber,14 the
various intermediates are well-known:15 In the so-called direct
route nitrobenzene is reduced to nitrosobenzene and then to
phenylhydroxylamine. In the nal step, phenylhydroxylamine is
reduced to aniline. In the condensation route, the intermediates
nitrosobenzene and phenylhydroxylamine react to form
azoxybenzene which is reduced subsequently to aniline. Recent
work has clearly revealed that in the presence of gold particles
as catalyst the reduction proceeds only along the direct route,
no traces of azoxybenzene and the following products are
found.1620 This nding has been explained by a strong adsorption of all intermediates to the surface of the nanoparticles.
Practically all published studies agree on this point and assume
that the catalysis takes place on the surface of the nanoparticles.
The work of Nigra et al. is the notable exception.21 These
authors state that the catalysis of the reduction of Nip in the
presence of gold nanoparticles is aected by a soluble species
leaching from the metal nanoparticles. The concentration of
this soluble gold species must be very low and its catalytic
activity in turn very high. However, the reduction of Nip by
BH4 is catalyzed by many other noble metals such as Pt,19
Ag,22,23 Pd,24 Ru,25 and alloys.26 Hence, one must postulate
soluble species of all these metals with similarly high catalytic
activity. Moreover, the recent study by Mahmoud et al.27 has
given clear evidence that the reaction is proceeding at the
surface. Additional evidence is given in recent experimental
work.13,25,27
Recently, we have analyzed the kinetics of this reaction in great
detail.19,20 Figure 1 shows a typical absorbance spectra measured

KINETICS
In analogy to the well-studied case of the reduction of nitrobenzene,1517 we formulate the reaction in terms of the direct
route15,16 shown in Figure 2.28 Two intermediates may be

Figure 2. Proposed mechanism (direct route) of the reduction of


4-nitrophenol by metallic nanoparticles: In step A, 4-nitrophenol
(Nip) is rst reduced to the nitrosophenol which is quickly converted
to 4-hydroxylaminophenol (Hx). This compound is the rst stable
intermediate. Its reduction to the nal product, namely 4-aminophenol
(Amp), takes place in step B, which is the rate-determining step. There
is an adsorption/desorption equilibrium for all compounds in all steps.
All reactions take place at the surface of the particles.

identied, namely 4-nitrosophenol and 4-hydroxylaminophenol. The rst stable intermediate is the 4-hydroxylaminophenol
as is well borne-out of the studies done on nitrobenzene.1517
Thus, we have three compounds that adsorb and desorb during
the reaction cycle, namely 4-nitrophenol (Nip), 4-hydroxylaminophenol (Hx) and 4-aminophenol (Amp). We assume furthermore that all three compounds compete for a xed number of
surface sites on the surface of the nanoparticles.
Let cNip, cHx, and camp be the actual concentrations of Nip,
4-hydroxylaminophenol, and of Amp, respectively. The surface
coverage Nip of Nip is modeled in terms of a Langmuir
Freundlich isotherm.19 Hence, we have

Figure 1. Typical time dependence of the absorption of


4-nitrophenolate ions at 400 nm. The blue portion of the line displays
the linear section, from which kapp is taken. The induction period t0
which 20 s in this case is marked with the black arrow.

as the function of time. At rst, there is a delay time in which


no reaction takes place. The induction period t0 was related to
a surface restructuring of the nanoparticles before the catalytic
reaction starts.19,20 Hence, a rearrangement of the surface
atoms seems to be necessary to create catalytically active sites
as, e.g., corners or edges on the surface. Subsequently, the reaction starts and after an intermediate period a stationary state is
reached that may last for many minutes. In our previous work,

Nip =

(KNipcNip)n
1 + (KNipcNip)n + KHxcHx + KBH4c BH4

(1)

where KNip, KHx, and KBH4 are the Langmuir adsorption constants of the respective compounds, and n is the Langmuir
Freundlich exponent. Following ref 19, n was set to 0.5. The
coverage Hx and BH4 of 4-hydroxylaminophenol, and of
18619

dx.doi.org/10.1021/jp5060606 | J. Phys. Chem. C 2014, 118, 1861818625

The Journal of Physical Chemistry C

Article

borohydride, respectively, are formulated using the classical


Langmuir isotherm, that is, n = 1. The reaction is now modeled
in two steps (cf. Figure 2) termed A and B: First Nip is reduced
to 4-hydroxylaminophenol in step A. The reduction of the
latter compound is done in step B. Hence, the rate of reaction
of Nip follows as

dcNip
dt

dc
= kappcNip = kaSNipBH4 = Hx
dt source

(2)

where S denotes the total surface of all nanoparticles in the


solution. This equation follows directly from the fact that SNip
is proportional to the number of all adsorbed molecules in the
system while BH4 denotes the conditional probability to nd an
adsorbed surface hydrogen atom near to an adsorbed Nip
molecule. As a tacit assumption in the entire LH kinetics, the
total number of adsorbed molecules is much smaller than the
total number of molecules of a given species in solution; that is,
adsorption on the surface of the catalyst does not shift the
concentration in the system in a detectable way. Moreover, it is
assumed that the adsorption equilibrium between the solution
and the surface of the catalyst is established quickly.
Given these assumptions and prerequisites, the reaction rate
for step A follows as

dcNip
dt

= kaS

Figure 3. Idealized time dependence of the concentration of


4-nitrophenol and denition of the dierent stages of the reaction
(see the discussion of eqs 5 to 10). The black line shows the concentration of 4-nitrophenol as the function of time whereas the red dashed
line corresponds to the concentration of 4-hydroxylaminophenol. The
early stage I, where 4-nitrophenol is reduced to 4-hydroxylaminophenol,
is mainly determined by step A of the reaction (see Figure 2) while the
concentration of the nal product 4-aminophenol is still small. The
decay rate in this stage is approximated by kapp,1 given by eq 6. Stage II,
starting at time tS (cf. eq 11), is the stationary state characterized
by kapp,II that can be approximated by eq 8. Here the concentration of
4-hydroxylaminophenol is approximately constant. In this idealized
picture, the stationary concentration cHx,stat of this compound follows
from the balance of its generation (step A; cf. Figure 2) and decay (step
B; cf. Figure 2) and may be approximated by eq 7. This concentration
equals the decay of the concentration of 4-nitrophenol at time tS.

(KNipcNip)n KBH4c BH4


[1 + (KNipcNip)n + KHxcHx + KBH4c BH4]2

dc
= Hx
dt source

(3)

nanoparticles and thus slow down the rate of reaction.


Moreover, it is evident that 4-hydroxylaminophenol is very
strongly adsorbed to the surface of the nanoparticles. This can
be argued from the fact that its formation slows down the rate
of reaction while not accumulating in solution.
The present model does not consider the adsorption/
desorption equilibrium of the nal product 4-aminophenol. In
principle, a strong adsorption of Amp on the surface of the
particles would strongly inuence the kinetics of the reaction, at
least in the nal state when the concentration of Amp has risen.
In case of strong adsorption, the reaction rate should decrease
markedly since more and more places are blocked by Amp.
This is not observed in any study so far. Hence, we disregard
this possibility in the present model.
A full solution of the kinetic problem thus dened consists in
the simultaneous solution of eq 3 and 4 which must be done
numerically. This procedure leads to the concentration of
nitrophenol cNip as the function of time that can directly be
compared to experimental data. However, in rst approximation a simple solution may be found: After the initial state,
we may postulate a stationary state in which

The intermediate 4-hydroxylaminophenol thus generated is


further reduced to the nal product Amp in step B and its rate
of decay may be formulated through
dc
Hx
dt decay
= kbS

KHxcHxKBH4c BH4
[1 + (KNipcNip)n + KHxcHx + KBH4c BH4]2

dcamp
dt

Hence, the full rate equation for the generation and decay of
the intermediate 4-hydroxyaminophenol is given by
dc
dc
dcHx
= Hx
Hx
dt source dt decay
dt
= kaS
kbS

(KNipcNip)n KBH4c BH4


[1 + (KNipcNip)n + KHxcHx + KBH4c BH4]2
KHxcHxKBH4c BH4
[1 + (KNipcNip)n + KHxcHx + KBH4c BH4]2

(4)

dcHx
=0
dt

Equations 3 and 4 now constitute a set of coupled rate


equations that allows us to discuss the entire kinetics of the
reaction. In the following we give a brief qualitative discussion
of these equations: First of all, it is evident that kA kB. This
can be seen directly from the fact that the initial rate of reaction
as determined from the tangent in Figure 1 for t > t0 is
much larger than the tangent in the stationary state. Hence,
4-hydroxylaminophenol is formed rather quickly but its further
reduction in step B is much slower. When its concentration
rises quickly in the early stage of the reaction, it will more and
more compete with nitrophenol for the surface places of the

(5)

In this approximation, the reaction kinetics after t0 may be


divided into two regimes depicted schematically in Figure 3:
(i) Early regime, ranging from t0 to a time ts: Here cHx 0
and

18620

dcNip
dt

kaS

(KNipcNip)n KBH4c BH4


[1 + (KNipcNip)n + KBH4c BH4]2

(6)

dx.doi.org/10.1021/jp5060606 | J. Phys. Chem. C 2014, 118, 1861818625

The Journal of Physical Chemistry C

Article

Figure 4. Fit of the concentration of Nip as the function of time by the numerical solution of eq 3 and 4. The concentration of Nip was normalized
to the respective starting concentration cNip,0. The experimental data have been taken from ref 20 and refer to a temperature of 10 C (data points
with error bars). The solid lines refer to the ts by the kinetic model.

In this regime, no isosbestic point can be expected since the


spectra of 3 compounds varying with time are superimposed.

In this stationary state, the amount of aminophenol generated


per unit time is exactly given by the decay rate of nitrophenol. If
the stationary concentration of the 4-hydroxylaminophenol is
small, the condition for the isosbestic point is restored. Hence,
the constants kapp, that is, the tangents of the absorbance as the
function of time, are given for the two limiting cases by

(ii) Stationary state (t > ts) in which cHx is approximately


constant (see Figure 3): eq 5 leads to the condition that

cHx , stat =

ka(KNipcNip)n
kbKHx

(7)

(I) Early regime from t0 to ts:

Thus,

dcNip
dt

= kaS

(KNipcNip)n KBH4c BH4

n
1 + (KNipcNip) 1 +

ka
kb

+ KBH4c BH4

kapp , I = kaS
2

(KNip)n (cNip)n 1KBH4c BH4


[1 + (KNipcNip)n + KBH4c BH4]2

(9)

where tS is the time where stationary state starts (see Figure 3).

dcamp
dt

(II) Stationary state for t > ts:

(8)
18621

dx.doi.org/10.1021/jp5060606 | J. Phys. Chem. C 2014, 118, 1861818625

The Journal of Physical Chemistry C


kapp , II = kaS

Article

n
KNip
(cNip)n 1KBH4c BH4

n
1 + (KNipcNip) 1 +

ka
kb

)+K

BH4c BH4

(10a)

The concept of a stationary state leads immediately to the


conclusion that the stationary concentration of 4-hydroxylaminophenol cHx,stat should be given in good approximation by the
amount of nitrophenol that has reacted at t = ts. Thus, we
assume that the subsequent conversion to the Hx has not taken
place to a notable degree. Hence, with cNip,0 being the concentration of nitrophenol at t = 0, we get for the time ts where the
stationary state has been reached
cNip ,0 cHx , stat
ln
= kapp , I(ts t0)
cNip ,0
which may be approximated through
ts t0 =

cHx , stat
cNip ,0kapp , I

[1 + (KNipcNip ,0)n + KBH4c BH4]2


kbSKBH4c BH4KHx
(11)

Equations 7 and 11 give the predictions for the onset of the


stationary state. Figure 3 summarizes all stages of this kinetic
scheme together with temporal evolution of the concentrations
of all reactants expected from this model.
It is interesting to compare this result to the previous version
of theory that did not take into account explicitly the intermediates. With this simplication we obtained for the stationary
state (see eq 3a of ref 19 or eq 5 of ref 20)
kapp , II = kaS

Figure 5. Fit of the concentration of Nip as the function of time by the


numerical solution of eq 3 and 4. The concentration of Nip was
normalized to the respective starting concentration cNip,0. The
experimental data have been taken from ref 20 and refer to a
temperature of 30 C (data points with error bars).The solid lines refer
to the ts by the kinetic model.

n
KNip
(cNip)n 1KBH4c BH4

[1 + (KNipcNip)n + KBH4c BH4]2

(10b)

which diers from eq 10a only by a factor 1 + ka/kb in the


denominator. The adsorption constant KHx does not appear in
eq 10a because of the stationary state condition of eq 5. There
are two limiting cases that can be derived from eq 10a: (i) ka
kb, that is, 4-hydroxylaminophenol reacts much faster than 4nitrophenol. In this case, eq 10b is a good approximation and
the reduction of 4-nitrophenol is the rate-determining step. (ii)
ka kb. Now the reduction of 4-hydroxylaminophenol becomes
the rate-determining step, and the reaction is slowed down due
to the additional factor in the denominator of eq 10b.

rst input for KNip, KBH4, ka, kb, and n. Then every theoretical
cNip,th as the function of time was compared to the corresponding experimental data cNip,exp. The calculation is repeated
until most of calculated data of cNip,th match the corresponding
experimental data sets cNip,exp.
Second, the reaction rate of steps A and B (ka, kb) may be
dierent at dierent initial reaction concentrations, so the
values of ka and kb were reoptimized using MatLab routine II
(see Supporting Information). This routine can only analyze
one cNip,exp at one time. The values of ka and kb were changed
while keeping KNip, KBH4, KHx, and n obtained by routine I constant until full agreement was reached.
Third, the error bars of these parameters were also checked
by MatLab routine II. Changing one parameter at one time,
cNip,th was compared to the corresponding experimental data
cNip,exp to check whether the value was within the error bars.
Evidently, the consumption of the Hx intermediate cannot be
measured directly and the t values for kb and KHx are less
precise to get from this t than the other parameters.

NUMERICAL SOLUTION OF THE KINETIC


EQUATIONS
All data analyzed here have been taken from ref 20. In all cases,
the delay time t0 has been subtracted as discussed previously.19,20 The concentration of Nip as the function of reaction
time, cNip,exp, was then analyzed by a numerical solution of eq 3
and 4 by two MatLab routines (see the full sheets in the
Supporting Information). The Matlab routines were used to
calculate the theoretical Nip concentration cNip,th as the function
of time for a given values of KNip, KBH4, KHx, ka, kb, and n. These
data are compared to the experimental results and the constants
are changed until agreement with the experiment is reached. In
the following we give the details of this procedure.
First, all cNip,exp data obtained for a given temperature were
put into MatLab routine I (see the Supporting Information).
Routine I calculates cNip,th for a given set of values of KNip, KBH4,
KHx, ka, kb and n. The parameters from ref.20 were used as a

RESULTS AND DISCUSSION


Figure 4 and 5 display examples of the ts of the experimental
data at temperature of 10 C (Figure 4) and 30 C (Figure 5)
obtained by a simultaneous numerical solution of eq 3 and 4.
The concentration of Nip normalized to the respective concentration cNip,0 at t = 0 is plotted as the function of time for
dierent initial concentrations of Nip and BH4. To ensure a
18622

dx.doi.org/10.1021/jp5060606 | J. Phys. Chem. C 2014, 118, 1861818625

The Journal of Physical Chemistry C

Article

Figure 6. Kinetic constants ka and kb obtained from the comparison of theory and experiment for (a, b) 10 and (c, d) 30 C, respectively. The
dashed lines give the average value of the constants.

Table 1. Constants Derived from the Fits of the Measurements at Dierent Temperatures
temp [C]
10
20
25
30

average ka [104mol/m2 s]
4.2
9.4
9.7
11.5

0.9
2.6
2.9
3.9

average kb [105mol/m2 s]
2.1
5.6
7.8
7.1

0.9
1.4
1.7
1.5

KNip [L/mol]
2700
3700
4600
5200

500
900
1200
1500

KBH4 [L/mol]
30
50
62
86

2
4
6
10

KHx [L/mol]

0.5
0.5
0.5
0.5

150000
160000
175000
200000

10000
15000
20000
25000

intermediates compete for free places at the surface of Au nanoparticles, and the reaction can occur only between species
adsorbed on the surface. If most places are occupied by a single
species, such as Hx, the reaction will be slowed down. For this
reason the accumulation of Hx slows down the apparent
reaction rate when the reaction approaches stage II.
Figure 6 gathers the reaction rates of steps A and B derived
from tting at 10 and 30 C. The rate constants ka and kb
scatter around a mean values indicated by a dashed line in
Figure 6. It should be noted that the constant kb is derived in an
indirect fashion since the experiment measures only the decay
of Nip. Given the various uncertainties of the analysis, the
agreement of theory and experiment may be regarded as
satisfactory. Table 1 gathers the resulting constants.
Figure 6 shows that kb is much smaller than ka. Moreover, the
adsorption constant of intermediate Hx is considerably greater
than that of the other components. Evidently, the reduction of
Hx is rate-determining step of the reaction and the accumulation
of Hx on the surface slows down the reaction when stage II is
reached. This strong adsorption of Hx on the surface of the
particles precludes the formation of other products as e.g. the
substituted azoxybenzenes. The formation of the latter compounds requires the presence of a sucient concentration of

meaningful comparison, all curves are plotted up to a conversion of 30%. The results of other ts taken at dierent concentrations and temperatures are given in the Supporting
Information. The solid lines are the ts by theory. It is clear
that the early stage and the transition to the state II which is
clearly seen for the data taken at 10 C (Figure 4) are wellmodeled by the kinetic scheme given in Figure 2. In general,
cNip,th deviates from cNip,exp more for higher Nip concentrations.
These deviations are clearly seen at longer reaction time. Partial
hydrolysis of BH4 is an unavoidable side reaction, in particular
at higher temperatures, and will shift its concentration during
the measurements. Moreover, the model assumes the strict
validity of the Langmuir adsorption isotherm which may not be
fully valid anymore when going to higher concentration of
4-nitrophenol.
The resulting t parameters are plotted in Figure 6 and are
summarized in Table 1. Obviously, a single set of constants
KNip, KBH4 and KHx is capable of describing the experimental
data at a given temperature, at least in the early stage up to
conversions of ca. 30%. The much larger value of KHx proves
that intermediate hydroxylamine is much stronger adsorbed on
the surface of the nanoparticles than the other components.
In the LangmuirHinshelwood model, the reactants and
18623

dx.doi.org/10.1021/jp5060606 | J. Phys. Chem. C 2014, 118, 1861818625

The Journal of Physical Chemistry C

Article

Figure 7. Dependence of the adsorption constants KNip of Nip (a), the adsorption constants KBH4 of borohydride (b), the adsorption constants KHx
of 4-hydroxyaminophenol (c), the reaction rate of step A ka (d) on the inverse of temperature.

the nitroso- and the hydroxylamino compounds which is not


the case.
Figure 7 displays that the adsorption constants and reaction
rate ka increase with an increasing temperature which can be
determined as a function of temperature. The enthalpies and
entropies for the adsorption of Nip, BH4 and Hx can be obtained from the dependence of the adsorption constants on
temperature through
ln K =

Table 2. Summary of Enthalpy and Entropy Values of the


Adsorption of Nip, BH4, and Hx
H[kJ/mol]
S[J/mol K]

KNip

KBH4

KHx

24 3
150 12

37 2
158 6

10 1
133 3

H
S
+
RT
R

Table 2 summarized the value of thermodynamic parameters.


All adsorption processes are endothermic. Here the compound
with larger adsorption constants has smaller enthalpy. The H
and S of the adsorption process of Nip and BH4 are larger
than those obtained from previous version of theory20 that did
not take into account the intermediates. The activation energy
for the reduction of Nip to Hx obtained from an Arrhenius plot
of ka, is 36.1 3.3 kJ/mol.

DISCUSSION OF THE STATIONARY STATE


ASSUMPTION
Figure 8 displays the temporal evolution of the concentration of
4-hydroxylaminophenol as calculated from the numerical
solution in comparison to the decay of 4-nitrophenol. Quite
evidently, the concentration of this main intermediate is rising
steadily throughout the time 300 s used for the evaluation of
the data. After reaching the maximum the concentration of
4-hydroxylaminophenol decreases slowly. For the range in

Figure 8. Calculated concentrations of Nip and Hx as the function


of time. The initial concentrations of Nip and BH4 are 0.04 mM and
5 mM, respectively.

which ln(cNip) varies linearly with time, the assumption of a


stationary state may be regarded as satisfactory. But clearly the
full kinetic scheme as developed here is superior and leads to a
more consistent description of all the data and should be
preferred for the analysis.
18624

dx.doi.org/10.1021/jp5060606 | J. Phys. Chem. C 2014, 118, 1861818625

The Journal of Physical Chemistry C

Article

(10) Herves, P.; Perez-Lorenzo, M.; Liz-Marzan, L. M.; Dzubiella, J.;


Lu, Y.; Ballauff, M. Catalysis by Metallic Nanoparticles in Aqueous
Solution: Model Reactions. Chem. Soc. Rev. 2012, 41, 55775587.
(11) Pradhan, N.; Pal, A.; Pal, T. Silver Nanoparticle Catalyzed
Reduction of Aromatic Nitro Compounds. Colloid Surf., A:
Physicochem. Eng. Asp. 2002, 196, 247257.
(12) Esumi, K.; Isono, R.; Yoshimura, T. Preparation of PAMAM
and PPIMetal (Silver, Platinum, and Palladium) Nanocomposites
and Their Catalytic Activities for Reduction of 4-Nitrophenol.
Langmuir 2004, 20, 237243.
(13) Fenger, R.; Fertitta, E.; Kirmse, H.; Thunemann, A. F.;
Rademann, K. Size Dependent Catalysis with CTAB-stabilized Gold
Nanoparticles. Phys. Chem. Chem. Phys. 2012, 14, 93439349.
(14) Haber, F. Z. Gradual Electrolytic Reduction of Nitrobenzene
with Limited Cathode Potential. Elektrochem. Angew. Phys. Chem.
1898, 22, 506514.
(15) Blaser, H. U. A Golden Boost to an Old Reaction. Science 2006,
313, 312313.
(16) Corma, A.; Concepcion, P.; Serna, P. A Different Reaction
Pathway for the Reduction of Aromatic Nitro Compounds on Gold
Catalysts. Angew. Chem. 2007, 119, 74047407.
(17) Corma, A.; Serna, P. Chemoselective Hydrogenation of Nitro
Compounds with Supported Gold Catalysts. Science 2006, 313, 332
334.
(18) Layek, K.; Kantam, L.; Shirai, M.; Nishio-Hamane, D.; Sasaki,
T.; Maheswarana, H. Gold Nanoparticles Stabilized on Nanocrystalline
Magnesium Oxide as an Active Catalyst for Reduction of Nitroarenes
in Aqueous Medium at Room Temperature. Green Chem. 2012, 14,
31643174.
(19) Wunder, S.; Polzer, F.; Lu, Y.; Mei, Y.; Ballauff, M. Kinetic
Analysis of Catalytic Reduction of 4-Nitrophenol by Metallic
Nanoparticles Immobilized in Spherical Polyelectrolyte Brushes. J.
Phys. Chem. C 2010, 114, 88148820.
(20) Wunder, S.; Lu, Y.; Albrecht, M.; Ballauff, M. Catalytic Activity of
Faceted Gold Nanoparticles Studied by a Model Reaction: Evidence for
Substrate-induced Surface Restructuring. ACS Catal. 2011, 1, 908916.
(21) Nigra, M. M.; Ha, J.; Katz, A. Identification of Site
Requirements for Reduction of 4-Nitrophenol Using Gold Nanoparticle Catalysts. Catal. Sci. Technol. 2013, 3, 29762983.
(22) Santos, K. O.; Elias, W. C.; Signori, A. M.; Giacomelli, F. C.;
Yang, H.; Domingos, J. B. Synthesis and Catalytic Properties of Silver
NanoparticleLinear Polyethylene Imine Colloidal Systems. J. Phys.
Chem. C 2012, 116, 45944604.
(23) Baruah, B.; Gabriel, G. J.; Akbashev, M. J.; Booher, M. E. Facile
Synthesis of Silver Nanoparticles Stabilized by Cationic Polynorbornenes and Their Catalytic Activity in 4-Nitrophenol Reduction.
Langmuir 2013, 29, 42254234.
(24) Johnson, J. A.; Makis, J. J.; Marvin, K. A.; Rodenbusch, S. E.;
Stevenson, K. J. Size-Dependent Hydrogenation of p-Nitrophenol with
Pd Nanoparticles Synthesized with Poly(amido)amine Dendrimer
Templates. J. Phys. Chem. C 2013, 117, 2264422651.
(25) Antonels, N. C.; Meijboom, R. Preparation of Well-Defined
Dendrimer Encapsulated Ruthenium Nanoparticles and Their
Evaluation in the Reduction of 4-Nitrophenol According to the
LangmuirHinshelwood Approach. Langmuir 2013, 29, 13433
13442.
(26) Kaiser, J.; Leppert, L.; Welz, H.; Polzer, F.; Wunder, S.;
Wanderka, N.; Albrecht, M.; Lunkenbein, T.; Breu, J.; Kumme, S.; Lu,
Y.; Ballauff, M. Catalytic Activity of Nanoalloys from Gold and
Palladium. Phys. Chem. Chem. Phys. 2012, 14, 64876495.
(27) Mahmoud, M. M.; Garlyyev, B.; El-Sayed, M. A. Determining
the Mechanism of Solution Metallic Nanocatalysis with Solid and
Hollow Nanoparticles: Homogeneous or Heterogeneous. J. Phys.
Chem. C 2013, 117, 2188621893.
(28) Haas, S.; Fenger, R.; Fertitta, E.; Rademann, K. Cascade
Catalysis of Highly Active Bimetallic Au/Pd Nanoclusters: Structure
function Relationship Investigation Using Anomalous Small-angle Xray Scattering and UVVis Spectroscopy. J. Appl. Crystallogr. 2013, 46,
13531360.

CONCLUSION
A kinetic scheme for the reduction of Nip by BH4 catalyzed by
metal nanoparticles in aqueous solution has been presented.
The analysis is based on the reaction shown in Figure 2:
4-nitrophenol is rst reduced to 4-hydroxylaminophenol which
subsequently is reduced to the nal product 4-aminophenol.
The kinetic scheme leads to the coupled dierential eqs 3 and 4
that upon numerical solution describe the decay of Nip with
time. Good agreement between theory and experiment is found.
In particular, the entire temporal evolution of the concentration
of Nip can be described while the earlier approach19,20 was only
capable of describing the stationary state. Moreover, the analysis
of the stationary state used in earlier analysis of this reaction can
be derived directly from this model. An isosbestic point is
predicted for the stationary state which is observed indeed for
the great majority of the published experimental data.

ASSOCIATED CONTENT

S Supporting Information
*

Tables of parameters from the simulation by MATLAB and


optimized ka and kb values and gures showing the ts of the
concentration of Nip and kinetic constants ka and kb, and the
MATLAB routine. This material is available free of charge via
the Internet at http://pubs.acs.org.

AUTHOR INFORMATION

Corresponding Author

*(M.B.) E-mail address: Matthias.Ballau@helmholtz-berlin.de.


Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
A.Z. gratefully acknowledges nancial support of the IAS at
TUM via the Moessbauer Fellowship.

REFERENCES

(1) Henglein, A. Small-particle Research: Physicochemical Properties


of Extremely Small Colloidal Metal and Semiconductor Particles.
Chem. Rev. 1989, 89, 18611873.
(2) Astruc, D. Nanoparticles and Catalysis; Wiley-VCH: New York,
2008.
(3) Ferrando, R.; Jellinek, J.; Johnston, R. L. Nanoalloys: From
Theory to Applications of Alloy Clusters and Nanoparticles. Chem.
Rev. 2008, 108, 845910.
(4) Burda, C.; Chen, X. B.; Narayanan, R.; El-Sayed, M. A. Chemistry
and Properties of Nanocrystals of Different Shapes. Chem. Rev. 2005,
105, 10251102.
(5) An, K.; Somorjai, G. A. Size and Shape Control of Metal
Nanoparticles for Reaction Selectivity in Catalysis. ChemCatChem.
2012, 4, 15121524.
(6) Carchini, G.; Almora-Barrios, N.; Blonski, P.; Lopez, N. How
Theoretical Simulations Can Address the Structure and Activity of
Nanoparticles. Top. Catal. 2013, 56, 12621272.
(7) Haruta, M.; Kobayashi, T.; Sano, H.; Yamada, N. Novel Gold
Catalysts for the Oxidation of Carbon Monoxide at a Temperature far
Below 0 C. Chem. Lett. 1987, 14, 405408.
(8) Zhang, Y.; Cui, X.; Shi, F.; Deng, Y. Nano-Gold Catalysis in Fine
Chemical Synthesis. Chem. Rev. 2012, 112, 24672505.
(9) Kaiser, J.; Szczerba, W.; Riesemeier, H.; Reinholz, U.; Radtke, M.;
Albrecht, M.; Lu, Y.; Ballauff, M. The Structure of AuPd Nanoalloys
Anchored on Spherical Polyelectrolyte Brushes Determined by X-ray
Absorption Spectroscopy. Faraday Discuss. 2013, 162, 4555.
18625

dx.doi.org/10.1021/jp5060606 | J. Phys. Chem. C 2014, 118, 1861818625

Vous aimerez peut-être aussi