Rate constants for elementary steps including adsorption, desorption, and chemical reactions on surfaces are calculated using the classi-cal collision theory and transition state theory.
Trang 1SurfKin: An Ab Initio Kinetic Code for Modeling Surface
Reactions
In this article, we describe a C/C11 program called SurfKin
(Surface Kinetics) to construct microkinetic mechanisms for
modeling gas–surface reactions Thermodynamic properties of
reaction species are estimated based on density functional
theory calculations and statistical mechanics Rate constants
for elementary steps (including adsorption, desorption, and
chemical reactions on surfaces) are calculated using the
classi-cal collision theory and transition state theory Methane
decomposition and water–gas shift reaction on Ni(111) surface
were chosen as test cases to validate the code implementa-tions The good agreement with literature data suggests this is
a powerful tool to facilitate the analysis of complex reactions
on surfaces, and thus it helps to effectively construct detailed microkinetic mechanisms for such surface reactions SurfKin also opens a possibility for designing nanoscale model cata-lysts.V C 2014 Wiley Periodicals, Inc
DOI: 10.1002/jcc.23704
Introduction
Microscopic understanding of gas–surface reactions has always
been an interest and also a challenge in surface chemistry,
specifically in determination of detailed reaction mechanisms
The molecular-level information of a reaction network is
essen-tially the starting point of developing a microkinetic model for
the understanding of the chemistry/physics occurring on
cata-lyst surfaces under realistic reaction conditions
Characteriza-tion of elusive surface intermediates is a very challenging task,
which cannot be easily done by performing experiments only
It is widely known that semiempirical kinetic models, or
power law kinetic models can provide a well-described picture
at the macroscopic scale, but the lack of detailed information of
reacting species at the molecular-level limits their applicability
to develop reliable kinetic models to capture a wide range of
reaction conditions For example, kinetic models for ammonia
decomposition over various transition metals were developed
based on experimental data.[1–3]In these studies, assumptions
are usually made on the rate-determining steps and dominant
surface coverages, which depend on actual conditions,[4]as
fit-ting parameters The unity bond index-quadratic exponential
potential (UBI-QEP) method is another semiempirical approach
that provides whole surface reaction energetics for constructing
microkinetic models.[5–11]In this approach, heats of adsorption,
reaction enthalpies, and activation energies were calculated
within 1–3 kcal/mol to the experimental thermodynamic
param-eters.[5,6] Due to its empirical nature, this practical method is
simple and effective to predict the energetics of surface
inter-mediates.[6] However, the UBI-QEP method cannot describe
accurately the nonenergetic contributions to rate coefficients,
and results from quantum mechanical methods are essential in
this aspect Additionally, compared to UBI-QEP, density
func-tional theory (DFT) calculations provide a more solid framework
to obtain reliable rate parameters; thus it can be effectively
extended to a wide range of reaction conditions Recent
advan-ces in DFT-based electronic structure calculations and
experi-mental techniques have validated the microkinetic models derived from the first-principles methods, which have become a bridge between microscopic properties and macroscopic per-formances.[12–18] Many microkinetic models have been devel-oped for a variety of important surface processes using common DFT-based computational tools, including SIESTA,[18] DACAPO,[12–16] and Vienna ab initio simulation package (VASP),[17]in combination with HREELS experiments.[12–14,16]The information derived from these calculations or experiments is the basis to create sequences of elementary reaction steps and estimate rate parameters for each step using statistical thermo-dynamics,[12,14–16] collision theory,[14,15] and transition state theory (TST).[14,16,17] The developed microkinetic models are essential for simulation of model reactors It can be seen that the development of microkinetic models from the first-principles calculations has become a powerful method for studying catalytic surface processes
In this article, we presented a C/C11 program called Surf-Kin (Surface Surf-Kinetics) for modeling gas–surface reactions from first-principles methods The code uses detailed kinetic mecha-nisms from DFT-based calculations Alternatively, it can be combined with the data obtained either from experiments or
[a] T N.-M Le , L K Huynh Molecular Science and Nano-Materials Laboratory, Institute for Computa-tional Science and Technology, Quang Trung Software Park, Dist 12, Ho Chi Minh City, Vietnam
E-mail: lamhuynh.us@gmail.com [b] B Liu
Department of Chemical Engineering, Kansas State University, 1005 Durland Hall, Manhattan, Kansas, 66506
[c] L K Huynh Applied Chemistry Department, School of Biotechnology, International Uni-versity, Vietnam National UniUni-versity, Ho Chi Minh City, Vietnam.
Contract grant sponsor: Department of Science and Technology, Ho Chi Minh City (L.K.H.); Contract grant sponsor: Kansas State University (B.L.)
V C 2014 Wiley Periodicals, Inc.
Trang 2simulations (if available) to model complex chemical processes
in real conditions Statistical mechanics is used to estimate
ther-modynamic properties, such as entropies and enthalpies for
both gas-phase and adsorbed species Kinetic analyses are
per-formed based on kinetic theories, such as the collision theory
and the canonical TST The analyses of methane decomposition
and water–gas shift reaction on a model Ni(111) surface were
performed to illustrate the applications of SurfKin as an
effec-tive tool that integrates quantum mechanical calculations and
statistical mechanics to study surface chemistry
Theoretical Methods
Statistical thermodynamic analysis
Thermodynamic properties (e.g., entropy and enthalpy) of the
reaction species were calculated using a well-established
statis-tical mechanical approach Details on thermodynamic property
calculations for gas-phase molecules can be found
else-where.[19] In this section, we only briefly describe the
imple-mentation for adsorbed molecules in SurfKin
The thermodynamic properties of adsorbed species can be
effectively derived from the total partition function, which can
be factored out into four corresponding components as follows
qtotal5qelectronic3qtranslation3qrotation3qvibration (1)
Electronic partition function (qelectronic) The contribution of
the electronic partition function depends on how high the
temperature and energy difference between ground state and
the first excited state Usually, the difference is too high
com-pared to kBT in the common temperature range of interest
(i.e., T < 2000 K) As a result, the electronic partition function is
restricted to the ground state Therefore, the focus is on the
contributions of translational, rotational and vibrational
parti-tion funcparti-tions to the total partiparti-tion funcparti-tion of a species of
interest
If a molecule strongly binds to the surface, translation and
rotation are considered as frustrated motions and thus
effec-tively treated as harmonic vibrations In the case of weakly
bound or indirect adsorption, the molecules create a precursor
state on the surface that translation on two dimensions and
rotation about direction perpendicular to the surface (defined
as z-axis) must be explicitly considered.[16]
Translational partition function (qtranslation) Translational
parti-tion funcparti-tion for a weakly bound species on surface takes the
following form
q2DtranslationðA; TÞ5 2pmkBT
h2
where m is the species mass, kB and h are Boltzmann and
Planck constants, respectively, and A is the surface area per
binding site, which depends on surface site density
character-izing for each single surface (a typical value of A for Ni(111) is
5.365 3 10220 m2 assuming for the p(2 3 2) cell and the fcc,
hcp, and atop binding sites[16])
Rotational partition function (qrotation) For adsorbed species, there is only the rotation about the z-axis of the center of mass, thus the rotational partition function takes the form
q2Drotation5p
1=2
rh ðIZZÞ1=28p2kBT1=2
where r is the symmetry number and IZZ is moment of inertia about the z-axis that passing through the center of mass of the species
Vibrational partition function (qvibration) For species with Nvib normal modes, the vibrational partition function is given by
qvibration5YN vib
i51
qvib
ð Þi5YN vib i51
1 12e2bhcm i
where b5 1
k B T, c is the speed of light and mi(cm21) denotes the ith vibrational frequency
Thermodynamic property calculations From the above partition functions, standard molar entropy (S0), standard molar enthalpy (H0) can be calculated using the following equations
S02D2translationðTÞ5R ln 2pmkBT
h2 A
11
(5)
S02D2rotationðTÞ5R ln p
1=2
rh ðIZZÞ1=28p2kBT1=2
11 2
(6)
S0vibrationðTÞ5RX
i
hcmi=kBT
ebhcm i212ln 12e
2bhcm i
(7)
H05Eelectronic 1 ZPE 1 U0corrections (8)
where Eelectronic is the total electronic energy, ZPE51
2
P
ihcmi is the zero point energy, the correction to the internal energy,
U0 corrections, includes all thermal corrections at standard molar state, namely
U0corrections5U0translation1U0rotation1U0vibration (10)
U02D2rotationðTÞ51
U0vibrationðTÞ5RTXN vib
i
bhcmi exp hcmi
k B T
21
(13)
Transition state theory Conventional TST was applied to calculate rate constants for elementary reactions/steps which have intrinsic reaction bar-riers,[20] which can be depicted in Figure 1 Within the TST framework, rate constant has the general form
Trang 3h
q0 TS
q0exp 2DETS
kBT
(14)
where q0
TS, q0 are the partition functions for the transition state (TS) and reactants with respect to its own ground states,
respectively The energy barrier DETSTSTS (or Ea in the
conven-tional notation) is the energy difference between the TS and
the reactant(s) On the surface, it is divided into three
individ-ual processes to calculate rate constants, namely adsorption,
desorption and reactions between adsorbed species
Reactions between adsorbed species The reaction scheme for
reactions between adsorbed species (cf Figure 1) can be
expressed as
A1 B
eq
ABTS m AB1 h (15)
The rate constants as a function of temperature for the forward
direction can be derived within the TST framework as follows
kforwardðTÞ5kBT
h
q0
AB TS
qAqB
expð2DETS
kBTÞ; (16) where DETS is defined as
DETS5ðE1ZPEÞAB
TS2 ðE1ZPEÞA2 ðE1ZPEÞB (17)
To fulfill the thermodynamic consistency, the reverse rate
constant is derived from Van’t Hoff equation,
KeqðTÞ5k
forwardðTÞ
kreverseðTÞ5exp 2
DGrxn
kBT
(18) and DGrxn5DHrxn2 TDSrxn (19)
Adsorption of molecules Two adsorption models, direct
adsorption and indirect adsorption, are considered The main
difference is in how strong or weak the molecules binding to
the surface If it is a strong binding case (e.g., adsorption energy is larger than DH0
rxn) or direct adsorption, the molecules immediately land on the surface with only vibrational motion Translation and rotation are considered as frustrated motions and treated as harmonic vibrations.[16] In the case of weakly bound or indirect adsorption, the molecules are in a precursor state on the surface that translation on two dimensions and rotation about z-axis are considered.[16]
Similar to the reaction on the surface, the adsorption pro-cess for molecule M in the gas-phase can be schematically presented as M 1 h
eq
M TS m
M For direct adsorption, rate constant can be expressed as
kdirectadsorptionðTÞ5 NAh
2 2pmMkBT
ffiffiffiffiffiffiffiffiffiffiffiffi
kBT 2pmM
s
rdirectð Þ;T (20)
where
rdirectð Þ5T q
vib
M TS
qrot
Mqvib M
exp 2DETS
kBT
(21)
and
DETS5ðE1ZPEÞM
TS2ðE1ZPEÞM2ðE1ZPEÞh
(22) For indirect adsorption, TS is weakly bound to the surface The molecules can freely move on surface, thus they can con-tribute to translational, rotational and vibrational partition functions The rate constants can be described as
kindirectadsorptionðTÞ5NAA2
ffiffiffiffiffiffiffiffiffiffiffiffi
kBT 2pmM
s
rindirectð ÞT (23)
rindirectð Þ5T q
rot
M
TSqvib
M TS
qrot
Mqvib M
exp 2DETS
kBT
(24)
As can be seen from (21) and (24), the indirect model takes into account the translation and rotation of the TS, while the direct model considers these degree of freedoms as vibration-like modes on the surface Therefore, it is important to determine the contribution of energetic degree of freedoms for the TS species The rate constant for desorption process can be calculated by equilibrium relation or using the following equation explicitly
kdesorptionðTÞ5kBT
h
q0
M TS
qM exp 2DETS; desorption
kBT
; (25)
where
DETS; desorption5ðE1ZPEÞM
TS2ðE1ZPEÞM (26)
Collision theory Within the simpler collision theory framework, the adsorption process can be written as following, with gas-phase M and
Figure 1 Schematic representation of a surface reaction with an intrinsic
barrier h* denotes an active surface site and A*, B*, and AB* are the
adsorbed species, and ABTSare the adsorbed species at the TS [Color
fig-ure can be viewed in the online issue, which is available at
wileyonlineli-brary.com.]
Trang 4adsorbed species M, M 1 h M(h is the active surface
site) Collision theory is used to calculate the rate of
adsorp-tion processes with the general formula[14,15]
radsorptioncollision 5Ar T; hcoverage
PM ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2pmMkBT
kBT
where r T; h coverage
is the sticking coefficient for the collision process The estimation of sticking coefficients is discussed
below PM is the partial pressure of gas-phase species M, and
DEfis the activation energy for adsorption process It is
essen-tial to extract an adsorption rate constant depending only on
temperature from the general adsorption rate It should be
noted the site density is N05N
S51for a surface with N adsorp-tion sites and total surface area of S, each with an area of A
The sticking coefficient rðT; hcoverageÞ depends on the
tempera-ture T and the free-site surface coverage hcoverage The sticking
coefficient can be written in the form of two factors, sticking
coefficient of clean surface rðTÞ and a function of surface
cov-erage r h coverage
.[21] The gas-phase species M is assumed as
an ideal gas, the relation between pressure and concentration
for ideal gas is widely known as
PM5 n V
h i
Substitute the expanded form of sticking coefficient and eq
(28) into eq (27), the adsorption rate expression becomes
radsorptioncollision ðT;½MÞ5NAAr Tð Þ
ffiffiffiffiffiffiffiffiffiffiffiffi
kBT 2pmM
s exp 2DEf
kBT
V
h i (29)
From eq (29), the adsorption rate constant can be extracted
as
kcollision
adsorptionðTÞ5NAArðTÞ
ffiffiffiffiffiffiffiffiffiffiffiffi
kBT 2pmM
s exp 2DEf
kBT
(30)
The adsorption processes are usually nonactivated,[14] that
is, DEf50, so the adsorption rate constant can be rewritten in
a simpler form
kadsorptioncollision ðTÞ5NAArðTÞ
ffiffiffiffiffiffiffiffiffiffiffiffi
kBT 2pmM
s
5rðTÞ C
ffiffiffiffiffiffiffiffiffiffiffiffi
kBT 2pmM
s
(31)
where C5 1
N A A is the surface site density and the typical value
of C is 3.095 3 1029(mol/cm2) for Ni(111); NA is the Avogadro
constant; A is the area of each adsorption site and the typical
value of A is 5.365 3 10220 m2.[16] The desorption rate
con-stants can be calculated through the equilibrium relations
rep-resented by eq (18)
Sticking coefficient for barrierless adsorption
Sticking coefficient for barrierless adsorption can be understood
as the ratio of the rate of adsorption onto surface to the rate of
collisions with surface The coefficient is controlled by both
enthalpic and entropic contributions which have opposing effects on the variational behavior of the TST rate coefficient For most exothermic adsorption processes, the sticking coefficients are typically unity The entropic effect (the tendency of adsorb-ate migration on surface) will affect the value of sticking coeffi-cient and make it less than unity For several reactions, the initial adsorption can be the rate controlling step; therefore, in this study, we tried to calculate the sticking coefficient for barrierless adsorption as a function of adsorbate–adsorbent perpendicular position z (A˚) using formula proposed by Pitt et al.[22]
In this approach, the reaction adsorption potential and the barrier height to migration on the surface as a function of z are needed and can be explicitly calculated from DFT calculations
SurfKin Program Interpretation
The structure of Surfkin program is schematically shown in Fig-ure 2 SurfKin program uses C/C11 language to take advan-tages of object-oriented programming pattern, which is convenient for defining properties of molecules, linking and processing data as well All molecules are held in a unique class type because of their similar data structure such as molecular names, masses, energies, vibration frequencies, geo-metries, and so forth The surface is a periodic structure, where reactions occur with active sites represented by h* In SurfKin,
an active site is treated as a reaction species, so it also has the properties of a surface adsorbate molecule This concept can
be conveniently adopted in SurfKin because it helps reduce processes for defining a new class, or creating linkage to others molecules The program is coded by a modulization method, which allows us to manipulate all the involving files properly Each module performs its individual functions, and then they are linked together for a specific calculation task
The first step is to prepare the input data for the program, including a database and a control file The database is stored
in a folder containing files of the information of all species, which characterizes the system of interest In the database, the files *.erg and *.freq contains information of molecular energies and frequencies, which is used to calculate energy barriers, par-tition functions, and thermodynamic quantities The ground state energy data of all species can be used to construct the potential energy surface (PES) Molecular geometries are stored
in file *.geom, which is used to calculate moments of inertia for
a specific species through its center of mass The control file is independent of the database It contains information of the databases for reactants, TSs, products as well as calculation parameters (e.g., temperatures and pressures)
The program starts by reading input from the control file to get required information It is also directed to the current database path The program will check if the species of inter-est from the control file are in the designated database If that
is the case, calculations are ready for the next steps in sepa-rate modules that calculate energy barriers, ZPE corrections, moments of inertia, partition functions, entropies, and enthal-pies Using these precalculated quantities, the rate constants and equilibrium constants are calculated at the conditions of interest
Trang 5Case study 1: Methane Decomposition on
Ni(111)
Methane decomposition on metal-based catalysts is a crucial
cess for methane steam reforming used mainly for hydrogen
pro-duction and fuel cell applications There are many successful
microkinetic models developed from both semiempirical UBI-QEP
method, in combination with experimental data[10,11] and
DFT-based calculations,[16,23]to investigate methane steam reforming
over nickel under realistic conditions These studies have
con-structed full microkinetic models for methane steam reforming
with detailed reaction mechanisms Simulations of reforming
reactor models have been performed as well In this study, this
methane decomposition system was used as a test case for the
SurfKin application Specifically, thermodynamic and kinetic
anal-yses were performed in the framework of DFT periodic
calcula-tions and classical statistical mechanics approach
Computational details
Periodic DFT calculations were performed using the VASP,[24–27]a
periodic, plane wave-based code The ionic cores are described
by the projector augmented wave method,[28,29]and the Kohn–
Sham valence states were expanded in the plane wave basis sets
up to 385 eV The exchange-correlation energy is described by
the generalized gradient approximation with the revised
Perdew-Burke-Ernzerhof (RPBE) functional.[29,30]
A three layer, close-packed Ni(111) surface with a vacuum of
12 A˚ between successive metal slabs The DFT-determined
lat-tice constant is found to be 3.52 A˚, which compares well with
the experimental bulk lattice constant (3.52 A˚ ).[31]
A p(2 3 2) unit cell equivalent to 1/4 monolayer was used The top layer is
relaxed for all geometry optimizations The surface Brillouin zone is sampled with a (6 3 6 3 1) mesh based on Mon-khorst–Pack scheme.[32] The ionic relaxation was stopped until the forces on all free atoms are less than 0.02 eV/A˚ A Methfes-sel–Paxton smearing of 0.2 eV was applied.[33] The total ener-gies are then extrapolated to kBT 5 0 eV The ZPE corrections were calculated from DFT vibrational analyses, and dipole cor-rections are also included.[34]The total energy of methane was calculated in a box with dimensions of 18 3 19 3 20 A˚ The gamma-point k point sampling is used The Gaussian smearing parameter is 0.01 eV To account for the magnetic properties of
Ni, all calculations were performed with spin polarizations The TS structures were initially estimated using the climbing image-nudged elastic band method.[35,36] The dimer method was then used to further refine the determined TSs.[37,38] Vibra-tional frequencies were calculated Each TS was confirmed to have only one imaginary (negative) vibrational mode
Reaction mechanism The sequence of elementary steps is constructed within the Langmuir–Hinshelwood framework The detailed reaction mechanism is given in Table 1, with four elementary reaction
steps (both directions on each step), including one gas-phase species (i.e.,
CH4), five adsorbed spe-cies (i.e., CH3*, CH2*, CH*, C*, and H*) and four TSs (i.e., HACH3*, HACH2*,
HACH*, and CAH*) Note that in this study we
Figure 2 Flowchart of calculation modules in SurfKin.
Table 1 Elementary reaction steps for methane dissociations on Ni(111).
h* represents an active surface site.
Trang 6focus our effort on the decomposition of methane; thus
sub-sequent important steps for the intermediate products in
methane steam reforming, such as the formation and
desorption of hydrogen, are not included These parameters
are derived from DFT calculations The adsorption of
meth-ane is treated as a dissociative adsorption process to directly
form CH3* and H* with a transition state (TS1 in Figure 3)
The remaining adsorbed species, including the other TSs, are
treated as strongly bound states with only vibrational
contri-bution to thermodynamic properties
Potential energy surface
The calculated PES for methane decomposition over Ni(111) is
shown in Figure 3, comparing to the values reported by Blaylock
et al.,[16]where the gas-phase methane and clean Ni(111) surface
is used as the reference state The vertical axis is the relative
energies at 0 K, the horizontal axis is the reaction coordinate
The energies with ZPE corrections are used for discussion,
other-wise it will be stated There are four TSs, that is, TS1 (HACH3*),
TS2 (HACH2*), TS3 (HACH*), and TS4 (CAH*), for a complete
decomposition of methane From the energy differences between species in the PES, it is easily seen that the reactions via TS1, TS2, and TS4 are endothermic, while the reaction via TS3 is exothermic The temperature dependence will be dis-cussed in the following sections The highest barrier occurs at the last step via TS4, breaking CAH bond; thus this step is prob-ably the slowest step at 0 K (113.0, 56.3, 29.9, and 127.4 kJ/mol for TS1, TS2, TS3, and TS4 routes, respectively) which is in good agreement with the trend proposed by Li et al.[39]However, the route via TS1 becomes the slowest step at higher temperature, which is consistent with earlier observations.[16,40,41] This issue will be discussed further in the kinetic analysis
Thermodynamic property analysis Table 2 presents the thermodynamic properties, namely DH0
rxn,
DS0 rxn, and Keq5exp 2DG0
rxn=RT
for each reaction step at
1073 K where the industrial steam methane reforming is actually performed If CH4* is treated as a weakly bound spe-cies, the DGo
rxnfor the adsorption is highly positive of 131.8 kJ/
mol (or Keq is much smaller than unity) This indicates low
Figure 3 Calculated PES (ZPE correction included) at 0 K for the decomposition of methane on Ni(111) The numbers in parentheses are the imaginary
fre-quencies of the TSs The inset figure plots the free energies associated with methane decomposition on Ni(111) at 600 and 1073 K.
Table 2 Comparisons of thermodynamic and kinetic parameters calculated in this work and results from Blaylock et al.[16]at 1073 K (800 C).
3.85 3 10210[b]
6.45 3 10 21[b]
1.28 3 1021[b]
1.79 3 10 23[b]
[a] This work Arrhenius prefactors A and activation energies E a are fitted from TST rate constants over a temperature range of 300–1500 K [b] Blaylock
et al [16] [c] Due to a small energy difference between [CH4(g) 1 h*] and CH4*, CH4* can be considered as a physisorbed state and the initial states are
CH4* (translation, rotation in 2D) and h*.
Trang 7coverage for CH4* and the desorption is more favorable at this
high temperature This is consistent with early observation by Lee
et al.[40]When compared with Blaylock’s data, which treated
meth-ane as an dissociative adsorption species, our numbers are in a
good agreement (e.g., 57.9 kJ/mol, 2123.3 J/mol-K for enthalpy
and entropy change of CH412h*! CH3* 1 H* compared with
the corresponding numbers of 59.8 kJ/mol, and 2124 J/mol-K,
respectively) In addition, it can be seen that equilibrium constants
of reactions 1–4 are the same orders of magnitude, similar to the
results from Blaylock et al.[16]at the same temperature, 1073 K The
differences are within a factor of two, which is likely due to the
uncertainty of different DFT parameters used in the two studies
Such a good agreement with available literature values for this
well-defined system provides us more confidence on our
calcu-lated numbers and our implementation
Kinetic analysis
The kinetic data for methane decomposition on Ni(111) is given
in Table 3 The parameters A and Eaare derived from fitting TST
rate constants to the simple Arrhenius expression,
k5A3expð2Ea=RTÞ, in a temperature range of 300–1500 K
Table 3 lists kinetic parameters at different temperatures for
methane decomposition process on Ni(111) In this process, the
rate constants for all elementary reactions show a trend that the
rate constants increase with temperature Because the decrease
of the equilibrium constant for the adsorption/desorption with
temperature (as discussed earlier), the methane adsorption
favors at low temperature energetically Comparison of
calcu-lated values to Blaylock’s results was presented in Table 2 The
highest activation energy, about 133 kJ/mol, occurs when
reac-tants are passing through TS4 (Reaction 4 in Table 2) This result
can be compared to the value 135 kJ/mol suggested by Blaylock
et al.[16]Moreover, the forward rate constants are much smaller
than the reverse ones with the variation of temperature
Ener-getically, the channel via TS4, CH*! C* 1 H*, is expected to be
the rate-limiting step, the slowest step in the reaction network,
due to its highest barrier of 127.4 kJ/mol at 0 K (cf Figure 3)
However, it is noticed that the entropy contribution to the rate
constants for this reaction is larger than that of reaction 1
(CH412h* ! CH3* 1 H*, via TS1), reflected by the A-factor of
5.20 3 1013 and 2.95 3 1011 for these two reactions,
respec-tively This make the later reaction (reaction 1) the slowest step
at high temperature (e.g., k(1073 K) 5 6.04 3 105vs 1.67 3 107
for reactions 1 and 4, respectively) This is a demonstration of
the importance of kinetic analysis in order to explain and/or
pre-dict surface reaction pathways This observation is consistent with previous study, where the first CAH bond cleavage of CH4
is the rate-limiting step.[16,40,41]
The good agreement with available data on thermodynamic properties and kinetics of methane dissociation on Ni(111) suggests SurfKin is an effective tool that integrates quantum mechanical calculations and statistical mechanics to study reactions on surfaces
Case study 2: Water-Gas Shift Reaction on Ni(111)
Water–gas shift reaction (WGSR), CO(g)1H2O(g) Nið111Þ
CO2(g)1H2(g), plays a key role in reducing the side product, carbon monoxide, as well as boosting hydrogen production in steam reforming of methane This reaction has been studied extensively on a wide range of transition metal systems show-ing overall energetic trends,[42]singly on Cu,[43,44]Cu- based,[45]
Pt,[46] Ni,[47] or coupling in steam reforming process on
Ni.[10,16,48]Among these candidates, it has been shown that Ni
is a good catalyst for this reaction though the reaction mecha-nism and kinetics on this surface are not fully understood.[47]
In this second test case, the reaction mechanism, based on DFT calculations, for WGSR on Ni(111) is proposed and the thermodynamic and kinetic properties are analyzed using the SurfKin program Because there is no research developing full microkinetic model for WGSR on Ni(111) surface, our results will be compared to DFT calculations from Catapan et al.,[47]
and the results extracting from the work of Blaylock et al.[16]
Reaction mechanism The WGSR mechanism is described within Langmuir–Hinshelwood framework The sequence of elementary reactions (both forward
and reverse directions) via the carboxyl species
is the main channel, showing in Table 4 There are four gas-phase species including CO(g),
H2O(g), CO2(g), and H2(g); six adsorbates, that is, CO*, H2O*, H*, OH*, COOH*, and COO*; and three TSs, that is,
HAOH*, COAOH*, and
Table 3 Prefactors A, activation energies E a , and rate constants for methane decomposition on Ni(111).[a]
Rate constants
[a] Prefactors A and activation energies E a are obtained by fitting calculated rate constants to the simple Arrhenius expression over a temperature range of 300–1500 K.
Table 4 Elementary reaction steps of water–gas shift reaction on Ni surface.
h* represents an active surface site
Trang 8HACOO* For the adsorption of gas-phase species apart from H2(g),
which is dissociative adsorption, the adsorption energies (including
ZPE corrections) for CO(g), H2O(g), CO2(g)on Ni(111) surface,
compar-ing to Blaylock’s results[16]given in the parentheses, are 2197.65
(2144.75), 20.34 (21.93), 0.82 (2.89) kJ/mol, respectively Small
binding energies of H2O(g)and CO2(g)suggesting that these surface
species will be treated as the free translational and rotational
spe-cies on Ni(111)
Potential energy surface
Figure 4 presents the PES for WGSR in the free energy
land-scape The gas-phase species, CO(g), and H2O(g), and the
clean Ni(111) surface are used as the reference state The
vertical axis is the relative electronic energies at 0 K
(includ-ing ZPE correction), the horizontal axis is the reaction
coor-dinate There are three TSs, namely TS1 (HAOH*), TS2
(COAOH*), and TS3 (COOAH*) for the carboxyl energetic
pathway From the energy difference between species in the
PES, it is shown that the reaction via TS2 is endothermic,
while the reactions via TS1 and TS3 are exothermic For the
forward WGSR, the highest barrier occurs at the step via TS2, forming COOH* from CO* and OH* This is probably the slowest step at 0 K (i.e., 127.63, 88.4, and 78.0 kJ/mol for TS2, TS1, and TS3 routes, respectively) It is worth mention-ing that the findmention-ing on the lowest reaction step is via the carboxyl formation and energetic trend of the PES for WGSR
on Ni(111) is in agreement with the recent DFT results from Catapan et al.[47]
Thermodynamic analysis Thermodynamic properties for the WGSR on Ni(111) at the practical condition of steam reforming are shown in Table 5
Only the carboxyl formation step via TS2 (reaction 4) is endo-thermic while the other reactions are exoendo-thermic The posi-tive entropy change and negaposi-tive enthalpy change for COO*
forming via TS3 (reaction 5) indicated that this reaction is the most thermodynamically favorable reaction on the Ni(111) surface in the current selected pathway There is large entropy loss, and large negative enthalpy change associated with CO adsorption, resulting in stable adsorbed CO species
Figure 4 Calculated PES at 0 K for WGSR on Ni(111) with the carboxyl formation pathway ZPE correction is included The free energy profiles along the
reaction coordinate at different temperatures are shown on the right subfigure.
Table 5 Thermodynamic data at standard molar state for WGSR on Ni(111) at 600 and 1073 K.
rxn (kJ/mol) DS 0
Trang 9on Ni(111) The equilibrium constant for this reaction is very
large (4.59 3 108at 600 K) and decreases by seven orders of
magnitude at a temperature of real steam reforming
condi-tion (1.47 3 101 at 1073 K), showing that CO conversion is
favorable at low temperature For the adsorption of H2O, H2,
and CO2, there is much entropy lost at the reaction
condi-tions comparing to enthalpy, which is equivalent to the
increase of free energy with increasing temperatures This
means that these processes are unfavorable at higher
tem-peratures, which can be seen clearer at free energy profiles
at 600 and 1073 K (cf Figure 4) The dissociation of H2O* via
TS1 (reaction 3) seems to be disadvantageous at a higher
temperature, but this reaction also has a rather high barrier
and the increasing temperature is favorable for the kinetics
The predicted thermodynamic properties are further
con-firmed by kinetic analysis
Kinetic analysis
Calculated kinetic information for WGSR on Ni(111) is given in
Table 6 The Arrhenius parameters, A and Ea, are independent
of temperature and derived from fitting rate constants to
Arrhenius expression in the temperature range of 300–1500 K
The highest activation energy (125.9 kJ/mol) occurs at the
car-boxyl forming step (reaction 4), corresponding to a low rate
constant (1.08 3 106 1/s) at 1073 K This rate constant is the
same order of the one for breaking OH group (1.05 3 1061/s)
from H2O* (reaction 3), which makes reactions 3 and 4 the
lowest reactions at the real reaction conditions The sticking
coefficients for barrierless adsorption of gas-phase species,
that is, CO, CO2, and H2O, were derived with the calculated
adsorption potential (V) and the barrier height to migration
(V0) on the surface with the similar assumptions made by
Gra-bow et al.[46] Our calculation results are consistent with the
common assumption of unity for all flat metal crystal faces as
V0 is much smaller than V and goes to zero much more
quickly than V reaches the reaction enthalpy, DH0.[22] Within the collision theory, the calculated rate constant values for these adsorption processes seem to have similar orders of magnitude The calculated activation energies are compared
to results from Blaylock et al.,[16] which are the values in the parentheses The maximum difference of activation energy is about 15 kJ/mol (125.9 vs 111 kJ/mol for this work and the lit-erature, respectively), occurring at the reaction 4 This discrep-ancy is much due to the DFT-calculated results for electronic energy barriers, that is, 124.7 versus 111.4 kJ/mol for this work and the literature, respectively In comparing between prefac-tors, the largest difference (3.78 3 1010 vs 1.4 3 1011 for this work and the literature, respectively) occurs at the breaking step of OH group from H2O* (reaction 3) This difference is in the acceptable range that gives the same order in the rate constant
Conclusions
The C/C11 SurfKin program has been successfully developed
in an attempt to construct microkinetic models for gas-surface reactions Thermodynamic properties of reaction spe-cies were estimated based on ab initio calculations and statis-tical mechanics Rate constants for elementary steps (including adsorption, desorption and chemical conversion on surfaces) can be obtained using kinetics/dynamics models from, that is, collision theory and TST The good agreement with available data in the literature for the methane decom-position and WGSR on Ni(111) surface suggests this is a powerful tool using DFT calculation data to explore complex gas-surface reactions in a wide range of conditions and it opens a possibility to effectively construct detailed microki-netic mechanisms for modeling real complex processes The code currently does not include simulation on reactor mod-els, as well as a graphical user interface (GUI) These features are being developed
Table 6 Prefactors (A),[a]activation energies (E a )[a]and rate constants for WGSR on Ni(111).
Rate constants
[a] Prefactors A and activation energies Ea(kJ/mol) of the simple Arrhenius expression, kðTÞ5A3exp ð2E a =RT Þ, are fitted from the calculated rate con-stants over a temperature range of 30021500 K unless otherwise noted [b] For the adsorption reactions 1, 2, 6, and 7, the rate concon-stants (cm 3 /mol-s) are calculated from collision theory [cf eq (31)] For desorption processes, the rate constants (1/s) are derived from equilibrium constants and the adsorption rate constants, to which normalizing factor [RT/p] is added [c] From Blaylock et al [16]
Trang 10The authors greatly appreciate the computing resources and
sup-port provided by the Institute for Computational Science and
Tech-nology—Ho Chi Minh City, International University—VNU-HCMC,
the high performance computing clusters hosted by Golden Energy
Computing Organization (GECO) at Colorado School of Mines, and
Fusion, a 320-node computing cluster operated by the Laboratory
Computing Resource Center at Argonne National Laboratory
Keywords: gas-surface reactionthermodynamicsrate
con-stantmicrokinetic mechanismmethane decompositionwater
gas shift reaction
How to cite this article: T Nguyen-Minh, B Liu, L K Huynh J
Comput Chem 2014, 35, 1890–1899 DOI: 10.1002/jcc.23704
[1] R W McCabe, J Catal 1983, 79, 445.
[2] G Papapolymerou, V Bontozoglou, J Mol Catal A: Chem 1997, 120,
165.
[3] S Armenise, E Garcia-Bordeje, J L Valverde, E Romeo, A Monzon,
Phys Chem Chem Phys 2013, 15, 12104.
[4] L J Broadbelt, R Q Snurr, Appl Catal A: General 2000, 200, 23.
[5] E Shustorovich, A T Bell, Surf Sci Lett 1991, 259, L791.
[6] E Shustorovich, H Sellers, Surf Sci Rep 1998, 31, 1.
[7] M Wolf, O Deutschmann, F Behrendt, J Warnatz, Catal Lett 1999, 61,
15.
[8] D Mantri, P Aghalayam, Catal Today 2007, 119, 88.
[9] S Appari, V M Janardhanan, S Jayanti, L Maier, S Tischer, O.
Deutschmann, Chem Eng Sci 2011, 66, 5184.
[10] L Maier, B Sch€ adel, K Herrera Delgado, S Tischer, O Deutschmann,
Top Catal 2011, 54, 845.
[11] E S Hecht, G K Gupta, H Zhu, A M Dean, R J Kee, L Maier, O.
Deutschmann, Appl Catal A 2005, 295, 40.
[12] A Andreasen, H Lynggaard, C Stegelmann, P Stoltze, Surf Sci 2003,
544, 5.
[13] S Linic, M A Barteau, J Catal 2003, 214, 200.
[14] A A Gokhale, S Kandoi, J P Greeley, M Mavrikakis, J A Dumesic,
Chem Eng Sci 2004, 59, 4679.
[15] S Kandoi, J Greeley, M Sanchez-Castillo, S Evans, A Gokhale, J.
Dumesic, M Mavrikakis, Top Catal 2006, 37, 17.
[16] D W Blaylock, T Ogura, W H Green, G J O Beran, J Phys Chem C
2009, 113, 4898.
[17] X.-M Cao, R Burch, C Hardacre, P Hu, Catal Today 2011, 165, 71.
[18] H.-F Wang, Y.-L Guo, G Lu, P Hu, J Phys Chem C 2009, 113, 18746.
[19] P Atkins, J D Paula, Physical Chemistry; Oxford University Press:
Oxford, 2006; pp 590–649.
[20] I Chorkendorff, J W Niemantsverdriet, Concepts of Modern Catal and Kinetics; Wiley-VCH Verlag GmbH & Co KGaA: Weinheim, 2003; p 108.
[21] R D Cortright, J A Dumesic, Advances in Catalysis; Academic Press:
San Diego, 2001; pp 161–264.
[22] I G Pitt, R G Gilbert, K R Ryan, J Phys Chem 1994, 98, 13001.
[23] B Liu, M T Lusk, J F.Ely, Surf Sci 2012, 606, 615.
[24] G Kresse, J Hafner, Phys Rev B 1993, 47, 558.
[25] G Kresse, J Hafner, Phys Rev B 1994, 49, 14251.
[26] G Kresse, J Furthm€ uller, Phys Rev B 1996, 54, 11169.
[27] G Kresse, J Furthm€ uller, Comput Mater Sci 1996, 6, 15.
[28] P E Bl€ ochl, Phys Rev B 1994, 50, 17953.
[29] B Hammer, L B Hansen, J K Nïrskov, Phys Rev B 1999, 59, 7413.
[30] J P Perdew, J A Chevary, S H Vosko, K A Jackson, M R Pederson,
D J Singh, C Fiolhais, Phys Rev B 1993, 48, 4978.
[31] C Kittel, Introduction to Solid State Physics; John Wiley: New York, 1996; p 23.
[32] H J Monkhorst, J D Pack, Phys Rev B 1976, 13, 5188.
[33] M Methfessel, A T Paxton, Phys Rev B 1989, 40, 3616.
[34] G Makov, M C Payne, Phys Review B 1995, 51, 4014.
[35] G Henkelman, G J ohannesson, H J onsson, In Theoretical Methods in Condensed Phase Chemistry; S Schwartz, Ed.; Springer: Netherlands, 2002; p 269.
[36] G Henkelman, B P Uberuaga, H J onsson, J Chem Phy 2000, 113, 9901.
[37] G Henkelman, H J onsson, J Chem Phys 1999, 111, 7010.
[38] R A Olsen, G J Kroes, G Henkelman, A Arnaldsson, H J onsson,
J Chem Phys 2004, 121, 9776.
[39] J Li, E Croiset, L Ricardez-Sandoval, J Mole Catal A: Chem 2012, 365, 103.
[40] M B Lee, Q Y Yang, S L Tang, S T Ceyer, J Chem Phys 1986, 85, 1693.
[41] J H Larsen, I Chorkendorff, Surf Sci Rep 1999, 35, 163.
[42] S.-C Huang, C.-H Lin, J H Wang, J Phys Chem C 2010, 114, 9826.
[43] C V Ovesen, P Stoltze, J K Nïrskov, C T Campbell, J Catal 1992,
134, 445.
[44] A A Gokhale, J A Dumesic, M Mavrikakis, J Am Chem Soc 2008, 130, 1402.
[45] C V Ovesen, B S Clausen, B S Hammershïi, G Steffensen, T.
Askgaard, I Chorkendorff, J K Nïrskov, P B Rasmussen, P Stoltze, P.
Taylor, J Catal 1996, 158, 170.
[46] L C Grabow, A A Gokhale, S T Evans, J A Dumesic, M Mavrikakis,
J Phys Chem C 2008, 112, 4608.
[47] R C Catapan, A A M Oliveira, Y Chen, D G Vlachos, J Phys Chem C
2012, 116, 20281.
[48] D W Blaylock, Y.-A Zhu, W Green, Top Catal 2011, 54, 828.
Received: 21 January 2014 Revised: 13 July 2014 Accepted: 15 July 2014 Published online on 11 August 2014
...in an attempt to construct microkinetic models for gas -surface reactions Thermodynamic properties of reaction spe-cies were estimated based on ab initio calculations and statis-tical mechanics... tool that integrates quantum mechanical calculations and statistical mechanics to study reactions on surfaces
Case study 2: Water-Gas Shift Reaction on Ni(111)
Water–gas...
for reactions and 4, respectively) This is a demonstration of
the importance of kinetic analysis in order to explain and/or
pre-dict surface reaction pathways This observation