Download as pdf or txt
Download as pdf or txt
You are on page 1of 218

University of South Carolina

Scholar Commons

Theses and Dissertations

Summer 2021

Solvent Effect Modeling in Heterogenous Catalysis


Mehdi Zare

Follow this and additional works at: https://scholarcommons.sc.edu/etd

Part of the Chemical Engineering Commons

Recommended Citation
Zare, M.(2021). Solvent Effect Modeling in Heterogenous Catalysis. (Doctoral dissertation). Retrieved
from https://scholarcommons.sc.edu/etd/6474

This Open Access Dissertation is brought to you by Scholar Commons. It has been accepted for inclusion in
Theses and Dissertations by an authorized administrator of Scholar Commons. For more information, please
contact digres@mailbox.sc.edu.
SOLVENT EFFECT MODELING IN HETEROGENOUS CATALYSIS
by

Mehdi Zare

Bachelor of Science
Shiraz University, 2011

Master of Science
University of Tehran, 2013

Submitted in Partial Fulfillment of the Requirements

For the Degree of Doctor of Philosophy in

Chemical Engineering

College of Engineering and Computing

University of South Carolina

2021

Accepted by:

Andreas Heyden, Major Professor

Donna A. Chen, Committee Member

John R. Monnier, Committee Member

John R. Regalbuto, Committee Member

Melissa Ann Moss, Committee Member

Tracey L. Weldon, Interim Vice Provost and Dean of the Graduate School
© Copyright by Mehdi Zare, 2021
All Rights Reserved.

ii
DEDICATION
To my wife, Mozhdeh Parizad.

“You are the best mercy of God I have ever received. I found love, tranquility and image

of God in you.”

iii
ACKNOWLEDGEMENTS

I wish to express my sincere appreciation to my supervisor, Dr. Andreas Heyden,

whose persistence help, patience, knowledge, intelligence, and encouragement guided me

to do the right thing even when the road got tough.

I wish to thank my committee members, Dr. Chen, Dr. Regalbuto, Dr. Monnier,

and Dr. Moss for their time and help. I would like to offer my special thanks to Dr.

Mohammad Saleheen, and Adam Yonge for their contribution to my thesis with their

knowledge and insightful comments and suggestions. I am thankful to all the Heyden Lab

group members, Dr. Salai Ammal, Dr. Osman Mamun, Dr. Yongjie Xi, Dr. Biplab

Rajbanshi, Dr. Wenqiang Yang, Charles Fricke, Kyung-Eun You, Subrata Kundu, Dia

Sahsah, Nicholas Szaro, and Olajide Bamidele for being great coworkers and friends. I am

grateful to the Chemical Engineering Department staffs, Marcia Rowen, Loretta

Hardcastle, Vernon Dorrell, Brian Loggans, and Shawn Hagan for assisting me with

administrative and technical issues. Finally, I would like to thank my family for their

unconditional love, encouragement, and support in all my endeavors.

iv
ABSTRACT
In recent years, the biorefining industry and biofuels have emerged as a major

American energy sector. Biofuels are fuels produced from plant and animal material, also

referred to as biomass. This includes wood products, manure, and corn, among other

materials. Compared to fossil fuels, biofuels are significantly more environmentally

friendly and thus pose less of a threat to environmental health. In 2019, the United States

consumed around 14.54 billion gallons of ethanol and around 1.81 billion gallons of

biodiesel. By 2030, the United States is expected to consume around 95 Mtoe of biofuels.

In order to meet current demand and expected growth in demand, it is necessary to advance

the efficiency of biomass processing. In this context, special characteristics of biomass

feedstock (highly reactive and water soluble, aqueous, and thermally unstable) demands

liquid-phase processing technologies for higher product selectivity and lower cost. As a

result, to design an efficient liquid-phase process, it is critical to understand the root causes

of solvent effects on surface properties and energetics. However, despite recent

improvements in studying reactions at gas-solid interfaces, methods capable of

investigating reactions occurring at solid-liquid interfaces are less developed; primarily due

to the inherent intricacies of a reaction system comprised of both a complex heterogeneous

catalyst and a condensed phase.

The objectives of this study are to develop and validate a hierarchy of multi-scale

methods for computing reaction and activation free energies of elementary processes

v
occurring at metal-solvent interfaces and to apply these methods to the rational design of

novel heterogeneous catalysts with exceptional activity and selectivity for the liquid-phase

conversion of lignocellulosic biomass into transportation fuels or commodity and specialty

chemicals. To gain a fundamental understanding of the role of metal identity (catalyst) on

the solvent effect, we studied the aqueous-phase effect on the initial C-H and O-H bond

cleavages of ethylene glycol (being a commonly studied surrogate molecule of biomass-

derived polyols) over the (111) facet of six transition metal surfaces (Ni, Pd, Pt, Cu, Ag,

Au) using our explicit solvation method, eSMS. We found a significant metal dependence

on aqueous solvation effects that can be traced back to a different amount of charge-transfer

between the adsorbed species and metals in the reaction and transition states for the

different metal surfaces. In addition, we developed a new hybrid QM/MM approach for

computing solvent effects on the free energy of adsorption and desorption processes that

enables us to compare our work with experimental findings quantitatively. Consequently,

comparison of computational predictions to experimental data of phenol adsorption on the

Pt(111) surface indicates that adsorption free energies in the aqueous phase can be

determined accurately within an error less than 0.10 eV.

vi
TABLE OF CONTENTS

Dedication .......................................................................................................................... iii

Acknowledgements ............................................................................................................ iv

Abstract ................................................................................................................................v

List of Tables ................................................................................................................... viii

List of Figure................................................................................................................... xvii

Chapter 1: Introduction ........................................................................................................1

Chapter 2: Theoretical Investigation of Solvent Effect on the


Hydrodeoxygenation of Propionic Acid over a Ni(111)
Catalyst Model ............................................................................................................5

Chapter 3: Dependency of Solvation effects in Metal Identity in Surface Reactions .......45

Chapter 4: Liquid Phase Adsorption Processes in Heterogenous Catalysis ......................82

Appendix A: Supporting Information for Theoretical


Investigation of Solvent Effect on the
Hydrodeoxygenation of Propionic Acid over a Ni(111)
Catalyst Model ........................................................................................................121

Appendix B: Supporting Information for Dependency of


Solvation effects in Metal Identity in Surface
Reactions ..............................................................................................................142

Appendix C: Supporting Information for Liquid Phase


Adsorption Processes in Heterogeneous Catalysis .................................................154

Appendix D: Copyright Permissions ...............................................................................193

vii
LIST OF TABLES

Table 2.1 Solvent effect on the stability of various adsorbed


species in the HDO of PAC to ethane and ethylene over
a Ni(111) catalyst surface model at a reaction
temperatureof 473 K. ΔΔGrxn indicates the difference
in the adsorption free energy of the corresponding
intermediate in the presence and the absence of the
solvent. Asterisk (*) represents a surface adsorption
site and multiple asterisks are indicative of the number
of occupied active sites. ............................................................................................30

Table 2.2 Reaction and activation free energies of all


elementary steps in the HDO of PAC to ethane and
ethylene over a Ni(111) catalyst surface model at 473
K. ΔΔGrxn and ΔΔGact indicate the reaction and the
activation free energy differences between
corresponding reaction in the presence of solvent and
in the gas phase, respectively ....................................................................................31

Table 2.3 Overall, decarbonylation, and decarboxylation


turnover frequencies as well as important steady state
surface coverages for the HDO of PAC over a Ni(111)
catalyst surface model in the gas phase, and in liquid
water and 1,4-dioxane at 473 K. Note that calculations
for solvents were performed with the help of the
COSMO-RS package with three different Ni cavity
radii: with default value, with a 10% increased value
and a 10% decreased value relative to the default ....................................................33

Table 2.4 Calculated turnover frequency (net rate) in the gas


phase and in the solvents for all elementary reaction
steps in the HDO of PAC over a Ni(111) catalyst
model surface at a temperature of 473 K and a
hydrogen partial pressure of 0.1 bar .........................................................................34

Table 3.1 Free energies of activation of O-H bond cleavage


(CH2OHCH2OH** + * ↔ CH2OCH2OH** + H*) and
C-H bond cleavage (CH2OHCH2OH** + * ↔
CHOHCH2OH ** + H*) of ethylene glycol in gas- and
aqueous-phase environments over the (111) surface

viii
facet of six transition metals at 423 K. QM/MM-FEP
indicates a free energy calculation in water between the
critical points identified by gas-phase calculations
(using the gas-phase vibrational partition function for
the reactant and transition states). QM/MM-MFEP-
OPT represents a free energy calculation in water for
the different cleavages between the respective reactant
and transition states that have been optimized in an
aqueous-phase environment. Here, the vibrational
partition functions are computed in the aqueous phase
assuming the timescale for reorientation of the solvent
molecules is much larger than the timescale for
molecular vibrations. For comparison, implicit
solvation calculations have also been performed using
both nonperiodic (iSMS) and periodic (VASPsol)
approaches. All numbers are in eV ...........................................................................65

Table 3.2 Average aqueous-phase effects on the activation


free energy (∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) as well
as three solvation effect descriptors for the O-H and C-
H bond cleavages of ethylene glycol over the (111)
surface facet of six transition metals at 423 K (see
Table 3.1 for 95% confidence intervals). ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 is
the free energy of activation in vapor phase, H-bond
denotes the change in mean of total hydrogen bonding
(acceptor + donor) going from RS to TS, MC
(molecular charge transfer) represents the change in
the absolute sum of partial charges on the reacting
moiety going from RS to TS, and finally BC (cleaving-
bond charge transfer) is the change in sum of partial
charges on the cleaving bond atoms going from RS to
TS ..............................................................................................................................67

Table 4.1 Aqueous-phase effect on the free energy of low


𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙
coverage adsorption (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 −
𝑔𝑎𝑠
∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) in eV for ethylene glycol over Cu(111)
and Pt(111) at 423 K, for CO on Cu(111) and Pt(111)
at 298 K, and for phenol over Cu(111) and Pt(111) at
298 K. The aqueous phase results are based on the
proposed scheme in this work (QM/MM-FEP) and two
implicit solvation schemes: iSMS and VASPsol. For
the QM/MM-FEP calculations, the 95% confidence
interval (based on limited water sampling and multiple
independent simulations) is also given ...................................................................106

ix
Table 4.2 Contributions to aqueous-phase effect on free
energy (in eV) of low coverage desorption
𝑙𝑖𝑞→𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞
(∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = −∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) of ethylene glycol at
423 K (see Figure 4.2c), of carbon monoxide at 298 K
(see Figure 4.2b), and of phenol at 298 K (see Figure
4.2d) over the (111) facets of Pt and Cu catalysts. All
numbers also contain a 95% confidence interval based
on limited water sampling obtained by performing
multiple independent simulations ...........................................................................107

Table 4.3 Temperature dependence of the aqueous-phase


𝑔𝑎𝑠→𝑙𝑖𝑞
effect on free energy (∆∆𝐺𝑃ℎ ), enthalpy
𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞
(∆∆𝐻𝑃ℎ ), and entropy (∆∆𝑆𝑃ℎ ) of low
coverage phenol adsorption on the Pt(111) catalyst
surface. Experimental data obtained by fitting equation
4.9 (𝑛 = 4 and 𝑐𝑜𝑛𝑠𝑡 = 0.1) to the raw data from
Singh et al. Computational data obtained by the
QM/MM-FEP scheme proposed in this study.
𝑔−𝑙 𝑔−𝑙
∆𝐻𝑎𝑑𝑠,𝑃ℎ and ∆𝑆𝑎𝑑𝑠,𝑃ℎ are the enthalpy and entropy of
phenol adsorption in liquid water from its free state in
the gas phase (𝑃ℎ(𝑔) + 𝑛 ∗ (𝑙𝑖𝑞) ↔ 𝑃ℎ𝑛∗ (𝑙𝑖𝑞)),
𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙 𝑔𝑎𝑠
respectively (∆∆𝑍𝑃ℎ = ∆𝑍𝑃ℎ − ∆𝑍𝑃ℎ , 𝑍 ≡
𝐺, 𝐻, 𝑆). Numbers in bracket [] in the QM/MM-FEP
row were calculated using the experimental gas-phase
enthalpy and entropy of phenol adsorption on the
𝑔𝑎𝑠
Pt(111) catalyst surface (∆𝐻𝑃ℎ = −200 𝑘𝐽⁄𝑚𝑜𝑙 ,
𝑔𝑎𝑠
∆𝑆𝑃ℎ = −14.7𝑅) ...................................................................................................108

Table A.1 Solvent effect on the stability of various adsorbed


species in the propionic acid conversion to ethane and
ethylene over a Ni (111) catalyst surface model at two
different temperatures of 298 K and 473 K based on
two implicit solvation schemes: iSMS and VASPsol.
ΔΔGrxn indicates the difference in the adsorption free
energy of the corresponding intermediates in the
presence and absence of water. Asterisk (*) represents
a surface adsorption site and multiple asterisks are
indicative of the number of occupied active sites ...................................................122

Table A.2 Solvent effect on the stability of transition states


in the propionic acid conversion to ethane and ethylene
over a Ni (111) catalyst surface model at two different
temperatures of 298 K and 473 K based on two implicit
solvation schemes: iSMS and VASPsol. Solvent effect
on TS indicates the difference in the adsorption free

x
energy of the corresponding transition state in the
presence and absence of water ................................................................................123

Table A.3 Solvent effect on the reaction free energies of all


elementary surface reactions of the propionic acid
conversion to ethane and ethylene over a Ni (111)
catalyst surface model at two different temperature of
298 K and 473 K based on two different implicit
solvation schemes: iSMS and VASPsol. ΔΔGrxn
indicates the reaction free energy difference between
corresponding reaction in liquid water and in a gas
phase .......................................................................................................................125

Table A.4 Solvent effect on the activation free energies of


all elementary surface reactions of the propionic acid
conversion to ethane and ethylene over a Ni (111)
catalyst surface model at two different temperature of
298 K and 473 K based on two different implicit
solvation schemes: iSMS and VASPsol. ΔΔGact
indicates the activation free energy differences
between corresponding reaction in liquid water and in
the gas phase ...........................................................................................................127

Table A.5 Overall turnover frequency of


hydrodeoxygenation of propionic acid over a Ni (111)
catalyst surface model at a temperature of 473 K in
vapor phase, and at different partial pressures of
hydrogen and carbon monoxide ..............................................................................129

Table A.6 Overall turnover frequency of


hydrodeoxygenation of propionic acid over a Ni (111)
catalyst surface model at a temperature of 473 K in
vapor phase, and at different partial pressures of
hydrogen and carbon monoxide based on considering
only 1 site for adsorbed CO and analogues change in
counting sites for other adsorbates attached to the
surface through their carbon atom. The number of sites
assigned to each adsorbate in this calculation is
included in Table A.7 ..............................................................................................129

Table A.7 Number of occupied sites used for microkinetic


modeling result displayed in Table A.6 ..................................................................130

Table A.8 Solvent effect on the stability of various adsorbed


species in the propionic acid conversion to ethane and
ethylene over a Ni (111) catalyst surface model at 473

xi
K in the presence of two different solvents, water and
1,4-dioxane, using iSMS methodology. ΔΔGrxn
indicates the difference in the adsorption free energy
of the corresponding intermediates in the presence and
absence of solvent. Asterisk (*) represents a surface
adsorption site and multiple asterisks are indicative of
the number of occupied active sites. Note that
calculations for solvents were performed with the help
of the COSMO-RS package with three different Ni
cavity radii: with default value, with a 10% increased
value and a 10% decreased value relative to the default
.................................................................................................................................131

Table A.9 Solvent effect on the reaction free energies of all


elementary surface reactions of the propionic acid
conversion to ethane and ethylene over a Ni (111)
catalyst surface model at 473 K based on iSMS
methodology in the presence of water and 1,4-dioxane.
ΔΔGrxn indicates the reaction free energy difference
between corresponding reaction in solvent and in a gas
phase. Note that calculations for solvents were
performed with the help of the COSMO-RS package
with three different Ni cavity radii: with default value,
with a 10% increased value and a 10% decreased value
relative to the default ..............................................................................................132

Table A.10 Solvent effect on the activation free energies of


all elementary surface reactions of the propionic acid
conversion to ethane and ethylene over a Ni (111)
catalyst surface model at 473 K based on iSMS
methodology in the presence of water and 1,4-dioxane.
ΔΔGact indicates the activation free energy difference
between corresponding reaction in solvent and in a gas
phase. Note that calculations for solvents were
performed with the help of the COSMO-RS package
with three different Ni cavity radii: with default value,
with a 10% increased value and a 10% decreased value
relative to the default ..............................................................................................134

Table A.11 Overall, decarbonylation, and decarboxylation


turnover frequencies as well as important steady state
surface coverages for the HDO of PAC over a Ni(111)
catalyst surface model in the gas phase and in liquid
water at 473 K. Calculations in water were performed
with the help of the iSMS and VASPsol schemes. Also,
𝑎𝑖 denotes the reaction order with respect to partial

xii
𝑛
pressure/fugacity of component i and 𝑋𝑟𝑐 indicates
Campbell’s degree of rate control for reaction step n .............................................136

Table B.1 Simulation box size used for (111) facet of each
metal surface ...........................................................................................................143

Table B.2 Lennard-Jones parameters of all elements used in


this study. The parameters are based on 12-6 Lennard-
𝜎 12 𝜎 6
Jones potential, 𝑉𝑙𝑗 = 4𝜀 [( 𝑟 ) − ( 𝑟 ) ]. Ow and Hw
represent oxygen and hydrogen of TIP3P water model,
respectively .............................................................................................................143

Table B.3 Average rotational correlation time (in


picosecond) of liquid water molecules residing in a 5
Å radius of an adsorbed ethylene glycol (C2H5OH*)
species on a (111) facet of six transition metal surfaces
studied in the present work. The predicted correlation
functions have been fitted with three exponential
functions to acquire average rotational correlation
times ........................................................................................................................144

Table B.4 Number of hydrogen bonds formed from the


interactions of the TIP3P water model and the reacting
moiety in reactant state (RS) and transition state (TS)
of O-H and C-H bond cleavage of ethylene glycol over
a (111) facet of six transition metal surfaces at 423 K.
Acceptor indicates that OH functional groups of
ethylene glycol accepts a hydrogen bond donated by
the oxygen of water while Donor indicates vice versa.
All numbers are computed based on three independent
simulations and hence possess 95% confidence
intervals (assuming a normal distribution) .............................................................145

Table B.5 Partial charges (in atomic units), computed based


on the NPA charge model, on the reacting moiety
(ethylene glycol) in the reactant (RS) and transition
states (TS) for O-H and C-H bond cleavages over the
(111) facet of six transition metal surfaces at 423 K.
Bold numbers indicate partial charges on the cleaving
bond atoms (O1 and H5 in O-H bond cleavage, and C1
and H4 in C-H bond cleavage) ................................................................................146

Table B.6 Pearson correlation coefficient (PCC) between


descriptors used in this study. It has a value between
+1 and −1, where +1 is total positive linear correlation,
0 is no linear correlation, and −1 is a total negative

xiii
linear correlation. ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 shows the free energy of
activation in vapor phase, H-bond denotes the change
in mean of total hydrogen bonding (acceptor + donor)
going from RS to TS, MC (molecular charge-transfer)
represents the change in the absolute sum of partial
charges on the reacting moiety going from RS to TS,
and finally BC (cleaving-bond charge-transfer)
expresses the change in sum of partial charges on the
cleaving bond going from RS to TS (see Table 3.2 in
the main text for the value of descriptors) ..............................................................148

Table B.7 Data of a linear model, (𝑦 − 𝑦̅) = 𝛼1 (𝑓1 − 𝑓̅1 ) +


𝛼2 (𝑓2 − 𝑓̅2 ), for estimating the solvent effect
(∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ), computed by
eSMS, on O-H and C-H bond cleavages of ethylene
glycol at 423 K. We have not considered the
combination of MC and BC since they are totally
correlated (see Table B.7). The fits are based on 6 data
points for each cleavage which correspond to six
transition metals. Also, 𝑀𝐴𝐸 represents the mean of
absolute errors associated with the fit, in eV, defined
as the difference between estimated and eSMS values ...........................................148

Table B.8 Data of a quadratic model, (𝑦 − 𝑦̅) = 𝛼1 (𝑓1 −


𝑓̅1 ) + 𝛼2 (𝑓2 − 𝑓̅2 ) + 𝛼3 (𝑓1 − 𝑓̅1 )2 + 𝛼4 (𝑓2 − 𝑓̅2 )2 , for
estimating the solvent effect (∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 −
∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ), computed by eSMS, on O-H bond cleavage
of ethylene glycol at 423 K .....................................................................................149

Table B.9 Solvent effects on the free energy of activation


(∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) of O-H and C-H
bond cleavage of ethylene glycol on the (111) facet of
six transition metal surfaces using the iSMS solvation
scheme. Note that calculations were performed with
the help of the COSMO-RS package with three
different metal cavity radii: with default value (2.2230
Å for all transition metal elements), with a 10%
increased value (2.4453 Å) and a 10% decreased
(2.0007 Å) value relative to the default ..................................................................150

Table B.10 Average fraction of TIP3P water molecules with


different orientations (see Figure B.3 for explanation
of the different orientations) within the first water layer
of the surface (see first peak in height distribution
function of the water O atom in Figure B.4). The data
indicate that water orients itself similar across the

xiv
metal surfaces and ~68% of them are in “parallel”
orientation, ~23% in H-up orientation, and only 9% in
H-down orientation (which lets the water hardly form
hydrogen bonds with the condensed phase)............................................................151

Table C.1 Lennard-Jones parameters of all elements in this


12 6
𝜎 𝜎
study based on 𝑉𝑖𝑗 = 4𝜀𝑖𝑗 [( 𝑟 𝑖𝑗 ) − ( 𝑟𝑖𝑗 ) ], where 𝜀
𝑖𝑗 𝑖𝑗

is the depth of the potential well and 𝜎 is the finite


distance at which the inter-particle potential is zero.
Ow and Hw represent oxygen and hydrogen of TIP3P
water model, respectively. ‘Ho’ in the phenol molecule
indicates the hydrogen of the OH functional group ................................................161

Table C.2 Average rotational correlation time (RCT) and


constants (in pico-second) of 3-Exponential fits
𝑡 𝑡 𝑡
−( ) −( ) −( )
(𝐹(𝑡) = 𝑎1 𝑒 + 𝑎2 𝑒 𝜏1 + (1 − 𝑎1 − 𝑎2 )𝑒 𝜏2 ) 𝜏3

in Figure C.1. The average rotational correlation time


𝑎1 𝜏1 2 +𝑎2 𝜏2 2 +(1−𝑎1 −𝑎2 )𝜏3 2
is computed as: 𝑅𝐶𝑇 = ...................................................162
𝑎1 𝜏1 +𝑎2 𝜏2 +(1−𝑎1 −𝑎2 )𝜏3

Table C.3 Charges of the QM system (in e) in the initial


state, where there exist electrostatic and van der Waals
interactions between the solvent and the adsorbate, and
in the final state, where all interactions have vanished
as if the adsorbate was removed completely, for
different adsorbates investigated in this work. The
initial and final states correspond to states I and V in
Figure C.1 in the main text, respectively. a) Reactant
and transition states of O-H bond cleavage of EG on a
Cu(111) surface model using the NPA charge model,
b) CO on a Pt(111) surface model using the NPA and
DDEC6 charge models, c) ethylene glycol (EG) on the
Pt(111) and Cu(111) surfaces using the NPA charge
model, d) CO on Pt(111) and Cu(111) surfaces using
the DDEC6 charge model, and e) phenol on the Pt(111)
and Cu(111) surfaces using the DDEC6 charge model
(Ho in the phenol molecule indicates the hydrogen of
the OH functional group) ........................................................................................169

Table C.4 Values for the parameter 𝛿 in 𝑉𝑖𝑗 = (1 −


𝐴 𝐵
𝜆) [(𝑟 2 +𝛿𝜆)6
− (𝑟 2 +𝛿𝜆)3
] for interactions used in this
𝑖𝑗 𝑖𝑗
study. As recommended by Zacharias et al., the
parameter δ was chosen as the square of the vdW
radius of the interacting atoms to allow for a smooth

xv
transition between an atom present on the surface and
filling the cavity after the molecule is annihilated. Ow
and Hw represent the oxygen and hydrogen of the
TIP3P water model, respectively. Ho in the phenol
molecule indicates the hydrogen of the OH functional
group .......................................................................................................................177

Table C.5 Contributions to the aqueous-phase effect on the


free energy of the low coverage desorption
𝑙𝑖𝑞→𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞
(∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = −∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) of the reactant and
transition states of the O-H bond cleavage of ethylene
glycol over a Cu(111) surface at 423 K (see Figure 4.2a
𝑔𝑎𝑠→𝑙𝑖𝑞
in the main text). ∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 also shows the overall
aqueous-phase effect on the free energy of the low
coverage adsorption of the states computed with
QM/MM-FEP scheme proposed in this study, as well
as two well-known implicit solvation schemes: iSMS
and VASPsol. For the QM/MM-FEP calculations, the
95% confidence interval (based on limited water
sampling) is given, assuming a normal distribution. All
numbers are in eV ...................................................................................................179

Table C.6 Net charge on adsorbate and number of sites


occupied by adsorbate (#Site) for all adsorbates studied
in this work. These two descriptors were then used in
𝑦−𝑦̅ ̅̅̅
𝑓1 −𝑓 1 ̅̅̅
𝑓2 −𝑓 2
a linear model, = 𝛼1 + 𝛼2 (𝑠 is standard
𝑠𝑦 𝑠𝑓1 𝑠𝑓2
deviation and bar sign shows mean), as 𝑓1 and 𝑓2
respectively, for estimating aqueous-phase effects on
𝑔𝑎𝑠→𝑙𝑖𝑞
the free energies of adsorption (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 =
𝑔−𝑙 𝑔𝑎𝑠
∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 − ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ), computed by the
proposed scheme in this work (QM/MM-FEP). 𝛼1 and
𝛼2 are 0.45 and 0.62, respectively, which indicates
almost equal importance of net charge and the size of
the adsorbate ...........................................................................................................180

xvi
LIST OF FIGURES

Figure 2.1 Reaction network for the HDO of PAC to ethane


and ethylene. Blue solid arrows belong to the
decarbonylation pathway, green dash-dotted arrows
depict the decarboxylation pathway, and yellow
dashed arrows show those steps that are in common
between the decarboxylation and the decarbonylation
mechanism. In addition, the number on each arrow
illustrates the corresponding elementary reaction in
Table 2.2, and the dashed circles show the
corresponding cleavage of fluid phase propanoic acid ............................................36

Figure 2.2 Schematic representation of TOF (s−1 ), of


various elementary reactions involved in the dominant
pathways. The left dashed box illustrates the dominant
pathway in the gas phase while the right blue drop
displays the dominant pathways in the solvent
environments. Numbers in parentheses () depict
reaction step numbers. The numbers in square brackets
[] in the right blue drop are the TOFs (s−1 ) in liquid
water and the numbers without square brackets are the
TOFs (s −1 ) in liquid 1,4-dioxane..............................................................................37

Figure 2.3 Apparent activation energy plot in the


temperature range of 473 to 623 K in gas and
condensed phase environments. The apparent
activation barrier is 2.38 eV in the gas phase, 2.71 eV
in liquid water, and 2.44 eV in liquid 1,4-dioxane ...................................................38

Figure 3.1 Aqueous-phase effects on the activation free


energy barrier of the O-H and C-H bond cleavages of
ethylene glycol over the (111) surface facet of six
transition metals at 423 K computed by eSMS (all
structures optimized in liquid water; see Table 3.1 for
specific numbers) ......................................................................................................68

Figure 3.2 Free-energy profiles for O-H bond cleavage of


ethylene glycol in vapor and aqueous phase over the
(111) surface facets of six transition metals at 423 K
without considering vibrational contributions to the

xvii
partition functions. See Table 3.1 for corresponding
data that include vibrational contributions. The number
of intermediate states between the reactant and
transition states for eSMS calculations is determined
by our desire to have an energy difference between
intermediates smaller than twice the thermal energy (<
2𝑘𝐵 𝑇). The aqueous-phase profiles portray the average
of three or more independent eSMS calculations
possessing 95% confidence intervals smaller than
±0.05 eV (see Table 3.1). The analogous plot for C-H
bond cleavage reaction is provided in the Appendix B ............................................69

Figure 3.3 Parity plot of the aqueous-phase effect on the free


energy of activation (∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 )
computed by eSMS (QM/MM-FEP in Table 3.1)
versus the model-predicted aqueous-phase effect. The
predicted values are calculated using the gas-phase free
energy of activation (∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) and cleaving-bond
charge transfer (BC) as two descriptors in (a) a linear
model (equation 3.1) for C-H bond cleavage and (b) a
quadratic model (equation 3.2) for O-H bond cleavage
of ethylene glycol over (111) surface facets of six
transition metals at 423 K. The average aqueous-phase
̅̅̅̅̅̅̅̅̅
effect, ∆∆𝐺 𝑎𝑐𝑡 , computed by eSMS, is -0.33 eV for O-

H bond cleavage and -0.25 eV for C-H bond cleavage.


The mean of absolute errors (𝑀𝐴𝐸) between eSMS
and model-predicted values is 0.03 eV in the linear
model (a) and 0.02 eV in the quadratic model (b). The
corresponding model parameters are listed in Tables
B.7 and B.8 ...............................................................................................................70

Figure 4.1 Free energy perturbation approach for computing


the difference in the adsorption free energy of any
adsorbed species in the presence of liquid water and in
a gas phase. States Ia and IIIa are “non-physical”
states. Between states I and II, the (classical)
electrostatic interaction between the solvent and the
adsorbate is slowly removed. Also, the coordinates of
the atoms described quantum chemically transition
from those in liquid to those in the vapor phase. From
state II to III, the non-classical contributions of water
molecules changing the electronic structure of the
quantum system transition from those of a metal-
adsorbate system to those without adsorbate, i.e., this
energy difference constitutes the main difference
between electrostatic and mechanical embedding of

xviii
the QM/MM system. In other words “gree circle”
symbolizes the adsorbate being only present in the
classical DL_POLY simulation and not in the quantum
system. Van der Waals interactions between the
solvent and the adsorbate are slowly removed between
state III and IV. The gas-phase desorption is described
between state IV and V. The (free) energy difference
between state IV and I is the solvent effect on the
stability of the adsorbate .........................................................................................109

Figure 4.2 (Average) free-energy profiles for aqueous-phase


𝑙𝑖𝑞→𝑔𝑎𝑠
effects on the desorption (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 =
𝑔𝑎𝑠→𝑙𝑖𝑞
−∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) of a) the reactant and transition states
of O-H bond cleavage of ethylene glycol over a
Cu(111) surface at 423 K, b) the CO molecule over the
Cu(111) and Pt(111) surfaces at 298 K, c) the ethylene
glycol molecule over Cu(111) and Pt(111) surfaces at
423 K, and d) the phenol molecule over the Cu(111)
𝑙𝑖𝑞→𝑔𝑎𝑠
and Pt(111) surfaces at 298 K. ∆∆𝐺 (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) is
the aqueous-phase effect on the low coverage
desorption of an adsorbate. “To OPT adsorbate in
Liquid” represents the FEP steps involved in the
transition of the metal-adsorbate QM system in gas to
that in liquid, “To OPT site in Gas” represents the FEP
steps involved in changing the coordinates of the metal
cluster atoms with dummy adsorbates that have zero
interaction with the solvent to those of the optimized
metal cluster in the gas phase, and “II-to-III” shows the
non-classical contribution of water molecules
changing the electronic structure of the quantum
system transitioning from those of a metal-adsorbate
system to those without adsorbate (see Figure 4.1,
states II and III). “Remove Electrostatic” and “Remove
Lennard-Jones” represent the removal of classical
electrostatic and Lennard-Jones interactions between
an adsorbate and the water molecules, respectively. For
the magnitude of each contribution, see Tables 4.2 and
C.5 ...........................................................................................................................110

Figure 4.3 Fits of equation 4.9 (adsorption isotherm) to the


experimental data of the aqueous-phase phenol
adsorption on a Pt(111) catalyst surface at 298 K from
Singh et al. (see Figures C.7 to C.9 for similar plots at
different temperatures). The “Const” parameter
represents the hydrogen surface coverage relative to
the free site coverage (const=0 shows zero H coverage

xix
approximation), 𝑛 represents the number of sites a
phenol molecule occupies when adsorbed on the
𝑙𝑖𝑞
surface, 𝛼 (𝛼𝑃ℎ,𝑃ℎ ) is a constant associated to the
phenol-phenol lateral interactions on the surface in
𝑔𝑎𝑠→𝑙𝑖𝑞
water, and ∆∆𝐺 (∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0)) is the
aqueous-phase effect on the low coverage phenol
adsorption on the surface. SSR is the sum of squared
residuals, which shows the error associated with each
fit .............................................................................................................................111

Figure 4.4 Temperature dependence of the aqueous-phase


effect on the free energy of low coverage adsorption of
phenol and carbon monoxide over a Pt(111) surface
𝑔𝑎𝑠→𝑙𝑖𝑞
(∆∆𝐺 = ∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 (𝜃𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = 0)). The error
bars indicate 95% confidence intervals based on
limited water sampling and multiple (three)
independent simulations..........................................................................................112

Figure A.1 Adsorption of different sized molecules (CO,


CH3CH2, and CH3CH2C) on a Ni (111) surface model
with various cluster sizes used in iSMS calculations in
liquid water at 473 K. We employed a 3x2 cluster
model for all surface reactions since the computational
cost for larger models increases dramatically while the
solvent effect does not change significantly. For
adsorption/desorption processes of saturated
molecules such as C2H4, C3H6, C2H2,
CH3CH2COOH, CO, CO2, H, H2O, where error
cancellation of solvation effects is smaller than for
surface reactions, a 3x4 model with 86 metal atoms
was used ..................................................................................................................137

Figure A.2 Lateral interaction correction terms for various


intermediates using a 1/4 ML coverage of CO and H.
Shown are the correction terms of the lateral
interaction effect of H (a) and CO (b) for each
individual adsorbed species ....................................................................................138

Figure A.3 Lateral interaction correction terms for the


transition state of each elementary step using 1/4 ML
coverages of CO and H. Shown are the correction
terms of the lateral interaction effect of H (a) and CO
(b) on the transition state of each elementary step ..................................................139

Figure A.4 Partial charges computed using the DDEC6


charge model on adsorbed CO (top left), adsorbed PAC

xx
(top right), and coadsorbed CO and PAC on a Ni (111)
catalyst surface model. In the coadsorbed state, CO is
a Lewis acid, which accepts electrons, while PAC is a
Lewis base, which donates electrons. This coadsorbed
Lewis acid and base pair explains the attractive
interaction between adsorbed CO and PAC............................................................140

Figure A.5 Differential zero-point energy corrected (ZPC)


energy of adding a) an extra CO as a function of
number of CO already adsorbed on a Ni(111) surface
containing 12 surface atoms with (3 × 2√3)
periodicity and b) an extra H as a function of number
of H already adsorbed on the same Ni surface. Adding
the fifth CO increases the energy significantly (by over
0.45 eV) which we attribute to the fact that the surface
is essentially fully covered with 4 CO and adding an
extra CO weakens the bond strength of adsorbed CO
molecules on the surface due to a strong repulsive
interaction between them. On the other hand, the
differential ZPC barely increases when adding an extra
hydrogen atom. Consequently, it can be concluded that
a CO molecule occupies 3 Ni sites on the surface while
H adsorbs on only 1 Ni site .....................................................................................141

Figure B.1 Free-energy profiles for C-H bond cleavage of


ethylene glycol in vapor and aqueous phases over the
(111) facet of six transition metal surfaces at 423 K
without considering vibrational contributions to the
partition function. See Table 3.1 for corresponding
data that include vibrational contributions. The
aqueous phase profile portrays the average of three or
more independent eSMS calculations possessing 95%
confidence intervals smaller than ±0.05 eV. The
analogous plot for O-H bond cleavage is provided in
the main text ............................................................................................................152

Figure B.2 Graphical representation of a geometric


hydrogen bond definition used in this study. In the
picture shown, water is the donor of hydrogen bonding,
and ethylene glycol is the acceptor of hydrogen
bonding ...................................................................................................................153

Figure B.3 Graphical representation of criterion used to


distinguish different water orientations at the surface.
A water molecule in the first layer next to the surface
is considered in H-up conformation if the difference in

xxi
z-coordinate of the H and O atom is larger than half the
H-H distance of the water molecule; this corresponds
for TIP3P water to an angle q show above larger than
52.26⁰. Water molecules that are not considered H-up
or H-down are labeled “parallel” ............................................................................154

Figure B.4 Height distribution function of water O and H


over (111) facet of six transition metal surface ......................................................155

Figure C.1 Rotational correlation time function of liquid


water molecules residing in a 5 Å radius of adsorbed
ethylene glycol (C2H6O2*) species on Cu(111) and
Pt(111) at 423K, of adsorbed carbon monoxide (CO*)
on Cu(111) and Pt(111) at 298 K, and of adsorbed
phenol (C6H5OH*) on Cu(111) and Pt(111) at 298 K.
The correlation time functions have been fitted using
three exponential functions to obtain the average
rotational correlation times (see Table C.2) ............................................................163

Figure C.2 Scaling of the Lennard-Jones potential


interaction between the oxygen of the CO molecule
and the oxygen of the TIP3P water model, O-Ow. 𝜆 =
0 indicates the regular LJ potential, increasing 𝜆
gradually decreases the interaction, and finally, 𝜆 = 1
leads to a complete removal of the LJ potential as if
there were no adsorbate on the surface ...................................................................178

Figure C.3 Average (of three independent simulations) free


energy profile of the desorption of a CO molecule from
a Pt(111) surface model at 298 K using the natural
population analysis (NPA) charge model. ∆∆𝐺
𝑙𝑖𝑞→𝑔𝑎𝑠
(∆∆𝐺𝐶𝑂 ) is the aqueous-phase effect on the low
coverage desorption free energy of CO ..................................................................181

Figure C.4 Vapor-phase optimized geometries of adsorbed


species studied in this work: a) ethylene glycol (EG)
over a Cu(111) surface model, b) transition state of O-
H bond cleavage of EG over the Cu(111) surface
model, c) EG over a Pt(111) surface model, d) CO over
the Cu(111) surface model, e) CO over the Pt(111)
surface model, f) phenol over the Cu(111) surface
model, and g) phenol over the Pt(111) surface model ............................................182

Figure C.5 Fits of equation 4.9 in the main text (adsorption


isotherm) to the experimental data of aqueous-phase
adsorption of phenol at 298 K on a Pt(111) catalyst

xxii
surface from Singh et al. for two different regions: low
and high phenol concentrations. The “Const”
parameter represents the hydrogen surface coverage
relative to the free site coverage (const=0 shows zero
H coverage approximation), 𝑛 represents the number
of sites a phenol molecule occupies when adsorbed on
𝑙𝑖𝑞
the surface, 𝛼 (𝛼𝑃ℎ,𝑃ℎ ) is a constant associated to
phenol-phenol lateral interactions on the surface in
𝑔𝑎𝑠→𝑙𝑖𝑞
water, and ∆∆𝐺 (∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0)) is the solvent
effect on the aqueous-phase low coverage phenol
adsorption on the surface. SSR is the sum of squared
residuals, which shows the error associated with each
fit. Fits with higher 𝑛 predict the lower concentration
region better, while fits with smaller n predict the
higher concentration region better. However, the fits
𝑔𝑎𝑠→𝑙𝑖𝑞
suggest that ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) is quite
independent of these two fit regions .......................................................................183

Figure C.6 Adsorbed phenol orientations at different


coverages in the gas phase. At low coverage, phenols
adsorb horizontally parallel to the surface, while at
high coverage, they adsorb slanted to occupy less space
on the surface ..........................................................................................................183

Figure C.7 Fits of equation 4.9 in the main text (adsorption


isotherm) to the experimental data of aqueous-phase
phenol adsorption on a Pt(111) catalyst surface at 283
K from Singh et al. The “Const” parameter represents
the hydrogen surface coverage relative to the free site
coverage (const=0 shows the zero H coverage
approximation), 𝑛 represents the number of sites a
phenol molecule occupies when adsorbed on the
𝑙𝑖𝑞
surface, 𝛼 (𝛼𝑃ℎ,𝑃ℎ ) is a constant associated to
phenol/phenol lateral interactions on the surface in
𝑔𝑎𝑠→𝑙𝑖𝑞
water, and ∆∆𝐺 (∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0)) is the
aqueous-phase effect on the low coverage phenol
adsorption free energy on the surface. SSR is the sum
of squared residuals, which shows the error associated
with each fit.............................................................................................................184

Figure C.8 Fits of equation 4.9 in the main text (adsorption


isotherm) to the experimental data of aqueous-phase
phenol adsorption on a Pt(111) catalyst surface at 288
K from Singh et al. See the caption of Figure C.7 for
detailed explanations ...............................................................................................185

xxiii
Figure C.9 Fits of equation 4.9 in the main text (adsorption
isotherm) to the experimental data of aqueous-phase
phenol adsorption on a Pt(111) catalyst surface at 314
K from Singh et al.62 See the caption of Figure C.7 for
detailed explanations ...............................................................................................186

xxiv
CHAPTER 1

INTRODUCTION
The critical role of solvents in catalytic transformations occurring at a solid-liquid

interfaces has given rise to an extensive range of studies aimed at understanding and

predicting the role of solvents. However, the inherent complexity of a reaction containing

both a complex heterogenous catalyst and a condensed phase has resulted in only a few

systematic experimental studies. In this context, theoretical calculations have the advantage

of being able to systematically investigate the root causes of such effect. There are

generally two major types of techniques to capture solvent effects computationally: implicit

(continuum) and explicit solvation methods. Implicit methods have the advantage of

computing free energies of reactions rapidly, but their reliability has often been questioned

because of their inability to capture the anisotropic site-specific interactions between the

solute and the solvent molecules. On the other hand, describing such system quantum

mechanically is limited to a few hundred atoms and a time scale of tens or a few hundred

picoseconds due to high computational cost. Finally, a compromise in efficiency and

accuracy a combined quantum mechanical/molecular mechanical (QM/MM) approach.

The objective of my research is to reveal how a solvent can affect the energetics of a

catalytic reaction using implicit and QM/MM solvation approaches and to test the accuracy

of these models against experimental studies.

1
This dissertation has been written following the manuscript style formatting,

meaning each chapter of this dissertation is an independent scientific publication. In my

first publication (chapter 2), which was published in July 2020 in The Journal of Physical

Chemistry C, we investigated hydrodeoxygenation (HDO) of propionic acid (PAC) over a

Ni(111) catalyst surface model in both vacuum and solvent environments with the help of

periodic DFT calculations coupled with our implicit solvation scheme for metal surfaces

(iSMS). We considered decarbonylation (DCN) and decarboxylation (DCX) in the reaction

network. Next, a microkinetic reaction model (MKM) was developed that permits

quantification of solvent effects on reaction kinetics. In this study, we focused on water

and 1,4-dioxane as solvents that are typically used protic and aprotic polar solvents. We

found that under all conditions the DCN is favored over the DCX. The dominant pathway

in gas phase starts with the removal of the hydroxyl group (OH) of PAC, followed by two

dehydrogenation and one hydrogenation steps leading to ethylene production. However,

the dominant pathway in the liquid phases shifts after first step (removal of OH group, i.e.,

C-OH bond cleavage of PAC). It proceeds with a dehydrogenation step to form a CH3CCO

surface intermediate, followed by decarbonylation to produce CH3C, and subsequent

hydrogenation and one last dehydrogenation to produce ethylene. In addition, from a

sensitivity analysis we found C-OH bond cleavage of PAC is the rate controlling step in

gas phase, while its dominance will be replaced with reaction step 9, where 𝛼-carbon of

CH3CHCO undergoes dehydrogenation. Finally, our computational results indicate that the

(111) facet of Ni is likely not the active facet for the HDO of propionic acid in both phases;

hence, other facets need to be investigated to identify the experimentally relevant active

site.

2
Chapter 3 of my dissertation contains my second publication, which has been

published recently (December 2020) in the Communications Chemistry journal. In this

study, we investigated the aqueous-phase effects on the initial C–H and O–H bond

cleavages of ethylene glycol (EG) over the (111) facet of six transition metal surfaces (Ni,

Pd, Pt, Cu, Ag, Au) using our explicit solvation method, eSMS, to gain an understanding

on the dependency of solvation effects on metal identity in surface reactions. We found

very different aqueous-phase effect on the activation free energy barriers across different

metal, indicating that our hypothesis that the nature of metal plays a key role for solvation

effects on surface reactions has been confirmed. Next, to understand the origin of this

dependency, we correlated solvent effect with some intuitive, physics-based descriptors,

including the difference in charge-transfer effects in the reactant and transition states, the

change of hydrogen bonding going from reactant to transition state, and the gas-phase free

energy of activation. The correlation results suggest that the aqueous-phase effect on the

free energy of activation of O–H and C–H bond cleavages of EG over the investigated

metal surfaces originates primarily from some form of a charge-transfer effect (more

charge transfer in the TS relative to the RS leads to more stabilization) and partly on another

descriptor that could be related to hydrogen bonding and gas-phase activation barrier.

Finally, we compared the eSMS result against two implicit solvation models, iSMS and

VASPsol and found that the implicit solvation models failed to capture the full solvent

stabilization during the O–H bond cleavage of EG. For the C–H bond cleavage, implicit

and explicit solvation models anticipate comparable aqueous-phase effects for the metals

that display a smaller charge transfer effect (Pd, Pt, Au).

3
In my third publication in chapter 4, we developed a hybrid QM/MM methodology

enabling quantification of solvent effect in adsorption/desorption processes. This enables

direct comparison of our computation with experimental studies. We first numerically

verified the applicability and precision of this scheme by computing aqueous-phase effect

‡ ‡ ‡
on the free energy activation barrier (∆∆𝐺𝑙𝑖𝑞 = ∆𝐺𝑙𝑖𝑞 − ∆𝐺𝑔𝑎𝑠 ) of O-H bond cleavage of

ethylene glycol (EG) over a Cu(111) surface at 423 K with two different approaches, eSMS

‡,𝑒𝑆𝑀𝑆
methodology for surface reactions (∆∆𝐺𝑙𝑖𝑞 = −0.56 ± 0.04 𝑒𝑉) and the adsorption


scheme proposed here (∆∆𝐺𝑙𝑖𝑞 = −0.50 ± 0.06 𝑒𝑉). We then applied this scheme to

compute solvent effect on adsorption of EG, carbon monoxide (CO), and phenol over (111)

facet of Cu and Pt. Our computation indicates that the aqueous phase has destabilized all

the three adsorbed species and has impacted phenol the most while CO the least

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞


irrespectively of metal surface (∆∆𝐺𝑃ℎ > ∆∆𝐺𝐸𝐺 > ∆∆𝐺𝐶𝑂 ). This is

attributed to a combination of various types of adsorbate-water interactions including

electrostatic, van der Waals, hydrogen bonding, and size of the adsorbate. Finally, to

confirm the accuracy of this proposed scheme, we computed the solvent effect on

𝑔𝑎𝑠→𝑙𝑖𝑞
adsorption free energy of phenol at low coverage (∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0)) over Pt(111) at

298 K to be 1.48 ± 0.04 eV. We found this to be in an excellent agreement with an

experimental study of phenol adsorption on Pt (111) at 298 K within an error of less than

0.10 eV.

4
CHAPTER 2

THEORETICAL INVESTIGATION OF SOLVENT EFFECTS ON THE

HYDRODEOXYGENATION OF PROPIONIC ACID OVER A NI(111)

CATALYST MODEL

Zare, M.; Solomon R. V.; Yang, W.; Yonge, A.; Heyden, A. The Journal of Physical
Chemistry C, 2020, 124, 16488-16500.

Reprinted here with the permission of the publisher

5
2.1 Abstract
The effect of two solvents, liquid water and 1,4-dioxane, has been studied from first

principles on the hydrodeoxygenation of propionic acid over a Ni (111) catalyst surface

model. A mean-field microkinetic model was developed to investigate these effects at a

temperature of 473 K. Under all reaction conditions, a decarbonylation mechanism is

favored significantly over a decarboxylation pathway. Although no significant solvent

effects were observed on the decarbonylation rate, a substantial solvent stabilization of two

key surface intermediates in the decarboxylation mechanism, CH3CCOO and

CH3CHCOO, lead to a notable increase of the decarboxylation rate by two orders of

magnitude in liquid water and by one order of magnitude in liquid 1,4-dioxane.

Furthermore, a significant solvent stabilization of the transition state of C-H bond cleavage

of the 𝛼-carbon of CH3CHCO, relative to the stabilization of the C-C bond cleavage of the

𝛼-carbon of CH3CHCO, leads to a change in dominant pathway in the liquid phase

environments. Finally, a sensitivity analysis shows that the C-OH bond cleavage of

propionic acid and C-C bond cleavage of the 𝛼-carbon of CH3CHCO are the most rate

controlling states in the gas phase. In contrast, in solvents the dehydrogenation of

CH3CHCO becomes the most influential step. This shift in rate controlling state is

attributed to the solvent effect on the dehydrogenation of CH3CHCO, which is facilitated

in aqueous phase. Overall, it is likely that the investigated (111) facet of Ni is not active

for the hydrodeoxygenation of propionic acid in neither the gas nor liquid phase and other

Ni facets or phases must be responsible for the experimentally observed kinetics.

6
2.2 Introduction

A consistent increase in fossil fuel demand and global warming concerns

originating from the release of carbon dioxide during fossil fuel combustion have drawn

increased attention toward utilization of alternative fuel sources and technologies such as

the conversion of biomass to fuels.1 Hydrodeoxygenation (HDO) is one of the more

promising routes for upgrading pyrolysis bio-oils for producing liquid transportation fuels.2

The hydrodeoxygenation of organic acids, present in e.g. pyrolysis oils, is often found to

be a rate controlling HDO process such that there is a desire to design supported transition

metal catalysts that can efficiently convert the organic acids into alcohols and alkanes. The

HDO catalysts must have good activity, be of low cost, and reasonably stable against coke

formation. Some studies have shown Ni-based catalysts to be promising candidates for the

HDO of bio-oils to fuels.3-6 For example, Bykova et al.3 tested a series of Ni-based catalysts

with different stabilizing components for the HDO of guaiacol as a bio-oil model

compound. They found Ni-based catalysts prepared by the sol-gel method and stabilized

with SiO2 and ZrO2 are highly active, and the high activity of these catalysts correlates

with nickel loading. In another study, Yin et al.5 used a bimetallic Ni-Cu catalyst with high

Ni loading (up to 50%) for catalytic hydrotreatment of fast pyrolysis liquids. Their results

show a low rate of undesired gas and coke/char formation when using a high Ni loading.

Similarly, in Ardiyanti et al.’s work4 the catalyst with the highest Ni loading (58 wt % Ni)

promoted with Pd (0.7 wt %) was the most active, yielding oil products with improved

properties such as low oxygen content and a lower tendency for coke formation.

Another factor playing an important role in the HDO of organic acids is the solvent

environment. Experimental studies use various solvents, yet how and why the solvent alters

7
the reaction mechanism remains unknown. For example, Hoelderich et al.7 reported that in

the catalytic deoxygenation of oleic acid (C18) over Pd/C, the presence of water can change

the selectivity towards C17 hydrocarbons by up to 20%. Also, Wan et al.8 discovered that

when water is replaced with n-heptane at otherwise similar conditions, the esterification

reaction is favored over ethanol reformation/hydrogenolysis, resulting in substantial

formation of ethyl acetate. Although the solvent effect in a heterogenous catalyst can

qualitatively be explained by the impact of solvent polarity, more work remains to be done

to explain this important effect quantitatively. This effect can be quantified

computationally using various solvation schemes such as implicit, explicit, or hybrid

explicit and implicit solvation models9-12.

In the present study, periodic DFT calculations coupled with our implicit solvation

scheme for metal surfaces (iSMS)13 are used for the investigation of the HDO of propionic

acid (PAC) over a Ni(111) catalyst surface model in both vacuum and solvent

environments. Decarboxylation (DCX) and decarbonylation (DCN) mechanisms are

considered in the reaction network bases on our prior studies of the HDO of PAC on various

transition metal surfaces.14-18 Figure 2.1 displays a schematic representation of all

elementary reactions investigated in this study. Next, a microkinetic reaction model was

developed under various reaction conditions that permits quantification of solvent effects

on turnover frequency (TOF), abundant surface intermediates, reaction orders, and

apparent activation barriers. This study specifically focuses on water and 1,4-dioxane as

solvents that are typically used protic and aprotic polar solvents. Finally, it is noted that the

lateral interaction effect of high surface coverage species such as adsorbed hydrogen and

8
CO are considered explicitly in our microkinetic models for all surface intermediates and

transition states.

2.3 Methods

2.3.1 DFT Calculations

All calculations for the catalyst slab model were performed using the DFT

implantation in the Vienna Ab Initio Simulation Package (VASP)19-20 code. The periodic

surface model was constructed using the optimized lattice constant of fcc-Ni bulk, 3.518

Å, and consists of 4 Ni layers separated by 15 Å to minimize interactions between the slab

and its periodic images. Each Ni layer contained 12 Ni atoms with (3 × 2√3) periodicity,

the bottom two layers were fixed, while the top two layers, as well as the adsorbates, were

relaxed.

Spin-polarized calculations have been carried out at the 4 × 4 × 1 Monkhorst-Pack

grid using Methfessel-Paxton smearing (σ=0.2 eV) with an energy cut off for plane waves

of 400 eV and a convergence criterion of a self-consistent field (SCF) of 1.0 × 10-7 eV.

The projector-augmented wave method (PAW) was utilized to describe the electron–ion

interactions along with the GGA-PW9121-22 functional for describing exchange and

correlation effects in the energy calculations. Transition state (TS) structures are optimized

using the NEB23 and Dimer methods.24-25 We note that all reported adsorption energies,

activation barriers, and reaction energies have been zero-point corrected.

Cluster model DFT calculations in vacuum were preformed using the

TURBOMOLE 7.0 program package.26-28 The cluster models were obtained by removal of

the periodic boundary conditions from the periodic slabs that were constructed from

9
previous plane-wave (VASP) calculations. A cluster model consisting of two layers and 51

Ni atoms was initially used for all surface reactions. For adsorption reactions that are

somewhat more sensitive to the size of the cluster, a 3-layer model containing 86 Ni atoms

was used. A convergence test of the size of the cluster is presented in the Appendix A

(Figure A.1). All adsorbates and metal atoms were represented by all-electron TZVP29-31

basis sets. The Coulomb potential was approximated with the RI-J approximation using

auxiliary basis sets.32-34 Single point energy calculations were performed with a self-

consistent field (SCF) energy convergence criterion of 1.0 × 10-7 Hartree. Energy

calculations for different spin states were carried out to identify the lowest energy spin

state. Afterwards, COSMO calculations for the cluster models in the liquid phase were

performed on the lowest energy spin state configuration at the same level of theory. The

dielectric constant was set to infinity as required for the Conductor-like Screening Model

for Real Solvents (COSMO-RS)35-36 calculations. Default cavity radii as well as 10%

increased and decreased cavity radii were used for the Ni atoms. We note that the cavity

radius is the most important solvent parameter in COSMO-RS that can, however, not be

accurately determined for Ni due to insufficient experimental data, explaining why are

varied the Ni cavity radius.

2.3.2 Solvation scheme

Solvent effects presented in this work were investigated using the implicit solvation

model for solid surfaces (iSMS) method.13 The Gibbs free energy of an intermediate

𝑙𝑖𝑞𝑢𝑖𝑑
adsorbed on the surface in a liquid phase environment, 𝐺𝑠𝑢𝑟𝑓𝑎𝑐𝑒+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 , is calculated

as:

10
𝑙𝑖𝑞𝑢𝑖𝑑 𝑣𝑎𝑐𝑢𝑢𝑚 𝑙𝑖𝑞𝑢𝑖𝑑
𝐺𝑠𝑢𝑟𝑓𝑎𝑐𝑒+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 = 𝐺𝑠𝑢𝑟𝑓𝑎𝑐𝑒+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 + (𝐺𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 −
𝑣𝑎𝑐𝑢𝑢𝑚
𝐸𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 ) (2.1)

𝑣𝑎𝑐𝑢𝑢𝑚
where 𝐺𝑠𝑢𝑟𝑓𝑎𝑐𝑒+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 is the free energy in the absence of a solvent, computed here

within the harmonic approximation using plane-wave DFT calculations for periodic slab

𝑙𝑖𝑞𝑢𝑖𝑑
models, 𝐺𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 is the free energy (without vibrational contributions) of the

surface species when the surface cluster model is immersed in an implicit solvent (which

is obtained by extracting selected metal atoms and removing the periodic boundary

𝑙𝑖𝑞𝑢𝑖𝑑
conditions). We note that 𝐺𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 does not contain vibrational contributions

𝑣𝑎𝑐𝑢𝑢𝑚
of the adsorbate that are already considered in the first term. Finally, 𝐸𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒

is the DFT energy of the same cluster in the absence of the solvent. To compute the

𝑙𝑖𝑞𝑢𝑖𝑑
𝐺𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 term, COSMO-RS calculations are performed using the

COSMOtherm program.37 The COSMOtherm program for solvent thermodynamic

properties requires COSMO calculations to be performed at the BP-TZVP level of theory.

In addition to implicit solvation calculations performed with the iSMS method, implicit

solvation calculations were also performed at 473 K using VASPsol38-39 with a relative

permittivity of water of 34.82 at reaction conditions.40 We used the default values for the

cutoff charge density, 𝑛𝐶 , and for the width of the diffuse cavity, 𝜎.38 We also employed

the default effective surface tension parameter, 𝜏, for describing the cavitation, dispersion,

and repulsive interaction between the solute and the solvent that are not captured by the

electrostatic terms.38 While the surface tension parameter is likely most accurate only for

simulations at 298 K and not at 473 K, it is an optimized parameter of the solvent model

that cannot easily be obtained at other temperatures. Due to the absence of a large number

of experimental solvation data at 473 K, we decided that the default parameter is likely the

11
most meaningful choice. All other computational details for periodic implicit solvation

calculations were kept the same as in our periodic vapor phase calculations. We note that

we used a reaction temperature of 473 K in this study since it is a typical reaction

temperature for the HDO of organic acids41-42.

2.3.3 Microkinetic modeling

For surface reactions, the forward rate constant (𝑘𝑓𝑜𝑟 ) of each reaction was

calculated as:


𝑘𝐵 𝑇 −∆𝐺
𝑘𝑓𝑜𝑟 = 𝑒 𝑘𝐵 𝑇 (2.2)

where 𝑘𝐵 is the Boltzmann constant, T represents the reaction temperature, h is the Planck

constant, and ∆𝐺 ‡ denotes the activation free energy at the corresponding reaction


condition. In the solvated environment, the activation free energy (∆𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 ) and the

𝑟𝑥𝑛
reaction free energy (∆𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 ) are calculated as:

‡ ‡ 𝑇𝑆 𝐼𝑆
∆𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 = ∆𝐺𝐺𝑎𝑠 + [𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 − 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 ] (2.3)

𝑟𝑥𝑛 𝑟𝑥𝑛 𝐹𝑆 𝐼𝑆
∆𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 = ∆𝐺𝐺𝑎𝑠 + [𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 − 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 ] (2.4)

𝐼𝑆 𝑇𝑆 𝐹𝑆
where 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 , 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 , and 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 represent the solvation free energies of initial,

‡ 𝑟𝑥𝑛
transition, and final states, respectively, and ∆𝐺𝐺𝑎𝑠 and ∆𝐺𝐺𝑎𝑠 are the respective activation

and reaction free energies under gas phase conditions. The reverse rate constants (𝑘𝑟𝑒𝑣 ) are

calculated from the thermodynamic equilibrium constants, 𝐾.

𝑘𝑓𝑜𝑟
𝑘𝑟𝑒𝑣 = (2.5)
𝐾

12
For an adsorption reaction, 𝐴(𝑔) + ∗ → 𝐴∗ , the adsorption rate is given by collision

theory with a sticking coefficient of 1, independent of solvent,

1
𝑘𝑓𝑜𝑟 = (2.6)
𝑁0 √2𝜋𝑚𝐴 𝑘𝐵 𝑇

where 𝑚𝐴 is the molecular weight of the adsorbent 𝐴 and 𝑁0 denotes the number of sites

per unit area (1.866 × 1019 𝑚−2). While the use of collision theory with a sticking

coefficient of 1 can be questioned, we found adsorption processes to be never rate

controlling in our models such that the choice of sticking coefficient has likely no effect

on the reported results. Also, Zhang et al.43 recently showed that collision theory is a

reasonable approximation for adsorption processes from a liquid as long as mass transfer

limitation in the liquid are negligible.

Adsorption equilibrium constants are calculated as

−∆𝐺 𝑎𝑑𝑠
𝐾= 𝑒 𝑘𝐵 𝑇 (2.7)

where ∆𝐺 𝑎𝑑𝑠 is the adsorption free energy which is obtained in the presence of a solvent

as,

𝑎𝑑𝑠 𝑎𝑑𝑠 𝐴 ∗ ∗
∆𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 = ∆𝐺𝐺𝑎𝑠 + [𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 − 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 ] (2.8)

𝑎𝑑𝑠 𝐴 ∗ ∗
where ∆𝐺𝐺𝑎𝑠 is the adsorption free energy in the gas phase and 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 and 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 are

the solvation free energies of the adsorbent 𝐴 on the surface and a clean surface immersed

in the solvent, respectively. With the forward and reverse rate constant calculated, all

nonlinear steady-state reactor equations are solved simultaneously under realistic reaction

conditions to compute the surface coverages and turnover frequency.

13
2.3.4 Lateral interaction effects

Preliminary results of the microkinetic model indicate that CO and H are the most

abundant surface species. Hence, a method which considers the lateral interaction effects

of CO and H on the stability of each adsorbed species and transition state in the

microkinetic model is needed. Our proposed method contains four steps.17, 44

Step 1: choose a reasonable coverage of high coverage species and perform DFT

calculations for all intermediates in the presence of the abundant species. We define surface

coverage of an adsorbate as the ratio of the number of sites it occupies to total number of

sites available on the surface. For example, 3 adsorbed H or 1 adsorbed CO on a Ni surface

(12 atoms) leads to 0.25 ML surface coverage of H or CO, respectively, since H occupies

1 site while CO occupies 3 sites (see below in this section for a detailed discussion).

Step 2: compute coverage dependent adsorption energies as,

𝐺𝑖 (𝜃𝑗 ) = 𝐺𝑖 (0) + ∑ 𝑎𝑖𝑗 𝜃𝑗 (2.9)


𝑗

where 𝐺𝑖 (𝜃𝑗 ) represents the free energy of species 𝑖 as a function of surface coverage of

abundant species, 𝜃𝑗 , 𝐺𝑖 (0) is the free energy of species 𝑖 on the clean surface without

considering the lateral interaction effect, and 𝑎𝑖𝑗 is the correction factor of species 𝑖 in the

presence of the abundant species 𝑗 which can be acquired from equation (2.9) when the

coverages of all other species are equal to zero,

𝐺𝑖 (𝜃𝑗 ) − 𝐺𝑖 (0)
𝑎𝑖𝑗 = (2.10)
𝜃𝑗

where 𝐺𝑖 (𝜃𝑗 ) is the free energy of species 𝑖 in the presence of only abundant species 𝑗.

14
Step 3: compute analogous to equation (2.9) for each transition state the free energy

in the presence of the high surface coverage species,

𝑇𝑆
𝐺𝑛𝑇𝑆 (𝜃𝑗 ) = 𝐺𝑛𝑇𝑆 (0) + ∑ 𝑎𝑛𝑗 𝜃𝑗 (2.11)
𝑗

where superscript 𝑇𝑆 indicates the transition state, subscript 𝑛 represents the reaction

𝑇𝑆
number and 𝑎𝑛𝑗 is the adsorption correction factor for the lateral interaction effect of the

abundant species 𝑗 on the transition state of reaction number 𝑛. To avoid optimizing

transition states in the presence of various adsorbates, we assumed that the transition states
𝑇𝑆
are affected as half the reactant and half the product state and hence, the coefficient 𝑎𝑛𝑗 is

estimated as

𝑇𝑆 1 𝑛
𝑎𝑛𝑗 = ∑ 𝑎𝑖𝑗 (2.12)
2
𝑖

𝑛
where 𝑎𝑖𝑗 represents the adsorption correction factor for the lateral interaction effect of the

abundant species 𝑗 on all species participating in reaction 𝑛.

Step 4: compute coverage dependent reaction and activation free energies as

follows for a generic surface reaction 𝑛: (𝐴𝐵 ∗ + ∗ ↔ 𝐴∗ + 𝐵 ∗ )

𝑟𝑥𝑛
∆𝐺𝑛𝑟𝑥𝑛 (𝜃𝑗 ) = ∆𝐺𝑛𝑟𝑥𝑛 (0) + ∑ ∆𝑎𝑛𝑗 𝜃𝑗 ,
𝑗
𝑟𝑥𝑛
∆𝑎𝑛𝑗 = (𝑎𝐴∗𝑗 + 𝑎𝐵∗𝑗 − 𝑎𝐴𝐵∗𝑗 − 𝑎∗𝑗 ) (2.13)

∆𝐺𝑛‡ (𝜃𝑗 ) = ∆𝐺𝑛‡ (0) + ∑ ∆𝑎𝑛𝑗



𝜃𝑗 , ‡
∆𝑎𝑛𝑗 𝑇𝑆
= (𝑎𝑛𝑗 − 𝑎𝐴𝐵∗𝑗 ) (2.14)
𝑗

and for a generic adsorption reaction: (𝐴(𝑔) + ∗ ↔ 𝐴∗ )

15
𝑎𝑑𝑠 𝑎𝑑𝑠
∆𝐺𝐴𝑎𝑑𝑠 (𝜃𝑗 ) = ∆𝐺𝐴𝑎𝑑𝑠 (0) + ∑ ∆𝑎𝐴𝑗 𝜃𝑗 , ∆𝑎𝐴𝑗 = (𝑎𝐴∗𝑗 − 𝑎∗𝑗 ) (2.15)
𝑗

where ∆𝐺𝑛𝑟𝑥𝑛 (𝜃𝑗 ) and ∆𝐺𝑛‡ (𝜃𝑗 ) are the reaction and activation free energies of reaction 𝑛,

respectively, and ∆𝐺𝐴𝑎𝑑𝑠 (𝜃𝑗 ) is the coverage dependent adsorption free energy of species

𝐴. Similarly, ∆𝐺𝑛𝑟𝑥𝑛 (0) and ∆𝐺𝑛‡ (0) are the reaction and activation free energies of reaction

𝑛, respectively, and ∆𝐺𝐴𝑎𝑑𝑠 (0) is the adsorption free energy of species 𝐴 in the dilute limit.

Specifically, preliminary microkinetic modeling results, that neglect lateral

interactions, show that the most abundant surface species are CO and H. Thus, we

computed the adsorption energy of all surface species in the presence of 0.25 ML H and

CO. Thus, equations 2.13 to 2.15 become:

𝑟𝑥𝑛 𝑟𝑥𝑛
∆𝐺𝑛𝑟𝑥𝑛 (𝜃𝐻 , 𝜃𝐶𝑂 ) = ∆𝐺𝑛𝑟𝑥𝑛 (0) + [∆𝑎𝑛𝐻 × 𝜃𝐻 ] + [∆𝑎𝑛𝐶𝑂 × 𝜃𝐶𝑂 ] (2.16)

∆𝐺𝑛‡ (𝜃𝐻 , 𝜃𝐶𝑂 ) = ∆𝐺𝑛‡ (0) + [∆𝑎𝑛𝐻


‡ ‡
× 𝜃𝐻 ] + [∆𝑎𝑛𝐶𝑂 × 𝜃𝐶𝑂 ] (2.17)

∆𝐺𝐴𝑎𝑑𝑠 (𝜃𝐻 , 𝜃𝐶𝑂 ) = ∆𝐺𝐴𝑎𝑑𝑠 (0) + [∆𝑎𝐴𝐻


𝑎𝑑𝑠 𝑎𝑑𝑠
× 𝜃𝐻 ] + [∆𝑎𝐴𝐶𝑂 × 𝜃𝐶𝑂 ] (2.18)

Figures A.2 and A.3 illustrate the correction terms for individual intermediates and

transition states, respectively. We note that we use a mean-field microkinetic model in this

study despite significant lateral interactions that can even be attractive, such as in the case

of CO and adsorbed PAC. This attractive interaction can be rationalized by the fact that the

interaction between a Lewis acid and Lewis base coadsorbed on the surface is surprisingly

strong.45 Charges on adsorbed CO and PAC, separately, on the surface, and on coadsorbed

CO and PAC are presented in the Appendix A (Figure A.4). Strong attractive lateral

interactions can result in island formation and the mean-field approximation46

overestimating reaction rates.

16
Figure A.5 displays the differential zero-point corrected energy of adding an extra

CO or H as a function of the number of CO or H already adsorbed on the surface. Adsorbing

a fifth CO molecule on a surface containing 12 surface Ni atoms leads to a significant

increase in energy such that practically only four CO molecules fit on this surface while up

to 12 hydrogen atoms fit. Thus, it can be concluded that on Ni (111) CO and H occupy 3

and 1 sites, respectively. We used analogue procedures to count occupied site(s) for other

surface species that are bonded through carbon atoms to the surface. Given that the

assumption that CO occupies 3 instead of 1 site is uncommon, we also present in the

Appendix A (Tables A.6 and A.7) all simulation results for the case that CO occupies a

single site.

2.4 Result and discussion

2.4.1 Solvent effect on the adsorption strength of surface species

Solvents influence the stability of each surface intermediate and consequently

facilitate certain types of cleavages. Hence, understanding the impact of the solvent

environment on each individual intermediate could shed light on how a solvent could

reshape the reaction mechanism. The change in the adsorption strength of different

intermediates can be quantified by considering the adsorption processes in the absence and

presence of a solvent as,

𝐴(𝑔) +∗ (𝑔) ↔ 𝐴∗ (𝑔) (2.19)

𝐴(𝑔) +∗ (𝑙𝑖𝑞) ↔ 𝐴∗ (𝑙𝑖𝑞) (2.20)

∆∆(𝐺𝑎𝑑𝑠,𝐴 (𝑙𝑖𝑞)) = ∆𝐺𝑎𝑑𝑠,𝐴 (𝑙𝑖𝑞) − ∆𝐺𝑎𝑑𝑠,𝐴 (𝑔)


= [𝐺 𝐴∗ (𝑙𝑖𝑞) − 𝐺 𝐴∗ (𝑔)] − [𝐺 ∗ (𝑙𝑖𝑞) − 𝐺 ∗ (𝑔)] (2.21)

17
where, ∆𝐺𝑎𝑑𝑠,𝐴 (𝑙𝑖𝑞) and ∆𝐺𝑎𝑑𝑠,𝐴 (𝑔) are the adsorption free energies of a “gas” molecule

𝐴 in the liquid (or solvent) and in the vacuum, respectively. 𝐺 𝐴∗ (𝑙𝑖𝑞) and 𝐺 𝐴∗ (𝑔) are the

free energies of 𝐴 adsorbed on the surface in liquid and in vacuum, respectively, and

𝐺 ∗ (𝑙𝑖𝑞) and 𝐺 ∗ (𝑔) are the free energies of the clean surface in liquid and in vacuum

environments, respectively.

Table 2.1 displays the change in the adsorption free energy of each adsorbed

intermediate in the presence and absence of the solvent molecules using the iSMS method.

The most stabilized adsorbed intermediates are CH3CCOO and CH3CHCOO. In water,

their adsorption strengths are increased by 0.55 and 0.31 eV, respectively, due to adsorption

through the carbon atom and the charged carboxyl group pointing in the direction of the

solvent. This trend can also be seen in the presence of 1,4-dioxane with a slightly lower

stabilization of 0.33 eV for CH3CCOO and 0.17 eV for CH3CHCOO.

In general, the solvent effect on the adsorption strength is pertinent to the strength

of attractive solvent-solvent interactions in solution (via e.g. hydrogen bonding in the

aqueous phase) relative to attractive solvent-adsorbate interaction. In the case of adsorbed

CH3CCOO and CH3CHCOO moieties in water, the negatively charged carboxyl group has

a strong electrostatic interaction with water molecules that accounts for the stabilization of

these adsorbed species in aqueous phase. Their reduced stabilization in 1,4-dixane is

ascribed to the fact that 1,4-dioxane is a non-polar solvent and hence incapable of strong

interaction with the charged carboxyl group.

Moreover, Table A.1 shows the aqueous phase effect on the adsorption energy of

intermediates at two different temperatures of 298 K and 473 K based on two implicit

18
solvation schemes: iSMS and VASPsol. The choice of room temperature is due to the fact

that all VASPsol parameters have been optimized based on experimental result at room

temperature. This means VASPsol results are most accurate at room temperature. The only

parameter adjusted in the VASPsol calculations at 473 K is the relative permittivity of

water. The data suggest that both iSMS and VASPsol predict (de)stabilization of adsorbed

moieties qualitatively similar.

2.4.2 Solvent effect on elementary surface reactions

By changing the stability of various surface intermediates and transition states, the

solvent environment plays an important role in altering the overall reaction pathway.

Therefore, to elucidate the solvent effect on each elementary reaction, the reaction and

activation free energies of all elementary steps in vacuum and in solvents are computed at

473 K using our iSMS method (see Table 2.2).

Propionic acid is consumed directly through three main reactions, which produce

three different intermediates: propanoyl (CH3CH2CO) as the product of OH removal of

𝑤𝑎𝑡𝑒𝑟 ) 1,4−𝑑𝑖𝑜𝑥𝑎𝑛𝑒
PAC in step 1 (∆(∆𝐺𝑟𝑥𝑛 = 0.00,∆(∆𝐺𝑟𝑥𝑛 ) = 0.03), CH3CHCOOH as the

𝑤𝑎𝑡𝑒𝑟 )
product of dehydrogenation of the 𝛼-carbon of PAC in step 2 (∆(∆𝐺𝑟𝑥𝑛 =

1,4−𝑑𝑖𝑜𝑥𝑎𝑛𝑒
−0.07,∆(∆𝐺𝑟𝑥𝑛 ) = −0.03), and propionate (CH3CH2COO) as the product of O-H

𝑤𝑎𝑡𝑒𝑟 ) 1,4−𝑑𝑖𝑜𝑥𝑎𝑛𝑒
bond dissociation of PAC in step 28 (∆(∆𝐺𝑟𝑥𝑛 = −0.03, ∆(∆𝐺𝑟𝑥𝑛 ) = 0.00).

Based on the solvent effect values on the reaction and activation free energies of these three

paths, it appears that the solvents do not impact the starting reactions of HDO of propionic

acid notably.

19
Furthermore, Table 2.2 illustrates a small change in the reaction and activation free

energies of decarbonylation steps in the presence of liquid water or 1,4-dioxane. For

decarbonylation steps that produce CO, the reaction energies in water and 1,4-dioxane are

around 0.1 eV more exergonic than in the gas phase as a result of CO adsorption being

stabilized in water and 1,4-dioxane by 0.25 and 0.18 eV, respectively (Table 2.1). In

contrast to the small effect of the liquid phase environments on the DCN reactions, the

solvent effect is pronounced on the elementary reactions of the DCX. For convenience of

comparison of the solvent effect on important elementary reaction steps of the DCX, we

classified them into three different classes of similar types of bond dissociations.

Class I: 𝛼-carbon dehydrogenation. The solvent effect on the reaction and

activation free energies of dehydrogenation of CH3CH2COO in step 30 is -0.27 and -0.17

eV in water and -0.14 and -0.09 eV in 1,4-dioxane, respectively. A slightly higher

exergonic solvent effect in water compared to in 1,4-dioxane originates from its polarity

that in turn stabilizes CH3CHCOO adsorption (the main product of this step) more in water

than in 1,4-dioxane. Another important dehydrogenation reaction occurs in step 34, where

CH3CHCOO, the second most stabilized intermediate in the liquid phases, is converted to

CH3CCOO, the most stabilized intermediate in the presence of the solvents. In particular,

a decline of 0.26 eV in the reaction free energy in water turns the endergonic reaction in

vacuum into an exergonic reaction in aqueous phase which facilitates the reaction

thermodynamically. In contrast, the endergonic solvent effect on the activation barrier of

this step by 0.12 and 0.09 eV in water and 1,4-dioxane, respectively, makes this reaction

kinetically unfavorable in the presence of solvents.

20
Class II: O-H bond cleavage: The O-H bond dissociation of CH3CHCOOH to

produce CH3CHCOO in step 31 and CH3CCOOH to generate CH3CCOO in step 35 are

facilitated remarkably in the presence of solvents as a result of significant stabilization of

CH3CHCOO and CH3CCOO on the surface.

Class III: C-C bond cleavage to produce CO2: Production of CH3CH and CO2 by

C-C bond cleavage of CH3CHCOO (step 33) and formation of CH3C and CO2 from

CH3CCOO (step 38) are some of the most influenced DCX elementary reactions by a liquid

phase environment. The solvents have endergonic effects on the reaction and activation

free energies of these steps since the reactant states of these steps (CH3CHCOO and

CH3CCOO) are stabilized significantly compared to the corresponding products and TS of

these steps in the presence of the solvent molecules.

Overall, both solvents altered the elementary reactions in the DCX mechanism

significantly while the solvent effect is small on the elementary reaction steps in the DCN

mechanism. In addition, the same calculations were performed using the VASPsol program

package and the results are compared to iSMS in Table A.3 and A.4 at 298 K and 473K.

In the case of the above-mentioned O-H bond cleavage reactions (step 31 and step 35),

VASPsol predicts endergonic solvent effects on activation barriers which contradicts both

our iSMS results and our intuition and experimental studies.47-53

2.4.3 Microkinetic modeling

To predict the overall TOF of our reaction network in both the presence and absence

of a solvent, a microkinetic model at a temperature of 473 K was developed. H2 and CO

partial pressures were set to 0.1 and 0.001 bar, respectively, and the partial pressure of

21
propionic acid, H2O (in the absence of liquid water), and CO2 were set to 1 bar. In liquid

water, the water chemical potential and partial pressure were computed assuming

equilibrium between the liquid and a gas phase, i.e.

𝑥𝐻2 𝑂 𝑓𝐻𝐿2 𝑂 = 𝑦𝐻2 𝑂 𝑃𝑡𝑜𝑡 = 𝑃𝐻2 𝑂 (2.22)

where 𝑥𝐻2 𝑂 is the mole fraction of water in the liquid phase (assumed to be close

to 1), 𝑓𝐻𝐿2 𝑂 is the pure water fugacity at 473 K, 𝑦𝐻2 𝑂 is the water mole fraction in the vapor

phase, 𝑃𝑡𝑜𝑡 is the total pressure of the system, and 𝑃𝐻2 𝑂 indicates the water partial pressure

assuming an ideal gas phase. The pure water fugacity at 473 K was obtained from a steam

table and a Lee/Kesler generalized-correlation table,54 the partial pressure of water was

calculated to be 14.36 bar and this value was used in all liquid water microkinetic models.

In the following sections, we represent our microkinetic results both in the gas phase and

in the solvent environments and clarify the solvent influence on important parameters of

the reaction network.

2.4.3.1 Dominant pathway and TOF

To determine the surface coverage, rate of each elementary surface reaction, and

the overall turnover frequency (defined here as the consumption of propionic acid), a

microkinetic model was developed taking the lateral interaction effects of the most

dominant surface species, CO and H, into account in both the gas phase and condensed

phase models. Table 2.3 summarizes the overall TOF, the decarboxylation and

decarbonylation rate, and the dominant species coverages during the HDO of propionic

acid over the Ni (111) catalyst surface in various reaction environments. In addition, the

TOF of each individual elementary reaction is listed in Table 2.4. Next, the dependence of

22
the overall vapor phase TOF on CO and H2 partial pressures is provided in Table A.5.

Finally, Figure 2.2 exhibits a schematic of the dominant pathways in the gas and liquid

phase environments.

2.4.3.1.1 Gas phase

The microkinetic model predicts that the dominant pathway begins with the

removal of the hydroxyl group of propionic acid in step 1, followed by dehydrogenation of

the 𝛼-carbon of propanoyl in step 4, decarbonylation of CH3CHCO in step 8,

hydrogenation of CH3CH in step 25, and finally dehydrogenation of the 𝛽-carbon in step

27 to produce ethylene. The most abundant species on the surface are H and CO with

coverages of 0.631 and 0.357, respectively. Moreover, the free site coverage is 0.011. The

overall TOF, which includes both DCX and DCN, is determined to be 3.46 × 10−8 𝑠 −1.

The decarbonylation rate is 3.46 × 10−8 𝑠 −1 , while the decarboxylation rate is predicted

to be 8.72 × 10−13 𝑠 −1 . Hence, the selectivity towards DCN is close to 100% in the vapor

phase. We note that an experimental study of catalytic HDO of propionic acid over

supported group VIII noble metals performed by Lugo-Jose et al.55 found a TOF of

1.5 × 10−4 𝑠 −1 over a Ni/SiO2 catalyst at 473 K. Therefore, it is possible that the

investigated (111) facet of Ni is not the most relevant active site for this catalytic

transformation and other Ni facets or phases could be responsible for the experimentally

observed kinetics.

2.4.3.1.2 Liquid water and 1,4-dioxane

The dominant pathway in the liquid phases is the same as in the gas phase until

CH3CHCO formation. Then, the dominant pathway follows a dehydrogenation in step 9 to

23
form the CH3CCO intermediate, a C-C bond cleavage in step 14 to produce CH3C, and

further hydrogenations through steps 24, 25, and the dehydrogenation of the 𝛽-carbon from

CH3CH2 in step 27 to produce ethylene. This shift in the dominant pathway is attributed to

the activation barriers of C-C bond dissociation of CH3CHCO (step 8) and C-H bond

cleavage of the 𝛼-carbon of CH3CHCO (step 9), which are 0.15 and 0.19 eV, respectively,

in the gas phase. We conclude that due to the lower activation barrier of step 8 relative to

step 9, step 8 is preferred in the gas phase. On the other hand, the activation barrier of step

8 is barely changed in the condensed phases while that of step 9 is decreased by 0.06 and

0.03 eV in water and 1,4-dioxane, respectively. Consequently, step 9 is more favorable in

the condensed phases and hence forces the catalytic pathway toward the production of

CH3CHCO rather than CH3CH. As shown in Table 2.4, the overall TOF in the liquid

phases is very similar to the one in the gas phase and does not strongly depend on the cavity

radius of Ni. However, the abundant adsorbed intermediates change in liquid phase. The

CO coverage increases while the hydrogen coverage significantly decreases. Furthermore,

the free site coverage decreases in the presence of the liquid phase. The overall effect on

the turnover frequency is not significant since the decarbonylation pathway is the dominant

HDO pathway and it is not strongly affected by the solvent. In contrast, the DCX pathways

are more affected by liquid water and the DCX rate is increased by approximately two

orders of magnitude in liquid water and one order of magnitude in liquid 1,4-dioxane.

2.4.3.2 Sensitivity analysis, apparent activation barrier, and reaction orders

To analyze the sensitivity of each individual elementary reaction, we used

Campbell’s degree of rate control,56-60 𝑋𝑅𝐶,𝑖 . This criterion describes which transition state

is the most influential on the overall reaction rate.

24
𝑘𝑖 𝜕𝑟
𝑋𝑅𝐶,𝑖 = ( ) (2.23)
𝑟 𝜕𝑘𝑖 𝐾 ,𝑘
𝑖 𝑗 ≠𝑘𝑖

where 𝑟 is the overall reaction rate, 𝑘𝑖 is the forward rate constant for step 𝑖, and 𝐾𝑖 is the

equilibrium constant for step 𝑖.

Next, the apparent activation barriers were computed in the temperature ranges of

473 to 623 K in all reaction environments.

𝜕ln (𝑟)
𝐸𝑎 = 𝑅𝑇 2 ( ) (2.24)
𝜕𝑇 𝑝𝑖

Figure 2.3 demonstrates the overall TOF as a function of inverse temperature in different

reaction environments at a hydrogen partial pressure of 0.1 bar.

Finally, the reaction order with respect to hydrogen, CO and propionic acid were

calculated at 473 K and a pressure ranges of 0.9 to 1.1 bar for H2, 7.5 × 10−4 to

12.5 × 10−4 bar for CO, and 0.9 to 1.1 bar for PAC.

𝜕 ln(𝑟)
𝑎𝑖 = ( ) (2.25)
𝜕 ln(𝑝𝑖 ) 𝑇,𝑝
𝑗≠𝑖

2.4.3.2.1 Gas Phase

At 473 K, microkinetic modeling results suggest that the initial C-OH bond

dissociation in step 1 is the most sensitive transition state in the gas phase with an 𝑋𝑅𝐶

value of 0.65. Another sensitive transition state belongs to step 8 (C-C bond cleavage of

𝛼-carbon of CH3CHCO), 𝑋𝑅𝐶 = 0.26. Also, step 9 (C-H bond cleavage of 𝛼-carbon of

CH3CHCO) with 𝑋𝑅𝐶 = 0.03 is the third most important step given that step 8 and 9 are

two competing steps in the reaction network. Finally, the apparent activation energy is

25
predicted to be 2.38 eV and the reaction rate is independent of CO partial pressure (𝑎𝐶𝑂 =

0), drops almost with a negative square of the hydrogen partial pressure (𝑎𝐻2 = −2.2), and

increases linearly with propionic acid partial pressure (𝑎𝑃𝐴𝐶 = 1.0 ).

2.4.3.2.2 Liquid water

The overall turnover frequency is highly sensitive to the reaction barrier of steps 9

(dehydrogenation of 𝛼-carbon of CH3CHCO) in liquid water. This is attributed to the shift

in the dominant pathway leading to step 9 becoming more favorable in the condensed

phase. The values of Campbell’s degree of rate control are 0.88 for step 9, 0.10 for step 1,

and 0.01 for step 8. The apparent activation energy obtained from the microkinetic model

is 2.71 eV. In contrast to the gas phase, the reaction rate decreases with CO partial pressure

(𝑎𝐶𝑂 = −0.7) in water. The reaction orders with respect to hydrogen and PAC remain

approximately the same as those in the gas phase (𝑎𝐻2 = −1.4, 𝑎𝑃𝐴𝐶 = 1.0).

2.4.3.2.3 Liquid 1,4-dioxane

In liquid 1,4-dioxane, at 473 K, steps 9 and 1 are the most sensitive transition states

with 𝑋𝑅𝐶 = 0.55 𝑎𝑛𝑑 𝑋𝑅𝐶 = 0.39, respectively, and step 8 has 𝑋𝑅𝐶 = 0.03. The decrease

in 𝑋𝑅𝐶 from 0.88 in water to 0.55 in 1,4-dioxane is related to a stronger stabilization of the

transition state of step 9 in liquid water compared to that in 1,4-dioxane. The apparent

activation barrier is 2.44 eV in the presence of 1,4-dioxane and the reaction orders are

almost similar to those in the liquid water environment (𝑎𝐶𝑂 = −0.4 𝑎𝐻2 = −1.7, 𝑎𝑃𝐴𝐶 =

1.0).

26
2.4.3.3 Comparison of implicit solvation models

While we have more confidence in our iSMS solvation model relative to VASPsol,

we also solved the microkinetic model using free energies computed with VASPsol in

(implicit) liquid water at 473 K. Table A.11 in the Appendix A provides a detailed

comparison between these two different solvent models. Overall, both solvation schemes

predict a similar solvent effect on the decarbonylation and decarboxylation of PAC. The

overall rate and the DCN rate are decreased in liquid water over Ni(111) while the DCX

rate is increased. While iSMS predicts a decrease of ~80% relative to the vapor phase for

the DCN rate, VASPsol only predicts a decrease of ~50%. Next, iSMS predicts an increase

in DCX rate by a factor 17 relative to the vapor phase, while VASPsol predicts an increase

by a factor 142. Nevertheless, the DCX hardly contributes to the overall rate in both

solvation schemes. The key differences between the solvation models are the surface

coverages of CO and H and a change in the rate controlling step being predicted by the

different solvation models. iSMS predicts a stronger solvent stabilization of CO on the

surface of 0.25 eV relative to a stabilization of only 0.08 eV predicted by VASPsol (see

Table A.1), explaining the difference in surface coverage. Next, iSMS predicts the C-H

bond cleavage of CH3CHCO in step 9 to be the rate determining step in aqueous phase,

while VASPsol predicts the C-C bond cleavage of CH3CHCO in step 8 to be the rate

determining step. This difference in rate controlling step is a result of the opposite solvent
𝑎𝑐𝑡
effect predicted with iSMS and VASPsol on the activation barrier of step 9 ( ∆∆𝐺𝑖𝑆𝑀𝑆 =

𝑎𝑐𝑡
−0.06 𝑒𝑉, ∆∆𝐺𝑉𝐴𝑆𝑃𝑠𝑜𝑙 = 0.07 𝑒𝑉, see Table A.4). We note here that all iSMS calculations

with different cavity radius for Ni predict a solvent stabilization in water, leading us to

have more confidence in the iSMS calculations (see Table A.10).

27
2.5 Conclusion

A microkinetic model has been developed for the decarboxylation and

decarbonylation of propanoic acid over Ni(111) that considers the lateral interaction effect

of the most dominant surface species, CO and H. In addition, the effect of two solvents,

liquid water and 1,4-dioxane, has been investigated with the help of periodic DFT

calculations, implicit solvation scheme, and microkinetic modeling. Mean-field

microkinetic models were developed for each solvent at a temperature of 473 K and a

hydrogen partial pressure of 0.1 bar. Under all conditions, the decarbonylation is favored

over the decarboxylation. The dominant pathway in gas phase begins with the removal of

the hydroxyl group of propionic acid in step 1 (C-OH bond cleavage), followed by two

dehydrogenations and one hydrogenation steps to produce ethylene, i.e., CH3CH2COOH

→ CH3CH2CO → CH3CHCO → CH3CH → CH3CH2 → CH2CH2. Next, the dominant

pathway in the condensed phases shifts after production of CH3CHCO. It continues with a

dehydrogenation step to form a CH3CCO surface intermediate, followed by

decarbonylation to produce CH3C, and subsequent hydrogenations and one last

dehydrogenation to reach ethylene, i.e., CH3CH2COOH → CH3CH2CO → CH3CHCO →

CH3CCO → CH3C → CH3CH → CH3CH2 → CH2CH2. In the presence of a significant

hydrogen pressure, ethylene can be hydrogenated to ethane.

Although no significant solvent effect was observed on the decarbonylation rate,

liquid water and 1,4-dioxane increase the decarboxylation rate by two orders of magnitude

and one order of magnitude, respectively, relative to the gas phase. This noticeable solvent

effect on the decarboxylation rate can be explained by the significant solvent stabilization

of two key surface intermediates in the decarboxylation mechanism, CH3CCOO and

28
CH3CHCOO. Next, a sensitivity analysis shows that C-OH bond cleavage of propionic

acid is the most rate controlling step in the gas phase (𝑋𝑅𝐶1 = 0.65), In solvent

environments, its dominance will be replaced with reaction step 9 (C-H bond cleavage of

𝛼-carbon of CH3CHCO). Finally, computations suggest that the (111) surface of Ni is

likely not the active facet for the HDO of propionic acid neither in gas phase nor the studied

solvents, and therefore other facets need to be investigated to identify the experimentally

relevant active site.

2.6 Acknowledgements

We gratefully acknowledge financial support from the U.S. Department of Energy,

Office of Basic Energy Science, Catalysis Science program under Award DE-SC0007167.

In addition, this work was partially supported by the South Carolina Smart State Center for

Strategic Approaches to the Generation of Electricity (SAGE). Computational resources

have been provided by the National Energy Research Scientific Computing Center

(NERSC) which is supported by the Office of Science of the U.S. Department of Energy

and in part by XSEDE under grant number TG-CTS090100. Computational resources from

the CASCADE cluster from the Environmental Molecular Sciences Laboratory (EMSL)

under Pacific Northwest National Laboratory (PNNL) are also used for selected DFT

calculations. Finally, computing resources from the USC High Performance Computing

Group are gratefully acknowledged.

29
2.7 Tables and Figures

Table 2.1: Solvent effect on the stability of various adsorbed species in the HDO of PAC
to ethane and ethylene over a Ni(111) catalyst surface model at a reaction temperature of
473 K. ΔΔGrxn indicates the difference in the adsorption free energy of the corresponding
intermediate in the presence and the absence of the solvent. Asterisk (*) represents a
surface adsorption site and multiple asterisks are indicative of the number of occupied
active sites.

ΔΔGrxn, eV
Adsorbed species
Water 1,4-Dioxane
CH2C*** -0.08 -0.05
CH2CH*** -0.01 0.00
CH2CH2** 0.10 0.12
CH2CHCO**** -0.09 -0.03
CH2CHCOOH**** -0.11 -0.04
CH3C*** -0.01 -0.01
CH3CCO**** -0.12 -0.06
CH3CCOO*** -0.55 -0.33
CH3CCOOH*** -0.12 -0.07
CH3CH*** 0.00 -0.01
CH3CH2** 0.01 -0.01
CH3CH2CO*** -0.05 -0.04
CH3CH2COO** -0.06 -0.04
CH3CH2COOH* -0.03 0.02
CH3CH3* 0.02 0.02
CH3CHCO** -0.09 -0.05
CH3CHCOO*** -0.31 -0.17
CH3CHCOOH** -0.10 -0.06
CHCH**** 0.05 0.08
CHCHCO**** -0.15 -0.05
CHCHCOOH**** -0.13 -0.06
CO*** -0.25 -0.18
CO2* -0.04 0.01
COOH** -0.16 -0.08
H* -0.01 0.00
H2O* 0.02 0.07
OH* 0.00 0.02

30
Table 2.2: Reaction and activation free energies of all elementary steps in the HDO of PAC to ethane and ethylene over a Ni(111)
catalyst surface model at 473 K. ΔΔGrxn and ΔΔGact indicate the reaction and the activation free energy differences between
corresponding reaction in the presence of solvent and in the gas phase, respectively.
Vacuum Water 1,4-Dioxane
# Reaction
ΔGrxn ΔGact ΔΔGrxn ΔΔGact ΔΔGrxn ΔΔGact
0 CH3CH2COOH + * → CH3CH2COOH* 0.69 -0.03 0.02
1 CH3CH2COOH* + 3* → CH3CH2CO*** + OH* -0.25 0.60 0.00 0.00 0.03 0.00
2 CH3CH2COOH* + 2* → CH3CHCOOH** + H* -0.05 0.51 -0.07 -0.03 -0.03 -0.01
3 CH3CH2CO***+ 2* → CH3CH2** + CO*** -0.61 0.78 -0.14 -0.03 -0.10 0.01
4 CH3CH2CO*** → CH3CHCO** + H* 0.19 0.49 -0.06 -0.03 -0.02 0.00
5 CH3CHCOOH** + * → CH3CHCO** + OH* -0.01 0.90 0.01 -0.03 0.04 0.00
6 CH3CHCOOH** + 3* → CH2CHCOOH**** + H* -0.42 0.57 -0.02 0.00 0.01 0.04
7 CH3CHCOOH** + 2*→CH3CCOOH*** + H* -0.19 0.35 -0.03 0.00 -0.01 0.01
8 CH3CHCO** + 4* → CH3CH*** + CO*** -1.14 0.15 -0.10 -0.01 -0.09 0.01
9 CH3CHCO** + 3* → CH3CCO**** + H* -0.73 0.19 -0.05 -0.06 -0.02 -0.03
31

10 CH3CHCO** + 3* → CH2CHCO**** + H* -0.43 0.49 -0.01 0.00 0.01 0.01


11 CH2CHCOOH**** + * → CH2CHCO**** + OH* -0.02 1.12 0.02 0.00 0.04 0.01
12 CH2CHCOOH**** + * → CHCHCOOH**** + H* -0.14 0.47 -0.05 0.01 -0.03 0.00
13 CH3CCOOH*** + 2* → CH3CCO**** + OH* -0.54 0.83 -0.01 -0.08 0.03 -0.03
14 CH3CCO**** + 2* → CH3C*** +CO*** -1.09 0.23 -0.08 -0.08 -0.08 -0.07
15 CH2CHCO**** + 2* → CH2CH*** + CO*** -0.71 0.41 -0.11 -0.01 -0.10 0.00
16 CH2CHCO**** + * → CHCHCO**** + H* -0.15 0.41 -0.08 0.00 -0.03 0.01
17 CHCHCOOH**** + * → CHCHCO**** + OH* -0.03 1.02 -0.01 -0.12 0.03 -0.06
18 CHCHCO**** + 3* → CHCH**** + CO*** -1.37 0.21 -0.06 -0.03 -0.08 -0.02
19 CH2CH*** + 2* → CHCH**** + H* -0.80 0.11 -0.03 0.02 -0.01 0.02
20 CH2CH2** + 2* → CH2CH*** + H* -0.10 0.48 -0.06 0.01 -0.03 0.01
21 CH2CH*** + * → CH2C*** + H* -0.54 0.18 -0.09 -0.02 -0.06 -0.01
22 CH3C*** + * → CH2C*** + H* 0.13 0.86 -0.08 -0.01 -0.05 0.01
23 CH3CH*** + * → CH2CH*** + H* 0.00 0.50 -0.03 -0.03 -0.01 0.00
24 CH3CH*** + * → CH3C*** + H* -0.67 0.09 -0.03 0.02 -0.01 0.01
25 CH3CH2** + 2* → CH3CH*** + H* -0.35 0.29 -0.03 0.01 -0.01 0.02
26 CH3CH3* + 2* → CH3CH2** + H* 0.22 0.94 -0.05 -0.02 -0.03 -0.01
27 CH3CH2** + * → CH2CH2** + H* -0.25 0.16 0.01 -0.01 0.02 0.02
28 CH3CH2COOH* + 2* → CH3CH2COO** + H* -0.95 0.00 -0.03 -0.07 0.00 -0.03
29 CH3CH2COO** + * → CH3CH2** + CO2* 0.71 1.88 0.06 -0.04 0.04 -0.03
30 CH3CH2COO** + 2* → CH3CHCOO*** + H* 0.65 1.52 -0.27 -0.17 -0.14 -0.09
31 CH3CHCOOH** + 2* → CH3CHCOO*** + H* -0.25 0.60 -0.23 -0.14 -0.11 -0.07
32 CH3CHCOOH** + 3* → CH3CH*** + COOH** -0.04 0.76 -0.06 -0.02 -0.02 0.02
33 CH3CHCOO*** + * → CH3CH*** + CO2* -0.29 0.80 0.31 0.12 0.17 0.10
34 CH3CHCOO*** + * → CH3CCOO*** + H* 0.14 0.85 -0.26 0.12 -0.17 0.09
35 CH3CCOOH*** + * → CH3CCOO*** + H* 0.08 0.97 -0.45 -0.18 -0.27 -0.07
36 CH3CCOOH*** + 2* → CH3C*** + COOH** -0.53 0.57 -0.06 -0.05 -0.02 -0.02
37 CH2CHCOOH**** + * → CH2CH*** + COOH** 0.38 0.95 -0.07 -0.01 -0.04 0.00
38 CH3CCOO*** + * → CH3C*** + CO2* -1.10 0.25 0.54 0.21 0.33 0.14
39 COOH** → CO2* + H* -0.50 0.75 0.15 -0.18 0.08 -0.14
40 COOH** + 2* → CO*** + OH* -1.11 0.34 -0.03 0.04 -0.03 0.03
32

41 H2O* + * → OH* + H* -0.43 0.75 0.00 -0.01 0.00 -0.01


42 CH3CH3 + * → CH3CH3* 0.81 0.02 0.02
43 CH2CH2 + 2* → CH2CH2** 0.13 0.10 0.12
44 H2O + * → H2O* 0.46 0.02 0.07
45 CO2 + * → CO2* 0.31 -0.04 0.01
46 CHCH + 4* → CHCH**** -1.86 0.05 0.08
47 CO + 3* → CO*** -0.71 -0.25 -0.18
48 H2 + 2* → H* + H* -0.42 -0.01 0.00
Table 2.3: Overall, decarbonylation, and decarboxylation turnover frequencies as well as important steady state surface coverages for
the HDO of PAC over a Ni(111) catalyst surface model in the gas phase, and in liquid water and 1,4-dioxane at 473 K. Note that
calculations for solvents were performed with the help of the COSMO-RS package with three different Ni cavity radii: with default
value, with a 10% increased value and a 10% decreased value relative to the default.

Water 1,4-Dioxane
Properties Gas
Default +10% -10% Default +10% -10%
DCN TOF 3.46×10-08 6.70×10-09 4.07×10-08 4.64×10-09 1.84×10-08 1.23×10-07 1.15×10-08
DCX TOF 8.72×10-13 1.47×10-11 1.47×10-10 2.12×10-11 5.24×10-13 6.64×10-12 1.02×10-12
Overall TOF (s-1) 3.46×10-08 6.71×10-09 4.09×10-08 4.66×10-09 1.84×10-08 1.23×10-07 1.15×10-08
θ* 0.011 0.003 0.003 0.003 0.004 0.005 0.004
θH* 0.631 0.206 0.154 0.149 0.296 0.253 0.207
θCO*** 0.357 0.789 0.834 0.849 0.699 0.741 0.789
θCH3C*** 0.000 0.000 0.000 0.000 0.000 0.000 0.000
33

θPAC* 0.000 0.000 0.004 0.000 0.000 0.000 0.000


θCH3CH2COO** 0.001 0.000 0.002 0.000 0.000 0.001 0.000
Table 2.4: Calculated turnover frequency (net rate) in the gas phase and in the solvents for all elementary reaction steps in the HDO of
PAC over a Ni(111) catalyst model surface at a temperature of 473 K and a hydrogen partial pressure of 0.1 bar.

TOF (s-1)
# Reaction
Gas Water 1,4-dioxane
0 CH3CH2COOH + * → CH3CH2COOH* 3.45×10-08 6.69×10-09 1.84×10-08
1 CH3CH2COOH* + 3* → CH3CH2CO*** + OH* 3.46×10-08 6.70×10-09 1.84×10-08
2 CH3CH2COOH* + 2* → CH3CHCOOH** + H* 1.37×10-12 8.98×10-12 4.04×10-13
3 CH3CH2CO***+ 2* → CH3CH2** + CO*** 2.65×10-09 2.48×10-11 2.75×10-10
4 CH3CH2CO*** → CH3CHCO** + H* 3.19×10-08 6.67×10-09 1.82×10-08
5 CH3CHCOOH** + * → CH3CHCO** + OH* 5.26×10-13 4.04×10-14 2.85×10-14
6 CH3CHCOOH** + 3* → CH2CHCOOH**** + H* 2.47×10-17 2.92×10-15 2.53×10-16
7 CH3CHCOOH** + 2*→CH3CCOOH*** + H* 7.63×10-13 1.24×10-12 1.87×10-13
8 CH3CHCO** + 4* → CH3CH*** + CO*** 2.91×10-08 5.84×10-11 8.26×10-10
9 CH3CHCO** + 3* → CH3CCO**** + H* 2.75×10-09 6.61×10-09 1.73×10-08
10 CH3CHCO** + 3* → CH2CHCO**** + H* 8.35×10-13 8.27×10-14 4.01×10-13
34

11 CH2CHCOOH**** + * → CH2CHCO**** + OH* 2.97×10-20 5.83×10-20 2.74×10-20


12 CH2CHCOOH**** + * → CHCHCOOH**** + H* 2.21×10-17 2.89×10-15 2.44×10-16
13 CH3CCOOH*** + 2* → CH3CCO**** + OH* 1.62×10-18 3.96×10-16 1.74×10-17
14 CH3CCO**** + 2* → CH3C*** +CO*** 2.75×10-09 6.61×10-09 1.73×10-08
15 CH2CHCO**** + 2* → CH2CH*** + CO*** 2.89×10-16 4.16×10-16 1.58×10-15
16 CH2CHCO**** + * → CHCHCO**** + H* 8.35×10-13 8.23×10-14 4.00×10-13
17 CHCHCOOH**** + * → CHCHCO**** + OH* 2.21×10-17 2.89×10-15 2.44×10-16
18 CHCHCO**** + 3* → CHCH**** + CO*** 8.35×10-13 8.52×10-14 4.00×10-13
19 CH2CH*** + 2* → CHCH**** + H* -8.35×10-13 -8.52×10-14 -4.00×10-13
20 CH2CH2** + 2* → CH2CH*** + H* -2.40×10-10 -4.76×10-10 -7.02×10-10
21 CH2CH*** + * → CH2C*** + H* -1.42×10-10 -3.74×10-10 -5.64×10-10
22 CH3C*** + * → CH2C*** + H* 1.42×10-10 3.74×10-10 5.64×10-10
23 CH3CH*** + * → CH2CH*** + H* 9.72×10-11 1.03×10-10 1.38×10-10
24 CH3CH*** + * → CH3C*** + H* -2.61×10-09 -6.24×10-09 -1.68×10-08
25 CH3CH2** + 2* → CH3CH*** + H* -3.17×10-08 -6.21×10-09 -1.75×10-08
26 CH3CH3* + 2* → CH3CH2** + H* -4.50×10-13 -2.58×10-14 -1.29×10-13
27 CH3CH2** + * → CH2CH2** + H* 3.43×10-08 6.24×10-09 1.77×10-08
28 CH3CH2COOH* + 2* → CH3CH2COO** + H* 3.73×10-14 5.68×10-12 1.35×10-13
29 CH3CH2COO** + * → CH3CH2** + CO2* 1.23×10-14 1.45×10-13 2.90×10-14
30 CH3CH2COO** + 2* → CH3CHCOO*** + H* 1.81×10-14 5.63×10-12 1.19×10-13
31 CH3CHCOOH** + 2* → CH3CHCOO*** + H* 7.00×10-14 7.70×10-12 1.88×10-13
32 CH3CHCOOH** + 3* → CH3CH*** + COOH** 7.90×10-15 1.00×10-15 2.41×10-16
33 CH3CHCOO*** + * → CH3CH*** + CO2* 7.52×10-14 1.28×10-11 2.90×10-13
34 CH3CHCOO*** + * → CH3CCOO*** + H* 1.29×10-14 5.50×10-13 1.76×10-14
35 CH3CCOOH*** + * → CH3CCOO*** + H* 1.06×10-15 4.70×10-14 7.65×10-16
36 CH3CCOOH*** + 2* → CH3C*** + COOH** 7.62×10-13 1.19×10-12 1.87×10-13
37 CH2CHCOOH**** + * → CH2CH*** + COOH** 2.61×10-18 2.76×10-17 9.43×10-18
38 CH3CCOO*** + * → CH3C*** + CO2* 1.39×10-14 5.97×10-13 1.83×10-14
39 COOH** → CO2* + H* -2.18×10-08 -1.03×10-07 -3.22×10-08
40 COOH** + 2* → CO*** + OH* 2.18×10-08 1.03×10-07 3.22×10-08
41 H2O* + * → OH* + H* -5.63×10-08 -1.10×10-07 -5.07×10-08
35

42 CH3CH3 + * → CH3CH3* -4.50×10-13 -2.58×10-14 -1.29×10-13


43 CH2CH2 + 2* → CH2CH2** -3.46×10-08 -6.71×10-09 -1.84×10-08
44 H2O + * → H2O* -5.64×10-08 -1.10×10-07 -5.07×10-08
45 CO2 + * → CO2* 2.18×10-08 1.03×10-07 3.22×10-08
46 CHCH + 4* → CHCH**** -5.87×10-23 -6.28×10-24 -2.99×10-23
47 CO + 3* → CO*** -5.63×10-08 -1.10×10-07 -5.07×10-08
48 H2 + 2* → H* + H* 2.18×10-08 1.03×10-07 3.22×10-08
Figure 2.1: Reaction network for the HDO of PAC to ethane and ethylene. Blue solid
arrows belong to the decarbonylation pathway, green dash-dotted arrows depict the
decarboxylation pathway, and yellow dashed arrows show those steps that are in common
between the decarboxylation and the decarbonylation mechanism. In addition, the number
on each arrow illustrates the corresponding elementary reaction in Table 2.2, and the
dashed circles show the corresponding cleavage of fluid phase propanoic acid.

36
Figure 2.2: Schematic representation of TOF (s −1 ), of various elementary reactions
involved in the dominant pathways. The left dashed box illustrates the dominant pathway
in the gas phase while the right blue drop displays the dominant pathways in the solvent
environments. Numbers in parentheses () depict reaction step numbers. The numbers in
square brackets [] in the right blue drop are the TOFs (s −1) in liquid water and the numbers
without square brackets are the TOFs (s−1 ) in liquid 1,4-dioxane.

37
Figure 2.3: Apparent activation energy plot in the temperature range of 473 to 623 K in gas
and condensed phase environments. The apparent activation barrier is 2.38 eV in the gas
phase, 2.71 eV in liquid water, and 2.44 eV in liquid 1,4-dioxane.

38
2.8 Bibliography
1. Simonetti, D. A.; Dumesic, J. A., Catalytic Production of Liquid Fuels from
Biomass‐Derived Oxygenated Hydrocarbons: Catalytic Coupling at Multiple
Length Scales. Catalysis Reviews 2009, 51 (3), 441-484.

2. He, Z.; Wang, X., Hydrodeoxygenation of model compounds and catalytic


systems for pyrolysis bio-oils upgrading. Catalysis for Sustainable Energy 2012,
1, 28-52.

3. Bykova, M. V.; Ermakov, D. Y.; Kaichev, V. V.; Bulavchenko, O. A.; Saraev, A.


A.; Lebedev, M. Y.; Yakovlev, V. А., Ni-based sol–gel catalysts as promising
systems for crude bio-oil upgrading: Guaiacol hydrodeoxygenation study. Applied
Catalysis B: Environmental 2012, 113-114, 296-307.

4. Ardiyanti, A. R.; Bykova, M. V.; Khromova, S. A.; Yin, W.; Venderbosch, R. H.;
Yakovlev, V. A.; Heeres, H. J., Ni-Based Catalysts for the Hydrotreatment of Fast
Pyrolysis Oil. Energy & Fuels 2016, 30 (3), 1544-1554.

5. Yin, W.; Kloekhorst, A.; Venderbosch, R. H.; Bykova, M. V.; Khromova, S. A.;
Yakovlev, V. A.; Heeres, H. J., Catalytic hydrotreatment of fast pyrolysis liquids
in batch and continuous set-ups using a bimetallic Ni–Cu catalyst with a high
metal content. Catalysis Science & Technology 2016, 6 (15), 5899-5915.

6. Ardiyanti, A. R.; Khromova, S. A.; Venderbosch, R. H.; Yakovlev, V. A.; Heeres,


H. J., Catalytic hydrotreatment of fast-pyrolysis oil using non-sulfided bimetallic
Ni-Cu catalysts on a δ-Al2O3 support. Applied Catalysis B: Environmental 2012,
117-118, 105-117.

7. Arend, M.; Nonnen, T.; Hoelderich, W. F.; Fischer, J.; Groos, J., Catalytic
deoxygenation of oleic acid in continuous gas flow for the production of diesel-
like hydrocarbons. Applied Catalysis A: General 2011, 399 (1), 198-204.

8. Wan, H.; Chaudhari, R. V.; Subramaniam, B., Aqueous Phase Hydrogenation of


Acetic Acid and Its Promotional Effect on p-Cresol Hydrodeoxygenation. Energy
& Fuels 2013, 27 (1), 487-493.

9. Saleheen, M.; Verma, A. M.; Mamun, O.; Lu, J.; Heyden, A., Investigation of
solvent effects on the hydrodeoxygenation of guaiacol over Ru catalysts.
Catalysis Science & Technology 2019, 9 (22), 6253-6273.

39
10. Saleheen, M.; Zare, M.; Faheem, M.; Heyden, A., Computational Investigation of
Aqueous Phase Effects on the Dehydrogenation and Dehydroxylation of Polyols
over Pt(111). The Journal of Physical Chemistry C 2019, 123 (31), 19052-19065.

11. Car, R.; Parrinello, M., Unified Approach for Molecular Dynamics and Density-
Functional Theory. Physical Review Letters 1985, 55 (22), 2471-2474.

12. Carloni, P.; Rothlisberger, U.; Parrinello, M., The Role and Perspective of Ab
Initio Molecular Dynamics in the Study of Biological Systems. Accounts of
Chemical Research 2002, 35 (6), 455-464.

13. Faheem, M.; Suthirakun, S.; Heyden, A., New Implicit Solvation Scheme for
Solid Surfaces. The Journal of Physical Chemistry C 2012, 116 (42), 22458-
22462.

14. Lu, J.; Behtash, S.; Heyden, A., Theoretical Investigation of the Reaction
Mechanism of the Decarboxylation and Decarbonylation of Propanoic Acid on
Pd(111) Model Surfaces. The Journal of Physical Chemistry C 2012, 116 (27),
14328-14341.

15. Behtash, S.; Lu, J.; Faheem, M.; Heyden, A., Solvent effects on the
hydrodeoxygenation of propanoic acid over Pd(111) model surfaces. Green
Chemistry 2014, 16 (2), 605-616.

16. Lu, J.; Faheem, M.; Behtash, S.; Heyden, A., Theoretical investigation of the
decarboxylation and decarbonylation mechanism of propanoic acid over a
Ru(0001) model surface. Journal of Catalysis 2015, 324, 14-24.

17. Yang, W.; Solomon, R. V.; Lu, J.; Mamun, O.; Bond, J. Q.; Heyden, A.,
Unraveling the mechanism of the hydrodeoxygenation of propionic acid over a Pt
(1 1 1) surface in vapor and liquid phases. Journal of Catalysis 2020, 381, 547-
560.

18. Behtash, S.; Lu, J.; Mamun, O.; Williams, C. T.; Monnier, J. R.; Heyden, A.,
Solvation Effects in the Hydrodeoxygenation of Propanoic Acid over a Model
Pd(211) Catalyst. The Journal of Physical Chemistry C 2016, 120 (5), 2724-2736.

19. Kresse, G.; Furthmüller, J., Efficiency of ab-initio total energy calculations for
metals and semiconductors using a plane-wave basis set. Computational
Materials Science 1996, 6 (1), 15-50.

40
20. Kresse, G.; Hafner, J., Ab initio molecular dynamics for liquid metals. Physical
Review B 1993, 47 (1), 558-561.

21. Perdew, J. P.; Wang, Y., Accurate and simple analytic representation of the
electron-gas correlation energy. Physical Review B 1992, 45 (23), 13244-13249.

22. Perdew, J. P.; Yue, W., Accurate and simple density functional for the electronic
exchange energy: Generalized gradient approximation. Physical Review B 1986,
33 (12), 8800-8802.

23. Henkelman, G.; Uberuaga, B. P.; Jónsson, H., A climbing image nudged elastic
band method for finding saddle points and minimum energy paths. The Journal of
Chemical Physics 2000, 113 (22), 9901-9904.

24. Henkelman, G.; Jónsson, H., A dimer method for finding saddle points on high
dimensional potential surfaces using only first derivatives. The Journal of
Chemical Physics 1999, 111 (15), 7010-7022.

25. Heyden, A.; Bell, A. T.; Keil, F. J., Efficient methods for finding transition states
in chemical reactions: Comparison of improved dimer method and partitioned
rational function optimization method. The Journal of Chemical Physics 2005,
123 (22), 224101.

26. Ahlrichs, R.; Bär, M.; Häser, M.; Horn, H.; Kölmel, C., Electronic structure
calculations on workstation computers: The program system turbomole. Chemical
Physics Letters 1989, 162 (3), 165-169.

27. Treutler, O.; Ahlrichs, R., Efficient molecular numerical integration schemes. The
Journal of Chemical Physics 1995, 102 (1), 346-354.

28. TURBOMOLE V6.0 2009, a development of University of Karlsruhe and


Forschungszentrum Karlsruhe GmbH, 1989−2007, TURBOMOLE GmbH, since
2007.

29. Weigend, F., Accurate Coulomb-fitting basis sets for H to Rn. Physical Chemistry
Chemical Physics 2006, 8 (9), 1057-1065.

30. Weigend, F.; Ahlrichs, R., Balanced basis sets of split valence, triple zeta valence
and quadruple zeta valence quality for H to Rn: Design and assessment of
accuracy. Physical Chemistry Chemical Physics 2005, 7 (18), 3297-3305.

41
31. Weigend, F.; Häser, M.; Patzelt, H.; Ahlrichs, R., RI-MP2: optimized auxiliary
basis sets and demonstration of efficiency. Chemical Physics Letters 1998, 294
(1), 143-152.

32. Eichkorn, K.; Treutler, O.; Öhm, H.; Häser, M.; Ahlrichs, R., Auxiliary basis sets
to approximate Coulomb potentials. Chemical Physics Letters 1995, 240 (4), 283-
290.

33. Eichkorn, K.; Weigend, F.; Treutler, O.; Ahlrichs, R., Auxiliary basis sets for
main row atoms and transition metals and their use to approximate Coulomb
potentials. Theoretical Chemistry Accounts 1997, 97 (1), 119-124.

34. Von Arnim, M.; Ahlrichs, R., Performance of parallel TURBOMOLE for density
functional calculations. Journal of Computational Chemistry 1998, 19 (15), 1746-
1757.

35. Klamt, A., Conductor-like Screening Model for Real Solvents: A New Approach
to the Quantitative Calculation of Solvation Phenomena. The Journal of Physical
Chemistry 1995, 99 (7), 2224-2235.

36. Klamt, A.; Jonas, V.; Bürger, T.; Lohrenz, J. C. W., Refinement and
Parametrization of COSMO-RS. The Journal of Physical Chemistry A 1998, 102
(26), 5074-5085.

37. Marsh, K. N., COSMO-RS from Quantum Chemistry to Fluid Phase


Thermodynamics and Drug Design. By A. Klamt. Elsevier: Amsterdam, The
Netherlands, 2005. 246 pp. $US 165. ISBN 0-444-51994-7. Journal of Chemical
& Engineering Data 2006, 51 (4), 1480-1480.

38. Mathew, K.; Sundararaman, R.; Letchworth-Weaver, K.; Arias, T. A.; Hennig, R.
G., Implicit solvation model for density-functional study of nanocrystal surfaces
and reaction pathways. The Journal of Chemical Physics 2014, 140 (8), 084106.

39. Fishman, M.; Zhuang, H. L.; Mathew, K.; Dirschka, W.; Hennig, R. G., Accuracy
of exchange-correlation functionals and effect of solvation on the surface energy
of copper. Physical Review B 2013, 87 (24), 245402.

40. Fernández, D. P.; Goodwin, A. R. H.; Lemmon, E. W.; Sengers, J. M. H. L.;


Williams, R. C., A Formulation for the Static Permittivity of Water and Steam at
Temperatures from 238 K to 873 K at Pressures up to 1200 MPa, Including
Derivatives and Debye–Hückel Coefficients. Journal of Physical and Chemical
Reference Data 1997, 26 (4), 1125-1166.

42
41. Lugo-José, Y. K.; Monnier, J. R.; Heyden, A.; Williams, C. T.,
Hydrodeoxygenation of propanoic acid over silica-supported palladium: effect of
metal particle size. Catalysis Science & Technology 2014, 4 (11), 3909-3916.

42. Lugo-José, Y. K.; Behtash, S.; Nicholson, M.; Monnier, J. R.; Heyden, A.;
Williams, C. T., Unraveling the mechanism of propanoic acid
hydrodeoxygenation on palladium using deuterium kinetic isotope effects.
Journal of Molecular Catalysis A: Chemical 2015, 406, 85-93.

43. Zhang, X.; Savara, A.; Getman, R. B., A Method for Obtaining Liquid–Solid
Adsorption Rates from Molecular Dynamics Simulations: Applied to Methanol on
Pt(111) in H2O. Journal of Chemical Theory and Computation 2020.

44. Demir, B.; Kropp, T.; Rivera-Dones, K. R.; Gilcher, E. B.; Huber, G. W.;
Mavrikakis, M.; Dumesic, J. A., A self-adjusting platinum surface for acetone
hydrogenation. Proceedings of the National Academy of Sciences 2020, 117 (7),
3446.

45. McFarland, E. W.; Metiu, H., Catalysis by Doped Oxides. Chemical Reviews
2013, 113 (6), 4391-4427.

46. I. Chorkendorff, J. W. N., Concepts of Modern Catalysis and Kinetics. third ed.;
Wiley-VCH, Weinheim: 2005.

47. Desai, S. K.; Neurock, M., First-principles study of the role of solvent in the
dissociation of water over a Pt-Ru alloy. Phys Rev B 2003, 68 (7), 075420:1-7.

48. Hibbitts, D. D.; Loveless, B. T.; Neurock, M.; Iglesia, E., Mechanistic Role of
Water on the Rate and Selectivity of Fischer-Tropsch Synthesis on Ruthenium
Catalysts. Angew. Chem. Int. Ed. Engl. 2013, 52 (47), 12273-12278.

49. Zhang, X. H.; Sewell, T. E.; Glatz, B.; Sarupria, S.; Getman, R. B., On the water
structure at hydrophobic interfaces and the roles of water on transition-metal
catalyzed reactions: A short review. Catal Today 2017, 285, 57-64.

50. Santana, J. A.; Mateo, J. J.; Ishikawa, Y., Electrochemical Hydrogen Oxidation on
Pt(110): A Combined Direct Molecular Dynamics/Density Functional Theory
Study. J Phys Chem C 2010, 114 (11), 4995-5002.

51. Skachkov, D.; Rao, C. V.; Ishikawa, Y., Combined First-Principles Molecular
Dynamics/Density Functional Theory Study of Ammonia Electrooxidation on
Pt(100) Electrode. J Phys Chem C 2013, 117 (48), 25451-25466.

43
52. Nie, X. W.; Luo, W. J.; Janik, M. J.; Asthagiri, A., Reaction mechanisms of CO2
electrochemical reduction on Cu(111) determined with density functional theory.
J Catal 2014, 312, 108-122.

53. Huang, Z. Q.; Long, B.; Chang, C. R., A theoretical study on the catalytic role of
water in methanol steam reforming on PdZn(111). Catal Sci Technol 2015, 5 (5),
2935-2944.

54. Smith, J. M.; Ness, H. C. V.; Abbott, M. M., Introduction to Chemical


Engineering Thermodynamics. McGraw-Hill: New York, 2005.

55. Lugo-José, Y. K.; Monnier, J. R.; Williams, C. T., Gas-phase, catalytic


hydrodeoxygenation of propanoic acid, over supported group VIII noble metals:
Metal and support effects. Applied Catalysis A: General 2014, 469, 410-418.

56. Campbell, C. T., Future Directions and Industrial Perspectives Micro- and macro-
kinetics: Their relationship in heterogeneous catalysis. Topics in Catalysis 1994, 1
(3), 353-366.

57. Campbell, C. T., Finding the Rate-Determining Step in a Mechanism: Comparing


DeDonder Relations with the “Degree of Rate Control”. Journal of Catalysis
2001, 204 (2), 520-524.

58. Stegelmann, C.; Andreasen, A.; Campbell, C. T., Degree of Rate Control: How
Much the Energies of Intermediates and Transition States Control Rates. Journal
of the American Chemical Society 2009, 131 (23), 8077-8082.

59. Kozuch, S.; Shaik, S., Kinetic-Quantum Chemical Model for Catalytic Cycles:
The Haber−Bosch Process and the Effect of Reagent Concentration. The Journal
of Physical Chemistry A 2008, 112 (26), 6032-6041.

60. Kozuch, S.; Shaik, S., A Combined Kinetic−Quantum Mechanical Model for
Assessment of Catalytic Cycles: Application to Cross-Coupling and Heck
Reactions. Journal of the American Chemical Society 2006, 128 (10), 3355-3365.

44
CHAPTER 3

DEPENDENCY OF SOLVATION EFFECTS IN METAL IDENTITY IN

SURFACE REACTIONS

Zare, M.; Saleheen, M.; Kundu, S. K.; Heyden, A. Communications Chemistry 3, 187,
2020.

Reprinted here with the permission of the publisher

45
3.1 Abstract
Solvent interactions with adsorbed moieties involved in surface reactions are often

believed to be similar for different metal surfaces. However, solvents alter the electronic

structures of surface atoms, which in turn affects their interaction with adsorbed moieties.

To reveal the importance of metal identity on aqueous solvent effects in heterogeneous

catalysis, we studied solvent effects on the activation free energies of the O-H and C-H

bond cleavages of ethylene glycol over the (111) facet of six transition metals (Ni, Pd, Pt,

Cu, Ag, Au) using an explicit solvation approach based on a hybrid quantum

mechanical/molecular mechanical (QM/MM) description of the potential energy surface.

A significant metal dependence on aqueous solvation effects was observed that suggests

solvation effects must be studied in detail for every reaction system. The main reason for

this dependence could be traced back to a different amount of charge-transfer between the

adsorbed moieties and metals in the reactant and transition states for the different metal

surfaces.

3.2 Introduction
The widespread use of solvents in applications varying from pharmaceutical1, 2, 3, 4,
5, 6
to electrochemistry7, 8, 9, 10 and catalysis11, 12, 13, 14, 15, 16 has given rise to an extensive

range of studies aimed at understanding and predicting the role of solvents. The concept

that a solvent can alter the performance of a catalyst, including its rate, selectivity, and

stability, is well known; yet predicting a specific solvation effect remains a challenge.17, 18,
19, 20, 21, 22, 23, 24
While impressive progress has been made in understanding the effects of

solvents in homogenous catalysis,25, 26, 27, 28, 29 for heterogeneously catalyzed processes that

benefit from easier separation of the catalyst relative to homogeneously catalyzed

46
processes,30 the role of solvents is hardly understood and only rarely studied. The inherent

complexity of a reaction system containing both a complex heterogeneous catalyst and a

condensed phase at a finite, often elevated, temperature has resulted in only few systematic

experimental (in situ and in operando)13, 15, 31, 32 and/or theoretical studies11, 12, 33, 34 of

solvation effects in heterogenous catalysis.

The role of solvents in catalytic transformations occurring at a solid-liquid interface

is typically ascribed to: heightened importance of mass transfer effects, nature of solvent

(polarity etc.)35, 36, 37, competitive adsorption between solvent molecules and adsorbed

moieties,32, 38, 39
direct participation of the solvent in the reaction coordinate,40 and/or

relative stabilization of reactant, transition and/or product state of elementary reactions.41,


42, 43
These effects in turn can lead to a change in reaction mechanism, reaction kinetics,

selectivity, and overall catalyst lifetime. In short, understanding and predicting solvent

effects on surface reactions requires detailed investigations of the direct and indirect

interactions between the solvent, catalyst, and reacting moieties on the surface under

reaction conditions.

In this regard, theoretical calculations have the advantage of being able to

systemically study the effect of an individual parameter on the effect of a solvent. Ab initio

molecular dynamics (AIMD) simulations have been used;44, 45, 46 however, due to the great

computational cost associated with the quantum mechanical calculations and the large

amount of phase space sampling necessary, AIMD simulations are currently limited to

simulation systems of a few hundred atoms and a time scale of tens or a few hundred

picoseconds.10, 47, 48

47
An alternative approach is to use implicit solvation models.49, 50 While they can

compute free energies of reactions at solid-liquid interfaces rapidly, their reliability has

often been questioned14 because of their inability to capture the anisotropic site-specific

interactions between the solute and the solvent molecules. A compromise in efficiency and

accuracy constitutes a combined quantum mechanical/molecular mechanical (QM/MM)

approach.51, 52, 53 In this class of simulations, the adsorbate and metal atoms involved in the

reaction are considered as a QM sub-system described from first principles, while the bulk

of the solvent and metal atoms distant to the active site are considered as an MM sub-

system described using classical molecular mechanics force fields. We have previously

developed such a hybrid QM/MM model, named eSMS (Explicit Solvation Model for

Metal Surfaces),54 that considers the long-range electrostatic interaction of the solvent

molecules in the electronic structure calculation of the active site and applied it to the free

energy calculation of the initial dehydrogenation and dehydroxylation of an adsorbed

ethylene glycol (EG) moiety on Pt(111) in the presence of liquid water.55

To design a liquid-phase surface-catalyzed reaction system for enhanced activity

and selectivity, the interrelation of energetic changes on variation of catalyst surface, nature

of solvent, and reacting moiety must be disclosed. In this context, several studies have been

dedicated to understanding how the nature of the solvent and reacting moiety direct

catalysis.23, 26, 56, 57, 58 Nonetheless, the role of the metal identity (catalyst) on the solvent

effect has to our knowledge not been investigated yet. While it could be argued that

solvation effects should be similar for the same bond cleavages or for the same adsorbates

on different metal surfaces such as in recent studies by Greely et al.59, 60, we hypothesize

that the electronic structure modification of the metal surface and reacting moiety as a

48
result of the nearby solvent is sufficiently significant that solvation effects can differ

significantly for different metals. To confirm this hypothesis, we have investigated the

aqueous-phase effects on the initial C-H and O-H bond cleavages of EG over the (111)

facet of six transition metal surfaces (Ni, Pd, Pt, Cu, Ag, Au) using our explicit solvation

method, eSMS. The choice of reaction systems is motivated by (i) EG being a commonly

studied surrogate molecule of biomass-derived polyols, (ii) the selected transition metals

are relatively stable and commonly used for aqueous-phase processing of biomass-derived

oxygenates,61 (iii) early dehydrogenation steps of EG over Pt and Ni/Pt catalysts have

previously been found to control the overall reaction rate,62, 63, 64 (iv) at least over Pt(111)

in the vapor phase, initial C-H and O-H bond cleavage are competitive (although O-H bond

cleavage is believed to be somewhat favored),34, 64


and finally (v), explicit solvation

approaches have recently been used to demonstrate that for bond cleavage reactions of

alcohols over Pt(111), aqueous solvation effects are large and can currently not be

described by implicit solvation models.55, 65

3.3 Results and Discussion


Figure 3.1 illustrates the aqueous-phase effects on the activation free energy barrier

of the O-H (CH2OHCH2OH** + * ↔ CH2OCH2OH** + H*) and C-H bond cleavages

(CH2OHCH2OH** + * ↔ CHOHCH2OH ** + H*) of EG at 423 K computed by eSMS

(see Table 3.1 for specific numbers). In addition, a graphical representation of the free-

energy (potential of mean force) profiles for the O-H bond cleavages is illustrated in Figure

3.2 and for the C-H bond cleavages in Figure B.1. We note that although some of the metals

(such as Ni) might get partially oxidized in liquid water environments with low reduction

potential, we chose to study the (111) facet of all metal surfaces for better comparison.

49
Generally, both O-H and C-H bond cleavages are somewhat facilitated in the presence of

water. For the O-H bond cleavage, Pt is the most active catalyst in both phases. However,

for the C-H bond cleavage, Ni is the most active catalyst in the presence of liquid water

because of strong aqueous-phase effect that is more than twice as large over Ni than Pt.

Interestingly, in liquid water Cu is predicted to be as active for O-H bond cleavage

(∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 = 0.50 𝑒𝑉) as Ni (∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 = 0.55 𝑒𝑉) and Pd (∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 = 0.58 𝑒𝑉). In short,

evidenced by very different aqueous-phase effects on the activation free energy barriers

across different metals, our hypothesis that the nature of metal plays a key role for solvation

effects on surface reactions has been confirmed.

3.3.1 Origin of dependency of solvent effect on metal identity


A heterogeneously catalyzed reaction occurring at a catalyst-solvent interface is at

least a three-body problem involving solvent, catalyst, and reacting moiety. Thus, the most

important factors contributing to solvation effects in such catalytic transformation arise

from direct and indirect interactions of solvent, catalyst, and reacting moiety. To explain

the origin for the variability of aqueous-phase effects on the free energy of activation

(∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) of O-H and C-H bond cleavages of EG across six

transition metals, we investigated some intuitive, physics-based descriptors.

One possible descriptor that could explain the role of the metal identity, which leads

to different ∆∆𝐺 𝑎𝑐𝑡 values across different metals and bond cleavages, is the difference in

charge-transfer effects (and thus solvent stabilization) in the reactant (RS) and transition

states (TS) (see Table B.5). This charge transfer effect is attributed to the indirect influence

of the water environment causing an electronic modification of the metal atoms and the

effective charge distribution in the reacting moieties. We note that, using AIMD

50
simulations, Siemer et al.17 have recently shown that water induced local charge transfer is

a leading contribution for the observed solvation effect on O2 activation at Au/TiO2

interface sites. We choose two suitable descriptors for quantifying this charge-transfer

effect. One is the cleaving-bond charge transfer (BC) defined as the change in sum of

partial NPA66 charges on the cleaving bond going from RS to TS; for example in O-H bond

cleavage: 𝐵𝐶 = |𝑄 𝑂 + 𝑄 𝐻 |𝑅𝑆 − |𝑄 𝑂 + 𝑄 𝐻 |𝑇𝑆 . The idea behind this formula is that a

change in charge from, e.g., -0.1 to +0.1, +0.1 to +0.1, or -0.1 to -0.1, should all result in

no significant net stabilization since water stabilizes the RS and TS similarly. Another

descriptor describing a charge-transfer effect is the molecular charge transfer (MC) defined

here as the change in absolute sum of charges on the reacting moiety going from RS to TS,

i.e., this descriptor describes the change in charge transfer from the metal surface to the

reacting moiety when going from the RS to the TS.

Table 3.2 lists BC and MC for the O-H and C-H bond cleavages over the (111)

facet of six transition metal surfaces (see Table B.5 for partial charges). According to our

definition of BC and MC, smaller BC values (more negative) or larger MC values (more

positive) correspond to an increased charge-transfer effect when going from reactant to

transition state. For instance, in the C-H bond cleavage, the charge-transfer effect over

Cu(111) (𝐵𝐶 = −0.46 𝑒 , 𝑀𝐶 = 0.53 𝑒) is higher than over Pt(111) (𝐵𝐶 =

−0.03 𝑒 , 𝑀𝐶 = 0.08 𝑒). As shown in Table 3.2, ∆∆𝐺 𝑎𝑐𝑡 is directly related to the charge-

transfer effect descriptors in the C-H bond cleavage; that is, solvent effects increase (more

negative ∆∆𝐺 𝑎𝑐𝑡 ) with increasing charge-transfer effect (more negative BC or more

positive MC). In contrast, for O-H bond cleavage, there is no linear relationship between

∆∆𝐺 𝑎𝑐𝑡 and the charge-transfer effect. Nevertheless, after investigating other solvation

51
effect descriptors, we will show that the charge-transfer effect is likely also here a key

descriptor for the solvent effect across metal surfaces.

Another commonly used criterion for describing adsorbate-solvent interaction in an

aqueous phase is hydrogen bonding.23, 41 Two main classes of hydrogen bonding definitions

commonly used in the literature are based on an energy criterion and a geometric definition.

Herein, we employed a geometric definition in which a hydrogen bond exists if the distance

between the donor oxygen (Od) and the acceptor oxygen (Oa), ROO, is less than 3.2 Å and

the angle ∠HOdOa is smaller than 20⁰67 (see Figure B.2). We note that EG with OH

functional groups can be either a donor or acceptor of hydrogen bonding (see Table B.4).

Hence, the change in mean of total hydrogen bonding (acceptor + donor) going from RS

to TS was chosen as a descriptor, named in the following H-bond, and is included in Table

3.2. Next, the gas-phase free energy of activation (∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ), a rough measure of change

in surface-adsorbate interaction going from reactant to transition state was selected as a

descriptor for describing the variability in solvation free energy effects across metal

surfaces. Given the absence of a linear ∆∆𝐺 𝑎𝑐𝑡 dependence on H-bond or ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 , we

also studied all pairwise combinations of descriptors. Finally, we note here that we also

analyzed the water orientation (H-up, H-down, parallel – see Figure B.3) within the first

water layer of the different surfaces in the reactant state (see height distribution function of

water O in Figure B.4); however, we did not observe any significant variation in water

orientations across metals, explaining why we did not further study water orientation as a

descriptor (see Table B.10). Also, we attempted to use the standard electrode potential of

the metal elements as descriptor but again no meaningful correlation could be obtained

such that it is not further discussed.

52
Next, we first examined the pairwise correlations between the descriptors using the

Pearson correlation coefficient (PCC).68 The results (see Table B.6) indicate that BC and

MC are totally correlated (PCC ~ -1.0) which is expected given that they both describe

charge transfer. In addition, H-bond is more correlated with MC and BC in the O-H bond

cleavage than in the C-H bond cleavage. We attribute this to the fact that hydrogen bonding

is obtained from the interaction of water molecules with the OH functional groups of EG,

and the O-H bond cleavage reaction causes a significant change in the charges on one of

the two O-H functional groups in EG in the TS. Finally, ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 is hardly correlated with

BC, MC or H-bond in both cleavages.

Second, using a simple linear model (equation 3.1), we investigated the relation

between the descriptors and the aqueous-phase effects on the free energies of activation

(∆∆𝐺 𝑎𝑐𝑡 ). We emphasize that we do not intend to quantitatively predict ∆∆𝐺 𝑎𝑐𝑡 using a set

of descriptors. Instead, we employ models based on physics-based descriptors to explain

what physical phenomena could explain the changes in solvation effects on the kinetics of

the studied surface-catalyzed reactions.

̅̅̅̅̅̅̅̅̅
(∆∆𝐺 𝑎𝑐𝑡 − ∆∆𝐺 𝑎𝑐𝑡 ) ̅ ̅
𝑚𝑜𝑑𝑒𝑙 = 𝛼1 (𝑓1 − 𝑓1 ) + 𝛼2 (𝑓2 − 𝑓2 ) (3.1)

In equation 3.1, bar signs show the mean of the corresponding variable, 𝛼1 and 𝛼2

are model parameters, and 𝑓1 and 𝑓2 represent descriptors (see Table 3.2). The best-fitting

parameters and mean absolute error (𝑀𝐴𝐸) of the linear model for different combinations

of descriptors are listed in Table B.7. Interestingly, the linear model can estimate ∆∆𝐺 𝑎𝑐𝑡

very well for the C-H bond cleavage but not for the O-H bond cleavage. As expected, the

dominant factor in C-H bond cleavage is the charge-transfer effect; that is, 𝛼2 ≫ 𝛼1 when

one of charge-transfer effect descriptors (BC or MC) used as 𝑓2 (see Table B.7). Finally, a

53
quadratic model with two descriptors was employed to explain ∆∆𝐺 𝑎𝑐𝑡 of the O-H bond

cleavage.

̅̅̅̅̅̅̅̅̅
(∆∆𝐺 𝑎𝑐𝑡 − ∆∆𝐺 𝑎𝑐𝑡 ) ̅ ̅ ̅ 2 ̅ 2
𝑚𝑜𝑑𝑒𝑙 = 𝛼1 (𝑓1 − 𝑓1 ) + 𝛼2 (𝑓2 − 𝑓2 ) + 𝛼3 (𝑓1 − 𝑓1 ) + 𝛼4 (𝑓2 − 𝑓2 ) (3.2)

The results of this model for different combinations of descriptors are included in Table

B.8.

Overall, when comparing the model parameters of the linear and quadratic fits, we

conclude that the aqueous-phase effect on the free energy of activation of O-H and C-H

bond cleavages of EG over the investigated metal surfaces originates primarily from some

form of a charge-transfer effect (BC or MC) (more charge transfer in the TS relative to the

RS leads to more stabilization) and partly on another descriptor that could be related to H-

bond and ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 (larger barriers are more stabilized). Figure 3.3 presents the computed

aqueous-phase effect by eSMS versus the estimated one from a linear fit for C-H bond

cleavage and a quadratic fit for O-H bond cleavage, in which BC and ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 were used

as descriptors.

3.3.2 Comparison to implicit solvation models


Finally, the implicit solvation models (iSMS69 and VASPsol70, 71) failed to capture

the full solvent stabilization during the O-H bond cleavage of EG (see Table 3.1). For the

C-H bond cleavage, implicit and explicit solvation models anticipate comparable aqueous-

phase effects for the metals that display a smaller charge transfer effect (Pd, Pt, Au). Table

B.9 illustrates that this observation holds even when considering typical uncertainties in

cavity radius of ±10% for transition metal elements. In short, the reliability of implicit

solvation calculations for heterogeneous (metal) catalysis applications is currently limited

54
(unknown) due to the very limited availability of experimental data that can be used in the

parameterization of the implicit solvation models. This appears to be currently an

advantage for explicit solvation models that rely “only” on a meaningful potential energy

description.

3.4 Conclusions
In summary, we hypothesized that since solvents can modify the electronic

structure of a metal surface, thereby affecting the stability of reacting moieties, the metal

identity plays a significant role for solvation effects on elementary reactions. We

investigated aqueous-phase effects on the free energy of activation of the initial O-H and

C-H bond cleavages of ethylene glycol over the (111) surface of six transition metals,

including Ni, Pd, Pt, Cu, Ag, and Au, to disclose the role of metal identity on solvation

effects. To compute the free energy of activation in the presence of water, we utilized our

explicit solvation approach, named eSMS, which is based on a hybrid quantum

mechanical/molecular mechanical (QM/MM) description of the potential energy surface.

Our hypothesis was confirmed by finding significantly different aqueous-phase effects on

the activation barrier (∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) for the same bond cleavage across

different metals, suggesting that solvation effects have to be studied in detail for any

specific reaction system. Subsequently, by introducing three intuitive, physics-based

descriptors, the charge-transfer effect, hydrogen bonding, and the gas-phase activation free

energy barrier, and studying correlations between the aqueous-phase effects and these

descriptors, we can conclude that the aqueous-phase effects originate primarily from

various charge-transfer effects across the different metals, and to a lesser extent to another

descriptor that could be related to a different amount of hydrogen bonding and the gas-

55
phase activation barriers across the different metals. This observation agrees with previous

AIMD simulations of oxygen activation over Au catalysts.17 Finally, implicit solvation

models are currently not able to capture these charge-transfer effects that can lead to

changes in preference of initial O-H versus C-H bond cleavage in alcohols over various

transition metal surfaces.

3.5 Methods

3.5.1 Planewave DFT Calculations


Vapor-phase DFT calculations were carried out using periodic boundary conditions

as implemented in the Vienna Ab Initio Simulation Package (VASP 5.4.4).72, 73 A frozen-

core, all-electron projector augmented-wave (PAW)74 method was utilized to avoid the

singularities of Kohn-Sham wavefunctions at the nuclear positions. The number of valence

electrons considered for Ni, Pd, Pt, Cu, Ag, Au, C, O, and H are 10 ,10, 10, 11, 11, 11, 4,

6, and 1, respectively. The unknown part of interaction energy between individual

electrons, i.e., the exchange-correlation functional, was expressed using the Perdew-Burke-

Ernzerhof (PBE)75, 76
functional within the semi-local generalized gradient

approximation.77 Brillouin zone integrations have been performed with a 4×4×1

Monkhorst-Pack78 k-point grid and electronic wavefunctions at each k-point were

expanded using a discrete plane-wave basis set with kinetic energies limited to 400 eV. A

first order smearing method (Methfessel-Paxton)79 with 0.10 eV smearing width was

employed, allowing to accurately calculate the entropic contributions due to the smearing.

Dipole and quadrupole corrections (along the surface normal) to the total energy have been

calculated using a modified version of the Makov-Payne80 method, and Harris corrections,

based on the non-self-consistent Harris-Foulkes81, 82 functional, have been applied to the

56
stress-tensor and forces. A 4×4-unit cell with four layers of metal atoms (bottom two layers

fixed in their bulk positions) has been employed. By introducing a 15 Å vacuum on top of

the surface, the interaction between the periodic images along the surface normal has been

curtailed. A self-consistent field (SCF) convergence criterion for the electronic degrees of

freedom of the valence electrons was set to 1.0×10-7 eV. Transition state structures for the

elementary processes were located using a combination of climbing-image nudged elastic

band83, 84 and dimer85, 86 methods. Finally, the minima and the first order saddle points were

validated by computing the Hessian matrix and vibrational spectra. We note that spin-

polarized calculations have been carried out for the Ni surface.

3.5.2 Non-periodic cluster calculations


Cluster model DFT calculations in vacuum have been carried out using the

TURBOMOLE 7.2 program package.87, 88, 89 To model the cluster surfaces, two layers of

metal atoms with a hexagonal shaped geometry (51 atoms) were chosen. The convergence

of the total QM/MM energy with respect to the lateral size and depth of the cluster

geometry can be found elsewhere.54 An improved version of the default TURBOMOLE

basis sets (def-bases) with split valence and polarization functions (def2-SVP)90, 91 were

employed to represent the adsorbate atoms and the metal atoms of Ni and Cu. Furthermore,

Ag, Pd, Pt, and Au atoms were represented using scalar relativistic effective core potentials

(ECPs) in conjunction with split valence basis sets augmented by polarization functions.91,
92
Electron exchange and correlation effects were accounted for by employing the PBE

functional.75, 76 To speed up the calculation as recommended by TURBOMOLE, the RI-J

approximation with auxiliary basis sets was used to approximate the coulomb integrals.93,

57
94
An SCF convergence criterion of 1.0 × 10−7 Hartree was established and a Gauss-

Chebyshev type spherical grid, m4, was employed to perform the numerical integrations.88

3.5.3 Molecular Dynamics (MD) Simulations


MD simulations were carried out using the DL_POLY 4.03 molecular simulation

program package.95 The initial 4×4 unit cell for each metal surface was augmented laterally

to a 16×20 surface with further vacuum added in the Z-direction resulting in a simulation

box comprising of 1280 metal atoms. The simulation box size for each metal is reported in

Table B.1. The simulation box height was selected based on the work from Behler et al.67

finding that simulations of metal-water interfaces should contain a water layer of ~40 Å

height. The experimental saturated liquid water density of ~0.9 g/cm3 at 423 K was

achieved by packing the simulation box of Ni, Pd, Pt, Cu, Ag, and Au with 2258, 2758,

2800, 2398, 3045, and 2985 water molecules, respectively. All metal and adsorbate atoms

were kept fixed while the geometry of water molecules was constricted to that of TIP3P 96

geometry with the RATTLE algorithm97. To solve the Newton’s equations of motion, a

velocity version of the SHAKE algorithm,98 in conjunction with the velocity Verlet (VV)

integrator99 were used. The TIP3P model was employed for the force field parameters of

liquid water while the van der Waals parameters for adsorbate atoms were obtained from

the OPLS force field.100, 101


In addition to the OPLS parameters, the Lennard-Jones

parameters from the Combined B3LYP/6-31_G*/AMBER Potential102 were used for the

hydrogen atoms of the adsorbed moieties. Lennard-Jones parameters for hydrogen atoms

are important in QM/MM optimizations that permit hydrogen atoms to approach water

molecules and leave the protective environment of a neighboring carbon or oxygen atom.

The Lennard-Jones metal potential103 was employed to describe the metal-water

58
interaction. The LJ cross-term of intermolecular parameters were calculated by Lorentz-
𝜎𝑖 +𝜎𝑗
Berthelot mixing rules through equations 𝜎𝑖𝑗 = and 𝜀𝑖𝑗 = √𝜀𝑖 𝜀𝑗 . All Lennard-Jones
2

parameters are included in Table B.2. The charges for the QM atoms were estimated using

the natural population analysis (NPA).66 To describe the interaction of the TIP3P water

point charges with the quantum chemically described cluster model, we employed the

periodic electrostatic embedded cluster method (PEECM)104 as implemented in

TURBOMOLE. Simulations were carried out in a canonical ensemble (NVT) with Nosé-

Hoover thermostat.105, 106 A 1 ps relaxation time constant for temperature fluctuations was

used to maintain the average system temperature. Electrostatic interactions were accounted

for by using the Smoothed Particle Mesh Ewald (SPME) method107 with automatic

parameter optimization for default SPME precision and a 12 Å cutoff radius was adopted

for the van der Waals interactions and the transition between short and long range

electrostatic interactions. Unless specified otherwise, for each free energy perturbation

step, all systems were equilibrated for 250 ps and sampled for 1000 ps (1ns) using a 1 fs

timestep to obtain 1000 MM conformations (1 ps apart). Thus, MD simulations for over

2.9 ms were performed for this study. To optimize structures in an aqueous reaction

environment, we utilized the fixed-size ensemble approximation with 5000 MM

conformations (250 ps equilibration and 5 ns sampling) recorded every 1 ps.

3.5.4 QM/MM Energy Calculation


The QM/MM minimum free energy path (QM/MM-MFEP)52, 53
method for

optimizing the intrinsic reaction coordinate on a potential of mean force (PMF) description

of the reaction system has been implemented in our program packages. A full description

of this methodology, eSMS (Explicit Solvation for Metal Surfaces) can be found

59
elsewhere.54. Briefly, the total energy function formulation of our eSMS method is given

by

𝑆𝑢𝑟𝑓𝑎𝑐𝑒 𝑐𝑙𝑢𝑠𝑡𝑒𝑟 𝑀𝑒𝑎𝑛𝐹𝑖𝑒𝑙𝑑(100)


𝐸𝑇 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 ) = 𝐸𝑄𝑀 (𝑟𝑄𝑀 ) − 𝐸𝑄𝑀 (𝑟𝑄𝑀 ) + 〈𝛹 |𝐻𝑒𝑓𝑓 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 )| 𝛹〉 +
1 𝑒𝑙𝑒𝑐+𝑣𝑑𝑊 𝑀𝑒𝑎𝑛𝐹𝑖𝑒𝑙𝑑(100)
[100 ∑100
𝑗=1 [𝐸𝑗,𝑀𝑀+𝑄𝑀⁄𝑀𝑀 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 )] −
𝑄𝑖 =0,𝑖∈𝑄𝑀
𝑒𝑙𝑒𝑐+𝑣𝑑𝑊 𝑀𝑒𝑎𝑛𝐹𝑖𝑒𝑙𝑑(100) 𝑒𝑙𝑒𝑐+𝑣𝑑𝑊
𝐸𝑗,𝑀𝑀+𝑄𝑀 ⁄𝑀𝑀 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 )] + 𝐸𝑗,𝑀𝑀+𝑄𝑀 ⁄𝑀𝑀 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 ) (3.3)

where the first term is evaluated for a periodic slab using the VASP program

package (planewave DFT calculation), the second term is a QM cluster calculation in

vacuum computed with the TURBOMOLE program package (non-periodic cluster

calculations), and the third term is a QM cluster calculation in a periodic mean field of

MM water molecules computed using the periodic electrostatic embedded cluster method

(PEECM) in TURBOMOLE under the fixed-charge approximation (fixed-charge

approximation has been validated for our eSMS approach54). We note that the number 100

in the equation indicates that 100 MM conformations, selected equally spaced from

equilibrated 1 ns molecular dynamic (MD) simulations (10 ps apart), were used to represent

the mean field of the MM water molecules. Finally, the last three terms account for the

classical (MM level of theory) electrostatic and van-der-Waals interaction energy of the

total system without overcounting the electrostatic interaction of the MM (water)

molecules with the QM cluster subsystem. We note that all solvent (water) molecules are

described in this study at the MM level of theory.

Free energy calculations require energy evaluation from uncorrelated

measurements of the system and ideally the energy estimator should also be capable of

minimizing the statistical bias and variance of the free energy differences of the physical

system being studied. Exponential averaging (EXP), also known as the Zwanzig

60
relationship108 has long been applied to study a variety of problems such as amino acid

recognition,109 RAS-RAF binding affinity,110 and octanol/water partition coefficients,111

etc. However, the EXP has been shown to represent poor efficiency and phase space

overlap,112, 113 and is also largely dependent on the distribution of the QM/MM energy.114

Here, we employed the Bennett Acceptance Ratio (BAR)115 as the free energy estimator

which uses both the forward and reverse distributions simultaneously in a more efficient

way than simply averaging the forward and reverse exponential estimators. BAR has been

demonstrated to benefit from a lower bias and variance of the free energy estimates in

practical atomistic simulations when compared to EXP and thermodynamic integration

(TI).112, 116 Finally, the whole free energy estimation procedure has been repeated three

times independently to establish 95% confidence intervals for evaluating the free energy

of activation, assuming a normal distribution.117 More independent simulations were

carried out only if these three experiments were not resulting in 95% confidence interval

smaller than 0.05 eV. All uncertainties reported in this study are 95% confidence intervals.

3.5.5 Average rotational correlation time


Adequately sampling of the potential energy surface for all relevant configurations

of the system is of great importance in any QM/MM approach for computing the liquid-

phase effect on the free energy of elementary processes114. Owing to a lack of consensus

on how much sampling of the configurational space is sufficient for a solvated adsorbed

carbohydrate species on a metal surface for an error smaller than 0.05 eV, we computed

the average rotational correlation time for water molecules in close proximity (up to 5 Å)

to adsorbed ethylene glycol in the reactant and transition states over the (111) facet of six

transition metals and reported them in Table B.3. The longest average correlation time is

61
computed to be ~200 ps. Hence, we decided to sample for 1000 ps to make sure that

relevant configurations of the systems are sampled adequately. As discussed above, for

free-energy calculations, the procedure was repeated at least three times with independent

MD trajectories to establish the confidence interval of the computation of the free energy

of activation in aqueous phase.

3.5.6 Non-Periodic Implicit Solvation Calculations


The implicit solvation model for solid surfaces (iSMS)69 was utilized to compute

the activation free energies in an aqueous reaction environment as

‡ ‡ 𝑇𝑆 𝐼𝑆
∆𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 = ∆𝐺𝐺𝑎𝑠 + [𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 − 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 ] (3.4)

‡ 𝐼𝑆
where ∆𝐺𝐺𝑎𝑠 is the respective activation free energy under gas phase conditions, and 𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 and
𝑇𝑆
𝐺𝑠𝑜𝑙𝑣𝑒𝑛𝑡 represent the solvation free energies of initial and transition states, respectively, computed
as
𝑙𝑖𝑞𝑢𝑖𝑑 𝑣𝑎𝑐𝑢𝑢𝑚 𝑙𝑖𝑞𝑢𝑖𝑑
𝐺𝑠𝑢𝑟𝑓𝑎𝑐𝑒+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 = 𝐺𝑠𝑢𝑟𝑓𝑎𝑐𝑒+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 + (𝐺𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 −
𝑣𝑎𝑐𝑢𝑢𝑚
𝐸𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 ) (3.5)
𝑣𝑎𝑐𝑢𝑢𝑚
where 𝐺𝑠𝑢𝑟𝑓𝑎𝑐𝑒+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 is the free energy of an intermediate (e.g. IS or TS) in the

absence of a solvent, computed here within the harmonic approximation using plane-wave

𝑙𝑖𝑞𝑢𝑖𝑑
DFT calculations for periodic slab models, 𝐺𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 is the free energy of the

metal cluster and surface intermediate in liquid without vibrational contributions (that are
𝑣𝑎𝑐𝑢𝑢𝑚
already considered in the first term), and 𝐸𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 is the DFT energy of the

𝑙𝑖𝑞𝑢𝑖𝑑
same cluster in the absence of the solvent. To compute the 𝐺𝑐𝑙𝑢𝑠𝑡𝑒𝑟+𝑖𝑛𝑡𝑒𝑟𝑚𝑒𝑑𝑖𝑎𝑡𝑒 term,

COSMO-RS118, 119 (Conductor-like Screening Model for Real Solvents) calculations are

performed using the COSMOtherm program with latest FINE parameterization.120 The

COSMOtherm program for solvent thermodynamic properties requires COSMO

calculations to be performed at the BP-TZVPD level of theory. Given the uncertainty in

62
cavity radius for transition metal elements, calculations are performed with default and

10% increased and decreased cavity radius for the transition metal elements (see Table

B.9).

3.5.7 Periodic Implicit Solvation Calculations


In addition to implicit solvation calculations performed with iSMS method, implicit

solvation calculations were also performed with VASPsol70, 71 using a relative permittivity

of water of 44.07 at 423 K.121 For VASPsol, we used the default values for the parameter

𝑛𝐶 that defines the value at which the dielectric cavity forms and for the width of the diffuse

cavity, 𝜎, and for effective surface tension parameter, 𝜏, describing the cavitation,

dispersion, and repulsive interaction between the solute and the solvent that are not

captured by the electrostatic terms.71 While the parameters are likely most accurate only

for simulations at 298 K and not at 423 K, they are the optimized parameters of the solvent

model that cannot easily be obtained at other temperatures. Due to the absence of adequate

experimental solvation data at 423 K, we decided that the default parameters are likely

most meaningful, i.e., the relative permittivity is the only temperature dependent solvent

parameter in our VASPsol model. All other computational details for periodic implicit

solvation calculations were kept the same as in our periodic vapor-phase calculations.

3.6 Acknowledgements
We gratefully acknowledge financial support from the U.S. Department of Energy,

Office of Basic Energy Science, Catalysis Science program under Award DE-SC0007167.

In addition, this work was partially supported by the South Carolina Smart State Center for

Strategic Approaches to the Generation of Electricity (SAGE). Computational resources

have been provided by the National Energy Research Scientific Computing Center

63
(NERSC) which is supported by the Office of Science of the U.S. Department of Energy

and in part by XSEDE under grant number TG-CTS090100. Computational resources from

the CASCADE cluster from the Environmental Molecular Sciences Laboratory (EMSL)

under Pacific Northwest National Laboratory (PNNL, Ringgold ID 130367, Grant Proposal

51163) are also used for selected DFT calculations. Finally, computing resources from the

USC High Performance Computing Group are gratefully acknowledged.

64
3.7 Tables and Figures

Table 3.1: Free energies of activation of O-H bond cleavage (CH2OHCH2OH** + * ↔


CH2OCH2OH** + H*) and C-H bond cleavage (CH2OHCH2OH** + * ↔ CHOHCH2OH
** + H*) of ethylene glycol in gas- and aqueous-phase environments over the (111) surface
facet of six transition metals at 423 K. QM/MM-FEP indicates a free energy calculation in
water between the critical points identified by gas-phase calculations (using the gas-phase
vibrational partition function for the reactant and transition states). QM/MM-MFEP-OPT
represents a free energy calculation in water for the different cleavages between the
respective reactant and transition states that have been optimized in an aqueous-phase
environment. Here, the vibrational partition functions are computed in the aqueous phase
assuming the timescale for reorientation of the solvent molecules is much larger than the
timescale for molecular vibrations. For comparison, implicit solvation calculations have
also been performed using both nonperiodic (iSMS)69 and periodic (VASPsol)70, 71
approaches. All numbers are in eV.
Metal Reaction Environment O-H cleavage C-H
cleavage
Vapor Phase 0.70 0.79
VASPsol 0.67 0.73
Ni(111) iSMS 0.62 0.75
QM/MM-FEP 0.49 ± 0.06 0.45 ± 0.00
QM/MM-MFEP-OPT 0.55 ± 0.06 0.43 ± 0.02
Vapor Phase 0.83 0.76
VASPsol 0.85 0.68
Pd(111) iSMS 0.86 0.75
QM/MM-FEP 0.59 ± 0.00 0.74 ± 0.02
QM/MM-MFEP-OPT 0.58 ± 0.00 0.76 ± 0.02
Vapor Phase 0.68 0.71
VASPsol 0.82 0.64
Pt(111) iSMS 0.62 0.62
QM/MM-FEP 0.23 ± 0.01 0.56 ± 0.01
QM/MM-MFEP-OPT 0.25 ± 0.01 0.55 ± 0.01
Vapor Phase 1.08 1.43
VASPsol 1.01 1.31
Cu(111) iSMS 0.99 1.31
QM/MM-FEP 0.52 ± 0.04 1.06 ± 0.00
QM/MM-MFEP-OPT 0.50 ± 0.03 1.02 ± 0.00
Vapor Phase 1.63 1.91

65
VASPsol 1.65 1.81
iSMS 1.57 1.88
Ag(111)
QM/MM-FEP 1.22 ± 0.04 1.57 ± 0.03
QM/MM-MFEP-OPT 1.14 ± 0.04 1.56 ± 0.03
Vapor Phase 1.67 1.69
VASPsol 1.75 1.43
Au(111) iSMS 1.64 1.52
QM/MM-FEP 1.55 ± 0.00 1.43 ± 0.01
QM/MM-MFEP-OPT 1.51 ± 0.01 1.41 ± 0.03

66
Table 3.2: Average aqueous-phase effects on the activation free energy (∆∆𝐺 𝑎𝑐𝑡 =
∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) as well as three solvation effect descriptors for the O-H and C-H
bond cleavages of ethylene glycol over the (111) surface facet of six transition metals at
423 K (see Table 3.1 for 95% confidence intervals). ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 is the free energy of
activation in vapor phase, H-bond denotes the change in mean of total hydrogen bonding
(acceptor + donor) going from RS to TS, MC (molecular charge transfer) represents the
change in the absolute sum of partial charges on the reacting moiety going from RS to TS,
and finally BC (cleaving-bond charge transfer) is the change in sum of partial charges on
the cleaving bond atoms going from RS to TS.

∆𝑮𝒂𝒄𝒕,𝒈𝒂𝒔,
Cleavage Surface ∆∆𝑮𝒂𝒄𝒕 , eV H-bond MC, e BC, e
eV
Ni -0.21 0.70 0.99 0.64 -0.61
Pd -0.24 0.76 0.34 0.13 -0.36
Pt -0.45 0.68 0.54 -0.10 -0.24
O-H
Cu -0.55 1.08 0.77 0.69 -0.74
Ag -0.41 1.63 0.92 0.78 -0.77
Au -0.12 1.67 0.95 0.62 -0.63
Ni -0.34 0.79 0.26 0.41 -0.36
Pd -0.02 0.83 -0.26 -0.03 0.06
Pt -0.15 0.71 0.12 0.08 0.03
C-H
Cu -0.37 1.43 0.21 0.53 -0.46
Ag -0.34 1.91 -0.27 0.28 -0.32
Au -0.26 1.69 -0.38 0.06 -0.12

67
Figure 3.1: Aqueous-phase effects on the activation free energy barrier of the O-H and C-
H bond cleavages of ethylene glycol over the (111) surface facet of six transition metals at
423 K computed by eSMS (all structures optimized in liquid water; see Table 3.1 for
specific numbers).

68
Figure 3.2: Free-energy profiles for O-H bond cleavage of ethylene glycol in vapor and
aqueous phase over the (111) surface facets of six transition metals at 423 K without
considering vibrational contributions to the partition functions. See Table 3.1 for
corresponding data that include vibrational contributions. The number of intermediate
states between the reactant and transition states for eSMS calculations is determined by our
desire to have an energy difference between intermediates smaller than twice the thermal
energy (< 2𝑘𝐵 𝑇). The aqueous-phase profiles portray the average of three or more
independent eSMS calculations possessing 95% confidence intervals smaller than ±0.05
eV (see Table 3.1). The analogous plot for C-H bond cleavage reaction is provided in the
Appendix B.

69
(a) (b)

Figure 3.3: Parity plot of the aqueous-phase effect on the free energy of activation
(∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) computed by eSMS (QM/MM-FEP in Table 3.1) versus
the model-predicted aqueous-phase effect. The predicted values are calculated using the
gas-phase free energy of activation (∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) and cleaving-bond charge transfer (BC) as
two descriptors in (a) a linear model (equation 3.1) for C-H bond cleavage and (b) a
quadratic model (equation 3.2) for O-H bond cleavage of ethylene glycol over (111)
̅̅̅̅̅̅̅̅̅
surface facets of six transition metals at 423 K. The average aqueous-phase effect, ∆∆𝐺 𝑎𝑐𝑡 ,

computed by eSMS, is -0.33 eV for O-H bond cleavage and -0.25 eV for C-H bond
cleavage. The mean of absolute errors (𝑀𝐴𝐸) between eSMS and model-predicted values
is 0.03 eV in the linear model (a) and 0.02 eV in the quadratic model (b). The corresponding
model parameters are listed in Tables B.7 and B.8.

70
3.8 Bibliography

1. Verdasco G, Martin MA, Delcastillo B, Lopezalvarado P, Menendez JC.


SOLVENT EFFECTS ON THE FLUORESCENT EMISSION OF SOME NEW
BENZIMIDAZOLE DERIVATIVES. Analytica Chimica Acta 303, 73-78 (1995).

2. Fung HL, Nealon T. SOLVENT EFFECTS ON COMPARATIVE


DISSOLUTION OF PHARMACEUTICAL SOLVATES. Chemical &
Pharmaceutical Bulletin 22, 454-458 (1974).

3. Jenke D, Odufu A, Poss M. The effect of solvent polarity on the accumulation of


leachables from pharmaceutical product containers. European Journal of
Pharmaceutical Sciences 27, 133-142 (2006).

4. Patil S, Patil S, Navale S. Effect of Solvent and Crystallization Method on


Physicochemical Properties of Aceclofenac and Fenofibrate. British Journal of
Pharmaceutical Research 12, (2016).

5. Beilin E, Baker LJ, Aikins J, Baryla NE. Effect of incomplete removal of the tert-
butoxycarbonyl protecting group during synthesis of a pharmaceutical drug
substance on the residual solvent analysis. Journal of Pharmaceutical and
Biomedical Analysis 52, 316-319 (2010).

6. Mota FL, Carneiro AR, Queimada AJ, Pinho SP, Macedo EA. Temperature and
solvent effects in the solubility of some pharmaceutical compounds:
Measurements and modeling. European Journal of Pharmaceutical Sciences 37,
499-507 (2009).

7. Chen LD, Urushihara M, Chan KR, Norskov JK. Electric Field Effects in
Electrochemical CO2 Reduction. Acs Catal 6, 7133-7139 (2016).

8. Cheng T, Xiao H, Goddard WA. Free-Energy Barriers and Reaction Mechanisms


for the Electrochemical Reduction of CO on the Cu(100) Surface, Including
Multiple Layers of Explicit Solvent at pH 0. J Phys Chem Lett 6, 4767-4773
(2015).

9. Goldsmith ZK, Secor M, Hammes-Schiffer S. Inhomogeneity of Interfacial


Electric Fields at Vibrational Probes on Electrode Surfaces. Acs Central Science
6, 304-311 (2020).

71
10. Otani M, Hamada I, Sugino O, Morikawa Y, Okamoto Y, Ikeshoji T. Structure of
the water/platinum interface - a first principles simulation under bias potential.
Phys Chem Chem Phys 10, 3609-3612 (2008).

11. Zope BN, Hibbitts DD, Neurock M, Davis RJ. Reactivity of the Gold/Water
Interface During Selective Oxidation Catalysis. Science 330, 74-78 (2010).

12. Yoon Y, Rousseau R, Weber RS, Mei DH, Lercher JA. First-Principles Study of
Phenol Hydrogenation on Pt and Ni Catalysts in Aqueous Phase. J Am Chem Soc
136, 10287-10298 (2014).

13. Tupy SA, et al. Correlating Ethylene Glycol Reforming Activity with In Situ
EXAFS Detection of Ni Segregation in Supported NiPt Bimetallic Catalysts. Acs
Catal 2, 2290-2296 (2012).

14. Saleheen M, Heyden A. Liquid-Phase Modeling in Heterogeneous Catalysis. Acs


Catal 8, 2188-2194 (2018).

15. Karim AM, et al. In Situ X-ray Absorption Fine Structure Studies on the Effect of
pH on Pt Electronic Density during Aqueous Phase Reforming of Glycerol. Acs
Catal 2, 2387-2394 (2012).

16. Miller KL, Lee CW, Falconer JL, Medlin JW. Effect of water on formic acid
photocatalytic decomposition on TiO2 and Pt/TiO2. J Catal 275, 294-299 (2010).

17. Siemer N, Muñoz-Santiburcio D, Marx D. Solvation-Enhanced Oxygen


Activation at Gold/Titania Nanocatalysts. Acs Catal 10, 8530-8534 (2020).

18. Singh N, Sanyal U, Fulton JL, Gutiérrez OY, Lercher JA, Campbell CT.
Quantifying Adsorption of Organic Molecules on Platinum in Aqueous Phase by
Hydrogen Site Blocking and in Situ X-ray Absorption Spectroscopy. Acs Catal 9,
6869-6881 (2019).

19. Akinola J, Barth I, Goldsmith BR, Singh N. Adsorption Energies of Oxygenated


Aromatics and Organics on Rhodium and Platinum in Aqueous Phase. Acs Catal
10, 4929-4941 (2020).

20. Behtash S, Lu JM, Mamun O, Williams CT, Monnier JR, Heyden A. Solvation
Effects in the Hydrodeoxygenation of Propanoic Acid over a Model Pd(211)
Catalyst. J Phys Chem C 120, 2724-2736 (2016).

72
21. Rajadhyaksha RA, Karwa SL. Solvent Effects in Catalytic-Hydrogenation. Chem
Eng Sci 41, 1765-1770 (1986).

22. Mukherjee S, Vannice MA. Solvent effects in liquid-phase reactions - I. Activity


and selectivity during citral hydrogenation on Pt/SiO2 and evaluation of mass
transfer effects. J Catal 243, 108-130 (2006).

23. Akpa BS, et al. Solvent effects in the hydrogenation of 2-butanone. J Catal 289,
30-41 (2012).

24. McManus I, et al. Effect of solvent on the hydrogenation of 4-phenyl-2-butanone


over Pt based catalysts. J Catal 330, 344-353 (2015).

25. Knowles WS. Asymmetric hydrogenation. Acc Chem Res 16, 106-112 (1983).

26. Valgimigli L, Banks JT, Ingold KU, Lusztyk J. Kinetic Solvent Effects on
Hydroxylic Hydrogen-Atom Abstractions Are Independent of the Nature of the
Abstracting Radical - 2 Extreme Tests Using Vitamin-E and Phenol. J Am Chem
Soc 117, 9966-9971 (1995).

27. Wiebus E, Cornils B. Industrial-Scale Oxo Synthesis with an Immobilized


Catalyst. Chem-Ing-Tech 66, 916-923 (1994).

28. Beller M, Cornils B, Frohning CD, Kohlpaintner CW. Progress in


Hydroformylation and Carbonylation. J Mol Catal a-Chem 104, 17-85 (1995).

29. Dyson PJ, Jessop PG. Solvent effects in catalysis: rational improvements of
catalysts via manipulation of solvent interactions. Catal Sci Technol 6, 3302-3316
(2016).

30. Cornils B, Herrmann WA, Eckl RW. Industrial aspects of aqueous catalysis. J
Mol Catal a-Chem 116, 27-33 (1997).

31. Zhang L, Karim AM, Engelhard MH, Wei ZH, King DL, Wang Y. Correlation of
Pt-Re surface properties with reaction pathways for the aqueous-phase reforming
of glycerol. J Catal 287, 37-43 (2012).

32. He R, Davda RR, Dumesic JA. In situ ATR-IR spectroscopic and reaction
kinetics studies of water-gas shift and methanol reforming on Pt/Al2O3 catalysts
in vapor and liquid phases. J Phys Chem B 109, 2810-2820 (2005).

73
33. Taylor CD, Neurock M. Theoretical insights into the structure and reactivity of
the aqueous/metal interface. Curr Opin Solid St M 9, 49-65 (2005).

34. Faheem M, Saleheen M, Lu JM, Heyden A. Ethylene glycol reforming on


Pt(111): first-principles microkinetic modeling in vapor and aqueous phases.
Catal Sci Technol 6, 8242-8256 (2016).

35. Bertero NM, Trasarti AF, Acevedo MC, Marchi AJ, Apesteguia CR. Solvent
effects in solid acid-catalyzed reactions: The case of the liquid-phase
isomerization/cyclization of citronellal over SiO2-Al2O3. Molecular Catalysis
481, (2020).

36. Faveere W, et al. Glycolaldehyde as a Bio-Based C-2 Platform Chemical:


Catalytic Reductive Amination of Vicinal Hydroxyl Aldehydes. Acs Catal 10,
391-404 (2020).

37. Saleheen M, Verma AM, Mamun O, Lu J, Heyden A. Investigation of solvent


effects on the hydrodeoxygenation of guaiacol over Ru catalysts. Catalysis
Science & Technology 9, 6253-6273 (2019).

38. Zhao JJ, et al. Suppressing Metal Leaching in a Supported Co/SiO2 Catalyst with
Effective Protectants in the Hydroformylation Reaction. Acs Catal 10, 914-920
(2020).

39. Kulal AB, Kasabe MM, Jadhav PV, Dongare MK, Umbarkar SB. Hydrophobic
WO3/SiO2 catalyst for the nitration of aromatics in liquid phase. Applied
Catalysis a-General 574, 105-113 (2019).

40. Staszak-Jirkovsky J, et al. Water as a Promoter and Catalyst for Dioxygen


Electrochemistry in Aqueous and Organic Media. Acs Catal 5, 6600-6607 (2015).

41. Wan H, Vitter A, Chaudhari RV, Subramaniam B. Kinetic investigations of


unusual solvent effects during Ru/C catalyzed hydrogenation of model
oxygenates. J Catal 309, 174-184 (2014).

42. Mellmer MA, Sener C, Gallo JMR, Luterbacher JS, Alonso DM, Dumesic JA.
Solvent Effects in Acid-Catalyzed Biomass Conversion Reactions. Angewandte
Chemie International Edition 53, 11872-11875 (2014).

43. Zare M, Solomon RV, Yang W, Yonge A, Heyden A. Theoretical Investigation of


Solvent Effects on the Hydrodeoxygenation of Propionic Acid over a Ni(111)
Catalyst Model. The Journal of Physical Chemistry C 124, 16488-16500 (2020).

74
44. Car R, Parrinello M. Unified Approach for Molecular-Dynamics and Density-
Functional Theory. Phys Rev Lett 55, 2471-2474 (1985).

45. Carloni P, Rothlisberger U, Parrinello M. The Role and Perspective of Ab Initio


Molecular Dynamics in the Study of Biological Systems. Accounts of Chemical
Research 35, 455-464 (2002).

46. Iftimie R, Minary P, Tuckerman ME. Ab initio molecular dynamics: Concepts,


recent developments, and future trends. P Natl Acad Sci USA 102, 6654-6659
(2005).

47. Mattsson TR, Paddison SJ. Methanol at the water-platinum interface studied by
ab initio molecular dynamics. Surf Sci 544, L697-L702 (2003).

48. Yang J, Dauenhauer PJ, Ramasubramaniam A. The role of water in the adsorption
of oxygenated aromatics on Pt and Pd. J Comput Chem 34, 60-66 (2013).

49. Klamt A. Conductor-Like Screening Model for Real Solvents - a New Approach
to the Quantitative Calculation of Solvation Phenomena. J Phys Chem-Us 99,
2224-2235 (1995).

50. Klamt A, Schuurmann G. Cosmo - a New Approach to Dielectric Screening in


Solvents with Explicit Expressions for the Screening Energy and Its Gradient. J
Chem Soc Perk T 2, 799-805 (1993).

51. Zhang YK, Liu HY, Yang WT. Free energy calculation on enzyme reactions with
an efficient iterative procedure to determine minimum energy paths on a
combined ab initio QM/MM potential energy surface. J Chem Phys 112, 3483-
3492 (2000).

52. Hu H, Lu ZY, Parks JM, Burger SK, Yang WT. Quantum mechanics/molecular
mechanics minimum free-energy path for accurate reaction energetics in solution
and enzymes: Sequential sampling and optimization on the potential of mean
force surface. J Chem Phys 128, 034105:034101-034118 (2008).

53. Hu H, Lu ZY, Yang WT. QM/MM minimum free-energy path: Methodology and
application to triosephosphate isomerase. J Chem Theory Comput 3, 390-406
(2007).

54. Faheem M, Heyden A. Hybrid Quantum Mechanics/Molecular Mechanics


Solvation Scheme for Computing Free Energies of Reactions at Metal-Water
Interfaces. J Chem Theory Comput 10, 3354-3368 (2014).

75
55. Saleheen M, Zare M, Faheem M, Heyden A. Computational Investigation of
Aqueous Phase Effects on the Dehydrogenation and Dehydroxylation of Polyols
over Pt(111). The Journal of Physical Chemistry C 123, 19052-19065 (2019).

56. Breslow R, Guo T. Diels-Alder Reactions in Nonaqueous Polar-Solvents - Kinetic


Effects of Chaotropic and Antichaotropic Agents and of Beta-Cyclodextrin. J Am
Chem Soc 110, 5613-5617 (1988).

57. Li Y, et al. Solvent effects on heterogeneous catalysis in the selective


hydrogenation of cinnamaldehyde over a conventional Pd/C catalyst. Catalysis
Science & Technology 8, 3580-3589 (2018).

58. Ebbesen SD, Mojet BL, Lefferts L. In situ ATR-IR study of CO adsorption and
oxidation over Pt/Al2O3 in gas and aqueous phase: Promotion effects by water
and pH. J Catal 246, 66-73 (2007).

59. Deshpande S, Greeley J. First-Principles Analysis of Coverage, Ensemble, and


Solvation Effects on Selectivity Trends in NO Electroreduction on Pt3Sn Alloys.
Acs Catal 10, 9320-9327 (2020).

60. Clayborne A, Chun H-J, Rankin RB, Greeley J. Elucidation of Pathways for NO
Electroreduction on Pt(111) from First Principles. Angewandte Chemie
International Edition 54, 8255-8258 (2015).

61. Kim S, et al. Recent advances in hydrodeoxygenation of biomass-derived


oxygenates over heterogeneous catalysts. Green Chemistry 21, 3715-3743 (2019).

62. Salciccioli M, Vlachos DG. Kinetic Modeling of Pt Catalyzed and Computation-


Driven Catalyst Discovery for Ethylene Glycol Decomposition. Acs Catal 1,
1246-1256 (2011).

63. Salciccioli M, Vlachos DG. Correction to Kinetic Modeling of Pt Catalyzed and


Computation-Driven Catalyst Discovery for Ethylene Glycol Decomposition. Acs
Catal 2, 306-306 (2012).

64. Salciccioli M, Yu W, Barteau MA, Chen JG, Vlachos DG. Differentiation of O–H
and C–H Bond Scission Mechanisms of Ethylene Glycol on Pt and Ni/Pt Using
Theory and Isotopic Labeling Experiments. J Am Chem Soc 133, 7996-8004
(2011).

65. Zhang X, DeFever RS, Sarupria S, Getman RB. Free Energies of Catalytic
Species Adsorbed to Pt(111) Surfaces under Liquid Solvent Calculated Using

76
Classical and Quantum Approaches. Journal of Chemical Information and
Modeling 59, 2190-2198 (2019).

66. Reed AE, Weinstock RB, Weinhold F. Natural-Population Analysis. J Chem Phys
83, 735-746 (1985).

67. Natarajan SK, Behler J. Neural network molecular dynamics simulations of solid-
liquid interfaces: water at low-index copper surfaces. Phys Chem Chem Phys 18,
28704-28725 (2016).

68. Pearson’s Correlation Coefficient. In: Encyclopedia of Public Health (ed Kirch
W). Springer Netherlands (2008).

69. Faheem M, Suthirakun S, Heyden A. New Implicit Solvation Scheme for Solid
Surfaces. J Phys Chem C 116, 22458-22462 (2012).

70. Fishman M, Zhuang HLL, Mathew K, Dirschka W, Hennig RG. Accuracy of


exchange-correlation functionals and effect of solvation on the surface energy of
copper. Phys Rev B 87, (2013).

71. Mathew K, Sundararaman R, Letchworth-Weaver K, Arias TA, Hennig RG.


Implicit solvation model for density-functional study of nanocrystal surfaces and
reaction pathways. J Chem Phys 140, 084106:084101-084108 (2014).

72. Kresse G, Furthmuller J. Efficiency of ab-initio total energy calculations for


metals and semiconductors using a plane-wave basis set. Comp Mater Sci 6, 15-
50 (1996).

73. Kresse G, Furthmuller J. Efficient iterative schemes for ab initio total-energy


calculations using a plane-wave basis set. Phys Rev B 54, 11169-11186 (1996).

74. Blochl PE. Projector Augmented-Wave Method. Phys Rev B 50, 17953-17979
(1994).

75. Perdew JP, Yue W. Accurate and Simple Density Functional for the Electronic
Exchange Energy - Generalized Gradient Approximation. Phys Rev B 33, 8800-
8802 (1986).

76. Perdew JP, Wang Y. Accurate and Simple Analytic Representation of the
Electron-Gas Correlation-Energy. Phys Rev B 45, 13244-13249 (1992).

77
77. Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made
simple. Phys Rev Lett 77, 3865-3868 (1996).

78. Monkhorst HJ, Pack JD. Special Points for Brillouin-Zone Integrations. Phys Rev
B 13, 5188-5192 (1976).

79. Methfessel M, Paxton AT. High-Precision Sampling for Brillouin-Zone


Integration in Metals. Phys Rev B 40, 3616-3621 (1989).

80. Makov G, Payne MC. Periodic Boundary-Conditions in Ab-Initio Calculations.


Phys Rev B 51, 4014-4022 (1995).

81. Harris J. Simplified Method for Calculating the Energy of Weakly Interacting
Fragments. Phys Rev B 31, 1770-1779 (1985).

82. Matthew W, Foulkes C, Haydock R. Tight-Binding Models and Density-


Functional Theory. Phys Rev B 39, 12520-12536 (1989).

83. Henkelman G, Jonsson H. Improved tangent estimate in the nudged elastic band
method for finding minimum energy paths and saddle points. J Chem Phys 113,
9978-9985 (2000).

84. Henkelman G, Uberuaga BP, Jonsson H. A climbing image nudged elastic band
method for finding saddle points and minimum energy paths. J Chem Phys 113,
9901-9904 (2000).

85. Henkelman G, Jonsson H. A dimer method for finding saddle points on high
dimensional potential surfaces using only first derivatives. J Chem Phys 111,
7010-7022 (1999).

86. Heyden A, Bell AT, Keil FJ. Efficient methods for finding transition states in
chemical reactions: Comparison of improved dimer method and partitioned
rational function optimization method. J Chem Phys 123, 224101:224101-224114
(2005).

87. Ahlrichs R, Bar M, Haser M, Horn H, Kolmel C. Electronic-Structure


Calculations on Workstation Computers - the Program System Turbomole. Chem
Phys Lett 162, 165-169 (1989).

88. Treutler O, Ahlrichs R. Efficient Molecular Numerical-Integration Schemes. J


Chem Phys 102, 346-354 (1995).

78
89. Von Arnim M, Ahlrichs R. Performance of parallel TURBOMOLE for density
functional calculations. J Comput Chem 19, 1746-1757 (1998).

90. Schafer A, Horn H, Ahlrichs R. Fully Optimized Contracted Gaussian-Basis Sets


for Atoms Li to Kr. J Chem Phys 97, 2571-2577 (1992).

91. Weigend F, Ahlrichs R. Balanced basis sets of split valence, triple zeta valence
and quadruple zeta valence quality for H to Rn: Design and assessment of
accuracy. Phys Chem Chem Phys 7, 3297-3305 (2005).

92. Russo TV, Martin RL, Hay PJ. Effective Core Potentials for Dft Calculations. J
Phys Chem-Us 99, 17085-17087 (1995).

93. Eichkorn K, Weigend F, Treutler O, Ahlrichs R. Auxiliary basis sets for main row
atoms and transition metals and their use to approximate Coulomb potentials.
Theor Chem Acc 97, 119-124 (1997).

94. Weigend F. Accurate Coulomb-fitting basis sets for H to Rn. Phys Chem Chem
Phys 8, 1057-1065 (2006).

95. Todorov IT, Smith W, Trachenko K, Dove MT. DL_POLY_3: new dimensions in
molecular dynamics simulations via massive parallelism. J Mater Chem 16, 1911-
1918 (2006).

96. Jorgensen WL, Chandrasekhar J, Madura JD, Impey RW, Klein ML. Comparison
of Simple Potential Functions for Simulating Liquid Water. J Chem Phys 79, 926-
935 (1983).

97. Andersen HC. Rattle - a Velocity Version of the Shake Algorithm for Molecular-
Dynamics Calculations. J Comput Phys 52, 24-34 (1983).

98. Smith W, Forester TR. Parallel Macromolecular Simulations and the Replicated
Data Strategy .2. The Rd-Shake Algorithm. Comput Phys Commun 79, 63-77
(1994).

99. Swope WC, Andersen HC, Berens PH, Wilson KR. A Computer-Simulation
Method for the Calculation of Equilibrium-Constants for the Formation of
Physical Clusters of Molecules - Application to Small Water Clusters. J Chem
Phys 76, 637-649 (1982).

100. Jorgensen WL. Optimized Intermolecular Potential Functions for Liquid


Alcohols. J Phys Chem-Us 90, 1276-1284 (1986).

79
101. Geerke DP, Van Gunsteren WF. The performance of non-polarizable and
polarizable force-field parameter sets for ethylene glycol in molecular dynamics
simulations of the pure liquid and its aqueous mixtures. Mol Phys 105, 1861-1881
(2007).

102. Freindorf M, Shao Y, Furlani TR, Kong J. Lennard–Jones parameters for the
combined QM/MM method using the B3LYP/6-31G*/AMBER potential. Journal
of Computational Chemistry 26, 1270-1278 (2005).

103. Heinz H, Vaia RA, Farmer BL, Naik RR. Accurate Simulation of Surfaces and
Interfaces of Face-Centered Cubic Metals Using 12-6 and 9-6 Lennard-Jones
Potentials. J Phys Chem C 112, 17281-17290 (2008).

104. Burow AM, Sierka M, Dobler J, Sauer J. Point defects in CaF2 and CeO2
investigated by the periodic electrostatic embedded cluster method. J Chem Phys
130, 174710:174711-174711 (2009).

105. Nose S. A Unified Formulation of the Constant Temperature Molecular-


Dynamics Methods. J Chem Phys 81, 511-519 (1984).

106. Hoover WG. Canonical Dynamics - Equilibrium Phase-Space Distributions. Phys


Rev A 31, 1695-1697 (1985).

107. Essmann U, Perera L, Berkowitz ML, Darden T, Lee H, Pedersen LG. A Smooth
Particle Mesh Ewald Method. J Chem Phys 103, 8577-8593 (1995).

108. Zwanzig RW. High-Temperature Equation of State by a Perturbation Method .1.


Nonpolar Gases. J Chem Phys 22, 1420-1426 (1954).

109. Archontis G, Simonson T, Moras D, Karplus M. Specific amino acid recognition


by aspartyl-tRNA synthetase studied by free energy simulations. J Mol Biol 275,
823-846 (1998).

110. Jun Z, Thomas S, Masha F, Hiroshi M, R. TH. Protein‐protein recognition: An


experimental and computational study of the R89K mutation in Raf and its effect
on Ras binding. Protein Science 8, 50-64 (1999).

111. Best SA, Merz KM, Reynolds CH. Free energy perturbation study of
octanol/water partition coefficients: Comparison with continuum GB/SA
calculations. J Phys Chem B 103, 714-726 (1999).

80
112. Shirts MR, Pande VS. Comparison of efficiency and bias of free energies
computed by exponential averaging, the Bennett acceptance ratio, and
thermodynamic integration. J Chem Phys 122, 144107:144101-144116 (2005).

113. Lu ND, Singh JK, Kofke DA. Appropriate methods to combine forward and
reverse free-energy perturbation averages. J Chem Phys 118, 2977-2984 (2003).

114. Ryde U. How Many Conformations Need To Be Sampled To Obtain Converged


QM/MM Energies? The Curse of Exponential Averaging. J Chem Theory Comput
13, 5745-5752 (2017).

115. Bennett C. Efficient Estimation of Free Energy Differences from Monte Carlo
Data. J Comput Phys 22, 245-268 (1976).

116. Shirts MR, Bair E, Hooker G, Pande VS. Equilibrium free energies from
nonequilibrium measurements using maximum-likelihood methods. Phys Rev Lett
91, (2003).

117. Kreyszig E. Advanced Engineering Mathematics, 9th edn. John Wiley & Sons,
Inc. (2006).

118. Klamt A, Jonas V, Bürger T, Lohrenz JCW. Refinement and Parametrization of


COSMO-RS. The Journal of Physical Chemistry A 102, 5074-5085 (1998).

119. Klamt A. Conductor-like Screening Model for Real Solvents: A New Approach to
the Quantitative Calculation of Solvation Phenomena. The Journal of Physical
Chemistry 99, 2224-2235 (1995).

120. Marsh KN. COSMO-RS from Quantum Chemistry to Fluid Phase


Thermodynamics and Drug Design. By A. Klamt. Elsevier: Amsterdam, The
Netherlands, 2005. 246 pp. $US 165. ISBN 0-444-51994-7. Journal of Chemical
& Engineering Data 51, 1480-1480 (2006).

121. Fernandez DP, Goodwin ARH, Lemmon EW, Sengers JMHL, Williams RC. A
formulation for the static permittivity of water and steam at temperatures from
238 K to 873 K at pressures up to 1200 MPa, including derivatives and Debye-
Huckel coefficients. J Phys Chem Ref Data 26, 1125-1166 (1997).

81
CHAPTER 4

LIQUID PHASE EFFECTS ON ADSORPTION PROCESSES IN

HETEROGENOUS CATALYSIS

Zare, M.; Saleheen, M.; Singh, N.; Uline, M. J.; Faheem, M.; Heyden, A.

Submitted to Nature Communications

82
4.1 Abstract

Aqueous solvation free energies of adsorption have recently been measured for

phenol adsorption on Pt(111) surfaces. Endergonic solvent effects of >1.4 eV suggest

solvents dramatically influence a metal catalysts activity with significant implications for

catalyst design. However, measurements are indirect and involve adsorption isotherm

models which potentially reduces the reliability of the measurements. Computational,

implicit solvation models predict exergonic solvation effects for phenol adsorption, failing

to agree with measurements even qualitatively. In this study, an explicit, hybrid QM/MM

approach for computing solvation free energies of adsorption is developed, solvation free

energies of phenol adsorption are computed, and experimental data for solvation free

energies of phenol adsorption are reanalyzed using multiple adsorption isotherm models.

Explicit solvation calculations predict an endergonic solvation free energy of 1.48±0.04eV

for phenol adsorption on Pt(111), that agrees with measurements to within the experimental

uncertainty. Computed adsorption free energies of solvation of carbon monoxide, ethylene

glycol, and phenol over the (111) facet of Pt and Cu suggests that liquid water destabilizes

all adsorbed species, with the largest impact on the largest adsorbates.

4.2 Introduction

A heterogeneously catalyzed process starts with the adsorption of at least one of the

reactants and ends with the desorption of the products.1 Activation barriers have also been

correlated to adsorption energies for about ninety years through, e.g., the well-known

Bronsted-Evans-Polanyi (BEP) relationships that connect thermodynamics with kinetics.2,


3, 4, 5
Thus, understanding adsorption and desorption processes is critical for heterogeneous

catalysis, and adsorption processes at gas-solid interfaces have been studied for many

83
decades.6, 7, 8, 9, 10 For solid-liquid interfaces, less progress has been made, largely due to

the added complexity of the solvent molecules that can both directly and indirectly interact

with the adsorbate through, e.g., hydrogen bonding and by changing the electronic structure

of the adsorbent. Nevertheless, it is well-known that solvents can affect important figures

of merit in catalysis, such as activity and selectivity, that go beyond reduced mass transfer

rates and non-ideality effects that determine a reactants activity/fugacity at a given liquid

concentration.11, 12, 13, 14 These intrinsic solvent effects originate from one or a combination

of the followings: nature of solvent (pH, polarity, H-bonding ability etc.),15, 16, 17 leading to

changes in relative stabilization of adsorbed moieties,18, 19, 20 participation of solvent in the

catalytic cycle or reaction coordinate,21 and competitive adsorption between solvent and

adsorbate molecules.22, 23, 24


Consequently, to design catalysts for enhanced activity,

selectivity, and stability, it is critical to understand the root causes of solvent effects on

adsorption processes in heterogenous catalysis. Such an understanding becomes the more

important the larger the solvation effects are on the adsorption free energy of the various

species in a catalytic system. Recently, aqueous solvation effects have been measured on

the free energy of adsorption for phenol, benzyl alcohol, benzaldehyde, and cyclohexanol

over Pt catalyst at 298 K25 and for phenol over Pt and Rh catalysts at 4 temperatures (283,

288, 298, and 314 K).26, 27 Importantly, for phenol adsorption on supported Pt catalysts, it

has become possible to deconvolute the adsorption isotherm data for the different surface

facets, which enabled the measurement of the aqueous solvent effect on the phenol

adsorption free energy on the Pt(111) facet, which again facilitates a correlation to

computational studies. Experimentally measured aqueous solvation effects are large (> 1.4

eV at 298 K) and endergonic for phenol on Pt(111), suggesting that the kinetic properties

84
of Pt nanoparticles for phenol catalysis are a strong function of solvent properties with

significant implications for catalyst design in condensed media. We highlight that at 298

K and at 500 K a change in a rate controlling free energy barrier of 0.1 eV can change the

overall rate by a factor 50 and factor 10, respectively. However, measurements are indirect

and involve adsorption isotherm models which potentially reduces the reliability of the

measurements. Also, computational, implicit solvation models predict significantly smaller

and exergonic solvation effects on the free energy of phenol adsorption on Pt(111), failing

to even qualitatively agree with the experimental measurements.26

Despite the availability of fairly established methods for computing the energetics

of adsorption reactions at gas-solid interfaces, methods capable of quantifying the

adsorption (free) energies at solid-liquid interfaces are less developed. This is partly due to

the limited availability of experimental24, 28, 29, 30 and theoretical studies31, 32, 33, 34, 35, 36, 37, 38,
39, 40, 41, 42, 43
and partly due to inherent intricacy of a reaction system comprised of both a

complex heterogeneous catalyst and a condensed phase. Although “on the fly” electronic

structure calculations in brute force ab initio molecular dynamics (AIMD)44, 45, 46

simulations are well-suited for this task, as a result of the large computational cost

associated with quantum mechanical calculations and configuration space sampling,

AIMD simulations for computing adsorption free energies from a condensed phase is

currently, practically impossible with small confidence interval from configuration space

sampling. Typically, the high computational cost of AIMD constrains both the size of the

simulation system to a few hundred atoms and the time scale of simulations to a few

picoseconds.47, 48, 49 Alternatively, implicit solvation schemes50, 51 and classical force field

simulations offer a practical approach that is computationally affordable. However, force

85
field simulations require a reliable potential for all fluid components with the adsorption

site/metal surface which, except for perhaps water, is hardly available due to a lack of

reliable experimental data. Similarly, performance reliability is largely unknown for

adsorption on metal surfaces with implicit solvation models due to the need to parameterize

the implicit solvation models against experimental data that are again hardly available or

possess unknown error bars. We highlight that for homogeneous systems, implicit

solvation models have been exceptionally successful despite their limitations in capturing

anisotropic site-specific interactions (e.g. hydrogen bonding).52, 53

To overcome these challenges a number of research groups have recently developed

combined quantum mechanical/molecular mechanical (QM/MM) approaches for

predicting free energy changes for processes on solid-liquid interfaces.36, 38, 39, 54, 55, 56 In

this class of simulations, the essential region of interest that can often not be properly

described by classical force fields is described quantum mechanically, while the remainder

of the system is described at the MM level of theory, enabling a balance of computational

cost and accuracy. For example, in a metal-catalyzed adsorption reaction, adsorbate, active

site and its immediate metal neighbors are treated from first principles, while the bulk of

the solvent molecules and the nonreactive part of the simulation system are treated using

classical force fields. In this regard, we previously developed a computationally affordable

and reliable method for computing free energy changes in the presence of solvents using a

hybrid (QM/MM) approach with electrostatic embedding, named eSMS (Explicit Solvation

for Metal Surfaces)57 and applied it to several surface reactions over various transition

metal surfaces.58, 59

86
A limitation of eSMS has been that it can only be applied for computing free energy

differences of surface reactions, that is, it cannot be employed to compute free energies of

adsorption processes from a liquid-phase environment due to (i) the long reaction

coordinate in adsorption processes that increases the computational expense, and (ii) the

need to be able to compute the free energy of the quantum mechanically described part of

the simulation system within the harmonic approximation, i.e., only the classically

described part of the simulation system is sampled extensively within eSMS. In this study,

we identify a reaction coordinate for adsorption processes that enables our eSMS approach

to predict adsorption free energies at a solid-liquid interface. Next, we validated our novel

computational scheme against previous eSMS calculations of the aqueous solvation effect

on the activation free energy barrier of O-H bond cleavage of ethylene glycol, EG

(C2H6O2), over a Cu(111) catalyst surface model59 by computing the adsorption free energy

of both the reactant and transition state. Next, we compute the aqueous solvation effect on

the free energy of phenol adsorption on the Pt(111) surface, and we reanalyzed the

experimental data for this solvation free energy using multiple adsorption isotherm models.

Computational predictions at 298 K are in excellent agreement with the experimental data

with an error less than 0.10 eV which is on the order of the experimental error. Thus, these

results confirm a large endergonic aqueous solvation free energy for phenol adsorption

with significant implications for catalyst design for conversion of aromatic molecules. To

elucidate the significance of these large, computed and measured, solvation effects for the

Pt-phenol system for various other catalytic processes, we computed the aqueous solvation

effect on the adsorption free energies of carbon monoxide (CO), ethylene glycol (EG), and

phenol (Ph) over the Pt(111) and Cu(111) surfaces.

87
4.3 QM/MM free energy scheme for adsorption processes at a solid-liquid

interface

The key challenge for computing the free energy of adsorption in a liquid-phase

environment with a hybrid QM/MM scheme is identifying a thermodynamic cycle that

requires extensive configuration space sampling only for the classically described part of

the simulation system, i.e., the harmonic approximation is meaningful for the quantum

mechanically described subsystem in the reactant and product state, and that involves a

relatively smooth reaction coordinate that does not involve overcoming large free energy

barriers, such that, only few free energy perturbation (FEP) steps are required for

computing the overall free energy difference. Figure 4.1 illustrates such a thermodynamic

cycle that is explained below. We note that a detailed computational methods section is

presented in the Supporting Information. The key idea of our thermodynamic cycle is to

not compute the desorption of an adsorbed moiety into a liquid phase but to only compute

the solvent effect on the desorption process. In this way, free energy differences for the

overall process are smaller, and desorption itself is only the gas-phase desorption which

can be described using the harmonic approximation or correction formulas to the harmonic

approximation.60, 61 Thus, a potential reaction coordinate for computing the solvent effect

on adsorption processes involves four steps:

Step I to II in Figure 4.1: Classical electrostatic interaction removal. In our

QM/MM description of the reaction system the electrostatic interaction is an electric force

between the adsorbed species and the metal atoms in direct vicinity of the adsorbed moiety

with the solvent molecules, governed by magnitude and distance of partial charges. To

88
remove the electrostatic potential, first, partial charges on the “metal-adsorbate cluster”

(see Figure C.3) and “metal cluster” (obtained by removing adsorbate atoms from the

metal-adsorbate cluster) are computed using a suitable charge model (we used NPA62 and

DDEC663 in this study). We note that the metal cluster is a “non-physical” state in which

the adsorbate is present but does not electrostatically interact with the surrounding water

molecules that are treated classically, i.e., the partial charges on the adsorbate are set to

zero (𝑄𝑎𝑑𝑠 = 0). Next, we insert adequate “non-physical” states in between by linearly

reducing the magnitude of partial charges on the QM atoms going from the metal-adsorbate

to the metal cluster. Table C.3 summarizes the charges on the QM atoms of these two states

for different adsorbates studied in this work.

We can use our conventional eSMS methodology for computing the solvent effect

owing to the removal of the classical electrostatic interactions between the adsorbed

species and the solvent molecules. The total energy function of our eSMS method is given

in Equation 4.1 whose derivation can be found elsewhere64.

𝑆𝑢𝑟𝑓𝑎𝑐𝑒 𝑐𝑙𝑢𝑠𝑡𝑒𝑟
𝐸𝑇 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 ) = 𝐸𝑄𝑀 (𝑟𝑄𝑀 ) − 𝐸𝑄𝑀 (𝑟𝑄𝑀 ) +

𝑀𝑒𝑎𝑛𝐹𝑖𝑒𝑙𝑑(100)
〈𝛹 |𝐻𝑒𝑓𝑓 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 )| 𝛹〉 +

1 𝑒𝑙𝑒𝑐+𝑣𝑑𝑊 𝑀𝑒𝑎𝑛𝐹𝑖𝑒𝑙𝑑(100)
[100 ∑100
𝑗=1 [𝐸𝑗,𝑀𝑀+𝑄𝑀⁄𝑀𝑀 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 )] −
𝑄𝑖 =0,𝑖∈𝑄𝑀

𝑒𝑙𝑒𝑐+𝑣𝑑𝑊 𝑀𝑒𝑎𝑛𝐹𝑖𝑒𝑙𝑑(100) 𝑒𝑙𝑒𝑐+𝑣𝑑𝑊


𝐸𝑗,𝑀𝑀+𝑄𝑀 ⁄𝑀𝑀 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 )] + 𝐸𝑗,𝑀𝑀+𝑄𝑀 ⁄𝑀𝑀 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 ) (4.1)

where the first two terms are evaluated for a periodic slab, using the VASP65, 66 program

package, and a QM cluster in vacuum, with the help of the TURBOMOLE program

package,67, 68, 69 respectively. The combination of the third to fifth term is a QM calculation

89
in a mean field of MM water molecules using the periodic electrostatic embedded cluster

method (PEECM)70 with the fixed-charge approximation (the fixed charge approximation

has been validated for our eSMS approach in reference64). We note that the number 100 in

the equation indicates that 100 MM conformations, selected equally spaced from 1 ns

molecular dynamic (MD) simulations (10 ps apart), were used to represent the mean field

of the MM water molecules. In other words, the first term constitutes the conventional gas

phase energy computed for a periodic slab model and the combination of the second to fifth

term is the non-classical part of the electrostatic interaction energy of a mean-field liquid

phase environment with the adsorbed moiety and metal surface described quantum

mechanically by a cluster model. We have previously shown that unlike gas-phase

adsorption energies on metal particles, this non-classical part of the electrostatic interaction

energy is fast converging with metal cluster size such that reliable energies can be

computed for relatively small metal clusters of ~50 metal atoms. Finally, the last term

accounts for the classical interaction energy within the MM subsystem and its interaction

with the QM subsystem. Since the QM geometry remains the same during the removal of
𝑒𝑙𝑒𝑐+𝑣𝑑𝑊
the electrostatic interaction, all terms except 𝐸𝑗,𝑀𝑀+𝑄𝑀 ⁄𝑀𝑀 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 ) remain the same in

the energy function of every intermediate state and hence, they do not need to be computed

in a FEP step calculation. In other words, removing the classical electrostatic interaction

involves no QM calculations but only classical free energy calculations.

The QM atom coordinates of the metal-adsorbate cluster state (state I in Figure 4.1)

are ideally those optimized in liquid water and not those optimized in vapor phase. At one

point in the thermodynamic cycle the coordinates need to be changed to those in vapor.

The free energy change for this change in QM atom coordinates involves standard eSMS

90
free energy calculations that involve a number of QM and MM calculations but that are

reasonably fast given the small change in QM atom coordinates during adsorption of

typical adsorbates on a metals surface.11, 57, 58, 59 We typically perform these calculations

before we remove the classical electrostatic interactions.

Step II to III in Figure 4.1: Electrostatic embedding. The non-classical contribution

of the solvent molecules changing the electronic structure of the quantum system transition

from those of a metal-adsorbate system to those without adsorbate, i.e., this energy

difference constitutes the main difference between electrostatic and mechanical embedding

of the QM/MM system. The energy difference is given by:

𝑚𝑒𝑡𝑎𝑙−𝑠𝑜𝑙𝑣𝑒𝑛𝑡 𝑚𝑒𝑡𝑎𝑙−𝑎𝑑𝑠−𝑠𝑜𝑙𝑣𝑒𝑛𝑡
𝐸 𝐼𝐼𝐼 − 𝐸 𝐼𝐼 = (𝐸𝑃𝐸𝐸𝐶𝑀 − 𝐸𝑃𝐸𝐸𝐶𝑀 )+
1
[(100 ∑100 𝑚𝑒𝑡𝑎𝑙−𝑠𝑜𝑙𝑣𝑒𝑛𝑡
𝑗=1 (𝐸𝑗,𝐷𝐿𝑃𝐿𝑂𝑌 ) 𝑚𝑒𝑡𝑎𝑙−𝑠𝑜𝑙𝑣𝑒𝑛𝑡
− 𝐸𝑗,𝐷𝐿 𝑃𝐿𝑂𝑌
)−
𝑄𝑖 =0,𝑖∈𝑄𝑀
1
(100 ∑100 𝑚𝑒𝑡𝑎𝑙−𝑎𝑑𝑠−𝑠𝑜𝑙𝑣𝑒𝑛𝑡
𝑗=1 (𝐸𝑗,𝐷𝐿𝑃𝐿𝑂𝑌 ) 𝑚𝑒𝑡𝑎𝑙−𝑎𝑑𝑠−𝑠𝑜𝑙𝑣𝑒𝑛𝑡
− 𝐸𝑗,𝐷𝐿 𝑃𝐿𝑂𝑌
𝑚𝑒𝑡𝑎𝑙−𝑠𝑜𝑙𝑣𝑒𝑛𝑡
)] − (𝐸𝑐𝑙𝑢𝑠𝑡𝑒𝑟 −
𝑄𝑖 =0,𝑖∈𝑄𝑀
𝑚𝑒𝑡𝑎𝑙−𝑎𝑑𝑠−𝑠𝑜𝑙𝑣𝑒𝑛𝑡
𝐸𝑐𝑙𝑢𝑠𝑡𝑒𝑟 ) (4.2)

where metal-solvent and metal-ads-solvent superscripts represent state III and II,

respectively (the first and last terms in parenthesis are performed using the TURBOMOLE

program package). This step involves a QM calculation; however, the free energy change

calculation is not a strong function of solvent coordinates (only the computational

description changes of a “non-physical” state changes) and although the free energy

differences are significant, multiple independent calculations suggest that this free energy

change can efficiently be computed without introducing intermediate states (which would

practically be impossible). Here, we also note that given the magnitude of the free energy

change from state II to III, electrostatic embedding is superior to mechanical embedding

91
(that does not consider these non-classical electrostatic contributions) for QM/MM

calculations of metallic systems where electrons move freely.

Step III to IV in Figure 4.1: Van der Waals interaction removal. Another interaction

between solvent and adsorbate-metal moiety are the classical van der Waals (vdW)

interactions. To remove the vdW interactions, we followed the classical approach of

scaling of the distance shifted Lennard-Jones (LJ) potentials.71 This approach allows for a

smooth transition in molecular simulations between real and dummy atoms, for which all

atomic interactions have vanished. The LJ potential that we implemented in the DL_POLY

program package to annihilate or create an atom has the following functional form:

𝐴 𝐵
𝑉𝑙𝑗 = (1 − 𝜆) [(𝑟2+𝛿𝜆)6 − (𝑟 2+𝛿𝜆)3] (4.3)

where 𝑟 is the interatomic distance between two atoms, and A and B are repulsive and

attractive LJ parameters, respectively, calculated as a product of LJ parameters specific for

each atom type of the interacting particles (𝐴 = 4𝜀𝜎12 , 𝐵 = 4𝜀𝜎 6 ; see Table C.1 for 𝜀 and

𝜎 values we used in this work). For 𝜆 = 0, the vdW interactions fully exist and an increase

in 𝜆 leads to a smooth disappearance of the vdW interactions. Finally, at 𝜆 = 1, the

interaction between the solvent molecules and the adsorbate is zero everywhere,

corresponding to the final state where the adsorbate atom is transformed to a dummy

particle. The shift parameter 𝛿 comes into play for 𝜆 larger than zero. It allows for a smooth

transition from the original LJ potential to no interaction. As recommended by Zacharias

et al.71, the parameter 𝛿 was chosen as the square of the vdW radius of the interacting atoms

to allow the smoothest transition between an atom present on the surface and filling the

cavity after molecule annihilation. Table C.4 summarizes all 𝛿 parameter values for all pair

92
interaction terms and, as an example, Figure C.2 illustrates the LJ potential as a function

of distance between the oxygen atom of an adsorbed CO and the oxygen atom of a TIP3P

water molecule for different values of λ.

To enable a reliable free energy difference calculation, the removal of the vdW

interactions between the solvent and the adsorbate is done in multiple steps by linearly

increasing 𝜆 from 0 to 1. Analogous to the removal of the classical electrostatic

𝑒𝑙𝑒𝑐+𝑣𝑑𝑊
interactions, in the energy function, equation 4.1, only the 𝐸𝑗,𝑀𝑀+𝑄𝑀 ⁄𝑀𝑀 (𝑟𝑄𝑀 , 𝑟𝑀𝑀 ) term

changes between two consecutive states and no QM calculations are required for this step,

i.e., only classical free energy calculations are performed at the MM level of theory. At this

point, all classical interactions between the adsorbate and the solvent have vanished.

Given that the quantum atom coordinates of the metal cluster are still those in the

gas phase adsorbed state and not those in the liquid-phase optimized metal cluster (free

site), the free energy change for this change in QM atom coordinates needs to be computed.

This involves again standard eSMS free energy calculations.11, 57, 58, 59 We note that for all

free energy calculations we employ the Bennett acceptance ratio (BAR) as free energy

estimator for all FEP steps.72 For practical convenience, we typically perform these

calculations at the very end of the solvation free energy calculation.

Step IV to V in Figure 4.1: Gas-phase desorption. We use for the adsorbate as

reference state the adsorbate in a gas-phase in equilibrium with the adsorbate in the liquid

phase. Thus, only a typical gas phase metal adsorption calculation at the QM level of theory

is required for this step. The free energy change from state I to state V constitutes the free

energy change from an adsorbed molecule surrounded by a solvent to the adsorbent (metal

93
surface) surrounded by a solvent and the adsorbate in a gas phase state. Thus, the free

energy change from state I to state IV constitutes the solvent effect on an adsorption

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙 𝑔𝑎𝑠


process, (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 − ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ), that is likely not as DFT functional

dependent as the gas phase adsorption energy.73 Also, the activity/fugacity calculation of

the adsorbate in vapor phase is often significantly easier than such a calculation in a solvent.

In this paper, we assumed the vapor phase to behave as an ideal gas. For adsorbates with

very low vapor pressure, where it is challenging to measure the fugacity, the gas-liquid

partition coefficient (equilibrium constant) can be computed from COSMO-RS

calculations that are quite reliable for this task.

4.4 Results and discussion

4.4.1 Numerical verification of the proposed thermodynamic cycle

To numerically verify the proposed methodology for adsorption/desorption

‡ ‡
processes, the aqueous-phase effect on the free energy of activation (∆∆𝐺𝑙𝑖𝑞 = ∆𝐺𝑙𝑖𝑞 −


∆𝐺𝑔𝑎𝑠 ) of the O-H bond cleavage of ethylene glycol over a Cu(111) surface at 423 K is

computed using two different approaches, the eSMS methodology for surface reactions59

and the adsorption scheme proposed here. In our proposed adsorption scheme, the solvent

effect on the activation barrier is given by Equation 4.4.

‡ 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞
∆∆𝐺𝑙𝑖𝑞 = ∆∆𝐺𝑇𝑆 − ∆∆𝐺𝑅𝑒𝑎𝑐𝑡𝑎𝑛𝑡 (4.4)

‡ 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞
where ∆∆𝐺𝑙𝑖𝑞 , ∆∆𝐺𝑇𝑆 , and ∆∆𝐺𝑅𝑒𝑎𝑐𝑡𝑎𝑛𝑡 indicate the solvent effect on the activation

barrier, on the adsorption of the transition state (TS), and on the adsorption of the reactant

state (RS), respectively.

94
Figure 4.2a illustrates the free-energy profile of the solvation effects on the

𝑙𝑖𝑞→𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞
desorption of the RS and TS (∆∆𝐺𝑃ℎ = −∆∆𝐺𝑃ℎ ). For both the RS and TS, the

transition of the QM atom coordinates leads only to a small free energy change relative to

the overall solvent effect. The reactant state is more destabilized in liquid water than the

transition state leading to a significantly reduced activation free energy barrier. Both, the

RS and TS are similarly destabilized through the creation of a solvent cavity on the surface

as shown in the change in free energy during the removal of the van-der-Waals/Lennord-

Jones interactions (see Table C.5). In contrast, the classical electrostatic interactions

𝑙𝑖𝑞→𝑔𝑎𝑠 𝑙𝑖𝑞→𝑔𝑎𝑠
stabilize the RS and TS in solution (∆∆𝐺𝐸𝐿−𝑅𝑚𝑣,𝑅𝑆 = 0.49, ∆∆𝐺𝐸𝐿−𝑅𝑚𝑣,𝑇𝑆 = 1.25 𝑒𝑉).

The difference in classical stabilization can be attributed to the significant change in partial

charges of the QM system going from RS to TS (see Table C.3a). In particular, the charge

of the cleaved hydrogen atom has transitioned from 0.47 e in RS to -0.16 e in TS, explaining

the 0.76 eV difference in solvent stabilization due to classical electrostatic interactions.

Overall, the electrostatic interaction effect is, however, reduced by the non-classical

electrostatic interactions from section “II-to-III”. The latter effect is ignored in molecular

𝑙𝑖𝑞→𝑔𝑎𝑠 𝑙𝑖𝑞→𝑔𝑎𝑠
embedding calculations (∆∆𝐺𝐼𝐼−𝑡𝑜−𝐼𝐼𝐼,𝑅𝑆 = −0.47, ∆∆𝐺𝐼𝐼−𝑡𝑜−𝐼𝐼𝐼,𝑇𝑆 = −0.68 𝑒𝑉).

Therefore, we conclude that electrostatic embedding of the QM/MM system is crucial in

any hybrid QM/MM procedure for metallic systems.

Finally, when considering the vibrational contributions, the overall aqueous solvent

𝑔𝑎𝑠→𝑙𝑖𝑞
effect on the adsorption free energy (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) is 0.60 ± 0.03 eV for the RS and


0.10 ± 0.05 eV for the TS, such that the solvent effect on the activation barrier (∆∆𝐺𝑙𝑖𝑞 ) is

95
−0.50 ± 0.06 eV. This agrees very well with our previous study, where we used eSMS for

‡,𝑒𝑆𝑀𝑆
surface reactions and found ∆∆𝐺𝑙𝑖𝑞 = −0.56 ± 0.04 𝑒𝑉.59

4.4.2 Phenol adsorption on Pt(111) at 298 K in liquid water

Using the proposed QM/MM-FEP procedure, we computed the solvent effect on

𝑔𝑎𝑠→𝑙𝑖𝑞
the adsorption free energy of phenol at low coverage (∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0)) and at 298 K

to be 1.48 ± 0.04 eV. To compare our result to the experimental work from Singh et al. 26,
74
, we used the adsorption model from Nitta et al.75, 76
that considers the ability of

adsorbates to cover multiple active/metal sites and that can be derived from statistical

mechanics. According to the experimental work, only hydrogen and phenol are present in

the aqueous phase and can adsorb on the Pt surface, hence

𝑔−𝑙 𝜃𝑃ℎ 𝜃𝑃ℎ


𝑃ℎ(𝑔) + 𝑛 ∗ (𝑙𝑖𝑞) ↔ 𝑃ℎ𝑛∗ (𝑙𝑖𝑞) 𝐾𝑎𝑑𝑠,𝑃ℎ = 𝑓𝑃ℎ = 𝑃𝑃ℎ (4.5)
( )𝜃𝑛 ( )𝜃𝑛
1 𝑏𝑎𝑟 ∗ 1 𝑏𝑎𝑟 ∗

2 2
𝑔−𝑙 𝜃𝐻 𝜃𝐻
𝐻2 (𝑔) + 2 ∗ (𝑙𝑖𝑞) ↔ 2𝐻 ∗ (𝑙𝑖𝑞) 𝐾𝑎𝑑𝑠,𝐻2 = 𝑓𝐻 = 𝑃𝐻 (4.6)
2 )𝜃 2 ( 2 )𝜃 2
( 1 𝑏𝑎𝑟 ∗
1 𝑏𝑎𝑟 ∗

1 = 𝜃∗ + 𝜃𝐻 + 𝑛𝜃𝑃ℎ (4.7)

where 𝜃𝑃ℎ , 𝜃𝐻 , and 𝜃∗ are the surface coverage of phenol, hydrogen and free site (water

everywhere), respectively, variable 𝑛 indicates the number of sites that a phenol molecule

𝑔−𝑙
can occupy when adsorbed on the surface, 𝐾𝑎𝑑𝑠 is the equilibrium constant of adsorption

of an adsorbate on the surface from its free state in the gas phase, 𝑓𝑃ℎ , and 𝑓𝐻 are phenol

and hydrogen fugacity in the gas phase, respectively, that can be reduced to the ideal gas

partial pressure under experimental conditions. We note that we have defined the “surface

coverage of species 𝑖” as

96
𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑠𝑢𝑟𝑓𝑎𝑐𝑒 𝑠𝑝𝑒𝑐𝑖𝑒𝑠 𝑖
𝜃𝑖 ≡ (4.8)
𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑠𝑢𝑟𝑓𝑎𝑐𝑒 𝑎𝑡𝑜𝑚𝑠

i.e., the “real” surface coverage is 𝑛𝑖 𝜃𝑖 . By performing some algebra, we reach as the

equation for the adsorption isotherm (see the Supporting Information for the full

derivation)

𝑛 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑙𝑖𝑞
𝑔𝑎𝑠 1−𝑛𝜃 −{∆∆𝐺𝑃ℎ (𝜃𝑃ℎ =0)+𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ }
𝑃ℎ
𝜃𝑃ℎ = 𝐾𝑎𝑑𝑠,𝑃ℎ (𝜃𝑃ℎ = 0)𝐾𝐻,𝑃ℎ 𝐶𝑃ℎ (1+𝑐𝑜𝑛𝑠𝑡 ) 𝑒𝑥𝑝 [ ] (4.9)
𝑅𝑇

𝑔𝑎𝑠
where 𝐾𝑎𝑑𝑠,𝑃ℎ (𝜃𝑃ℎ = 0) is gas-phase adsorption equilibrium constant of phenol at low

𝑔𝑎𝑠 77 𝑔𝑎𝑠 78
coverage (at 298 K: ∆𝐻𝑃ℎ = −200 𝑘𝐽⁄𝑚𝑜𝑙 , ∆𝑆𝑃ℎ = −14.7𝑅 ), 𝐾𝐻,𝑃ℎ is phenol

Henry’s constant in water (at 298 K: 𝐾𝐻,𝑃ℎ = 5 × 10−4 𝑏𝑎𝑟⁄𝑀 79


), 𝐶𝑃ℎ is phenol

concentration in the aqueous phase, 𝑐𝑜𝑛𝑠𝑡 is a constant taking into account a hydrogen

𝑙𝑖𝑞
coverage effect (𝑐𝑜𝑛𝑠𝑡 = 0 indicates zero hydrogen coverage approximation), 𝛼𝑃ℎ,𝑃ℎ is a

constant associated with adsorbed phenol-phenol lateral interactions in liquid, 𝑅 is the gas

𝑔𝑎𝑠→𝑙𝑖𝑞
constant, 𝑇 is the system temperature, and finally ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) is the solvent

effect on the low coverage phenol adsorption defined as

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙 𝑔
∆∆𝐺𝑃ℎ = ∆𝐺𝑎𝑑𝑠,𝑃ℎ − ∆𝐺𝑎𝑑𝑠,𝑃ℎ (4.10)

𝑔−𝑙 𝑙 𝑠𝑜𝑙𝑣𝑎𝑡𝑖𝑜𝑛
∆𝐺𝑎𝑑𝑠,𝑃ℎ = ∆𝐺𝑎𝑑𝑠,𝑃ℎ + ∆𝐺𝑃ℎ (4.11)

𝑙 𝑔−𝑙
We note here that ∆𝐺𝑎𝑑𝑠,𝑃ℎ and ∆𝐺𝑎𝑑𝑠,𝑃ℎ are the adsorption free energy of phenol in the

aqueous phase from its solvated state in solution and from its free state in the gas phase,

𝑠𝑜𝑙𝑣𝑎𝑡𝑖𝑜𝑛
respectively, and ∆𝐺𝑃ℎ is the solvation free energy of phenol which can be obtained

from Henry’s constant at relevant temperatures79.

97
Next, we fitted equation 4.9 to the experimental data at different phenol

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑙𝑖𝑞
concentration at 298 K to extract ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) and 𝛼𝑃ℎ,𝑃ℎ . Figure 4.3 shows the

best fits for different 𝑛 and 𝑐𝑜𝑛𝑠𝑡 that are experimentally not known. A change of 𝑐𝑜𝑛𝑠𝑡

𝑔𝑎𝑠→𝑙𝑖𝑞
from 0 to 1 fortunately only changes ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) by less than 0.10 eV; however,

for a large 𝑐𝑜𝑛𝑠𝑡 (high hydrogen coverage) such as 10, this change is about 0.25 eV.

Experimental work suggests that 𝑐𝑜𝑛𝑠𝑡 < 1.0 since the partial pressure of hydrogen is

quite low such that the assumption of no hydrogen coverage (𝑐𝑜𝑛𝑠𝑡 = 0) is acceptable.

Furthermore, the plots in Figure 4.3 indicate that the fits with lower 𝑛 predict the higher

concentration data better, while a higher 𝑛 leads to better fits at low phenol concentration

(see also Figure C.5). We attribute this observation and the fact that, for a fixed 𝑛 value,

we fit a negative lateral phenol-phenol interaction parameter 𝛼, to different phenol

orientations on the surface at low and high coverages. At low coverage, phenol adsorbs

horizontally parallel to the surface, while increasing the phenol coverage forces the

adsorbed phenol molecules to slant upwards to occupy less space on the surface for

𝑔𝑎𝑠→𝑙𝑖𝑞
additional adsorbed phenol molecules (see Figure C.6). Fortunately, ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0)

is not very sensitive to the 𝑛 value and we conclude that independent of these intricacies,

𝑔𝑎𝑠→𝑙𝑖𝑞
the experimental ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) value is in an excellent agreement with our

computational prediction to within an error of less than 0.10 eV. Aqueous solvation effects

are indeed large for phenol adsorption on Pt(111) with significant consequences for metal

catalyst design for conversion of reactants such as phenol.

98
4.4.3 Aqueous-phase effect across adsorbates and metals

To elucidate the significance of these large, computed and measured, solvation

effects for the Pt-phenol system for various other catalytic processes, we computed the

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙
aqueous solvation effect on the adsorption free energies (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 −
𝑔𝑎𝑠
∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) of three different adsorbates, ethylene glycol (EG), carbon monoxide (CO),

and phenol (Ph), over the (111) surface facet of Cu and Pt. In addition to our QM/MM-

FEP calculations, we also used two different well-known implicit solvation methods,

iSMS80 and VASPsol81, 82 (see Table 4.1). Positive solvent effects on the adsorption free

energies indicate that liquid water destabilizes the adsorbed moiety. In this regard, we

found that an aqueous phase destabilizes all three adsorbed species and that the impact is

most pronounced for phenol and smallest for CO, irrespective of the metal surface

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞


(∆∆𝐺𝑃ℎ > ∆∆𝐺𝐸𝐺 > ∆∆𝐺𝐶𝑂 ). We attribute the destabilization of the

adsorbates on the surface primarily to the solvent cavity formation energy and the desire

of water molecules to adsorb on a metal surface.

Furthermore, the metal identity also plays a key role in the solvation effect,

primarily through an indirect influence of the solvent environment on the electronic

structure of the metal atoms that in turn affects the stability of the adsorbed moieties59.

Table 4.1 shows that all three adsorbed moieties are more destabilized on the Pt(111)

surface relative to the Cu(111) surface. To elucidate the root cause for the different solvent

effects across the metals and the adsorbates, we developed a linear model for describing

𝑦−𝑦̅ ̅̅̅
𝑓1 −𝑓 1 ̅̅̅
𝑓2 −𝑓 2
the variability of the solvation free energy of adsorption, = 𝛼1 + 𝛼2 (𝑠 is
𝑠𝑦 𝑠𝑓1 𝑠 𝑓2

standard deviation and bar sign shows mean), in which 𝑓1 and 𝑓2 are the net charge on the

99
adsorbate and number of sites occupied by the adsorbate (#Site), respectively (see Table

C.6). Based on the best fit parameters (𝛼1 = 0.45, 𝛼2 = 0.62), we conclude that different

solvent effects arise almost equally from different charge-transfer effects (net charge on

adsorbate) across metals and adsorbates and the size of the molecule (number of sites

occupied). We note that we also tested hydrogen bond formation between the adsorbate

and TIP3P water molecules as well as gas phase adsorption energy as possible descriptors,

but we found the best fit using the net charge and #Site.

4.4.4 Analysis of the specific contributions to the solvation free energy of


adsorption

Figure 4.2 illustrates the free-energy profiles of the aqueous solvent effect on the

desorption of the various adsorbed moieties. The profiles are comprised of five sub-

sections: To OPT adsorbate in Liquid, Remove Electrostatic potential, II-to-III

(Electrostatic Embedding), Remove Lennard-Jones potential, and To OPT site in Gas,

whose individual values are listed in Table 4.2.

Classical electrostatic contribution to the solvent effect. Removal of the classical

electrostatic interactions between the adsorbate and the liquid water molecules generally

increases the total free energy of the system, i.e., the classical electrostatic interactions

stabilize all adsorbed species. As expected, this effect is most dominant for ethylene glycol,

somewhat smaller for phenol and least relevant for CO, which is attributed to the fact that

attractive forces between charged species decrease the energy of the system while repulsive

forces between charged species have opposite effect. In this context, perturbing charges of

the adsorbate atoms from its original magnitude to zero decreases the attractive forces

100
between the adsorbate and the water molecules and in turn increases the energy of the total

system.

Electrostatic embedding contributions to the solvent effect (II-to-III). As described

above, removal of the non-classical electrostatic contribution of the water molecules

changing the electronic structure of the quantum system generally reduces the free energy

of system, reducing the adsorbate stabilization from the classical electrostatic

contributions.

Lennard-Jones contribution to the solvent effect. Removing the LJ interactions

between the adsorbed species and the water molecules generally decreases the free energy

of the system as shown in Table 4.2 and Figure 4.2. Slow removal of the LJ interaction

energy leads to a reduction of the water cavity for the adsorbates. Cavity formation is

generally an endergonic process and given that any adsorbate requires a similar-sized water

cavity on any metal surface, we found this LJ removal contribution to be almost

independent of metal surface (see Table 4.2). This again explains why larger adsorbates

that require larger water cavities are often also more destabilized on a metal surface.

Transition of metal cluster coordinates (To OPT site in Gas) contribution to the

solvent effect. This contribution is significant for adsorption of molecules that bind strongly

to the surface in the gas phase and thus distort the coordinates of the surface metal atoms

upon adsorption. For EG, CO, and phenol on Pt(111), the zero-point corrected adsorption

energies, -0.35, -1.73,and -2.29 eV, respectively, correlate with this contribution to the

overall solvent effect of 0.14, 0.16, and 0.45 eV, respectively.

101
4.4.5 Implicit (iSMS and VASPsol) vs. explicit (QM/MM-FEP) solvation
models

Finally, we compare the results of the solvent effect on the adsorption free energies

of EG, CO and phenol over Pt(111) and Cu(111) computed with our QM/MM-FEP scheme

against two different implicit solvation schemes: VASPsol and our iSMS methodology.

Table 4.1 shows that both implicit solvation methods fail to capture the true solvent effect,

probably because they underestimate the endergonic cavity formation energy. In contrast,

our explicit solvation scheme agrees quite well with experimental data probably because

the metal-water force field is specifically optimized for this free energy contribution.83

Interestingly, both implicit solvation schemes predict similar solvation free energies at 298

K, the temperature for which the VASPsol parameters have been optimized. While iSMS

permits a straightforward change in system temperature, we could only adjust the dielectric

constant for water with temperature in VASPsol and otherwise used the default VASPsol

parameters. Thus, at 423 K the solvation free energy prediction deviate significantly

between the two implicit solvation methods. To improve implicit solvation models for

metallic surface systems, we recommend focusing on the cavity formation energy

description.

4.4.6 Enthalpic and entropic contributions: Van’t Hoff plots

To compute the enthalpic and entropic contributions to the solvent effect on the

adsorption free energies, we repeated our explicit solvation calculations (QM/MM-FEP)

for CO and phenol on Pt(111) at five different temperatures within the temperature interval

of 285 to 353 K and constructed van’t Hoff plots shown in Figure 4.4. The results indicate

that the enthalpic and entropic contributions to the solvation free energy of adsorption are

102
of equal size for phenol and that phenol adsorption on Pt(111) in liquid water is destabilized

𝑔𝑎𝑠→𝑙𝑖𝑞
both enthalpically and entropically (∆∆𝐻𝑃ℎ = 0.71 ± 0.12 𝑒𝑉,

𝑔𝑎𝑠→𝑙𝑖𝑞
−𝑇∆∆𝑆𝑃ℎ @298𝐾 = 0.77 ± 0.09 𝑒𝑉). For CO adsorption, our error bars for the free

energy of solvation are too large to unambiguously determine the enthalpic and entropic

𝑔𝑎𝑠→𝑙𝑖𝑞
contributions with a van’t Hoff plot (∆∆𝐻𝐶𝑂 = 0.11 ± 0.22 𝑒𝑉,

𝑔𝑎𝑠→𝑙𝑖𝑞
−𝑇∆∆𝐻𝐶𝑂 @298𝐾 = 0.38 ± 0.19 𝑒𝑉). Still it appears that the entropic solvation

effect is not negligible relative to the enthalpic contribution.

Phenol on Pt(111), experimental vs. computational: Singh et al.25, 27, 74 recently also

studied phenol adsorption in water at four different temperatures (283, 288, 298, 314 K) to

acquire experimental enthalpic and entropic solvation contributions. We also fitted

equation 4.9 to their measured data (see Figures 4.3 and C.7 to C.9) and Table 4.3 shows

𝑔𝑎𝑠→𝑙𝑖𝑞
the extracted ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) values at the different temperatures. Overall, the

agreement between the computed (QM/MM-FEP) and experimental solvation free energies

of adsorption is excellent (an error of ~0.10 eV) for the low coverage adsorption free energy

of phenol on Pt(111) at all temperatures. However, the experimental solvation enthalpies

and entropies of phenol adsorption in water disagree with our computational data. Very

large enthalpic solvation effects are predicted that are reduced by entropic contributions

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞
(∆∆𝐻𝑃ℎ = 2.33 𝑒𝑉, ∆∆𝑆𝑃ℎ = 3.10 𝑚𝑒𝑉/𝐾). Intuitively, we fear that the

solvation effect on the enthalpy of adsorption is too large and an artifact of the small

experimental temperature interval in the Van’t Hoff plot together with the sizable error

bars of the experimental solvation effects on the Gibbs free energies of adsorption.

103
4.5 Conclusion

In this study, we have developed an explicit solvation scheme, based on hybrid

QM/MM-FEP calculations, for quantifying solvation effects on adsorption free energies

for heterogeneous catalysis applications. To numerically verify the proposed scheme, the

solvent effect on the activation free energy of O-H bond cleavage in EG over Cu(111) has

been computed through two adsorption free energy of solvation calculations (reactant and

transition state) and by performing conventional eSMS calculations for surface reactions.

Using the adsorption calculations, we predict a solvation effect on the activation free

energy of −0.50 ± 0.06 eV which agrees very well with our conventional eSMS


calculations (∆∆𝐺𝑙𝑖𝑞 = −0.56 ± 0.04 𝑒𝑉) and which highlights that the sampling error is

similarly small in the new scheme that has the added benefit of being able to describe

adsorption events.

Next, we computed solvation free energies of phenol adsorption on Pt(111) and

reanalyzed experimental data for solvation free energies of phenol adsorption on this

surface using multiple adsorption isotherm models. Our explicit solvation calculations

predict an endergonic solvation free energy of 1.48±0.04eV which is in excellent

agreement with measurements to within the experimental uncertainty. Implicit solvation

schemes such as VASPsol and iSMS fail to capture the true solvent effect, probably because

they underestimate the endergonic cavity formation energy. To elucidate the significance

of these large, computed and measured, solvation effects for the Pt-phenol system for

various other catalytic processes, we computed the aqueous solvation effect on the

adsorption free energies of carbon monoxide, ethylene glycol, and phenol over the Pt(111)

104
and Cu(111) surfaces. Irrespective of metal surface, the water environment destabilized all

𝑔𝑎𝑠→𝑙𝑖𝑞
three adsorbed species with the largest impact on the largest adsorbate (∆∆𝐺𝑃ℎ >

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞
∆∆𝐺𝐸𝐺 > ∆∆𝐺𝐶𝑂 ) which is attributed to both the size of the molecule (number of

sites occupied) and corresponding solvent cavity formation and the different charge-

transfer effects (net charge on adsorbate) across the metals and adsorbates.

4.6 Acknowledgements

We gratefully acknowledge financial support from the U.S. Department of Energy,

Office of Basic Energy Science, Catalysis Science program under Award DE-

SC0007167. This work was also supported by the South Carolina State Center for

Strategic Approaches to the Generation of Electricity (SAGE). Computational resources

have been provided by XSEDE facilities located at Texas Advanced Computing Center

(TACC) and San Diego Supercomputer Center (SDSC) under grant number TG-

CTS090100 and High Performance Computing clusters located at the University of South

Carolina.

105
4.7 Tables and Figures

Table 4.1: Aqueous-phase effect on the free energy of low coverage adsorption
𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙 𝑔𝑎𝑠
(∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 − ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) in eV for ethylene glycol over Cu(111) and
Pt(111) at 423 K, for CO on Cu(111) and Pt(111) at 298 K, and for phenol over Cu(111)
and Pt(111) at 298 K. The aqueous phase results are based on the proposed scheme in this
work (QM/MM-FEP) and two implicit solvation schemes: iSMS and VASPsol. For the
QM/MM-FEP calculations, the 95% confidence interval (based on limited water sampling
and multiple independent simulations) is also given.

𝒈𝒂𝒔→𝒍𝒊𝒒
∆∆𝑮𝑨𝒅𝒔𝒐𝒓𝒃𝒂𝒕𝒆
Adsorbate / 𝒈𝒂𝒔 ⁄𝑴𝑴−𝑭𝑬𝑷 QM/MM- VASPs
∆𝑮𝑨𝒅𝒔𝒐𝒓𝒃𝒂𝒕𝒆 ∆𝑮𝒈−𝒍,𝑸𝑴
𝑨𝒅𝒔𝒐𝒓𝒃𝒂𝒕𝒆
iSMS
Surface / T(K) FEP ol

Ethylene Glycol /
0.45 1.05 ± 0.03 0.60 ± 0.03 0.02 -0.15
Cu(111) / 423

Ethylene Glycol /
0.35 1.30 ± 0.05 0.95 ± 0.05 -0.19 -0.21
Pt(111) / 423

Carbon Monoxide /
-0.22 0.16 ± 0.06 0.38 ± 0.06 -0.03 -0.02
Cu(111) / 298

Carbon Monoxide /
-1.24 -0.71 ± 0.04 0.53 ± 0.04 0.01 -0.05
Pt(111) / 298

Phenol / Cu(111) /
-0.59 0.69 ± 0.05 1.28 ± 0.05 -0.19 -0.22
298

Phenol / Pt(111) /
-1.61 -0.13 ± 0.04 1.48 ± 0.04 -0.41 -0.47
298

106
𝑙𝑖𝑞→𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞
Table 4.2: Contributions to aqueous-phase effect on free energy (in eV) of low coverage desorption (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = −∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 )
of ethylene glycol at 423 K (see Figure 4.2c), of carbon monoxide at 298 K (see Figure 4.2b), and of phenol at 298 K (see Figure 4.2d)
over the (111) facets of Pt and Cu catalysts. All numbers also contain a 95% confidence interval based on limited water sampling
obtained by performing multiple independent simulations.

Remove
To OPT Remove II-to-III
Lennard- To OPT site in
Surface Adsorbate adsorbate in Electrostatic (Electrostatic
Jones Gas
Liquid potential Embedding)
potential

Ethylene glycol
0.07 ± 0.00 0.49 ± 0.00 -0.47 ± 0.02 -0.64 ± 0.00 -0.05 ± 0.00
@ 423K
Carbon Monoxide
Cu(111) 0.00 ± 0.00 0.00 ± 0.00 -0.05 ± 0.07 -0.30 ± 0.02 -0.04 ± 0.00
@ 298 K
107

Phenol
-0.02 ± 0.00 0.13 ± 0.00 -0.45 ± 0.05 -0.90 ± 0.01 -0.01 ± 0.00
@ 298 K
Ethylene glycol
-0.01 ± 0.00 0.46 ± 0.00 -0.51 ± 0.05 -0.72 ± 0.01 -0.14 ± 0.00
@ 423 K
Carbon monoxide
Pt(111) 0.00 ± 0.00 0.04 ± 0.00 -0.04 ± 0.05 -0.37 ± 0.02 -0.16 ± 0.01
@ 298 K
Phenol
-0.01 ± 0.00 0.40 ± 0.00 -0.40 ± 0.03 -1.02 ± 0.01 -0.45 ± 0.00
@ 298 K
𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞
Table 4.3: Temperature dependence of the aqueous-phase effect on free energy (∆∆𝐺𝑃ℎ ), enthalpy (∆∆𝐻𝑃ℎ ), and entropy
𝑔𝑎𝑠→𝑙𝑖𝑞
(∆∆𝑆𝑃ℎ ) of low coverage phenol adsorption on the Pt(111) catalyst surface. Experimental data obtained by fitting equation 4.9 (𝑛 =
4 and 𝑐𝑜𝑛𝑠𝑡 = 0.1) to the raw data from Singh et al.25, 26, 74 Computational data obtained by the QM/MM-FEP scheme proposed in this
𝑔−𝑙 𝑔−𝑙
study. ∆𝐻𝑎𝑑𝑠,𝑃ℎ and ∆𝑆𝑎𝑑𝑠,𝑃ℎ are the enthalpy and entropy of phenol adsorption in liquid water from its free state in the gas phase
𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙 𝑔𝑎𝑠
(𝑃ℎ(𝑔) + 𝑛 ∗ (𝑙𝑖𝑞) ↔ 𝑃ℎ𝑛∗ (𝑙𝑖𝑞)), respectively (∆∆𝑍𝑃ℎ = ∆𝑍𝑃ℎ − ∆𝑍𝑃ℎ , 𝑍 ≡ 𝐺, 𝐻, 𝑆). Numbers in bracket [] in the QM/MM-
FEP row were calculated using the experimental gas-phase enthalpy and entropy of phenol adsorption on the Pt(111) catalyst surface
𝑔𝑎𝑠 𝑔𝑎𝑠
(∆𝐻𝑃ℎ = −200 𝑘𝐽⁄𝑚𝑜𝑙 77, ∆𝑆𝑃ℎ = −14.7𝑅 78).

𝒈−𝒍
𝒈𝒂𝒔→𝒍𝒊𝒒 𝒈𝒂𝒔→𝒍𝒊𝒒 𝒈𝒂𝒔→𝒍𝒊𝒒 𝒈−𝒍 ∆𝑺𝒂𝒅𝒔,𝑷𝒉 (meV/K
𝑻(K) ∆∆𝑮𝑷𝒉 (eV) ∆∆𝑯𝑷𝒉 (eV) ∆∆𝑺𝑷𝒉 (meV/K) ∆𝑯𝒂𝒅𝒔,𝑷𝒉 (eV)
)
283 1.46
288 1.44
Experimental 2.33 3.10 0.26 1.80
108

298 1.42
314 1.36
285 1.45±0.02
298 1.48±0.04
-1.60±0.12 -4.90±0.3
QM/MM-FEP 315 1.50±0.00 0.71±0.12 -2.60±0.3
[-1.36±0.12] [-3.90±0.3]
335 1.56±0.02
353 1.63±0.01
I Ia II III IIIa IV V

FEP FEP FEP FEP

+ + + + + + +

Figure 4.1: Free energy perturbation approach for computing the difference in the
adsorption free energy of any adsorbed species, , in the presence of liquid water ( (𝑔) +
∗ (𝑙) ↔ ∗ (𝑙)) and in a gas phase ( (𝑔) + ∗ (𝑔) ↔ ∗ (𝑔)). States Ia and IIIa are
“non-physical” states. Between states I and II, the (classical) electrostatic interaction
between the solvent and the adsorbate is slowly removed. Also, the coordinates of the
atoms described quantum chemically transition from those in liquid to those in the vapor
phase. From state II to III, the non-classical contributions of water molecules changing the
electronic structure of the quantum system transition from those of a metal-adsorbate
system to those without adsorbate, i.e., this energy difference constitutes the main
difference between electrostatic and mechanical embedding of the QM/MM system. In
other words, symbolizes the adsorbate being only present in the classical DL_POLY
simulation and not in the quantum system. Van der Waals interactions between the solvent
and the adsorbate are slowly removed between state III and IV. The gas-phase desorption
is described between state IV and V. The (free) energy difference between state IV and I
is the solvent effect on the stability of the adsorbate.

109
Figure 4.2: (Average) free-energy profiles for aqueous-phase effects on the desorption
𝑙𝑖𝑞→𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞
(∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = −∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) of a) the reactant and transition states of O-H bond
cleavage of ethylene glycol over a Cu(111) surface at 423 K, b) the CO molecule over the
Cu(111) and Pt(111) surfaces at 298 K, c) the ethylene glycol molecule over Cu(111) and
Pt(111) surfaces at 423 K, and d) the phenol molecule over the Cu(111) and Pt(111)
𝑙𝑖𝑞→𝑔𝑎𝑠
surfaces at 298 K. ∆∆𝐺 (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) is the aqueous-phase effect on the low coverage
desorption of an adsorbate. “To OPT adsorbate in Liquid” represents the FEP steps
involved in the transition of the metal-adsorbate QM system in gas to that in liquid, “To
OPT site in Gas” represents the FEP steps involved in changing the coordinates of the metal
cluster atoms with dummy adsorbates that have zero interaction with the solvent to those
of the optimized metal cluster in the gas phase, and “II-to-III” shows the non-classical
contribution of water molecules changing the electronic structure of the quantum system
transitioning from those of a metal-adsorbate system to those without adsorbate (see Figure
4.1, states II and III). “Remove Electrostatic” and “Remove Lennard-Jones” represent the
removal of classical electrostatic and Lennard-Jones interactions between an adsorbate and
the water molecules, respectively. For the magnitude of each contribution, see Tables 4.2
and C.5.

110
Figure 4.3: Fits of equation 4.9 (adsorption isotherm) to the experimental data of the
aqueous-phase phenol adsorption on a Pt(111) catalyst surface at 298 K from Singh et al.26,
74
(see Figures C.7 to C.9 for similar plots at different temperatures). The “Const”
parameter represents the hydrogen surface coverage relative to the free site coverage
(const=0 shows zero H coverage approximation), 𝑛 represents the number of sites a phenol
𝑙𝑖𝑞
molecule occupies when adsorbed on the surface, 𝛼 (𝛼𝑃ℎ,𝑃ℎ ) is a constant associated to the
𝑔𝑎𝑠→𝑙𝑖𝑞
phenol-phenol lateral interactions on the surface in water, and ∆∆𝐺 (∆∆𝐺𝑃ℎ (𝜃𝑃ℎ =
0)) is the aqueous-phase effect on the low coverage phenol adsorption on the surface. SSR
is the sum of squared residuals, which shows the error associated with each fit.

111
Figure 4.4: Temperature dependence of the aqueous-phase effect on the free energy of low
coverage adsorption of phenol and carbon monoxide over a Pt(111) surface (∆∆𝐺 =
𝑔𝑎𝑠→𝑙𝑖𝑞
∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 (𝜃𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = 0)). The error bars indicate 95% confidence intervals based
on limited water sampling and multiple (three) independent simulations.

112
4.8 Bibliography

1. Králik M. Adsorption, chemisorption, and catalysis. Chemical Papers 68, 1625-


1638 (2014).

2. Viñes F, Vojvodic A, Abild-Pedersen F, Illas F. Brønsted–Evans–Polanyi


Relationship for Transition Metal Carbide and Transition Metal Oxide Surfaces.
The Journal of Physical Chemistry C 117, 4168-4171 (2013).

3. Bligaard T, Nørskov JK, Dahl S, Matthiesen J, Christensen CH, Sehested J. The


Brønsted–Evans–Polanyi relation and the volcano curve in heterogeneous
catalysis. Journal of Catalysis 224, 206-217 (2004).

4. Bronsted JN. Acid and Basic Catalysis. Chemical Reviews 5, 231-338 (1928).

5. Dumesic JA, Huber GW, Boudart M. Principles of Heterogeneous Catalysis. In:


Handbook of Heterogeneous Catalysis).

6. Wisniak J. The History of Catalysis. From the Beginning to Nobel Prizes.


Educación Química 21, 60-69 (2010).

7. Satterfield CN. Heterogeneous catalysis in practice. McGraw-Hill (1980).

8. Campbell CT. Energies of Adsorbed Catalytic Intermediates on Transition Metal


Surfaces: Calorimetric Measurements and Benchmarks for Theory. Accounts of
Chemical Research 52, 984-993 (2019).

9. Silbaugh TL, Campbell CT. Energies of Formation Reactions Measured for


Adsorbates on Late Transition Metal Surfaces. The Journal of Physical Chemistry
C 120, 25161-25172 (2016).

10. Lytken O, Lew W, Campbell CT. Catalytic reaction energetics by single crystal
adsorption calorimetry: hydrocarbons on Pt(111). Chemical Society Reviews 37,
2172-2179 (2008).

11. Saleheen M, Heyden A. Liquid-Phase Modeling in Heterogeneous Catalysis. ACS


Catalysis 8, 2188-2194 (2018).

113
12. Varghese JJ, Mushrif SH. Origins of complex solvent effects on chemical
reactivity and computational tools to investigate them: a review. Reaction
Chemistry & Engineering 4, 165-206 (2019).

13. Wilke TJ, Barteau MA. Development of thermodynamic activity coefficients to


describe the catalytic performance of supported polyoxometalate catalysts.
Journal of Catalysis 382, 286-294 (2020).

14. Deshpande S, Greeley J. First-Principles Analysis of Coverage, Ensemble, and


Solvation Effects on Selectivity Trends in NO Electroreduction on Pt3Sn Alloys.
ACS Catalysis 10, 9320-9327 (2020).

15. Bertero NM, Trasarti AF, Acevedo MC, Marchi AJ, Apesteguia CR. Solvent
effects in solid acid-catalyzed reactions: The case of the liquid-phase
isomerization/cyclization of citronellal over SiO2-Al2O3. Molecular Catalysis
481, (2020).

16. Faveere W, et al. Glycolaldehyde as a Bio-Based C-2 Platform Chemical:


Catalytic Reductive Amination of Vicinal Hydroxyl Aldehydes. Acs Catalysis 10,
391-404 (2020).

17. Saleheen M, Verma AM, Mamun O, Lu J, Heyden A. Investigation of solvent


effects on the hydrodeoxygenation of guaiacol over Ru catalysts. Catalysis
Science & Technology 9, 6253-6273 (2019).

18. Wan H, Vitter A, Chaudhari RV, Subramaniam B. Kinetic investigations of


unusual solvent effects during Ru/C catalyzed hydrogenation of model
oxygenates. Journal of Catalysis 309, 174-184 (2014).

19. Mellmer MA, Sener C, Gallo JMR, Luterbacher JS, Alonso DM, Dumesic JA.
Solvent Effects in Acid-Catalyzed Biomass Conversion Reactions. Angewandte
Chemie International Edition 53, 11872-11875 (2014).

20. Zare M, Solomon RV, Yang W, Yonge A, Heyden A. Theoretical Investigation of


Solvent Effects on the Hydrodeoxygenation of Propionic Acid over a Ni(111)
Catalyst Model. The Journal of Physical Chemistry C 124, 16488-16500 (2020).

21. Staszak-Jirkovsky J, et al. Water as a Promoter and Catalyst for Dioxygen


Electrochemistry in Aqueous and Organic Media. Acs Catalysis 5, 6600-6607
(2015).

114
22. Zhao JJ, et al. Suppressing Metal Leaching in a Supported Co/SiO2 Catalyst with
Effective Protectants in the Hydroformylation Reaction. Acs Catalysis 10, 914-
920 (2020).

23. Kulal AB, Kasabe MM, Jadhav PV, Dongare MK, Umbarkar SB. Hydrophobic
WO3/SiO2 catalyst for the nitration of aromatics in liquid phase. Applied
Catalysis a-General 574, 105-113 (2019).

24. He R, Davda RR, Dumesic JA. In situ ATR-IR spectroscopic and reaction
kinetics studies of water-gas shift and methanol reforming on Pt/Al2O3 catalysts
in vapor and liquid phases. J Phys Chem B 109, 2810-2820 (2005).

25. Singh N, Campbell CT. A Simple Bond-Additivity Model Explains Large


Decreases in Heats of Adsorption in Solvents Versus Gas Phase: A Case Study
with Phenol on Pt(111) in Water. ACS Catalysis 9, 8116-8127 (2019).

26. Akinola J, Barth I, Goldsmith BR, Singh N. Adsorption Energies of Oxygenated


Aromatics and Organics on Rhodium and Platinum in Aqueous Phase. ACS
Catalysis 10, 4929-4941 (2020).

27. Akinola J, Singh N. Temperature dependence of aqueous-phase phenol adsorption


on Pt and Rh. Journal of Applied Electrochemistry 51, 37-50 (2021).

28. Zhang L, Karim AM, Engelhard MH, Wei ZH, King DL, Wang Y. Correlation of
Pt-Re surface properties with reaction pathways for the aqueous-phase reforming
of glycerol. Journal of Catalysis 287, 37-43 (2012).

29. Tupy SA, et al. Correlating Ethylene Glycol Reforming Activity with In Situ
EXAFS Detection of Ni Segregation in Supported NiPt Bimetallic Catalysts. Acs
Catalysis 2, 2290-2296 (2012).

30. Karim AM, et al. In Situ X-ray Absorption Fine Structure Studies on the Effect of
pH on Pt Electronic Density during Aqueous Phase Reforming of Glycerol. Acs
Catalysis 2, 2387-2394 (2012).

31. Zope BN, Hibbitts DD, Neurock M, Davis RJ. Reactivity of the Gold/Water
Interface During Selective Oxidation Catalysis. Science 330, 74-78 (2010).

32. Taylor CD, Neurock M. Theoretical insights into the structure and reactivity of
the aqueous/metal interface. Curr Opin Solid St M 9, 49-65 (2005).

115
33. Yoon Y, Rousseau R, Weber RS, Mei DH, Lercher JA. First-Principles Study of
Phenol Hydrogenation on Pt and Ni Catalysts in Aqueous Phase. J Am Chem Soc
136, 10287-10298 (2014).

34. Faheem M, Saleheen M, Lu JM, Heyden A. Ethylene glycol reforming on


Pt(111): first-principles microkinetic modeling in vapor and aqueous phases.
Catal Sci Technol 6, 8242-8256 (2016).

35. Zhang X, Savara A, Getman RB. A Method for Obtaining Liquid–Solid


Adsorption Rates from Molecular Dynamics Simulations: Applied to Methanol on
Pt(111) in H2O. Journal of Chemical Theory and Computation 16, 2680-2691
(2020).

36. Clabaut P, Schweitzer B, Götz AW, Michel C, Steinmann SN. Solvation Free
Energies and Adsorption Energies at the Metal/Water Interface from Hybrid
Quantum-Mechanical/Molecular Mechanics Simulations. Journal of Chemical
Theory and Computation 16, 6539-6549 (2020).

37. Steinmann SN, Sautet P, Michel C. Solvation free energies for periodic surfaces:
comparison of implicit and explicit solvation models. Physical Chemistry
Chemical Physics 18, 31850-31861 (2016).

38. Zhang X, DeFever RS, Sarupria S, Getman RB. Free Energies of Catalytic
Species Adsorbed to Pt(111) Surfaces under Liquid Solvent Calculated Using
Classical and Quantum Approaches. Journal of Chemical Information and
Modeling 59, 2190-2198 (2019).

39. Ward BM, Getman RB. Molecular simulations of physical and chemical
adsorption under gas and liquid environments using force field- and quantum
mechanics-based methods. Molecular Simulation 40, 678-689 (2014).

40. Gould NS, et al. Understanding solvent effects on adsorption and protonation in
porous catalysts. Nature Communications 11, 1060 (2020).

41. Mushrif SH, Caratzoulas S, Vlachos DG. Understanding solvent effects in the
selective conversion of fructose to 5-hydroxymethyl-furfural: a molecular
dynamics investigation. Physical Chemistry Chemical Physics 14, 2637-2644
(2012).

116
42. Bregante DT, et al. Cooperative Effects between Hydrophilic Pores and Solvents:
Catalytic Consequences of Hydrogen Bonding on Alkene Epoxidation in Zeolites.
J Am Chem Soc 141, 7302-7319 (2019).

43. Kandoi S, Greeley J, Simonetti D, Shabaker J, Dumesic JA, Mavrikakis M.


Reaction Kinetics of Ethylene Glycol Reforming over Platinum in the Vapor
versus Aqueous Phases. The Journal of Physical Chemistry C 115, 961-971
(2011).

44. Car R, Parrinello M. Unified Approach for Molecular-Dynamics and Density-


Functional Theory. Phys Rev Lett 55, 2471-2474 (1985).

45. Carloni P, Rothlisberger U, Parrinello M. The role and perspective of a initio


molecular dynamics in the study of biological systems. Acc Chem Res 35, 455-
464 (2002).

46. Iftimie R, Minary P, Tuckerman ME. Ab initio molecular dynamics: Concepts,


recent developments, and future trends. P Natl Acad Sci USA 102, 6654-6659
(2005).

47. Mattsson TR, Paddison SJ. Methanol at the water-platinum interface studied by
ab initio molecular dynamics. Surf Sci 544, L697-L702 (2003).

48. Otani M, Hamada I, Sugino O, Morikawa Y, Okamoto Y, Ikeshoji T. Structure of


the water/platinum interface - a first principles simulation under bias potential.
Physical Chemistry Chemical Physics 10, 3609-3612 (2008).

49. Yang J, Dauenhauer PJ, Ramasubramaniam A. The role of water in the adsorption
of oxygenated aromatics on Pt and Pd. J Comput Chem 34, 60-66 (2013).

50. Klamt A. Conductor-Like Screening Model for Real Solvents - a New Approach
to the Quantitative Calculation of Solvation Phenomena. J Phys Chem-Us 99,
2224-2235 (1995).

51. Klamt A, Schuurmann G. Cosmo - a New Approach to Dielectric Screening in


Solvents with Explicit Expressions for the Screening Energy and Its Gradient. J
Chem Soc Perk T 2, 799-805 (1993).

52. Dyson PJ, Jessop PG. Solvent effects in catalysis: rational improvements of
catalysts via manipulation of solvent interactions. Catalysis Science &
Technology 6, 3302-3316 (2016).

117
53. Valgimigli L, Banks JT, Ingold KU, Lusztyk J. Kinetic Solvent Effects on
Hydroxylic Hydrogen Atom Abstractions Are Independent of the Nature of the
Abstracting Radical. Two Extreme Tests Using Vitamin E and Phenol. J Am
Chem Soc 117, 9966-9971 (1995).

54. Zhang YK, Liu HY, Yang WT. Free energy calculation on enzyme reactions with
an efficient iterative procedure to determine minimum energy paths on a
combined ab initio QM/MM potential energy surface. J Chem Phys 112, 3483-
3492 (2000).

55. Hu H, Lu ZY, Parks JM, Burger SK, Yang WT. Quantum mechanics/molecular
mechanics minimum free-energy path for accurate reaction energetics in solution
and enzymes: Sequential sampling and optimization on the potential of mean
force surface. J Chem Phys 128, 034105:034101-034118 (2008).

56. Hu H, Lu ZY, Yang WT. QM/MM minimum free-energy path: Methodology and
application to triosephosphate isomerase. J Chem Theory Comput 3, 390-406
(2007).

57. Faheem M, Heyden A. Hybrid Quantum Mechanics/Molecular Mechanics


Solvation Scheme for Computing Free Energies of Reactions at Metal–Water
Interfaces. Journal of Chemical Theory and Computation 10, 3354-3368 (2014).

58. Saleheen M, Zare M, Faheem M, Heyden A. Computational Investigation of


Aqueous Phase Effects on the Dehydrogenation and Dehydroxylation of Polyols
over Pt(111). The Journal of Physical Chemistry C 123, 19052-19065 (2019).

59. Zare M, Saleheen M, Kundu SK, Heyden A. Dependency of solvation effects on


metal identity in surface reactions. Communications Chemistry 3, 187 (2020).

60. Sprowl LH, Campbell CT, Árnadóttir L. Hindered Translator and Hindered Rotor
Models for Adsorbates: Partition Functions and Entropies. The Journal of
Physical Chemistry C 120, 9719-9731 (2016).

61. Bajpai A, Mehta P, Frey K, Lehmer AM, Schneider WF. Benchmark First-
Principles Calculations of Adsorbate Free Energies. ACS Catalysis 8, 1945-1954
(2018).

62. Reed AE, Weinstock RB, Weinhold F. Natural-Population Analysis. J Chem Phys
83, 735-746 (1985).

118
63. Manz TA, Limas NG. Introducing DDEC6 atomic population analysis: part 1.
Charge partitioning theory and methodology. RSC Advances 6, 47771-47801
(2016).

64. Faheem M, Heyden A. Hybrid Quantum Mechanics/Molecular Mechanics


Solvation Scheme for Computing Free Energies of Reactions at Metal-Water
Interfaces. Journal of Chemical Theory and Computation 10, 3354-3368 (2014).

65. Kresse G, Furthmuller J. Efficiency of ab-initio total energy calculations for


metals and semiconductors using a plane-wave basis set. Comp Mater Sci 6, 15-
50 (1996).

66. Kresse G, Furthmuller J. Efficient iterative schemes for ab initio total-energy


calculations using a plane-wave basis set. Phys Rev B 54, 11169-11186 (1996).

67. Ahlrichs R, Bar M, Haser M, Horn H, Kolmel C. Electronic-Structure


Calculations on Workstation Computers - the Program System Turbomole. Chem
Phys Lett 162, 165-169 (1989).

68. Treutler O, Ahlrichs R. Efficient Molecular Numerical-Integration Schemes. J


Chem Phys 102, 346-354 (1995).

69. Von Arnim M, Ahlrichs R. Performance of parallel TURBOMOLE for density


functional calculations. J Comput Chem 19, 1746-1757 (1998).

70. Burow AM, Sierka M, Dobler J, Sauer J. Point defects in CaF2 and CeO2
investigated by the periodic electrostatic embedded cluster method. J Chem Phys
130, 174710:174711-174711 (2009).

71. Zacharias M, Straatsma TP, McCammon JA. Separation‐shifted scaling, a new


scaling method for Lennard‐Jones interactions in thermodynamic integration. The
Journal of Chemical Physics 100, 9025-9031 (1994).

72. Bennett C. Efficient Estimation of Free Energy Differences from Monte Carlo
Data. J Comput Phys 22, 245-268 (1976).

73. Hensley AJR, et al. DFT-Based Method for More Accurate Adsorption Energies:
An Adaptive Sum of Energies from RPBE and vdW Density Functionals. The
Journal of Physical Chemistry C 121, 4937-4945 (2017).

119
74. Singh N, Sanyal U, Fulton JL, Gutiérrez OY, Lercher JA, Campbell CT.
Quantifying Adsorption of Organic Molecules on Platinum in Aqueous Phase by
Hydrogen Site Blocking and in Situ X-ray Absorption Spectroscopy. ACS
Catalysis 9, 6869-6881 (2019).

75. Nitta T, Shigetomi T, Kuro-Oka M, Katayama T. AN ADSORPTION


ISOTHERM OF MULTI-SITE OCCUPANCY MODEL FOR HOMOGENEOUS
SURFACE. Journal of Chemical Engineering of Japan 17, 39-45 (1984).

76. Nitta T, Kuro-Oka M, Katayama T. AN ADSORPTION ISOTHERM OF


MULTI-SITE OCCUPANCY MODEL FOR HETEROGENEOUS SURFACE.
Journal of Chemical Engineering of Japan 17, 45-52 (1984).

77. Carey SJ, Zhao W, Mao Z, Campbell CT. Energetics of Adsorbed Phenol on
Ni(111) and Pt(111) by Calorimetry. The Journal of Physical Chemistry C 123,
7627-7632 (2019).

78. Kudchadker SA, Kudchadker AP, Wilhoit RC, Zwolinski BJ. Ideal gas
thermodynamic properties of phenol and cresols. J Phys Chem Ref Data 7, 417-
423 (1978).

79. Sander R. Compilation of Henry's law constants (version 4.0) for water as solvent.
Atmospheric Chemistry and Physics 15, 4399-4981 (2015).

80. Faheem M, Suthirakun S, Heyden A. New Implicit Solvation Scheme for Solid
Surfaces. J Phys Chem C 116, 22458-22462 (2012).

81. Fishman M, Zhuang HLL, Mathew K, Dirschka W, Hennig RG. Accuracy of


exchange-correlation functionals and effect of solvation on the surface energy of
copper. Phys Rev B 87, (2013).

82. Mathew K, Sundararaman R, Letchworth-Weaver K, Arias TA, Hennig RG.


Implicit solvation model for density-functional study of nanocrystal surfaces and
reaction pathways. J Chem Phys 140, 084106:084101-084108 (2014).

83. Heinz H, Vaia RA, Farmer BL, Naik RR. Accurate Simulation of Surfaces and
Interfaces of Face-Centered Cubic Metals Using 12-6 and 9-6 Lennard-Jones
Potentials. J Phys Chem C 112, 17281-17290 (2008).

120
APPENDIX A

SUPPORTING INFORMATION FOR THEORETICAL INVESTIGATION

OF SOLVENT EFFECTS ON THE HYDRODEOXYGENATION OF

PROPIONIC ACID OVER A NI(111) CATALYST MODEL

Zare, M.; Solomon R. V.; Yang, W.; Yonge, A.; Heyden, A. The Journal of Physical
Chemistry C, 2020, 124, 16488-16500.

Reprinted here with the permission of the publisher

121
Table A.1: Solvent effect on the stability of various adsorbed species in the propionic acid
conversion to ethane and ethylene over a Ni (111) catalyst surface model at two different
temperatures of 298 K and 473 K based on two implicit solvation schemes: iSMS and
VASPsol. ΔΔGrxn indicates the difference in the adsorption free energy of the
corresponding intermediates in the presence and absence of water. Asterisk (*) represents
a surface adsorption site and multiple asterisks are indicative of the number of occupied
active sites.

ΔΔGrxn, eV

Adsorbed species T=298 K T=473 K


iSMS VASPsol iSMS VASPsol
CH2C*** -0.04 0.02 -0.08 0.05
CH2CH*** 0.02 -0.08 -0.01 -0.07
CH2CH2** 0.06 -0.08 0.10 -0.08
CH2CHCO**** -0.05 -0.07 -0.09 -0.05
CH2CHCOOH**** -0.10 -0.34 -0.11 -0.29
CH3C*** 0.02 -0.04 -0.01 -0.03
CH3CCO**** -0.09 -0.07 -0.12 -0.06
CH3CCOO*** -0.63 -0.29 -0.55 -0.26
CH3CCOOH*** -0.12 -0.28 -0.12 -0.26
CH3CH*** 0.03 -0.05 0.00 -0.03
CH3CH2** 0.03 -0.06 0.01 -0.05
CH3CH2CO*** -0.02 -0.11 -0.05 -0.12
CH3CH2COO** -0.05 -0.11 -0.06 -0.10
CH3CH2COOH* -0.07 -0.44 -0.03 -0.40
CH3CH3* 0.05 -0.04 0.02 -0.06
CH3CHCO** -0.06 -0.11 -0.09 -0.11
CH3CHCOO*** -0.34 -0.12 -0.31 -0.10
CH3CHCOOH** -0.10 -0.30 -0.10 -0.28
CHCH**** 0.02 -0.06 0.05 -0.05
CHCHCO**** -0.11 -0.12 -0.15 -0.10
CHCHCOOH**** -0.13 -0.31 -0.13 -0.26
CO*** -0.15 -0.08 -0.25 -0.08
CO2* 0.05 -0.10 -0.04 -0.07
COOH** -0.18 -0.25 -0.16 -0.21
H* -0.02 0.00 -0.01 0.00
H2O* -0.04 -0.31 0.02 -0.25
OH* -0.02 -0.21 0.00 -0.18

122
Table A.2: Solvent effect on the stability of transition states in the propionic acid conversion to ethane and ethylene over a Ni (111)
catalyst surface model at two different temperatures of 298 K and 473 K based on two implicit solvation schemes: iSMS and VASPsol.
Solvent effect on TS indicates the difference in the adsorption free energy of the corresponding transition state in the presence and
absence of water.

Solvent effect on TS, eV


# Reaction T=298 K T=473 K
iSMS VASPsol iSMS VASPsol
1 CH3CH2COOH* + 3* → CH3CH2CO*** + OH* -0.03 -0.16 -0.05 -0.11
2 CH3CH2COOH* + 2* → CH3CHCOOH** + H* -0.10 -0.39 -0.09 -0.37
3 CH3CH2CO***+ 2* → CH3CH2** + CO*** -0.04 -0.10 -0.07 -0.10
4 CH3CH2CO*** → CH3CHCO** + H* -0.05 -0.10 -0.08 -0.07
5 CH3CHCOOH** + * → CH3CHCO** + OH* -0.12 -0.17 -0.14 -0.15
6 CH3CHCOOH** + 3* → CH2CHCOOH**** + H* -0.11 -0.36 -0.10 -0.32
7 CH3CHCOOH** + 2*→CH3CCOOH*** + H* -0.10 -0.30 -0.10 -0.25
CH3CHCO** + 4* → CH3CH*** + CO***
123

8 -0.06 -0.11 -0.10 -0.11


9 CH3CHCO** + 3* → CH3CCO**** + H* -0.13 -0.07 -0.15 -0.04
10 CH3CHCO** + 3* → CH2CHCO**** + H* -0.05 -0.13 -0.09 -0.10
11 CH2CHCOOH**** + * → CH2CHCO**** + OH* -0.08 -0.19 -0.11 -0.16
12 CH2CHCOOH**** + * → CHCHCOOH**** + H* -0.10 -0.34 -0.10 -0.30
13 CH3CCOOH*** + 2* → CH3CCO**** + OH* -0.18 -0.19 -0.20 -0.17
14 CH3CCO**** + 2* → CH3C*** +CO*** -0.17 -0.09 -0.20 -0.07
15 CH2CHCO**** + 2* → CH2CH*** + CO*** -0.05 -0.12 -0.09 -0.10
16 CH2CHCO**** + * → CHCHCO**** + H* -0.05 -0.11 -0.09 -0.10
17 CHCHCOOH**** + * → CHCHCO**** + OH* -0.24 -0.17 -0.26 -0.14
18 CHCHCO**** + 3* → CHCH**** + CO*** -0.13 -0.12 -0.17 -0.11
19 CH2CH*** + 2* → CHCH**** + H* 0.05 -0.08 0.01 -0.07
20 CH2CH2** + 2* → CH2CH*** + H* 0.07 -0.09 0.04 -0.09
21 CH2CH*** + * → CH2C*** + H* 0.01 -0.05 -0.03 -0.05
22 CH3C*** + * → CH2C*** + H* 0.01 -0.08 -0.02 -0.07
23 CH3CH*** + * → CH2CH*** + H* 0.00 -0.10 -0.03 -0.09
24 CH3CH*** + * → CH3C*** + H* 0.04 -0.05 0.01 -0.04
25 CH3CH2** + 2* → CH3CH*** + H* 0.05 -0.05 0.02 -0.02
26 CH3CH3* + 2* → CH3CH2** + H* 0.04 -0.08 0.02 -0.08
27 CH3CH2** + * → CH2CH2** + H* 0.03 -0.13 0.00 -0.11
28 CH3CH2COOH* + 2* → CH3CH2COO** + H* -0.11 -0.15 -0.12 -0.16
29 CH3CH2COO** + * → CH3CH2** + CO2* -0.06 -0.12 -0.10 -0.09
30 CH3CH2COO** + 2* → CH3CHCOO*** + H* -0.24 -0.12 -0.23 -0.11
31 CH3CHCOOH** + 2* → CH3CHCOO*** + H* -0.24 -0.10 -0.25 -0.09
32 CH3CHCOOH** + 3* → CH3CH*** + COOH** -0.12 -0.29 -0.12 -0.23
33 CH3CHCOO*** + * → CH3CH*** + CO2* -0.18 -0.12 -0.19 -0.10
34 CH3CHCOO*** + * → CH3CCOO*** + H* -0.20 -0.09 -0.20 -0.08
35 CH3CCOOH*** + * → CH3CCOO*** + H* -0.32 -0.16 -0.30 -0.15
36 CH3CCOOH*** + 2* → CH3C*** + COOH** -0.16 -0.30 -0.17 -0.25
37 CH2CHCOOH**** + * → CH2CH*** + COOH** -0.11 -0.32 -0.11 -0.26
38 CH3CCOO*** + * → CH3C*** + CO2* -0.34 -0.16 -0.35 -0.15
124

39 COOH** → CO2* + H* -0.34 -0.17 -0.34 -0.15


40 COOH** + 2* → CO*** + OH* -0.12 -0.25 -0.12 -0.23
41 H2O* + * → OH* + H* -0.04 -0.53 -0.03 -0.58
Table A.3: Solvent effect on the reaction free energies of all elementary surface reactions of the propionic acid conversion to ethane and
ethylene over a Ni (111) catalyst surface model at two different temperature of 298 K and 473 K based on two different implicit solvation
schemes: iSMS and VASPsol. ΔΔGrxn indicates the reaction free energy difference between corresponding reaction in liquid water and
in a gas phase.

ΔΔGrxn
# Reaction T=298 K T=473 K
iSMS VASPsol iSMS VASPsol
1 CH3CH2COOH* + 3* → CH3CH2CO*** + OH* 0.03 0.13 0.00 0.10
2 CH3CH2COOH* + 2* → CH3CHCOOH** + H* -0.04 0.14 -0.07 0.12
3 CH3CH2CO***+ 2* → CH3CH2** + CO*** -0.10 -0.04 -0.14 -0.01
4 CH3CH2CO*** → CH3CHCO** + H* -0.05 0.00 -0.06 0.01
5 CH3CHCOOH** + * → CH3CHCO** + OH* 0.02 -0.01 0.01 -0.01
6 CH3CHCOOH** + 3* → CH2CHCOOH**** + H* -0.02 -0.03 -0.02 -0.01
7 CH3CHCOOH** + 2*→CH3CCOOH*** + H* -0.03 0.02 -0.03 0.02
CH3CHCO** + 4* → CH3CH*** + CO***
125

8 -0.07 -0.03 -0.10 0.00


9 CH3CHCO** + 3* → CH3CCO**** + H* -0.05 0.04 -0.05 0.06
10 CH3CHCO** + 3* → CH2CHCO**** + H* -0.01 0.04 -0.01 0.06
11 CH2CHCOOH**** + * → CH2CHCO**** + OH* 0.03 0.06 0.02 0.06
12 CH2CHCOOH**** + * → CHCHCOOH**** + H* -0.05 0.04 -0.05 0.03
13 CH3CCOOH*** + 2* → CH3CCO**** + OH* 0.01 0.01 -0.01 0.02
14 CH3CCO**** + 2* → CH3C*** +CO*** -0.04 -0.06 -0.08 -0.05
15 CH2CHCO**** + 2* → CH2CH*** + CO*** -0.07 -0.09 -0.11 -0.10
16 CH2CHCO**** + * → CHCHCO**** + H* -0.08 -0.04 -0.08 -0.05
17 CHCHCOOH**** + * → CHCHCO**** + OH* 0.00 -0.02 -0.01 -0.03
18 CHCHCO**** + 3* → CHCH**** + CO*** -0.02 -0.02 -0.06 -0.03
19 CH2CH*** + 2* → CHCH**** + H* -0.02 0.02 -0.03 0.02
20 CH2CH2** + 2* → CH2CH*** + H* -0.05 0.00 -0.06 0.01
21 CH2CH*** + * → CH2C*** + H* -0.08 0.10 -0.09 0.13
22 CH3C*** + * → CH2C*** + H* -0.07 0.06 -0.08 0.08
23 CH3CH*** + * → CH2CH*** + H* -0.02 -0.02 -0.03 -0.04
24 CH3CH*** + * → CH3C*** + H* -0.02 0.01 -0.03 0.01
25 CH3CH2** + 2* → CH3CH*** + H* -0.02 0.02 -0.03 0.02
26 CH3CH3* + 2* → CH3CH2** + H* -0.04 -0.02 -0.05 0.02
27 CH3CH2** + * → CH2CH2** + H* 0.01 -0.01 0.01 -0.02
28 CH3CH2COOH* + 2* → CH3CH2COO** + H* 0.01 0.33 -0.03 0.30
29 CH3CH2COO** + * → CH3CH2** + CO2* 0.13 -0.05 0.06 -0.02
30 CH3CH2COO** + 2* → CH3CHCOO*** + H* -0.31 -0.01 -0.27 0.00
31 CH3CHCOOH** + 2* → CH3CHCOO*** + H* -0.26 0.19 -0.23 0.18
32 CH3CHCOOH** + 3* → CH3CH*** + COOH** -0.05 0.01 -0.06 0.04
33 CH3CHCOO*** + * → CH3CH*** + CO2* 0.42 -0.03 0.31 0.00
34 CH3CHCOO*** + * → CH3CCOO*** + H* -0.30 -0.18 -0.26 -0.14
35 CH3CCOOH*** + * → CH3CCOO*** + H* -0.52 0.00 -0.45 0.00
36 CH3CCOOH*** + 2* → CH3C*** + COOH** -0.04 -0.01 -0.06 0.02
37 CH2CHCOOH**** + * → CH2CH*** + COOH** -0.05 0.02 -0.07 0.01
38 CH3CCOO*** + * → CH3C*** + CO2* 0.70 0.15 0.54 0.16
126

39 COOH** → CO2* + H* 0.22 0.15 0.15 0.14


40 COOH** + 2* → CO*** + OH* 0.01 -0.04 -0.03 -0.06
41 H2O* + * → OH* + H* 0.00 0.11 0.00 0.07
Table A.4: Solvent effect on the activation free energies of all elementary surface reactions of the propionic acid conversion to ethane
and ethylene over a Ni (111) catalyst surface model at two different temperature of 298 K and 473 K based on two different implicit
solvation schemes: iSMS and VASPsol. ΔΔGact indicates the activation free energy differences between corresponding reaction in liquid
water and in the gas phase.

ΔΔGact
# Reaction T=298 K T=473 K
iSMS VASPsol iSMS VASPsol
1 CH3CH2COOH* + 3* → CH3CH2CO*** + OH* 0.04 0.29 0.00 0.29
2 CH3CH2COOH* + 2* → CH3CHCOOH** + H* -0.03 0.05 -0.03 0.03
3 CH3CH2CO***+ 2* → CH3CH2** + CO*** -0.02 0.01 -0.03 0.02
4 CH3CH2CO*** → CH3CHCO** + H* -0.03 0.01 -0.03 0.05
5 CH3CHCOOH** + * → CH3CHCO** + OH* -0.02 0.13 -0.03 0.13
6 CH3CHCOOH** + 3* → CH2CHCOOH**** + H* -0.01 -0.06 0.00 -0.04
7 CH3CHCOOH** + 2*→CH3CCOOH*** + H* 0.00 0.00 0.00 0.02
CH3CHCO** + 4* → CH3CH*** + CO***
127

8 -0.01 0.00 -0.01 0.00


9 CH3CHCO** + 3* → CH3CCO**** + H* -0.08 0.04 -0.06 0.07
10 CH3CHCO** + 3* → CH2CHCO**** + H* 0.01 -0.02 0.00 0.01
11 CH2CHCOOH**** + * → CH2CHCO**** + OH* 0.03 0.15 0.00 0.13
12 CH2CHCOOH**** + * → CHCHCOOH**** + H* 0.01 0.00 0.01 -0.01
13 CH3CCOOH*** + 2* → CH3CCO**** + OH* -0.06 0.10 -0.08 0.09
14 CH3CCO**** + 2* → CH3C*** +CO*** -0.08 0.21 -0.08 0.18
15 CH2CHCO**** + 2* → CH2CH*** + CO*** 0.00 -0.04 -0.01 -0.05
16 CH2CHCO**** + * → CHCHCO**** + H* 0.00 -0.04 0.00 -0.05
17 CHCHCOOH**** + * → CHCHCO**** + OH* -0.11 0.14 -0.12 0.12
18 CHCHCO**** + 3* → CHCH**** + CO*** -0.02 0.00 -0.03 0.00
19 CH2CH*** + 2* → CHCH**** + H* 0.02 -0.01 0.02 0.00
20 CH2CH2** + 2* → CH2CH*** + H* 0.01 -0.02 0.01 -0.01
21 CH2CH*** + * → CH2C*** + H* -0.02 0.03 -0.02 0.02
22 CH3C*** + * → CH2C*** + H* -0.01 -0.04 -0.01 -0.04
23 CH3CH*** + * → CH2CH*** + H* -0.02 -0.05 -0.03 -0.06
24 CH3CH*** + * → CH3C*** + H* 0.02 0.00 0.02 -0.01
25 CH3CH2** + 2* → CH3CH*** + H* 0.02 0.02 0.01 0.03
26 CH3CH3* + 2* → CH3CH2** + H* -0.01 -0.04 -0.02 -0.01
27 CH3CH2** + * → CH2CH2** + H* 0.00 -0.07 -0.01 -0.06
28 CH3CH2COOH* + 2* → CH3CH2COO** + H* -0.04 0.29 -0.07 0.24
29 CH3CH2COO** + * → CH3CH2** + CO2* -0.01 -0.01 -0.04 0.01
30 CH3CH2COO** + 2* → CH3CHCOO*** + H* -0.19 -0.01 -0.17 -0.01
31 CH3CHCOOH** + 2* → CH3CHCOO*** + H* -0.14 0.21 -0.14 0.19
32 CH3CHCOOH** + 3* → CH3CH*** + COOH** -0.02 0.01 -0.02 0.05
33 CH3CHCOO*** + * → CH3CH*** + CO2* 0.16 0.01 0.12 0.00
34 CH3CHCOO*** + * → CH3CCOO*** + H* 0.14 0.03 0.12 0.03
35 CH3CCOOH*** + * → CH3CCOO*** + H* -0.20 0.12 -0.18 0.11
36 CH3CCOOH*** + 2* → CH3C*** + COOH** -0.04 -0.01 -0.05 0.01
37 CH2CHCOOH**** + * → CH2CH*** + COOH** 0.00 0.02 -0.01 0.03
38 CH3CCOO*** + * → CH3C*** + CO2* 0.28 0.13 0.21 0.11
128

39 COOH** → CO2* + H* -0.15 0.08 -0.18 0.06


40 COOH** + 2* → CO*** + OH* 0.06 -0.01 0.04 -0.02
41 H2O* + * → OH* + H* 0.00 -0.22 -0.01 -0.33
Table A.5: Overall turnover frequency of hydrodeoxygenation of propionic acid over a Ni
(111) catalyst surface model at a temperature of 473 K in vapor phase, and at different
partial pressures of hydrogen and carbon monoxide.

PCO \
PH2 0.01 0.1 1 10 30
(bar)
0.0001 3.99×10-6 1.98×10-8 8.72×10-11 7.23×10-13 7.72×10-14

0.001 2.66×10-6 3.46×10-8 1.30×10-10 7.41×10-13 7.76×10-14

0.01 6.92×10-7 1.63×10-8 1.86×10-10 8.76×10-13 8.10×10-14

0.1 8.98×10-8 5.09×10-9 9.26×10-11 1.20×10-12 1.02×10-13

1 5.72×10-9 9.47×10-10 3.30×10-11 8.55×10-13 1.24×10-13

Table A.6: Overall turnover frequency of hydrodeoxygenation of propionic acid over a Ni


(111) catalyst surface model at a temperature of 473 K in vapor phase, and at different
partial pressures of hydrogen and carbon monoxide based on considering only 1 site for
adsorbed CO and analogues change in counting sites for other adsorbates attached to the
surface through their carbon atom. The number of sites assigned to each adsorbate in this
calculation is included in Table A.7.

PCO \ PH2
0.01 0.1 1 10 30
(bar)

0.0001 1.14×10-8 1.02×10-8 6.12×10-9 1.77×10-9 6.63×10-10

0.001 9.91×10-12 9.61×10-12 8.87×10-12 6.99×10-12 5.54×10-12

0.01 8.50×10-15 7.46×10-15 7.07×10-15 6.74×10-15 6.50×10-15

0.1 2.99×10-17 1.27×10-17 7.33×10-18 5.59×10-18 5.24×10-18

1 4.20×10-19 1.35×10-19 4.49×10-20 1.64×10-20 1.09×10-20

129
Table A.7: Number of occupied sites used for microkinetic modeling result displayed in
Table A.6.

Adsorbed Species # of occupied site


CH2C 2
CH2CH 3
CH2CH 2
CH2CHCO 3
CH2CHCOOH 3
CH3C 1
CH3CCO 3
CH3CCOO 3
CH3CCOOH 3
CH3CH 2
CH3CH2 1
CH3CH2CO 3
CH3CH2COO 2
CH3CH2COOH 1
CH3CH3 1
CH3CHCO 2
CH3CHCOO 3
CH3CHCOOH 2
CHCH 3
CHCHCO 4
CHCHCOOH 3
CO 1
CO2 1
COOH 2
H 1
H2 O 1
OH 1

130
Table A.8: Solvent effect on the stability of various adsorbed species in the propionic acid
conversion to ethane and ethylene over a Ni (111) catalyst surface model at 473 K in the
presence of two different solvents, water and 1,4-dioxane, using iSMS methodology.
ΔΔGrxn indicates the difference in the adsorption free energy of the corresponding
intermediates in the presence and absence of solvent. Asterisk (*) represents a surface
adsorption site and multiple asterisks are indicative of the number of occupied active sites.
Note that calculations for solvents were performed with the help of the COSMO-RS
package with three different Ni cavity radii: with default value, with a 10% increased value
and a 10% decreased value relative to the default.

ΔΔGrxn, eV @ 473 K

Adsorbed species Water 1,4-dioxane


Default +10% -10% Default +10% -10%
CH2C*** -0.08 -0.09 -0.06 -0.05 -0.04 -0.03
CH2CH*** -0.01 -0.03 0.05 0.00 0.00 0.05
CH2CH2** 0.10 0.02 0.20 0.12 0.06 0.18
CH2CHCO**** -0.09 -0.13 -0.04 -0.03 -0.04 0.01
CH2CHCOOH**** -0.11 -0.16 -0.03 -0.04 -0.07 0.01
CH3C*** -0.01 -0.03 0.03 -0.01 0.00 0.02
CH3CCO**** -0.12 -0.15 -0.09 -0.06 -0.06 -0.04
CH3CCOO*** -0.55 -0.51 -0.58 -0.33 -0.27 -0.36
CH3CCOOH*** -0.12 -0.17 -0.08 -0.07 -0.09 -0.04
CH3CH*** 0.00 -0.02 0.07 -0.01 0.01 0.05
CH3CH2** 0.01 -0.02 0.07 -0.01 0.00 0.03
CH3CH2CO*** -0.05 -0.08 0.01 -0.04 -0.04 0.00
CH3CH2COO** -0.06 -0.09 -0.03 -0.04 -0.04 -0.04
CH3CH2COOH* -0.03 -0.11 0.06 0.02 -0.04 0.08
CH3CH3* 0.02 -0.02 0.07 0.02 -0.01 0.04
CH3CHCO** -0.09 -0.13 -0.05 -0.05 -0.06 -0.02
CH3CHCOO*** -0.31 -0.30 -0.33 -0.17 -0.13 -0.18
CH3CHCOOH** -0.10 -0.15 -0.04 -0.06 -0.07 -0.03
CHCH**** 0.05 -0.01 0.13 0.08 0.05 0.13
CHCHCO**** -0.15 -0.19 -0.14 -0.05 -0.08 -0.05
CHCHCOOH**** -0.13 -0.19 -0.09 -0.06 -0.09 -0.03
CO*** -0.25 -0.25 -0.25 -0.18 -0.16 -0.20
CO2* -0.04 -0.08 0.04 0.01 -0.02 0.04
COOH** -0.16 -0.17 -0.13 -0.08 -0.07 -0.05
H* -0.01 0.02 0.03 0.00 0.03 0.03
H2O* 0.02 -0.03 0.09 0.07 0.03 0.13
OH* 0.00 -0.03 0.06 0.02 0.01 0.08

131
Table A.9: Solvent effect on the reaction free energies of all elementary surface reactions of the propionic acid conversion to ethane and
ethylene over a Ni (111) catalyst surface model at 473 K based on iSMS methodology in the presence of water and 1,4-dioxane. ΔΔGrxn
indicates the reaction free energy difference between corresponding reaction in solvent and in a gas phase. Note that calculations for
solvents were performed with the help of the COSMO-RS package with three different Ni cavity radii: with default value, with a 10%
increased value and a 10% decreased value relative to the default.

ΔΔGrxn, eV @ 473 K
Water 1,4-dioxane

# Reaction Default +10% -10% Default +10% -10%


1 CH3CH2COOH* + 3* → CH3CH2CO*** + OH* 0.00 0.01 0.05 0.03 0.04 0.08
2 CH3CH2COOH* + 2* → CH3CHCOOH** + H* -0.07 -0.04 -0.07 -0.03 0.00 -0.03
3 CH3CH2CO***+ 2* → CH3CH2** + CO*** -0.14 -0.12 -0.18 -0.10 -0.07 -0.15
4 CH3CH2CO*** → CH3CHCO** + H* -0.06 -0.07 -0.07 -0.02 -0.02 -0.03
5 CH3CHCOOH** + * → CH3CHCO** + OH* 0.01 -0.02 0.05 0.04 0.02 0.09
132

6 CH3CHCOOH** + 3* → CH2CHCOOH**** + H* -0.02 -0.02 0.00 0.01 0.01 0.02


7 CH3CHCOOH** + 2*→CH3CCOOH*** + H* -0.03 -0.03 -0.06 -0.01 -0.02 -0.03
8 CH3CHCO** + 4* → CH3CH*** + CO*** -0.10 -0.07 -0.14 -0.09 -0.04 -0.12
9 CH3CHCO** + 3* → CH3CCO**** + H* -0.05 -0.03 -0.06 -0.02 0.00 -0.03
10 CH3CHCO** + 3* → CH2CHCO**** + H* -0.01 0.00 -0.01 0.01 0.02 0.01
11 CH2CHCOOH**** + * → CH2CHCO**** + OH* 0.02 0.00 0.04 0.04 0.04 0.08
12 CH2CHCOOH**** + * → CHCHCOOH**** + H* -0.05 -0.04 -0.07 -0.03 -0.03 -0.04
13 CH3CCOOH*** + 2* → CH3CCO**** + OH* -0.01 -0.02 0.04 0.03 0.04 0.09
14 CH3CCO**** + 2* → CH3C*** +CO*** -0.08 -0.06 -0.13 -0.08 -0.05 -0.12
15 CH2CHCO**** + 2* → CH2CH*** + CO*** -0.11 -0.09 -0.15 -0.10 -0.07 -0.14
16 CH2CHCO**** + * → CHCHCO**** + H* -0.08 -0.08 -0.11 -0.03 -0.04 -0.07
17 CHCHCOOH**** + * → CHCHCO**** + OH* -0.01 -0.04 0.00 0.03 0.02 0.06
18 CHCHCO**** + 3* → CHCH**** + CO*** -0.06 -0.04 -0.08 -0.08 -0.03 -0.09
19 CH2CH*** + 2* → CHCH**** + H* -0.03 -0.03 -0.03 -0.01 0.00 -0.02
20 CH2CH2** + 2* → CH2CH*** + H* -0.06 -0.05 -0.09 -0.03 -0.02 -0.06
21 CH2CH*** + * → CH2C*** + H* -0.09 -0.08 -0.13 -0.06 -0.05 -0.09
22 CH3C*** + * → CH2C*** + H* -0.08 -0.07 -0.11 -0.05 -0.04 -0.06
23 CH3CH*** + * → CH2CH*** + H* -0.03 -0.02 -0.03 -0.01 0.00 -0.01
24 CH3CH*** + * → CH3C*** + H* -0.03 -0.02 -0.05 -0.01 -0.01 -0.04
25 CH3CH2** + 2* → CH3CH*** + H* -0.03 -0.02 -0.02 -0.01 0.01 0.01
26 CH3CH3* + 2* → CH3CH2** + H* -0.05 -0.03 -0.05 -0.03 0.00 -0.02
27 CH3CH2** + * → CH2CH2** + H* 0.01 0.01 0.04 0.02 0.02 0.05
28 CH3CH2COOH* + 2* → CH3CH2COO** + H* -0.03 0.02 -0.06 0.00 0.04 -0.05
29 CH3CH2COO** + * → CH3CH2** + CO2* 0.06 0.03 0.17 0.04 0.03 0.13
30 CH3CH2COO** + 2* → CH3CHCOO*** + H* -0.27 -0.22 -0.32 -0.14 -0.10 -0.15
31 CH3CHCOOH** + 2* → CH3CHCOO*** + H* -0.23 -0.17 -0.30 -0.11 -0.06 -0.17
32 CH3CHCOOH** + 3* → CH3CH*** + COOH** -0.06 -0.05 -0.02 -0.02 0.00 0.03
33 CH3CHCOO*** + * → CH3CH*** + CO2* 0.31 0.24 0.47 0.17 0.13 0.29
133

34 CH3CHCOO*** + * → CH3CCOO*** + H* -0.26 -0.23 -0.27 -0.17 -0.14 -0.19


35 CH3CCOOH*** + * → CH3CCOO*** + H* -0.45 -0.36 -0.51 -0.27 -0.19 -0.32
36 CH3CCOOH*** + 2* → CH3C*** + COOH** -0.06 -0.04 -0.02 -0.02 0.01 0.02
37 CH2CHCOOH**** + * → CH2CH*** + COOH** -0.07 -0.05 -0.05 -0.04 -0.01 -0.01
38 CH3CCOO*** + * → CH3C*** + CO2* 0.54 0.44 0.68 0.33 0.26 0.44
39 COOH** → CO2* + H* 0.15 0.12 0.19 0.08 0.07 0.09
40 COOH** + 2* → CO*** + OH* -0.03 -0.04 -0.06 -0.03 -0.03 -0.05
41 H2O* + * → OH* + H* 0.00 0.01 0.00 0.00 0.02 0.01
Table A.10: Solvent effect on the activation free energies of all elementary surface reactions of the propionic acid conversion to ethane
and ethylene over a Ni (111) catalyst surface model at 473 K based on iSMS methodology in the presence of water and 1,4-dioxane.
ΔΔGact indicates the activation free energy difference between corresponding reaction in solvent and in a gas phase. Note that
calculations for solvents were performed with the help of the COSMO-RS package with three different Ni cavity radii: with default
value, with a 10% increased value and a 10% decreased value relative to the default.

ΔΔGact, eV @ 473 K
Water 1,4-dioxane

# Reaction Default +10% -10% Default +10% -10%


1 CH3CH2COOH* + 3* → CH3CH2CO*** + OH* 0.00 0.03 0.02 0.00 0.01 0.02
2 CH3CH2COOH* + 2* → CH3CHCOOH** + H* -0.03 -0.01 -0.02 -0.01 0.01 0.00
3 CH3CH2CO***+ 2* → CH3CH2** + CO*** -0.03 -0.02 -0.01 0.01 0.00 0.01
4 CH3CH2CO*** → CH3CHCO** + H* -0.03 -0.03 -0.02 0.00 0.00 0.01
5 CH3CHCOOH** + * → CH3CHCO** + OH* -0.03 -0.03 -0.05 0.00 -0.01 -0.01
134

6 CH3CHCOOH** + 3* → CH2CHCOOH**** + H* 0.00 -0.01 0.02 0.04 0.02 0.05


7 CH3CHCOOH** + 2*→CH3CCOOH*** + H* 0.00 0.00 0.01 0.01 0.00 0.02
8 CH3CHCO** + 4* → CH3CH*** + CO*** -0.01 0.00 -0.04 0.01 0.01 -0.02
9 CH3CHCO** + 3* → CH3CCO**** + H* -0.06 -0.05 -0.08 -0.03 -0.02 -0.05
10 CH3CHCO** + 3* → CH2CHCO**** + H* 0.00 0.00 -0.02 0.01 0.01 0.00
11 CH2CHCOOH**** + * → CH2CHCO**** + OH* 0.00 0.01 -0.03 0.01 0.02 -0.01
12 CH2CHCOOH**** + * → CHCHCOOH**** + H* 0.01 0.00 0.01 0.00 -0.01 0.01
13 CH3CCOOH*** + 2* → CH3CCO**** + OH* -0.08 -0.05 -0.08 -0.03 -0.01 -0.02
14 CH3CCO**** + 2* → CH3C*** +CO*** -0.08 -0.06 -0.11 -0.07 -0.05 -0.09
15 CH2CHCO**** + 2* → CH2CH*** + CO*** -0.01 -0.01 -0.01 0.00 0.00 -0.01
16 CH2CHCO**** + * → CHCHCO**** + H* 0.00 0.00 0.00 0.01 0.00 0.00
17 CHCHCOOH**** + * → CHCHCO**** + OH* -0.12 -0.07 -0.17 -0.06 -0.01 -0.11
18 CHCHCO**** + 3* → CHCH**** + CO*** -0.03 -0.01 -0.04 -0.02 0.00 -0.03
19 CH2CH*** + 2* → CHCH**** + H* 0.02 0.01 0.04 0.02 0.02 0.04
20 CH2CH2** + 2* → CH2CH*** + H* 0.01 -0.01 0.01 0.01 0.01 0.01
21 CH2CH*** + * → CH2C*** + H* -0.02 -0.02 -0.02 -0.01 -0.01 -0.01
22 CH3C*** + * → CH2C*** + H* -0.01 -0.03 -0.01 0.01 -0.01 0.02
23 CH3CH*** + * → CH2CH*** + H* -0.03 -0.03 -0.06 0.00 -0.02 -0.03
24 CH3CH*** + * → CH3C*** + H* 0.02 0.01 0.02 0.01 0.01 0.02
25 CH3CH2** + 2* → CH3CH*** + H* 0.01 0.01 0.03 0.02 0.02 0.05
26 CH3CH3* + 2* → CH3CH2** + H* -0.02 -0.02 0.00 -0.01 0.00 0.01
27 CH3CH2** + * → CH2CH2** + H* -0.01 -0.02 0.01 0.02 0.01 0.04
28 CH3CH2COOH* + 2* → CH3CH2COO** + H* -0.07 -0.03 -0.09 -0.03 -0.01 -0.06
29 CH3CH2COO** + * → CH3CH2** + CO2* -0.04 -0.06 -0.02 -0.03 -0.06 0.01
30 CH3CH2COO** + 2* → CH3CHCOO*** + H* -0.17 -0.16 -0.19 -0.09 -0.07 -0.06
31 CH3CHCOOH** + 2* → CH3CHCOO*** + H* -0.14 -0.10 -0.18 -0.07 -0.04 -0.09
32 CH3CHCOOH** + 3* → CH3CH*** + COOH** -0.02 -0.03 -0.02 0.02 0.00 0.03
33 CH3CHCOO*** + * → CH3CH*** + CO2* 0.12 0.08 0.17 0.10 0.06 0.13
135

34 CH3CHCOO*** + * → CH3CCOO*** + H* 0.12 0.07 0.15 0.09 0.06 0.11


35 CH3CCOOH*** + * → CH3CCOO*** + H* -0.18 -0.14 -0.26 -0.07 -0.04 -0.13
36 CH3CCOOH*** + 2* → CH3C*** + COOH** -0.05 -0.04 -0.05 -0.02 -0.01 -0.02
37 CH2CHCOOH**** + * → CH2CH*** + COOH** -0.01 -0.01 -0.04 0.00 0.00 -0.02
38 CH3CCOO*** + * → CH3C*** + CO2* 0.21 0.18 0.24 0.14 0.12 0.17
39 COOH** → CO2* + H* -0.18 -0.13 -0.25 -0.14 -0.09 -0.20
40 COOH** + 2* → CO*** + OH* 0.04 0.01 0.05 0.03 0.02 0.04
41 H2O* + * → OH* + H* -0.01 0.02 -0.05 -0.01 0.01 -0.03
Table A.11: Overall, decarbonylation, and decarboxylation turnover frequencies as well as
important steady state surface coverages for the HDO of PAC over a Ni(111) catalyst
surface model in the gas phase and in liquid water at 473 K. Calculations in water were
performed with the help of the iSMS and VASPsol schemes. Also, 𝑎𝑖 denotes the reaction
𝑛
order with respect to partial pressure/fugacity of component i and 𝑋𝑟𝑐 indicates Campbell’s
degree of rate control for reaction step n.

Properties Gas iSMS VASPsol

DCN TOF 3.46×10-08 6.70×10-09 1.68×10-08

DCX TOF 8.72×10-13 1.47×10-11 1.24×10-10

Overall TOF (s-1) 3.46×10-08 6.71×10-09 1.69×10-08

θ* 0.011 0.003 0.008


0.418
θH* 0.631 0.206

θCO*** 0.357 0.789 0.474

θCH3C*** 0.000 0.000 0.000

θPAC* 0.000 0.000 0.021

θCH3CH2COO** 0.001 0.000 0.010

𝒂𝑪𝑶 0.0 -0.7 -1.0

𝒂𝑯𝟐 -2.2 -1.4 -1.2

𝒂𝑷𝑨𝑪 1.0 1.0 1.0

𝑿𝟏𝒓𝒄 0.65 0.10 0.03

𝑿𝟖𝒓𝒄 0.26 0.01 0.68

𝑿𝟗𝒓𝒄 0.03 0.88 0.16

136
Figure A.1: Adsorption of different sized molecules (CO, CH3CH2, and CH3CH2C) on a
Ni (111) surface model with various cluster sizes used in iSMS calculations in liquid water
at 473 K. We employed a 3x2 cluster model for all surface reactions since the
computational cost for larger models increases dramatically while the solvent effect does
not change significantly. For adsorption/desorption processes of saturated molecules such
as C2H4, C3H6, C2H2, CH3CH2COOH, CO, CO2, H, H2O, where error cancellation of
solvation effects is smaller than for surface reactions, a 3x4 model with 86 metal atoms
was used.

137
a)

b)

Figure A.2: Lateral interaction correction terms for various intermediates using a 1/4 ML
coverage of CO and H. Shown are the correction terms of the lateral interaction effect of
H (a) and CO (b) for each individual adsorbed species.

138
a)

b)

Figure A.3: Lateral interaction correction terms for the transition state of each elementary
step using 1/4 ML coverages of CO and H. Shown are the correction terms of the lateral
interaction effect of H (a) and CO (b) on the transition state of each elementary step.

139
Figure A.4: Partial charges computed using the DDEC6 charge model on adsorbed CO (top
left), adsorbed PAC (top right), and coadsorbed CO and PAC on a Ni (111) catalyst surface
model. In the coadsorbed state, CO is a Lewis acid, which accepts electrons, while PAC is
a Lewis base, which donates electrons. This coadsorbed Lewis acid and base pair explains
the attractive interaction between adsorbed CO and PAC.

140
a) b)

Figure A.5: Differential zero-point energy corrected (ZPC) energy of adding a) an extra
CO as a function of number of CO already adsorbed on a Ni(111) surface containing 12
surface atoms with (3 × 2√3) periodicity and b) an extra H as a function of number of H
already adsorbed on the same Ni surface. Adding the fifth CO increases the energy
significantly (by over 0.45 eV) which we attribute to the fact that the surface is essentially
fully covered with 4 CO and adding an extra CO weakens the bond strength of adsorbed
CO molecules on the surface due to a strong repulsive interaction between them. On the
other hand, the differential ZPC barely increases when adding an extra hydrogen atom.
Consequently, it can be concluded that a CO molecule occupies 3 Ni sites on the surface
while H adsorbs on only 1 Ni site.

141
APPENDIX B

SUPPORTING INFORMATION FOR DEPENDENCY OF SOLVATION

EFFECTS IN METAL IDENTITY IN SURFACE REACTIONS

Zare, M.; Saleheen, M.; Kundu, S. K.; Heyden, A. Communications Chemistry 3, 187,
2020.

Reprinted here with the permission of the publisher

142
Table B.1: Simulation box size used for (111) facet of each metal surface.
Metal Simulation box size, Å3
Ni 39.75×43.03×49.01
Pd 44.53×48.20×49.01
Pt 44.98×48.69×49.01
Cu 41.04×44.43×49.01
Ag 46.85×50.71×49.01
Au 47.00×50.88×49.01

Table B.2: Lennard-Jones parameters of all elements used in this study. The parameters are
𝜎 12 𝜎 6
based on 12-6 Lennard-Jones potential, 𝑉𝑙𝑗 = 4𝜀 [( 𝑟 ) − ( 𝑟 ) ]. Ow and Hw represent
oxygen and hydrogen of TIP3P water model, respectively.

Element 𝜺 (kJ/mol) 𝝈 (Å)


Ni 23.640 2.274
Pd 25.732 2.511
Pt 32.635 2.535
Cu 19.748 2.331
Ag 19.079 2.633
Au 22.133 2.629
C 0.495 3.911
O 0.710 3.071
H 0.126 1.978
Ow 0.636 3.151
Hw 0.192 0.400

143
Table B.3: Average rotational correlation time (in picosecond) of liquid water molecules
residing in a 5 Å radius of an adsorbed ethylene glycol (C2H5OH*) species on a (111) facet
of six transition metal surfaces studied in the present work. The predicted correlation
functions have been fitted with three exponential functions to acquire average rotational
correlation times.
O-H cleavage C-H cleavage
Surface RS TS TS
Ni 218 147 79
Pd 227 101 223
Pt 183 46 53
Cu 115 67 21
Ag 82 111 196
Au 73 43 32

144
Table B.4: Number of hydrogen bonds formed from the interactions of the TIP3P water model and the reacting moiety in reactant state
(RS) and transition state (TS) of O-H and C-H bond cleavage of ethylene glycol over a (111) facet of six transition metal surfaces at 423
K. Acceptor indicates that OH functional groups of ethylene glycol accepts a hydrogen bond donated by the oxygen of water while
Donor indicates vice versa. All numbers are computed based on three independent simulations and hence possess 95% confidence
intervals (assuming a normal distribution).

O-H cleavage C-H cleavage

RS TS RS TS

Surface Acceptor Donor Acceptor Donor Acceptor Donor Acceptor Donor

Ni 1.50±0.08 0.34±0.06 2.66±0.13 0.17±0.03 1.50±0.08 0.34±0.06 1.96±0.06 0.14±0.04


145

Pd 1.23±0.08 0.63±0.10 2.14±0.10 0.05±0.04 1.23±0.08 0.63±0.10 1.37±0.05 0.22±0.03

Pt 1.17±0.10 0.11±0.02 1.82±0.04 0.00±0.00 1.17±0.10 0.11±0.02 1.23±0.10 0.17±0.02

Cu 1.47±0.08 0.35±0.03 2.47±0.12 0.12±0.05 1.47±0.08 0.35±0.03 1.79±0.12 0.24±0.03

Ag 1.51±0.11 0.40±0.05 2.81±0.10 0.01±0.00 1.51±0.11 0.40±0.05 1.38±0.06 0.26±0.02

Au 1.58±0.10 0.05±0.01 2.53±0.08 0.04±0.02 1.58±0.10 0.05±0.01 1.09±0.08 0.15±0.03


Table B.5: Partial charges (in atomic units), computed based on the NPA charge model, on the reacting moiety (ethylene glycol) in the
reactant (RS) and transition states (TS) for O-H and C-H bond cleavages over the (111) facet of six transition metal surfaces at 423 K.
Bold numbers indicate partial charges on the cleaving bond atoms (O1 and H5 in O-H bond cleavage, and C1 and H4 in C-H bond
cleavage).

Cleavage Surface State C1 C2 O1 O2 H1 H2 H3 H4 H5 H6

RS -0.11 -0.10 -0.75 -0.79 0.21 0.21 0.21 0.15 0.45 0.48
Ni
TS -0.11 -0.09 -0.80 -0.78 0.21 0.22 0.21 0.10 -0.12 0.47
RS -0.08 -0.10 -0.75 -0.72 0.21 0.20 0.22 0.17 0.50 0.47
Pd
TS -0.06 -0.11 -0.71 -0.71 0.23 0.18 0.22 0.15 0.10 0.47
RS
146

-0.11 -0.10 -0.75 -0.71 0.22 0.23 0.22 0.19 0.49 0.51
Pt
TS -0.12 -0.06 -0.68 -0.69 0.24 0.14 0.23 0.20 0.17 0.48
O-H
RS -0.10 -0.10 -0.75 -0.79 0.21 0.21 0.20 0.17 0.47 0.48
Cu
TS -0.10 -0.10 -0.86 -0.78 0.19 0.20 0.21 0.21 -0.16 0.48
RS -0.10 -0.09 -0.75 -0.77 0.20 0.21 0.20 0.18 0.48 0.48
Ag
TS -0.11 -0.09 -0.78 -0.79 0.20 0.20 0.19 0.14 -0.27 0.47
RS -0.13 -0.12 -0.71 -0.70 0.19 0.18 0.18 0.17 0.48 0.47
Au
TS -0.13 -0.12 -0.78 -0.69 0.20 0.18 0.20 0.12 -0.08 0.48
RS -0.11 -0.10 -0.75 -0.79 0.21 0.21 0.21 0.15 0.45 0.48
Ni
TS -0.23 -0.12 -0.73 -0.77 0.22 0.20 0.22 -0.17 0.45 0.48
RS -0.08 -0.10 -0.75 -0.72 0.22 0.20 0.22 0.17 0.50 0.47
Pd
TS -0.06 -0.12 -0.75 -0.68 0.25 0.21 0.23 0.03 0.50 0.48
RS -0.11 -0.10 -0.75 -0.71 0.22 0.23 0.22 0.19 0.49 0.51
Pt
TS -0.07 -0.11 -0.72 -0.71 0.25 0.24 0.24 0.13 0.50 0.51
C-H
RS -0.11 -0.10 -0.75 -0.79 0.21 0.21 0.21 0.17 0.47 0.48
Cu
TS
147

-0.21 -0.12 -0.72 -0.77 0.21 0.21 0.21 -0.32 0.47 0.49
RS -0.10 -0.09 -0.75 -0.77 0.20 0.21 0.20 0.18 0.48 0.48
Ag
TS -0.06 -0.13 -0.68 -0.75 0.21 0.23 0.22 -0.35 0.50 0.48
RS -0.13 -0.12 -0.71 -0.70 0.19 0.18 0.18 0.17 0.48 0.47
Au
TS 0.03 -0.18 -0.64 -0.68 0.21 0.20 0.20 -0.19 0.50 0.48
Table B.6: Pearson correlation coefficient (PCC) between descriptors used in this study. It has a value between +1 and −1, where +1 is
total positive linear correlation, 0 is no linear correlation, and −1 is a total negative linear correlation. ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 shows the free energy
of activation in vapor phase, H-bond denotes the change in mean of total hydrogen bonding (acceptor + donor) going from RS to TS,
MC (molecular charge-transfer) represents the change in the absolute sum of partial charges on the reacting moiety going from RS to
TS, and finally BC (cleaving-bond charge-transfer) expresses the change in sum of partial charges on the cleaving bond going from RS
to TS (see Table 3.2 in the main text for the value of descriptors).
MC,
Cleavage MC, BC H-bond, MC H-bond, BC H-bond, ∆𝑮𝒂𝒄𝒕,𝒈𝒂𝒔 BC, ∆𝑮𝒂𝒄𝒕,𝒈𝒂𝒔
∆𝑮𝒂𝒄𝒕,𝒈𝒂𝒔
O-H -0.99 0.83 -0.78 0.56 -0.67 0.65
C-H -0.96 0.66 -0.43 -0.54 -0.45 0.22

Table B.7: Data of a linear model, (𝑦 − 𝑦̅) = 𝛼1 (𝑓1 − 𝑓̅1 ) + 𝛼2 (𝑓2 − 𝑓̅2 ), for estimating the solvent effect (∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 −
∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ), computed by eSMS, on O-H and C-H bond cleavages of ethylene glycol at 423 K. We have not considered the combination
148

of MC and BC since they are totally correlated (see Table B.7). The fits are based on 6 data points for each cleavage which correspond
to six transition metals. Also, 𝑀𝐴𝐸 represents the mean of absolute errors associated with the fit, in eV, defined as the difference between
estimated and eSMS values.
O-H cleavage C-H cleavage
𝒇𝟏 , 𝒇𝟐 𝜶𝟏 𝜶𝟐 𝑀𝐴𝐸 𝜶𝟏 𝜶𝟐 𝑀𝐴𝐸
𝑎𝑐𝑡,𝑔𝑎𝑠
∆𝐺 , BC 0.13 0.27 0.14 -0.04 0.54 0.03
𝑎𝑐𝑡,𝑔𝑎𝑠
∆𝐺 , MC 0.06 -0.03 0.14 -0.10 -0.48 0.04
𝑎𝑐𝑡,𝑔𝑎𝑠
∆𝐺 , H-bond 0.02 0.10 0.14 -0.27 -0.44 0.04
H-bond, BC 0.43 0.48 0.12 0.03 0.61 0.04
H-bond, MC 0.30 -0.15 0.14 0.19 -0.69 0.04
Table B.8: Data of a quadratic model, (𝑦 − 𝑦̅) = 𝛼1 (𝑓1 − 𝑓̅1 ) + 𝛼2 (𝑓2 − 𝑓̅2 ) + 𝛼3 (𝑓1 − 𝑓̅1 )2 + 𝛼4 (𝑓2 − 𝑓̅2 )2 , for estimating the solvent
effect (∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ), computed by eSMS, on O-H bond cleavage of ethylene glycol at 423 K.

𝒇𝟏 , 𝒇𝟐 𝜶𝟏 𝜶𝟐 𝜶𝟑 𝜶𝟒 𝑀𝐴𝐸
∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 , BC -0.08 0.49 0.92 -4.38 0.02
∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 , MC -0.18 -0.44 1.29 -2.11 0.01
𝑎𝑐𝑡,𝑔𝑎𝑠
∆𝐺 , H-bond 0.11 0.44 -0.44 2.12 0.08
H-bond, BC 0.56 0.59 1.82 -2.68 0.04
H-bond, MC 0.90 -0.62 2.26 -1.19 0.04
149
Table B.9: Solvent effects on the free energy of activation (∆∆𝐺 𝑎𝑐𝑡 = ∆𝐺 𝑎𝑐𝑡,𝑙𝑖𝑞 − ∆𝐺 𝑎𝑐𝑡,𝑔𝑎𝑠 ) of O-H and C-H bond cleavage of ethylene
glycol on the (111) facet of six transition metal surfaces using the iSMS solvation scheme. Note that calculations were performed with
the help of the COSMO-RS package with three different metal cavity radii: with default value (2.2230 Å for all transition metal
elements), with a 10% increased value (2.4453 Å) and a 10% decreased (2.0007 Å) value relative to the default.
default +10% -10%
Ni -0.08 -0.06 -0.15
Cu -0.09 -0.08 -0.09
Ag -0.06 -0.06 -0.08
O-H cleavage
Pt -0.06 -0.05 -0.12
Pd 0.03 -0.01 0.02
Au -0.03 0.00 -0.05
Ni -0.04 -0.04 -0.04
Cu -0.12 -0.09 -0.15
Ag -0.03 -0.03 -0.05
150

C-H cleavage
Pt -0.09 -0.12 -0.15
Pd -0.01 -0.06 0.00
Au -0.17 -0.19 -0.08
Table B.10: Average fraction of TIP3P water molecules with different orientations (see Figure B.3 for explanation of the different
orientations) within the first water layer of the surface (see first peak in height distribution function of the water O atom in Figure B.4).
The data indicate that water orients itself similar across the metal surfaces and ~68% of them are in “parallel” orientation, ~23% in H-
up orientation, and only 9% in H-down orientation (which lets the water hardly form hydrogen bonds with the condensed phase).
Water Orientation Ni Cu Ag Pt Au Pd
Parallel 68.17±0.07 68.10±0.40 67.85±0.11 68.08±0.29 68.25±0.12 67.93±0.25
H-up 22.93±0.10 23.15±0.25 22.78±0.07 23.05±0.15 22.63±0.07 22.81±0.26
H-down 8.90±0.16 8.75±0.17 9.38±0.07 8.87±0.17 9.12±0.12 9.26±0.15
151
Figure B.1: Free-energy profiles for C-H bond cleavage of ethylene glycol in vapor and
aqueous phases over the (111) facet of six transition metal surfaces at 423 K without
considering vibrational contributions to the partition function. See Table 3.1 for
corresponding data that include vibrational contributions. The aqueous phase profile
portrays the average of three or more independent eSMS calculations possessing 95%
confidence intervals smaller than ±0.05 eV. The analogous plot for O-H bond cleavage is
provided in the main text.

152
Figure B.2: Graphical representation of a geometric hydrogen bond definition used in this
study. In the picture shown, water is the donor of hydrogen bonding, and ethylene glycol
is the acceptor of hydrogen bonding.

153
Figure B.3: Graphical representation of criterion used to distinguish different water
orientations at the surface. A water molecule in the first layer next to the surface is
considered in H-up conformation if the difference in z-coordinate of the H and O atom is
larger than half the H-H distance of the water molecule; this corresponds for TIP3P water
to an angle q show above larger than 52.26⁰. Water molecules that are not considered H-
up or H-down are labeled “parallel”.

154
Figure B.4: Height distribution function of water O and H over (111) facet of six transition
metal surfaces

155
APPENDIX C

SUPPORTING INFORMATION FOR LIQUID PHASE EFFECTS ON

ADSORPTION PROCESSES IN HETEROGENEOUS CATALYSIS

Zare, M.; Saleheen, M.; Singh, N.; Uline, M. J.; Faheem, M.; Heyden, A.

Submitted to Nature Communications

156
C.1 Computational Details
Planewave DFT Calculations. Vapor phase DFT calculations were carried out

employing periodic boundary conditions and using the Vienna Ab Initio Simulation

Package (VASP 5.4.4).1, 2 The frozen-core, all-electron projector augmented-wave (PAW)3

method was utilized to avoid the singularities of Kohn-Sham wavefunctions at the nuclear

positions. The exchange-correlation energy is calculated within the generalized gradient

approximation (GGA)4 using the Perdew-Burke-Ernzerhof (PBE)5, 6 functional. For phenol

adsorption, we used Grimme’s DFT-D3 methodology7 as implemented in VASP to

describe the van-der-Waals interactions. Brillouin zone integrations have been performed

with a 4×4×1 Monkhorst-Pack8 k-point grid, and electronic wavefunctions at each k-point

were expanded using a discrete plane-wave basis set with kinetic energies limited to 400

eV. For phenol adsorption, we increased the energy cutoff to 450 eV. Fractional

occupancies of bands were allowed within a window of 0.10 eV using a first-order

Methfessel–Paxton9 smearing method, which permitted us to calculate the entropic

contributions due to the smearing accurately. The modified version of the Makov-Payne10

method with Harris corrections were applied to the stress-tensor and forces to calculate

dipole and quadrupole corrections (along the surface normal) to the total energy. Using the

supercell approach, a 4×4-unit cell with four layers of metal atoms, in which the bottom

two layers were fixed while the top two layers were relaxed, has been constructed

employing a 15 Å vacuum on top of the surface, which restricted the interaction between

the periodic images along the surface normal. The self-consistent field (SCF) calculations

were converged to 1.0×10-7 eV. A force criterion of 0.02 eV Å-1 was used on relaxed atoms

for geometry optimizations. The transition state structure of the O-H bond cleavage of

157
ethylene glycol (EG) over a Cu(111) surface model was located using a combination of

climbing-image nudged elastic band11, 12 and dimer13, 14 methods. Finally, the minima and

the first order saddle points were validated by computing the Hessian matrix and

vibrational spectra.

Non-periodic cluster calculations. Cluster model DFT calculations in vacuum have

been carried out using the TURBOMOLE 7.2 program package.15, 16, 17 We used a cluster

model consisting of two layers with a total number of 51 metal atoms. Convergence of the

total QM/MM energy with respect to the lateral size and depth of the cluster geometry can

be found elsewhere.18 To represent the adsorbate atoms and the metal atoms, an improved

version of the default TURBOMOLE basis sets (def-bases) with split valence and

polarization functions (def2-SVP)19, 20 was used. Pt atoms were represented using scalar

relativistic effective core potentials (ECPs) in conjunction with split valence basis sets

augmented by polarization functions.20, 21 Exchange-correlation effects were calculated

using the PBE functional5, 6 (for phenol adsorption Grimme’s DFT-D3 methodology7 was

used). Finally, the RI-J approximation with auxiliary basis sets was used to approximate

the coulomb integrals.22, 23 Single point energy calculations were performed with an SCF

energy convergence criterion of 1.0 × 10-7 Hartree, and a Gauss-Chebyshev type spherical

grid, m4, was employed to perform the numerical integrations.16

C.2 Molecular Dynamics (MD) Simulations


The DL_POLY 4.03 molecular simulation program package24 was utilized for the

MD simulations. Using the supercell approach, the initial 4×4 unit cell for each clean

surface metal was augmented laterally to a 16×20 surface with further vacuum added in

the Z-direction resulting in a simulation box comprising of 1280 metal atoms. As a result,

158
a simulation box size of 44.98 Å × 48.69 Å × 49.01 Å for a Pt(111) surface and

41.04 Å × 44.43 Å × 49.01 Å for a Cu(111) surface were achieved. The QM cluster atoms

were then substituted into the constructed surface. The simulation box height was selected

based on the work from Behler et al.,25 who found that simulations of metal-water

interfaces should contain a water layer of ~40 Å height. Subsequently, the experimental

saturated liquid water density of ~1.0 g/cm3 at 298 K for CO and phenol adsorption

calculations on Pt(111) and Cu(111) surfaces and of ~0.92 g/cm3 at 423 K for O-H bond

cleavage’s calculations of ethylene glycol (EG) on Cu(111) as well as EG adsorption on

Pt(111) were achieved by packing the simulation box of the Pt(111) and Cu(111) surfaces

with ~3000 and ~2400 water molecules, respectively. All metal and adsorbate atoms were

kept fixed while the geometry of water molecules was constricted to that of TIP3P 26

geometry with the RATTLE algorithm,27 a velocity version of the SHAKE algorithm,28 in

conjunction with the velocity Verlet (VV) integrator29 to solve Newton’s equations of

motion. The force field parameters of liquid water were obtained from the TIP3P model.

The Lennard-Jones parameters were obtained from the OPLS force field30, 31, 32 for EG and

phenol and from Straub et al.33 for CO. In addition to the OPLS parameters, the Lennard-

Jones parameters from the combined B3LYP/6-31_G*/AMBER Potential34 were used for

the hydrogen atoms of EG. Lennard-Jones parameters for hydrogen atoms are important in

QM/MM optimizations that permit hydrogen atoms to approach water molecules and leave

the protective environment of a neighboring carbon or oxygen atom. The metal-water

interaction was represented by a Lennard-Jones metal potential.35 The LJ cross-term of the

intermolecular parameters are calculated by Lorentz-Berthelot mixing rules through


𝜎𝑖 +𝜎𝑗
equations 𝜎𝑖𝑗 = and 𝜀𝑖𝑗 = √𝜀𝑖 𝜀𝑗 . All Lennard-Jones parameters are included in Table
2

159
C.1. For O-H bond breaking of EG on the Cu(111) surface, and EG adsorption on the

Pt(111) surface, the charges for the QM atoms were estimated using the natural population

analysis (NPA)36 while for CO and phenol adsorptions on the Pt(111) and Cu(111) surfaces

the DDEC637 (a refinement of the Density Derived Electrostatic and Chemical (DDEC)

approach) charge model was used (to compare the performance of NPA and DDEC6 in our

scheme, we also repeated the calculations of CO adsorption on Pt(111) with the NPA

charge model). To describe the interaction of the TIP3P water point charges with the

quantum chemically described cluster model, we employed the periodic electrostatic

embedded cluster method (PEECM)38 as implemented in TURBOMOLE. Simulations

were carried out in a canonical ensemble (NVT) with Nosé-Hoover thermostat39, 40 with 1

ps relaxation time constant. Electrostatic interactions were accounted for by using the

Smoothed Particle Mesh Ewald (SPME) method41 with automatic parameter optimization

with default SPME precision. A 12 Å cutoff radius was adopted for the van-der-Waals

interactions and the transition between short and long range electrostatic interactions. If

not specified differently, all systems were equilibrated for 250 ps and sampled for 1000 ps

(1ns) using a 1 fs timestep to obtain 1000 MM conformations, which are 1 ps apart, for

each experiment. To optimize structures in an aqueous phase environment, we utilized the

fixed-size ensemble approximation with 5000 MM conformations (5ns sampling) recorded

every 1 ps.

160
Table C.1: Lennard-Jones parameters of all elements in this study based on 𝑉𝑖𝑗 =
12 6
𝜎 𝜎
4𝜀𝑖𝑗 [( 𝑟𝑖𝑗 ) − ( 𝑟 𝑖𝑗 ) ], where 𝜀 is the depth of the potential well and 𝜎 is the finite distance
𝑖𝑗 𝑖𝑗

at which the inter-particle potential is zero. Ow and Hw represent oxygen and hydrogen of
TIP3P water model, respectively. ‘Ho’ in the phenol molecule indicates the hydrogen of
the OH functional group.

Molecule or Metal Element 𝜺 (kJ/mol) 𝝈 (Å)


Pt 32.635 2.535
Metal35
Cu 19.748 2.331
C 0.495 3.911
Ethylene glycol30, 31 O 0.710 3.071
H34 0.126 1.978
Ow 0.636 3.151
TIP3P water
Hw 0.192 0.400
C 0.264 3.491
Carbon monoxide33
O 0.651 3.136
C 0.293 3.550
O 0.711 3.070
Phenol32
H 0.126 2.420
Ho34 0.126 0.944

C.3 Rotational Correlation time


Adequately sampling of the potential energy surface for all relevant configurations

of the system is a key for a reliable estimate of ensemble averages and a prediction of the

liquid phase effect on free energies of elementary processes.42, 43 Yet, extensive exploration

of the configuration space for the relevant configurations is challenging due to the

substantial number of solvent molecules in the system. In addition, it is not a priori obvious

how much water sampling is required for adsorbed species on metal surfaces.

161
Consequently, the average rotational correlation time for water molecules in proximity (up

to 5 Å) of the adsorbed moieties for each free energy calculation is computed. And the

configuration space is then sampled long enough, relative to the average rotational

correlation time, to assure that relevant configurations of the system are sampled

sufficiently. A three-exponential fit to the rotational correlation time function of water

molecules in 5 Å radius of each adsorbed species, investigated in this study, is presented in

Figure C.1. Table C.2 also lists the average rotational correlation time and constants of the

3-exponential fits of Figure C.1.

Table C.2: Average rotational correlation time (RCT)44 and constants (in pico-second) of
𝑡 𝑡 𝑡
−( ) −( ) −( )
3-Exponential fits (𝐹(𝑡) = 𝑎1 𝑒 𝜏1 + 𝑎2 𝑒 𝜏2 + (1 − 𝑎1 − 𝑎2 )𝑒 𝜏3 ) in Figure C.1. The
𝑎1 𝜏1 2 +𝑎2 𝜏2 2 +(1−𝑎1 −𝑎2 )𝜏3 2
average rotational correlation time is computed as: 𝑅𝐶𝑇 = .
𝑎1 𝜏1 +𝑎2 𝜏2 +(1−𝑎1 −𝑎2 )𝜏3

Phenol Phenol
EG on EG on CO on CO on
on on
Cu(111) Pt(111) Cu(111) Pt(111)
Cu(111) Pt(111)
𝒂𝟏 0.127 0.102 0.520 0.534 0.305 0.394
𝒂𝟐 0.447 0.451 0.354 0.360 0.512 0.436
𝝉𝟏 122.491 86.866 18.707 46.349 54.223 60.506
𝝉𝟐 2.176 2.416 2.330 2.143 4.106 4.102
𝝉𝟑 0.271 0.290 0.069 0.047 0.188 0.186
RCT 115.0 76.6 17.4 45.0 48.5 56.5

162
Figure C.1: Rotational correlation time function of liquid water molecules residing in a 5
Å radius of adsorbed ethylene glycol (C2H6O2*) species on Cu(111) and Pt(111) at 423K,
of adsorbed carbon monoxide (CO*) on Cu(111) and Pt(111) at 298 K, and of adsorbed
phenol (C6H5OH*) on Cu(111) and Pt(111) at 298 K. The correlation time functions have
been fitted using three exponential functions to obtain the average rotational correlation
times (see Table C.2).

163
C.4 QM/MM Free Energy Calculation
The explicit Solvation model for Metal Surfaces (eSMS) methodology18, based on

QM/MM-(M)FEP (free energy perturbation)45, 46, which has been implemented in our

program packages and which was successfully employed in our previous studies,47, 48 was

used to carry out the free energy calculations. Computation of the free energy difference

between two states was carried out using the Bennett Acceptance Ratio (BAR),49 which is

a more efficient and precise way than exponential averaging (EXP) when two states are

sampled50 since it uses both the forward and reverse distributions simultaneously. Finally,

the whole free energy estimation procedure has been repeated three times, independently,

by different initialization of the velocities using different seeds for the pseudorandom

number generator51 implemented in the DL_POLY program package, to establish 95%

confidence intervals for evaluating the adsorption free energy of the investigated

adsorbates (assuming a normal distribution).52 More than three independent simulations

were carried out only if these three experiments could not result in a 95% confidence

interval of lower than 0.05 eV. All uncertainties reported in this study are 95% confidence

intervals.

C.5 Implicit Solvation Calculations


Implicit solvation calculations were performed with the iSMS method,53 and

VASPsol54, 55 with a relative permittivity of water at desired reaction conditions.56 We used

the default values for the VASPsol parameters. While these parameters are likely most

accurate only for simulations at 298 K and not at 423 K, they are optimized parameters of

the solvent model that cannot easily be obtained at other temperatures. All other

164
computational details for periodic implicit solvation calculations were kept the same as in

our periodic vapor phase calculations.

C.6 Liquid-phase phenol adsorption isotherm derivation


57, 58
Equations C.1 and C.2 are from Nitta et al. and were derived solely from

statistical mechanics.

𝑔−𝑙 𝜃𝑃ℎ 𝜃𝑃ℎ


𝑃ℎ(𝑔) + 𝑛 ∗ (𝑙𝑖𝑞) ↔ 𝑃ℎ𝑛∗ (𝑙𝑖𝑞) 𝐾𝑎𝑑𝑠,𝑃ℎ = 𝑓𝑃ℎ = 𝑃𝑃ℎ (C.1)
( )𝜃𝑛 ( )𝜃𝑛
1 𝑏𝑎𝑟 ∗ 1 𝑏𝑎𝑟 ∗

2 2
𝑔−𝑙 𝜃𝐻 𝜃𝐻
𝐻2 (𝑔) + 2 ∗ (𝑙𝑖𝑞) ↔ 2𝐻 ∗ (𝑙𝑖𝑞) 𝐾𝑎𝑑𝑠,𝐻2 = 𝑓𝐻 = 𝑃𝐻 (C.2)
( 2 )𝜃∗2 ( 2 )𝜃∗2
1 𝑏𝑎𝑟 1 𝑏𝑎𝑟

1 = 𝜃∗ + 𝜃𝐻 + 𝑛𝜃𝑃ℎ (C.3)

where 𝜃𝑃ℎ , 𝜃𝐻 , and 𝜃∗ are the surface coverage of phenol, hydrogen, and the free site,

respectively, variable 𝑛 indicates the number of sites that a phenol molecule can occupy

𝑔−𝑙
when adsorbed on the surface; 𝐾𝑎𝑑𝑠 is the equilibrium constant of adsorption of an

adsorbate on the surface from its free state in the gas phase; 𝑓𝑃ℎ , and 𝑓𝐻 are the phenol and

hydrogen fugacities in the gas phase, respectively, which can be reduced to partial

pressures with ideal gas assumption. We note that in the following, we define the “surface

coverage” as the number of surface species divided by the number of surface atoms, which

is strictly speaking only the coverage for 𝑛=1. From Henry’s law we have

𝑚𝑜𝑙
𝑓𝑃ℎ ⁄1𝑏𝑎𝑟 = 𝑃𝑃ℎ ⁄1𝑏𝑎𝑟 = 𝐾𝐻,𝑃ℎ 𝐶𝑃ℎ ⁄( 𝐿
) (C.4)

where 𝐾𝐻,𝑃ℎ is the Henry’s constant of phenol and 𝐶𝑃ℎ is the phenol concentration in the

aqueous phase. In the following, we neglect the concentration and pressure unit but work

in bar and mol/L.

Substitute (C.4) in (C.1)

165
𝑔−𝑙 𝜃𝑃ℎ
𝐾𝑎𝑑𝑠,𝑃ℎ 𝐾𝐻,𝑃ℎ = (C.5)
𝐶𝑃ℎ 𝜃∗𝑛

We also have
𝑔−𝑙 𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙 𝑔 𝑔𝑎𝑠→𝑙𝑖𝑞
𝐾𝑎𝑑𝑠,𝑃ℎ = 𝐾𝑎𝑑𝑠,𝑃ℎ 𝐾𝑃ℎ or ∆𝐺𝑎𝑑𝑠,𝑃ℎ = ∆𝐺𝑎𝑑𝑠,𝑃ℎ + ∆∆𝐺𝑃ℎ (C.6)

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙
where ∆∆𝐺𝑃ℎ is the solvent effect on phenol adsorption, ∆𝐺𝑎𝑑𝑠,𝑃ℎ is the adsorption

𝑔
free energy of phenol in the aqueous phase from its free state in the gas phase, and ∆𝐺𝑎𝑑𝑠,𝑃ℎ

is free energy of phenol adsorption in the gas phase. Hence, (C.5) becomes

𝑔−𝑙 𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞 𝜃𝑃ℎ 𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞


𝐾𝑎𝑑𝑠,𝑃ℎ 𝐾𝐻,𝑃ℎ = 𝐾𝑎𝑑𝑠,𝑃ℎ 𝐾𝐻,𝑃ℎ 𝐾𝑃ℎ =𝐶 𝑛 or 𝜃𝑃ℎ = 𝐾𝑎𝑑𝑠,𝑃ℎ 𝐾𝐻,𝑃ℎ 𝐾𝑃ℎ 𝐶𝑃ℎ 𝜃∗𝑛 (C.7)
𝑃ℎ 𝜃∗

In the same way:

2
𝑔−𝑙 𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞 𝜃𝐻 𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞
𝐾𝑎𝑑𝑠,𝐻2 = 𝐾𝑎𝑑𝑠,𝐻2 𝐾𝐻2 = 𝑃𝐻 → 𝜃𝐻 = √𝐾𝑎𝑑𝑠,𝐻2 𝐾𝐻2 𝑃𝐻2 𝜃∗ (C.8)
( 2 )𝜃∗2
1 𝑏𝑎𝑟

Substitute (C.7) and (C.8) into (C.3):

𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞


1 = 𝜃∗ + √𝐾𝑎𝑑𝑠,𝐻2 𝐾𝐻2 𝑃𝐻2 𝜃∗ + 𝑛[𝐾𝑎𝑑𝑠,𝑃ℎ 𝐾𝐻,𝑃ℎ 𝐾𝑃ℎ 𝐶𝑃ℎ 𝜃∗𝑛 ] (C.9)

Taking into account lateral interactions, we have:

𝑔𝑎𝑠
𝑔𝑎𝑠 −∆𝐺𝑃ℎ 𝑔𝑎𝑠 𝑔𝑎𝑠 𝑔𝑎𝑠 𝑔𝑎𝑠
𝐾𝑎𝑑𝑠,𝑃ℎ = 𝑒𝑥𝑝( 𝑅𝑇
) and ∆𝐺𝑃ℎ = ∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) + 𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ + 𝛼𝑃ℎ,𝐻 𝜃𝐻 (C.10)

𝑔𝑎𝑠
𝑔𝑎𝑠 −∆𝐺𝐻 𝑔𝑎𝑠 𝑔𝑎𝑠 𝑔𝑎𝑠 𝑔𝑎𝑠
2
𝐾𝑎𝑑𝑠,𝐻2 = 𝑒𝑥𝑝( 𝑅𝑇
) and ∆𝐺𝐻2 = ∆𝐺𝐻2 (𝜃𝐻 = 0) + 𝛼𝐻,𝐻 𝜃𝐻 + 𝛼𝑃ℎ,𝐻 𝜃𝑃ℎ (C.11)

𝑔𝑎𝑠 𝑔𝑎𝑠 𝑔𝑎𝑠


where 𝛼𝑃ℎ,𝑃ℎ , 𝛼𝑃ℎ,𝐻 , and 𝛼𝐻,𝐻 are constants associated with adsorbed phenol-phenol,

phenol-hydrogen, and hydrogen-hydrogen lateral interactions, respectively. If 𝜃𝐻 is small

(because 𝑃𝐻2 is sufficiently small), (C.9) reduces to:

𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞
1 = 𝜃∗ + 𝑛[𝐾𝑎𝑑𝑠,𝑃ℎ 𝐾𝐻,𝑃ℎ 𝐾𝑃ℎ 𝐶𝑃ℎ 𝜃∗𝑛 ] (C.12)

Substituting (C.10) in (C.7):

166
𝑔𝑎𝑠
𝑔𝑎𝑠 𝑔𝑎𝑠→𝑙𝑖𝑞 −𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ
𝜃𝑃ℎ = 𝐾
⏟𝑎𝑑𝑠,𝑃ℎ (𝜃𝑃ℎ = 0)𝐾𝐻,𝑃ℎ 𝐾𝑃ℎ 𝐶𝑃ℎ 𝜃∗𝑛 𝑒𝑥𝑝( 𝑅𝑇
) (C.13)
𝛽

Substituting (C.10) in (C.12):

𝑔𝑎𝑠
𝑔𝑎𝑠→𝑙𝑖𝑞 −𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ
1 = 𝜃∗ + 𝑛 [𝛽𝐾𝑃ℎ 𝐶𝑃ℎ 𝜃∗𝑛 𝑒𝑥𝑝( 𝑅𝑇
)] (C.14)

We can also write (C.10) for the liquid phase:

𝑔−𝑙
𝑔−𝑙 −∆𝐺𝑃ℎ 𝑔−𝑙 𝑔−𝑙 𝑙𝑖𝑞 𝑙𝑖𝑞
𝐾𝑎𝑑𝑠,𝑃ℎ = 𝑒𝑥𝑝( ) and ∆𝐺𝑃ℎ = ∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) + 𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ + 𝛼𝑃ℎ,𝐻 𝜃𝐻 (C.15)
𝑅𝑇

Subtracting (C.10) from (C.15)

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞


∆∆𝐺𝑃ℎ = ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) + ∆𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ + ∆𝛼𝑃ℎ,𝐻 𝜃𝐻 (C.16)

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑙𝑖𝑞 𝑔𝑎𝑠


where ∆𝛼𝑃ℎ,𝑃ℎ = 𝛼𝑃ℎ,𝑃ℎ − 𝛼𝑃ℎ,𝑃ℎ . Hence, for low 𝑃𝐻2 and 𝜃𝐻 (C.14) becomes:

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠
𝑔𝑎𝑠→𝑙𝑖𝑞 −∆𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ −𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ
1 = 𝜃∗ + 𝑛 𝛽𝐾𝑃ℎ (𝜃𝑃ℎ = 0)𝐶𝑃ℎ 𝜃∗𝑛 𝑒𝑥𝑝(
⏟ 𝑅𝑇
)𝑒𝑥𝑝( 𝑅𝑇
) or
𝑙𝑖𝑞
−𝛼 𝜃
𝑃ℎ,𝑃ℎ 𝑃ℎ
[ 𝑒𝑥𝑝(
𝑅𝑇
) ]
𝑙𝑖𝑞
𝑔𝑎𝑠→𝑙𝑖𝑞 −𝛼 𝜃𝑃ℎ
1 = 𝜃∗ + 𝑛𝜃𝑃ℎ = 𝜃∗ + 𝑛 [𝛽𝐾𝑃ℎ (𝜃𝑃ℎ = 0)𝐶𝑃ℎ 𝜃∗𝑛 𝑒𝑥𝑝( 𝑃ℎ,𝑃ℎ
𝑅𝑇
)] (C.17)

and (C.13) becomes:

𝑙𝑖𝑞
𝑔𝑎𝑠→𝑙𝑖𝑞 −𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ
𝜃𝑃ℎ = 𝛽𝐾𝑃ℎ (𝜃𝑃ℎ = 0)𝐶𝑃ℎ 𝜃∗𝑛 𝑒𝑥𝑝( 𝑅𝑇
) (C.18)

Possibly the easiest way of solving equations (C.17) and (C.18) is rewriting them into a

self-consistent loop:

𝑙𝑖𝑞
𝑔𝑎𝑠→𝑙𝑖𝑞 −𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ
𝜃𝑃ℎ = 𝛽𝐾𝑃ℎ (𝜃𝑃ℎ = 0)𝐶𝑃ℎ (1 − 𝑛𝜃𝑃ℎ )𝑛 𝑒𝑥𝑝( 𝑅𝑇
) (C.19)

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑙𝑖𝑞
𝑔𝑎𝑠 −{∆∆𝐺𝑃ℎ (𝜃𝑃ℎ =0)+𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ }
or: 𝜃𝑃ℎ = 𝐾𝑒𝑞,𝑃ℎ (𝜃𝑃ℎ = 0)𝐾𝐻,𝑃ℎ 𝐶𝑃ℎ (1 − 𝑛𝜃𝑃ℎ )𝑛 𝑒𝑥𝑝 [ 𝑅𝑇
] (C.20)

167
Equation (C.20) can be solved iteratively using fixed-point iteration that has unfortunately

only a small convergence radius. Alternatively, we can use a one-dimensional root finder

on Eq. C.20.

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑙𝑖𝑞
𝑔𝑎𝑠 −{∆∆𝐺𝑃ℎ (𝜃𝑃ℎ =0)+𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ }
0 = 𝐾𝑒𝑞,𝑃ℎ (𝜃𝑃ℎ = 0)𝐾𝐻,𝑃ℎ 𝐶𝑃ℎ (1 − 𝑛𝜃𝑃ℎ )𝑛 𝑒𝑥𝑝 [ ] − 𝜃𝑃ℎ (C.21)
𝑅𝑇

We know a bracket for 𝜃𝑃ℎ (between 0 and 1/n) such that any bracketing method such as

bisection or better Brent’s method should be ideal. Note that the equations cannot be solved

analytically for n≠1. Specifically, we solved Eq. C.21 for n = 1 to 5. Once we have a

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑙𝑖𝑞
function that can solve for 𝜃𝑃ℎ for any given 𝐶𝑃ℎ , ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0), and 𝛼𝑃ℎ,𝑃ℎ , we

𝑔𝑎𝑠→𝑙𝑖𝑞 𝑙𝑖𝑞
can optimize for ∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) and 𝛼𝑃ℎ,𝑃ℎ .

Next, we can also test how the presence of 𝜃𝐻 invalidates the results. Since Eq. C.18

remains valid and 𝜃𝐻 = 𝑐𝑜𝑛𝑠𝑡 × 𝜃∗ remains a good approximation at constant temperature

and hydrogen partial pressure, we have 1 = 𝜃∗ + 𝜃𝐻 + 𝑛𝜃𝑃ℎ = 𝜃∗ (1 + 𝑐𝑜𝑛𝑠𝑡) + 𝑛𝜃𝑃ℎ or

1−𝑛𝜃𝑃ℎ
𝜃∗ = 1+𝑐𝑜𝑛𝑠𝑡 and we just have to solve Eq. C.22 for possible values of the constant ‘𝑐𝑜𝑛𝑠𝑡’

of 0.1, 1, and 10:

𝑛 𝑔𝑎𝑠→𝑙𝑖𝑞 𝑙𝑖𝑞
𝑔𝑎𝑠 1−𝑛𝜃 −{∆∆𝐺𝑃ℎ (𝜃𝑃ℎ =0)+𝛼𝑃ℎ,𝑃ℎ 𝜃𝑃ℎ }
𝑃ℎ
0 = 𝐾𝑎𝑑𝑠,𝑃ℎ (𝜃𝑃ℎ = 0)𝐾𝐻,𝑃ℎ 𝐶𝑃ℎ (1+𝑐𝑜𝑛𝑠𝑡 ) 𝑒𝑥𝑝 [ 𝑅𝑇
] − 𝜃𝑃ℎ (C.22)

168
Table C.3: Charges of the QM system (in e) in the initial state, where there exist
electrostatic and van der Waals interactions between the solvent and the adsorbate, and in
the final state, where all interactions have vanished as if the adsorbate was removed
completely, for different adsorbates investigated in this work. The initial and final states
correspond to states I and V in Figure C.1 in the main text, respectively. a) Reactant and
transition states of O-H bond cleavage of EG on a Cu(111) surface model using the NPA
charge model, b) CO on a Pt(111) surface model using the NPA and DDEC6 charge
models, c) ethylene glycol (EG) on the Pt(111) and Cu(111) surfaces using the NPA charge
model, d) CO on Pt(111) and Cu(111) surfaces using the DDEC6 charge model, and e)
phenol on the Pt(111) and Cu(111) surfaces using the DDEC6 charge model (Ho in the
phenol molecule indicates the hydrogen of the OH functional group).

a)

Reactant State Transition State


Atom
Initial Final Initial Final
Cu 0.042 0.046 0.042 0.045
Cu -0.020 -0.021 -0.025 -0.019
Cu 0.059 0.049 0.082 0.046
Cu 0.017 0.015 0.011 0.011
Cu -0.009 -0.026 -0.003 -0.023
Cu -0.011 -0.025 -0.007 -0.019
Cu 0.011 0.012 0.009 0.013
Cu 0.027 0.031 0.047 0.029
Cu 0.004 0.004 0.007 0.002
Cu -0.056 -0.081 -0.030 -0.082
Cu -0.003 0.004 0.009 -0.001
Cu 0.026 0.030 0.010 0.034
Cu -0.021 -0.010 -0.027 -0.005
Cu -0.024 -0.022 -0.003 -0.024
Cu -0.039 -0.025 -0.036 -0.022
Cu -0.004 -0.005 -0.004 -0.011
Cu -0.007 -0.012 -0.002 -0.015
Cu 0.029 0.035 0.045 0.037
Cu -0.019 -0.015 -0.009 -0.011
Cu 0.063 0.065 0.057 0.064
Cu 0.057 0.062 0.036 0.066
Cu 0.018 0.011 0.014 0.007
Cu 0.025 0.012 0.023 0.014
Cu -0.009 0.005 -0.012 0.001
Cu 0.019 0.034 0.023 0.030
Cu -0.051 -0.027 -0.034 -0.025
Cu -0.013 -0.032 -0.040 -0.032
Cu 0.038 0.030 0.038 0.030

169
Cu 0.040 0.025 0.036 0.020
Cu -0.058 -0.030 -0.067 -0.026
Cu 0.128 0.019 0.150 0.003
Cu -0.043 -0.028 -0.061 -0.032
Cu 0.036 0.021 0.009 0.025
Cu 0.003 -0.002 0.000 -0.003
Cu -0.031 -0.020 -0.009 -0.025
Cu -0.074 -0.034 0.197 -0.011
Cu -0.069 -0.036 0.083 -0.035
Cu -0.029 -0.018 -0.017 -0.019
Cu 0.005 -0.007 -0.001 0.008
Cu 0.014 0.015 0.014 0.015
Cu -0.030 -0.037 0.018 -0.040
Cu -0.023 0.019 0.107 0.023
Cu -0.033 -0.042 0.003 -0.045
Cu 0.007 0.021 0.009 0.018
Cu 0.011 0.012 0.017 0.014
Cu -0.041 -0.047 -0.034 -0.050
Cu -0.053 -0.046 -0.043 -0.047
Cu 0.019 0.014 0.025 0.013
Cu 0.018 0.021 0.019 0.024
Cu 0.024 0.024 0.013 0.020
Cu 0.015 0.014 0.015 0.006
C -0.106 0.000 -0.101 0.000
C -0.097 0.000 -0.096 0.000
O -0.749 0.000 -0.856 0.000
O -0.790 0.000 -0.780 0.000
H 0.208 0.000 0.193 0.000
H 0.216 0.000 0.199 0.000
H 0.206 0.000 0.212 0.000
H 0.173 0.000 0.212 0.000
H 0.472 0.000 -0.164 0.000
H 0.483 0.000 0.479 0.000

b)

NPA DDEC6
Atom
Initial Final Initial Final
Pt 0.075 0.081 0.000 -0.001
Pt 0.076 0.080 -0.003 0.000
Pt 0.084 0.078 0.007 0.006
Pt -0.093 -0.082 -0.001 -0.003
Pt 0.081 0.073 0.003 0.003
Pt 0.089 0.087 0.008 0.000

170
Pt -0.069 -0.030 -0.019 0.010
Pt -0.010 -0.018 0.010 0.006
Pt 0.077 0.083 -0.002 -0.001
Pt 0.096 0.094 0.003 0.001
Pt -0.082 -0.043 -0.019 0.009
Pt -0.054 0.002 -0.011 0.016
Pt -0.039 -0.038 0.005 0.005
Pt 0.086 0.097 -0.001 0.000
Pt 0.029 0.034 -0.005 -0.007
Pt -0.029 -0.040 0.010 0.006
Pt -0.047 -0.044 0.005 0.004
Pt 0.035 0.029 -0.010 -0.010
Pt 0.103 0.092 0.003 0.002
Pt 0.056 0.061 -0.008 -0.006
Pt 0.092 0.092 0.003 0.002
Pt -0.064 -0.049 -0.020 -0.018
Pt 0.047 0.036 -0.010 -0.009
Pt -0.054 -0.051 -0.011 -0.013
Pt 0.002 0.012 -0.042 -0.043
Pt -0.021 -0.029 0.042 0.043
Pt -0.034 -0.030 0.043 0.044
Pt 0.003 0.009 -0.041 -0.043
Pt -0.014 -0.011 -0.049 -0.046
Pt -0.051 -0.034 0.017 0.040
Pt -0.063 -0.074 -0.020 -0.002
Pt -0.026 -0.030 0.036 0.038
Pt -0.003 -0.011 -0.049 -0.047
Pt -0.035 -0.041 -0.007 -0.005
Pt -0.029 -0.022 0.026 0.045
Pt 0.048 -0.030 0.028 -0.041
Pt -0.066 -0.048 -0.016 0.004
Pt -0.023 -0.019 0.040 0.039
Pt -0.039 -0.047 -0.001 -0.002
Pt 0.003 0.001 -0.027 -0.029
Pt -0.032 -0.033 0.014 0.035
Pt -0.073 -0.062 -0.022 -0.002
Pt -0.033 -0.023 0.029 0.030
Pt 0.001 -0.006 -0.031 -0.029
Pt 0.007 0.009 -0.030 -0.028
Pt -0.012 -0.015 0.043 0.044
Pt -0.022 -0.018 0.047 0.048
Pt 0.009 0.000 -0.029 -0.029
Pt -0.050 -0.041 -0.013 -0.012
Pt 0.017 0.009 -0.039 -0.038
Pt -0.043 -0.040 -0.012 -0.014

171
C 0.480 0.000 0.241 0.000
O -0.387 0.000 -0.115 0.000

c)

EG on Pt(111) EG on Cu(111)
Atom
Initial Final Initial Final
Metal 0.076 0.079 0.042 0.046
Metal 0.074 0.074 -0.020 -0.021
Metal 0.082 0.080 0.059 0.049
Metal -0.097 -0.094 0.017 0.015
Metal 0.078 0.081 -0.009 -0.026
Metal 0.096 0.088 -0.011 -0.025
Metal -0.026 -0.029 0.011 0.012
Metal -0.010 -0.024 0.027 0.031
Metal 0.082 0.084 0.004 0.004
Metal 0.095 0.091 -0.056 -0.081
Metal -0.044 -0.045 -0.003 0.004
Metal 0.004 0.000 0.026 0.030
Metal -0.040 -0.044 -0.021 -0.010
Metal 0.093 0.096 -0.024 -0.022
Metal 0.031 0.033 -0.039 -0.025
Metal -0.028 -0.041 -0.004 -0.005
Metal -0.045 -0.044 -0.007 -0.012
Metal 0.036 0.033 0.029 0.035
Metal 0.095 0.091 -0.019 -0.015
Metal 0.067 0.075 0.063 0.065
Metal 0.094 0.096 0.057 0.062
Metal -0.051 -0.059 0.018 0.011
Metal 0.038 0.040 0.025 0.012
Metal -0.048 -0.049 -0.009 0.005
Metal 0.002 0.014 0.019 0.034
Metal -0.038 -0.024 -0.051 -0.027
Metal -0.026 -0.027 -0.013 -0.032
Metal -0.004 0.000 0.038 0.030
Metal -0.019 -0.013 0.040 0.025
Metal -0.032 -0.035 -0.058 -0.030
Metal -0.086 -0.049 0.128 0.019
Metal -0.042 -0.025 -0.043 -0.028
Metal -0.013 -0.010 0.036 0.021
Metal -0.047 -0.043 0.003 -0.002
Metal -0.006 -0.018 -0.031 -0.020
Metal 0.014 -0.039 -0.074 -0.034
Metal -0.133 -0.048 -0.069 -0.036

172
Metal -0.032 -0.019 -0.029 -0.018
Metal -0.044 -0.042 0.005 -0.007
Metal 0.000 -0.001 0.014 0.015
Metal -0.033 -0.036 -0.030 -0.037
Metal -0.097 -0.056 -0.023 0.019
Metal -0.041 -0.015 -0.033 -0.042
Metal -0.007 -0.005 0.007 0.021
Metal 0.004 0.010 0.011 0.012
Metal -0.033 -0.018 -0.041 -0.047
Metal -0.029 -0.020 -0.053 -0.046
Metal -0.003 -0.010 0.019 0.014
Metal -0.044 -0.038 0.018 0.021
Metal 0.000 0.002 0.024 0.024
Metal -0.040 -0.046 0.015 0.014
C -0.107 0.000 -0.106 0.000
C -0.096 0.000 -0.097 0.000
O -0.748 0.000 -0.749 0.000
O -0.720 0.000 -0.790 0.000
H 0.215 0.000 0.208 0.000
H 0.226 0.000 0.216 0.000
H 0.219 0.000 0.206 0.000
H 0.196 0.000 0.173 0.000
H 0.486 0.000 0.472 0.000
H 0.505 0.000 0.483 0.000

d)

CO on Pt(111) CO on Cu(111)
Atom
Initial Final Initial Final
Metal 0.000 -0.001 0.020 0.010
Metal -0.003 0.000 -0.001 -0.003
Metal 0.007 0.006 0.006 0.011
Metal -0.001 -0.003 -0.007 -0.013
Metal 0.003 0.003 0.005 0.002
Metal 0.008 0.000 -0.008 0.003
Metal -0.019 0.010 -0.003 -0.016
Metal 0.010 0.006 0.002 -0.003
Metal -0.002 -0.001 0.014 0.019
Metal 0.003 0.001 -0.034 -0.016
Metal -0.019 0.009 0.006 0.024
Metal -0.011 0.016 -0.005 -0.006
Metal 0.005 0.005 -0.012 0.000
Metal -0.001 0.000 -0.002 0.008
Metal -0.005 -0.007 -0.004 0.011

173
Metal 0.010 0.006 0.008 0.000
Metal 0.005 0.004 -0.001 0.000
Metal -0.010 -0.010 0.008 0.002
Metal 0.003 0.002 0.008 -0.002
Metal -0.008 -0.006 0.013 0.011
Metal 0.003 0.002 -0.007 0.011
Metal -0.020 -0.018 0.012 0.015
Metal -0.010 -0.009 -0.006 -0.005
Metal -0.011 -0.013 0.018 0.013
Metal -0.042 -0.043 -0.001 -0.010
Metal 0.042 0.043 -0.021 -0.017
Metal 0.043 0.044 -0.021 -0.017
Metal -0.041 -0.043 -0.003 -0.008
Metal -0.049 -0.046 0.004 0.001
Metal 0.017 0.040 -0.009 0.002
Metal -0.020 -0.002 0.016 0.004
Metal 0.036 0.038 0.006 0.006
Metal -0.049 -0.047 -0.004 0.001
Metal -0.007 -0.005 0.023 0.027
Metal 0.026 0.045 -0.026 -0.024
Metal 0.028 -0.041 -0.014 0.009
Metal -0.016 0.004 -0.002 -0.007
Metal 0.040 0.039 -0.016 -0.019
Metal -0.001 -0.002 0.020 0.026
Metal -0.027 -0.029 0.000 -0.009
Metal 0.014 0.035 -0.015 -0.011
Metal -0.022 -0.002 0.010 0.010
Metal 0.029 0.030 -0.007 -0.008
Metal -0.031 -0.029 0.000 -0.009
Metal -0.030 -0.028 -0.008 -0.014
Metal 0.043 0.044 -0.021 -0.019
Metal 0.047 0.048 -0.026 -0.020
Metal -0.029 -0.029 -0.013 -0.011
Metal -0.013 -0.012 0.033 0.029
Metal -0.039 -0.038 -0.022 -0.015
Metal -0.012 -0.014 0.036 0.028
C 0.241 0.000 0.169 0.000
O -0.115 0.000 -0.112 0.000

174
e)

Phenol on Phenol on
Pt(111) Cu(111)
Atom Initial Final Initial Final
Metal -0.006 -0.006 0.009 0.009
Metal 0.002 -0.005 0.002 -0.002
Metal -0.005 -0.006 0.015 0.008
Metal -0.013 -0.011 -0.006 -0.014
Metal 0.005 0.014 -0.009 0.001
Metal 0.011 0.015 0.002 0.002
Metal -0.017 -0.012 0.004 -0.012
Metal -0.003 -0.003 -0.004 -0.008
Metal -0.016 0.012 0.017 0.022
Metal -0.101 0.013 -0.048 -0.016
Metal -0.007 0.012 0.021 0.022
Metal -0.006 -0.004 -0.006 -0.008
Metal 0.006 -0.003 -0.002 0.000
Metal -0.047 0.009 -0.010 0.009
Metal -0.047 0.009 -0.013 0.009
Metal 0.004 -0.003 -0.005 0.001
Metal 0.009 0.007 0.000 -0.001
Metal -0.002 0.006 -0.003 0.000
Metal 0.008 0.006 -0.002 -0.001
Metal -0.014 -0.009 0.010 0.012
Metal -0.013 -0.009 0.012 0.012
Metal -0.006 -0.004 0.014 0.015
Metal -0.062 -0.056 -0.006 -0.004
Metal -0.003 -0.004 0.014 0.015
Metal -0.035 -0.028 -0.013 -0.009
Metal 0.014 0.049 -0.003 -0.016
Metal -0.005 0.047 -0.043 -0.017
Metal -0.041 -0.029 -0.021 -0.010
Metal -0.054 -0.032 -0.013 0.001
Metal 0.009 0.040 0.001 0.006
Metal 0.024 -0.028 0.017 0.004
Metal -0.052 0.044 -0.067 0.006
Metal -0.048 -0.032 -0.013 0.001
Metal -0.004 -0.001 0.021 0.026
Metal 0.003 0.047 -0.035 -0.019
Metal 0.134 -0.017 0.023 -0.003
Metal 0.086 -0.016 0.026 -0.001
Metal 0.045 0.047 -0.034 -0.021
Metal -0.004 0.000 0.020 0.027
Metal -0.069 -0.050 -0.017 -0.008

175
Metal 0.004 0.054 -0.008 -0.010
Metal 0.032 -0.038 0.037 0.015
Metal 0.013 0.055 -0.006 -0.011
Metal -0.069 -0.050 -0.020 -0.008
Metal -0.051 -0.048 -0.021 -0.011
Metal 0.016 0.053 -0.041 -0.021
Metal 0.017 0.053 -0.040 -0.022
Metal -0.051 -0.050 -0.023 -0.012
Metal -0.010 -0.007 0.025 0.027
Metal -0.033 -0.029 -0.022 -0.014
Metal -0.007 -0.004 0.026 0.028
C -0.201 0.000 -0.188 0.000
C -0.032 0.000 -0.035 0.000
C -0.053 0.000 -0.102 0.000
C -0.071 0.000 -0.026 0.000
C -0.134 0.000 -0.219 0.000
C 0.336 0.000 0.329 0.000
O -0.406 0.000 -0.417 0.000
H 0.146 0.000 0.130 0.000
H 0.127 0.000 0.106 0.000
H 0.127 0.000 0.107 0.000
H 0.133 0.000 0.107 0.000
H 0.145 0.000 0.119 0.000
Ho 0.341 0.000 0.331 0.000

176
𝐴 𝐵
Table C.4: Values for the parameter 𝛿 in 𝑉𝑖𝑗 = (1 − 𝜆) [(𝑟 2 +𝛿𝜆)6
− (𝑟 2 +𝛿𝜆)3
] for
𝑖𝑗 𝑖𝑗

interactions used in this study. As recommended by Zacharias et al.59, the parameter δ was
chosen as the square of the vdW radius of the interacting atoms to allow for a smooth
transition between an atom present on the surface and filling the cavity after the molecule
is annihilated. Ow and Hw represent the oxygen and hydrogen of the TIP3P water model,
respectively. Ho in the phenol molecule indicates the hydrogen of the OH functional group.

Adsorbate Interacting term 𝛅(Å𝟐 )


C-Ow 12.468
O-Ow 9.678
Ethylene glycol
H-Ow 6.574
H-Hw 1.414
C-Ow 12.184
Carbon monoxide
O-Ow 9.831
C-Ow 11.223

O-Ow 9.672

H-Ow 7.756
Phenol
Ho-Ow 4.190

H-Hw 1.988

Ho-Hw 0.452

177
Figure C.2: Scaling of the Lennard-Jones potential interaction between the oxygen of the
CO molecule and the oxygen of the TIP3P water model, O-Ow. 𝜆 = 0 indicates the regular
LJ potential, increasing 𝜆 gradually decreases the interaction, and finally, 𝜆 = 1 leads to a
complete removal of the LJ potential as if there were no adsorbate on the surface.

178
𝑙𝑖𝑞→𝑔𝑎𝑠
Table C.5: Contributions to the aqueous-phase effect on the free energy of the low coverage desorption (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 =
𝑔𝑎𝑠→𝑙𝑖𝑞
−∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ) of the reactant and transition states of the O-H bond cleavage of ethylene glycol over a Cu(111) surface at 423 K (see
𝑔𝑎𝑠→𝑙𝑖𝑞
Figure 4.2a in the main text). ∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 also shows the overall aqueous-phase effect on the free energy of the low coverage adsorption
of the states computed with QM/MM-FEP scheme proposed in this study, as well as two well-known implicit solvation schemes: iSMS
and VASPsol. For the QM/MM-FEP calculations, the 95% confidence interval (based on limited water sampling) is given, assuming a
normal distribution. All numbers are in eV.

𝒍𝒊𝒒→𝒈𝒂𝒔 𝒈𝒂𝒔→𝒍𝒊𝒒
Contributions to ∆∆𝑮𝑨𝒅𝒔𝒐𝒓𝒃𝒂𝒕𝒆 in QM/MM-FEP scheme ∆∆𝑮𝑨𝒅𝒔𝒐𝒓𝒃𝒂𝒕𝒆

Remove
To OPT Remove II-to-III
Lennard- To OPT site QM/MM- VASPs
Adsorbate adsorbate in Electrostatic (Electrostatic iSMS
Jones in Gas FEP ol
Liquid potential Embedding)
potential
Reactant
179

0.07 ± 0.00 0.49 ± 0.00 -0.47 ± 0.02 -0.64 ± 0.00 -0.05 ± 0.00 0.60 ± 0.03 0.02 -0.15
state
Transition
0.02 ± 0.00 1.25 ± 0.00 -0.68 ± 0.05 -0.56 ± 0.01 -0.11 ± 0.00 0.10 ± 0.05 0.00 -0.21
state
Table C.6: Net charge on adsorbate and number of sites occupied by adsorbate (#Site) for all adsorbates studied in this work. These two
𝑦−𝑦̅ ̅̅̅
𝑓1 −𝑓 1 ̅̅̅
𝑓2 −𝑓 2
descriptors were then used in a linear model, = 𝛼1 + 𝛼2 (𝑠 is standard deviation and bar sign shows mean), as 𝑓1 and 𝑓2
𝑠𝑦 𝑠 𝑓1 𝑠𝑓2
𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔−𝑙 𝑔𝑎𝑠
respectively, for estimating aqueous-phase effects on the free energies of adsorption (∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 = ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 − ∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 ),
computed by the proposed scheme in this work (QM/MM-FEP). 𝛼1 and 𝛼2 are 0.45 and 0.62, respectively, which indicates almost equal
importance of net charge and the size of the adsorbate.

∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒 − ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
𝑔𝑎𝑠→𝑙𝑖𝑞 𝑔𝑎𝑠→𝑙𝑖𝑞
∆∆𝐺𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒
Net charge on 𝑠∆∆𝐺𝑔𝑎𝑠→𝑙𝑖𝑞
Adsorbate / Surface / T(K) #Site 𝐴𝑑𝑠𝑜𝑟𝑏𝑎𝑡𝑒
adsorbate, e
QM/MM-FEP Fit |Error|

EG / Cu(111) / 423 0.02 2 -0.61 -0.61 0.00


180

EG / Pt(111) / 423 0.18 2 0.18 -0.16 0.34

CO / Cu(111) / 298 0.06 1 -1.11 -0.95 0.16

CO / Pt(111) / 298 0.13 1 -0.77 -0.75 0.02

Phenol / Cu(111) / 298 0.24 4 0.93 0.93 0.00

Phenol / Pt(111) / 298 0.46 4 1.38 1.54 0.16


Figure C.3: Average (of three independent simulations) free energy profile of the
desorption of a CO molecule from a Pt(111) surface model at 298 K using the natural
𝑙𝑖𝑞→𝑔𝑎𝑠
population analysis (NPA)36 charge model. ∆∆𝐺 (∆∆𝐺𝐶𝑂 ) is the aqueous-phase effect
on the low coverage desorption free energy of CO.

181
Figure C.4: Vapor-phase optimized geometries of adsorbed species studied in this work: a)
ethylene glycol (EG) over a Cu(111) surface model, b) transition state of O-H bond
cleavage of EG over the Cu(111) surface model, c) EG over a Pt(111) surface model, d)
CO over the Cu(111) surface model, e) CO over the Pt(111) surface model, f) phenol over
the Cu(111) surface model, and g) phenol over the Pt(111) surface model.

182
Figure C.5: Fits of equation 4.9 in the main text (adsorption isotherm) to the experimental
data of aqueous-phase adsorption of phenol at 298 K on a Pt(111) catalyst surface from
Singh et al.60, 61 for two different regions: low and high phenol concentrations. The “Const”
parameter represents the hydrogen surface coverage relative to the free site coverage
(const=0 shows zero H coverage approximation), 𝑛 represents the number of sites a phenol
𝑙𝑖𝑞
molecule occupies when adsorbed on the surface, 𝛼 (𝛼𝑃ℎ,𝑃ℎ ) is a constant associated to
𝑔𝑎𝑠→𝑙𝑖𝑞
phenol-phenol lateral interactions on the surface in water, and ∆∆𝐺 (∆∆𝐺𝑃ℎ (𝜃𝑃ℎ =
0)) is the solvent effect on the aqueous-phase low coverage phenol adsorption on the
surface. SSR is the sum of squared residuals, which shows the error associated with each
fit. Fits with higher 𝑛 predict the lower concentration region better, while fits with smaller
n predict the higher concentration region better. However, the fits suggest that
𝑔𝑎𝑠→𝑙𝑖𝑞
∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0) is quite independent of these two fit regions.

Figure C.6: Adsorbed phenol orientations at different coverages in the gas phase. At low
coverage, phenols adsorb horizontally parallel to the surface, while at high coverage, they
adsorb slanted to occupy less space on the surface.

183
Figure C.7: Fits of equation 4.9 in the main text (adsorption isotherm) to the experimental
data of aqueous-phase phenol adsorption on a Pt(111) catalyst surface at 283 K from Singh
et al.62 The “Const” parameter represents the hydrogen surface coverage relative to the free
site coverage (const=0 shows the zero H coverage approximation), 𝑛 represents the number
𝑙𝑖𝑞
of sites a phenol molecule occupies when adsorbed on the surface, 𝛼 (𝛼𝑃ℎ,𝑃ℎ ) is a constant
associated to phenol/phenol lateral interactions on the surface in water, and ∆∆𝐺
𝑔𝑎𝑠→𝑙𝑖𝑞
(∆∆𝐺𝑃ℎ (𝜃𝑃ℎ = 0)) is the aqueous-phase effect on the low coverage phenol adsorption
free energy on the surface. SSR is the sum of squared residuals, which shows the error
associated with each fit.

184
Figure C.8: Fits of equation 4.9 in the main text (adsorption isotherm) to the experimental
data of aqueous-phase phenol adsorption on a Pt(111) catalyst surface at 288 K from Singh
et al.62 See the caption of Figure C.7 for detailed explanations.

185
Figure C.9: Fits of equation 4.9 in the main text (adsorption isotherm) to the experimental
data of aqueous-phase phenol adsorption on a Pt(111) catalyst surface at 314 K from Singh
et al.62 See the caption of Figure C.7 for detailed explanations.

186
C.7 Bibliography
1. Kresse G, Furthmuller J. Efficiency of ab-initio total energy calculations for
metals and semiconductors using a plane-wave basis set. Comp Mater Sci 6, 15-
50 (1996).

2. Kresse G, Furthmuller J. Efficient iterative schemes for ab initio total-energy


calculations using a plane-wave basis set. Phys Rev B 54, 11169-11186 (1996).

3. Blöchl PE. Projector augmented-wave method. Phys Rev B 50, 17953-17979


(1994).

4. Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made


simple. Phys Rev Lett 77, 3865-3868 (1996).

5. Perdew JP, Yue W. Accurate and Simple Density Functional for the Electronic
Exchange Energy - Generalized Gradient Approximation. Phys Rev B 33, 8800-
8802 (1986).

6. Perdew JP, Wang Y. Accurate and Simple Analytic Representation of the


Electron-Gas Correlation-Energy. Phys Rev B 45, 13244-13249 (1992).

7. Grimme S, Antony J, Ehrlich S, Krieg H. A consistent and accurate ab initio


parametrization of density functional dispersion correction (DFT-D) for the 94
elements H-Pu. The Journal of Chemical Physics 132, 154104 (2010).

8. Monkhorst HJ, Pack JD. Special Points for Brillouin-Zone Integrations. Phys Rev
B 13, 5188-5192 (1976).

9. Methfessel M, Paxton AT. High-precision sampling for Brillouin-zone integration


in metals. Phys Rev B 40, 3616-3621 (1989).

10. Makov G, Payne MC. Periodic Boundary-Conditions in Ab-Initio Calculations.


Phys Rev B 51, 4014-4022 (1995).

11. Henkelman G, Jonsson H. Improved tangent estimate in the nudged elastic band
method for finding minimum energy paths and saddle points. J Chem Phys 113,
9978-9985 (2000).

187
12. Henkelman G, Uberuaga BP, Jonsson H. A climbing image nudged elastic band
method for finding saddle points and minimum energy paths. J Chem Phys 113,
9901-9904 (2000).

13. Henkelman G, Jonsson H. A dimer method for finding saddle points on high
dimensional potential surfaces using only first derivatives. J Chem Phys 111,
7010-7022 (1999).

14. Heyden A, Bell AT, Keil FJ. Efficient methods for finding transition states in
chemical reactions: Comparison of improved dimer method and partitioned
rational function optimization method. J Chem Phys 123, 224101:224101-224114
(2005).

15. Ahlrichs R, Bar M, Haser M, Horn H, Kolmel C. Electronic-Structure


Calculations on Workstation Computers - the Program System Turbomole. Chem
Phys Lett 162, 165-169 (1989).

16. Treutler O, Ahlrichs R. Efficient Molecular Numerical-Integration Schemes. J


Chem Phys 102, 346-354 (1995).

17. Von Arnim M, Ahlrichs R. Performance of parallel TURBOMOLE for density


functional calculations. J Comput Chem 19, 1746-1757 (1998).

18. Faheem M, Heyden A. Hybrid Quantum Mechanics/Molecular Mechanics


Solvation Scheme for Computing Free Energies of Reactions at Metal-Water
Interfaces. Journal of Chemical Theory and Computation 10, 3354-3368 (2014).

19. Schafer A, Horn H, Ahlrichs R. Fully Optimized Contracted Gaussian-Basis Sets


for Atoms Li to Kr. J Chem Phys 97, 2571-2577 (1992).

20. Weigend F, Ahlrichs R. Balanced basis sets of split valence, triple zeta valence
and quadruple zeta valence quality for H to Rn: Design and assessment of
accuracy. Physical Chemistry Chemical Physics 7, 3297-3305 (2005).

21. Russo TV, Martin RL, Hay PJ. Effective Core Potentials for Dft Calculations. J
Phys Chem-Us 99, 17085-17087 (1995).

22. Eichkorn K, Weigend F, Treutler O, Ahlrichs R. Auxiliary basis sets for main row
atoms and transition metals and their use to approximate Coulomb potentials.
Theor Chem Acc 97, 119-124 (1997).

188
23. Weigend F. Accurate Coulomb-fitting basis sets for H to Rn. Physical Chemistry
Chemical Physics 8, 1057-1065 (2006).

24. Todorov IT, Smith W, Trachenko K, Dove MT. DL_POLY_3: new dimensions in
molecular dynamics simulations via massive parallelism. J Mater Chem 16, 1911-
1918 (2006).

25. Natarajan SK, Behler J. Neural network molecular dynamics simulations of solid-
liquid interfaces: water at low-index copper surfaces. Physical Chemistry
Chemical Physics 18, 28704-28725 (2016).

26. Jorgensen WL, Chandrasekhar J, Madura JD, Impey RW, Klein ML. Comparison
of Simple Potential Functions for Simulating Liquid Water. J Chem Phys 79, 926-
935 (1983).

27. Andersen HC. Rattle - a Velocity Version of the Shake Algorithm for Molecular-
Dynamics Calculations. J Comput Phys 52, 24-34 (1983).

28. Smith W, Forester TR. Parallel Macromolecular Simulations and the Replicated
Data Strategy .2. The Rd-Shake Algorithm. Comput Phys Commun 79, 63-77
(1994).

29. Swope WC, Andersen HC, Berens PH, Wilson KR. A Computer-Simulation
Method for the Calculation of Equilibrium-Constants for the Formation of
Physical Clusters of Molecules - Application to Small Water Clusters. J Chem
Phys 76, 637-649 (1982).

30. Jorgensen WL. Optimized Intermolecular Potential Functions for Liquid


Alcohols. J Phys Chem-Us 90, 1276-1284 (1986).

31. Geerke DP, Van Gunsteren WF. The performance of non-polarizable and
polarizable force-field parameter sets for ethylene glycol in molecular dynamics
simulations of the pure liquid and its aqueous mixtures. Mol Phys 105, 1861-1881
(2007).

32. Jorgensen WL, Laird ER, Nguyen TB, Tirado-Rives J. Monte Carlo simulations
of pure liquid substituted benzenes with OPLS potential functions. Journal of
Computational Chemistry 14, 206-215 (1993).

189
33. Straub JE, Karplus M. Molecular dynamics study of the photodissociation of
carbon monoxide from myoglobin: Ligand dynamics in the first 10 ps. Chemical
Physics 158, 221-248 (1991).

34. Freindorf M, Shao Y, Furlani TR, Kong J. Lennard–Jones parameters for the
combined QM/MM method using the B3LYP/6-31G*/AMBER potential. Journal
of Computational Chemistry 26, 1270-1278 (2005).

35. Heinz H, Vaia RA, Farmer BL, Naik RR. Accurate Simulation of Surfaces and
Interfaces of Face-Centered Cubic Metals Using 12-6 and 9-6 Lennard-Jones
Potentials. J Phys Chem C 112, 17281-17290 (2008).

36. Reed AE, Weinstock RB, Weinhold F. Natural-Population Analysis. J Chem Phys
83, 735-746 (1985).

37. Manz TA, Limas NG. Introducing DDEC6 atomic population analysis: part 1.
Charge partitioning theory and methodology. RSC Advances 6, 47771-47801
(2016).

38. Burow AM, Sierka M, Dobler J, Sauer J. Point defects in CaF2 and CeO2
investigated by the periodic electrostatic embedded cluster method. J Chem Phys
130, 174710:174711-174711 (2009).

39. Nose S. A Unified Formulation of the Constant Temperature Molecular-


Dynamics Methods. J Chem Phys 81, 511-519 (1984).

40. Hoover WG. Canonical Dynamics - Equilibrium Phase-Space Distributions. Phys


Rev A 31, 1695-1697 (1985).

41. Essmann U, Perera L, Berkowitz ML, Darden T, Lee H, Pedersen LG. A Smooth
Particle Mesh Ewald Method. J Chem Phys 103, 8577-8593 (1995).

42. Hu H, Yang WT. Development and application of ab initio QM/MM methods for
mechanistic simulation of reactions in solution and in enzymes. J Mol Struc-
Theochem 898, 17-30 (2009).

43. Ryde U. How Many Conformations Need To Be Sampled To Obtain Converged


QM/MM Energies? The Curse of Exponential Averaging. Journal of Chemical
Theory and Computation 13, 5745-5752 (2017).

190
44. Pal S, Chakraborty K, Khatua P, Bandyopadhyay S. Microscopic dynamics of
water around unfolded structures of barstar at room temperature. The Journal of
Chemical Physics 142, 055102 (2015).

45. Hu H, Lu ZY, Yang WT. QM/MM minimum free-energy path: Methodology and
application to triosephosphate isomerase. Journal of Chemical Theory and
Computation 3, 390-406 (2007).

46. Hu H, Lu ZY, Parks JM, Burger SK, Yang WT. Quantum mechanics/molecular
mechanics minimum free-energy path for accurate reaction energetics in solution
and enzymes: Sequential sampling and optimization on the potential of mean
force surface. J Chem Phys 128, 034105:034101-034118 (2008).

47. Saleheen M, Heyden A. Liquid-Phase Modeling in Heterogeneous Catalysis. ACS


Catalysis 8, 2188-2194 (2018).

48. Saleheen M, Zare M, Faheem M, Heyden A. Computational Investigation of


Aqueous Phase Effects on the Dehydrogenation and Dehydroxylation of Polyols
over Pt(111). The Journal of Physical Chemistry C 123, 19052-19065 (2019).

49. Bennett C. Efficient Estimation of Free Energy Differences from Monte Carlo
Data. J Comput Phys 22, 245-268 (1976).

50. Shirts MR, Bair E, Hooker G, Pande VS. Equilibrium Free Energies from
Nonequilibrium Measurements Using Maximum-Likelihood Methods. Phys Rev
Lett 91, 140601 (2003).

51. James F. A review of pseudorandom number generators. Comput Phys Commun


60, 329-344 (1990).

52. Kreyszig E. Advanced Engineering Mathematics, 9th edn. John Wiley & Sons,
Inc. (2006).

53. Faheem M, Suthirakun S, Heyden A. New Implicit Solvation Scheme for Solid
Surfaces. J Phys Chem C 116, 22458-22462 (2012).

54. Fishman M, Zhuang HLL, Mathew K, Dirschka W, Hennig RG. Accuracy of


exchange-correlation functionals and effect of solvation on the surface energy of
copper. Phys Rev B 87, (2013).

191
55. Mathew K, Sundararaman R, Letchworth-Weaver K, Arias TA, Hennig RG.
Implicit solvation model for density-functional study of nanocrystal surfaces and
reaction pathways. J Chem Phys 140, 084106:084101-084108 (2014).

56. Fernandez DP, Goodwin ARH, Lemmon EW, Sengers JMHL, Williams RC. A
formulation for the static permittivity of water and steam at temperatures from
238 K to 873 K at pressures up to 1200 MPa, including derivatives and Debye-
Huckel coefficients. J Phys Chem Ref Data 26, 1125-1166 (1997).

57. Nitta T, Shigetomi T, Kuro-Oka M, Katayama T. AN ADSORPTION


ISOTHERM OF MULTI-SITE OCCUPANCY MODEL FOR HOMOGENEOUS
SURFACE. Journal of Chemical Engineering of Japan 17, 39-45 (1984).

58. Nitta T, Kuro-Oka M, Katayama T. AN ADSORPTION ISOTHERM OF


MULTI-SITE OCCUPANCY MODEL FOR HETEROGENEOUS SURFACE.
Journal of Chemical Engineering of Japan 17, 45-52 (1984).

59. Zacharias M, Straatsma TP, McCammon JA. Separation‐shifted scaling, a new


scaling method for Lennard‐Jones interactions in thermodynamic integration. The
Journal of Chemical Physics 100, 9025-9031 (1994).

60. Singh N, Sanyal U, Fulton JL, Gutiérrez OY, Lercher JA, Campbell CT.
Quantifying Adsorption of Organic Molecules on Platinum in Aqueous Phase by
Hydrogen Site Blocking and in Situ X-ray Absorption Spectroscopy. ACS
Catalysis 9, 6869-6881 (2019).

61. Akinola J, Barth I, Goldsmith BR, Singh N. Adsorption Energies of Oxygenated


Aromatics and Organics on Rhodium and Platinum in Aqueous Phase. ACS
Catalysis 10, 4929-4941 (2020).

62. Akinola J, Singh N. Temperature dependence of aqueous-phase phenol adsorption


on Pt and Rh. Journal of Applied Electrochemistry, (2020).

192
APPENDIX D:

COPYRIGHT PERMISSIONS

193

You might also like