conversion chenlu xie

90
Rational Design and Synthesis of Inorganic Nanostructures for Tandem Catalysis and CO2 Conversion by Chenlu Xie A dissertation submitted in partial satisfaction of the requirements for the degree of Doctor of Philosophy in Chemistry in the Graduate Division of the University of California, Berkeley Committee in charge: Professor Peidong Yang, Chair Professor Gabor A. Somorjai Professor David B. Graves Spring 2018

Upload: others

Post on 23-Mar-2022

11 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Conversion Chenlu Xie

Rational Design and Synthesis of Inorganic Nanostructures for Tandem Catalysis

and CO2 Conversion

by

Chenlu Xie

A dissertation submitted in partial satisfaction of the

requirements for the degree of

Doctor of Philosophy

in

Chemistry

in the

Graduate Division

of the

University of California, Berkeley

Committee in charge:

Professor Peidong Yang, Chair

Professor Gabor A. Somorjai

Professor David B. Graves

Spring 2018

Page 2: Conversion Chenlu Xie

Rational Design and Synthesis of Inorganic Nanostructures for Tandem Catalysis

and CO2 Conversion

© Copyright 2018

by

Chenlu Xie

Page 3: Conversion Chenlu Xie

1

Abstract

Rational Design and Synthesis of Inorganic Nanostructures for Tandem Catalysis

and CO2 Conversion

by

Chenlu Xie

Doctor of Philosophy in Chemistry

University of California, Berkeley

Professor Peidong Yang, Chair

The subject of this dissertation focuses on the design and synthesis of new catalysts with well-

defined structures and superior performance to meet the new challenges in heterogenous catalysis.

The past decade has witness the development of nanoscience as well as the inorganic catalysts for

industrial applications, however there are still fundamental challenges and practical need for

catalysis. Specifically, it is desirable to have the ability to selectivity produce complex molecules

from simple components. Another great challenge faced by the modern industry is being

environmentally friendly, and going for a carbon neutral economy would require using CO2 as

feedstock to produce valuable products. The work herein focuses on the design and synthesis of

inorganic nanocrystal catalysts that address these challenges by achieving selective and sequential

chemical reactions and conversion of CO2 to valuable products.

Chapter 1 introduces the development of heterogenous catalysis and the colloidal synthesis of

metal nanoparticles catalysts with well-controlled structure. Tremendous efforts have been

devoted to understanding the nucleation and growth process in the colloidal synthesis and

developing new methods to produce metal nanoparticles with controlled sizes, shapes, composition.

These well-defined catalytic system shows promising catalytic performance, which can be

modulated by their structure (size, shape, compositions and the metal-oxide interfaces). The

chapters hereafter explore the synthesis of new catalysts with controlled structures for catalysis.

Chapter 2 presents the design and synthesis of a three dimensional (3D) nanostructured

catalysts CeO2-Pt@mSiO2 with dual metal-oxide interfaces to study the tandem hydroformylation

reaction in gas phase, where CO and H2 produced by methanol decomposition (catalyzed by CeO2-

Pt interface) were reacted with ethylene to selectively yield propyl aldehyde (catalyzed by Pt-SiO2

interface). With the stable core-shell architecture and well-defined metal-oxide interfaces, the

origin of the high propyl aldehyde selectivity over ethane, the dominant byproduct in conventional

hydroformylation, was revealed by in-depth mechanism study and attributed to the synergy

Page 4: Conversion Chenlu Xie

2

between the two sequential reactions and the altered elementary reaction steps of the tandem

reaction compared to the single-step reaction. The effective production of aldehyde through the

tandem hydroformylation was also observed on other light olefin system, such as propylene and

1-butene.

Chapter 3 expands the strategy of tandem catalysis into conversion of CO2 with hydrogen to

value-added C2-C4 hydrocarbons, which is a major pursuit in clean energy research. Another well-

defined 3D catalyst CeO2–Pt@mSiO2–Co was designed and synthesized, and CO2 was converted

to C2-C4 hydrocarbons with 60% selectivity on this catalyst via reverse water gas shift reaction

and subsequent Fischer–Tropsch process. In addition, the catalysts is stable and shows no obvious

deactivation over 40 h. The successful production of C2−C4 hydrocarbons via a tandem process on

a rationally designed, structurally well-defined catalyst demonstrates the power of sophisticated

structure control in designing nanostructured catalysts for multiple-step chemical conversions.

Chapter 4 turns to electrochemistry and apply the precision in catalyst structural design to the

development of electrocatalysts for CO2 reduction. Herein, atomic ordering of bimetallic

nanoparticles were synthetically tuned, from disordered alloy to ordered intermetallic, and it

showed that this atomic level control over nanocrystal catalysts could give significant performance

benefits in electrochemical CO2 reduction to CO. Atomic-level structural investigations revealed

the atomic gold layers over the intermetallic core to be sufficient for enhanced catalytic behavior,

which is further supported by DFT analysis.

Page 5: Conversion Chenlu Xie

i

To my parents and my fiancé

Page 6: Conversion Chenlu Xie

ii

Table of Contents

List of Figures................................................................................................................................. v

List of Tables and Schemes .........................................................................................................viii

Acknowledgement .........................................................................................................................ix

Vita................................................................................................................................................. x

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

Section 1.1 Heterogenous Catalysis in Modern Industry............................................................ 1

Section 1.2 Colloidal Nanocrystal Synthesis with Controlled Structure .................................... 1

Section 1.2.1 Basic Principle in Monodispersed Nanocrystal Synthesis ................................ 1

Section 1.2.2 Synthesis of Nanocrystals with Controlled Size, Shape and Composition ....... 3

Section 1.3 Synthesis of Supported Metal Catalysts ................................................................. 6

Section 1.3.1 Conventional Synthetic Method: Precipitation and Impregnation.................... 6

Section 1.3.2 Emerging Synthetic Method: Preparation of Catalysts with Well-controlled

Structure and Interface ............................................................................................................ 7

Section 1.4 Catalytic Reaction Studies Using Nanoparticle Catalysts ....................................... 9

Section 1.5 References .............................................................................................................. 11

Chapter 2. Tandem Alkene Hydroformylation over a Three Dimensional Nanostructured

Catalyst with Multiple Interfaces .............................................................................................. 15

Section 2.1 Introduction ............................................................................................................ 15

Section 2.2 Experimental Section ............................................................................................. 16

Section 2.2.1 Methods and Materials .................................................................................... 16

Section 2.2.2 Synthesis of 3D tandem catalysts ................................................................... 17

Section 2.3 Results and Discussion .......................................................................................... 19

Section 2.3.1 Synthesis and Characterizations of 3D Tandem Catalysts. ............................ 19

Section 2.3.2 Catalytic Performance of CeO2−Pt@mSiO2 in the Hydroformylation of

Ethylene with Methanol ........................................................................................................ 21

Page 7: Conversion Chenlu Xie

iii

Section 2.4 Conclusion ............................................................................................................. 29

Section 2.5 References .............................................................................................................. 29

Section 2.6 Appendix ................................................................................................................ 31

Section 2.6.1 Stability of Tandem Catalyst CeO2-Pt@miSO2 .............................................. 31

Section 2.6.2 TEM Images of Single Interface Catalysts. .................................................... 32

Section 2.6.3 Detailed Data of Conversion and Selectivity of Each Step in the Tandem

Process and the Single Tandem Hydroformylation .............................................................. 32

Chapter 3. Tandem Catalysis for CO2 Hydrogenation to C2−C4 Hydrocarbons.................. 35

Section 3.1 Introduction ............................................................................................................ 35

Section 3.2 Experimental Section ............................................................................................. 35

Section 3.2.1 Synthetic Methods and Materials.................................................................... 35

Section 3.2.2 Characterization and Catalytic Study of CeO2-Pt@mSiO2-Co Tandem

Catalysts ................................................................................................................................ 36

Section 3.3 Results and Discussion .......................................................................................... 37

Section 3.3.1 Structure Design and Synthesis of the CeO2-Pt@mSiO2-Co Tandem Catalysts

............................................................................................................................................... 37

Section 3.3.2 Catalytic Performance for CO2 hydrogenation ............................................... 42

Section 3.4 Conclusion ............................................................................................................. 46

Section 3.5 References .............................................................................................................. 46

Charpter 4. Controlling Atomic Ordering of AuCu Nanoparticles for Electrochemical CO2

Reduction ..................................................................................................................................... 49

Section 4.1 Introduction ............................................................................................................ 49

Section 4.2 Experimental Section ............................................................................................. 49

Section 4.2.1 Synthetic Methods and Materials.................................................................... 49

Section 4.2.2 XRD and Electron Microscopic Characterization of AuCu Nanoparticles .... 50

Section 4.2.3 Spectroscopic Characterization of AuCu Nanoparticles................................. 50

Section 4.2.4 Electrochemical Study of AuCu Nanoparticles .............................................. 51

Section 4.2.5 Estimation of the Long-range Order Parameter (S) ........................................ 51

Page 8: Conversion Chenlu Xie

iv

Section 4.2.6 XPS and EXAFS Study of AuCu Nanoparticles ............................................ 52

Section 4.2.7 Computational Methods .................................................................................. 53

Section 4.3 Result and Discussion ............................................................................................ 54

Section 4.3.1 Control of AuCu NPs with Different Atomic Ordering.................................. 54

Section 4.3.2 AuCu NPs with Different Ordering Degree for Electrochemical CO2

Reduction .............................................................................................................................. 57

Section 4.3.3 Further Structural Investigation at both Nanoscale and Atomic Scale ........... 61

Section 4.3.4 Extended X-ray Absorption Fine Structure (EXAFS): Coordination

Environment Surrounding Au and Cu .................................................................................. 63

Section 4.3.5 Further Comparison with Pure Au NPs on CO2 Reduction ........................... 67

Section 4.3.6 Density Function Theory (DFT) Calculations: Insights into the Origin of

Enhanced CO2 Reduction Activity ....................................................................................... 70

Section 4.4 Conclusion ............................................................................................................. 73

Section 4.5 References .............................................................................................................. 73

Page 9: Conversion Chenlu Xie

v

List of Figures

Chapter 1

Figure 1.1. The LaMer model for nucleation and growth............................................................... 2

Figure 1.2. Synthesis of noble with controlled size. ....................................................................... 4

Figure 1.3. Synthesis of transition metal nanoparticles with different size .................................... 4

Figure 1.4. Pt nanoparticles with different shape of cubes, cubotahedra and octahedra using silver

ions as modulator. ........................................................................................................................... 5

Figure 1.5. Preparation of CeO2-Pt-SiO2 catalysts by Layer-by-layer deposition method ............. 8

Figure 1.6. Synthesis of metal catalysts by directly coating of porous silica shell around the metal

nanoparticles. .................................................................................................................................. 9

Figure 1.7. Catalytic activity and selectivity depends on the metal-oxide interfaces. .................. 10

Chapter 2

Figure 2.1. The schematic of the reactor....................................................................................... 17

Figure 2.2. Synthesis and characterization of the 3D nanocrystal tandem catalyst CeO2-

Pt@mSiO2. .................................................................................................................................... 19

Figure 2.3. Elemental distribution of the 3D tandem catalyst. ..................................................... 20

Figure 2.4. Porosity characterization of CeO2−Pt@mSiO2 nanoparticles calcined at 350 °C...... 20

Figure 2.5. Ethylene hydroformylation with methanol carried out at 150 °C over CeO2-

Pt@mSiO2 and single interface catalysts CeO2-Pt and Pt@mSiO2. ............................................. 22

Figure 2.6. Methanol decomposition over 3D catalyst CeO2−Pt@mSiO2 and single interface

catalysts: CeO2−Pt and Pt@mSiO2. .............................................................................................. 22

Figure 2.7. Catalytic performance of CeO2-Pt@mSiO2 for the hydroformylation of ethylene with

methanol at an ethylene pressure of 7.5 torr. ................................................................................ 23

Figure 2.8. The inhibiting effect of ethylene on methanol decomposition over the CeO2-

Pt@mSiO2 catalyst ........................................................................................................................ 24

Figure 2.9. H2 and CO effect on the tandem hydroformylation of ethylene. ................................ 26

Figure 2.10. Proposed reaction pathway for the tandem ethylene hydroformylation. .................. 28

Page 10: Conversion Chenlu Xie

vi

Chapter 3

Figure 3.1. Synthesis and characterization of the CeO2-Pt@mSiO2-Co tandem catalyst............. 38

Figure 3.2. Tuning the size of CeO2 NPs by changing the ethanol and water ratio. .................... 39

Figure 3.3. TEM images of CeO2-Pt NPs with Pt loading amount around .................................. 40

Figure 3.4. Silica shell thickness tuning of CeO2-Pt@mSiO2 NPs by changing the added volume

of TEOS. ....................................................................................................................................... 40

Figure 3.5. TEM image of monodisperse cobalt NPs. .................................................................. 40

Figure 3.6. Elemental distribution of the catalyst. ........................................................................ 41

Figure 3.7. Catalytic reaction study on the single-interface catalysts, tandem catalyst and the

physical mixture. ........................................................................................................................... 42

Figure 3.8. Catalytic performance of CeO2-Pt@mSiO2 under different temperatures. ................ 43

Figure 3.9. Catalytic performance of CeO2@mSiO2-Co under different temperatures. ............... 43

Figure 3.10. Catalytic performance of tandem catalyst CeO2-Pt@mSiO2-Co under different

temperatures. ................................................................................................................................. 44

Figure 3.11. Catalytic performance of physical mixture of CeO2-Pt@mSiO2 and CeO2@mSiO2-

Co under different temperatures. .................................................................................................. 45

Figure 3.12. Stability test of CeO2-Pt@mSiO2-Co catalyst ......................................................... 45

Figure 3.13. Stability test of CO selectivity over the tandem catalyst CeO2-Pt@mSiO2-Co ....... 46

Figure 3.14. TEM images of tandem catalyst CeO2-Pt@mSiO2-Co after reaction. ..................... 46

Chapter 4

Figure 4.1. TEM images of Au NP seeds (size 6 nm) .................................................................. 55

Figure 4.2. Scheme of synthesis process of AuCu NPs with different ordering degree. .............. 55

Figure 4.3. TEM images of AuCu bimetallic nanoparticles synthesized at various conditions

allowing for systematic tuning of the degree of ordering. ............................................................ 56

Figure 4.4. Size distribution and Au:Cu ratio of AuCu nanoparitcles with different ordering .... 56

Figure 4.5. XRD of AuCu bimetallic nanoparticles. .................................................................... 57

Figure 4.6. Representative SEM images of carbon paper electrodes decorated with AuCu NPs. 58

Figure 4.7. Ligand mass fraction of AuCu NPs before electrochemical testing........................... 58

Figure 4.8. Electrochemical CO2 study on the AuCu nanoparticles. ............................................ 59

Page 11: Conversion Chenlu Xie

vii

Figure 4.9. Formate FE of AuCu NPs at -0.77 V vs. RHE measured by qNMR. ........................ 59

Figure 4.10. CO FE at various potentials for d-AuCu and o-AuCu NP catalysts......................... 60

Figure 4.11. Elemental maps of Au and Cu for d-AuCu and o-AuCu NPs. ................................. 60

Figure 4.12. Structural investigation of o-AuCu at atomic resolution. ......................................... 61

Figure 4.13. Aberration-corrected HAADF-STEM images of o-AuCu NPs ................................ 62

Figure 4.14. Aberration-corrected HAADF-STEM images of d-AuCu NPs. ............................... 62

Figure 4.15. X-ray absorption spectra at the Au L3-edge and Cu K-edge collected from o-AuCu

NPs and d-AuCu NPs.................................................................................................................... 63

Figure 4.16. X-ray absorption spectroscopy of AuCu NPs. .......................................................... 64

Figure 4.17. Aberration-corrected HAADF-STEM images of d-AuCu NPs exhibiting short-range

order. ............................................................................................................................................. 65

Figure 4.18. Au 4f and Cu 2p XPS spectra of AuCu NPs. ........................................................... 66

Figure 4.19. UV-Vis absorption spectra of AuCu NPs. ................................................................ 66

Figure 4.20. Au NPs observed from TEM and their size distribution. ......................................... 67

Figure 4.21. Cu underpotential deposition performed on o-AuCu NP and Au NP electrodes in

0.5M H2SO4 solution with 0.1M CuSO4 and Ar purge................................................................. 68

Figure 4.22. Electrocatalytic CO2 reduction activity of AuCu NPs. ............................................. 69

Figure 4.23. CO specific current and CO production rate of o-AuCu NPs and d-AuCu NPs. ..... 69

Figure 4.24. CO partial current densities and CO mass activities of d-AuCu NP, o-AuCu NP, and

Au NP for electrochemical reduction of CO2 at -0.85 V vs. RHE. ............................................... 70

Figure 4.25. Stability test of o-AuCu NPs under extended operation of 12 hours for

electrochemical reduction of CO2 ................................................................................................. 70

Figure 4.26. Computational results of CO2 reduction on AuCu surfaces. .................................... 71

Figure 4.27. Surface energy of mixed Au/Cu 211 slabs as a function of Au density at the surface.

....................................................................................................................................................... 72

Figure 4.28. d-projected density of states for the model systems of d-AuCu, o-AuCu, and Au. . 73

Page 12: Conversion Chenlu Xie

viii

List of Tables

Chapter 2

Table 2.1. Pt loading amount for the tandem catalyst CeO2−Pt@mSiO2 and single interface

catalysts according to ICP-AES. ................................................................................................... 21

Table 2.2. Tandem hydroformylation of different alkenes. ........................................................... 27

Table 2.3. The inhibiting effect of different alkenes on methanol decomposition over

CeO2−Pt@mSiO2 catalyst ............................................................................................................. 27

Chapter 3

Table 3.1. Pt and Co loading amount for the catalysts used for CO2 hydrogenation according to

ICP-AES. ...................................................................................................................................... 41

Chapter 4

Table 4.1. EXAFS fitting parameters for the fit to the o-AuCu NP displayed in Fig. 4.16a of the

main text........................................................................................................................................ 64

Table 4.2. EXAFS fitting parameters for the fit to the o-AuCu NP displayed in Fig. 4.16b of the

main text........................................................................................................................................ 65

Page 13: Conversion Chenlu Xie

ix

Acknowledgement

I would first and foremost thank my advisor Prof. Peidong Yang for giving me this opportunity

to work in Berkeley and his lab. He is enormously hardworking, persistent and his hands-on and

effective mentoring throughout the last five years greatly accelerated my intellectual growth and

sharpened my mind. His mentorship and support, both direct and indirect, have fundamentally

shaped my perspective of research.

I am also deeply grateful to Prof. Gabor Somorjai, who I worked with for the first four years in

my Ph.D study. I collaborated extensively with his group on tandem catalysis with three

dimensional nanostructures with multiple interfaces. I learnt much of my knowledge in surface

science and gas phase catalysis from him, and his keen interest and passion in science at his eighties

are very inspiring and made him a role model for me. He is also my qualifying exam committee

member. I would also want to express my gratitude to Prof. Jinghua Guo at Advance Light Source,

who has been my long-term collaborator in synchrotron-based experiments. I would also like to

thank Prof. Omar Yaghi and Prof. William Lester for being my qualifying exam committee and

Prof. David Graves for being my qualifying exam outside committee.

It would not be possible for me to carry out the research projects presented herein without the

help of all my collaborators: Prof. Jens Nørskov, and Dr. Karen Chan at Stanford University, Dr.

Ethan Crumlin, Dr. Yi-de Chuang and Dr. Sirine Fakra in the Advanced Light Source, Dr. Ji Su in

Somorjai group, Dr. Karen Bustillo in the NCEM. I am also grateful for the helpful discussion

from Dr. Yifan Ye in Advanced Light Source, Dr. Selim Alayoglu and Mr. Wenchi Liu in the

Somorjai group.

The Yang lab members have always been helpful and supportive during my Ph.D study. I am

especially thankful to Dr. Chen Chen as my mentor when I joined the Yang group and started my

first project. He patiently taught me a lot on nanocrystal synthesis and characterization techniques.

Dr. Dachao Hong is an incredibly nice person and we had many helpful discussions and

collaborations in my early days. Mr. Dohyung Kim collaborated with me on the AuCu

intermetallic nanoparticles project. Dr. Yi Yu always helped me with the TEM characterization

and also taught me a lot. Dr. Nigel Becknell helped a lot on the experiments at Advanced Light

Source, and we have spent many day and night shifts for the experiments, together with Mr. Qiao

Kong and Ms. Hao Zhang. Dr. Zhiqiang Niu and Dr. Fan Cui are always there for helpful

discussion.

I owe my greatest gratitude to my parents, who raised me up and always stand behind me with

their everlasting love. Without their unconditionally support, both materially and spiritually, I

would not be who I am. It is my greatest fortune to meet Yingbo at Berkeley, who is my soul mate

and is becoming my family. He is extraordinary supportive, encouraging, understanding and

considerate, and his passion in his research, determined and positive attitude have deeply

influenced me. He shared all my sadness, happiness, failure and success during my Ph.D., my life

would be completely different without him.

Page 14: Conversion Chenlu Xie

x

Vita

November 1992 Born, Shaoyang, Hunan, China

2013 B.S., Chemistry

University of Science and Technology of China

Hefei. Anhui, China

2015 Ph.D. Candidate, Chemistry

University of California, Berkeley

Berkeley, California

2013-2018 Graduate Student Instructor and Researcher

University of California, Berkeley

Berkeley, California

PUBLICATIONS

Su, J., Xie, C., Chen, C., Yu, Y., Kennedy, G., Somorjai, G. A., Yang, P. Insights into the

Mechanism of Tandem Alkene Hydroformylation over Nanocrystal Catalyst with Multiple

Interfaces. J. Am. Chem. Soc. 2016, 138, 11568-11574.

Xie, C., Chen, C., Yu, Y., Su, J., Li, Y., Somorjai, G. A., Yang, P. Tandem Catalysis for CO2

Hydrogenation to C2-C4 Hydrocarbons. Nano Lett., 2017, 17, 3798-3802.

Kim, D., Xie, C., Becknell, N., Yu, Y., Karamad, M., Chan, K., Crumlin, E. J., Norskov, J. K.,

Yang, P. Electrochemical Activation of CO2 Through Atomic Ordering Transformations of AuCu

Nanoparticles. J. Am. Chem. Soc. 2017, 139, 8329-8336.

Page 15: Conversion Chenlu Xie

1

Chapter 1. Introduction

Section 1.1 Heterogenous Catalysis in Modern Industry

From Haber–Bosch nitrogen fixation to fossil fuel reforming, chemical conversions catalyzed

by inorganic catalysts have been elemental to modern industrial civilization and are playing central

roles in the pursuit of green and sustainable energy. Haber-Bosch process produced ammonia from

nitrogen and revolutionize the modern agriculture and detonate the global populations. After

entering the 20th century, petroleum-based fuels became the blood of modern industry and this

relies on many catalytic reactions such as Fischer-Tropsch synthesis, which alone contribute 2%

of the world’s fuels. 1 Moreover, the methane reformation and the water gas shift reactions supplied

the syngas and hydrogen feedstock to the industrial reactions and also laid the foundation to the

quickly developing hydrogen-based clean technologies.2-3 The application of the heterogeneous

catalysis covers the range of the chemical manufacturing, energy harvesting, conversion and

storage, to environmental technology.

The central pieces of these important industrial reactions are the inorganic catalyst, which are

metal nanoparticles deposited on stable, high surface area supports in most cases. The catalytic

performance of the metal nanoparticles is governed by their size, shape, composition and their

interface with the oxide support.4-10 The development of these inorganic catalysts has greatly

accelerated with the advances in nanoscience during the past two decades. Especially, colloidal

nanocrystal synthesis provides a viable route for accessing a variety of inorganic nanocrystals with

controlled compositions, sizes and shapes and enables further control over the spatial arrangements

of the nanoparticles to form well-defined interfaces. 11-16

These well-defined catalytic systems have been used to prepare novel catalysts with great

promises for industrial application, and tremendous efforts have been made to understand the

effect of particles size, shape, composition and the metal-support interface on their catalytic

performance. With these understandings and the toolbox of colloidal nanocrystals, we could

pursue new catalysts with controlled structure and better performance and to meet the new

challenges in catalysis.

Section 1.2 Colloidal Nanocrystal Synthesis with Controlled Structure

Section 1.2.1 Basic Principle in Monodispersed Nanocrystal Synthesis

During the past decades, much attention has been paid to understand the nucleation and growth

process in the synthesis and gave rise to the better controlled synthetic route.14-18 And it is well

known that in order to get monodispersed nanocrystals, it is necessary to induce a single nucleation

event and prevent additional nucleation during the growth process. In other words, separation of

nucleation from the subsequent growth is of key significance in homogeneous nucleation. This can

be achieved by a ‘burst of nucleation’ which can be described by the LaMer model,18 which

involves three stages. Herein we describe the three stages using metal nanocrystal synthesis as an

example (Figure 1.1). Metal nanocrystals are typically generated by reduction or decomposition

of the metal precursor, and the nucleation process can be triggered by injection of the reagent (e.g.

reducing agent to the metal precursor) or heating up the reaction mixture to a certain temperature

where the metal precursor starts to thermally decompose. In the first stage, the metal ions were

Page 16: Conversion Chenlu Xie

2

reduced to generate reduced atomic species in the solutions. Due to the extremely high energy

barrier for spontaneous homogenous nucleation, the precipitation would not happen during the

stage even under the supersaturated conditions (Cs). With time increase, the concentration of these

atomic species constantly increases, and once it is above the critical concentration, the degree of

supersaturation is high enough to overcome the energy barrier for nucleation, then the atoms start

to aggregate into small nuclei via homogeneous nucleation. In this stage, the nucleation results in

a rapid depletion of the reactants, so the concentration of atomic species decreases. As long as the

concentration is kept below the critical level, no additional nucleation event will occur, and the

system enters the growth stage, which is stage III. In the growth stage, the remaining metal

precursor will decompose and deposit on the formed nuclei and increase their size. At the same

time, the smaller nuclei with higher surface energy could still be dissolved and further facilitate

the growth of larger nuclei. This Ostwald ripening process would further improve the size

distribution of the nanocrystals.

Figure 1.1. The LaMer model for nucleation and growth. It illustrates the generation of atoms, nucleation and

subsequent growth of nanocrystals as a function of reaction time and concentration of precursor atoms (reproduced

with permission from ref. 18, Copyright © 1950 American Chemical Society).

Page 17: Conversion Chenlu Xie

3

Section 1.2.2 Synthesis of Nanocrystals with Controlled Size, Shape and Composition

As mentioned in the above section, separation of nucleation and growth process is critical to

forming nanocrystals with controlled and uniform size. Many of the monodispersed metal

nanoparticles have been prepared mainly by alcohol reduction or thermal decomposition of metal

precursors in the presence of organic surfactants.16, 19 Noble metals such as Pt, Pd, Rh, Ru, Ag and

Au can be obtained by alcohol reduction,20-22 where high boiling point alcohols such as ethylene

glycol are used as both solvent to dissolve metal precursor and weak reducing agent to generate

metal nanoparticles. In this process, poly(vinylpyrrolidone) (PVP) is commonly used as a surfacant

to passivate the surface of the nanocrystals and stabilize small nuclei. These nanocrystals can also

be made in aqueous solutions using alkylammonium halides as a capping agent and reducing

agents such as sodium borohydride, citric acid, and ascorbic acid. Alternatively, when hydrophobic

nanocrystals are desired, oleylamine and oleic are used as surfactants and weak reducing agent,

and the synthesis can be carried out both at high temperature (rely on metal precursor

decomposition or reduction by the solvent/capping agent) or low temperature with additional

stronger reductant.23-24 The surfactant has both hydrophobic hydrocarbon chains and hydrophilic

functional groups that can bind to and stabilize nanoparticle surfaces in colloidal solution.

Practically, the size control of the nanoparticles can be achieved by choosing different types of

solvents, ratio of reagent, reaction temperature and time. For example, Pt nanoparticles with

controlled size from 3.5 nm to 9nm are synthesized using (NH4)2Pt(II)Cl4 or H2Pt(IV)Cl6 as

precursor, PVP or tetramethylammonium bromide (TTAB) as capping ligands under temperature

range of 140 °C to 180 °C, and the size can be tuned by the oxidation state of the Pt precursor and

reaction temperature.11 The synthesis of transition metal nanoparticles such as Cu and Co could be

achieved by thermal decomposition of the reactive metal precursor [e.g. Co2(CO)8] by hot injection

or heating up method. The hot injection method involves rapid injection of excess precursor into

a hot surfactant solution, leading to fast nucleation and yield narrow size distribution.23 While in

the heating up method, metal precursors, surfactants and solvent are mixed at a low temperature

and then heated up to a certain high temperature to initiate the crystallization process. For examples,

the monodispersed Co nanoparticles can be synthesized by injecting organometallic precursor into

the surfactant solution at high temperature, and the size of Co nanocrystals can be tuned from 3

nm to 10 nm by tuning the injection temperature.23 The copper nanoparticles can be produced by

thermal decomposition of Cu (I) acetate in trioctylamine in the presence of tetradecylphosphonic

acid (TDPA) under high temperature, and the size of Cu can be tuned from 8 nm to 14 nm by

changing the ration between the copper precursor and the surfactant TDPA. 24

Page 18: Conversion Chenlu Xie

4

Figure 1.2. Synthesis of noble with controlled size. Representative transmission electron microscopy (TEM) images

of Pt nanoparticles with controlled size from 3.5nm to 9nm (adopted with permission from ref. 11, Copyright © 2009

American Chemical Society).

Figure 1.3. Synthesis of transition metal nanoparticles with different size. TEM and HRTEM images of Co

nanoparticles with different sizes (3.2, 4.8, 6.8, and 10.2 nm), the precursor injection temperature was controlled to

tune the size (adopted with permission from ref. 23, Copyright © 2012 American Chemical Society).

Page 19: Conversion Chenlu Xie

5

In addition to the rapid and instantaneously nucleation process, synthesizing monodispersed

nanocrystals with well-defined shape requires thermodynamic driving force to stabilize specific

crystallographic facets selectively.15 This is typically achieved by employing molecular

modulators during the growth process, which selectively adsorb to specific crystal planes and

stabilize particular facets. The growth would be then slowed down for these crystal planes and the

crystals will grow on the planes where binding is week. A wide variety of molecules can act as

such modulators and can facilitate the shape control. These molecules includes long chain

surfactants, polymers and biomolecules, small molecules such as adsorbed gas, and even atomic

species, such as different metal ions.25-32 For example, PVP has been demonstrated as an excellent

shape-control agent for noble metal, such as Ag, Au and Pt.25, 32-33 The chain length of PVP will

significantly affects the shape and size dispersity of the products. Some small gas molecules such

as NO2 and CO were also demonstrated as shape-control agent.34-35 Introduction of NO2 leads to

the formation of Pd cuboctahedra and octahedra, since NO2 can selectively stabilize the {111}

facets of the Pd nanocrystals, while Pd nanocubes would be produced without NO2.35 Small

amount of metal ions have also been used to achieve the shape control.25, 36 Pt nanoparticles with

various shapes including cubes, cuboctahedra and octahedra can be synthesized selectivity by

adding silver ions, since the amount of silver ions in the solution can control the crystal growth

rate along <100>.25 Besides the homogeneous nucleation method to achieve shape control, seed-

mediated growth methods are also a common way, where seed particles are added to a growth

medium to facilitate the reduction of metal ions. This way allows a wide range of growth

conditions and higher synthetic flexibility.35, 37

Figure 1.4. Pt nanoparticles with different shape of cubes, cubotahedra and octahedra using silver ions as modulator.

(reproduced with permission from ref. 25, Copyright © 2005 American Chemical Society).

For metallic nanocrystals with more than one elements (e.g. bimetallic nanocrystals) the

composition of the bimetallic nanoparticles can be controlled by adjusting the ratio of two metal

precursor. In bimetallic nanocrystals, different metal atoms can have various spatial arrangements

within the nanocrystal, which is thermodynamically related to the phase diagram of the two

elements and can be synthetically tuned.38-41 For metals that have alloy phase (e.g. Au and Cu),

low-temperature synthesis would generally give nanocrystals with different metal atoms randomly

distributed within the nanocrystal. 38, 40, 42 At higher temperatures, thermodynamically more stable

phases, if exist in the phase diagram, can be achieved where different metal atoms often occupy

crystallographically distinct sites. When noble metals and transition metals form bimetallic

particles, as noble metal atoms are comparably more inert, they tend to occupy positions with

higher surface energy, such as the corner, edge and surface of the nanocrystal.43-44 This effect is

Page 20: Conversion Chenlu Xie

6

especially pronounced when harsh reaction conditions are used (high temperature, slight acidic

media or oxidative environment). The spatial distribution of different metal atoms in nanocrystals

can thus be manipulated synthetically to achieve more desirable catalytic performance such as

higher stability, higher mass activity for the noble metal atoms, and different surface electronic

structure that interact differently with small molecules.

Section 1.3 Synthesis of Supported Metal Catalysts

In many industries including chemical manufacturing, energy-related applications and

environmental remediation, metallic nanoparticles are used as heterogeneous catalysts for

chemical conversions. To prevent nanocrystal aggregation and achieve maximum accessible

surface area, metallic nanocrystals are usually finely deposited over supports with high surface

area.7, 20, 45 The common supports materials are oxides such as SiO2, Al2O3 or TiO2 and porous

carbon based materials, which exhibit high specific surface areas, thermal and mechanical stability

and tunable pore sizes.46-49 Zeolite is also commonly used, mostly in the oil-refining and petroleum

based applications.50 Recently other emerging class of porous materials such as metal organic

frameworks have also been used as supports for metal nanoparticles.51 The supports can disperse

and stabilize the nanoparticles, and more importantly, the interact with the metal nanocrystal

through the interface and tune the catalytic performance. Fundamentally, it is the metal-oxide

interface that act as active sites in many reactions (e.g. CO oxidation over Au-TiO2 catalyst).52-55

Thus, it is also of great importance to synthesis these supported metal catalysts with elegantly

controlled interfaces and structure. Here we highlight three most common synthetic routes.

Section 1.3.1 Conventional Synthetic Method: Precipitation and Impregnation

Precipitation has long been used to produced supported catalysts, where the precursors of

nanocrystals are subject to change of temperature, pH or chemical environment and give rise to

the metal nanocrystals. The controlled precipitation from a precursor solution follows the principle

of nucleation and growth, where the separation of burst of nucleation and growth is critical to

produce monodispersed nanoparticles. Coprecipitation is the most common used method, where

the precursor of metal and supports are dissolved and mixed together in a very high concentration,

and all components nucleate at a high rate simutaenously. In this way, the nucleation and growth

of a combined solid precursor of the metal nanocrystal and support can be completed in a single

step. Typically, highly soluble metal salts of the metal nanocrystal and oxide support, such as

nitrates and chlorides, are dissolved at high concentration and are mixed with the basic reagents

such as alkali carbonate or hydroxides to form the metal carbonates or hydroxides, which

precipitate from the solution.56-57 In order to get monodispersed nanoparticles, sufficient mixing,

uniform temperature and concentration are necessary to avoid additional nucleation events and

inhomogeneous growth. After precipitation, washing steps are necessary to remove the residual

metal salts and subsequent thermal treatment (mostly in reducing environment) is also needed to

yield the oxide support and metallic nanocrystals.

Impregnation is another frequently used preparation method of supported metal catalysts. The

process involves diffusing the solution of metal precursor into the porous supports followed by

calcination to convert the absorbed metal precursor to corresponding nanocrystalline catalysts.49,

58-59 Most common used precursor includes metal chloride, sulfate, nitrates, acetate, carbonate or

acetylacetonate. Water is mostly used solvent for inorganic salts due to its high solubility of these

precursors, while organic solvents are mainly used for organometallic precursor. The adsorption

Page 21: Conversion Chenlu Xie

7

of the metal salts from solution onto the support plays a significant role in the impregnation, which

will affect the metal loading amount as well as the metal distribution in the support.60 Electrostatic

adsorption is the most widely employed to load the metal precursor, where the surface charge of

the metal oxide support would attract ions with opposite polarity, and the pH of the solution and

point of zero charge of the support is also crucial in this process. After drying, the deposited metal

precursor is converted into active metal nanoparticles by high temperature calcination/reduction

step. Both drying step and calcination/reduction step can also affect the state of final catalysts.

Section 1.3.2 Emerging Synthetic Method: Preparation of Catalysts with Well-controlled

Structure and Interface

Deposition of the as-synthesized nanocrystals on the support materials is another way to obtain

supported nanocrystals as heterogeneous catalysts. The method has attracted much attention as it

can reliably produce catalysts with well-defined structure due to the structurally well-controlled

nanocrystals prepared by colloidal synthesis.

Wet impregnation is one of the most simple and versatile way to deposit nanoparticles onto

supports, where nanoparticle solutions are mixed with the supports so that the particles can diffuse

into the pores of the supports or absorb to their surface with sonication and stirring.61-62 The

deposition of nanoparticles often relies on attractive interactions between the support and the

nanocrystal, such as electrostatic interactions, van der Waals interactions, or dipole-charge

interactions. This method is highly tunable on the types of metal nanoparticle, oxide supports as

well as loading amount.

Atomic layer deposition is another efficient deposition method for the production of finely

dispersed metal particles on supports.63 Basically, ALD is a thin-film deposition method where the

surface of a substrate is exposed to vapors of the precursors in alternating pulse, and the precursor

molecules reacts on the surface in a self-limiting way to form uniform monolayers per cycle. ALD

method could be used to deposit variety types of oxide layers, such as TiO2, ZnO and Al2O3. For

the deposition of transition metal and especially noble metals, for example, Pt and Pd, the metals

tend to agglomerate to form small particles and thus can be used as nanocrystal catalysts.64 Layer-

by-layer deposition method is another preparation method for metal-oxide catalysts, where metal

nanocrystals are assembled and deposited on top of the deposited oxide support layer.52 Both of

the methods have advantage to form a wide range of oxide supported metal catalysts with well-

controlled structure, however, the stability of the catalysts synthesized by these two methods limits

their practical applications.

Page 22: Conversion Chenlu Xie

8

Figure 1.5. Preparation of CeO2-Pt-SiO2 catalysts by Layer-by-layer deposition method (reproduced with permission

from ref. 52, Copyright © 2011 Nature Publishing Group)

Supported metal nanocrystals can also be synthesized by directly growth or coating of porous

oxide support shell around metal nanoparticles core. This core-shell configuration could provide

high thermal stability and prevent metal particles from coalescence and sintering during catalytic

reactions. For example, Pt@mSiO2 core shell catalysts were designed as a thermally stable catalyst,

where the mesoporous silica shells encapsulating the Pt cores can stabilize the Pt nanocrystal up

to 750C in air.12 Besides mesoporous silica shell, TiO2 shell and a variety of metal organic

framework shell can also be synthesized on the metal particle core with tunable thickness and pore

size. 65-67

Page 23: Conversion Chenlu Xie

9

Figure 1.6. Synthesis of metal catalysts by directly coating of porous silica shell around the metal nanoparticles. (a)

The schematic of synthesis process of Pt@mSiO2 catalysts. (b) TEM images of Pt nanoparticles. (c) TEM images of

as-synthesized Pt@mSiO2 catalysts (reproduced with permission from ref. 12, Copyright © 2009 Nature Publishing

Group).

Section 1.4 Catalytic Reaction Studies Using Nanoparticle Catalysts

Metal nanoparticles with controlled size, shape and composition provides well-defined

platform for systematic study of their catalytic performance. In 1969, Boudart first classified

heterogeneous catalytic reactions as structure-sensitive and structure-insensitive according to

whether the reaction activity or selectivity depends on the catalyst structure.68 Since then,

tremendous efforts have been devoted to study the effects of particle size, shape, composition and

the metal-support interface on the catalytic performance of these nanoparticle-based catalysts. The

reduction of size would increase the surface to volume ratio, and the number of chemically

unsaturated sites on the surface (highly active edge and corner sites), as well as increase the Fermi

level of the nanoparticles which changes the binding energy of adsorbates, which can influence

both reaction activity and selectivity.69-71 For example, for Fischer-Tropsch reaction catalyzed by

Co nanoparticles, the turn over frequency of the reaction reach optimal value when particle size is

above 6nm, and a higher CH4 selectivity is found for Co nanoparticle with size smaller than 6nm.

Also, the selectivity to olefin and paraffins is also influence by the cobalt particle size, where larger

Co nanoparticles gives lower olefin/paraffin ratio.46

The shape of the nanoparticles also has a significant effect on their catalytic properties, which

is related to the exposed facets of the particles and their surface structure and the fraction of atoms

at corners and edges. Earlier studies using the single-crystal model surface have demonstrated that

different catalytic properties were found using single crystals with different exposed facets.72-74

Adding another element into the metal nanoparticle can form bimetallic nanoparticles, the

electronic structure of the bimetallic system will be different from the parent single metal system,

Page 24: Conversion Chenlu Xie

10

which will tune the binding strength of the adsorbates on the surface.40-41 The catalytic active site

of the bimetallic structure can also be different from the parent metal surfaces, where different

atomic distribution would give different active site configurations of the new bimetallic system.

The metal-support interaction also has profound effect on the catalytic properties of the metal

nanocrystal, and it was discovered that the actual active sites were at the metal-oxide interface in

many catalytic reactions.53 Further study give rise to the strong metal-support interaction (SMSI)

theory to describe the drastic change in chemisorption properties of Group VIII metals such as Fe,

Ni, Rh, Pt, Pd and Ir, when in contact with certain oxides (TiOx, CeOx, TaOx and NbOx). After

that, substantial efforts have been devoted to understanding the SMSI, which indicates the

interfacial charge transfer could be the origin of different catalytic performance.75 For example, in

the CO oxidation via Pt nanoparticles deposited on different types of mesoporous oxides (Co3O4,

NiO, MnO2, Fe2O3, and CeO2), the activity varies greatly with different oxides and Pt/SiO2 shows

the lowest activity.62 Also different metal-oxide support interface can affect the reaction selectivity,

for example in the n-hexane isomerization catalyzed by Pt nanoparticles on different kinds of oxide

supports, the product distribution is highly dependent on the type of oxides.76

Figure 1.7. Catalytic activity and selectivity depends on the metal-oxide interfaces. (a) TOFs of Pt nanoparticles

supported on different mesoporous oxide catalysts in CO oxidation (reproduced with permission from ref. 62,

Copyright © 2013 American Chemical Society). (b) Product selectivity in n-hexane reforming over Pt nanoparticles

supported on different kinds of oxide supports (reproduced with permission from ref. 76, Copyright © 2014 American

Chemical Society).

Page 25: Conversion Chenlu Xie

11

Section 1.5 References

(1) Schulz, H. Appl. Catal., A 1999, 186 (1), 3-12.

(2) Barelli, L.; Bidini, G.; Gallorini, F.; Servili, S. Energy 2008, 33 (4), 554-570.

(3) Newsome, D. S. Catal. Rev. 1980, 21 (2), 275-318.

(4) Herzing, A. A.; Kiely, C. J.; Carley, A. F.; Landon, P.; Hutchings, G. J. Science 2008, 321

(5894), 1331-1335.

(5) Narayanan, R.; El-Sayed, M. A. Nano Lett. 2004, 4 (7), 1343-1348.

(6) Haruta, M. Gold Bull 2004, 37 (1), 27-36.

(7) An, K.; Somorjai, G. A. Catal. Lett. 2015, 145 (1), 233-248.

(8) Somorjai, G. A.; Tao, F.; Park, J. Y. Top. Catal. 2008, 47 (1), 1-14.

(9) Li, Y.; Somorjai, G. A. Nano Lett. 2010, 10 (7), 2289-2295.

(10) Xia, Y.; Yang, H.; Campbell, C. T. Acc. Chem. Res. 2013, 46 (8), 1671-1672.

(11) Tsung, C.-K.; Kuhn, J. N.; Huang, W.; Aliaga, C.; Hung, L.-I.; Somorjai, G. A.; Yang, P. J.

Am. Chem. Soc. 2009, 131 (16), 5816-5822.

(12) Joo, S. H.; Park, J. Y.; Tsung, C.-K.; Yamada, Y.; Yang, P.; Somorjai, G. A. Nat. Mater. 2008,

8, 126.

(13) Lee, H.; Habas, S. E.; Kweskin, S.; Butcher, D.; Somorjai, G. A.; Yang, P. Angew. Chem. Int.

Ed. 2006, 118 (46), 7988-7992.

(14) Xia, Y.; Xiong, Y.; Lim, B.; Skrabalak, S. E. Angew. Chem. Int. Ed. 2009, 48 (1), 60-103.

(15) Tao, A. R.; Habas, S.; Yang, P. Small 2008, 4 (3), 310-325.

(16) Park, J.; Joo, J.; Kwon, S. G.; Jang, Y.; Hyeon, T. Angew. Chem. Int. Ed. 2007, 46 (25), 4630-

4660.

(17) Yau, S. T.; Vekilov, P. G. Nature 2000, 406, 494.

(18) LaMer, V. K.; Dinegar, R. H. J. Am. Chem. Soc. 1950, 72 (11), 4847-4854.

(19) Somorjai, G. A.; Park, J. Y. Angew. Chem. Int. Ed. 2008, 47 (48), 9212-9228.

(20) Song, H.; Rioux, R. M.; Hoefelmeyer, J. D.; Komor, R.; Niesz, K.; Grass, M.; Yang, P.;

Somorjai, G. A. J. Am. Chem. Soc 2006, 128 (9), 3027-3037.

(21) Zhang, Y.; Grass, M. E.; Kuhn, J. N.; Tao, F.; Habas, S. E.; Huang, W.; Yang, P.; Somorjai,

G. A. J. Am. Chem. Soc 2008, 130 (18), 5868-5869.

(22) Joo, S. H.; Park, J. Y.; Renzas, J. R.; Butcher, D. R.; Huang, W.; Somorjai, G. A. Nano Lett.

2010, 10 (7), 2709-2713.

(23) Iablokov, V.; Beaumont, S. K.; Alayoglu, S.; Pushkarev, V. V.; Specht, C.; Gao, J.; Alivisatos,

A. P.; Kruse, N.; Somorjai, G. A. Nano Lett. 2012, 12 (6), 3091-3096.

Page 26: Conversion Chenlu Xie

12

(24) Hung Ling‐, I.; Tsung, C. K.; Huang, W.; Yang, P. Adv. Mater. 2010, 22 (17), 1910-1914.

(25) Song, H.; Kim, F.; Connor, S.; Somorjai, G. A.; Yang, P. J. Phys. Chem. B 2005, 109 (1),

188-193.

(26) Cui, F.; Yu, Y.; Dou, L.; Sun, J.; Yang, Q.; Schildknecht, C.; Schierle-Arndt, K.; Yang, P.

Nano Lett. 2015, 15 (11), 7610-7615.

(27) Huo, Z.; Tsung, C.-k.; Huang, W.; Zhang, X.; Yang, P. Nano Lett. 2008, 8 (7), 2041-2044.

(28) Zhang, Y.; Grass, M. E.; Habas, S. E.; Tao, F.; Zhang, T.; Yang, P.; Somorjai, G. A. J. Phys.

Chem. C 2007, 111 (33), 12243-12253.

(29) Grzelczak, M.; Perez-Juste, J.; Mulvaney, P.; Liz-Marzan, L. M. Chem. Soc. Rev. 2008, 37

(9), 1783-1791.

(30) Huang, X.; Tang, S.; Mu, X.; Dai, Y.; Chen, G.; Zhou, Z.; Ruan, F.; Yang, Z.; Zheng, N. Nat.

Nanotechnol. 2010, 6, 28.

(31) Jana, N. R.; Gearheart, L.; Murphy, C. J. J. Phys. Chem. B 2001, 105 (19), 4065-4067.

(32) Xiong, Y.; Xia, Y. Adv. Mater. 2007, 19 (20), 3385-3391.

(33) Henzie, J.; Grünwald, M.; Widmer-Cooper, A.; Geissler, P. L.; Yang, P. Nat. Mater. 2011,

11, 131.

(34) Chen, M.; Wu, B.; Yang, J.; Zheng, N. Adv. Mater. 2012, 24 (7), 862-879.

(35) Habas, S. E.; Lee, H.; Radmilovic, V.; Somorjai, G. A.; Yang, P. Nat. Mater. 2007, 6, 692.

(36) Kim, F.; Connor, S.; Song, H.; Kuykendall, T.; Yang, P. Angew. Chem. Int. Ed 2004, 43 (28),

3673-3677.

(37) Sun, Y.; Xia, Y. Science 2002, 298 (5601), 2176.

(38) Kim, D.; Xie, C.; Becknell, N.; Yu, Y.; Karamad, M.; Chan, K.; Crumlin, E. J.; Nørskov, J.

K.; Yang, P. J. Am. Chem. Soc 2017, 139 (24), 8329-8336.

(39) Musselwhite, N.; Alayoglu, S.; Melaet, G.; Pushkarev, V. V.; Lindeman, A. E.; An, K.;

Somorjai, G. A. Catal. Lett. 2013, 143 (9), 907-911.

(40) Kim, D.; Resasco, J.; Yu, Y.; Asiri, A. M.; Yang, P. Nat. Commun. 2014, 5, 4948.

(41) Yu, W.; Porosoff, M. D.; Chen, J. G. Chem. Rev. 2012, 112 (11), 5780-5817.

(42) Sra, A. K.; Schaak, R. E. J. Am. Chem. Soc 2004, 126 (21), 6667-6672.

(43) Tao, F.; Grass, M. E.; Zhang, Y.; Butcher, D. R.; Renzas, J. R.; Liu, Z.; Chung, J. Y.; Mun,

B. S.; Salmeron, M.; Somorjai, G. A. Science 2008, 322 (5903), 932.

(44) Chen, C.; Kang, Y.; Huo, Z.; Zhu, Z.; Huang, W.; Xin, H. L.; Snyder, J. D.; Li, D.; Herron, J.

A.; Mavrikakis, M.; Chi, M.; More, K. L.; Li, Y.; Markovic, N. M.; Somorjai, G. A.; Yang, P.;

Stamenkovic, V. R. Science 2014, 343 (6177), 1339.

Page 27: Conversion Chenlu Xie

13

(45) Shekhar, M.; Wang, J.; Lee, W.-S.; Williams, W. D.; Kim, S. M.; Stach, E. A.; Miller, J. T.;

Delgass, W. N.; Ribeiro, F. H. J. Am. Chem. Soc 2012, 134 (10), 4700-4708.

(46) Bezemer, G. L.; Bitter, J. H.; Kuipers, H. P. C. E.; Oosterbeek, H.; Holewijn, J. E.; Xu, X.;

Kapteijn, F.; van Dillen, A. J.; de Jong, K. P. J. Am. Chem. Soc 2006, 128 (12), 3956-3964.

(47) Pan, X.; Fan, Z.; Chen, W.; Ding, Y.; Luo, H.; Bao, X. Nat. Mater. 2007, 6, 507.

(48) Wan, Y.; Zhao, Y. Chem. Rev. 2007, 107 (7), 2821-2860.

(49) Munnik, P.; de Jongh, P. E.; de Jong, K. P. Chem. Rev. 2015, 115 (14), 6687-6718.

(50) Shu, Y.; Ichikawa, M. Catal. Today 2001, 71 (1), 55-67.

(51) Rungtaweevoranit, B.; Baek, J.; Araujo, J. R.; Archanjo, B. S.; Choi, K. M.; Yaghi, O. M.;

Somorjai, G. A. Nano Lett. 2016, 16 (12), 7645-7649.

(52) Yamada, Y.; Tsung, C.-K.; Huang, W.; Huo, Z.; Habas, S. E.; Soejima, T.; Aliaga, C. E.;

Somorjai, G. A.; Yang, P. Nat. Chem. 2011, 3, 372.

(53) Green, I. X.; Tang, W.; Neurock, M.; Yates, J. T. Science 2011, 333 (6043), 736.

(54) Matsubu, J. C.; Zhang, S.; DeRita, L.; Marinkovic, N. S.; Chen, J. G.; Graham, G. W.; Pan,

X.; Christopher, P. Nat. Chem. 2016, 9, 120.

(55) Liu, P.; Qin, R.; Fu, G.; Zheng, N. J. Am. Chem. Soc 2017, 139 (6), 2122-2131.

(56) Kaluza, S.; Behrens, M.; Schiefenhövel, N.; Kniep, B.; Fischer, R.; Schlögl, R.; Muhler, M.

Chemcatchem 2010, 3 (1), 189-199.

(57) Kaluza, S.; Schröter Marie, K.; Naumann d'Alnoncourt, R.; Reinecke, T.; Muhler, M. Adv.

Funct. Mater. 2008, 18 (22), 3670-3677.

(58) Eggenhuisen, T. M.; Friedrich, H.; Nudelman, F.; Zečević, J.; Sommerdijk, N. A. J. M.; de

Jongh, P. E.; de Jong, K. P. Chem. Mater. 2013, 25 (6), 890-896.

(59) Park, J.; Regalbuto, J. R. J. Colloid Interface Sci. 1995, 175 (1), 239-252.

(60) Bourikas, K.; Kordulis, C.; Lycourghiotis, A. Cat. Rev. 2006, 48 (4), 363-444.

(61) Zheng, N.; Stucky, G. D. J. Am. Chem. Soc 2006, 128 (44), 14278-14280.

(62) An, K.; Alayoglu, S.; Musselwhite, N.; Plamthottam, S.; Melaet, G.; Lindeman, A. E.;

Somorjai, G. A. J. Am. Chem. Soc 2013, 135 (44), 16689-16696.

(63) Cao, K.; Cai, J.; Liu, X.; Chen, R. J. Vac. Sci. Technol 2017, 36 (1), 010801.

(64) Lu, J.; Liu, B.; Greeley, J. P.; Feng, Z.; Libera, J. A.; Lei, Y.; Bedzyk, M. J.; Stair, P. C.; Elam,

J. W. Chem. Mater. 2012, 24 (11), 2047-2055.

(65) Kuo, C.-H.; Tang, Y.; Chou, L.-Y.; Sneed, B. T.; Brodsky, C. N.; Zhao, Z.; Tsung, C.-K. J.

Am. Chem. Soc 2012, 134 (35), 14345-14348.

(66) Zhang, Q.; Shu, X.-Z.; Lucas, J. M.; Toste, F. D.; Somorjai, G. A.; Alivisatos, A. P. Nano

Lett. 2014, 14 (1), 379-383.

Page 28: Conversion Chenlu Xie

14

(67) An, K.; Zhang, Q.; Alayoglu, S.; Musselwhite, N.; Shin, J.-Y.; Somorjai, G. A. Nano Lett.

2014, 14 (8), 4907-4912.

(68) Boudart, M. Adv. Catal. 1969, 20, 153-166.

(69) Alayoglu, S.; Pushkarev, V. V.; Musselwhite, N.; An, K.; Beaumont, S. K.; Somorjai, G. A.

Top Catal. 2012, 55 (11), 723-730.

(70) Kleis, J.; Greeley, J.; Romero, N. A.; Morozov, V. A.; Falsig, H.; Larsen, A. H.; Lu, J.;

Mortensen, J. J.; Dułak, M.; Thygesen, K. S.; Nørskov, J. K.; Jacobsen, K. W. Catal. Lett. 2011,

141 (8), 1067-1071.

(71) Schauermann, S.; Nilius, N.; Shaikhutdinov, S.; Freund, H.-J. Acc. Chem. Res. 2013, 46 (8),

1673-1681.

(72) Mahmoud, M. A.; Narayanan, R.; El-Sayed, M. A. Acc. Chem. Res. 2013, 46 (8), 1795-1805.

(73) Bratlie, K. M.; Kliewer, C. J.; Somorjai, G. A. J. Phys. Chem. B 2006, 110 (36), 17925-17930.

(74) Bratlie, K. M.; Lee, H.; Komvopoulos, K.; Yang, P.; Somorjai, G. A. Nano Lett. 2007, 7 (10),

3097-3101.

(75) Tauster, S. J.; Fung, S. C. J. Catal. 1978, 55 (1), 29-35.

(76) An, K.; Alayoglu, S.; Musselwhite, N.; Na, K.; Somorjai, G. A. J. Am. Chem. Soc 2014, 136

(19), 6830-6833.

Page 29: Conversion Chenlu Xie

15

Chapter 2. Tandem Alkene Hydroformylation over a

Three Dimensional Nanostructured Catalyst with

Multiple Interfaces

Section 2.1 Introduction

Nanocrystalline inorganic catalysts for heterogeneous reactions have recently been the subject

of extensive research due to their essential role in modern chemical industrial processes.1−3 Thus

far, the design and development of these catalysts has been exclusively carried out at the single-

interface length scale, where the atomic arrangements on the catalytic interface are tuned to alter

the activity and selectivity of a single chemical conversion process.4−8 Recently, however, tandem

catalysis was demonstrated in a heterogeneous gas-phase reaction,9 wherein two metal−oxide

interfaces in a single nanostructure catalyzed sequential chemical conversions with high product

selectivity. This work has inspired us to explore the rational design of nanocrystalline

heterogeneous catalysts beyond the single interface length scale. Specifically, instead of

optimizing a single catalytic interface, the tandem catalyst design takes advantage of the synergy

between different chemical conversions on multiple spatially arranged interfaces to achieve

desirable product distributions. In this context, harmonized reaction kinetics at two interfaces can

facilitate the overall sequential reactivity wherein one reactive species modulates the chemical

conversion of another. Thus, a thorough study of heterogeneous tandem reactions, which is yet to

be carried out, would help to elucidate the underlying principles of such catalysis and open up new

opportunities for application in heterogeneous reactions.

A model system for tandem catalysis is the ethylene hydroformylation reaction, where the first

chemical conversion is the production of hydrogen (H2) and carbon monoxide (CO) by methanol

decomposition, followed by ethylene hydroformylation. This reaction is ideal not only because the

production of aldehydes via alkene hydroformylation is an essential industrial process10,11 but also

due to the fact that the initial decomposition of methanol and subsequent hydroformylation are

chemically orthogonal and compatible reactions. Previous research on alkene hydroformylation

via heterogeneous catalysis led to the development of nanocrystaline Rh- and Pt-based

catalysts.12−16 However, these catalysts gave low aldehyde selectivity due to dominance of the

competing alkene hydrogenation reaction.17−19 We previously showed that tandem catalysis is a

plausible approach to overcome this low selectivity, where propanal could be preferentially

produced by reacting ethylene with H2 and CO formed in situ over a nanocrystal bilayer

CeO2−Pt−SiO2.9 The origin of the high selectivity has yet to be elucidated, however, and an in-

depth study of this tandem reaction would bring new insights concerning tandem catalysis.

An essential prerequisite for successful implementation of tandem catalysis is the synthesis of

the complex nanostructured catalyst. Although the drop-casted bilayer CeO2−Pt−SiO2 catalyst was

adequate to demonstrate tandem catalysis, it exhibited poor stability toward aggregation and low

active site density.20,21 In order to thoroughly study the tandem hydroformylation reaction and gain

insights into the fundamental principles behind this sort of catalysis, a stable, well-defined catalyst

is necessary. Such a catalyst could be obtained by engineering both catalytic interfaces into one

particle. Enclosing this particle in mesoporous silica (mSiO2) would further provide thermal

stability and enhance active-site accessibility by favoring the high surface area, powdered form of

Page 30: Conversion Chenlu Xie

16

the catalyst.17, 21 Fundamentally, the fabrication of such a complex nanostructure is of synthetic

significance and would be a great example of utilizing the synthetic control of inorganic

nanocrystals to pursue the next generation heterogeneous catalysts.

Herein, we report the design and synthesis of a new generation tandem catalyst, the three-

dimensional CeO2−Pt@mSiO2, with a core−shell configuration that consists of a CeO2−Pt core

and mesoporous silica shell. Tandem hydroformylation reactions carried out by this 3D catalyst

show greatly enhanced propanal selectivity compared to the singlestep ethylene hydroformylation

with CO and H2. Significantly, the catalytic tandem hydroformylation can also be extended to

propylene and 1-butene. Further study of the tandem hydroformylation indicates that the superior

propanal selectivity can be attributed to the synergy between the two sequential reactions and the

altered reaction pathway afforded by the tandem reaction compared to the single-step reaction.

Section 2.2 Experimental Section

Section 2.2.1 Methods and Materials

Cerium nitrate hexahydrate (99% trace metals basis), poly(vinylpyrrolidone) (PVP, Mw =

360,000), PVP (Mw = 29000), tetradecyltrimethylammonium bromide (TTAB), cetrimonium

bromide (CTAB) (99%), and tetraethyl orthosilicate (TEOS) (99.999% trace metal basis) were

purchased from Sigma-Aldrich. Ammonium hexachloroplatinate (IV) [(NH4)2PtCl6, Pt 43.4% min]

was purchased from Alfa Aesar. Ethanol and ethylene glycol were purchased from Fisher

Chemical. Ammonia solution (28−30%) was purchased from EMD Millipore. All chemicals were

used as received without further purification.

The structural analysis of the composite nanoparticles was performed using transmission

electron microscopy (TEM) on a FEI Tecnai F20 at an accelerating voltage of 200 kV. High-angle

annular dark-field scanning transmission electron microscopy (HAADF-STEM) and energy

dispersive X-ray spectroscopy (EDS) mapping was carried out with an FEI TitanX 60−300, which

provided the elemental distribution of the catalyst. The surface area and pore size distribution of

the catalyst were obtained by nitrogen physisorption experiments, which were carried out on a

Quantachrome Autosorb-1 analyzer. Platinum quantitative analysis by inductively coupled plasma

atomic emission spectroscopy (ICP-AES) was carried out on a PerkinElmer optical emission

spectrometer (Optima 7000 DV). Before the ICP-AES measurement, the catalyst was digested in

aqua regia for 24 h, and then deionized water was added to dilute the solution. A clear solution

was obtained for the ICP-AES measurement by centrifuging at 4000 rpm to remove sediment.

The reactions were carried out in a batch mode reactor equipped with a boron nitride substrate

heater and a metal bellows circulation pump for gas mixing. The batch reactor consists of a

stainless steel chamber equipped with a boron nitride heater, a turbo pump, a mechanical pump

and a circulator pump (Figure 2.1). In the gas feeding system, H2, CO, ethylene and He lines as

well as a methanol container were connected to the reactor, and were delivered via a series of

independent needle valves. The molar ratio of each reactant gas was precisely controlled via their

partial pressure. Pressure was monitored by a digital pressure gauge, and temperature was

monitored by the thermocouple underneath the catalyst. The gases exiting the reactor were

analyzed using a Hewlett Packard HP 5890 Series II chromatograph equipped with both flame

ionization detector (FID) and a thermal conductivity detector (TCD). For the alkene

hydroformylation with CO and H2 formed by methanol decomposition, the reactor was typically

filled with 35 Torr of methanol, 7.5 Torr of alkene, and 727.5 Torr of helium. For single step

Page 31: Conversion Chenlu Xie

17

alkene hydroformylation, the reactor was typically filled with 35 Torr of CO, 70 Torr of H2, 7.5

Torr of alkene, and 657.5 Torr of helium. The products were analyzed approximately every 20 min

by the GC, the carrier gas was helium, the FID was used to quantify the alkene and aldehyde

products, and the TCD was used for CO quantification.

Figure 2.1. The schematic of the reactor.

The turnover frequency (TOF) values are derived by the number of Pt active sites and the

number of propanal molecules produced as monitored by GC. The TOF was calculated based on

the experiment data with a total conversion of reactants below 20%. The propanal selectivity was

calculated on a carbon basis and defined as follows: propanal selectivity = (propanal

formed)/(ethylene converted) × 100%. Number of active sites was calculated through two different

methods. The first method was based on the geometric dispersion of Pt nanoparticles, by taking

particle size from TEM and weight % Pt loading from ICP-AES, assuming the same atomic density

on the Pt nanoparticle surface as Pt (111) and all the surface Pt atoms are accessible. The second

method was based on actual measurement. Ethylene hydrogenation reaction, which was known to

be structure-insensitive22-24 was performed to estimate the accessible number of Pt sites of the

catalysts. The number of active sites derived was found to be 70% of the active sites from

geometric determination, likely due to silica over coating. Therefore, the number of active sites

estimated in this study was based on the second method.

Section 2.2.2 Synthesis of 3D tandem catalysts

Synthesis of CeO2 Nanoparticles. Cerium nitrate hexahydrate (0.85 g) was dissolved in a

mixture of deionized water (5 mL) and ethanol (5 mL). To this solution was added 30 mL of a

PVP (Mw = 360000) ethanol solution (60 mg/mL). This reaction mixture was heated in a stainless-

steel autoclave to 140 °C for 24 h. The as-synthesized CeO2 nanoparticles were collected by

centrifugation (12000 rpm, 60 min) and then washed twice with water and ethanol and stored in

ethanol for further synthesis.

Page 32: Conversion Chenlu Xie

18

Overgrowth of Pt NPs on As-Synthesized CeO2 NPs. The as-synthesized CeO2 nanoparticles

(40 mg) were dispersed in 20 mL ethanol. TTAB (36.8 mg) and PVP (Mw = 29000, 21.8 mg)

dissolved in ethylene glycol (16 mL) were added to the CeO2 NPs/ethanol solution.

(NH4)2Pt(IV)Cl6 (9.75 mg, this amount can be changed accordingly to tune the Pt loading amount)

was dissolved into ethylene glycol (4 mL) in a 25 mL three-neck round flask at 80 °C under argon

protection with magnetic stirring. This Pt precursor solution was mixed with the CeO2 NPs/ethanol

solution and heated to 140 °C for 6 h in a stainless-steel autoclave. The as-synthesized CeO2−Pt

NPs were separated by centrifugation (12000 rpm, 45 min) and then re-dispersed in 40 mL

deionized water.

Synthesis of CeO2−Pt@mSiO2 Tandem Catalyst. The CeO2−Pt@mSiO2 core−shell

nanoparticles were prepared by a reported sol−gel method with some modification. The solution

of pre-synthesized CeO2−Pt (40 mg in 45 mL deionized water) was mixed with a solution of CTAB,

which was prepared by dissolving 225 mg CTAB in 30 mL of ethanol. An ammonia solution (0.2

mL) was added to the above solution with stirring. A controlled amount of 1 vol % TEOS diluted

with ethanol was then added under continuous magnetic stirring at room temperature. After 6 h,

the as-synthesized CeO2−Pt@SiO2 nanoparticles were obtained by centrifugation (6000 rpm, 5

min). The product was calcined at 350 °C for 1 h in static air to remove the CTAB template and

other surfactants (PVP, TTAB) to generate CeO2−Pt@mSiO2 particles with clean interfaces.

Section 2.2.3 Synthesis of single interface catalysts

Synthesis of Pt@mSiO2 single interface catalyst. The Pt@mSiO2 core shell nanoparticles

were prepared using the similar method as CeO2−Pt@mSiO2. The solution of pre-synthesized Pt

nanoparticles (3.5 mg in 25 mL deionized water) was mixed with a solution of CTAB, which was

prepared by dissolving 125 mg CTAB in 25 ml ethanol. An ammonia solution (0.1 mL) was added

to the above solution with stirring. Then a controlled amount of 1 vol % TEOS diluted with ethanol

was added under continuous magnetic stirring at room temperature. After 6 hours, the as-

synthesized Pt@SiO2 nanoparticles were obtained by centrifugation (6000 rpm, 5 minutes). The

product was calcined at 350 °C for 1 hour in static air to remove ligands and generate Pt@mSiO2

particles with clean interfaces.

Synthesis of Pt nanoparticles. For the synthesis of Pt nanoparticles, 44.4 mg (NH4)2PtCl6 and

36.8 mg TTAB and 21.8 mg PVP (Mw = 29000) were mixed with 20 mL of ethylene glycol in a 50

mL round bottom flask at room temperature. The mixture was heated at 80 °C for 20 min under

argon protection and then quickly heated to 180 °C. The solution was kept at 180 °C for 1 h before

it was cooled to room temperature, and the product was centrifuged at 12000 rpm for 10 min. The

supernatant solution was separated and centrifuged again at 12000 rpm for 10 min, twice. The Pt

nanoparticle colloids were collected and re-dispersed in 5 mL of deionized water by sonication for

further use.

Page 33: Conversion Chenlu Xie

19

Section 2.3 Results and Discussion

Section 2.3.1 Synthesis and Characterizations of 3D Tandem Catalysts.

The CeO2−Pt@mSiO2 catalyst consists of a CeO2−Pt core and mesoporous silica shell and thus

features both Pt/CeO2 and Pt/SiO2 functional interfaces. Importantly, while Pt/CeO2 is a typical

catalytic interface for methanol decomposition to produce CO and H2,24,25 Pt/SiO2 has been shown

to exhibit catalytic activity for the hydroformylation of ethylene with CO and H2.12 Thus, with

these integrated interfaces, this catalyst could convert methanol and ethylene to propanal through

a tandem process. Indeed, methanol could diffuse through the mesoporous silica shell to the

Pt/CeO2 core, whereby it would be decomposed to CO and H2. The subsequent outward diffusion

of the CO an H2 through the mSiO2 channels and reaction with ethylene would then result in the

formation of propanal at the nearby Pt/SiO2 interface, completing the two step reaction.

Figure 2.2. Synthesis and characterization of the 3D nanocrystal tandem catalyst CeO2-Pt@mSiO2. (A) Synthesis of

the tandem catalyst. (B) TEM image of well-dispersed CeO2 nanoparticles. (C) TEM image of CeO2-Pt nanoparticles

synthesized the overgrowth of Pt. (D) TEM image of core-shell CeO2-Pt@mSiO2 nanoparticles. Scale bar: 100 nm.

The synthesis of the CeO2−Pt@mSiO2 was carried out over the course of three steps (Figure

2.2). Well-dispersed and uniform CeO2 nanoparticles were first synthesized via a solvothermal

method, and subsequently, Pt nanocrystals were grown directly on the CeO2 surface.19 Finally, the

CeO2−Pt nanocrystal were coated with mesoporous silica and annealed in air to obtain

CeO2−[email protected],25 The as-synthesized nanoparticle was characterized by transmission electron

microscopy (TEM) and energy-dispersive X-ray spectroscopy (EDS), which clearly showed its

complex core−shell structure (Figure 2.2 and Figure 2.3). The Pt loading was further calculated to

be 5.14% from inductively coupled plasma atomic emission spectroscopy (ICP-AES) (Table 2.1).

The porosity of CeO2−Pt@mSiO2 was confirmed by nitrogen physisorption, while the pore size

distribution curve calculated from the adsorption branch of the isotherms exhibited a maximum at

Page 34: Conversion Chenlu Xie

20

2.4 nm. The Brunauer−Emmett−Teller surface area of CeO2−Pt@mSiO2 was calculated to be 236

m2 g−1 , indicating the highly mesoporous nature of the silica shell (Figure 2.4). Importantly, the

3D nature of CeO2−Pt@mSiO2 imparts a higher surface area than would be accessible in a bilayer

catalyst. Finally, the CeO2−Pt core maintained its original shape after annealing at 350 °C and the

hydroformylation reactions.(Figure A2.1). Thus, CeO2−Pt@ mSiO2 clearly possesses high thermal

stability imparted by the mesoporous SiO2 shell.

Figure 2.3. Elemental distribution of the 3D tandem catalyst. (A) Imaging of CeO2-Pt@mSiO2 via high-angle annular

dark-field scanning transmission electron microscopy. (B) Elemental mapping of CeO2-Pt@mSiO2 with energy

dispersive X-ray spectroscopy (EDS). Corresponding EDS elemental mapping for (C) Ce, (D) Pt, (E) Si, and (F) O,

respectively. Scale bar: 30 nm.

Figure 2.4. Porosity characterization of CeO2−Pt@mSiO2 nanoparticles calcined at 350 °C. (A) Nitrogen adsorption-

desorption isotherms. (B) Pore size distribution calculated from the adsorption branch of the isotherms.

Page 35: Conversion Chenlu Xie

21

Section 2.3.2 Catalytic Performance of CeO2−Pt@mSiO2 in the Hydroformylation of

Ethylene with Methanol

2.3.2.1. Catalytic Performance of Individual Interfaces in CeO2−Pt@mSiO2: Tandem

Reaction Is Sequentially Catalyzed

Prior to examining the tandem reaction facilitated by CeO2−Pt@ mSiO2, control experiments

were performed to understand the role of each interface in catalysis. The single interface catalysts

were synthesized using the same capping ligands and have the similar Pt loading amount of the

tandem catalysts. (Figure A2.2 and Table 2.1). Accordingly, the tandem hydroformylation of

ethylene with methanol was carried out at 150 °C over the two single interface catalysts CeO2−Pt

and Pt@mSiO2 (Figure 2.5). The CeO2−Pt catalyst exhibited obvious activity for methanol

decomposition (Figure 2.6), with a turnover frequency (TOF) of 4.2 × 10−3 s−1 per Pt atom at

150 °C. However, this catalyst exhibited a very low activity for ethylene hydroformylation to

propanal and instead was highly active for ethylene hydrogenation to ethane. The Pt@mSiO2

catalyst showed almost no activity for the tandem hydroformylation due to the fact it is not

catalytically active for methanol decomposition, which is the first step of the tandem process

(Figure 2.6). Gratifyingly, a much higher activity for propanal formation was achieved by

integrating the Pt/CeO2 and Pt/SiO2 interfaces into a tandem catalyst. Indeed, the TOF for ethylene

hydroformylation by CeO2−Pt@mSiO2 was determined to be 4.1 × 10−3 s−1 per Pt atom, which is

about 13 times greater than the TOF exhibited by CeO2−Pt alone (3.0 × 10−4 s−1 per Pt atom). This

significant enhancement in the hydroformylation activity with CeO2−Pt@mSiO2 clearly

demonstrates that the designed interfaces in the tandem catalyst can be used to carry out sequential

chemical reactions effectively and selectively.

Table 2.1. Pt loading amount for the tandem catalyst CeO2−Pt@mSiO2 and single interface catalysts according to

ICP-AES.

Catalyst Pt loading amount

CeO2−Pt@mSiO2 5.14%

CeO2−Pt 5.38%

Pt@mSiO2 5.78%

Page 36: Conversion Chenlu Xie

22

Figure 2.5. Ethylene hydroformylation with methanol carried out at 150 °C over CeO2-Pt@mSiO2 and single interface

catalysts CeO2-Pt and Pt@mSiO2. Methanol, ethylene, and helium partial pressures were 35 torr, 7.5 torr and 727.5

torr, respectively.

Figure 2.6. Methanol decomposition over 3D catalyst CeO2−Pt@mSiO2 and single interface catalysts: CeO2−Pt and

Pt@mSiO2. (Methanol is 35 Torr, He is 735 Torr, reaction temperatures are 150 °C, 170 °C, 190 °C)

Page 37: Conversion Chenlu Xie

23

2.3.2.2. Comparison of Single-Step Hydroformylation and Tandem Reaction: Synergy

between the Sequential Reactions.

The production of propanal via sequential chemical conversions at two different neighboring

metal-oxide interfaces in CeO2−Pt@mSiO2 illustrates the unique advantage of the tandem catalysis.

In order to unravel the underlying mechanism of propanal formation, we further studied the

“decoupled” single-step hydroformylation reaction as a control experiment, wherein ethylene was

directly reacted with CO and H2 present in stoichiometry equal to one equivalent of methanol. The

single-step ethylene hydroformylation and the tandem hydroformylation were examined over

CeO2−Pt@mSiO2 for a range of temperatures from 150 to 230 °C (Figure 2.7). As shown in Figure

2.7, the selectivity for propanal formation in the tandem reaction was much higher than for the

single-step hydroformylation for all temperatures. The highest selectivity achieved with tandem

hydroformylation was an impressive ∼50% at 150 °C compared to only 2.2% for the single-step

hydroformylation with CO and H2. Also considering that ethylene hydrogenation is much more

favorable than the ethylene hydroformylation on conventional Pt catalysts,17,19 intuitively ethane

should almost always be the dominant product in this reaction. Thus, the effective and highly

selective production of propanal with CeO2−Pt@mSiO2 suggests an important synergy between

the two sequential reactions facilitated with this catalyst.

Figure 2.7. Catalytic performance of CeO2-Pt@mSiO2 for the hydroformylation of ethylene with methanol at an

ethylene pressure of 7.5 torr. (A) Comparison of tandem hydroformylation of ethylene with methanol and single-step

hydroformylation with CO and H2. (B) The influence of temperature on tandem hydroformylation of ethylene. (C)

Catalytic performance of single-step hydroformylation under a deficiency of CO and H2 at 150 °C.

To gain more insights into this synergistic effect, we examined the temperature dependence of

the reaction with in situ methanol decomposition and found that the selectivity for propanal

formation decreased with increasing temperature. At 150 °C, the rates of methanol decomposition,

ethylene hydroformylation, and ethylene hydrogenation in the tandem process were found to be

identical, suggesting that all the CO and H2 formed in situ from methanol decomposition (1:2 CO

to H2 ratio) was simultaneously and fully consumed by reacting with ethylene, which yields

propanal at 50% selectivity without accumulating H2 and CO on the Pt surface. With elevated

temperatures, all three reactions proceeded more rapidly with relative rates as follows: methanol

decomposition > ethylene hydrogenation > hydroformylation. Consequently, the selectivity for

propanal declined because the rate of ethylene hydrogenation was higher than the

hydroformylation and also due to an accumulation of H2 that disfavored the production of propanal.

These results suggest that when the methanol decomposition, and thus H2 gas accumulation,

greatly exceeds ethylene hydroformylation, the selectivity of hydroformylation drops significantly.

Page 38: Conversion Chenlu Xie

24

In other words, a slower methanol decomposition rate favors propanal production and imparts

selectivity.

Notably, the high selectivity toward propanal could not be obtained in the single-step

hydroformylation even upon artificially creating a deficiency of CO and H2 to simulate the tandem

conditions. When a very small amount of CO and H2 (equivalent to the stoichiometric amount of

actual converted methanol in the tandem process) was introduced to the catalyst with ethylene,

ethane was the only product and the formation of propanal was less than the detection limit

(<0.01%) (Figure 2.4C and Table A2.1). This result led us to consider possible molecular-level

mechanisms behind the tandem process, which is likely distinct from the single-step

hydroformylation. In the case of tandem reaction, the methanol could decompose to form

intermediates H* and CHO* on the Pt/CeO2 interface,26, 28 which would be adsorbed on the

platinum surface and diffuse to the nearby Pt/SiO2 interface. At this interface, the intermediates

would then be fully consumed in the ethylene hydroformylation and hydrogenation reactions,

yielding the observed 50% propanal selectivity. Thus, we can conclude that the absence of H2

accumulation on the Pt surface is the key to the effective formation of propanal, which could be

achieved by slow methanol decomposition and the subsequent consumption of CO and H2 by

ethylene in the tandem process.

2.3.2.3. Effect of Ethylene on Methanol Decomposition: Another Aspect of the Synergy and

a Pathway toward Higher Selectivity.

We also investigated the influence of ethylene on the methanol decomposition and subsequent

hydroformylation. Notably, for all temperatures investigated, we found that methanol

decomposition was impeded by the presence of ethylene (Figure 2.8). For instance, at 150 °C the

TOF for methanol decomposition in the absence of ethylene was found to be ∼0.49 s−1 per Pt atom

(Figure 2.6). Upon introduction of ethylene, however, this TOF decreased significantly and

reached a minimum of 4.2 × 10−3 s −1 per Pt atom when the ratio of ethylene to methanol was 0.21.

Further addition of ethylene to achieve ethylene: methanol ratios >0.21 had no effect on the

methanol decomposition rate.

Figure 2.8. The inhibiting effect of ethylene on methanol decomposition over the CeO2-Pt@mSiO2 catalyst: changes

in methanol decomposition rate (purple squares), ethylene hydroformylation rate (green circles), and propanol

selectivity (empty blue squares) Methanol partial pressure was 35 torr and reaction temperature was 150 °C.

Page 39: Conversion Chenlu Xie

25

We attributed this inhibiting effect to the adsorption of ethylene on the Pt surface and

corresponding blockage of the active sites for methanol decomposition. To test this possibility, H2

gas was cofed to the catalyst with methanol and ethylene (Figure 2.9, panels A and C), which

resulted in a significant increase in the TOF for methanol decomposition. Indeed, the excess H2

reacted with adsorbed ethylene thereby opening up active sites and enhancing the methanol

decomposition rate. This data further supports the observation that variations in the selectivity and

TOF for the ethylene hydrogenation and hydroformylation reactions can be attributed to

accumulation of H2, from externally added and the accelerated decomposition of methanol. The

presence of ethylene conversely slows this decomposition, and upon achieving an

ethylene:methanol ratio > 0.21, the TOF for ethylene hydroformylation becomes identical with

methanol decomposition, yielding the propanal selectivity of 50%.

A similar inhibiting effect due to ethylene was observed for temperatures above 150 °C (Figure

2.8), and notably the rate of ethylene hydroformylation was significantly enhanced at these higher

temperatures. For instance, the hydroformylation TOF reached a maximum of 9.4 × 10−3 s −1 per

Pt atom at 230 °C, more than double the TOF achieved at 150 °C for the same ethylene: methanol

ratio. Because methanol decomposition was accelerated at higher temperatures, however, a larger

ethylene to-methanol ratio was necessary to achieve the maximum selectivity for propanal when

compared with the lower temperature data. For instance, at 230 °C the ethylene hydroformylation

reached its maximum activity and selectivity for an ethylene: methanol ratio of 0.4, compared with

the maximum ratio of 0.21 necessary at 150 °C. This observed reactivity with ethylene thus stands

as further support of the dual reaction synergy achieved with CeO2−Pt@mSiO2 catalysis.

Given this comprehensive understanding of the tandem hydroformylation, it is possible to tune

the reaction conditions to further increase the selectivity for propanal and the reaction rate

simultaneously. As hydroformylation involves CO, ethylene, and H2, increasing the partial

pressure of CO facilitates hydroformylation to compete favorably with hydrogenation and gives

higher aldehyde selectivity. As illustrated in Figure 2.9 (panels B and D), an initial increase in fed

CO results in significant enhancement of the TOFs for methanol decomposition and ethylene

hydroformylation, whereas the reaction rate of ethylene hydrogenation correspondingly decreased.

Upon cofeeding with 70 Torr of CO, the TOF for ethylene hydroformylation increased by a factor

of 3 and notably a selectivity as high as 80% could be achieved.

Page 40: Conversion Chenlu Xie

26

Figure 2.9. H2 and CO effect on the tandem hydroformylation of ethylene. (Top) The effect of co-feeding H2 or CO

on the tandem hydroformylation of ethylene. (A) Co-feeding with H2; (B) Co-feeding with CO. (Bottom) Propanal

selectivity of tandem hydroformylation. (C) Co-feeding with H2; (D) Co-feeding with CO.

Gratifyingly, we also found that the catalytic tandem hydroformylation reaction was also

applicable to other light olefin systems. When propylene or 1-butene were fed over the

CeO2−Pt@mSiO2 catalyst instead of ethylene, hydroformylation occurred to produce butyl

aldehyde or pentanaldehyde (Table 2.2). In both cases, the tandem reaction gave much better

selectivity compared to the single-step hydroformylation reaction where propylene and 1-butene

were reacted with CO and H2. The temperature difference for different alkene is due to the fact

that larger alkenes show greater inhibition of methanol decomposition based on the experiment

results, and the order of inhibition effect is ethylene < propylene < 1-butene ( Table 2.3). At 150 oC, no decomposition of methanol was observed when co-feeding with propylene and 1-butene

(see the column under 150 oC). Thus, in the tandem hydroformylation with propylene and 1-

butene, due to the greater inhibiting effect, higher reaction temperature (i.e., 190 oC for propylene

and 230 oC for 1-butene) was needed to trigger the methanol decomposition, which is the

prerequisite for the further hydrogenation and hydroformylation. Once the methanol

decomposition reaction was triggered, further increase of reaction temperature would lead to a

decline on aldehyde selectivity. As shown in the Table 2.3, the selectivity toward aldehyde in each

alkene was optimum (i.e., around 50%) at the temperature that is just enough to trigger the

methanol decomposition (i.e., 150 oC for ethylene, 190 oC for propylene and 230 oC for 1-butene),

where the rates of methanol decomposition were very slow among the three alkenes. These results

support the generality of a desired slow methanol decomposition rate favoring aldehyde formation.

Page 41: Conversion Chenlu Xie

27

Table 2.2. Tandem hydroformylation of different alkenes.

Reaction conditions a Aldehyde selectivity (%)

Alkene Temperature (oC) Tandem

Hydroformylation

Single-Step

Hydroformylation

Ethylene 150 48.9 2.2

Propylene 190 47.7 1.3

1-butene 230 48.2 0.9

a, Tandem alkene hydroformylation: 35 Torr of methanol, 7.5 Torr of alkene, and 727.5 Torr of helium. Single-step

hydroformylation with CO and H2: 35 Torr of CO, 70 Torr of H2, 7.5 Torr of alkene, and 657.5 Torr of helium.

Table 2.3. The inhibiting effect of different alkenes on methanol decomposition over CeO2−Pt@mSiO2 catalyst

Alkenes Methanol decomposition TOF (s−1) a

150 oC 190 oC 230 oC

No alkene 0.49 0.86 2.6

Ethylene 0.0042 0.02 0.15

Propylene Trace amount 0.0035 0.018

1-Butene Trace amount Trace amount 0.0023

a. Reaction condition: 35 Torr of methanol, 7.5 Torr of alkene, and 727.5 Torr of helium.

2.3.2.4 Proposed Reaction Mechanism

Combining the insights provided by preceding results, we proposed a mechanism for the

tandem reaction with methanol and ethylene catalyzed by CeO2−Pt@mSiO2 (Figure 2.10). In this

proposed mechanism, ethylene molecules are adsorbed on the platinum surface and occupy the

majority of active sites, hindering the dissociation of methanol on the Pt/CeO2 interface. At limited

active sites, methanol molecules will dissociate to form hydrogen species (Hads) and intermediate

products (e.g., CH3O*, CH2O*, CHO*, hereafter abbreviated as “CHOads species”) in microscopic

amounts,26,28−31 which are adsorbed on the surface. The presence of these “CHOads species” in

methanol decomposition has long been known.28−40 Both experimental and theoretical studies have

found that the methanol decomposition on platinum surface proceeds via methoxy (CH3O), as a

first intermediate, then by stepwise hydrogen abstraction via formaldehyde (CH2O), and then

formyl (CHO). The formation of these intermediates from methanol on the Pt surface has been

detected by energy electron loss spectroscopy (EELS),33−36,40 thermal desorption spectroscopy

(TDS),22−36,38,39 low-energy electron diffraction (LEED),33,40 infrared reflection−absorption

spectroscopy (IRAS),37 and X-ray photoelectron spectroscopy.28,40

Page 42: Conversion Chenlu Xie

28

Figure 2.10. Proposed reaction pathway for the tandem ethylene hydroformylation.

In our proposed tandem reaction pathway, the ratio of the Hads and CHOads intermediates could

be determined from the methanol decomposition stoichiometry. These adsorbed intermediates then

diffuse from the Pt/CeO2 interface to the nearby Pt/SiO2 interface, where they are consumed by

adsorbed ethylene. The ethylene then reacts with these CHOads species to generate propanal or

with Hads to produce ethane as a byproduct.12 At low temperatures, the CHOads and Hads species

are produced in very small amounts, and thus the adsorbed ethylene molecules are in great excess.

Consequently, we can assume each CHOads species and Hads are consumed by the adsorbed

ethylene and do not accumulate on the platinum surface, resulting in a propanal selectivity of

∼50%. With elevated temperatures, the coverage of CHOads and Hads on the Pt surface increases

along with the methanol decomposition rate, thereby increasing the opportunity for reactions

between ethylene and both Hads and the CHOads species. Thus, the reaction rates for

hydroformylation and hydrogenation should both increase, which is consistent with our results.

However, as the amounts of CHOads and Hads increase, it is no longer valid to assume that they are

consumed immediately by ethylene, and they likely begin to accumulate on the Pt surface.

Considering that ethylene hydrogenation is faster than hydroformylation, the overall

hydrogenation rate is higher than that of hydroformylation, which is responsible for the decline of

propanal selectivity. However, in the case of the single-step hydroformylation, the reaction

pathway is different. Macroscopic CO and H2 are introduced to the catalyst and adsorbed on the

Pt surface, which leads to an inevitable accumulation of Hads. As the ethylene hydrogenation is

more favorable than the ethylene hydroformylation reaction, ethane will always be the dominant

product.

Page 43: Conversion Chenlu Xie

29

Section 2.4 Conclusion

A stable 3D tandem catalyst CeO2−Pt@mSiO2 with well-defined catalytic interfaces was

developed, and its catalytic performance was studied in tandem alkene hydroformylation with

methanol. Importantly, the tandem ethylene hydroformylation exhibited greatly enhanced

propanal selectivity compared to the single-step ethylene hydroformylation with CO and H2, which

could further be improved by cofeeding with CO. This effective production of propanal results

from synergy between the two sequential chemical conversions facilitated by CeO2−Pt@mSiO2

and the altered reaction pathway, compared to the single-step reaction. Ultimately, this in-depth

study highlights the benefits of tandem catalysis and paves the way for further rational design of

complex nanostructured catalysts.

Section 2.5 References

(1) Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjaer, C. F.; Hummelshoj, J. S.; Dahl, S.;

Chorkendorff, I.; Norskov, J. K. Nat. Chem. 2014, 6, 320.

(2) Behrens, M.; Studt, F.; Kasatkin, I.; Kuhl, S.; Havecker, M.; Abild-Pedersen, F.; Zander, S.;

Girgsdies, F.; Kurr, P.; Kniep, B. L.; Tovar, M.; Fischer, R. W.; Norskov, J. K.; Schlogl, R. Science

2012, 336, 893.

(3) Jiao, F.; Li, J.; Pan, X.; Xiao, J.; Li, H.; Ma, H.; Wei, M.; Pan, Y.; Zhou, Z.; Li, M.; Miao, S.;

Li, J.; Zhu, Y.; Xiao, D.; He, T.; Yang, J.; Qi, F.; Fu, Q.; Bao, X. Science 2016, 351, 1065.

(4) An, K.; Alayoglu, S.; Musselwhite, N.; Plamthottam, S.; Melaet, G.; Lindeman, A. E.;

Somorjai, G. A. J. Am. Chem. Soc. 2013, 135, 16689.

(5) Schwab, G. M.; Koller, K. J. Am. Chem. Soc. 1968, 90, 3078.

(6) Fu, Q.; Saltsburg, H.; Flytzani-Stephanopoulos, M. Science 2003, 301, 935.

(7) Chen, G.; Zhao, Y.; Fu, G.; Duchesne, P. N.; Gu, L.; Zheng, Y.; Weng, X.; Chen, M.; Zhang,

P.; Pao, C.-W.; Lee, J.-F.; Zheng, N. Science 2014, 344, 495.

(8) Cargnello, M.; Doan-Nguyen, V. V. T.; Gordon, T. R.; Diaz, R. E.; Stach, E. A.; Gorte, R. J.;

Fornasiero, P.; Murray, C. B. Science 2013, 341, 771.

(9) Yamada, Y.; Tsung, C.-K.; Huang, W.; Huo, Z.; Habas, S. E.; Soejima, T.; Aliaga, C. E.;

Somorjai, G. A.; Yang, P. Nat. Chem. 2011, 3, 372.

(10) Ojima, I.; Tsai, C.-Y.; Tzamarioudaki, M.; Bonafoux, D. Organic Reactions; John Wiley &

Sons, Inc.: 2004.

(11) Dharmidhikari, S.; Abraham, M. A. J. Supercrit. Fluids. 2000, 18, 1.

(12) Naito, S.; Tanimoto, M. J. Chem. Soc., Chem. Commun. 1989, 1403.

(13) Hanaoka, T.; Arakawa, H.; Matsuzaki, T.; Sugi, Y.; Kanno, K.; Abe, Y. Catal. Today 2000, 58,

271.

(14) Gao, R.; Tan, C. D.; Baker, R. T. K. Catal. Today 2001, 65, 19.

(15) Takahashi, N.; Kobayashi, M. J. Catal. 1984, 85, 89.

Page 44: Conversion Chenlu Xie

30

(16) Arai, H.; Tominaga, H. J. Catal. 1982, 75, 188.

(17) Rioux, R. M.; Song, H.; Hoefelmeyer, J. D.; Yang, P.; Somorjai, G. A. J. Phys. Chem. B 2005,

109, 2192.

(18) Song, H.; Rioux, R. M.; Hoefelmeyer, J. D.; Komor, R.; Niesz, K.; Grass, M.; Yang, P.;

Somorjai, G. A. J. Am. Chem. Soc. 2006, 128, 3027.

(19) Tsung, C.-K.; Kuhn, J. N.; Huang, W.; Aliaga, C.; Hung, L.-I.; Somorjai, G. A.; Yang, P. J.

Am. Chem. Soc. 2009, 131, 5816.

(20) Somorjai, G. A.; Frei, H.; Park, J. Y. J. Am. Chem. Soc. 2009, 131, 16589.

(21) Joo, S. H.; Park, J. Y.; Tsung, C.-K.; Yamada, Y.; Yang, P.; Somorjai, G. A. Nat. Mater. 2009,

8, 126.

(22) Zaera, F., Somorjai, G. A., J. Am. Chem. Soc. 1984, 106, 2288-2293;

(23) Cremer, P. S., Su, X., Shen, Y. R., Somorjai, G. A., J. Am. Chem. Soc. 1996, 118, 2942-

2949;

(24) Zaera, F., J. Phys. Chem. 1990, 94, 5090-5095

(25) Croy, J. R.; Mostafa, S.; Liu, J.; Sohn, Y.; Heinrich, H.; Cuenya, B. R. Catal. Lett. 2007, 119,

209.

(26) Imamura, S.; Higashihara, T.; Saito, Y.; Aritani, H.; Kanai, H.; Matsumura, Y.; Tsuda, N.

Catal. Today 1999, 50, 369.

(27) Deng, Y.; Qi, D.; Deng, C.; Zhang, X.; Zhao, D. J. Am. Chem. Soc. 2008, 130, 28.

(28) Matolín, V.; Johánek, V.; Škoda, M.; Tsud, N.; Prince, K. C.; Skála, T.; Matolínová, I.

Langmuir 2010, 26, 13333.

(29) Cao, D.; Lu, G. Q.; Wieckowski, A.; Wasileski, S. A.; Neurock, M. J. Phys. Chem. B 2005,

109, 11622.

(30) Greeley, J.; Mavrikakis, M. J. Am. Chem. Soc. 2002, 124, 7193.

(31) Greeley, J.; Mavrikakis, M. J. Am. Chem. Soc. 2004, 126, 3910.

(32) Waszczuk, P.; Lu, G. Q.; Wieckowski, A.; Lu, C.; Rice, C.; Masel, R. I. Electrochim Acta

2002, 47, 3637.

(33) Wang, J.; Masel, R. I. J. Vac. Sci. Technol., A 1991, 9, 1879.

(34) Franaszczuk, K.; Herrero, E.; Zelenay, P.; Wieckowski, A.; Wang, J.; Masel, R. I. The Journal

of Physical Chemistry 1992, 96, 8509.

(35) Diekhöner, L.; Butler, D. A.; Baurichter, A.; Luntz, A. C. Surf. Sci. 1998, 409, 384.

(36) Wang, J.; Masel, R. I. Surf. Sci. 1991, 243, 199.

(37) Peremans, A.; Maseri, F.; Darville, J.; Gilles, J. M. Surf. Sci. 1990, 227, 73.

(38) Davis, J. L.; Barteau, M. A. Surf. Sci. 1987, 187, 387.

Page 45: Conversion Chenlu Xie

31

(39) Akhter, S.; White, J. M. Surf. Sci. 1986, 167, 101.

(40) Attard, G. A.; Chibane, K.; Ebert, H. D.; Parsons, R. Surf. Sci. 1989, 224, 311.

Section 2.6 Appendix

Section 2.6.1 Stability of Tandem Catalyst CeO2-Pt@miSO2

Figure A2.1. (A−D) TEM images of CeO2−Pt@mSiO2 particles after calcination at 350 °C and hydroformylation

reactions. Scale bar: 20 nm.

Page 46: Conversion Chenlu Xie

32

Section 2.6.2 TEM Images of Single Interface Catalysts.

Figure A2.2. (A) CeO2−Pt catalyst. (B) Pt@mSiO2 catalyst. All the catalysts were calcined at 350 °C for 1 hour in

static air to remove ligands.

Section 2.6.3 Detailed Data of Conversion and Selectivity of Each Step in the Tandem Process

and the Single Tandem Hydroformylation

Figure A2.3. Time evolution of methanol and ethylene conversions and propanal and ethane concentrations in an

effluent. Propanal yield = produced propanal/reactants fed in, ethane yield = produced ethane/reactants fed in. (A)

Tandem ethylene hydroformylation with methanol. Reaction condition: methanol 35 Torr, ethylene 7.5 Torr, reaction

temperature 150 °C. (B) Single-step hydroformylation with CO and H2. Reaction condition: ethylene 7.5 Torr, CO 35

Torr, H2 70 Torr, reaction temperature 150 °C.

Page 47: Conversion Chenlu Xie

33

Table A2.1. Conversion of methanol and ethylene of tandem hydroformylation, single-step hydroformylation and

tandem hydroformylation cofed with CO or H2.

Reaction conditions a Conversion

Date refer

to Methanol

(Torr)

Ethylene

(Torr)

H2

(Torr)

CO

(Torr)

Time

(h)

Methanol

(%)

Ethylene

(%)

- 7.5 70 35 5 - 87.2 Figure 4A

35 7.5 - - 15 2.1 9.8 Figure 4B

- 7.5 0.7 0.35 5 - Trace amount Figure 4C

- 7.5 3.5 1.75 5 - 4.3 Figure 4C

35 7.5 0.1 - 15 4.5 22.5 Figure 6A

35 7.5 0.2 - 15 10.7 41.6 Figure 6A

35 7.5 1 - 5 63.3 84.2 Figure 6A

35 7.5 1 - 15 100 100 Figure 6A

35 7.5 - 1 15 2.1 9.9 Figure 6B

35 7.5 - 10 15 3.0 13.9 Figure 6B

35 7.5 - 70 15 3.9 18.0 Figure 6B

a, Reaction conditions: 150 °C, total pressure is 1 atm.

Page 48: Conversion Chenlu Xie

34

Table A2.2 Selectivity of single-step ethylene hydroformylation with CO and H2.

a, Reaction conditions: 150 °C, total pressure is 1 atm. b, Entry 1 to 3 are the same results with Figure 2.7 C

Entry b Catalyst Reaction conditions a Products selectivity (%)

H2 CO Ethylene Ethane Propanal

1 CeO2-Pt@mSiO2 0.7 0.35 7.5 99.9 trace amount

2 CeO2-Pt@mSiO2 3.5 1.75 7.5 98.6 1.4

3 CeO2-Pt@mSiO2 70 35 7.5 97.8 2.2

4 Pt@mSiO2 0.7 0.35 7.5 99.9 trace amount

5 Pt@mSiO2 3.5 1.75 7.5 99.2 0.8

6 Pt@mSiO2 70 35 7.5 98.0 2.0

Page 49: Conversion Chenlu Xie

35

Chapter 3. Tandem Catalysis for CO2 Hydrogenation

to C2−C4 Hydrocarbons

Section 3.1 Introduction

Transformation of CO2 into transportable chemicals has been spurred by emerging

environmental issues and rising energy demands.1-5 Of particular interest is the synthesis of

hydrocarbons containing two to four carbon atoms (C2-C4 hydrocarbons), which are key chemical

feedstocks to synthesize a wide range of products such as polymers, solvents, drugs and detergents.

This is a challenging task because of the difficulties associated with the chemical inertness of CO2

and the competing formation of methane.6-8

Currently, the conversion of CO2 to C2-C4 hydrocarbons mainly relies on iron-based catalysts

(via the Fischer-Tropsch process(F-T process)) and composite oxide catalysts such as Cu-ZnO-

Al2O3/zeolite (via a methanol mediated pathway).9-15 However, these catalysts are mostly

synthesized through co-precipitation, impregnation or physical mixing, and contain multiple

components such as structural promoters. Thus, they exhibit large morphological variations and

great uncertainties on the spatial arrangements of active sites.10-12, 16 We believe the development

of structurally well-defined catalysts would not only facilitate discovery of new catalysts, but also

enable the fundamental study of reaction mechanisms to unravel principles for rational catalyst

design. Specifically, the importance of spatial control of catalytic interfaces was emphasized in a

recent development in tandem catalysis, where two metal-oxide interfaces in a single nanostructure

were employed to catalyze two sequential chemical reactions.17, 18

Herein, we demonstrate the rational design and controlled synthesis of a nanostructured catalyst

with well-defined architecture for CO2 conversion to C2-C4 hydrocarbons via tandem catalysis.

The designed catalyst CeO2-Pt@mSiO2-Co (mSiO2 denotes mesoporous silica) has two types of

metal-oxide interfaces that catalyze two sequential reactions. The CeO2/Pt interface converts CO2

and H2 into CO through the reverse water gas shift (RWGS) reaction, and the Co/mSiO2 interface

subsequently reacts the formed CO with H2 through the Fischer-Tropsch process. Thus, this

catalyst could carry out the CO2 conversion to C2-C4 hydrocarbons through a two-step tandem

process. As a result, a C2-C4 selectivity up to 60% of all hydrocarbons (carbon atom-based) through

this tandem process was achieved.

Section 3.2 Experimental Section

Section 3.2.1 Synthetic Methods and Materials

Chemicals. Cerium nitrate hexahydrate (99 % trace metals basis), poly(vinylpyrrolidone) (PVP,

Mw =360,000), PVP (Mw=29000), Tetradecyltrimethylammonium bromide (TTAB), cetrimonium

bromide (CTAB) (99 %), tetraethyl orthosilicate (TEOS) (99.999% trace metal basis), cobalt

carbonyl (moistened with hexanes, (hexanes 1-10%), >90% Co), oleic acid (99%), o-

dichlorobenzene (DCB) (anhydrous ,99%, DCB) were purchased from Sigma-Aldrich.

Ammonium hexachloroplatinate (IV) ((NH4)2PtCl6, Pt 43.4 % min) was purchased from Alfa

Aesar. Ethanol, ethylene glycol, hexanes, 2-propanol were purchased from Fisher Chemical.

Ammonia solution (28-30 %) was purchased from EMD Millipore. All chemicals were used as

received without further purification.

Page 50: Conversion Chenlu Xie

36

Synthesis of CeO2-Pt@mSiO2 nanoparticles (NPs). The CeO2-Pt@mSiO2 NPs were

synthesized following our previous report in Chapter 2 via a three-step procedure. (1). Synthesis

of CeO2 NPs: Cerium nitrate hexahydrate (0.85 g) was dissolved in the mixture solution of

deionized water (5 mL) and ethanol (5 mL). 30 ml of PVP (Mw = 360,000)-ethanol solution (60

mg/ml) was added to the solution. This reaction mixture was then transferred to a stainless-steel

autoclave and heated to 140 °C for 24 hours. The as-synthesized CeO2 nanoparticles were collected

by centrifugation (12000 rpm, 60 minutes) and then washed twice with water and ethanol and

stored in ethanol for further use; (2). Overgrowth of Pt NPs on as-synthesized CeO2 NPs: The pre-

synthesized CeO2 NPs (40 mg) were dispersed in 20 mL ethanol. To this solution, TTAB (36.8 mg)

and PVP (Mw=29000, 21.8 mg) dissolved in ethylene glycol (16 mL) were added. (NH4)2Pt(IV)Cl6

(9.75 mg) was dissolved into ethylene glycol (4 mL) at 80 °C in a 25 mL three-neck round flask

under Ar atmosphere. This Pt precursor solution was then mixed with the CeO2 NPs/ethanol

solution and heated to 140 °C for 6 hours in a stainless-steel autoclave. The as-synthesized CeO2-

Pt NPs were separated by centrifugation (12000 rpm, 45 minutes) and re-dispersed in 40 mL

deionized water; (3). Synthesis of CeO2-Pt@mSiO2: The CeO2-Pt@mSiO2 core-shell NPs were

prepared by a sol-gel method. 225mg CTAB in ethanol (30 mL) was added to the solution of pre-

synthesized CeO2-Pt (40 mg in 45 mL deionized water). An ammonia solution (0.2 mL) was added

to the above solution. Then a controlled amount of 1 vol % TEOS diluted with ethanol was slowly

added under continuous magnetic stirring at room temperature. After 6 hours, the as-synthesized

CeO2-Pt@SiO2 nanoparticles were obtained by centrifugation (6000 rpm, 5 minutes). The product

was calcined at 350 °C for 1 hour in static air to remove CTAB template to generate CeO2-

Pt@mSiO2 particles.

Synthesis of oleic acid-capped Cobalt NPs. The monodisperse cobalt NPs were synthesized

by a reported method with some modification.2 In a typical synthesis, oleic acid (130 mg) in a 250

mL round bottom flask was evacuated for 30 minutes, and anhydrous o-dichlorobenzene (15 ml)

was added under Argon. The solution was heated at 172°C for 30 minutes under magnetic stirring.

Then Co2(CO)8 (512 mg) dissolved in o-dichlorobenzene (3 ml) was quickly injected into this

solution. This colloidal solution was kept at 172°C for 20 min prior to cooling down to room

temperature. The particles were precipitated with 2-propanol and by centrifugation ( 12000 rpm,

5 mins) and then re-dispersed in hexane for further use.

Synthesis of CeO2-Pt@mSiO2-Co NPs. The CeO2-Pt@mSiO2 nanoparticle powders (100 mg)

were first dispersed in hexanes (20 mL) and sonicated for 15 minutes. The as-synthesized oleic

acid-capped Co NPs were dissolved in hexanes to form a solution with a concentration of ~ 0.5

mg/mL. Then the Co NPs/hexanes solution was added to the CeO2-Pt@mSiO2 solution and stirred

for 3 hours before the solids were centrifuged. The amount of Co NPs/hexanes solution added was

calculated from the desired loading of Co. The product was then dried at 100°C and calcined in

static air at 350 °C for 1 hour to remove the organic ligands.

Section 3.2.2 Characterization and Catalytic Study of CeO2-Pt@mSiO2-Co Tandem

Catalysts

Characterization. The morphology of CeO2-Pt@mSiO2-Co NPs was analyzed using

transmission electron microscopy (TEM) on a Hitachi H7650 and on a FEI Tecnai F20. High-angle

annular dark-field scanning transmission electron microscopy (HAADF-STEM) and energy-

dispersive X-ray spectroscopy (EDS) mapping were carried out with an FEI TitanX 60-300, which

provided the elemental distribution of the catalysts. Surface area and pore size distribution of the

Page 51: Conversion Chenlu Xie

37

catalyst were obtained by nitrogen physisorption experiments, which were carried out on a

Quantachrome Autosorb-1 analyser. Platinum and cobalt quantitative analysis by inductively

coupled plasma atomic emission spectroscopy (ICP-AES) was carried out on a PerkinElmer

optical emission spectrometer (Optima 7000 DV). Before ICP-AES measurement, catalyst was

digested in aqua regia for 24 hours and diluted in deioninzed water. The clear solution for ICP-

AES measurement was obtained by centrifuging at 4000 rpm to remove the sediment.

Catalytic Reaction Testing. Catalytic reactions were performed using a commercial tubular

plug flow reactor from Parr Company (i.d. 7 mm). The catalyst sample (50 mg) was retained

between plugs of quartz wool, and the reactor temperature monitored with a J type thermocouple.

The catalysts were conditioned at 350 °C in flowing 50 vol. % H2 in balance with He for 1 hour

for in situ reduction prior to reaction. The reaction was carried out at 90 psi of CO2: H2: He (CO2

to H2 ratio was tuned from 0.3 to 7) with a total flow of 42 sccm (standard cubic centimeters per

minute), which was delivered carefully via calibrated Brooks mass flow controllers. The gas hourly

space velocity was kept constant at 50,400 cm3/h·gcal. Products were analyzed by an online gas

chromatograph (Shimadzu 2010), which was equipped with a thermal conductivity detector (TCD)

and a flame ionization detector (FID). Calibration of retention times and peak intensities was made

directly with all observed reactants and products.

CO2 conversion was calculated on a carbon atom basis:

𝐶𝑜𝑛𝐶𝑂2 =𝐶𝑂2(𝑖𝑛𝑙𝑒𝑡) − 𝐶𝑂2(𝑜𝑢𝑡𝑙𝑒𝑡)

𝐶𝑂2(𝑖𝑛𝑙𝑒𝑡)100%

Where CO2 (inlet) and CO2 (outlet) represent moles of CO2 at the inlet and outlet, respectively.

CO selectivity (𝑆𝑒𝑙𝐶𝑂) was calculated according to:

𝑆𝑒𝑙𝐶𝑂 =𝐶𝑂(𝑜𝑢𝑡𝑙𝑒𝑡)

𝐶𝑂2(𝑖𝑛𝑙𝑒𝑡) − 𝐶𝑂2(𝑜𝑢𝑡𝑙𝑒𝑡)100%

Where CO (𝑜𝑢𝑡𝑙𝑒𝑡) denotes moles of CO at the outlet.

The selectivity of individual hydrocarbon CnHm (𝑆𝑒𝑙𝐶𝑛𝐻𝑚) was obtained according to:

𝑆𝑒𝑙𝐶𝑛𝐻𝑚 =𝑛𝐶𝑛𝐻𝑚𝑜𝑢𝑡𝑙𝑒𝑡

∑ 𝑛𝐶𝑛𝐻𝑚𝑜𝑢𝑡𝑙𝑒𝑡𝑛1

100%

The carbon balance was over 95%.

Section 3.3 Results and Discussion

Section 3.3.1 Structure Design and Synthesis of the CeO2-Pt@mSiO2-Co Tandem Catalysts

To create these two interfaces, the catalyst was designed with a CeO2-Pt core and a mesoporous

silica shell, which is further decorated with cobalt nanoparticles. Considering the compatibility of

synthetic conditions of different types of nanoparticles, the integration of all the four components

in one single nanoparticle is challenging and requires an elegant synthetic design. In particular,

synthesis of monodisperse cobalt nanoparticles could only be conducted under hydrophobic

conditions as they are prone to oxidation in aqueous solution, whereas the silica shell is typically

synthesized in aqueous solution.19-21 This incompatibility requires the preparation of monodisperse

Page 52: Conversion Chenlu Xie

38

CeO2-Pt@mSiO2, and subsequent homogeneous loading of cobalt nanoparticles on the silica

surface.

Figure 3.1. Synthesis and characterization of the CeO2-Pt@mSiO2-Co tandem catalyst. (a) Schematic of synthetic

process. TEM images of each step: (b, c) CeO2 nanoparticles, (d, e) overgrowth of Pt nanoparticles on CeO2, (f, g)

silica shell coating on CeO2-Pt composite nanoparticles, (h, i) deposition of Co nanoparticles on CeO2-Pt@mSiO2, (j)

scaled up preparation of CeO2-Pt@mSiO2-Co nanoparticles. One-pot synthesis can yield 300 mg catalyst.

The optimized synthesis involves four steps. The first step is to synthesize well-dispersed and

uniform CeO2 nanoparticles. Considering that the subsequent silica coating step is typically

performed in aqueous solution, the CeO2 nanoparticles need to be dispersible in aqueous media.

Thus, we carried out the synthesis of CeO2 in ethanol and water solution with

poly(vinylpyrrolidone) (PVP) as a capping ligand.18 This procedure enabled the production of

uniform CeO2 nanoparticles with very good dispersity, which is confirmed by transmission

electron microscopy (TEM) (Figure 3.1(b), (c), Figure 3.2). The size of CeO2 could be tuned by

changing the ratio between ethanol and water. A higher ethanol/water ratio yields CeO2

nanoparticles with smaller size (Figure 3.2). In order to gain a higher dispersity of Pt nanoparticles

on the CeO2 support, a smaller CeO2 nanoparticle size around 35 nm was chosen. The second step

is to load Pt nanoparticles onto the pre-synthesized CeO2 nanoparticles by Pt overgrowth.18, 22

Tetradecyltrimethylammonium bromide (TTAB) and PVP were used as capping ligands to give 3

nm loaded Pt NPs (Figure 3.1(d), (e)). Compared to the electrostatic absorption method, where the

pre-synthesized Pt nanoparticles were absorbed on the CeO2 surface, the overgrowth method gave

stronger interaction between Pt and CeO2, which helped maintain the structure of the nanocrystals

in the silica coating step. Moreover, the loading amount of Pt could be easily tuned by adding

different amount of (NH4)2Pt(IV)Cl6 (Figure 3.3). Subsequently, a sol-gel approach was adopted

Page 53: Conversion Chenlu Xie

39

to coat a mesoporous SiO2 shell on the CeO2-Pt core.18, 20, 23-24 The as-synthesized CeO2-Pt@SiO2

was calcined at 350 °C in air to remove the cetrimonium bromide (CTAB) template to generate

the mesopores and clean interfaces. The average thickness of the silica layer surrounding the CeO2-

Pt core was 25 nm (Figure 3.1(f), (g)). With different amount of tetraethyl orthosilicate (TEOS)

added, the thickness of the mesoporous silicon shell can be tuned (Figure 3.4). Finally,

homogeneous loading of cobalt NPs on the silica shell was achieved by an approach utilizing the

weak interactions between cobalt and silica in aprotic solvents.25 Monodisperse Co nanoparticles

were synthesized first by decomposing dicobalt carbonyl under the protection of oleic acid, which

was dispersed in hexanes (Figure 3.5).26 The as-synthesized CeO2-Pt@mSiO2 powder was also

dispersed in hexanes, and the cobalt-hexanes solution was added slowly under stirring to gain a

uniform distribution of cobalt nanoparticles on the silica shell. The dispersion of Co nanoparticles

was then locked in place by calcination under air at 350 ºC (Figure 3.1(h), (i)), which

simultaneously removed the oleic acid ligands.25, 27 This four-step synthesis of CeO2-Pt@mSiO2-

Co can be readily scaled up to produce the tandem catalysts at gram scale (Figure 3.1(j)). This

highly tunable and versatile synthetic protocol can be generalized toward the synthesis of other

systems with multiple metal-oxide interfaces, which paves the way for the development of other

sophisticated multifunctional catalysts for multi-step chemical reactions.

Figure 3.2. Tuning the size of CeO2 NPs by changing the ethanol and water ratio. TEM images (scale bar: 100 nm)

of CeO2 NPs with different ethanol and water amount: (a) 10 mL H2O and 30 mL ethanol and (b) corresponding

diameter distribution statistics; (c) 5 mL H2O and 35 mL ethanol and (d) corresponding diameter distribution statistics.

Page 54: Conversion Chenlu Xie

40

Figure 3.3. TEM images of CeO2-Pt NPs with Pt loading amount around (a), 1%; (b), 5% and (c), 10%. The loading

amount of Pt was calculated based on the mass of CeO2-Pt NPs. Scale bar: 20 nm.

Figure 3.4. Silica shell thickness tuning of CeO2-Pt@mSiO2 NPs by changing the added volume of TEOS. TEM

images of CeO2-Pt@mSiO2 when the added volume of TEOS is (a) 25 µL, scale bar: 20 nm; (b) 50 µL, scale bar: 20

nm and (c) 120 µL, scale bar: 50 nm.

Figure 3.5. TEM image of monodisperse cobalt NPs. Scale bar: 20 nm.

Page 55: Conversion Chenlu Xie

41

High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM)

and the corresponding energy-dispersive X-ray spectroscopy (EDS) mapping further confirmed

the elemental distribution and structure of the catalyst (Figure 3.6). The Pt and Co loading were

determined to be 4.3% and 5.8 % (Table 3.1), respectively, from inductively coupled plasma

atomic emission spectroscopy (ICP-AES).

Figure 3.6. Elemental distribution of the catalyst. (a) Imaging of CeO2-Pt@mSiO2-Co by high-angle annular dark-

field scanning transmission electron microscopy. (b) Elemental mapping of CeO2−Pt@mSiO2−Co with energy

dispersive X-ray spectroscopy (EDS). Corresponding EDS elemental mapping for (c) Ce, (d) Pt, (e) O, (f) Si, and (g)

Co, respectively. Scale bar: 20 nm.

Table 3.1. Pt and Co loading amount for the catalysts used for CO2 hydrogenation according to ICP-AES.

Catalyst Pt loading amount (%) Co loading amount (%)

CeO2−Pt@mSiO2-Co 4.3 5.8

CeO2−Pt@mSiO2 4.4 0

CeO2@mSiO2−Co 0 5.3

Physical mixture 4.4 5.3

Page 56: Conversion Chenlu Xie

42

Section 3.3.2 Catalytic Performance for CO2 hydrogenation

The catalytic performance of the tandem catalyst for CO2 hydrogenation was examined under a

series of temperatures with a pressure of 90 psi and a H2/CO2 ratio of 3. To monitor the role of

each interface individually, we prepared single-interface catalysts CeO2-Pt@mSiO2 and

CeO2@mSiO2-Co and tested their catalytic performance under the same conditions. Pt and Co

loading amounts of these catalysts were controlled to be identical to the tandem catalyst (Table

3.1). It is known that Pt loaded on CeO2 shows high activity and selectivity towards CO2

hydrogenation to produce CO through the RWGS reaction,28, 29 while supported Co catalysts are

widely used for hydrocarbons production from CO via the F-T process.30-34 However, supported

Co catalysts were reported to be almost only active for methane formation when replacing CO by

CO2 and a high CO partial pressure is necessary for the production of hydrocarbons beyond

methane.32, 35-37 As shown in Figure 3.7(a) and Figure 3.8, the CeO2-Pt@mSiO2 catalyst indeed

produced CO with excellent selectivity (>99%) at all temperatures, which substantiated that CO2

and H2 could be converted into CO via RWGS reaction on Pt/CeO2 interface in the tandem catalyst.

As to the CeO2@mSiO2-Co catalyst, CO2 hydrogenation over the catalyst led to methane as the

only hydrocarbon product at 250 ºC (Figure 3.7(a)). With elevated temperatures, methane was still

the dominant product with selectivity > 99%, with a very small amount of C2-C4 hydrocarbon

produced (Figure 3.9).

Figure 3.7. Catalytic reaction study on the single-interface catalysts, tandem catalyst and the physical mixture. (a)

Catalytic performance of single-interface catalysts CeO2-Pt@mSiO2 and CeO2@mSiO2-Co, physical mixture catalyst

and tandem catalyst CeO2-Pt@mSiO2-Co. (H2/CO2 ratio is 3, reaction temperature is 250°C). (b) CO2 conversion and

hydrocarbons distribution at different H2/CO2 ratios over the tandem catalyst at 250°C.

Page 57: Conversion Chenlu Xie

43

Figure 3.8. Catalytic performance of CeO2-Pt@mSiO2 under different temperatures. (a) products distribution, (b) CO2

conversion at H2/CO2 ratio of 3.0 and total pressure of 90 psi.

Figure 3.9. Catalytic performance of CeO2@mSiO2-Co under different temperatures. (a) hydrocarbons distribution

(left axes) and CO selectivity (right axes). (b) CO2 conversion at H2/CO2 ratio of 3.0 and total pressure of 90 psi.

Page 58: Conversion Chenlu Xie

44

Figure 3.10. Catalytic performance of tandem catalyst CeO2-Pt@mSiO2-Co under different temperatures. (a)

hydrocarbons distribution (left axes) and CO selectivity (right axes). (b) CO2 conversion at H2/CO2 ratio of 3.0 and

total pressure of 90 psi.

Upon controlled integration of Pt/CeO2 interface and Co/mSiO2 interface into a tandem catalyst,

the products shifted and the formation of C2-C4 hydrocarbons was clearly observed with a

selectivity of 40% whereas methane selectivity dropped to 60% at 250 ºC (Figure 3.7(a)). At

increased temperatures, the selectivity toward C2-C4 hydrocarbons decreased but was still well

above the C2-C4 hydrocarbons production over the single- interface catalysts (Figure 3.8, Figure

3.9 and Figure 3.10). The decline of selectivity toward C2-C4 hydrocarbons under higher

temperature is due to the more favorable methanation of CO at higher temperature.38 This

outstanding selectivity toward C2-C4, not observed for the single-interface catalysts, suggests that

CO2 hydrogenation to C2-C4 hydrocarbons on the CeO2-Pt@mSiO2-Co catalyst undergoes a

tandem process. The uniqueness of the tandem catalyst is further confirmed by the comparison

with a physical mixture of Pt-CeO2 and Co-SiO2, which produced methane as a primary product

and only a small amount of C2-C4 hydrocarbons at all temperatures (Figure 3.7(a) and Figure 3.11).

As the Co catalysts produce mainly CH4 from CO2 and a high CO/CO2 ratio is necessary for C2-

C4 hydrocarbons formation,32, 35-37 we speculated that favorable formation of C2-C4 in the tandem

system could be attributed to the locally high CO partial pressure at the Co/mSiO2 interface due to

the well-controlled spatial arrangement of Pt/CeO2 and cobalt nanocrystals. Considering the well-

defined core-shell structure and confinement of Pt/CeO2 interface within the mesoporous silica

shell, all produced CO molecules on Pt/CeO2 interface can be transported to the neighboring cobalt

surface before diffusing out of the shell. Moreover, CO molecules have shown a higher sticking

probability to cobalt than CO2 and could be strongly adsorbed on a cobalt surface.39 Consequently,

a CO-rich local environment at the Co/mSiO2 interface was created, which favors the production

of C2-C4 hydrocarbons. In the case of a physical mixture, meanwhile, the uncontrolled spatial

arrangement of the Pt-CeO2 and Co-SiO2 interfaces resulted in a very low chance for CO from Pt-

CeO2 to be involved in the second reaction. Thus, the low CO partial pressure on the Co/mSiO2

interface in the physical mixture catalysts resulted in low selectivity toward C2-C4 hydrocarbons.

In accordance with our hypothesis, any measures that could further increase the localized CO

partial pressure at Co/SiO2 interface should benefit the selectivity of C2-C4 hydrocarbons, for

example, decreasing the H2/CO2 ratio. This in turn would decrease the H2/CO ratio for the second

Page 59: Conversion Chenlu Xie

45

F-T reaction, and thus lead to a more localized CO environment in Co/SiO2 interface. As

demonstrated in Figure 3.7(b), with the H2/CO2 ratio decreased from 7.0 to 0.3, the selectivity

toward C2-C4 hydrocarbons increased from 23% to 59%, while the selectivity of methane dropped

from 77% to 41%.

Figure 3.11. Catalytic performance of physical mixture of CeO2-Pt@mSiO2 and CeO2@mSiO2-Co under different

temperatures. (a) hydrocarbons distribution (left axes) and CO selectivity (right axes). (b) CO2 conversion at H2/CO2

ratio of 3.0 and total pressure of 90 psi.

The harsh reaction conditions employed to study these reactions prompted us to examine the

stability of the catalyst and its performance. As shown in Figure 3.12, Figure 3.13 and Figure 3.14,

the CeO2-Pt@mSiO2-Co tandem catalyst produced no significant change in catalytic activity and

product selectivity while running the catalysts for up to 40 hours. Moreover, the tested catalysts

were also evaluated by TEM, and no obvious morphology change was observed (shown in Figure

3.14), indicating good structural and chemical stability of the tandem catalysts.

Figure 3.12. Stability test of CeO2-Pt@mSiO2-Co catalyst with H2/CO2 of 3.0 at 250 °C.

Page 60: Conversion Chenlu Xie

46

Figure 3.13. Stability test of CO selectivity over the tandem catalyst CeO2-Pt@mSiO2-Co at temperature 250 °C and

H2/CO2 ratio of 3.0 and total pressure of 90 psi.

Figure 3.14. TEM images of tandem catalyst CeO2-Pt@mSiO2-Co after reaction. Scale bar: 20 nm.

Section 3.4 Conclusion

In conclusion, we developed a highly tunable method to synthesize a well-defined

nanostructured catalyst CeO2-Pt@mSiO2-Co for the selective production of C2-C4 hydrocarbons

from CO2. This catalyst achieved a selectivity of 60% towards C2-C4 hydrocarbons with two

interfaces Pt/CeO2 and Co/SiO2 in close proximity, catalyzing the RWGS reaction and the F-T

reaction respectively. This high C2-C4 hydrocarbons selectivity is attributed to the unique spatial

arrangement of two metal-oxide interfaces, which creates local environments conducive for

multistep reactions that a physical mixture fails to achieve. Our synthetic protocol offers a highly

generalizable method to integrate different metal-oxide interfaces for design and synthesis of next

generation nanostructured catalysts. These advances give impetus to the rational design and

development of high performance, multifunctional catalysts for multiple-step chemical

conversions.

Section 3.5 References

(1) Dresselhaus, M. S.; Thomas, I. L. Nature 2001, 414, 332-337.

Page 61: Conversion Chenlu Xie

47

(2) Lewis, N. S.; Nocera, D. G. Proc. Natl. Acad. Sci. U. S. A 2006, 103, 15729-15735.

(3) Aresta, M., Carbon Dioxide: Utilization Options to Reduce its Accumulation in the

Atmosphere. In Carbon Dioxide as Chemical Feedstock, Wiley-VCH Verlag GmbH & Co. KGaA:

2010; pp 1-13.

(4) Graciani, J.; Mudiyanselage, K.; Xu, F.; Baber, A. E.; Evans, J.; Senanayake, S. D.; Stacchiola,

D. J.; Liu, P.; Hrbek, J.; Sanz, J. F.; Rodriguez, J. A. Science 2014, 345, 546-550.

(5) Li, C.-S.; Melaet, G.; Ralston, W. T.; An, K.; Brooks, C.; Ye, Y.; Liu, Y.-S.; Zhu, J.; Guo, J.;

Alayoglu, S.; Somorjai, G. A. Nat. Commun. 2015, 6, 6538.

(6) Rodemerck, U.; Holeňa, M.; Wagner, E.; Smejkal, Q.; Barkschat, A.; Baerns, M.

ChemCatChem 2013, 5, 1948-1955.

(7) Müller, K.; Mokrushina, L.; Arlt, W. Chem. Ing. Tech. 2014, 86, 497-503.

(8) Dorner, R. W.; Hardy, D. R.; Williams, F. W.; Willauer, H. D. Energy Environ. Sci. 2010, 3,

884-890.

(9) Sai Prasad, P. S.; Bae, J. W.; Jun, K.-W.; Lee, K.-W. Catal. Surv. Asia 2008, 12, 170-183.

(10) Dorner, R. W.; Hardy, D. R.; Williams, F. W.; Willauer, H. D. Appl. Catal., A 2010, 373,

112-121.

(11) Kishan, G.; Lee, M.-W.; Nam, S.-S.; Choi, M.-J.; Lee, K.-W. Catal. Lett. 1998, 56, 215-219.

(12) Fujiwara, M.; Kieffer, R.; Ando, H.; Souma, Y. Appl. Catal., A 1995, 121, 113-124.

(13) Jeon, J.-K.; Jeong, K.-E.; Park, Y.-K.; Ihm, S.-K. Appl. Catal., A 1995, 124, 91-106.

(14) Satthawong, R.; Koizumi, N.; Song, C.; Prasassarakich, P. Catal. Today 2015, 251, 34-40.

(15) Niemelä, M.; Nokkosmäki, M. Catal. Today 2005, 100, 269-274.

(16) Riedel, T.; Schulz, H.; Schaub, G.; Jun, K.-W.; Hwang, J.-S.; Lee, K.-W. Top. Catal. 2003,

26, 41-54.

(17) Yamada, Y.; Tsung, C.-K.; Huang, W.; Huo, Z.; Habas, S. E.; Soejima, T.; Aliaga, C. E.;

Somorjai, G. A.; Yang, P. Nat. Chem. 2011, 3, 372-376.

(18) Su, J.; Xie, C.; Chen, C.; Yu, Y.; Kennedy, G.; Somorjai, G. A.; Yang, P. J. Am. Chem. Soc.

2016, 138, 11568-11574.

(19) Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Nature 1992, 359,

710-712.

(20) Joo, S. H.; Park, J. Y.; Tsung, C.-K.; Yamada, Y.; Yang, P.; Somorjai, G. A. Nat. Mater. 2009,

8, 126-131.

(21) Qi, Z.; Xiao, C.; Liu, C.; Goh, T. W.; Zhou, L.; Maligal-Ganesh, R.; Pei, Y.; Li, X.; Curtiss,

L. A.; Huang, W. J. Am. Chem. Soc. 2017, 139, 4762-4768.

Page 62: Conversion Chenlu Xie

48

(22) Tsung, C.-K.; Kuhn, J. N.; Huang, W.; Aliaga, C.; Hung, L.-I.; Somorjai, G. A.; Yang, P. J.

Am. Chem. Soc. 2009, 131, 5816-5822.

(23) Zhao, E. W.; Maligal-Ganesh, R.; Xiao, C.; Goh, T.-W.; Qi, Z.; Pei, Y.; Hagelin-Weaver, H.

E.; Huang, W.; Bowers, C. R. Angew. Chem. Int. Ed. 2017, 56, 3925-3929.

(24) Maligal-Ganesh, R. V.; Xiao, C.; Goh, T. W.; Wang, L.-L.; Gustafson, J.; Pei, Y.; Qi, Z.;

Johnson, D. D.; Zhang, S.; Tao, F.; Huang, W. ACS Catal. 2016, 6, 1754-1763.

(25) Zheng, N.; Stucky, G. D. J. Am. Chem. Soc. 2006, 128, 14278-14280.

(26) Iablokov, V.; Beaumont, S. K.; Alayoglu, S.; Pushkarev, V. V.; Specht, C.; Gao, J.; Alivisatos,

A. P.; Kruse, N.; Somorjai, G. A. Nano Lett. 2012, 12, 3091-3096.

(27) Ma, G.; Binder, A.; Chi, M.; Liu, C.; Jin, R.; Jiang, D.-e.; Fan, J.; Dai, S. Chem. Commun.

2012, 48, 11413-11415.

(28) Fu, Q.; Saltsburg, H.; Flytzani-Stephanopoulos, M. Science 2003, 301, 935-938.

(29) Goguet, A.; Meunier, F. C.; Tibiletti, D.; Breen, J. P.; Burch, R. J. Phys. Chem. B 2004, 108,

20240-20246.

(30) Iglesia, E. Appl. Catal., A 1997, 161, 59-78.

(31) den Breejen, J. P.; Radstake, P. B.; Bezemer, G. L.; Bitter, J. H.; Frøseth, V.; Holmen, A.;

Jong, K. P. d. J. Am. Chem. Soc. 2009, 131, 7197-7203.

(32) Melaet, G.; Ralston, W. T.; Li, C.-S.; Alayoglu, S.; An, K.; Musselwhite, N.; Kalkan, B.;

Somorjai, G. A. J. Am. Chem. Soc. 2014, 136, 2260-2263.

(33) Jacobs, G.; Das, T. K.; Zhang, Y.; Li, J.; Racoillet, G.; Davis, B. H. Appl. Catal., A 2002, 233,

263-281.

(34) Khodakov, A. Y.; Chu, W.; Fongarland, P. Chem. Rev. 2007, 107, 1692-1744.

(35) Riedel, T.; Claeys, M.; Schulz, H.; Schaub, G.; Nam, S.-S.; Jun, K.-W.; Choi, M.-J.; Kishan,

G.; Lee, K.-W. Appl. Catal., A 1999, 186, 201-213.

(36) Zhang, Y.; Jacobs, G.; Sparks, D. E.; Dry, M. E.; Davis, B. H. Catal. Today 2002, 71, 411-

418.

(37) Akin, A. N.; Ataman, M.; Aksoylu, A. E.; Önsan, Z. I. React. Kinet. Catal. Lett. 2002, 76,

265-270.

(38) Dry, M. E. Catal. Today 2002, 71, 227-241.

(39) Visconti, C. G.; Lietti, L.; Tronconi, E.; Forzatti, P.; Zennaro, R.; Finocchio, E. Appl. Catal.,

A 2009, 355, 61-68.

Page 63: Conversion Chenlu Xie

49

Charpter 4. Controlling Atomic Ordering of AuCu

Nanoparticles for Electrochemical CO2 Reduction

Section 4.1 Introduction

Multimetallic nanoparticles (NPs) have great tunability in their structure and properties because

of their widely tunable compositions. In this regard, they have been extensively studied for various

applications, where their properties overcome the drawbacks of single-component systems.1−3

Beyond simply controlling the ratio between the elements where each is randomly distributed,

precise positioning of atoms in a crystal lattice becomes possible when there exists a

thermodynamically accessible phase. This results in a distinction between a disordered alloy and

an ordered intermetallic structure, and the transition between the two different phases is the so-

called order−disorder transformation.4

Recent works on ordered intermetallic nanoparticles have focused on developing synthetic

strategies to promote ordering transformations from the initial alloy and investigating their

structure and properties for applications in magnetism and catalysis.5−12 Structurally ordered Pt-

based (with Fe, Co, Cu, etc.) NPs have been under investigation in the past few years, as the

ordered structure has been known to offer enhanced activity and durability for the oxygen

reduction reaction (ORR).6−9,12 On the other hand, for the other well-known bimetallic system of

Au−Cu, discovery and use of its enhanced properties, attributable to its having an ordered lattice,

have not been well studied, with only a few reports present regarding the synthetic approach to

create ordered NPs.5,10,11,13,14 Recently, it has been shown that Au−Cu bimetallic systems exhibit

interesting electrocatalytic properties for CO2 reduction controlled by their composition.15,16 In this

work, we investigate the effects of order−disorder transformation in AuCu NPs on electrochemical

CO2 reduction. We show that the ordering transformation of a disordered AuCu (1:1 atomic ratio)

NP turns an inactive catalyst, which mainly generates hydrogen gas, to an active catalyst for

reducing CO2. This was found through systematic control of the ordering degree of AuCu NPs and

observation of its effect on the catalytic behavior. We perform detailed structural investigations of

ordered AuCu NPs down to the atomic level to find ordered lattice configurations, together with a

few-atoms-thick Au layer on the outer surface. Furthermore, thermochemical DFT calculations

suggest that the enhanced CO2 reduction activities observed for ordered intermetallic AuCu NPs

arise from the formation of compressively strained Au overlayers.

Section 4.2 Experimental Section

Section 4.2.1 Synthetic Methods and Materials

Chemicals. Copper(II) acetate monohydrate (98%), oleic acid (90%), tri-n-octylamine (TOA,

98%), oleylamine (70%), borane tertbutylamine complex (TBAB, 97%), gold(III) chloride

trihydrate (HAuCl4·3H2O, 49% Au basis), and o-xylene were purchased from Sigma-Aldrich.

Potassium carbonate (99.997%) was purchased from Alfa Aesar. All chemicals were used as

received without further purification. Carbon dioxide (5.0 UHP) and argon (5.0 UHP) gas were

purchased from Praxair.

Synthesis of Au Seed Nanoparticles. Au NPs that act as seeds during elemental copper

incorporation were prepared using modified procedures from earlier work.5 HAuCl4·3H2O (0.1 g)

was dissolved in 10 mL of o-xylene and 10 mL of oleylamine at 10 °C under Ar flow and vigorous

Page 64: Conversion Chenlu Xie

50

magnetic stirring. TBAB (0.5 mmol) dissolved in oxylene (1 mL) and oleylamine (1 mL) via

sonication was injected into the HAuCl4 solution. The mixed solution was then stirred for 1 h at

10 °C before ethanol was added to precipitate Au NPs via centrifugation. The precipitate was

redispersed in hexane, followed by addition of ethanol and centrifugation. The final product was

dispersed in hexane for AuCu bimetallic NP synthesis.

Synthesis of AuCu Bimetallic Nanoparticles with Different Ordering Degree. The as-

synthesized Au NPs (0.25 mmol) were dispersed in 20 mL of hexane and used as seeds for AuCu

NP synthesis. A typical synthesis procedure was carried out in a Schlenk line. For the synthesis of

d-AuCu (disordered), i1-AuCu, and i2-AuCu NPs (i1 and i2 at intermediate ordering),

Cu(CH3COO)2·H2O (0.25 mmol) was mixed with oleic acid (0.25 mL) and TOA (1.125 mL) in a

25 mL three-neck flask and heated at 80 °C for 30 min to form a clear solution. Au NPs in hexane

was added to this solution. Once hexane completely evaporated, the mixture was heated at 120 °C

for 20 min under Ar and then quickly heated to 200 or 280 °C, and kept for 50−100 min (i.e.,

200 °C 50 min, 200 °C 100 min, 280 °C 100 min for the synthesis of d-AuCu, i1-AuCu and i2-

AuCu NPs, respectively) before it was cooled to room temperature. The product was collected by

ethanol addition, centrifuged, and redispersed in 20 mL of hexane. Particularly, for the synthesis

of ordered AuCu NPs (o-AuCu NPs), Cu(CH3COO)2· H2O (0.25 mmol) was dissolved in oleic

acid (0.5 mL), TOA (2.25 mL), and oleylamine (0.3 mL) at 80 °C. After the addition of Au NPs/

hexane solution and the evaporation of hexane, the solution was heated at 120 °C for 20 min and

then quickly heated to 300 °C for 50 min. NP separation and purification steps similar to those

used for the other NPs were followed.

Section 4.2.2 XRD and Electron Microscopic Characterization of AuCu Nanoparticles

Powder X-ray Diffraction. Powder X-ray diffraction (XRD) patterns of the AuCu NPs were

obtained using a Bruker AXS D8 Advance diffractometer with a Co Kα source (λCo Kα = 1.79 Å).

Transmission Electron Microscopy and Energy dispersive X-ray Spectroscopy. The

morphology of AuCu NPs was probed using transmission electron microscopy (TEM) on a Hitachi

H-7650 and a FEI Tecnai F20 instrument. The energy-dispersive X-ray spectroscopy (EDS)

elemental mapping images were recorded using an FEI Titan microscope. This instrument was

equipped with an FEI Super-X Quad windowless detector that is based on silicon drift technology.

Aberration-corrected high-angle annular dark field (HAADF) scanning transmission

electron microscopy (STEM). HAADF-STEM was performed on a double aberration-corrected

TEAM0.5 microscope at 300 kV using a high-angle annular detector, resulting in “Z-contrast”

images. Experimental STEM images shown in this work were deconvoluted using the maximum

entropy method.17

Section 4.2.3 Spectroscopic Characterization of AuCu Nanoparticles

ICP-AES. Gold and copper quantitative analysis by inductively coupled plasma atomic

emission spectroscopy (ICP-AES) was carried out on a PerkinElmer optical emission spectrometer

(Optima 7000 DV). Before the ICP-AES measurement, the catalyst was digested in aqua regia for

24 h, and then deionized water was added to dilute the solution to a suitable concentration.

Page 65: Conversion Chenlu Xie

51

Ultraviolet−Visible Spectroscopy. UV−vis absorption spectra (Vernier) were measured with

NPs dispersed in solution.

Section 4.2.4 Electrochemical Study of AuCu Nanoparticles

Electrochemical Reduction of Carbon Dioxide. Working electrodes were prepared by drop-

casting 4 μg of each AuCu NP (dispersed in hexanes) onto a carbon paper electrode (1 cm2 ) and

drying under vacuum at room temperature overnight. All electrochemical measurements were

carried out in our customized setup, which has two compartments separated by an anion-exchange

membrane (Selemion AMV). The working electrode and the counter electrode compartment hold

15 mL of electrolyte each, and the working compartment is sealed in order to allow measurements

of gas products. Platinum wire was used as a counter electrode and Ag/AgCl (3 M KCl) as a

reference electrode. All potentials in this work are converted to the RHE scale by E(vs RHE) =

E(vs Ag/AgCl) + 0.210 V + 0.0591 × pH. The 0.1 M KHCO3 electrolyte was prepared from K2CO3

saturated with CO2 (pH 6.8). During electrochemistry, CO2 flowed through the working

compartment at a rate of 30 sccm. Linear sweep voltammetry was conducted initially at a scan rate

of 50 mV/s, and multiple working electrodes prepared for each sample were similar in performance.

Gas Chromatography. During chronoamperometry, effluent gas from the cell went through

the sampling loop of a gas chromatograph (SRI) to analyze the concentration of gas products. The

gas chromatograph was equipped with a molecular sieve 13X (1/8 in. × 6 ft) and hayesep D (1/8

in. × 6 ft) column with Ar flowing as a carrier gas. The separated gas products were analyzed by

a thermal conductivity detector (for H2) and a flame ionization detector (for CO). Quantification

of the products was performed with the conversion factor derived from the standard calibration

gases.

NMR. Liquid products were analyzed afterward by qNMR (Bruker AV-500) using dimethyl

sulfoxide as an internal standard. Solvent presaturation technique was implemented to suppress

the water peak. Faradaic efficiencies were calculated from the amount of charge passed to produce

each product, divided by the total charge passed at a specific time or during the overall run.

Section 4.2.5 Estimation of the Long-range Order Parameter (S)

The long-range order parameter S, indicating the deviation from perfect order, is defined as:

𝑆 =𝑟𝐴 − 𝐹𝐴1 − 𝐹𝐴

where 𝑟𝐴= fraction of A sites occupied by the “right” atoms, and 𝐹𝐴= fraction of A atoms in the

alloy.

Due to the fact that any departure from perfect long-range order causes the superlattice lines to

be weaker, S could be determined experimentally by comparing the integrated intensity ratio of a

superlattice and a fundamental line.

where F is the structure factor, Z is the Lorentz polarizability factor and P is the multiplicity factor.

𝐼 ∝ 𝑆2|𝐹|2 ∙ 𝑃 ∙ 𝑍

𝐼𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦 (𝑠𝑢𝑝𝑒𝑟𝑙𝑎𝑡𝑡𝑖𝑐𝑒 𝑙𝑖𝑛𝑒)

𝐼𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦 (𝑓𝑢𝑛𝑑𝑎𝑚𝑒𝑛𝑡𝑎𝑙 𝑙𝑖𝑛𝑒)=𝑆2|𝐹|2 ∙ 𝑃 ∙ 𝑍 (𝑠𝑢𝑝𝑒𝑟𝑙𝑎𝑡𝑡𝑖𝑐𝑒 𝑙𝑖𝑛𝑒)

|𝐹|2 ∙ 𝑃 ∙ 𝑍 (𝑓𝑢𝑛𝑑𝑎𝑚𝑒𝑛𝑡𝑎𝑙 𝑙𝑖𝑛𝑒)

Page 66: Conversion Chenlu Xie

52

Here, the calculation of long-range order parameter S was based on the normalized peak

intensity ratio of (111) (i.e., fundamental line) and (001) (i.e., superlattice line) (with parameters

shown below). The normalized peak intensity was determined from Gaussian fitting of each peak.

Peak Parameters

F Z P

(111) 151.6 10.05 8

(001) 108.4 30.41 2

Section 4.2.6 XPS and EXAFS Study of AuCu Nanoparticles

X-ray Photoelectron Spectroscopy. XPS( PHI 5400) measurements were carried out using

an Al Kα source, with the pressure inside the chamber maintained below ∼4 × 10−9 Torr. The

spectra were collected at a pass energy of 17.9 eV. All the spectra collected were corrected using

a Shirley background. Compositions were determined by considering the atomic sensitivity

factors.

X-ray absorption spectroscopy (XAS) data collection and refinement. Extended X-ray

absorption fine structure (EXAFS) data were collected at the Advanced Light Source Beamline

10.3.2. The X-ray wavelength was monochromatized by a Si(111) double-crystal, fixed exit

monochromator. The intensity of the incident X-ray radiation, I0, was monitored with a nitrogen

filled ionization chamber. All data were collected in fluorescence mode with a 7-element Ge

detector (Canberra). The data at the Cu K-edge and Au L3-edge were calibrated to a Cu foil and

Au foil, respectively. All spectra at a given edge were aligned according to a glitch in I0 near each

absorption edge.

EXAFS data reduction and EXAFS fitting. The data analysis were performed using the

IFEFFIT based programs, Athena and Artemis. Edge step normalization for each spectra was

performed by subtracting the pre-edge and post-edge backgrounds in Athena. For EXAFS

background removal, a cubic spline was fit to the data and the k-space data was Fourier

transformed resulting in an R-space spectrum which was fit in Artemis using multiple k-weight

fitting (k1, k2, k3). EXAFS data was fit to the following EXAFS function:

𝜒(𝑘) = 𝑆02 ∑

𝑁𝑖𝑘𝑅𝑖

2 𝐹𝑖(𝑘)𝑒−2𝑘2𝜎𝑖

2𝑒−2𝑅𝑖/𝜆(𝑘) sin[2𝑘𝑅𝑖 − 𝜑𝑖(𝑘)]

𝑠ℎ𝑒𝑙𝑙𝑠

𝑖=1

The amplitude of the contribution from each coordination shell in the EXAFS function is

summed to generate a fit to the data. 𝑆02 represents an amplitude reduction factor which is typically

assumed to be chemically transferable and is affected by shake-up effects at the absorbing atom.

It was estimated by performing a fit to EXAFS transmission measurements on Cu and Au metal

foils. Values for 𝑆02 were determined to be 0.83 for Au and 0.75 for Cu. 𝑁𝑖 and 𝑅𝑖 are the

coordination number and half-path length between the central absorbing atom and a scattering

atom, respectively. The mean-square disorder in the distance from the central absorbing atom to a

given shell due to thermal fluctuation and structural disorder is represented by 𝜎𝑖2. These

parameters are calculated by fitting the experimental data. The photoelectron mean free path is

Page 67: Conversion Chenlu Xie

53

represented by 𝜆. Lastly, 𝐹𝑖(𝑘) is the backscattering amplitude and 𝜑𝑖(𝑘) is the the phase factor

for a given coordination shell. These parameters are calculated through ab initio methods using

FEFF6 as embedded in Artemis and Atoms.

The o-AuCu NP EXAFS data was fit to the L10 (face centered tetragonal) structure of AuCu.

The data at the Cu K-edge and Au L3-edge were co-fit in order to best represent the actual

nanoparticle structure. A multiple-shell fit was performed in order to account for the long-range

ordering of the o-AuCu NP sample. The following restraints to coordination numbers were applied

in order to best represent the ordering of the face centered tetragonal intermetallic:

Au L3-edge

NAu1 = NCu1/2

NCu1-Au5 = 2×NAu5

NCu1-Au5-Cu1 = NAu5

NAu6 = NAu1-Au6/2

NAu1-Au6-Au1 = NAu1-Au6/2

Cu K-edge

NCu1 = NAu1/2

NAu1-Cu5 = 2×NCu5

NAu1-Cu5-Au1 = NCu5

NCu6 = NCu1-Cu6/2

NCu1-Cu6-Cu1 = NCu1-Cu6/2

Path lengths between identical pairs of lattice sites were restricted to be equal such that the

nearest-neighbor distance between an Au and Au atom was the same as a Cu and Cu atom. This

was extended out to further shells such that, for example, the distance from a Cu to the Cu4 shell

was the same as the distance from an Au to the Au4 shell. These requirements are dictated by the

ordered nature of the structure. Lastly, the mean-square disorder in the half-path length, 𝜎2, was

restricted to be the same only if the half-path length and the terminal scattering atoms were the

same.

Analysis of the o-AuCu NP EXAFS fit. Even with the rigid restraints applied to the fit to

model the ordered lattice, the co-fit provided scientifically reasonable and insightful values for

each fitted parameter. Furthermore, Au-enrichment on the surface compared to Cu is demonstrated

by comparing the CN of each, which supports the observation using aberration-corrected HAADF-

STEM. However, the error bars are quite large for photoelectron scattering paths at large distances

(i.e. involving fifth or sixth scattering atoms). This is due to very low amplitude of the Au EXAFS

signal in the outer shell regions that limits the quality of the co-fit. Still, the large amplitudes seen

in the outer shells of the Cu EXAFS signal originating from the ordered structure enabled us to

apply constraints in performing the co-fit that supported the existence of the ordered lattice with

physically reasonable values, except the large errors seen at outer shells. Furthermore, our

conclusion of lower CN for Au relative to Cu is unaffected by these large errors as the nearest

neighboring atoms are fit well. When the model was used to co-fit just to the nearest neighbors, it

led to the same conclusion.

Section 4.2.7 Computational Methods

Density Functional Theory (DFT) calculations were performed using the Quantum Espresso

(QE)18 code integrated with the Atomic Simulation Environment (ASE).19 The revised

Page 68: Conversion Chenlu Xie

54

Perdew−Burke−Ernzerhof (RPBE) functional was used for exchange and correlation,20 which has

been shown to provide better adsorption energies than the PBE functional. The ion−electron

interaction was described by ultrasoft pseudo potentials.21 The Kohn− Sham (KS) wave functions

were expanded in a series of plane waves with a converged energy cutoff of 550 eV and a density

cutoff of 5500 eV. The electron density was determined by the iterative diagonalization of the KS

Hamiltonian using Pulay mixing of electronic densities at an electronic temperature of 0.1 eV. All

total energies were extrapolated to 0 K.

The lattice parameters for the ordered L10 AuCu intermetallic within this crystal structure were

calculated as a = 4.09 Å and c = 3.73 Å, in a reasonable agreement with the experimental lattice

parameters of a = 3.98 Å and c = 3.72 Å in the current study and previous studies.11

To determine adsorption energies, the surfaces were modeled using four-layer, 3×2 slabs

repeated in a supercell geometry, with at least 17 Å of vacuum between successive slabs and the

bottom two layers fixed. To determine surface energies, nine-layer symmetric slabs were used with

five layers fixed.22 Slabs were relaxed until all force components were less than 0.05 eV/Å. We

used a Monkhorst−Pack grid with dimensions of 4×6×1 to sample the first Brillouin zones of the

surfaces.23 In determining the adsorption energies, all sites were considered, and only the most

stable ones are reported here.

To determine the thermochemical reaction energetics, we applied the computational hydrogen

electrode (CHE) model,24 which does not require explicit consideration of the solvent or solvated

protons. To determine the reaction free energy ΔG of a given proton electron transfer, e.g.,

* + CO2 + (H+ + e-) → COOH*

we can determine the free energy Go of the corresponding chemical hydrogenation reaction

* + CO2 + H2 → COOH*

At standard pressure, the reaction H+ + e- 1/2 H2(g) is at equilibrium at 0 VRHE. Therefore,

ΔG = ΔG0 at 0 VRHE, and in general, ΔG(U) = ΔG0 + eU. We also define a limiting potential UL,

the potential at which the reaction step becomes exergonic: UL = −ΔG0/e. Since barriers scale with

reaction energies,25 trends in UL should follow trends in activity. In the case of CO2 reduction

toward methane and methanol, this potential has been shown previously to correlate well with

experimental activities on transition metals.26,27 We have applied a simple −0.25 eV correction to

*COOH adsorption energies to account for solvation, as was determined in previous works.27

Section 4.3 Result and Discussion

Section 4.3.1 Control of AuCu NPs with Different Atomic Ordering

AuCu bimetallic nanoparticles with varying degrees of atomic ordering were realized by

synthetic procedures modified from earlier work.5 Mainly, temperature and time were controlled

during the incorporation of elemental copper into the gold NP seed (Figure 4.1 and Figure 4.2),

where higher temperatures and longer times promote the formation of ordered lattices.

Page 69: Conversion Chenlu Xie

55

Figure 4.1. TEM images of Au NP seeds (size 6 nm)

Figure 4.2. Scheme of synthesis process of AuCu NPs with different ordering degree. The heating temperature and

time are controlled to tune the atomic ordering degree.

Since we start from identical seed particles, AuCu bimetallic NPs with similar sizes (∼7 nm)

could be produced that look identical in their morphologies (Figure 4.2, Figure 4.3 and Figure

4.4a). Au-to-Cu ratios, identified from ICP-AES (Figure S3), are all around 1, confirming their

compositional invariance.

Page 70: Conversion Chenlu Xie

56

Figure 4.3. TEM images of AuCu bimetallic nanoparticles synthesized at various conditions allowing for systematic

tuning of the degree of ordering. Scale bar, 20 nm.

Figure 4.4. Size distribution and Au:Cu ratio of AuCu nanoparitcles with different ordering (a). Size distribution;

(b). ICP-AES analysis of AuCu bimetallic nanoparticles.

The structural difference between AuCu NPs could be identified from XRD. As shown in

Figure 4.5, the peak positions of 111 and 200 diffractions further confirmed the formation of a

bimetallic system, as they lie between those of pure Au and Cu. However, there was a trend where

the superlattice peaks, which should be absent in a disordered face-centered cubic (FCC) structure

of an alloy, became more evident in the NPs synthesized at higher temperatures for extended

periods. Moreover, peak splitting became clearer as well at the 200 and 220 positions, indicating

that the crystal structure was transitioning from FCC to face-centered tetragonal. All of these

indicated an increase in the degree of atomic ordering following greater input of energy (controlled

with temperature and time) during Cu diffusion into Au NP.5,10,11

Page 71: Conversion Chenlu Xie

57

Figure 4.5. XRD of AuCu bimetallic nanoparticles. Diffractograms were recorded using Co K radiation (λCo-K =

1.79 Å).

The degree of ordering was further quantified by the longrange order parameter (S), estimated

from the relative intensity ratio between the superlattice peak and the fundamental peak.28 The

AuCu NP synthesized at 300 °C had the highest S value of 0.611, which corresponded to ∼80%

of atoms in the ordered lattice positions. After the second most ordered AuCu NP (synthesized at

280 °C 100 min), the degree of ordering could not be quantified by S for the other two NPs, which

had only short-range order. Combining the qualitative and quantitative comparisons of degree of

ordering among the four AuCu NPs, we could conclude that the AuCu NP synthesized at 200 °C

for 50 min (hereafter d-AuCu) was disordered (or the least ordered) and the one synthesized at

300 °C for 50 min (hereafter o-AuCu) was ordered (or the most ordered). The AuCu NPs at

intermediate stages of ordering were named as i1-AuCu and i2-AuCu.

Section 4.3.2 AuCu NPs with Different Ordering Degree for Electrochemical CO2 Reduction

To study their activity toward electrochemical reduction of CO2, identical numbers of each type

of AuCu NP (mass of 4 μg total) were loaded on carbon paper electrodes (Figure 4.6). Each NP

sample was washed multiple times to the point just before the NP solution became unstable, and

the mass fraction of ligands in each sample was kept similar (Figure 4.7). Any thermal or chemical

treatments were avoided so as not to alter their inherent structure produced. Each electrode was

then tested in a customized electrochemical setup under conditions of 0.1 M KHCO3 (pH 6.8) at 1

atm CO2 and room temperature. Linear sweep voltammetry of AuCu NPs resulted in similar

current densities from utilizing same number of NPs with identical morphology for each and

providing equal surface area (Figure 4.8a). During chronoamperometry, gas products were

measured in situ using gas chromatography, and liquid products were analyzed from quantitative

nuclear magnetic resonance (qNMR) spectroscopy.

Page 72: Conversion Chenlu Xie

58

Figure 4.6. Representative SEM images of carbon paper electrodes decorated with AuCu NPs. NPs are well-dispersed

and individually isolated to ensure activity comparison assuming equivalent surface area. After electrochemistry,

morphological changes to the electrodes were not observed.

Figure 4.7. Ligand mass fraction of AuCu NPs before electrochemical testing.

Page 73: Conversion Chenlu Xie

59

The major products formed were CO and H2, and their faradaic efficiencies (FE) were

measured at −0.77 V vs RHE (Figure 4.8b). The only liquid product found was formate at relatively

low FE < 5% (Figure 4.9). For the d-AuCu NP, H2 was the dominant product (H2 FE ≈ 61% and

CO FE ≈ 34%), which meant it was considered rather inactive for electrochemical CO2 reduction.

However, a continuous rise was observed in CO FE with increase in the ordering degree of AuCu

NPs, where o-AuCu NP exhibited the highest CO selectivity. CO became the major product for o-

AuCu NP, with FE ≈ 80%. Previous study on the effects of composition in Au−Cu alloys for

electrocatalytic CO2 reduction has shown that H2 is the major product for AuCu (1:1 ratio),

consistent with what we observe from d-AuCu or i1-AuCu.15 Surprisingly, without altering

composition, we found that the ordering transformation could enhance CO formation, transitioning

a CO2 inactive catalyst to an active catalyst with high CO selectivity. The CO FEs measured at a

range of potentials for d-AuCu and o-AuCu NP further confirmed the activity enhancement

following disorder-to-order transformation, where CO formation could be obtained at ∼200 mV

lower overpotentials (Figure 4.10).

Figure 4.8. Electrochemical CO2 study on the AuCu nanoparticles. (a). Linear sweep voltammetry of AuCu NPs under

conditions of 0.1 M KHCO3 (pH 6.8) and 1 atm CO2 with scan rate 50 mV/s; (b). Electrochemical CO2 reduction

activities of AuCu bimetallic nanoparticles evaluated by faradaic efficiencies of CO and H2. Measurements were

conducted at -0.77 V vs. RHE.

Figure 4.9. Formate FE of AuCu NPs at -0.77 V vs. RHE measured by qNMR.

Page 74: Conversion Chenlu Xie

60

Figure 4.10. CO FE at various potentials for d-AuCu and o-AuCu NP catalysts.

Figure 4.11. Elemental maps of Au and Cu for d-AuCu and o-AuCu NPs.

Page 75: Conversion Chenlu Xie

61

Section 4.3.3 Further Structural Investigation at both Nanoscale and Atomic Scale

In order to understand this drastic change in activity that arises from the ordering

transformation of a AuCu bimetallic NP, we probed the structure of d-AuCu and o-AuCu NP using

various methods. Elemental analysis mapping of the distribution of Au and Cu showed that both

elements were uniformly distributed in a single NP for both d-AuCu and o-AuCu NP (Figure 4.11).

In order to investigate down to the atomic level, aberration-corrected HAADF-STEM was

implemented. Figure 4.12a,b show aberration-corrected HAADF-STEM images of o-AuCu NPs

in two different orientations. From the images, a periodic oscillation of intensity was observed,

which can be attributed to the Z-contrast differences between different elements that should be

present in an ordered lattice. As Au (Z = 79) and Cu (Z = 29) differ greatly in atomic number, a

column of Au atoms should appear much brighter than that of Cu atoms in an ordered structure.

By measuring the distance between the planes and distinguishing elements by their contrast,

certain unit cell orientations of the AuCu L10 structure were found (Figure 4.12c,d).

Figure 4.12. Structural investigation of o-AuCu at atomic resolution. Aberration-corrected HAADF-STEM images of

o-AuCu (a, b), and magnified STEM images of the center of the particle together with the overlapping schematics of

structure projections (c, d). Atoms in orange and blue color represent gold and copper, respectively. (e) Intensity profile

across the particle measured from the yellow box shown in (a). The distance between the strong intensities matches

with the separation between gold atoms. Arrows indicate alternating high and low intensities, which represent gold

and copper atoms, respectively. (f) Average intensity profile from the interior to the outer surface of the particle along

rows A and B shown in (b). The distance between the peaks shown in A matches with the known separation between

gold atoms.

Page 76: Conversion Chenlu Xie

62

In addition to the atomically ordered structures that were verified, certain elemental

characteristics at the surface were also observed. As can be directly seen from the images, atoms

near the NP surface contained only brighter intensities of gold, in contrast to the alternating

intensities at the core. This suggested a few-atoms-thick gold layer over the structurally ordered

core. With the intensity profile measured across the particle (Figure 4.12e), a three-atoms-thick

gold layer could be identified. Intensities were also measured along the rows of atoms which

should contain only one element (Figure 4.12f). While the intensity along the row of gold atoms

decayed to the surface (from the thickness decreasing), the intensity profile along what is supposed

to be only copper atoms had spiking intensities starting from three atoms beneath the surface. All

of these pointed to the fact that o-AuCu NP contained around three-atoms-thick gold layers grown

directly over the ordered lattice. This could be observed in multiple images as well (Figure 4.13).

In contrast, aberration-corrected HAADF-STEM images of d-AuCu NPs exhibited uniform

intensities (Figure 4.14) from the random distribution of both elements, as expected. From

atomically resolved imaging of o-AuCu NPs, we could conclude that the ordering transformation

of a AuCu NP induced three-atoms-thick gold layers at its surface, in addition to the crystal

structure transition to intermetallic ordered phase throughout its interior.

Figure 4.13. Aberration-corrected HAADF-STEM images of o-AuCu NPs (a, b, c) and an intensity profile along

yellow line shown in c (d).

Figure 4.14. Aberration-corrected HAADF-STEM images of d-AuCu NPs.

Page 77: Conversion Chenlu Xie

63

Section 4.3.4 Extended X-ray Absorption Fine Structure (EXAFS): Coordination

Environment Surrounding Au and Cu

XAS was used to further probe the structure of AuCu NPs. Absorption spectra at the gold L3-

edge and copper K-edge were collected from o-AuCu NPs and d-AuCu NPs (Figure 4.15). Figure

4.16 presents extended X-ray absorption fine structure (EXAFS) data for both NPs, where the

coordination environment surrounding Au and Cu for each NP can be observed. At the Au L3-edge

(Figure 4.16a), o-AuCu NP exhibits a larger first-shell scattering amplitude, which can be

attributed to the stronger constructive interference of photoelectron scattering in the ordered

structure. The atomic order is more evident at the Cu K-edge (Figure 4.16b), where third and fourth

shell scattering (at radial distances from 4 to 6 Å) can be clearly observed due to the unique

multiple scattering paths present in the ordered structure. On the other hand, d-AuCu NP at the Cu

K-edge shows many variations of scattering paths at various lengths, together with destructive

interference from the disordered structure, which lowers the amplitude.

Figure 4.15. X-ray absorption spectra at the Au L3-edge and Cu K-edge collected from o-AuCu NPs and d-AuCu NPs.

By fitting the Au L3-edge and Cu K-edge of the o-AuCu NPs simultaneously to an ordered

face-centered tetragonal (L10) model, information regarding the coordination environment

surrounding gold and copper atoms in o-AuCu NPs could be extracted (Tables 4.1 and 4.2). When

comparing the number of nearest neighbor atoms to gold and copper, copper had a coordination

number (CN) of 9.0 ± 1.2, while gold had a CN of only 6.9 ± 1.1. The lower first shell CN for gold

indicated that gold atoms are relatively undercoordinated compared to copper, which resulted from

gold being enriched at the surface29 as observed from HAADF-STEM. Therefore, from STEM

imaging and XAS analysis, we could conclude that the o-AuCu NP surface contains a three-atoms-

thick gold layer over the ordered AuCu intermetallic core. Furthermore, we speculate NPs at

intermediate stages of ordering to have gold overlayers partially covering their surface. Rise in the

degree of ordering may lead to increased coverage of gold overlayers associated with the ordered

phase (Figure 4.17). XPS (Figure 4.18) and UV−vis spectroscopy (Figure 4.19) were used to

further prove that the gold overlayer is only a few atoms thick and not significant enough to be

detected by these techniques.30

Page 78: Conversion Chenlu Xie

64

Figure 4.16. X-ray absorption spectroscopy of AuCu NPs. EXAFS spectra at the Au L3-edge (a) and Cu K-edge (b)

of o-AuCu NPs (orange circles) and d-AuCu NPs (blue circles). Orange solid lines are the fitting results to o-AuCu

NPs.

Table 4.1. EXAFS fitting parameters for the fit to the o-AuCu NP displayed in Fig. 4.16a of the main text.

Au L3-edge

N R (Å) σ2

Cu1 4.6 (0.7) 2.68(1) 0.0077(14)

Au1 2.3 (0.4) 2.79(1) 0.0038(17)

Au4 3.3 (4.7) 4.67(6) 0.0043(99)

Cu2 4.8 (6.9) 4.78(4) 0.0063(105)

Au5 3.1 (28.8) 5.38(6) 0.0101(672)

Cu1-Au5 6.2 (57.6) 5.38(6) 0.0101(672)

Cu1-Au5-Cu1 3.1 (28.8) 5.38(6) 0.0101(672)

Au6 1.1 (7.3) 5.62(7) 0.0073(606)

Au1-Au6 2.2 (14.6) 5.62(7) 0.0073(606)

Au1-Au6-Au1 1.1 (7.3) 5.62(7) 0.0073(606)

∆E 4.1(1.2)

R-factor 7.04%

Page 79: Conversion Chenlu Xie

65

Table 4.2. EXAFS fitting parameters for the fit to the o-AuCu NP displayed in Fig. 4.16b of the main text.

Cu K-edge

N R (Å) σ2

Au1 6.0 (0.8) 2.68(1) 0.0077(14)

Cu1 3.0 (0.4) 2.79(1) 0.0149(61)

Cu4 8.0 (0.0) 4.67(6) 0.0160(108)

Au2 2.8 (4.0) 4.78(4) 0.0063(105)

Cu5 1.6 (3.0) 5.38(6) 0.0049(144)

Au1-Cu5 3.2 (6.0) 5.38(6) 0.0049(144)

Au1-Cu5-Au1 1.6 (3.0) 5.38(6) 0.0049(144)

Cu6 3.4 (11.2) 5.62(7) 0.0067(214)

Cu1-Cu6 6.8 (22.3) 5.62(7) 0.0067(214)

Cu1-Cu6-Cu1 3.4 (11.2) 5.62(7) 0.0067(214)

∆E 4.0(0.6)

R-factor 7.04%

Figure 4.17. Aberration-corrected HAADF-STEM images of d-AuCu NPs exhibiting short-range order. Ordered

atomic arrangements can be observed at some portions of the NPs. Order-to-disorder transition in an AuCu bimetallic

system is considered as a first-order transition,2 which involves nucleation and growth of the ordered phase. Increase

in the degree of ordering will result in the growing volume fraction of the ordered phase within the disordered matrix,

which may lead to increased surface coverage by the gold overlayers associated with the ordered domains.

Page 80: Conversion Chenlu Xie

66

Figure 4.18. Au 4f and Cu 2p XPS spectra of AuCu NPs. Both exhibit Au-to-Cu atomic ratio close to 1. d-AuCu NP

shows weak signals (~935eV and ~943.5eV) from oxidized copper (Cu2+), which indicates that there is more

likelihood of copper atoms being exposed to the surface of d-AuCu NP.

Figure 4.19. UV-Vis absorption spectra of AuCu NPs. Surface plasmonic resonance (SPR) results in an absorption

peak at ~540nm for Au-Cu bimetallic NP of 1:1 atomic ratio. Degree of atomic ordering does not shift the SPR peak

(also expected from their same visual appearance in color) and this indicates that the gold overlayer formed on o-

AuCu NP is only few atoms thick. If the thickness of the overlayer becomes significant (beyond ~1nm), the original

SPR characteristic decays and the new features from the overlayer appear.49

From the structural investigation of o-AuCu NPs, we found the formation of a three-atoms-

thick gold overlayer following ordering transformation. This suggests that surface enrichment of

gold at atomic levels may be the origin of enhanced catalytic behavior. Identifying the precise

structures of surface active sites is necessary to gain catalytic insights. Here, we have identified

the catalytically active surface structure of a ordered AuCu NP and this may well be extended to

other ordered bimetallic systems involving noble and non-noble element mixtures, such as the

Cu−Pd system recently demonstrated for CO2 reduction.31

Page 81: Conversion Chenlu Xie

67

Section 4.3.5 Further Comparison with Pure Au NPs on CO2 Reduction

To further investigate electrocatalytic activities for reducing carbon dioxide, CO production

rates measured in terms of specific current density (mA/cm2) and mass activity (A/gAu) were

probed for AuCu bimetallic NPs, in addition to the FE (Figure 4.22). Catalytic activity of pure Au

NPs (Figure 4.20) of same size was compared as well to assess the effects of having a few-atoms-

thick Au layer directly over the ordered core. Beyond having NPs of the same dimension with

identical numbers to match active surface areas, surface areas were estimated using

electrochemical methods. With o-AuCu NPs having a three-atoms-thick gold layer on the surface,

underpotential deposition of copper in acidic media could be stably performed to estimate their

surface area (Figure 4.21).32,33 The same procedure was conducted on Au NPs as well and yielded

comparable surface areas.

Figure 4.20. Au NPs observed from TEM and their size distribution.

Page 82: Conversion Chenlu Xie

68

Figure 4.21. Cu underpotential deposition performed on o-AuCu NP and Au NP electrodes in 0.5M H2SO4 solution

with 0.1M CuSO4 and Ar purge. Cyclic voltammetry was conducted from 0.04 to 0.44 V vs Ag/AgCl (1M KCl) at a

scan rate 50 mV/s and repeated till traces converged. The charge associated with the anodic stripping wave at ~ 0.27

V vs Ag/AgCl was integrated. The surface characteristics of o-AuCu NP having a few atom thick gold layers allowed

Cu underpotential deposition on gold in acidic media. By using conversion factors of 92.4 C/cm2, the surface areas

estimated for o-AuCu NP electrode and Au NP electrode were 1.8 cm2 and 1.68 cm2, respectively.50, 51 The comparable

surface areas further supported the validity of comparison between NP electrodes with identical number of NP

catalysts. The surface area measured for o-AuCu NP electrode was used to other AuCu NP electrodes. In addition, in

comparison to the theoretical estimate of 2.31 cm2 (based on NPs of 7nm diameter), the NP electrodes are able to use

~75% of its actual surface area, which may be due to the activity inhibited at sites associated with surface ligands and

the area under contact with the substrate.

Figure 4.22a shows CO FEs at −0.77 V vs RHE, where atomic ordering drastically converts

AuCu NPs to be selective for CO production as mentioned above. Indeed, this trend is

accompanied by rise in the specific CO current density from 0.43 mA/cm2 for d-AuCu NP to 1.39

mA/cm2 for o-AuCu NP (Figure 4.22b). Therefore, we find that atomic ordering in a AuCu

bimetallic NP enhanced the intrinsic activity for CO2-to-CO conversion by 3.2-fold (Figure 4.23),

suggesting the importance of precise control of atomic arrangements in bimetallic NPs for catalytic

purposes.

Interestingly, when o-AuCu NPs were compared to Au NPs, Au NPs exhibited higher CO

selectivity, while o-AuCu NPs showed higher activity. Au NPs had a CO FE of 87% and a specific

current density of 1.01 mA/cm2. A similar trend in CO FE and specific current density was

observed at other potentials as well (Figure 4.24). This indicated that a three-atoms-thick gold

layer over the intermetallic core possesses different catalytic properties, such as the intermediate

binding strength, to a pure gold lattice.

Here, we also probed mass activities (A/gAu) for CO formation and, as shown in Figure 4.22c,

AuCu NPs are clearly advantageous for CO production with even NPs at intermediate ordering,

such as i1-AuCu NP, reaching comparable activities to Au NP at ∼320 A/g Au. For the o-AuCu

NP, CO mass activity reaches up to ∼830 A/gAu, which is 2.6-fold over that of Au NP. The o-

AuCu NP catalyst was also relatively stable for extended periods of 12 h (Figure 4.25). However,

there was still a steady loss of activity (∼15%) and this might be linked to the structural stability

Page 83: Conversion Chenlu Xie

69

of gold overlayers induced in the ordered structure, as some loss of gold was observed. For

electrochemically producing carbon monoxide, which is considered as the most economically

viable so far among various CO2 reduction products for downstream processing to useful

chemicals,34 various types of gold-based nanostructures have been demonstrated.32, 33, 35−38 We find

that creating an Au−Cu bimetallic system with induced atomic ordering could be an effective

strategy for promoting efficient processing of CO2, where the structural motifs found here can also

be translated to other Au-based morphologies that provide higher current output.38 Certainly,

further efforts are needed to ensure catalytic stability of the motifs found here to the levels required

for industrially relevant applications.

Figure 4.22. Electrocatalytic CO2 reduction activity of AuCu NPs. CO faradaic efficiencies (a), CO partial current

densities (b) and CO mass activities (c) of d-AuCu NP, i1-AuCu NP, i2-AuCu NP, o-AuCu NP, and Au NP for

electrochemical reduction of CO2 at -0.77 V vs. RHE. CO partial current densities are specific current densities based

on actual surface area of NP catalysts measured by Cu underpotential deposition on o-AuCu NP and Au NP.

Electrochemically active surface areas of d-AuCu NP, i1-AuCu NP and i2-AuCu NP were estimated from the value

measured of o-AuCu NP. CO mass activities are based on the actual mass of gold (gAu) loaded for each catalyst. Green

circles in (b, c) indicate enhancements over the activity of d-AuCu NP, which is set to one.

Figure 4.23. CO specific current and CO production rate of o-AuCu NPs and d-AuCu NPs. (a) CO specific current

density at various potentials of o-AuCu NPs and d-AuCu NPs. (b) CO production rate enhancements of o-AuCu NPs

over d-AuCu NPs.

Page 84: Conversion Chenlu Xie

70

Figure 4.24. CO partial current densities and CO mass activities of d-AuCu NP, o-AuCu NP, and Au NP for

electrochemical reduction of CO2 at -0.85 V vs. RHE. (a) CO partial current density. (b) CO mass activities. Numbers

in (a) indicate CO faradaic efficiency (%).

Figure 4.25. Stability test of o-AuCu NPs under extended operation of 12 hours for electrochemical reduction of

CO2, exhibiting high CO2-to-CO conversion activity.

Section 4.3.6 Density Function Theory (DFT) Calculations: Insights into the Origin of

Enhanced CO2 Reduction Activity

The origin of enhanced CO2 reduction activity by atomic ordering transformations and the

resulting three-atoms-thick gold overlayer of AuCu NPs was investigated using DFT calculations.

Here, we modeled a (211) facet, since it has been shown that step sites are considerably more

energetically favorable for the reactions involved than terrace sites39 and catalytic activity should

be dominated by the step sites.36 Figure 4.26a shows the model systems considered for AuCu NPs.

o-AuCu NP was modeled as a strained Au(211) surface, since the underlying intermetallic AuCu

lattice compressively strains (∼6%) the few-atoms-thick Au layer from its original lattice

Page 85: Conversion Chenlu Xie

71

parameter. In principle, the underlying AuCu phase could also have an electronic effect on the Au

overlayers. However, previous theoretical studies have shown that the electronic effect is

essentially negligible at transition metal overlayers that are approximately three monolayers

thick.40−43 For the d-AuCu NPs, surface stability of various site motifs was investigated (Figure

4.27). Here, we assumed that the surface of d-AuCu NPs would be in a metastable state with mixed

Au/Cu elemental configuration that has local, short-range order. Of the various (211) facets, the

Au-rich (211) step had the lowest surface energy. A model Au(211) surface in its original lattice

configuration was also investigated as a comparison.

Figure 4.26. Computational results of CO2 reduction on AuCu surfaces. (a) Top and side views of model 211 slabs for

d-AuCu NP, o-AuCu NP, and Au NP. (b) Calculated limiting potentials for CO2 reduction and H2 evolution on model

systems. (c) The difference in limiting potentials for CO2 reduction and H2 evolution on model systems.

Page 86: Conversion Chenlu Xie

72

Figure 4.27. Surface energy of mixed Au/Cu 211 slabs as a function of Au density at the surface. Au and Cu atoms

are shown in yellow and orange, respectively. Dashed lines indicate step sites. The Cu-rich surface, mixed Au/Cu

surface and Au-rich surface are shown by red, blue and black frames, respectively. Surface energies of 20 different

slabs generated by varying the position of Au and Cu on a stepped (211) surface in a cubic lattice were investigated.

The analysis suggests a significant stabilization with an increase Au density at the surface, consistent with the lower

surface energy of Au. Assuming that the disordered AuCu NP would be in a metastable state with mixed Au and Cu

elemental configuration that would possess local order, the surface of d-AuCu NP could be considered as Au-rich

(211) surface.

The reduction of CO2 to CO occurs through a COOH intermediate bound to the active site.44−46

For the metals that bind CO weakly, desorption of CO readily occurs, and the formation of COOH*

is the potential-limiting step, where * represents a surface site. On the other hand, hydrogen

evolution involves H* as the only adsorbed intermediate, and the binding energy of H* has been

shown to be a reasonable descriptor of hydrogen evolution activity.47,48

Figure 4.26b compares the thermodynamic limiting potentials UL(CO2) and UL(H2) for the three

model systems. Since activation barriers scale linearly with binding energies, trends in limiting

potentials reflect trends in catalytic turnover.25 Based on this analysis, UL(CO2) of o-AuCu NP is

most positive followed by Au NP and d-AuCu NP and therefore CO2-to-CO conversion activity

would decrease in the order of o-AuCu NP > Au NP > d-AuCu NP, which is in agreement to the

trend observed experimentally. Recent DFT calculations on transition metals have suggested that

the difference between the limiting potentials for CO2 reduction and H2 evolution (i.e., UL(CO2)−

UL(H2)) can be linked to the trends in experimentally observed CO2 reduction selectivity.39 With

UL(CO2) being more negative than UL(H2) for all transition metals, more positive UL(CO2)−UL(H2)

corresponds to higher selectivity toward CO2 reduction. The trend observed for UL(CO2)−UL(H2),

as shown in Figure 4.26c, suggests that selectivity toward CO2 reduction decreases in the order of

Au NP > o-AuCu NP > d-AuCu NP, consistent with the experimental trends.

From the thermochemical DFT analysis, we find that the activity of o-AuCu NP originates from

the compressively strained three-atoms-thick gold layers formed over the AuCu intermetallic core.

We note that, while the differences in energetics appear to be only ∼0.1 eV, such shifts can provide

order of magnitude changes in activity, since the turnover frequencies vary exponentially with

activation barriers, which scale linearly with binding energies. Furthermore, projected density of

Page 87: Conversion Chenlu Xie

73

states of the model systems predict a similar trend as well (Figure 4.28), which shows that the

ordering transformation in AuCu should favor CO2 reduction.

Figure 4.28. d-projected density of states for the model systems of d-AuCu, o-AuCu, and Au. Associated d-band

centers are -3.52, -3.19, and -3.14 eV, respectively. The d-band model for metal surfaces provides a simple correlation

between binding strength of adsorbates and their d-band centers;52-54 it is expected that o-AuCu would have stronger

binding to *COOH (a major intermediate to CO) compared to d-AuCu due to its higher d-band center relative to the

Fermi level. This would facilitate the formation of CO from CO2. The d-band centers for o-AuCu and Au model

systems are close and within DFT error,55 so that further considerations of the d-band shape and adsorbate geometry

are necessary to rationalize the difference in binding between these two systems.

Section 4.4 Conclusion

The exquisite control of atomic arrangements in unit cells of bimetallic systems to an individual

nanoparticle level could be a promising approach to induce interesting properties and understand

their structural origin. The effect of ordering transformations in AuCu bimetallic NPs on their

electrocatalytic activity for CO2 reduction was studied, where ordered AuCu NPs selectively

reduced CO2, in contrast to disordered AuCu NPs that favored hydrogen evolution. With structure

probing techniques down to the atomic level, the enhanced activity was attributed to compressively

strained three-atoms thick gold overlayers that formed over the intermetallic core, resulting from

the disorder-to-order transformation. With a large number of multimetallic systems accessible, we

expect the method of regulating phase transformation between disordered alloys and ordered

intermetallic in nanomaterials to have great potential in areas where control over the level of

atomic precision is desired.

Section 4.5 References

(1) Gilroy, K. D.; Ruditskiy, A.; Peng, H.-C.; Qin, D.; Xia, Y. Chem. Rev. 2016, 116 (18), 10414.

(2) Wang, D.; Li, Y. Adv. Mater. 2011, 23 (9), 1044.

(3) Sankar, M.; Dimitratos, N.; Miedziak, P. J.; Wells, P. P.; Kiely, C. J.; Hutchings, G. J. Chem.

Soc. Rev. 2012, 8099.

(4) Porter, D. A.; Easterling, K. E.; Sherif, M. Y. Phase Transformations in Metals and Alloys, 3rd

ed.; CRC Press, 2009.

Page 88: Conversion Chenlu Xie

74

(5) Chen, W.; Yu, R.; Li, L.; Wang, A.; Peng, Q.; Li, Y. Angew. Chemie Int. Ed. 2010, 49 (16),

2917.

(6) Jia, Q.; Caldwell, K.; Ramaker, D. E.; Ziegelbauer, J. M.; Liu, Z.; Yu, Z.; Trahan, M.; Mukerjee,

S. J. Phys. Chem. C 2014, 118 (35), 20496.

(7) Kim, J.; Lee, Y.; Sun, S. J. Am. Chem. Soc. 2010, 132 (14), 4996.

(8) Li, Q.; Wu, L.; Wu, G.; Su, D.; Lv, H.; Zhang, S.; Zhu, W.; Casimir, A.; Zhu, H.; Mendoza-

Garcia, A.; Sun, S. Nano Lett. 2015, 15 (4), 2468.

(9) Prabhudev, S.; Bugnet, M.; Bock, C.; Botton, G. a. ACS Nano 2013, 7 (7), 6103.

(10) Sra, A. K.; Ewers, T. D.; Schaak, R. E. Chem. Mater. 2005, 17 (4), 758.

(11) Sra, A. K.; Schaak, R. E. J. Am. Chem. Soc. 2004, 126 (21), 6667.

(12) Wang, D.; Xin, H. L.; Hovden, R.; Wang, H.; Yu, Y.; Muller, D. a.; DiSalvo, F. J.; Abruña,

H. D. Nat. Mater. 2012, 12 (1), 81.

(13) Wang, G.; Xiao, L.; Huang, B.; Ren, Z.; Tang, X.; Zhuang, L.; Lu, J. J. Mater. Chem. 2012,

22 (31), 15769.

(14) Zhang, N.; Chen, X.; Lu, Y.; An, L.; Li, X.; Xia, D.; Zhang, Z.; Li, J. Small 2014, 10 (13),

2662.

(15) Kim, D.; Resasco, J.; Yu, Y.; Asiri, A. M.; Yang, P. Nat. Commun. 2014, 5 (May), 4948.

(16) Xu, Z.; Lai, E.; Shao-Horn, Y.; Hamad-Schifferli, K. Chem Commun 2012, 48 (111), 5626.

(17) Grillo, V.; Rotunno, E. Ultramicroscopy 2013, 125, 97.

(18) Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.;

Chiarotti, G. L.; Cococcioni, M.; Dabo, I.; Dal Corso, A.; de Gironcoli, S.; Fabris, S.; Fratesi, G.;

Gebauer, R.; Gerstmann, U.; Gougoussis, C.; Kokalj, A.; Lazzeri, M.; Martin-Samos, L.; Marzari,

N.; Mauri, F.; Mazzarello, R.; Paolini, S.; Pasquarello, A.; Paulatto, L.; Sbraccia, C.; Scandolo, S.;

Sclauzero, G.; Seitsonen, A. P.; Smogunov, A.; Umari, P.; Wentzcovitch, R. M. J. Phys. Condens.

Matter 2009, 21, 395502.

(19) Bahn, S. R.; Jacobsen, K. W. Comput. Sci. Eng. 2002, 4 (3), 56.

(20) Hammer, B.; Hansen, L.; Nørskov, J. Phys. Rev. B 1999, 59 (11), 7413.

(21) Vanderbilt, D. Phys. Rev. B 1990, 41 (11), 7892.

(22) Dannenberg, A.; Gruner, M. E.; Hucht, A.; Entel, P. Phys. Rev. B - Condens. Matter Mater.

Phys. 2009, 80 (24), 1.

(23) Monkhorst, H. J.; Pack, J. D. Phys. Rev. B 1976, 13 (12), 5188.

(24) Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J. R.; Bligaard, T.;

Jónsson, H. J. Phys. Chem. B 2004, 108 (46), 17886.

Page 89: Conversion Chenlu Xie

75

(25) Wang, S.; Petzold, V.; Tripkovic, V.; Kleis, J.; Howalt, J. G.; Skúlason, E.; Fernández, E. M.;

Hvolbæk, B.; Jones, G.; Toftelund, A.; Falsig, H.; Björketun, M.; Studt, F.; Abild-Pedersen, F.;

Rossmeisl, J.; Nørskov, J. K.; Bligaard, T. Phys. Chem. Chem. Phys. 2011, 13 (46), 20760.

(26) Kuhl, K. P.; Hatsukade, T.; Cave, E. R.; Abram, D. N.; Kibsgaard, J.; Jaramillo, T. F. J. Am.

Chem. Soc. 2014, 136 (40), 14107.

(27) Peterson, A. A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Nørskov, J. K. Energy Environ.

Sci. 2010, 3 (9), 1311.

(28) Cullity, B. D.; Stock, S. R. Elements of X-Ray Diffraction, 3rd ed.; Pearson, 2001.

(29) Sinfelt, J. H.; Via, G. H.; Lytle, F. W. J. Chem. Phys. 1980, 72 (9), 4832.

(30) Knauer, A.; Eisenhardt, A.; Krischok, S.; Koehler, J. M. Nanoscale 2014, 6 (10), 5230.

(31) Ma, S.; Sadakiyo, M.; Heima, M.; Luo, R.; Haasch, R. T.; Gold, J. I.; Yamauchi, M.; Kenis,

P. J. A. J. Am. Chem. Soc. 2016, jacs. 6b10740.

(32) Feng, X.; Jiang, K.; Fan, S.; Kanan, M. W. J. Am. Chem. Soc. 2015, 137 (14), 4606.

(33) Chen, Y.; Li, C. W.; Kanan, M. W. J. Am. Chem. Soc. 2012, 134 (49), 19969.

(34) Hansen, H. A.; Varley, J. B.; Peterson, A. A.; Nørskov, J. K. J. Phys. Chem. Lett. 2013, 4 (3),

388.

(35) Kauffman, D. R.; Alfonso, D.; Matranga, C.; Qian, H.; Jin, R. J. Am. Chem. Soc. 2012, 134

(24), 10237.

(36) Zhu, W.; Michalsky, R.; Metin, Ö.; Lv, H.; Guo, S.; Wright, C. J.; Sun, X.; Peterson, A. a;

Sun, S. J. Am. Chem. Soc. 2013, 135 (45), 16833.

(37) Zhu, W.; Zhang, Y.-J.; Zhang, H.; Lv, H.; Li, Q.; Michalsky, R.; Peterson, A. a.; Sun, S. J.

Am. Chem. Soc. 2014, 136 (46), 16132.

(38) Liu, M.; Pang, Y.; Zhang, B.; De Luna, P.; Voznyy, O.; Xu, J.; Zheng, X.; Dinh, C. T.; Fan,

F.; Cao, C.; de Arquer, F. P. G.; Safaei, T. S.; Mepham, A.; Klinkova, A.; Kumacheva, E.; Filleter,

T.; Sinton, D.; Kelley, S. O.; Sargent, E. H. Nature 2016, 537 (7620), 382.

(39) Verma, S.; Kim, B.; Jhong, H. R. M.; Ma, S.; Kenis, P. J. A. ChemSusChem 2016, No. 2,

1972.

(40) Shi, C.; Hansen, H. a; Lausche, A. C.; Nørskov, J. K. Phys. Chem. Chem. Phys. 2014, 16 (10),

4720.

(41) Escudero-Escribano, M.; Verdaguer-Casadevall, A.; Malacrida, P.; Grønbjerg, U.; Knudsen,

B. P.; Jepsen, A. K.; Rossmeisl, J.; Stephens, I. E. L.; Chorkendorff, I. J. Am. Chem. Soc. 2012,

134 (40), 16476.

(42) Stephens, I. E. L.; Bondarenko, A. S.; Grønbjerg, U.; Rossmeisl, J.; Chorkendorff, I. Energy

Environ. Sci. 2012, 5 (5), 6744.

(43) Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.; Yu, C.; Liu, Z.; Kaya, S.; Nordlund,

D.; Ogasawara, H.; Toney, M. F.; Nilsson, A. Nat. Chem. 2010, 2 (6), 454.

Page 90: Conversion Chenlu Xie

76

(44) Kitchin, J. R.; N??rskov, J. K.; Barteau, M. A.; Chen, J. G. J. Chem. Phys. 2004, 120 (21),

10240.

(45) Hansen, H. A.; Shi, C.; Lausche, A.; Peterson, A.; Nørskov, J. K. Phys. Chem. Chem. Phys.

2016, No. 111, 9194.

(46) Peterson, A. A.; Nørskov, J. K. J. Phys. Chem. Lett. 2012, 3 (2), 251.

(47) Greeley, J.; Nørskov, J. K.; Kibler, L. A.; El-Aziz, A. M.; Kolb, D. M. ChemPhysChem 2006,

7 (5), 1032.

(48) Greeley, J.; Jaramillo, T. F.; Bonde, J.; Chorkendorff, I. B.; Nørskov, J. K. Nat. Mater. 2006,

5 (11), 909.

(49) Knauer, A.; Eisenhardt, A.; Krischok, S.; Koehler, J. M. Nanoscale 2014, 6 (10), 5230.

(50) Feng, X.; Jiang, K.; Fan, S.; Kanan, M. W. J. Am. Chem. Soc. 2015, 137 (14), 4606.

(51) Chen, Y.; Li, C. W.; Kanan, M. W. J. Am. Chem. Soc. 2012, 134 (49), 19969.

(52) Kitchin, J. R.; Nørskov, J. K.; Barteau, M. A.; Chen, J. G. Phys. Rev. Lett. 2004, 93 (15),

156801.

(53) Mavrikakis, M.; Hammer, B.; Nørskov, J. K. Phys. Rev. Lett. 1998, 81 (13), 2819.

(54) Hammer, B.; Norskov, J. K. Nature 1995, 376 (6537), 238.