universidade estadual de campinas instituto de … · 2018-08-31 · “ainda que eu falasse as...

137
UNIVERSIDADE ESTADUAL DE CAMPINAS INSTITUTO DE QUÍMICA PRISCILA DESTRO GOLD-COPPER COLLOIDAL NANOPARTICLES: INSIGHTS ABOUT THE ALLOY FORMATION AND APPLICATION IN CO OXIDATION REACTION NANOPARTÍCULAS COLOIDAIS DE COBRE-OURO: CONSIDERAÇÕES SOBRE A FORMAÇÃO DA LIGA E APLICAÇÃO NA REAÇÃO DE OXIDAÇÃO DE CO CAMPINAS 2016

Upload: duongnhi

Post on 18-Jan-2019

213 views

Category:

Documents


0 download

TRANSCRIPT

UNIVERSIDADE ESTADUAL DE CAMPINAS

INSTITUTO DE QUÍMICA

PRISCILA DESTRO

GOLD-COPPER COLLOIDAL NANOPARTICLES:

INSIGHTS ABOUT THE ALLOY FORMATION AND APPLICATION IN CO OXIDATION REACTION

NANOPARTÍCULAS COLOIDAIS DE COBRE-OURO:

CONSIDERAÇÕES SOBRE A FORMAÇÃO DA LIGA E APLICAÇÃO NA REAÇÃO DE OXIDAÇÃO DE CO

CAMPINAS 2016

PRISCILA DESTRO

GOLD-COPPER COLLOIDAL NANOPARTICLES:

INSIGHTS ABOUT THE ALLOY FORMATION AND APPLICATION IN CO OXIDATION REACTION

NANOPARTÍCULAS COLOIDAIS DE COBRE-OURO: CONSIDERAÇÕES SOBRE A FORMAÇÃO DA LIGA E APLICAÇÃO NA REAÇÃO

DE OXIDAÇÃO DE CO

Doctoral Thesis presented to the Institute of Chemistry of the

University of Campinas as part of the requirements to obtain the

title of Doctor in Sciences.

Supervisor: Profa. Dra. Daniela Zanchet

THIS COPY CORRESPONDS TO THE FINAL VERSION OF THE PhD THESIS

DEFENDED BY PRISCILA DESTRO AND SUPERVISED BY PROF. DR. DANIELA

ZANCHET.

CAMPINAS 2016

“Ainda que eu falasse as línguas dos homens e dos anjos,

e não tivesse amor,

seria como o metal que soa ou como o sino que tine.

E ainda que tivesse o dom de profecia,

e conhecesse todos os mistérios e toda a ciência,

e ainda que tivesse toda a fé, de maneira tal que transportasse os montes,

e não tivesse amor, nada seria.”

1 Cor 13, 1-2

Aos meus pais, Valdivino e Maria de Lourdes,

por me ensinarem o que há de mais importante na vida.

Às minhas avós, Benedita e Joana, exemplos de

mulheres fortes sempre tão presentes nesta caminhada.

Com muito amor,

Dedico.

AGRADECIMENTOS

Acredito que escrever os agradecimentos de uma tese seja tão difícil quanto escrever a própria

tese. Foram anos de aprendizado que vão muito além de um título, foram aprendizados de

vida. Termino este trabalho com a plena certeza de que tudo que valeu a pena.

Agradeço primeiramente a Deus, por me dar fé e força para seguir sempre em frente, mesmo

nas vezes em que o caminho à minha frente era tão sinuoso. Por me manter de pé quando eu

achei que não conseguiria mais e por, acima de tudo, se fazer presente no olhar de tantas

pessoas que cruzaram o meu caminho.

Aos meus pais, pelo amor e apoio em todos os momentos. Por acreditarem em mim e fazerem

do meu sonho o sonho deles. Ao meu pai seu pelo exemplo de caráter e honestidade que me

acompanha sempre, à minha mãe por seu exemplo de dedicação e amor que faz toda a

diferença. Aos dois por me ensinarem a seguirem sempre no caminho do bem. À minha irmã

pela sua doçura e delicadeza me apoiando em cada passo que eu dou.

À minha orientadora, Profa. Daniela Zanchet. Acredito que o exemplo é a melhor forma de

passar uma mensagem e agradeço pela honra de conviver com você nestes anos e ver seu

exemplo de mulher e profissional tão ética e dedicada. Você fez total diferença na minha

formação! Obrigada por me motivar e acreditar sempre em mim! Por me ensinar muito mais

do aquilo que está na minha tese e me incentivar a ir sempre muito mais além. Se fosse para

recomeçar, tinha que ser sob a sua impecável orientação novamente!

Aos meus amigos do Grupo de Catálise e Nanomateriais. Crescemos muito nestes últimos

anos e levarei para sempre os maravilhosos momentos com vocês... a projeção das linhas

catalíticas, as madrugadas no LNLS, os cafés das 16:00, os churrascos, almoços, o grupo

caminhada e o apoio diário. Agradeço muito à primeira geração do nosso grupo. Ao Felipão

por estar ao meu lado desde a entrevista para a seleção do doutorado em Lavras. Desculpe por

te desesperar naquele dia, mas você ainda vai me agradecer por isso! Ao Zay pela sua

sinceridade e apoio em todas as horas, principalmente pelos conselhos práticos e gambiarras

geniais! À Tathi por compartilhar tantos momentos importantes comigo e, mesmo quando

desesperada, me ajudar a manter a calma. Ao Babá pela sua tranquilidade e paciência em tudo

que faz, me ensinando a pensar antes de falar ou tomar qualquer decisão. Ao Lulu pelo seu

bom humor e amizade que fizeram toda a diferença no meu dia-a-dia. À Monique por seu

apoio, carinho e por nos ajudar a organizara bagunça. Sem palavras para agradecer tudo que

vocês fizeram por mim! Agradeço à Danielle pela amizade e por me deixar ainda mais curiosa

no estudo de ligas metálicas. Ao Doctor Daniel por sua incrível memória científica e pelo seu

grande apoio neste trabalho. À Debs por me ajudar sempre que precisei com alegria e

amizade. Ao Arthur pela ótima convivência com todo o grupo neste último ano. À Tanna por

dar continuidade a este trabalho. Aos alunos de iniciação científica que passaram pelo

laboratório, Guilherme, Pedro, Thaís, Lanousse, Ana Carmem, Henrique, Débora, Luigi e

Guilherme por tornarem o trabalho muito mais divertido. Também agradeço muito Lili e

Amandinha, nossas agregadas tão queridas.

Ao Instituto de Química da Unicamp, funcionários e professores, por fornecerem toda a infra-

estrutura para a realização deste trabalho.

Ao Prof. José Maria Corrêa Bueno do LabCat-UFSCar e seus alunos por toda a colaboração

com nosso grupo.

Aos pesquisadores e funcionários do LNLS pelo apoio em tantas semanas de medidas durante

estes anos de doutorado.Em especial à Daniela Coelho, Cristiane Rodella e Fábio Zambello

por toda a ajuda nos momentos de desespero.

Aos membros da banca, por aceitarem o convite e por dedicarem o tempo tão corrido de vocês

para a este trabalho.

Aos meus amigos e familiares por recarregarem minhas energias a cada encontro e abraço,

tornando a caminhada muito mais leve. Às minhas amadas avós pela torcida de sempre.

Agradeço em especial ao Felipe, por aguentar meu desespero nos últimos meses e por seu

carinho, atenção e paciência que tanto me ajudaram.

Às queridas Amanda,Larissa,Eugênia e Renata por me receberem tão bem em casa.

Ao CNPq pela concessão da minha bolsa de estudos no Instituto de Química. À CAPES/

Programa Ciência Sem Fronteiras pela concessão da bolsa para a realização de parte deste

trabalho no Istituto Italiano di Tecnologia, parte muito importante da minha formação.

For all the amazing people that I met during my year abroad. For all the technicians,

researchers and students from the Nanochemistry Department of the Italian Institute of

Technology. My special thanks to Dr. Liberato Manna for the opportunity and to Dr.

Massimo Colombo for his great contribution in the development of this work. Grazie Mille!!

Este trabalho é fruto do apoio de todos vocês! Muito obrigada!!!

RESUMO

A catálise é uma área de plena ascensão no cenário científico atual. Desta forma, o

desenvolvimento de novos catalisadores e a compreensão das modificações do material

durante o processo catalítico tornam-se áreas de pesquisa extremamente atrativas. Diante

deste panorama, o objetivo geral deste trabalho foi a utilização do sistema bimetálico cobre-

ouro (AuCu) como um catalisador modelo para aplicação em catálise heterogênea, buscando a

compreensão dos mecanismos envolvidos na formação da liga AuCu e sua posterior aplicação

em catálise. Primeiramente buscou-se a síntese, caracterização e compreensão do mecanismo

de formação de nanopartículas daliga AuCu, pelo método coloidal utilizando a estratégia one-

pot. Durante esta etapa do projeto foram realizados ensaios de espectroscopia da estrutura fina

da borda de absorção de raios X (XAFS) in situ durante a formação das nanopartículas

identificando as etapas cruciais no processo de formação de AuCu. Após a identificação dos

pontos críticos envolvidos durante a síntese foi possível aperfeiçoar o processo obtendo

partículas esféricas e homogêneas com diâmetro em torno de 14 nm e composição de

Au0,75Cu0,25, Au0,60Cu0,40 e Au0,50Cu0,50 alterando a temperatura final de reação. Utilizando

diversas técnicas de caracterização, foi possível identificar o mecanismo de formação como

um processo digestion ripening-like, onde uma fase enriquecida com cobre é formada

juntamente com nanopartículas de ouro e, com o andamento da reação, esta fase é

completamente digerida e a partícula enriquecida com cobre pela difusão em estado sólido.

Após a compreensão dos aspectos fundamentais envolvidos na síntese, as partículas

produzidas foram suportadas em sílica e alumina e aplicadas em reação de oxidação de CO,

onde as amostras apresentaram elevada conversão e estabilidade em condições reacionais.

Ensaios alternando ciclos reacionais e pré-tratamentos redox comprovaram o processo de

segregação da liga AuCu, fato que está intimamente relacionado à sua capacidade catalítica

bem como à estabilidade do catalisador em condições reacionais Combinando as técnicas de

XAFS e difração de raios X in situ foi possível comprovar o efeito do suporte no processo de

segregação/ re-formação da liga AuCu durante condições reacionais e pré-tratamentos, uma

característica extremamente importante que abre novos horizontes para a aplicação de ligas

metálicas na catálise heterogênea.

ABSTRACT

Catalysis is a strategic area in the current scientific scenario. Thus, the development of new

catalysts and understanding of the material changes during the catalytic process becomes an

extremely attractive research area. Based on that, the aim of this work was the exploitation of

gold-copper (AuCu) nanoparticles as a model heterogeneous catalyst, seeking to understand

the mechanisms involved in the AuCu alloy formation and its subsequent application in

catalysis. Firstly, the AuCu alloy nanoparticles were synthesized and characterized to

understand the alloy formation mechanism by the colloidal method using the one-pot

approach. During this part of the project, the formation of the nanoparticles was probed by in

situ X-ray absorption of fine structure (XAFS) spectroscopy, identifying the crucial steps in

the AuCu nanoparticles formation process. After the identification of the main steps involved

in the synthesis, it was possible to improve the process obtaining spherical and homogeneous

nanoparticles with a diameter of around 14 nm and composition of Au0.75Cu0.25, Au0.60Cu0.40

and Au0.50Cu0.50, by changing the reaction final temperature. Using different characterization

techniques, it was possible to identify the mechanism of formation as a digestion ripening-like

process where a copper-riched phase is formed together with the gold nanoparticles; with the

progress of the reaction, this phase is completely digested and the alloy nanoparticles are

formed by a solid state diffusion process. After understanding fundamental aspects involved

on the synthesis, the AuCu nanoparticles were supported on alumina and silica and applied to

CO oxidation reaction, where the samples showed high conversion and stability at reaction

conditions. Tests alternating reaction cycles and redox pretreatments revealed a partially

reversible dealloying/alloying process, which is closely related to the catalytic potential and

the stability of the catalyst under reaction conditions. By combining the in situ techniques

XAFS and X ray diffraction, it was possible to highlight the effect of the support in the

process of alloying/ dealloying of the AuCu nanoparticles under reaction conditions and

pretreatments, an extremely important feature that opens up new horizons for the application

of metallic alloys in heterogeneous catalysis.

Abbreviations and Symbols

AuCu NPs Gold-copper nanoparticles

BF-TEM Bright-field transmission electron microscopy

EDS Energy-dispersive X-ray spectroscopy

EXAFS Extended X-ray absorption fine structure

HAADF-STEM High-angle annular dark field– scanning transmission electron

microscopy

ICP Inductively coupled plasma optical emission spectroscopy

LNLS Brazilian Synchrotron Light Laboratory

NPs Nanoparticles

STEM Scanning transmission electron microscopy

TEM Transmission electron microscopy

UV-Vis Electronic absorption spectroscopy in ultraviolet-visible range

XAFS X-Ray Absorption of fine structure

XANES X-ray absorption near edge structure

XAS X-Ray absorption

XPS X-ray photoelectron spectroscopy

XRD X-ray diffraction

Summary

1. Introduction ...................................................................................................................... 14

1.1 General aspects about nanomaterials ........................................................................... 14

1.2 Colloidal synthesis of nanoparticles............................................................................. 15

1.3 The gold-copper alloy ................................................................................................. 18

1.6 The heterogeneous catalysis ........................................................................................ 22

1.4 CO Oxidation reaction................................................................................................. 26

1.5 Goals and thesis presentation ....................................................................................... 29

1.6 Overview .................................................................................................................... 29

1.7 Summary of each chapter ............................................................................................ 29

Chapter 2 .............................................................................................................................. 31

Synthesis of gold-copper nanoparticles by colloidal method varying the compositions as a

function of the synthesis final temperature ............................................................................ 31

2.1 Introduction................................................................................................................. 32

2.2 Experimental section ................................................................................................... 33

2.2.1 Chemicals ............................................................................................................ 33

2.2.2 Synthesis of bimetallic gold-copper NPs .............................................................. 33

2.3 Characterization .......................................................................................................... 34

2.3. Results ....................................................................................................................... 36

2.4. Discussion .................................................................................................................. 44

Chapter 3 .............................................................................................................................. 49

Formation of bimetallic copper-gold alloy nanoparticles probed by in situ XAFS ................. 49

3.1 Introduction................................................................................................................. 50

3.2 Experimental session ................................................................................................... 51

3.2.1 Chemicals ............................................................................................................ 51

3.2.2 Synthesis of bimetallic copper-gold nanoparticles ................................................ 51

3.2.3 Characterization ................................................................................................... 52

3.3 Results and discussion ................................................................................................. 52

3.3.1 Nanoparticles formation ....................................................................................... 54

3.3.2 Mechanistic aspects .............................................................................................. 57

3.4 Conclusions................................................................................................................. 63

Chapter 4 .............................................................................................................................. 65

Incorporation of AuCu nanoparticles on silica: Studies about the stability of nanoparticles and

the application in CO oxidation reaction ............................................................................... 65

4.1 Introduction................................................................................................................. 66

4.2 Experimentalsession .................................................................................................... 67

4.2.1Chemicals and materials ........................................................................................ 67

4.2.2 Synthesis of bimetallic copper-gold NPs .............................................................. 67

4.2.3 CatalyticTests....................................................................................................... 68

4.2.4 Characterization ................................................................................................... 69

4.3 Results ........................................................................................................................ 69

4.4 Discussion ................................................................................................................... 76

4.5 Conclusion .................................................................................................................. 78

Chapter 5 .............................................................................................................................. 79

Au1-xCux supported nanoparticles applied on CO oxidation: Insights about the support effect

using in situ techniques ......................................................................................................... 79

5.1 Introduction................................................................................................................. 80

5.2 Experimental Session .................................................................................................. 80

5.2.1 Chemicals. ........................................................................................................... 80

5.2.2 Synthesis of bimetallic copper-gold NPs. ............................................................. 80

5.2.3 Catalysis preparation ............................................................................................ 81

5.2.4 Characterization ................................................................................................... 81

5.2.5 In situ measurements ............................................................................................ 81

5.2.6 Catalytic Tests...................................................................................................... 82

5.3 Results and discussion ................................................................................................. 83

5.3.1 Catalytic results .................................................................................................... 84

5.3.2 In situ measurements during the activation ........................................................... 87

5.3.3 In situ measurements during the CO oxidation ..................................................... 93

5.3.4 In situ measurements during the reduction ............................................................ 94

Chapter 6 .............................................................................................................................. 98

Global discussion and conclusions ........................................................................................ 98

Bibliography....................................................................................................................... 100

Appendix ............................................................................................................................ 116

14

1. Introduction

1.1 General aspects about nanomaterials

The manipulation of matter at increasingly smaller scales has aroused great fascination by

the humanity since the ancient times. The ability to control different substances and thus

rearrange atoms to produce new materials with fully adjustable features to specific

applications was firstly seen as magic or something close to mythical; nevertheless, the

scientific research and advanced characterization methods led to demystification of matter,

making science the fundamental pillar for the development of society.

In this context, advances in nanotechnology has enabled the development of innovative

particles with revolutionary applications in several important areas such as the development

of new types of cleaners and more efficient ways to storage energy1, the fast and secure

transmission of data and the development of new drugs2 associated to treatments that provide

a better quality of life to the world population. Thus, the development of new synthesis routes

to produce nanoparticles as well as the thorough understanding of the processes behind the

synthesis is of fundamental importance for technological development3.

To fully exploit the nanoparticles, it is necessary to develop efficient and innovative

synthetic routes that lead to a fine-tuning of the properties at the nanoscale. Two different

approaches can be highlighted, the “top-down” and the “bottom-up”4. The “top-down”

approach often uses traditional techniques of micro fabrication, using controlled tools to cut

and shape the materials into the desired shape or array. The most common method is the

lithography, using UV radiation or an ion beam. On other hand, the “bottom-up” approach

uses the physico-chemical properties of materials to assemble the building blocks into a

desired conformation, for example, chemical vapor deposition and colloidal methods. It is

difficult to elect the most efficient approach, each one has its advantages and disadvantages;

however, according to Chaudhuri and Paria,5 the bottom-up approach allows producing

smaller particles and can be more cost-effective in the future because of its accuracy, control

of the entire process and minimizing the energy loss. The colloidal synthesis can generate a

broad library of particles of different compositions, shapes and sizes, to be applied in several

areas.

15

1.2 Colloidal synthesis of nanoparticles

Colloidal nanoparticles have attracted the interest of many research groups because of

its versatility that finds application in various areas of research.6–8This work highlights the use

of chemical synthesis, in which metal precursor is in the cationic form that is reduced in the

presence of protective agents, producing particles stable in a dispersion. By controlling

parameters such as metal/protective agent or metal/reducing agent ratios, reaction temperature

and time it is possible to obtain particles with different shapes, sizes and compositions. Figure

1.1 shows schematically the colloidal synthetic process.

Figure 1.1: Schematic representation of the colloidal synthesis of metallic nanoparticles.

(Adapted from Jia and Schüth9)

According to Berti and Palazzo10, a colloidal system can be defined as particles with at

least one dimension between 1 and 100 nm dispersed in a solvent. These particles are small

enough for the Brownian forces being greater than the gravitational force, maintaining the

particles stable in dispersion.

The first scientific studies of the stability of colloidal dispersions were described by

Faraday in the 19th century,11 describing the reduction of an aqueous solution of chloroaurate

(AuCl4-) and investigating the optical properties. 12,13 An important work was developed by

Turkevich and co-workers14in the50´s, describing a stable dispersion of 13 nm gold

nanoparticles using sodium citrate as reducing agent and stabilizer. The schematic approach

used in Turkevich´s work is shown in equation (1.1).

푥푀 + 푛푒 + 푠푡푎푏푖푙푖푧푒푟 → 푀 (1.1)

16

In this approach, the reducing agent (e.g. hydrogen, alcohol, hydrazine or borohydride)

is mixed with the metal precursor salt in the presence of stabilizing agents (ligands, polymers

or surfactants). The stabilizer prevents the undesired agglomeration, helps the formation of

well dispersed metal particles and may provide a way to functionalize the surface of the

nanoparticles targeting specific applications.9

Kudera and Manna10 listed the main points that determine the success of a colloidal

synthesis route, which are: the control of particles growth by preventing their agglomeration,

an efficient way to collect or isolate the final product, and minimization of by-products

formation.

The study of the mechanisms of nucleation and growth of colloidal nanoparticles

started in the mid-twentieth century by the seminal work by LaMer and was detailed by Reiss

in the 50´s.15,16These authors described the colloidal synthesis as a process with three main

stages. The first one is the formation of monomers, species that are formed from the cationic

precursors connected to the stabilizing molecules. The binding strength and the size of the

monomers influence directly the formation of the colloidal nanoparticles. The amount of

monomers in solution increases until the supersaturation regime is reached, and the second

stage of the LaMer process initiates: the nucleation. This is the most important stage and

influences directly the final structure of the nanoparticles. Once the nuclei are formed, the

growing step initiates. The growth can be carried by aggregation of nuclei or by incorporation

of atoms to the nuclei. Figure 1.2 shows the main stages of the formation of colloidal particles

proposed by LaMer.

Figure1.2: The classic LaMer mechanism for nanoparticles formation. The first stage is the

monomer formation followed by nucleation and growth. (Adapted from Reiss, Vreeland and

Kudera15)

17

The strategy to synthesize metal nanoparticles by colloidal methods is widely explored

in several applications because of its accurate control of size, shape and composition17–19.

Particles with 1-10 nm are considered by some authors as intermediate size between small

molecules and bulk metal.2The figure 1.3 shows the UV-Vis absorption spectra of Au

nanoparticles synthesized by Piella et al.20that illustrate the size-dependence of their optical

properties.

Figure 1.3: UV-Vis absorption spectra of Au nanoparticles as a function of size (From Piella

et. al 20)

Another approach that has been widely used in the synthesis of the metal nanoparticles

is the introduction of a second metal, producing bimetallic particles. Some of the most studied

bimetallic systems are PtPd21, PtNi22, PdCo23 and AuAg24. Figure 1.4 shows some

possibilities in the formation of bimetallic particles.

Figure 1.4: Possible structures presented by bimetallic particles. Type (a) heterostructure, (b)

core-shell structure (c) alloy with random substitution and (d) chemically ordered alloy.

(Adapted from An et al.25)

18

Bimetallic structures can be denominate as ordered when there is a spatial organization

between the components A and B constituting the particle. The first case described in the

figure 1.4a is a heterostructure type (also known as Janus particle) where there is a solid

interface between the two domains, each one formed exclusively by compounds A or B.

Figure 1.4b illustrates the core-shell structures, where one element is in the center of the

particle while the other forms a shell on the surface.

Metal alloys may be formed by two different ways: a random substitution of atoms in

the crystal lattice of the major component (Figure 1.4.c) or by replacement of atoms at

specific positions, usually forming a different crystal structure (Figure 1.4d). The latest is also

known as chemically ordered alloy.

According to Johnsoton et al26, some factors can directly influence the atomic

distribution of a AB bimetallic nanoparticle, such as:

The bond strength between A-A, B-B and A-B: If the A-B bonding is stronger, it

favors the formation of the alloy, if the homonuclear bonds are stronger, a segregated

structure is favored.

Surface energy of A and B bulk phases: The element with the lowest surface energy

on the bulk phase tends to segregate to the surface of the particle.

Charge transfer: Electrons are transferred from the least to the most electronegative

element, favoring the mixture by maximizing attractive coulombic interactions.

Bonding strength of the metal with the surface ligands: For nanoparticles with a ligand

layer on the surface (especially colloidal particles), the element that binds more

strongly to the ligand may remain on the surface of the nanoparticles.

The final structure of the nanoparticle depends on the balance among the various factors

mentioned above as well as on the preparation method and experimental conditions which can

control the kinetics involved in the process.

1.3 The gold-copper alloy

The AuCu is one of the most studied alloy nowadays, however, its potentialities are

exploited since pre-Columbian civilizations of Central America that used the mixture of

copper and gold27, getting the "Tumbaga", an alloy with improved ductility and strength.

Currently, the AuCu alloy is still widely used in jewelry and emerged as a potential system of

19

study in the nanoscience.28 One feature of the AuCu alloy is the alloy formation of any

composition, with the generation of chemically ordered alloys for specific Au:Cu ratios

(known as superlattices)29.

The miscibility between metals was studied thoroughly by Hume-Rothery, and the

main factors are known as the “Hume-Rothery Rule”30. According to this rule, the main

factors that determine the miscibility of two metals are the similarity between atomic radius,

crystal structure, valence and electronegativity. The combination of Au and Cu satisfy two

these requirements since the difference in atomic radius is smaller than 15% (Au = 134pm

and Cu = 117pm) and they have the same crystal structure (face-centered cubic), causing the

random replacement at elevated temperatures30. In the case of nanostructures, studies

developed by Guisbier and co-workers31 for AuCu alloys stability considers another important

factor to the of metal nanoalloys formation: the molar heat of vaporization, which relates the

cohesion energy of the materials. The difference between the molar heat of vaporization of Au

and Cu is less than 15% (ΔHvapAu= 334 kJ.mol-1 and ΔHvapCu=300 kJ/mol-1,indicating

possible formation of the AuCu alloy nanoparticles.

In the study of metal alloys, the Vegard´s Law32 is a very useful tool for quantitative

analysis of each metal content in the disordered AuCu alloy (randomic substitution in the

crystal lattice). The Vegard's Law shows a linear dependence of the lattice parameter with the

composition, as described in the Figure 1.5. In the case of AuCu alloy, a small deviation in

the lattice parameter values can be related to the interaction of Au and Cu within the crystal

lattice. This work was originally described to bulk systems, but the concept of the Vegard's

law is also used for nanostructured systems.

Figure 1.5: Relationship of lattice parameters and composition in the disordered fcc AuCu

system. The symbols on the continuous line indicate some experimental values used by

Okamoto at. al.33.

20

As previously mentioned, another factor that makes the AuCu system extremely

interesting is the formation of superlattices, chemically ordered structures determined by

composition and temperatures that assume different crystalline structure. The superlattices

formation may occur in the following proportions: Au3Cu, AuCu and AuCu3. These structures

are shown in Figure 1.6 in the phase diagram of the AuCu alloy. The phase diagram indicates

the transformations occurred in the AuCu system by different compositions and temperatures.

The structural data are described on the table 1.1and represented at figure 1.7.

Figure 1.6: Phase diagram of Au-Cu (Adapted from Okamoto33)

Table 1.1: AuCu structural data.

Phase Composition

(at. % Cu)

Pearson symbol Space group

(Au,Cu) 0 to 100 cF4 Fm3m

Au3Cu 10 to 38.5 cP4 Pm3m

AuCu (I) 42 to 57 tP4 P4/mmm

AuCu (II) 38.5 to 63 oI40 Imma

AuCu3 (I) 67 to 81 Cp4 Pm3m

AuCu3 (II) 66 to ? tP28 P4mm

21

Figure 1.7: Representative unit cells structures of (a) Au1-xCux (disordered alloy), (b)Au3Cu,

(c) and (d) AuCu and (e) AuCu3. The orange balls represented in (a) indicates the random

substitution of Cu in the Au lattice. The red balls indicate the Cu atoms and the yellow balls

indicate the gold atoms in the structures (b) to (e).

Theoretical studies have provided significant advances in the understanding of AuCu

system. 34,35 Studies evaluating small metal particles (454 atoms) observed that AuCu

particles are not fully homogeneous, as predicted by the bulk phase diagram, but shows a

certain degree of segregation with the copper in the core and a gold shell due to the

minimization of surface energy and maximizing strongest interactions as Au-Au36,37. Guisbier

and colleagues also found a similar this behavior with the segregation of Au to the surface for

various polyhedral nanoparticles, in the size range of 4-10 nm. 31 It is important to remark that

these studies are performed on bare nanoparticles and for other systems (e.g. colloidal

nanoparticles) also the interaction with ligands or solvents has an impact on the particle.

The alloy formation can also interfere in the symmetry of the clusters. Darby and co-

workers36 described the effect of alloying in cluster of 55 atoms. Clusters of Cu55 have an

icosahedral symmetry whileAu55 is amorphous-like; however, the replacement of only one

atom of Au by Cu leads to a reorganization of the structure, with icosahedral symmetry, as

shown in figure 1.8.

Figure 1.8: Simulation Au55, CuAu54 and Cu55 structures presented by Darby et al.38

22

Clusters structure with compositions CuAu, CuAu3 and Cu3Au have also been studied.

Au and Cu atoms were added one by one to a defined seed and the CuAu showed a strong

dependence on the size in the 100 and 200 atoms range. 39 The theoretical studies indicate that

the composition and/or size of the metal alloy can directly change its structure, resulting in

structural and electronic effects that have a direct impact in their applications.

From the catalytic point of view, one of the most explored applications of metal alloys,

the impact of the size and composition have a fundamental importance for the understanding

of the effects involved. 37 The changes in the structure can modify structurally the alloy,

generating preferential adsorption sites or low coordination sites that interfere directly in the

catalytic properties of the material. The electronic effects are related to the band structure that

depends on the composition and particle size. Studies by Kim and co-workers40 showed the

relationship between the composition of Au1-xCux particles and the reactivity in CO2

reduction. The distinct Au:Cu ratio has a direct impact on the d-band position, determining

the reactivity of the catalyst. Figure 1.9 is the surface valence band spectra of Au–Cu

bimetallic nanoparticles collected by high-resolution X-ray photoemission spectroscopy

(XPS)

Figure 1.9: Surface valence band photoemission spectra of Au–Cu bimetallic nanoparticles.

(Available at Kim, D. et. al.40)

1.6 The heterogeneous catalysis

In heterogeneous catalysts, the fundamental principle is the ability of the reactants to

diffuse and to adsorb on the surface of the catalyst and, after reacting, being desorbed

releasing the active site to react again with another reactant molecule. To illustrate the

catalytic process, a generic cycle is showed at figure 1.10. In this process, A and B are the

reactants that adsorb on the catalyst surface, react, and desorb as a product P. The catalyst

23

provides a new potential energy pathway to the process, as described in Figure 1.10, which

compares a non-catalytic reaction versus a catalytic reaction.

Figure 1.10: Potential energy diagram of a heterogeneous catalytic reaction with gaseous

reactants and products and a solid catalyst. The red line indicates the non-catalyzed reaction

and the green line, the catalyzed pathway.41

It is possible to observe in Figure 1.10 that the non-catalytic reaction, in which A and

B has to collide with sufficient energy to overcome the activation barrier forming P, has an

energy barrier much higher that the catalytic route.

In the catalytic route, the initial step is the adsorption of A and B on the catalyst

surface in a spontaneous process, followed by formation of a complex between A-B- catalyst

surface resulting the product P adsorbed on the catalyst surface. The process of formation of

this complex has a significantly lower activation energy compared to the non-catalyzed

reaction. Finally, the product P formed is desorbed from the catalyst leaving the catalytic site

free for a new cycle.41

Most of the heterogeneous catalysts are solids, composed by a metallic phase highly

dispersed on a matrix with high surface area.42 The matrix used to support the metallic phase

can be composed of metal oxides, aluminosilicates, carbonaceous compounds among others.

The metal surface is normally the active phase, and therefore is the crucial part in the design

of a catalyst. The highly dispersed metal leads to a greater accessibility of reactants to the

active phase, improving the catalytic capacity of the material.

There are different methods to incorporate the metal phase to supports with high

surface area. One of the most applied is the deposition- precipitation method43, where the pH

of the solution containing the metallic precursors and the support is altered inducing the

24

formation of the metallic phase on the support surface. For the incipient-wetness method, only

the pores of the support are filled with precursor solution, to prevent deposition on the

external surface of the catalyst44. Another used approach is the impregnation and drying of the

metallic precursors in solution on the support.44 The deposition-precipitation and incipient

wetness are defined as conventional methods, usually used for monometallic material,

however for bimetallic alloys are not the best option due to the difficult to keep the size and

composition homogeneity for all the supported particles. In this context, the great advances in

the development of nanomaterials fit perfectly to this goal: increase the available area by

decreasing the size of the metal particles with high control of their properties.46–48To achieve a

good dispersion and homogeneity of the bimetallic phase it is possible to highlight the use of

colloidal metal nanoparticles, since the characteristics of the metallic phase are defined before

the incorporation, as described in the Figure 1.11.

Figure 1.11: Scheme comparing conventional incorporation versus the colloidal method.

(Adapted from Sonström et. al. 49)

Besides the fact that the available area increases by decreasing the particle size, it is

described in the literature that some reactions are highly sensitive to particle size (structure

sensitive reactions).50 The work by Haruta in the 80´s51 represents a breakthrough in the

history of catalysis, showing the catalytic activity of Au particles in the CO oxidation

reaction, revolutionizing the application of nanomaterials in catalysis and opening a new and

extensive research field.

25

For a better comprehension of the catalytic process, the development of model catalysis is

extremely important, since it allows a deep understanding of the parameters that determine the

catalyst activity. Ideally, a model supported catalyst has homogeneous particles on its surface,

as described in figure 1.12, to evaluate each factor individually.

Figure 1.12: Scheme of model supported catalyst.

The composition of the metal phase is an important parameter to be evaluated in

model catalysts. The versatility of metal alloys makes this system extremely attractive for

applications in catalysis52. Many of the works disclose the use of alloys with enhanced

specific catalytic activity and selectivity when compared to the corresponding monometallic

catalysts, as shown in Table 1.3

Table 1.3: Examples of metal alloys used in catalysis.

Metal Reaction

Temperature

(°C)

Conversion

(%) Reference

Pt

Water-Gas Shift

Reaction 300 70 Xu et al., Journal of

Catalysis, 201253

Ru 300 80

PtRu 300 85

Pt CO oxidation 150 10 Zhou et al., Advanced

Functional Materials,

200754 AuPt 150 80

26

Au 100 50 Liu et al., Catalysis Today,

201155

Cu CO oxidation 100 0

Au3Cu 100 100

Au CO oxidation 350 20 Liu et al., Journal of

Physical Chemistry B,

200556

Ag 350 0

Au3Ag 350 100

1.4 CO Oxidation reaction

A probe reaction for studies in catalysis is the CO oxidation. The catalytic oxidation of

CO has been studied extensively because of its great importance in fundamental research as

well as in practical applications, such as in air clean-up exhaust-gas control in cars, and in the

removal of CO traces in H2 stream for proton-exchange membrane fuel cells. 57–60

For many years, gold was considered catalytically inactive because of its high

stability, being unexplored in catalysis.61 However, Haruta showed the size dependence of

Au62 nanoparticles in the CO conversion, revolutionizing the application of Au in catalysis.63–

65

퐶푂 + 12푂 ⥨ 퐶푂 + 1 2푂 (1.2)

This reaction is normally catalyzed by noble metals or small Au nanoparticles. The

mechanism is described by the following reactions.41

27

An important step in the CO oxidation reaction is the activation of the O2. Thereby,

the choice of an appropriate support plays an important role in the catalyst performance.

Remediakis57 and co-workers presented a theoretical study showing that TiO2 is an efficient

support in the catalytic CO oxidation reaction, since it facilitates the activation of O2 at the

metal-support interfacial sites. (Figure 1.13).

Figure 1.13: CO oxidation mechanism for Au10 supported on TiO2 (110) proposed by

Remediakis et. al. 57

The use of supports considered "inerts" for oxidation reactions such as Al2O3 and SiO2

was studied by Schubert et al.66 and in this case, Au particles are less active because the

activation of O2 molecules occur in Au low coordination sites, competing with the adsorption

of CO molecules and decreasing its catalytic performance.67,68

Since the successful application of Au nanoparticles in CO oxidation reaction, the use

of bimetallic nanoparticles involving gold has also become an important research field.69,70

The main motivation is related to the fact that the O2 is not able to adsorb well on Au surface,

so the alloy formation can generate different sites on the surface that improves the O2

adsorption. Metals as silver, cobalt and copper are mostly used, facilitating the adsorption and

activation of oxygen on the surface.

28

Figure 1.14: Model structure of CO and O2 adsorption on the Au25Ag30 alloy surface. Au

atoms are represented by yellow balls, Ag atoms are represented in gray, oxygen in red and

carbon in white.71

In the case of AuCu alloy, the catalyst is more active than monometallic ones,

subjected to the same reaction conditions.72 A list of factors can contribute to the greater

activity, such as the formation of Cu sites that facilitates the adsorption of oxygen, as well as

the change of alloy electronic properties. 73,74

Another fact that affects the catalytic ability of the material is the nature of the metal-

support interaction75,76 and, in the case AuCu, this is an additional factor that makes this

system even more attractive from the catalytic point of view. It is described in the literature

that, under oxidizing conditions, Cu in the alloy can be oxidized forming CuOx on the surface

of the material that can act as (a) a layer between the metal particle and the support,

enhancing the support-metal interaction and thus preventing sintering,72 (b) forming sites of

CuOx between the metal and the support that facilitate activation of O2 or even (c) forming a

layer of CuOx on surface of the nanoparticle preventing the sintering77. All these factors are

extremely useful and can be assessed also in the application of other existing bimetallic

systems, opening a large field of study for the application of bimetallic nanoparticles under

redoxconditions.78,79

29

1.5 Goals and thesis presentation

The general goal of this thesis was to study the formation of colloidal AuCu alloy

nanoparticles and their application in heterogeneous catalysis, using the CO oxidation

reaction as a model reaction.

Specific goals:

Understand the steps involved in the AuCu nanoparticles formation controlling

the size and composition by the colloidal method.

Produce catalysts by impregnation of AuCu nanoparticles on silica and

alumina and evaluate the structure-activity relationship in CO oxidation.

1.6 Overview

The main idea of this work was to explore the potential of colloidal approach for the

synthesis of bimetallic nanoparticles for applications in catalysis. The metal system chosen

was AuCu, a highly versatile alloy for catalytic applications.

After understanding the mechanism involved during colloidal synthesis, AuCu

nanoparticles produced were supported on silica and alumina to evaluate the stability and

catalytic potential of the material produced.

After the identification of critical points of AuCu colloidal synthesis and the

application of this system in CO oxidation reaction, in situ characterizations experiments

using X-ray absorption of fine structure (XAFS) spectroscopy and X ray diffraction (XRD) at

the Brazilian Synchrotron Light Laboratory (LNLS) were conducted for a deeper

characterization of our system.

1.7 Summary of each chapter

For the AuCu nanoparticles synthesis, the method described by Motl and co-workers81

was optmized to obtain a better control of size and composition depending on the final

temperature of the synthesis. The main results of this part are presented in Chapter 2, using

different characterization techniques and evaluating aliquots taken at different reaction stages

to infer about the mechanism of the synthesis.

30

For a deeper understanding of the synthesis mechanism, AuCu nanoparticles were

synthesized and evaluated in situ using XAFS at LNLS. For this step, it was designed a home-

made reactor adapted for XAFS measurements, allowing following the modifications of Au

and Cu precursors in real time. These results are presented in Chapter 3.

The synthesis of nanoparticles with well-defined size and composition were used to

produce heterogeneous catalysts and Chapter 4 presents the application of the

Au0.75Cu0.25/SiO2catalyst in CO oxidation, focusing in the impact of the redox pretreatments

in the performance and stability of the catalyst.

Finally, the impact of alloy composition and support nature of several catalysts were

addressed. This part of the work explore in detail by in situ XAFS and XRD allowing a deep

understanding of the effects involved during the pretreatments and catalytic cycles as well as

to assess the effect of the support in the alloy/dealloying processes of supported AuCu

nanoparticles. These results are presented in the Chapter 5.

Chapter 6 presents a general discussion of the work, the main conclusions and

perspectives.

31

Chapter 2

Synthesis of gold-copper nanoparticles by colloidal method varying the compositions

as a function of the synthesis final temperature

The content of this chapter is an adaptation of the article entitled

“Au1-xCux colloidal nanoparticles synthesized via a one-pot

approach: understanding the temperature effect on the Au:Cu

ratio” by Priscila Destro, Massimo Colombo, Mirko Prato,

Rosaria Brescia, Liberato Manna and Daniela Zanchet published

for RSC Advances.

Reference: RSC Adv.,2016, 6, 22213.

My contribution to this work was the planning of the

experiment, synthesis of Au1-xCux nanoparticles, the

characterizations by UV-Vis, BF-TEM, XRD and ICP, the data

analysis and writing the publication.

32

2.1 Introduction

Advances in the synthesis of nanomaterials have contributed to significant development in

fields such as sensors, photonics, drug-delivery systems and, especially, to fulfill the

increasing demand for new materials and technologies for energy.81,82

Among the huge variety of nanostructured materials, it is important to highlight the use of

metallic nanoparticles (NPs), where the metal confined in small portion has distinct properties

from those of the bulk. In this context, the control of size, shape and composition of the metal

NPs represents an important way to obtain new materials for many applications.58,83

The synthesis of gold NPs by colloidal methods has been extensively described using

different approaches, enabling accurate control of their properties.84,85 The use of Au-based

bimetallic NPs has also been presented in the literature as an interesting strategy for the

design of new materials.86,87 In this context, the AuCu alloy is an interesting example of how

the electronic properties of the alloy differ from the simple sum of the properties of Au and

Cu.29,88 For example, the Au:Cu ratio reflects directly on the position of the d-band center

relative to the Fermi level, which has a strong correlation with the reactivity of materials,

especially for applications in catalysis.40 Another point to be featured in the case of the AuCu

alloy is the possible formation of chemically ordered alloys. Some metallic nanoparticles such

as CoPt89, FePt90 and AuCo91 are able to form chemically ordered structures at specific

compositions. This ordering can affect different properties of the materials since the distinct

symmetry alters the resulting density of states.92 In the case of AuCu, it is possible to observe

chemically ordered alloys of Au3Cu, AuCu and AuCu3 compositions.

AuCu NPs can find interest in many fields of applications. Their tunable optical properties,

which depend on the shape and composition of nanoparticles, have been explored, for

example, in biosensors applications,93–95 SERS detection88 and photothermal therapy97,

showing exciting results. Another important research field in which AuCu NPs have showed

interesting and unique properties is heterogeneous catalysis, where supported bimetallic NPs

are more active and stable than the monometallic ones.33,97

The study of nanoalloys formation and growth mechanisms has been investigated by

different groups because of the applicability of these systems in several areas.18,37,99 Among

the methods to obtain nanometric metallic alloys reported in the literature, a widely used

strategy is the formation of a metallic seed followed by the incorporation of a second metal, in

a two-step growth mechanism.100,101 By using a seed of the metal with lowest reduction

33

potential, the second metal can be incorporated by galvanic replacement as nicely explored by

Xia and co-workers to produce a large library of nanoalloys starting from Ag seeds.102 In the

case of CuAu system, Liu et al.103 firstly synthesized Cu NPs, and then by partial replacement

of Cu by Au they obtained a Cu@Au core-shell structure. The alloy formation took place by

heating the core-shell NPs, through a unidirectional diffusion mechanism in solid state. In a

more general case, when the galvanic replacement is not favored, the reduction of the second

metal can be done by using a reducing agent. Chen et al.104 successfully obtained intermetallic

CuAu NPs using Au seeds.

In this work, we explore the one-pot strategy (i.e. all reagents are brought into contact at

the beginning of the reaction) to synthesize AuCu NPs with different compositions. In this

approach, characteristics such as the final size, composition and chemical ordering can be

varied by tuning the reaction parameters, i.e. metal: metal ratio, metal: ligand ratio and

temperature. We produced Au1-xCux spherical NPs of 14 nm in diameter, with a fine control

of their composition obtained by optimization of the final reaction temperature; we probed in

detail the alloy formation by combining different methods of characterization.

2.2 Experimental section

2.2.1 Chemicals

HAuCl4.3H2O (99%), and Cu(acac)2 (99%) were used as metal precursors. Oleylamine

(Oley, 70%) and oleic acid (OlAc, 90%) were used as protective agents. 1,2-hexadecanediol

(90%) was used as reducing agent and 1-octadecene (1-ODE, 90%) as solvent. All reactants

were purchased from Sigma-Aldrich and used without further purification.

2.2.2 Synthesis of bimetallic gold-copper NPs

The synthesis of bimetallic AuCu NPs was adapted from the literature.80 Briefly, 45 mg of

HAuCl4.3H2O and 30 mg of Cu(acac)2 (molar ratio Au:Cu 1:1) were added to a three-necked

round-bottom flask containing 5 mL of 1-ODE, 800 L of OlAc, 600 L of Oley and 100 mg

of 1,2-hexadecanediol. The mixture was magnetically stirred at room temperature under

vacuum for30 min, generating a green solution. The mixture was heated up under vacuum up

to 120 °C. After holding at 120 °C for 30 min (step 1), the solution was heated to the final

temperature. At about 200oC, N2 was fed to the system. The final temperature was set at 225,

260 or 280°C to obtain different compositions of Au1-xCux NPs (Figure 2.1). The reaction was

34

annealed at the corresponding final temperature for 30 min (step 2) and then cooled down.

The purification was carried out by adding a toluene: isopropanol (1:5 volume ratio) mixture,

followed by centrifugation at 4000 rpm for 10 min. Precipitated NPs were dispersed by

adding 2 mL of hexane and centrifuged at 2000 rpm for 10 min to remove any undispersed

material. This procedure was repeated twice and the particles were then dispersed in hexane.

The samples were named based on the final synthesis temperature as AuCu_225, AuCu_260

and AuCu_280. To follow the alloy formation, aliquots were taken at different stages of the

synthesis, and named by the temperature (100, 120_I, 120_F, 225_I, 225_F, 260_I, 260_F,

280_I, 280_F). The letters I and F indicate respectively the beginning and end of steps (1 and

2), on which the temperature was hold for 30 min.

Figure 2.1: Heating protocol adopted in the Au1-xCux NPs synthesis. The letters I and F

indicate the initial and final points of each step. Step 1 corresponds to 120 °C for 30 min and

step 2to the final temperature for 30 min. See text for more details.

2.3 Characterization.

Aliquots of approximately 20 µL were collected along the synthesis and diluted in hexane

using a quartz cell of 10 mm path length to characterize the Au1-xCux NP growth by UV-Vis

analysis, using a Varian Cary 5000 UV-Visible-NIR spectrophotometer in single path

configuration within the wavelength range of 200-800 nm with a scanning rate of 10 nm/s.

Overview bright-field transmission electron microscopy (BF-TEM) analyses were carried

out using a JEOL JEM-1011 instrument (thermionic W source, 100 kV high tension).

analyses were performed using a JEOL JEM-2200FS microscope (Schottky emitter, 200 kV

high tension) equipped with a CEOS spherical aberration corrector of the High-angle annular

dark field–scanning TEM (HAADF- STEM) objective lens and an in-column image filter (-

35

type). The shown values for NP diameter and uncertainty were obtained as average and

standard deviation over 500 NPs, respectively. Elemental mapping over the Au1-xCux NP

samples was carried out by energy-dispersive X-ray spectroscopy (EDS) performed in STEM

mode with a Bruker Quantax 400 system with a 60 mm2 silicon-drift detector (SDD). The

reported elemental maps of Cu and Au were obtained by integration of the Cu Kα and Au Lα

after background subtraction and peak deconvolution. For BF-TEM analyses, ~10 μL of the

samples were deposited onto carbon-coated Cu grids. For EDS analyses, carbon-coated

molybdenum grids and an analytical holder with a Be specimen retainer were used.

The elemental analysis of the colloidal NPs was done by Inductively Coupled Plasma

Optical Emission Spectroscopy (ICP) using an iCAP 6000 Thermo Scientific spectrometer.

Each sample was obtained by quenching the reaction at the selected temperature, followed by

purification. Samples were dissolved in HCl/HNO3 3/1 (v/v) overnight, diluted with deionized

water (14 μS), and filtered using a PTFE filter before measurement. The yield of Au and Cu

was defined as %Metal and calculated by the Equation 1.

% = × 100 (Eq.2.1)

where %Metal is the final yield of Au or Cu, Nfinal is the molar amount of metal determined by

ICP and Ninitial the molar amount of metal initially added.

X-ray diffraction (XRD) measurements were performed using a RigakuSmartLab X-ray

diffractometer equipped with a 9 kW Cu Kα (= 1.542 Å) rotating anode, operating at 40 kV

and 150 mA. A zero diffraction silicon substrate was used to collect XRD data on colloidal

NPs. The diffraction patterns were collected at room temperature over an angular range of

20−90°, with a step size of 0.05°. XRD data analysis was carried out using PDXL 2.1

software from Rigaku. As for the ICP measurements, each sample was obtained by quenching

the reaction at the selected temperature, followed by purification. XRD data were used to

estimate the Au1-xCux alloy composition. We used the equation (Eq.2) proposed by Okamoto

et al.32, based on experimental values described in the literature for the Au1-xCux solid solution

phase with face-centered cubic (fcc) structure:

푎 = 0.40784(1− 푥) + 0.36149푥 + 0.01198푥(1− 푥)(Eq.2.2)

Where a is the Au1-xCux lattice parameter and x is the copper atomic fraction, defined as

NCu/(NCu+NAu), where Nx is the number of atoms of the element x.

36

Since the Au1-xCuxalloy system shows a positive deviation from the Vegard’s law, the

equation proposed by Okamoto et al.33 gives a more accurate estimation of the alloy

composition.

X-ray photoelectron spectroscopy (XPS) measurements were performed on a Kratos Axis

Ultra DLD spectrometer, using a monochromatic Al Kα source (15 kV, 20 mA). Wide scans

were acquired at an analyzer pass energy of 160 eV. High-resolution narrow scans were

performed at constant pass energy of 10 eV and steps of 0.1 eV. Photoelectrons were detected

at a takeoff angle of Φ = 0° with respect to the surface normal. The pressure in the analysis

chamber was maintained below 7× 10−9 Torr for data acquisition. The samples were prepared

by drop-casting a few microliters of purified NPs solutions onto a graphite substrate (HOPG,

ZYB quality, NTMDT), which was then transferred to the XPS setup. Data were converted to

VAMAS format and processed using CasaXPS 2.3.16 software. The binding energy scale was

internally referenced to the C 1s peak (BE for C−C = 285 eV). Semi-quantitative analysis was

performed by measuring the areas under selected Au and Cu XPS peaks and by applying

appropriate atomic sensitivity factors, as obtained by the instrument manufacturer.

2.3. Results

Au and Cu NPs exhibit a characteristic absorption band in the visible region due to

plasmon excitation.105 Plasmon band arises from the collective oscillation of the free electrons

induced by the interaction with visible light.17 While Au NPs characteristic plasmon band is

centered at approximately 520 nm, the Cu NPs plasmon appears at 560 nm, and the Au1-xCux

NPs alloy plasmon is usually observed between these two values.80 For this reason, we firstly

followed the evolution of synthesis products by UV-Vis spectroscopy, as shown in Figure 2.2.

At 100 °C (black lines in Figure 2.2a, b and c) the characteristic plasmon resonance of Au

NPs is observed around 520 nm, indicating that the formation of Au NPs already occurred

before reaching this temperature. In fact, around 70 °C the colour of the solution already

changes from orange to deep red and TEM images at 100 °C confirms the formation of NPs

(Figure A1). By increasing the temperature, it is possible to observe a similar profile at 120°C

in all the three tests performed. Above this temperature a red shift occurs and the plasmon

peak becomes asymmetric, indicating the incorporation of copper into the Au NPs. Among

the samples, the red shift increases as a function of final temperature and annealing time:

AuCu_225 shows a plasmon band at 540 nm (Figure 2.2a), AuCu_260 at 550 nm (Figure

2.2b) and AuCu_280 at 552 nm (Figure 2.2c). In contrast, there is no significant change in the

peak width, indicating that the NPs remain stable by the incorporation of Cu.29

37

Figure.2.2 UV-Vis spectra during Au1-xCux NPs synthesis at different final temperatures: (a)

AuCu_225, (b) AuCu_260 and (c) AuCu_280.The red dashed line indicates the position of

Au plasmon and the blue dashed line the Cu plasmon position.

Figure 2.3 shows the BF-TEM images and corresponding size distribution of the final

particles. According to the BF-TEM analyses, the AuCu_225 sample shows spherical and

uniform NPs, with a diameter of 14± 1 nm. The average particle size did not significantly

change with the reaction temperature: we measured 13±1 nm for AuCu_260 and 14±1 nm for

AuCu_280. For AuCu_260, it is possible to observe also some less isotropic particles, which

become the dominant population for the AuCu_280 sample. The lower isotropy in these

samples may be due to a structural rearrangement upon incorporation of a higher fraction of

Cu atoms in the initial Au structure and/or to the higher final temperature.

Elemental analysis by ICP was performed on the final products and on all the aliquots

collected during the syntheses to get more information about the growth mechanism of the

Au1-xCux NPs. The results are presented in Table 2.1.

38

Table 2.1%Cu (at.) at different temperatures by ICP.

Temperature/Sample AuCu_225 AuCu_260 AuCu_280

100* 9.4

120_I* 14.3

120_F* 22.3

225_I 31.6

225_F 24.5 - -

260_I - 27.7 -

260_F - 37.8 -

280_I - - 36.1

280_F - - 48.8

*common steps for the three samples.

ICP values at 100, 120_I and 120_F are common for the three syntheses. It is possible to

verify that around 100°C, only a small amount of Cu is detected in the final product, reaching

22% after 30 min at 120oC.This is about 44% of the Cu added at the beginning of reaction. In

line with UV-Vis spectroscopy data, the increase of the synthesis final temperature results in

higher amounts of copper in the final product, going from 24.5% at 225°C up to 48.8% at

280°C.The annealing time at the final temperature is also important to increase the Cu

incorporation. This is evident especially for the AuCu_260 and AuCu_280 samples, where

an increase of the copper content of about 11% is found with the annealing. It is interesting to

mention that the yield of Au is also affected by the final temperature, decreasing from 40% in

case of the AuCu_225 sample, to 20% for AuCu_280. At a given temperature, longer

annealing increases further the Cu fraction in the final product, but the yield of the synthesis

decreases considerably. The dependence of the yield with time and temperature can be related

to an increased collision of the formed particles, forming larger aggregates that were removed

in the purification step and/or to a poor stability of the NPs with higher content of copper.

Attempts to form Cu enriched alloy NPs, such as Au0.25Cu0.75 by changing the amount of Cu

added at the beginning of the synthesis failed under these synthesis conditions. It is worth to

39

point out that other authors report the formation of Au0.25Cu0.75 NPs31 and Au0.33Cu0.77

35 using

different conditions, but the yields were not reported.

Figure2.3:BF-TEM images and size distribution of (a) AuCu_225, (b) AuCu_260 and (c)

AuCu_280.

Table 2.3 shows the quantitative XPS results, wherein the amount of Cu increases with

temperature. Interestingly, although XPS is a surface technique, probing a few atomic layers

(< 2 nm for Cu and Au)106, it shows Cu atomic % similar to those measured by ICP. These

results indicate a homogeneous distribution of both metals in the synthesis product, at all

stages, ruling out the formation of a core-shell structure with, for example, a Cu-rich shell.

However, it is important to note that a similar result could also be found, for example, if Au

and Cu single-metal particles were formed. This hypothesis is excluded by STEM-EDS

mapping over individual particles (see Figure A2), showing a homogeneous distribution of

both metals in the NPs.

40

The amount of Cu that was effectively alloyed with Au was quantified by XRD analyses.

Both metals exhibit a fcc structure, and the variation in lattice parameter is related to the

Au:Cu ratio in the alloy (Eq. 2.2) 33 Figure 4 shows the corresponding diffractograms in which

a clear shift of the peak positions can be seen as the synthesis proceeds. The dashed lines in

Figure 2.4 correspond to the (111) peak position of metallic Au (a = 4.08 Å, red line) and

metallic Cu (a = 3.61 Å, blue line). It can be seen that during the first reaction step (120°C),

common to all samples, XRD confirms the formation of monometallic Au NPs, in agreement

with the UV-VIS results. By increasing the temperature, the peak positions shift to larger

Bragg angle values showing a decrease of the lattice parameter in agreement with the

incorporation of Cu into the Au lattice. The results also show that the higher the temperature,

the larger the shift, indicating higher Cu incorporation. Interestingly, in the case of

AuCu_280 sample the superlattice peaks related to the chemically ordered tetragonal

structure of AuCu can also be seen.107 Therefore, although the yield of NPs decreases with

temperature and annealing time, keeping the system at 280°C for 30 min lead to NPs with

Au:Cu molar ratio of 1:1, matching the initial metals ratio, and favoring the formation of the

chemically ordered bimetallic AuCu phase.

Table 2.2: %Cu (at.) at different temperatures obtained by XPS.

Temperature/Sample AuCu_225 AuCu_260 AuCu_280

100* 3.4

120_I* 5.7

120_F* 22.3

225_I 29.5

225_F 26.5 - -

260_I - 28.6 -

260_F - 40.3 -

280_I - - 36.2

280_F - - 40.4

*common steps for the three samples.

41

Figure 2.4: XRD patterns of Au1-xCux NPs at different stages of the synthesis: (a) AuCu_225,

(b) AuCu_260 and (c) AuCu_280. The red dashed line indicates the (111) peak position of

Au (a= 4.08 Å, JCPDS N° 00-004-0784) and the blue one indicates the (111) position of Cu

(a= 3.61 Å, JCPDS N°. 00-901-3023). The superlattice peaks of the chemically ordered AuCu

phase (JCPDS N°. 01-071-5026) are indicated by * in (c).

Table 2.3 presents the Cu at.% obtained by the analysis of the lattice parameter. It is

possible to confirm the increase in the copper amount as a function of the temperature,

similarly to what observed by ICP and XPS analyses. However, while ICP/XPS showed that

42

the total amount of Cu raises steadily, e.g. from about 10% when reaching 120°C (120_I) up

to 20% after 30 min at this temperature (120_F), XRD gives no indication that the Cu is

incorporated in the Au lattice at this temperature. In all stages, the Cu contents found by XRD

were smaller than those measured by ICP/XPS, with the exception of the final stages at 260 °C (260_F) and 280 °C (280_F).

Table 2.3: %Cu (at.) derived from the lattice parameter, obtained by XRD (Eq.2.1)

Temperature/Sample AuCu_225 AuCu_260 AuCu_280

100* 0

120_I* 0

120_F* 0

225_I 15.0

225_F 19.2 - -

260_I - 21.9 -

260_F - 38.1 -

280_I - - 30.1

280_F - - 50.0

*common steps for the three samples

Figure 2.5 compares the results found by ICP and XRD. It is possible to observe that,

as the syntheses proceed, the values obtained by the two techniques tend to converge toward

the same values. These results likely indicate that part of the Cu species present at the

beginning of the synthesis, which precipitate together with the Au NPs during the purification

process, are not detectable by XRD, not even as a separate phase (e.g. metallic Cu or

crystalline copper oxides).

43

Figure2.5: Comparison of the %Cu (at.) obtained by ICP and XRD as a function of the

synthesis temperature of (a) AuCu_225, (b) AuCu_260 and (c) AuCu_280.

In order to understand this difference, a detailed analysis was performed by HAADF-

STEM imaging and STEM-EDS elemental analysis on samples 280_I and 280_Fas shown in

Figure A.6 and Figure A2. Besides the high-contrast (i.e., high thickness and/or mean atomic

number) of the Au1-xCux NPs, a fainter contrast region can be observed in HAADF-STEM

images around the particles for the sample 280_I, as indicated by arrows in Figure 2.6. This

region is mainly composed by copper, as shown by the corresponding STEM-EDS analysis.

Although this type of analysis is local, this result, associated to the similar quantification of

Au and Cu provided by ICP and XPS, is strong evidence that the reduction of Cu leads

initially to the formation of a Cu-rich phase that cannot be identified by means of XRD.

44

Figure 2.6: HAADF-STEM image of 280_I sample, showing the presence of a lower contrast

phase (indicated by arrows) surrounding the Au1-xCux NPs, mainly composed by copper as

demonstrated by the comparison between STEM-EDS spectra collected on NPs and on the

surrounding phase (identical area and acquisition time).

2.4. Discussion

In the one-pot synthetic method, both metals precursors are added together to the initial

solution and, depending on the reaction conditions, one may expected the formation of an

alloy already in the early stages of the synthesis process. Previous works with the Au1-xCux

system found, however, that in fact Au nucleates first. 72,108 Destro et al.109 probed the

formation of the AuCu alloy (in air) by in situ X-ray absorption fine structure spectroscopy

45

(XAFS) and confirmed that Au reduction starts already at 70oC. In the present work, where

the synthesis was performed in vacuum/inert atmosphere, the results confirmed that Au

precursor is reduced first, and alloy formation is highly dependent on the temperature.

Considering the experimental conditions implemented in the present work, the early reduction

of Au can be rationalized considering that 1,2-hexadecanediol is not a strong reducing agent,

in particular at low temperature, and that Au3+ is reduced easier than Cu2+ due to the higher

reduction potential (+1.5 and +0.3 eV respectively).110

One interesting consequence of the formation of monometallic Au NPs at the early stages

of the synthesis is that theAu1-xCuxNPs, characterized by different compositions, show similar

sizes. Similar results were reported by Motl et al.72 they implemented a one-pot synthesis

approach where the amount of Cu in the NPs was controlled by the Au:Cu ratio of the

precursors. They analyzed eight samples with copper contents ranging from 0 to 50% (at.) and

seven of them showed an average size in the10-13 nm range. Recently, Sinha et al.,100

explored the one-pot strategy for the synthesis of Au1-xCuxNPs by varying the (Oley +

OlAc)/metal and Oley/OlAc ratios seeking to control the average particles size. They reported

that a 40:1 total ligand: metal ratio led to the smallest average size of the NPs (i.e. about 14

nm). This value is in line with our results, where the total ligand: metal ratio is 50:1.

According to Sinha et al.,100 it was not possible to decrease the particle size below this value

just by increasing the amount of ligands. Smaller Au1-xCuxNPs were only obtained by using a

seed mediated approach where pre-formed Au NPs are added to solution containing the

copper precursor.104 We have also found similar results, where the final temperature of the

synthesis impact both the Au:Cu ratio and the eventual chemical order of the alloy without

affecting significantly the average particles size. In fact, at 100oC when most of the Au ions

have already been reduced forming the NPs, the mean size is 14 nm (see Figure S1). It is

worth to point out that the increase of the Cu content in the NPs from 25 % (AuCu_225) to

50 % (AuCu_280) should lead to an increase of about 11 % in size. This small increase

would be hardly detected by TEM due to the size distribution of the samples. Different

parameters such as total ligand/metal ratio, heating ramp and temperature of step 1 were also

varied but did not impact significantly in the final particle size (results not reported).

More interestingly, how the Cu atoms are generated and incorporated to form the alloy is

still not clear and there are some possible mechanisms that can lead to the Au1-xCux alloy

formation. Chen et. al.96 proposed that in the seed mediated synthesis, a diffusion-based

mechanism takes place. They proposed that Cu2+ ions were reduced to Cu atoms and/or small

46

cluster in solution. Then, by colliding with the Au NPs Cu species diffuses into the core of the

NPs forming the alloy. No evidence of a core-shell structure was found. On the other hand,

Shore et al.111 observed a core-shell structure in AuAg system at low temperature (< 200oC)

and the alloy formation was achieved by increasing the temperature. They also showed that

similar AuAg alloy NPs could be directly obtained by heating Au NPs in the presence of Ag

precursor at higher temperatures (~250 o C).

In our work, the combination of XPS, ICP, TEM, XRD and STEM-EDS does not indicate

the formation of a core-shell structure or the direct formation of atomic species of copper that

would collide and diffuse into the Au NPs. We showed that the Au NPs are formed at low

temperature and that the amount of Cu in the final product progressively increases with the

reaction temperature. Moreover, the copper content inferred from XRD analysis was

systematically smaller than that measured by ICP and XPS at all reaction stages except at the

end of the annealing step. Our results strongly indicate that a Cu-rich phase, undetectable by

XRD, was initially formed and then progressively digested in the reaction medium at high

temperatures. This Cu phase, amorphous or formed by few nm-sized Cu or CuOx

nanocrystals, acts as a reservoir of Cu species that are released as the synthesis proceeds and

that are incorporated into the Au NPs through a solid state diffusion process.

The growth mechanism of our Au1-xCuxNPs resembles the mechanism called digestion

ripening described by Smetana et al.112 in which AuCu and AuAg alloy NPs were obtained by

mixing and heating preformed single metal colloidal NPs in the presence of alkanethiol under

reflux. This mechanism is highly dependent on the ligand and temperature, and core-shell

NPs have also been prepared by this method at lower temperatures.113–115 We speculate that

the large excess of ligands and the high temperature achieved in our synthesis can promote

the digestion of the Cu-rich phase leading to the formation of the Au1-xCuxalloy, as proposed

in Figure 2.7.

47

Figure 2.7: Proposed growth mechanism of Au1-xCux alloy NPs. Yellow = Au atom, Red=Cu

atom. Gray = solid Cu-rich phase.

It is interesting to note that we did not found indication of the formation of core-shell

Au@Cu NPs in any stage of the synthesis (Figure A1).Recently, Liu et al.103 studied Cu@Au

NPs and nicely proposed the formation of CuAu alloy through a unidirectional diffusion route

of Au atoms into to Cu core. The driving force for the Au diffusion was proposed to be the

structural disorder of the shell, enhanced by temperature. The unidirectional diffusion

mechanism was associated instead to the structural stability of an Au-diluted CuAu alloy

front. In our case, since Au NPs formed first, we could expect initially the formation of

Au@Cu NPs that would then be converted to Au1-xCux alloy. This would happen if enough

copper species were available in solution at low temperatures and metal diffusivities were

limiting the alloy formation. In fact, Destro et. al.109 showed by in situ XAFS that the

reduction of copper starts already at about 120 oC. As a consequence, our results suggest that

the digestion of a Cu-rich phase forming species available in solution is the limiting factor,

and the temperature controls the final composition of the Au1-xCux NPs.

The possibility of an accurate composition control by the colloidal method keeping the

same size distribution is an interesting and useful approach for several applications, especially

in heterogeneous catalysis,98 since it makes possible to analyze the effect of the alloy

composition using particles of similar size and shape.

48

2.5. Conclusions

We evaluated the effect of the temperature on the synthesis of Au1-xCuxalloy colloidal NPs

via one-pot approach. By changing the final reaction temperature it was possible to obtain

uniform spherical particles, with size of about 14 nm, with different composition: Au0.75Cu0.25

at 2250C, Au0.60Cu0.40 at 2600C and Au0.50Cu0.50 at 280°C. By combining different techniques,

it was possible to propose a growth mechanism, where Au NPs are formed at low

temperatures together with a Cu-rich phase. This intermediary Cu-rich phase acts as a

reservoir of Cu atoms that are incorporated into the Au NPs at a given temperature through a

digestive ripening like-process.

The results presented here clearly point out the important role played by the temperature in

the formation of bimetallic Au1-xCux NPs. Despite the efforts, we still lack deeply studies on

the phase diagrams at nanoscale, as well as the main mechanisms that determine atomic

diffusion and phase segregation at nanoscale. This is of crucial importance in several areas,

such as catalysis, where the interaction with supports and reaction medium can further affect

the stability of nanoalloys. We hope that our results will contribute to motivate further

theoretical and experimental studies in this direction.

49

Chapter 3

Formation of bimetallic copper-gold alloy nanoparticles probed by in situ XAFS

This chapter is an adaptation of the paper “Formation of

bimetallic copper-gold alloy nanoparticles probed by in situ

XAFS” by Priscila Destro, Daniel Augusto Cantane, Guilherme

dos Santos Honório, Luelc Souza da Costa, Débora M. Meira,

José Maria C. Bueno and Daniela Zanchet that is in the final

revision to be submitted.

My contribution to this work was the synthesis of AuCu

nanoparticles, the ex situ characterizations of AuCu

nanoparticles, participation in the XAFS data analysis and

writing process.

50

3.1 Introduction

Bimetallic gold-copper alloy nanoparticles (Au1-xCux NPs) have attracted much

attention not only from a scientific perspective but also from an applied one, especially in the

field of catalysis.34 Extensive efforts have been devoted to building synthesis strategies for

obtaining Au1-xCux NPs with well-controlled chemical composition80,116,117 crystalline

structure-ordered/disordered87,118, size and shape (morphology).97,119,120Among many

synthesis methods, the chemical reduction of metal precursors via one-pot approach in

organic-phase has revealed to be one of the most powerful ways to access these desired Au1-

xCux NPs at nanoscale. 87,104,121 In particular, through controlled synthesis parameters such as

precursor/co-surfactants concentration and reaction temperature/time Au1-xCux NPs from

nanospheres to nanowires as well as from disordered to atomically ordered alloys with

different compositions have been made.87,119 Although a significant progress has been reached

in synthesizing Au1-xCux NPs in organic-phase via one-pot approach102 the mechanism

(nucleation and growth) which govern its formation has not yet been satisfactorily understood

and is currently the greatest hurdle to design bimetallic Au1-xCux materials with well-

optimized proprieties.34,122 To overcome this drawback, ex situ characterizations by sampling

procedures have been used to unraveling the mechanism involved during the evolution of

Au1-xCux NPs formation, especially taking out aliquots from reaction solution at different

times.80,117 Continuing from these understandings, we have recently revealed the temperature

effect on the final Au:Cu composition, demonstrating the Au1-xCuxalloy NPs formation occurs

through a digestive ripening-like process.4 However, from these ex situ procedures all shed

light on the synthesis mechanism cannot be deeply extracted, once that the stability of the

intermediates formed is critically influenced by interactions of potential in solution, i.e., the

solvation effects. As another strategy, in situ synchrotron radiation-based X-ray

characterizations have shown to be capable of identifying and monitoring the chemical

evolution of the NPs formation in solution. 123 Among them, in situ X-ray absorption fine

structure (XAFS) spectroscopy have received considerable attention124, since it offers

versatility, such as monitoring the changes in oxidation states, coordination, bond formation,

already at the initial stages of nucleation and growth of the NPs, as well as information about

local environment with chemical sensibility, even in the absence of a crystalline phase.125,126

By this means, many in situ studies have dedicated in monitoring the formation of bimetallic

NPs, however in aqueous-phase 127–129 with few limited reports in understanding the

formation of bimetallic NPs in organic-phase.130

51

Herein, we report new insights into the mechanism of one-pot bimetallic Au0.75Cu0.25

alloy NPs formation in organic-phase by time-resolved in situ XAFS spectroscopy. For that, a

home-made system was built enabling the direct monitoring of all events involved during the

bimetallic NPs formation in organic-phase. Therefore, the reaction sequence of Au0.75Cu0.25

alloy NPs formation with time resolution was mapped, providing strong evidences that

digestive ripening-like process is driven by a solid state diffusion mechanism at the metallic

gold/copper-oxide-rich interface.

3.2 Experimental session

3.2.1 Chemicals

HAuCl4.3H2O (99%), and Cu(acac)2 (99%) were used as metal precursors. Oleylamine

(Oley, 70%) and oleic acid (OlAc, 90%) were used as protective agents. 1,2-hexadecanodiol

(HDD, 90%) was used as reducing agent and 1-octadecene (1-ODE, 90%) as solvent. All

reactants were purchased from Sigma-Aldrich and used without further purification.

3.2.2 Synthesis of bimetallic copper-gold nanoparticles

Bimetallic NPs were synthesized using a protocol previously reported.80 In a typical

procedure used for in situ XAFS measurements, HAuCl4.3H2O (45 mg) and Cu(acac)2 (68

mg)were added to a home-made reactor (see details below) containing 1-ODE (5 mL),Oley

(600 L), Olac (800 L) and HDD (100 mg) (Cu:Au atomic ratio ~ 2.3:1). The mixture was

magnetically stirred at room temperature to generate a green solution and then heated to 120

ºC. After holding at 120 ºC for 30 min in air, the solution was heated to 225 ºC and

maintained at this temperature for 30 min. After that, the solution was cooled to room

temperature naturally. Similar procedure was repeated at laboratory, to perform ex situ

characterization by a larger set of techniques and test the reproducibility. The purification was

carried out by adding acetone: hexane (3:1 volume ratio, 10 mL), followed by centrifugation

at 12000 rpm for 10 min. The precipitate was dispersed in hexane (10 mL) containing Oley

(100 L) and Olac (100 L) and the NPs were further purified by centrifugation at 3000 rpm

for 10 min, to remove large agglomerates. After discarding the precipitated, the NPs were

precipitated out by adding ethanol (10 mL), followed by centrifugation at 8500 rpm for 8 min.

The final NPs were dispersed in hexane (5 mL) for characterization.

52

3.2.3 Characterization

Inductively coupled plasma-atomic emission spectroscopy (ICP-AES) measurements

were carried out on 3000DV-Perkin Elmer atomic-emission spectrometer. X-ray diffraction

(XRD) patterns were collected using a Shimadzu X-ray diffractometer XDR7000 with Cu Kα

radiation (λ = 1.5418 Å), 40 kV and 30mA. Transmission electron microscopy (TEM) images

were obtained using a JEOL JEM 2100 equipment (200 kV and point resolution of 2.3 Å)

available at the Brazilian Nanotechnology National Laboratory (LNNano), Campinas, Brazil.

TEM samples were prepared by depositing a single drop of diluted NPs dispersion (~5 L) on

ultra-thin amorphous carbon-coated copper grids. EDS-STEM (energy dispersive

spectroscopy- scanning TEM) was performed with a 2100 FEG-TEM, operating with a probe

of 1 nm at LNNano. Ultraviolet–visible spectra were acquired with an Agilent 8453

spectrophotometer. In situ time-resolved X-ray absorption fine structure (XAFS) spectra in

the near edge region (XANES) were acquired at theD06A-DXAS beamline of the Brazilian

Synchrotron Light Laboratory (LNLS), Campinas, Brazil, operating in dispersive mode. A

curved Si(111) monochromator was used to focus the beam at the reactor window center and

a CCD camera was used as detector (time per spectrum ~ 4-300 s depending on the

concentration). Data at both absorption edges, Au-L3 (11919 eV) and Cu-K (8979 eV), were

collected by running twice the experiment. The final spectra were obtained by subtracting

each collected spectrum from the background signal originated by a solution containing all

the reactants except the metal precursor of interest. A home-made reactor was built based on a

round bottom flask with two adjustable hollow Teflon® pistons capped with Kapton®

windows to allow the X-rays passing through the reaction medium. The optical pathway could

be adjusted from about 0.5 to 10 mm. The dispersion was stirred with a stir plate and heated

by a metallic block, controlled via a thermocouple system. Details about the setup for

monitoring the NPs synthesis by in situ XAFS are presented in the Supporting Information

(Figure B1).

3.3 Results and discussion

Far from the best practice in controlled-synthesis, the mapping in time-resolved

formation of Au0.75Cu0.5 alloy NPs was carried out under air, which could impact on the final

Cu ratio incorporated into alloy material. Therefore, prior to in situ XAFS measurements, the

NPs were characterized by ICP, TEM, UV-vis absorption, and XRD to confirm the

53

reproducibility and robustness of the synthesis approach under air condition, besides

providing a basis for comparison of the obtained nanostructure.

400 500 600 700 800N

orm

aliz

ed a

bsor

ptio

n (A

rb. U

nits

)

Wavelength (nm)

20 30 40 50 60 70 80Nor

mal

ized

inte

nsity

(Arb

. Uni

ts)

2 (°)

Figure 3.1. (a) Representative TEM image;(b) UV-Vis spectrum. The red and blue dashed

lines indicates the plasmon peak position of Au and Cu NPs, respectively(c) XRD pattern of

the synthesized bimetallic Au1-xCux NPs The red dashed line indicates the(111) peak position

of Au (a= 4.08 Å, JCPDS N° 00-004-0784) and the blue one indicates the (111) position of

Cu (a= 3.61 Å, JCPDS N°. 00-901-3023).

From Figure 3.1a, it is observed that the as-prepared NPs are substantially spherical in

shape, with an average diameter size of 14.3± 1.3 nm (see statistical information in Table B1).

UV-vis absorption spectrum in the Figure 3.1b confirmed the formation of Au1-xCux NPs

alloy; it shows a centered band at 545 nm corresponding to a maximum plasmon resonance

frequency between the plasmon bands of Au NPs in this size range (red dashed line), at 520

nm, and Cu NPs (blue dashed line), at 560 nm.80,116 According to ICP data, the purified NPs

are formed by ~ 1:3 Cu:Au atomic ratio (namely Cu:Au 27.4:72.6 ±0.3%). XRD pattern in the

Figure 3.1c corresponds to face-centered cubic (fcc) nanostructure with peak positions in

between of gold and copper standards. Based on the lattice parameter obtained from XRD

pattern, it is possible to quantify the amount of Cu incorporated into the alloy using the

Vegard´s Law33 and the results confirm the Cu amount of about 26 %, in accordance with the

formation of Au0.75Cu0.25 NPs. The elemental line-scan EDS-STEM analysis (Figure C2) is in

agreement with a homogenous distribution of Cu and Au indicated by UV-Vis and XRD

measurements. Therefore, the results confirm the formation of Au0.75Cu0.25 NPs. It is

important noting that a large fraction of Cu precursor was placed to compensate the favorable

copper oxide formation under air condition (initial Cu:Au 3:1 atomic ratio for generating

incorporated Cu:Au 1:3). In fact, we have recently demonstrated that under rigorous

54

atmosphere control (oxygen-free), the final Cu:Au composition is the same that one obtained,

however a smaller initial amount of metal precursorswas required (1:1 Cu:Au ratio).117. A

fully Cu incorporation into the bimetallic NPs is only reached at the higher temperature

synthesis, as pointed in previous works80,104,108,117 Our results showed that this method is also

reproductive under air condition and robust to investigate the mechanism of nucleation and

growth for a bimetallic system by in situ XAFS spectroscopy.

3.3.1 Nanoparticles formation

After optimizing the operational conditions for in situ XAFS measurements, the metal

precursors were added to 1-ODE in the presence of the colloidal stabilizers (Oley and OlAc)

and reducing agent (HDD). Figures 3.2a,b shows time-resolved in situ XAFS spectra at both

the Au L3 (E0 = 11919 eV) and Cu K (E0 = 8979 eV) absorption edges, respectively, during

the growing of the bimetallic Au0.75Cu0.25 NPs. Changes in the spectra can beclearlyseen as

the synthesis proceeds.

(a)

(b)

Figure 3.2. Time-resolved in situ XAFS spectra at both (a) Au L3- and (b) Cu K-edges

showing the chemical evolution of Au and Cu species during the synthesis of the bimetallic

Au1-xCux NPs.

In the Figure 3.2a, the peak observed just above the threshold of the Au L3-edge,

known as white-line, is related to dipole transitions 2p3/25d and its intensity is directly

related to the unoccupied d-projected density of states. 131 Metallic gold as well as cooper has

nominally a full d band and the changes of the spectra as a function of temperature in the

Figure 3.2a is in agreement with the evolution of Au(III) oxidation state to Au(0) during the

55

formation of the bimetallic NPs. In the case of Cu K-edge (Figure 3.2b), which mainly

involves 1s4p dipole transitions, the change in intensity of the white-line is also visible and

correspond to the evolution of Cu(II) oxidation state to a more reduced state. A pre-edge

feature emerged upon heating, which is related to dipole forbidden transition from 1s to d

states. 131 To better understand these changes, selected in situ XANES spectra at different

temperatures (30, 50, 100, 120, 160, 200 and 225 ºC) are shown in Figure 3.3a a set of

XANES spectra of standards materials, including Au and Cu foils, HAuCl4 and Cu(acac)2

precursors and Cu2O powder, are displayed in Figure C3 for reference.

11900 11925 11950 11975 12000

11920 11925 11930 11935

i

iiiii

Nor

mal

ized

abs

orpt

ion

Energy (eV)

30 ºC 100 ºC 120 ºC 160 ºC 200 ºC 225 ºC

(a)i

8960 8980 9000 9020

8980 8985 8990

i

iii

ii

Nor

mal

ized

abs

orpt

ion

Energy (eV)

30 ºC 100 ºC 120 ºC 160 ºC 200 ºC 225 ºC

(b)

i

Figure 3.3 In situ XAFS spectra for (a) Au L3-edge and (b) Cu K-edge, taken at different

reaction temperatures: 30 ºC (black-line), 100 ºC (dot), 120 ºC (dash dot–spectrum taken at

the when reaching 120oC; open-circle–spectrum taken after 30 min at 120oC, respectively),

56

160 ºC (solid-circle line), 200 ºC (star), and 225 ºC (red-line). Insets show a magnified view

of the edge regions. See text for details.

From both set of spectra, a particular and interesting trend can be observed. The black-

line spectra obtained at 30 ºC in the Figure.3.3a and b match well with those ones observed

for HAuCl4 and Cu(acac)2 precursors, having Au(III) and Cu(II) oxidation states, respectively

(see precursors spectra for comparison in the Figure. B3a,b). Focusing on the spectra profile

at the Au L3-edge, in the Figure 3.3a, it is noted that there is a rapid depletion of signal

intensity related to the Au(III)species (peak i, at ~ 11922 eV) when the temperature of

reaction is still mild (lower than 100 ºC temperature). Moreover, two shoulders concomitantly

emerge at around 11944 eV (peak ii) and 11966 eV (peak iii), which are attributed to a

metallic gold phase (see pure gold foil spectrum in the Figure B3a), suggesting the formation

of Au−Au bonding as in the metal phase already at the earliest stage of the reaction. This

reduction process is also supported by the color changes of the reaction medium from yellow

to deep-red. It is well-known that the formation of deep-red solution takes place in the

presence of gold nanoparticles in colloidal dispersions. Therefore, in situ XAFS spectroscopy

results at the Au L3 absorption edge revealed a nearly complete depletion of the HAuCl4

precursor as the reaction temperature is increased above 100 ºC.

In contrast, there is no evidence of Cu(II) converting (from the Cu(acac)2 precursor) to

Cu(I) or to Cu(0) species up to 100 ºC temperature (dot line spectrum in the Figure 3b). Upon

reaching the temperature to 120 ºC (dash-dot line spectrum in the Figure 3.3b), however, a

small pre-peak is detected at about 8981 eV (peak i) and a shift in the position of the main

peak from (iii) to (ii) position occurred (dash-dot line spectrum in the Figure 3.3b). This

suggests that the copper reduction start to take place in this temperature. More interestingly

however, after holding the temperature for 30 min at 120 ºC (open-circle line spectrum in the

Figure 3.3b), a profile more similar to Cu2O spectrum is visible (see Figure B3b for

comparison).When the temperature is further increased to 225 ºC, the modifications are small

but going into the direction of metallic copper. Previously XANES studies of bulk AuCu

alloy at both Au L3 and Cu K absorption edges have shown qualitatively changes in the

XANES spectra features when compared with pure Au and Cu foils.131 For example, Kuhn

and Sham131 found a small increase in absorption just above Au L3-edge threshold (~ 11927

eV) for bulk Au0.75Cu0.25alloy when compared to Au foil. In the case of Cu K-edge, the study

has reported a decrease of the absorption pre-edge peak(9879 eV) and an increase of the first

57

main peak for the alloy material (~8994 eV) when compared to Cu foil.131 The red-line

spectrum in the Figure 3.3a matched well with the one reported for bulk alloy (Figure C3b).

In contrast, final Cu K-edge spectrum (Figure 3.3b) is very different from both metallic

copper (Figure C4b) and bulk Au0.75Cu0.25 alloy. These results together reveal that, although

the Au1-xCux NPs formation occurred, a large fraction of unreduced copper remains in

solution overcoming the typical features of the absorption spectrum of the bulk alloy. This is

agreement with the ICP results, that showed that although the initial Cu:Au atomic ratio was

3:1 the final Cu:Au atomic ratio in the NPs was only 1:3Cu:Au.

3.3.2 Mechanistic aspects

An illustrative analysis can be drawn from in situ XANES data by looking at the

profile of absorption intensity at characteristic energies. Figure 3.4a shows the evolution of

the Au(III) white-line intensity as function of temperature at 11922 eV (peak i in Figure 3.3a)

and Figure 3.4b shows the corresponding graphic for Cu(II)looking at 8998 eV (peak iii in

Figure 3.3b). Additionally, the evolution of the intensity of absorption pre-edge peak at ~

8981 eV is also shown in the Figure 3.4b, for comparison.

50

100

150

200

Temperature (°C) White line intensity

(at 11922 eV)

Synthesis evolution

Tem

pera

ture

(ºC

)

0,90

0,95

1,00

1,05

1,10

1,15

(t5) (t1) (t4) (t

3) (t

2)

Whi

te li

ne In

tens

ity (a

t 119

22 e

V)(a)

58

50

100

150

200

Temperature (°C) 8981 eV 8998 eV

Synthesis evolution

Tem

pera

ture

(°C

)

Pre-

peak

and

whi

te li

ne in

tens

ities

(at 8

981

and

8998

eV

)

0,0

0,3

0,6

0,9

1,2

(t4)(t

3)(t

1) (t2)

(b)

(t5)

Figure 3.4.: Evolution of the X ray absorption intensity at chosen energies for both (a) Au L3-

(a) and (b) Cu K- edges showing the species evolution as function of temperature. See text for

details.

We labeled from t1 to t5five main regions in the Figure 3.4. In the Figure 3.4a, the

large variation in the intensity at the t1 region (< ~60 oC), is due to the poor solubility of the

HAuCl4in the reaction medium at low temperature, introducing artifacts in the spectra. More

interesting, proceeding by the t2 region (~60-90oC) the X-ray absorption intensity decreases

significantly, in agreement with reduction of Au(III) and formation of Au NPs. In t3region,

from 90 oC to the end of the step at 120oC, there is not significant variation in the X-ray

absorption intensity. Only during t4 region, when the temperature is again increased the

intensity decreases. Above ~195 oC (t5 region), the intensity slowly increases again. These

results can be interpreted as followed: up to 90oC mostly of the gold precursor is already

reduced forming Au NPs; temperature above 120 oC is necessary to finally reduced residual

metal ions in solution and/or promote the growth/dissolution of small nanoparticles; above

~195oC the smoothly increase of intensity indicates a more extensive incorporation of Cu into

the Au NPs. It is important noting that the spectra intensities obtained at region (t3) and (t4)

matched well with those ones reported for both bulk Au foil and Au0.75Cu0.5 alloy,

respectively.

The interpretation of the reduction pattern of Cu(II) species, Figure 3.4b,was not as

clear due to the large amount of oxidized species that remained in solution. An attempt to

interpret the variation was made by looking at two energies, 8981 eV (close to the pre-peak

region of Cu(I)/Cu(0)) and 8998 eV (Cu(II) white line). At low temperatures (~40 oC), in the

59

t1 region, the increasing solubility of Cu(acac)2 accounts for the slight increase of the X-ray

absorption intensity. In the t2 region (~ up to 90 oC) there is no significant variation, in

contrast to the Au case. Also, different from the Au, during t3 region variations occur; from 90 oC up to the beginning-middle of 120 oC step the main modification takes place, with increase

of the Cu(I)/Cu(0) pre-peak and concomitant decrease of the Cu(II) white-line showing the

beginning of copper reduction. The next modification will only take place in the t4 region, by

increasing the temperature, indicating that more copper species are reduced. Interesting, the

intensity profile of the pre-peak (8981 eV) shows an abrupt increase followed by a decrease.

Since the pre-peak of Cu(I) is much more pronounced than in the metal, these results reveal

that above ~160oC a more extensive formation of Cu(0) takes place. By this means, the

reduction process from Cu(II)state to Cu(0) was followed by two slow-stages, one determined

by the formation of Cu(I) species and other driven by the generation of CuAu alloy.

To gain further understand the time-evolution chemical nature of the alloy

nanostructure formed and correlated to the Au L3- and Cu K-edges data, UV-vis, TEM and

XRD measurements were performed by taking out aliquots (~ 20 L) at different reaction

temperatures (90, 120, 160, 200 and 225 ºC)during the reaction. For this purpose, the NPs

were analyzed without further purification; these results were matched with those carried out

by in situ XAFS measurements.

Figure 3.5 shows UV-vis absorption spectra as function of the synthesis evolution

collected at different temperatures. As it can be seen from Figure 3.5, the UV-vis absorption

spectrum at 90 ºC readily revealed a plasmon band at 520 nm (ii), corresponding to quasi

spherical gold nanoparticles.80 By increasing the temperature, the UV-vis absorption spectra

show a clear shift to higher value (~540 nm at 225 oC), indicating the incorporation of copper

species into the Au crystal lattice.80,116At160 ºC and above, a small but noticeable band at

around 488 nm, (i), can also be seen. Bulk Cu2O has an optical band gap at 570 nm that

significantly blue shifts for nanocrystals, depending on size132 and shape133. The presence of

this band might indicate the concomitant formation of Cu2O nanocrystallites or a thin-layer of

Cu2O. Therefore, as the temperature is raised to 225 ºC, the shift of the main plasmon band

and the broadening of the band at around 488 nm may indicate the decreasing of Cu2O species

and an increase of Cu species into the final NPs. The UV-vis absorption spectrum of the

purified nanoparticles NPs shows a fully depletion of Cu2O signal and matches well with that

one reported for AuCu alloy nanoparticles, with no visible presence of residual Cu2O signal.80

60

450 500 550 600 650 700

purified NPs

225 oC

200 oC

160 oC

120 oC

ii iii

Nor

mal

ized

abs

orpt

ion

Wavelenght (nm)

i

90 oC

Figure 3.5: UV-Vis spectra evolution during the synthesis of Au0.75Cu0.25 NPs. NPs spectrum

obtained after purification was inserted for comparison. The lines are guide to the eyes

indicating the center absorption bands for Cu2O (i), Au (ii), and AuCu (iii) NPs. An offset

was applied to all spectra for sake of clarity.

The UV-Vis results further confirm that the growth of bimetallic CuAu NPs is guided

by multi-stages; firstly the Au NPs formation, followed by the Cu2O formation, and then the

incorporation of Cu species into the Au crystal lattice. This trend is in agreement with those

ones earlier observed by time-resolved in situ XAFS measurements.

Figure 3.6a,b display diameter distribution histograms during the synthesis evolution.

A set of representatives TEM images is shown in Figure B5. Most of nanoparticles are

spherical in shape and well-formed nanoparticles, with mean diameter of 11.2 ± 1.2 nm can

be already seen at 70 oC (Table B1), and increase up to 13.8 ± 1.8 nm at 225°C. At

intermediates temperatures the size is around 14 nm, considering the error associated to these

measurements.

61

6 8 10 12 14 16 18 200

70

140

Diameter (nm)

70°C11.2±1.2 nm

120°C_final13.2±1.2 nm

160°C14.5±0.9 nm

225°C_final13.8±1.8 nm

6 8 10 12 14 16 18 200

306090

Num

ber o

f par

ticle

s

6 8 10 12 14 16 18 200

10

20

6 8 10 12 14 16 18 2005

101520

Figure 3.6: Diameter distribution histograms showing the evolution of synthesis obtained at

70, 120, 160, and 225 ºC.

Figure 3.7 shows the temporal evolution of the synthesized bimetallic AuCu NPs at

different temperature (90, 120, 160 and 225 ºC) by XRD measurements. The XRD pattern

confirms the formation of Au NPs at low-temperature (90ºC), since the peak positions

relatively matched with that one for bulk Au pattern. This result is much plausible with UV-

Vis data and further confirms the observation obtained by XAFS measurements at the Au L3-

edge. After aging for 30 min at 225 ºC a noticeable shift to higher 2 values in the peak

positions is observed. This implies that Cu species are incorporated into Au crystal lattice,

shortening Au-Au bond distance due to the formation of Cu-Au bonding.

62

30 40 50 60 70 80

225 °C_final

160 °C

120 °C_final

120 °C_initial

Nor

mal

ized

Inte

nsity

(Arb

. Uni

ts)

2

90 °C

Figure 3.7: XRD patterns during the formation of AuCu NPs, obtained at different

temperatures: 90 ºC (black line), 120 ºC (red line), after 30 minutes at 120°C (blue line), 160

ºC (green line) and 225 ºC (purple line).The red dashed line indicates the (111) plane position

of Au and the blue dashed line indicates the (111) plane position of Cu.

From the above results, we can now get a deeper insight about the mechanism behind

the formation of bimetallic AuCu NPs in organic-phase, as it is illustrated in Figure 3.8.

Figure 3.8: Representation of the synthesis evolution in the bimetallic AuCu NPs formation

in organic-phase.

Destro et al.117 studied the formation of Au1-xCux NPs with different composition in

oxygen-free conditions by combining different ex situ techniques, suggesting the alloy

formation takes place by a digestion ripening-like process of a Cu-rich phase that it is formed

63

together with initial Au NPs. Continuing from this investigation, herein we demonstrate the

Cu intermediate phase (CuOx species) incorporates into the Au NPs through a solid state

diffusion mechanism at a metallic gold/copper-oxide-rich interface. In situ XAFS clearly

shows that the Au precursor is mostly reduced when the temperature reaches 90 0C, releasing

Au NPs with sizes close to the final alloy NPs. Meanwhile, Cu(II) species start to reduce only

above 90oC forming Cu(I)/Cu(0) species as an intermediate step. These species are not

detected by XRD but UV-Vis spectra suggest the formation of Cu2O species at these early

stages. Only above 160 oC it is possible to observe by in situ XAFS a modification on Au and

Cu profile that can be related to the incorporation of Cu species into the Au nanoparticles pre-

formed at the first step.

In summa, these results indicates that the formation of Cu2O species (nuclei or

clusters) and, consequently, their diffusion from Au NPs surface towards inner layers appears

to be the driving force for the generation of a bimetallic AuCu nanostructure in organic-phase.

Probably, there are enough thermal energy at higher temperature for supplying both the

diffusion of solid-structures at nanoscale and the generation of bimetallic nanostructures one,

as the AuCu NPs. These results also are in agreement with recently study reported on the

formation of intermetallic AuCu NPs from copper ions diffusion route into gold-seed at high

temperature.104

Therefore, we demonstrate that employing a time-resolved mapping by in situ XAFS

analysis from fundamental chemical processes during the bimetallic nanostructures formation

can be an interesting tool, not only for controlled-chemical synthesis but also applied-science

of the nanostructured materials, bringing useful information for scientists to design novel

bimetallic nanoparticles with well-optimized proprieties.

3.4 Conclusions

In the present contribution, we have reported the time-resolved evolution of formation

of colloidal gold-cooper nanoparticles using in situ XAFS spectroscopy. Characterization data

reveals an intermediate stage during the formation of bimetallic AuCu nanoparticles in

organic-phase, involving the diffusion of the Cu species into the Au crystal lattice upon

heating to a specific temperature. The UV-vis, TEM, and XRD measurements further

confirmed that the formation of bimetallic AuCu NPs occurred by multi-stages, including the

initial formation of Au NPs, followed by the diffusion of copper species (Cu2O) from the Au

64

NPs surface to their inner lattice, which appears to be the driving force for the generation of a

bimetallic AuCu nanostructure in organic-phase. The presented results provide more

information for designing one-pot synthesis of bimetallic nanoparticles, thus manipulating

monometallic species diffusion processes, for example, by turning reaction time and

temperature plateau.

Since other bimetallic alloys of paramount importance can be synthesized by a similar

protocol in organic-phase, we believe this setup could be extended to build an understanding

about the processes involved in the initial events of growing and nucleation during the

formation of a wide range of bimetallic nanostructures.

65

Chapter 4

Incorporation of AuCu nanoparticles on silica: Studies about the stability of

nanoparticles and the application in CO oxidation reaction

The content of this chapter is an adaptation of the article

entitled “AuCu alloy nanoparticles supported on SiO2:

Impact of redox pretreatments in the catalyst performance in

CO oxidation” by Priscila Destro; Sergio Marras; Liberato

Manna; Massimo Colombo and Daniela Zanchet accepted to

Catalysis Today in August 2016.

Reference: http://dx.doi.org/10.1016/j.cattod.2016.08.003

My contribution to this work was the synthesis and of AuCu

nanoparticles and AuCu supported on SiO2, the XRD, TEM,

and ICP characterizations and writing the paper for

publication.

66

4.1 Introduction

Metallic nanoparticles (NPs) are interesting materials that can present strategic

properties of interest in a large variety of applications. The impressive improvements made in

the synthetic methods in the last 30 years have permitted an easy access to high quality

samples of metal NPs for a large number of groups, working in different fields. In this

context, a significant contribution was made by the colloidal methods, in which the use of

protective agents and easy handling facilitates the use of the NPs. 26

In catalysis, the accurate control on size, shape, structure and composition achieved by

colloidal NPs have been used to address the complex interplay of different factors, such as

support effect, metal NPs size, metal-oxide interface among others134–138. In several cases, this

high control and stability of the colloidal NPs are originated by the use of organic capping

agents that form a protective layer on the metallic surface which prevents the aggregation of

the NPs and directly participate in the nucleation and grow mechanism. Ligands with

functional groups (i.e. amines, phosphines, thiols, carboxylic acids), which bind to the metal

surface and nonpolar tails that provides steric stabilization have been largely used in colloidal

synthesis of metal NPs.4,138.

The use of colloidal NPs in catalysis has significantly contributed to understand

fundamental aspects in several reactions.9,19,90 In the case of bimetallic NPs, the colloidal

approach is particularly promising, since it enables a fine control of NPs composition and

homogeneity.48,50 In addition, by using preformed metallic NPs supported on metallic oxides

(i.e. silica, alumina, titania, ceria, etc.) it is possible to control and stabilize the metal phase

prior the deposition on the support, which allows a greater stability of the metallic phase, and

the control of activity and selectivity of the catalyst. 44,140

Nevertheless, one of the main challenges in the case of supported catalysts produced

from colloidal NPs is the removal of the organic layer that covers the metallic phase. Their

presence usually blocks the active sites and thus requires an additional step to activate the

catalyst, in which the ligands have to be removed without causing sintering or undesired

modification of the metal NPs.141–143In particular for the case of bimetallic NPs, such as AuCu

NPs, the conditions required to remove the ligands may induce dealloying, changing

significantly the NPs properties.

Several methods have been presented in the literature, such as oxidation by ozone and

UV radiation144–146, heat treatments in an oxidizing atmosphere and redox cycling at mild

67

temperature 141,147. These procedures may induce alterations in the metal NPs, such as

sintering, oxidation, migration on the support surface, etc., which has to be evaluated with

great care to not irreversibly modify the NPs. In general, mild temperatures (< 400oC) are

used 148,149 but Cargnello et al. recently showed that a fast treatment at high temperatures (500

–700oC) can also have a similar effect, preserving the NPs.149

Therefore, understanding the different aspects involved in the activation of colloidal

NPs for catalytic applications is of fundamental importance. 151,152 In addition to the presence

of capping ligands on the metallic surface, much less understood is the impact and mechanism

of removal of other contaminants that may be present in the synthesis medium and that are not

fully eliminated during the purification steps, such as counter ions and other residues153. In

conventional methods of catalyst preparation, such as wet and dry impregnation of metal

precursors, this is much better understood. An important case is the presence of residual

chlorine from the metallic precursors that can block active sites by different mechanisms and

thus interfere in the catalyst performance.154,155

Within this framework, the aim of this work is to evaluate the effect of redox

pretreatments in the activation and CO oxidation catalytic activity of SiO2 supported AuCu

colloidal NPs. The AuCu system has been indeed reported as one of the most promising

metallic system in CO oxidation in which the properties are much superior than the

monometallic Au and Cu phases, especially in CO oxidation. 156

4.2 Experimentalsession

4.2.1Chemicals and materials

HAuCl4.3H2O (99%) and Cu(acac)2 (99%) were used as metal precursors. Oleylamine

(Oley, 70%) and oleic acid (OlAc, 90%) were used as protective agents. 1,2-hexadecanediol

(90%) was used as reducing agent and 1-octadecene (1-ODE, 90%) as solvent. All reactants

were purchased from Sigma-Aldrich and used without further purification. Silica AEROSIL®

380(BET surface area of 350-410 m2/g) was purchased from Evonik Industries.

4.2.2 Synthesis of bimetallic copper-gold NPs

The synthesis of bimetallic AuCu NPs was adapted from the literature72 and described

in detail by Destro et al.117 Briefly, 45 mg HAuCl4.3H2O and 30 mg of Cu(acac)2 (molar ratio

Au:Cu 1:1) were added to a three-necked round-bottom flask containing 5 mL of 1-ODE, 800

L of OlAc, 600 L of Oley and 100 mg of 1,2-hexadecanediol. The mixture was

68

magnetically stirred at room temperature under vacuum for 30 min, generating a green

solution. The mixture was heated up under vacuum up to 120 ºC. After holding at 120 ºC for

30 min, the solution was heated at about 200oC, N2 was fed to the system and the system was

heated up to 225°C, kept at this temperature for 30 min. The final composition using these

synthesis parameters was Au0.75Cu0.25, determined by Inductively Coupled Plasma Optical

Emission Spectroscopy (ICP) using an iCAP 6000 Thermo Scientific spectrometer (Cu at% of

24,89±0,26%).After purification, a given amount of Au0.75Cu0.25 NPs were dispersed in 5 mL

of hexane/15 mL of toluene and left in contact with SiO2 support at room temperature and

under stirring for 12 h. After this period, the solvent became clear indicating that most of the

Au0.75Cu0.25 NPs impregnated the SiO2 support. The material was washed 3 times using

hexane and centrifuged at 4000 rpm for 10 min. After the successive washes the material was

left overnight at 80 °C to dry. Elemental analysis of the Au0.75Cu0.25/SiO2 catalyst gave a total

metal loading of 1.93±0.20 wt%.

4.2.3 CatalyticTests

Catalytic activity data for CO oxidation were collected using a micro reactor system

(internal diameter = 6 mm) coupled with a NDIR analyzer (ABB Uras 26)for the continuous

measurement of CO and CO2. The feed gas consisted of a mixture of 1% v/v CO and 6% v/v

O2 balanced with He with a flow rate of 80 Ncc/min corresponding to a gas hourly space

velocity (GHSV) of 3000000 Ncc/h (g of (Au + Cu)). Once loaded in the reactor, the catalyst

was initially activated at 350 °C for 10 h in 6% v/v O2 in He to remove the organic ligands

present on the NPs surface. Independent thermogravimetric analysis showed that weight loss

due to organic ligand removal was completed at 350°C. Each catalytic test cycle was

performed at constant flow rate and inlet gas composition. For each cycle the reactor was

heated from room temperature to 300°C with a heating rate of 2 °C/min, kept at 300 °C for 30

min, then cooled to 100 °C with the same rate and kept at this temperature for 30 min. Before

each cycle the sample was usually pretreated for 1 h at 350 °C either in an oxidizing (6% v/v

O2, balance He)or in a reducing (5% v/v H2, balance He) atmosphere. Where not specified, no

pretreatment was performed. In this case, a new test cycle was started immediately after the

conclusion of the previous one, without changing the gas composition. The heating rate used

in activation, reducing, and oxidizing pretreatments was set to 5 °C/min.

69

4.2.4 Characterization

Overview bright-field transmission electron microscopy (BF-TEM) analyses were

carried out using a JEOL JEM-1011 instrument (thermionic W source, 100 kV high tension).

HRSEM-EDX analyses were carried out using a JEOL JSM-7500FA instrument,

equipped with an Oxford X-max LN2-free SDD, with 80 mm2 of sensor active area and 129

eV of energy resolution at 5.9 keV (MnK). The Extended Pouchou and Pichoir (XPP) matrix

correction algorithm included in the Oxford AZtec software was used to analyze the data. In

case of Au0.75Cu0.25NPs colloidal solution, the sample was prepared by dropcasting the alloy

NCs solution on an ultra-smooth silicon wafer. In case of Au0.75Cu0.25/SiO2, few mg of

catalyst powder were dispersed in isopropanol by sonication, then the suspension was

dropcasted on an ultra-smooth silicon wafer.

X-ray diffraction (XRD) measurements were performed using a RigakuSmartLab X-

ray diffractometer equipped with a 9 kW Cu Kα (= 1.542 Å) rotating anode, operating at 40

kV and 150 mA. XRD data on colloidal NPs were collected using a zero diffraction silicon

substrate while a glass sample holder was used for the powdered sample. The diffraction

patterns were collected at room temperature over an angular range of 20−90°, with a step size

of 0.05°. XRD data analysis was carried out using PDXL 2.1 software from Rigaku. XRD

data were used to confirm the Au0.75Cu0.25 alloy composition using the equation proposed by

Okamoto et al.33

4.3 Results

Figure 4.1a shows a representative BF-TEM image of the Au0.75Cu0.25/SiO2 confirming

that the synthesized NPs are homogeneous and well dispersed on the support (Figure 4.1a).

The XRD patterns of the colloidal Au0.75Cu0.25NPs and Au0.75Cu0.25/SiO2 catalyst are shown in

Figure 4.1b.According to the Figure 4.1 it is possible to observe that there is no significant

modification of the alloy NPs composition in the catalyst. The analysis of peak positions

confirmed that the average composition of the alloy NPs is Au0.75Cu0.25. It is also possible to

note that the peaks width are maintained, confirming that there was no sintering after the

incorporation.

70

Figure 4.1: (a) BF-TEM image of as prepared Au0.75Cu0.25 /SiO2 catalyst and (b) XRD

patterns of Au0.75Cu0.25 colloidal NPs and Au0.75Cu0.25/SiO2.

The Au0.75Cu0.25/SiO2 sample was used as catalyst in CO oxidation. As described in

the Experimental section, the sample underwent a series of cycles in order to evaluate the

impact of the pretreatments in the performance of the catalyst. The test protocol followed in

the first set of experiments is shown in Figure 4.2a. Firstly, an activation step is performed in

oxidative atmosphere. This treatment was necessary to burn off from the catalyst surface the

organic ligands (i.e. Oley and/or OlAc) that acted as stabilizing agents during the colloidal

synthesis. After the activation, the catalytic activity in CO oxidation is evaluated (Run #1) and

shown as the black curve in Figure 4.2b.The catalyst is inactive below 220 °C; then, the

activity slowly increases with temperature and the conversion reaches about 16% at 300 °C.

During the cooling phase the activity is slightly higher but still significantly limited. The

sample is then subjected to an oxidative treatment and the reaction cycle repeated (Run #2,

Figure 4.2b). The CO conversion resulted is slightly higher compared to the first reaction

cycle, approaching 19 % at 300°C. Repeating again the same oxidative pretreatment resulted

in a further enhancement of the activity, with a maximum of about 23% at 300°C (Run #3, in

Figure 4.2b). At the end of the Run#3, two more reaction cycles (Run # 4 and #5) are

performed without any pretreatment. These two tests results in a different trend in which the

activity during the heating phase is higher compared to the one measured during the cooling

transient. Higher CO conversions are also achieved, namely 30 % at 300 °C in Run #4 and

34% at 275 °C in case Run #5. Despite this different trend during the heating and cooling

transients when no pretreatments were performed, the main information from the data

reported in Figure 4.2b is represented by the non-reproducibility of the catalyst performance

among the different reaction cycles, even when performing the same pretreatment.

71

Figure 4.2: (a) Test protocol for Runs #1 to #5. Grey segments represent the pre-pretreatment

in oxidizing atmosphere, blue segment the reaction cycle, (b) CO conversion over

Au0.75Cu0.25/SiO2 during the test protocol depicted in (a).

The same sample was then exposed to a reducing atmosphere at high temperature

(Figure 4.3). The following reaction cycle resulted in a dramatic change of catalyst activity

compared to the previous tests (Run # 6, Figure 4.3b). At 75°C the CO conversion started to

increase and finally reached 100% when approaching 300°C. The activity during the cooling

cycle resulted even higher if compared to the heating transient (e.g. at 200°C we measured

66% CO conversion during the cooling transient against 45% measured during the heating

phase). The reducing pretreatment was repeated, followed by a reaction cycle (Run # 7,

Figure 4.3b) resulting in the same CO conversion profile. The same result was obtained in

Run #8. When no pre-treatment was applied before the reaction cycle (Runs # 9 and 10,

Figure 4.3b), the activity substantially overlapped with the previous run and no hysteresis

between the heating and cooling transients were observed.

72

Figure 4.3: Test protocol for Runs #6 to #10. Red segments represent the pretreatment in

reducing atmosphere, blue segment the reaction cycle, (b) CO conversion over

Au0.75Cu0.25/SiO2 during the test protocol depicted in (a).

After these series of tests, the sample underwent again a sequence of reaction cycles,

each preceded by an oxidative treatment (Runs #11 to 13, Figure 4.4a). The CO conversion

measured during the reaction cycles is dramatically higher if compared to the data reported in

Figure 4.3, approaching 100% at 300°C. Noteworthy, the three tests resulted in the same

activity profile, indicating that the activity data are reproducible and the catalyst approached a

stable behavior. The same situation is observed if no pretreatment is performed: four reaction

cycles (Runs #14 to 17) results in the same CO conversion profile as a function of

temperature, as reported in Figure 4.4b.

73

Figure 4.4: CO conversion over Au0.75Cu0.25/SiO2 during (a) Runs #11 to #13 performed after

pretreatment in oxidizing atmosphere, (b) Runs #14 to #17 performed without any

pretreatment.

The catalytic activity data presented in Figures 4.2 to 4.4 clearly point out the pivotal

role of a reducing treatment in order to obtain a stable and considerable catalytic activity over

the tested Au0.75Cu0.25/SiO2 catalyst. It is evident that the oxidative pretreatment, necessary to

burn off the organic ligands from the NPs surface, is not sufficient to fully activate the

catalyst and bring to a stable performance. On the opposite, one reducing treatment is enough

to improve the activity and provide stable performances. Impressively, the CO conversion

profiles measured in all the reaction cycles performed after the first exposure to a reducing

treatment do not exhibit any memory of the previous reaction cycles and pretreatment history:

the results were influenced exclusively by the pretreatment applied just before the test cycle.

These experimental evidences likely suggest that the first exposure to a reducing

environment caused an irreversible change on the catalyst surface that dramatically affected

the performances of the Au0.75Cu0.25 /SiO2catalyst.

74

In order to shed light on this phenomenon, the Au0.75Cu0.25/SiO2 sample is

characterized by means of different techniques. We firstly measure the size of the Au0.75Cu0.25

NPs on the as prepared catalyst (i.e. FRESH) and compare it with the size of the NPs after the

initial oxidative activation (i.e. OXI) and after a reducing treatment (i.e. RED). Graphical

evaluation of NPs size from TEM images shows no statistically relevant difference in the

three samples (Figure 4.5).

Figure 4.5: Size distribution histograms of Au0.75Cu0.25/SiO2 of (a) as prepared catalyst, (b)

after activation in oxidative atmosphere and (c) after first reduction.

We further characterize the same samples by means of ICP and SEM-EDX, to verify

their chemical composition. As control experiments we characterize with the same technique

Au0.75Cu0.25 colloidal NPs and the bare SiO2 support. Results of quantitative SEM-EDX

analyses are summarized in Table 4.1.

75

Table 4.1: Chemical composition obtained by ICP and SEM-EDX.

wt. %

(Au+Cu)/SiO2

from SEM-

EDX*

wt. %

(Au+Cu)/SiO2

from ICP

atomic %

Cu/(Au+Cu)

from SEM-

EDX

atomic %

Cu/(Au+Cu)

from ICP

wt.%

Cl/SiO2

wt.%

Cl/(Au+Cu)

FRESH 1.596 1.93 ± 0.2 29.24 24.89 0.0313 1.961± 0.654

OXI 1.357 1.93 ± 0.2 26.40 24.89 0.0201 1.481± 0.741

RED 1.492 1.93 ± 0.2 25.95 24.89 Not

detected

0.000

SiO2–Aerosil®

380

- - - - Not

detected

-

* SiO2 content evaluated based on O signal. The Si signal was influenced by the use of a Si

wafer for sample deposition.

Data reported in Table 4.1 confirm the presence of Au and Cu in all samples, without

any significant variation of their content or of their ratio with the applied pretreatment. Au

and Cu contents obtain by SEM-EDX and their Au: Cu ratio are comparable to those

measured by ICP. Interesting, besides these two elements, the SEM-EDX analysis of the

FRESH sample highlights the clear presence of chlorine. It was not possible to detect chlorine

by ICP due to the use of HCl for sample digestion. This element can be associated to both the

Au0.75Cu0.25 NPs and the SiO2 support: the Au0.75Cu0.25NPs are synthesized using HAuCl4 as

precursor, while the fumed SiO2Aerosil is produced using SiCl4. Analysis of unsupported

Au0.75Cu0.25colloidal NPs and of the bare SiO2 support clearly point out that the chlorine

observed in the FRESH sample was related to the Au0.75Cu0.25NPs. After the activation

treatment (OXI sample), the chlorine was still clearly detected, at levels comparable to the

ones measure over the FRESH sample. On the other hand, after the reducing treatment (RED

sample) no traces of chlorine could be detected on the sample. Considering the well-known

poisoning effect of chlorine in CO oxidation over Au catalysts 155,157 we can speculate that the

oxidative treatment resulted indeed in the removal of organic surfactants, but was not

76

effective for the removal of residual chlorine from the surface of the Au0.75Cu0.25NPs. On the

opposite, the high temperature treatment in a reducing atmosphere effectively removed the

chlorine from the Au0.75Cu0.25NPs, thus enhancing significantly the catalytic performances

and leading to a stable activity.

4.4 Discussion

The catalytic performance of the Au0.75Cu0.25/SiO2 after been submitted to different

pretreatments and reactions show a very intriguing and interesting behavior. Firstly, the CO

conversion achieved after the activation and the following oxidative treatments is typical low,

and slowly improves cycle after cycle (Figure 4.2). This could indicate, for example that

ligand residues remain firmly attached to the Au0.75Cu0.25 NPs surface and are slowly released

in the several cycles. However, the more significant increase in the CO conversion obtained

during subsequent reactions, without oxidative treatments in between, does not agree with this

explanation: the reaction atmosphere is indeed less oxidative than the oxidizing pretreatment

and thus should be less effective in the removal of residual ligands. This is further

corroborated by the remarkable effect of the reductive pretreatment, which should not have a

huge impact in removing residual ligands at these mild temperatures. Even more impressive is

the stability achieved by the catalyst after being submitted to the reducing treatment.

The detail characterization of the catalysts showed that the size distribution of the NPs

are not affected in this whole processes. The chemical analysis, on the other hand, showed the

presence of small amounts of residual chlorine that is eliminated only by the reductive

pretreatment.

The presence of chlorine in the colloidal NPs and the FRESH catalyst suggests that the

purification steps performed in the synthesis of the NPs117 were not able to successfully

remove it from the surface of the NPs. In fact, the percentage of chlorine detected by SEM-

EDX is of the order of metal atoms percentage at the surface of the NPs, suggesting that the

surface of the NPs is significantly covered by chlorine species. This result was quite

unexpected since the NPs go through several precipitation- redispersion before being

deposited on the SiO2 support. However, it is important to remark that the synthesis and

purification is performed in organic medium, with a much smaller affinity by chlorinated

species than aqueous medium.

77

Several authors have reported the poisoning effect of chlorine in catalysts, but the

mechanism is not totally clear. There are different mechanisms described in the literature to

explain such a poisoning effect. In the case of Au catalysts, Ho et al.155 proposed that the

presence of chloride ions can block electrostatically the metal surface, preventing the

reactants to access the active sites, and favor agglomeration. In other systems, the presence of

an [MxClyOz] complex has been described especially for Pd-Pt systems using chlorinated

metal precursors158 Gracia et al. 154 described interesting results regarding the chlorine

poisoning on platinum supported catalysts and systematically observed the effect of the

pretreatment for oxidation reactions. They proposed that the chlorinated species, initially on

the surface, can form a complex phase described as [Mx(OH)yClz] or [MxOyClz] after the

oxidation. It means that, even if the organic ligands on the surface are removed, after the

oxidative treatment this chlorinated phase can block the active metal sites. After a reductive

treatment, the metallic sites that are available can adsorb dissociative H2, which reacts with

the [MxOyClz] complex forming HCl that can migrate to the support or desorbs as a gas,

removing the Cl from the surface. This may be one possible mechanism behind the

considerable increase of the catalytic activity that we detected after the reductive

pretreatment.

It is important, however, to consider that the presence of chlorinate species is not the

only mechanism that can explain the increase of the CO conversion until the catalyst achieved

a stable state. It is important to highlight that after the oxidative treatment, the AuCu NPs

undergo a dealloying process, in which the less noble metal, in this case Cu, segregates

forming CuOx species at the interface with the support.95 The formation of CuOx species may

facilitate the dispersion and removal of Cl species that were previously on the metal surface

of the AuCu NPs. Najafishirtari et al.90 described the influence of different pretreatments on

an AuCu/Al2O3 catalyst and observed that the activation under oxidizing atmosphere removes

the organic ligand and induces Cu segregation, resulting in the formation of CuOx species

dispersed on the support. After a reductive treatment these CuOx species can be reduced and

realloy. This means that by exposing the catalyst to different atmospheres it is possible to

generate a new chemical environment on the surface of the catalyst, which can directly impact

the catalytic performances. In this case, a significant role of the support is most likely

expected and should be also investigated.

Despite this complex process, one of the most interesting results found in this system

is the stability achieved after the first reductive treatment. Once stable, the performance

78

depends only on the last pretreatment before the test, and remains constant even after a large

number of cycles. These results reveal that the AuCu catalyst do not exhibit a memory of the

previous reaction cycles or pretreatments, showing a remarkable stability.

4.5 Conclusion

In this work we synthesize Au0.75Cu0.25 colloidal NPs and use them to prepare a SiO2

supported catalyst characterized by a homogenous size distribution of the metallic phase. The

pretreatments performed before the catalytic oxidation of CO have a remarkable influence on

the catalytic performance. After the initial oxidative activation, the CO conversion is very low

but increases considerably after a reductive pretreatment. After that, the catalyst becomes

highly stable. This behavior can be related to the removal of Cl residue, detected on the

catalysts, by the reductive pretreatment and/or the formation of a new chemical environment

generating stable species on the surface that can improve the CO conversion. The most

interesting point to be highlighted is that after alternate oxidative/reductive pretreatments the

CO conversion is totally stable, showing that the material reaches an equilibrium state and

high activity.

79

Chapter 5

Au1-xCux supported nanoparticles applied on CO oxidation:

Insights about the support effect using in situ techniques

This chapter is an adaptation of the paper “Au1-xCux

supported nanoparticles applied on CO oxidation:

Insights about the support effect using in situ techniques”

by Priscila Destro, Tathiana M. Kokumai, Alice

Scarpellini, Liberato Manna, Massimo Colombo, Daniela

Zanchet that is under preparation.

My contribution to this work was planning the

experiments, the synthesis of AuCu nanoparticles and the

catalysts, the ex situ characterization by ICP, XRD and

BF-TEM, the planning and execution of in situ XRD and

XAFS measurements, XRD data analysis and writing the

draft for publication.

80

5.1 Introduction

Metal alloys are an extremely promising system for the application in catalysis.52,58,159 The

possibility of mixing two metals allows the modulation of their electronic properties and, by

synergism between the two metals, the properties of the resulting material are not merely an

average of the two metals involved, but a new material with distinct and innovative properties.30,32

In heterogeneous catalysts, the metal phase is supported on a matrix with high surface area 47,160

Thus, the use of supported nanoalloys becomes an interesting strategy because it permits control

of size, shape and composition of the metal phase. It is described in the literature that catalytic

supports are not just a phase where the metallic particles are deposited, but a phase that can

actively participate in the catalytic cycle.161 In many reactions the support acts by activating one of

the reagents involved, such as in oxidation reactions. In oxidation reactions, such as oxidation of

alcohols162, methane163 and carbon monoxide164,165 the support plays a fundamental role, where the

oxygen is activated on the metal/support interface. An interesting strategy of the design of new

and promissory catalysts is the application of metal alloys and, by different pretreatments, to

control the alloying/ dealloying process166,167, allowing the control of metal-oxide contact area.

Thus, the objective of this work is the deposition of AuCu nanoparticles previously synthesized by

the method colloidal on silica and alumina and, by redox pretreatments, evaluate the support role

by in situ techniques.

5.2 Experimental Session

5.2.1 Chemicals.

HAuCl4.3H2O (99%), and Cu(acac)2 (99%), Oleylamine (Oley, 70%) Oleic acid (OlAc,

90%), 1,2-hexadecanediol (90%), 1-octadecene (1-ODE, 90%). All reactants were purchased from

Sigma-Aldrich and used without further purification. Silica AEROSIL® 380 (BET surface area of

350-410 m2/g) was purchased from Evonik Industries and γ-Al2O3 from Strem Chemicals (Surface

area 190 m2/g)

5.2.2 Synthesis of bimetallic copper-gold NPs.

The synthesis of bimetallic AuCu NPs was performed as described in details by Destro et

al.117 Briefly, 45 mg HAuCl4.3H2O and 30 mg of Cu(acac)2 (molar ratio Au:Cu 1:1) were added to

a three-necked round-bottom flask containing 5 mL of 1-ODE, 800 mL of OlAc, 600 mL of Oley

and 100 mg of 1,2-hexadecanediol. The mixture was heated up under vaccum to a final

81

temperature between 225 and 280°C to obtain different Au1-xCux compositions. The purification

was performed by adding a toluene/ isopropanol (1:5 volume ratio) mixture, followed by

centrifugation/ dispersion in hexane. The samples were labeled based on the final composition,

AuxCuy where x + y = 1.

5.2.3 Catalysis preparation

Purified Au1-xCux NPs from the same batch were dispersed in a mixture of hexane toluene

(1:3 vol.), then mixed with the support (SiO2 or Al2O3) for 12 h under stirring to obtain 2 %wt (Cu

+ Au). After deposition, the material was washed 3 times by centrifugation at 4000 rpm with

hexane. Finally the sample was dried at 80 °C overnight in an oven.

5.2.4 Characterization

X-ray diffraction (XRD) ex situ measurements were performed using a Rigaku SmartLab

X-ray diffractometer equipped with a 9 kW Cu Kα (= 1.542 Å) rotating anode, operating at 40

kV and 150 mA. XRD data on colloidal NPs were collected using a zero diffraction silicon

substrate while a glass sample holder was used for the powdered sample. The diffraction patterns

were collected at room temperature over an angular range of 20−90°, with a step size of 0.05°.

XRD data analysis was carried out using PDXL 2.1 software from Rigaku. For analyze of the

metal phase of the catalyst were performed XRD measurements only on the supports Al2O3 and

SiO2and then subtracted from the from the catalyst XRD pattern originally obtained from the

catalyst.

The elemental analysis was done by Inductively Coupled Plasma Optical Emission

Spectroscopy (ICP) using an iCAP 6000 Thermo Scientific spectrometer. Colloidal samples and

the catalysts were dissolved in HCl/HNO3 3/1 (v/v) overnight, diluted with deionized water (14

mS), and filtered using a PTFE filter before each measurement. The analyses were performed in

triplicate.

Overview bright-field transmission electron microscopy (BF-TEM) analyseswere carried

out using a JEOL JEM-1011 instrument (thermionic W source, 100 kVhigh tension).

5.2.5 In situ measurements

In situ measurements were performed at the beamlines XAFS1 and XPD of the Brazilian

Synchrotron Light Laboratory (LNLS), Campinas, Brazil. The X-ray absorption fine structure

spectroscopy (XAFS) measurements were acquired atXAFS1 beamline and a Si(200)

82

monochromator. Spectra in the near edge region (XANES) and in the extended region (EXAFS)

were acquired at the XAFS1 beamline. The samples were packed in pellets and mounted in a

home-made quartz reactor coupled to a gas flow system. XANES spectra were collected at Cu K-

edge (8979 eV) and Au L3-edge (11919 eV) in transmission mode with ion chambers detectors

during the heating and EXAFS measurements collected at 298K.Data at both absorption edges,

Au-L3 (11919 eV) and Cu-K (8979 eV), were collected by running twice the experiment. The data

collected was analyzed using Athena code within Ifeffit package.

The in situ X-ray diffraction measurements were performed at the XPD beamline using a

Si (100) monochromator in the range 2θ=20-50° using the furnace “Arara” that allows the heating

up to 350°C using a rate of 5°C/ min and the detector Mythen. The calibration was performed by

the measurement of the pattern of α-Al2O3 and the wavelength defined as λ=1.5478 Å. Were also

performed measurements of the bare supports Al2O3 and SiO2 used for the catalyst incorporation.

The supports were heated up to 350 °C and acquired patterns during the heating. For the data

analysis, the signal from the bare supports at respective temperature was subtracted from the

catalyst signal, to analyze only the supported phase.

5.2.6 Catalytic Tests

The CO oxidation catalytic activity was measured using a micro reactor system coupled

with a NDIR analyzer that allowed the continuous detection of CO and CO2 concentrations at

reactor outlet. Typically, 80 mg of catalyst powder was loaded into a quartz reactor (internal

diameter = 4 mm). The feed gas was a mixture of 1 % v/v CO and 6 % v/v O2 balanced with He,

with a flowrate of 80 Ncc/min, corresponding to a gas hourly space velocity of 3E+6

Ncc/h/g(Au+Cu).

Once loaded in the reactor, the sample was treated for 10h at 350 °C in 6% v/v O2, balance

He in order to remove organic ligands, as confirmed by thermogravimetricanalysis.167The activity

measurements were performed by heating up the reactor from room temperature to 300 °C at a

heating rate of 2°C/min. The temperature was kept constant for 30 min, then the system was

cooled down at 2°C/min. Before each activity measurement the catalyst was exposed either to a

reducing (5% v/v H2 in He) or to an oxidizing (6% v/v O2 in He) atmosphere at a temperature of

350 °C for 1 h. The heating rate used in the initial activation, in the reducing and in the oxidizing

pretreatments was set to 5 °C/min. Scheme 1 depicts the sequence of pre-treatments and activity

measurements performed on each catalyst sample.

83

Scheme 5.1: Catalytic test protocol followed in this work.

5.3 Results and discussion

The synthesis final temperature was adjusted to obtain colloidal NPs with different

compositions.116The final compositions were determined by ICP and confirmed by XRD (Figure

5.1). The results for three sets of NPs (Au0.75Cu0.25, Au0.60Cu0.40and Au0.50Cu0.50) are shown in

Table C1 and Figure 5.1.The composition analysis by XRD was performed considering the lattice

parameter variation related do the Au (4.07 Å) and Cu (3.61 Å) based on the Vegard´s Law32.

Figure 1 shows the XRD patterns of the colloidal NPs.

20 30 40 50 60 70 80

*

Inte

nsity

(Arb

Uni

ts)

Au0.50Cu0.50

Au0.60Cu0.40

Au0.75Cu0.25

*

Figure 5.1: Diffraction patterns of Au0.75Cu0.25, Au0.60Cu0.40 and Au0.50Cu0.50 colloidal NPs. The

green dashed line indicates the peak at 2θ=38.2°related to the (111) position of Au (JCPDS no. 00-

004-0784) and the purple dashed line indicates the peak at 2θ=44.6° related to the (111) position

of Cu (JCPDS no. 00-901-3023). The red asterisks peaks in the Au0.50Cu0.50patternindicates the

(001) and (100) peaks related to the chemically ordered Au0.50Cu0.50 phase (JCPDS no. 01-071-

5026).

TEM images confirm the formation of Au1-xCuxcolloidal NPs with 14 nm size, for all

samples (Figure 5.2a and C1a,d).15 In addition, according to the TEM images it can be observed

that the Au1-xCuxNPs are preserved after incorporation in the supports and are well-dispersed on

84

the surface of the Al2O3 and SiO2 (Figures 5.2b,c and C1b,c,e,f).The total metal loading was

determined by ICP (Table S1), and was close to the 2 %wt nominal value for all supported

materials.

(a) (b) (c)

Figure 5.2: TEM images of (a) Au0.75Cu0.25colloidalNPs,(b) Au0.75Cu0.25/Al2O3and (c)

Au0.75Cu0.25/SiO2.

5.3.1 Catalytic results

Figure 5.3a-c shows the catalytic activity in CO oxidation for the alumina supported

catalysts after the oxidative and reductive pretreatments (Runs# 10 and 12 of scheme 5.1). Test

results obtained during the previous runs of the experimental protocol (see Figure C2) show that

the catalytic activity was reproducible and stable between Runs #6 and 12. Independently from the

NPs composition, the same qualitative trend was observed when changing the pre-treatment

environment. The samples pre-treated in an oxidizing atmosphere show a hysteresis between the

heating and the cooling phase of the test, with the activity being higher during the latter phase.

The reductive pre-treatment results in a reverse hysteresis, with a higher CO conversion measured

during the heating transient compared to the subsequent cooling. Exposing the sample to the

reducing atmosphere resulted in the best CO oxidation performances. Noteworthy, the activity

during the cooling phase of the experiments were substantially overlapped, independently from

the applied pre-treatment. The results of Figures 3a-c are surprisingly in line with what recently

reported by some of us. Najafishirtari et al.98 studied the effect of oxidizing or reducing pre-

treatments on the CO oxidation catalytic activity of ~ 6nm Au0.50Cu0.50 NPs supported on Al2O3.

Despite the different synthesis procedure and the significantly different size of the AuCu NPs (6

nm vs 14 nm), the qualitative effect of the pre-treatment was identical. On the other hand, the

different Cu content in the NPs affected the CO conversion level: as shown in Table 1, the

temperature at which 50% conversion is achieved resulted to be minimized (i.e. activity

85

maximized) in the case of the Au0.60Cu0.40 sample, independently from the pre-treatment and the

transient phase of the test considered.

A completely different picture was observed when the same Au1-xCux NPs were deposited

on SiO2 (Figure 5.3d-f) and subjected to the same testing protocol. In analogy with alumina, the

catalytic activity resulted to be strongly dependent from the pre-treatment atmosphere, while the

qualitative effect of the pre-treatment was independent from the Cu content in the alloy NPs.

When a stable activity was approached (see Figure C3 and related discussion), the oxidizing pre-

treatment resulted in almost the same CO conversion during both the heating and the cooling

phase of the test. A complex, “8-shaped” pattern of the CO conversion vs temperature was instead

observed after high temperature exposure to a reducing environment: at low temperature the

activity was similar to the one observed after oxidation. Then, the CO conversion increased less

markedly with temperature compared to the pre-oxidized system, resulting in lower CO

conversion values before eventually approaching 100% conversion between 200-250°C. Finally,

the activity during the cooling transient resulted to be almost overlapped with the one measured

after oxidation, in analogy with what observed for the same NPs supported on alumina. Also in

this case, the copper fraction in the starting NPs affected the CO conversion level. Table 5.2

shows how, similarly to what was observed over Al2O3, the best catalytic performances were

achieved over the Au0.60Cu0.40 sample.

86

(a) (b) (c)

(d)

(e)

(f)

Figure 5.3: Catalytic activity in CO oxidation after oxidative and reductive pretreatments for (a)

Au0.75Cu0.25/Al2O3, (b) Au0.60Cu0.40/Al2O3, (c) Au0.50Cu0.50/Al2O3,(d) Au0.75Cu0.25/SiO2, (e)

Au0.60Cu0.40/SiO2 and (f) Au0.50Cu0.50/SiO2.

Table 5.1: T50 for the Au1-xCux supported samples under CO oxidation and after being exposed to

oxidative or reductive pre-treatments.

Sample Oxidative pretreatment Reductive pretreatment

Heating phase Cooling phase Heating phase Cooling phase

Au0.75Cu0.25/Al2O3 242 °C 188 °C 163 °C 174 °C

Au0.60Cu0.40/Al2O3 199 °C 168 °C 150 °C 164 °C

Au0.50Cu0.50/Al2O3 202 °C 173 °C 150 °C 165 °C

Au0.75Cu0.25/SiO2 144 °C 147 °C 153 °C 147 °C

50 100 150 200 250 300

0102030405060708090

100

Oxidized Reduced

CO

con

vers

ion

(%)

Temperature (°C)

Au0.75Cu0.25/Al2O3

50 100 150 200 250 300

0102030405060708090

100

Oxidized Reduced

CO

con

vers

ion

(%)

Temperature (°C)

Au0.60Cu0.40/Al2O3

50 100 150 200 250 300

0102030405060708090

100

Oxidized Reduced

CO

con

vers

ion

(%)

Temperature (°C)

Au0.50Cu0.50/Al2O3

0 50 100 150 200 250 300

0102030405060708090

100

Au0.75Cu0.25/SiO2

Oxidized Reduced

CO

conv

ersi

on (%

)

Temperature (°C)0 50 100 150 200 250 300

0102030405060708090

100

Au0.60Cu0.40/SiO2

Oxidized Reduced

CO

conv

ersi

on (%

)

Temperature (°C)0 50 100 150 200 250 300

0102030405060708090

100

Au0.50Cu0.50/SiO2

Oxidized Reduced

CO

conv

ersi

on (%

)

Temperature (°C)

87

Au0.60Cu0.40/SiO2 110 °C 110°C 141 °C 116 °C

Au0.50Cu0.50/SiO2 129 °C 126°C 157 °C 126 °C

5.3.2 In situ measurements during the activation

Since a common pattern was found for all catalysts prepared using the same support,

independently of the Au:Cu ratio, a series of in situ experiments were performed with catalysts

prepared using the same batch of colloidal NPs (Au0.60Cu0.40), to understand the transformations

involved during pre-treatments and to identify the species formed during the catalytic cycles. As

mentioned before and discussed by Najafishirtari el al.98, after deposited on the support, the Au1-

xCux NPs are still capped by the organic ligands, so it is fundamental an oxidative treatment to

remove this organic layer and activate the catalyst exposing the metallic phase. Figure 5.4 shows

the in situ XRD patterns collected during the activation (Run#1) of Au0.60Cu0.40/Al2O3 and

Au0.60Cu0.40/SiO2 catalysts. The diffraction patterns presented in Figure 5.4 are obtained after

subtraction of the bare support signals, as described in the experimental section and indicated in

Figure C4.

30 35 40 45 50

350°C325°C

300°C290°C

220°C250°C

200°C150°C

100°C

Inte

nsity

(Arb

. Uni

ts)

2 (°)

25°C

Au0.70Cu0.60/Al2O3

(a)

30 35 40 45 50

Inte

nsity

(Arb

. Uni

ts.)

2 (°)

350°C

325°C300°C290°C250°C220°C200°C150°C100°C25°C

Au0.60Cu0.40/SiO2

(b)

Figure 5.4: (a) XRD patterns of Au0.60Cu0.40/Al2O3 and (b) Au0.60Cu0.40/SiO2 during the activation

(Run#1).

88

Figure 5.4 a,b shows the main (111) diffraction peak of the Au0.60Cu0.40 alloy phase at 2θ ~

40°. Comparing to the colloidal NPs (see Figure 5.1), the diffraction peaks at room temperature

become asymmetric after the incorporation of the NPs on both supports. Since the NPs sizes are

preserved after incorporation, the results suggest that the interaction with the supports surface

generates this asymmetry that seems more pronounced for the Au0.60Cu0.40/Al2O3catalyst. The

asymmetric peak can be simulated by considering the presence of two crystalline domains, with

different Cu contends. Interesting, by increasing the temperature during activation, it is clear that

the NPs go through a structural Figure C4 shows the XRD patterns and best fits using two

Gaussians in the range of 2θ=36-42° at selected temperatures.

For the Au0.60Cu0.40/Al2O3catalyst (Figure 5.4a and C5a,c), the fit obtained at 25°C

confirms that the (111) peak of the alloy can be simulated using two Gaussians, one centered at

38.9° (domain #1) and another at 40.0° (domain #2). According to the Vegard´s Law, domain #1

correspond to an Cu-poor alloy domain (6.6% of copper)and the domain#2 roughly matches the

original alloy ( 40.9% of copper). By increasing the temperature, both domains show a significant

decrease of Cu content and, at approximately 200 °C, there is only one metallic domain. At 350°C

the peak at 2θ =38.2° matches the Au metallic phase, indicating that the NPs go through a

dealloying process and at the end of the activation what remains are Au NPs and Cu species

highly dispersed on alumina. This newly formed Cu species are not detected by XRD.

Figure 5.4b shows the evolution of Au0.60Cu0.40/SiO2 during the activation. Comparing to

the behavior of on the Au0.60Cu0.40/Al2O3, it is also possible to observe a similar profile with two

different domains that evolved to a single one during activation. In this case, the initial pattern

domain #1 corresponds to pure Au and domain #2 to roughly the original alloy composition

(36.7% of Cu). At approximately 200°C, only one domain is observed, corresponding to Au NPs.

Similar to the previous case, a dealloying process takes place with the formation of CuOx species,

not detected by XRD. Interesting, a similar temperature of dealloying was observed by Bauer et

al.76 for the Au0.50Cu0,50 NPs in silica, indicating that the composition does not seem to affect the

temperature range of dealloying. Figure C4c,d shows the evolution of Cu% in each domain for

both catalysts during the activation.

To shade light in the Cu species formed during the activation, we perform in situ X-ray

absorption measurements. Figure 5.5 shows the XANES spectra at Cu K-edge of

Au0,60Cu0,40/Al2O3, measured during activation (Run#1).

89

8900 8950 9000 9050 9100 9150 9200

0,0

0,2

0,4

0,6

0,8

1,0

1,2

Nor

mal

ized

abs

orpt

ion

Tempe

ratur

e (°C

)

Activation

350250

150

Energy (eV)

50Cu_K

Au0.60Cu0.40/Al2O3

(a)

8980 9000 9020

Nor

mal

ized

abs

orpt

ion

E nergy (eV )

C u0

350 °C_ f

200 °C

CuOCu

2O

350 °C_ i

250 °C

150 °C

R T

(b)

8900 8950 9000 9050 9100 9150 9200

0,0

0,2

0,4

0,6

0,8

1,0

1,2

Au0.60

Cu0.40

/SiO2

Energy (eV)

Nor

mal

ized

abs

orpt

ion

50150

250350

Tempe

ratur

e (°C

)

Cu_KActivation

(c)

8980 9000 9020

Nor

mal

ized

Abs

orpt

ion

Energy (eV)

Cu0

350 °C_f

200 °C

CuO

Cu2O

350 °C_i

250 °C

150 °C

RT

(d)

Figure 5.5: In situ XANES at Cu K-edge during activation of (a, b) Au0,60Cu0,40/Al2O3.and (c,d)

Au0,60Cu0,40/SiO2. The black, red, green and blue dashed lines in (b) and (d) indicate the 8980,

8993, 8997 and 9002 eV energies, respectively.

Comparing the initial spectra, at room temperature, of Au0.60Cu0.40/Al2O3 and

Au0.60Cu0.40/SiO2 with the Cu metallic standard it is possible to observe a pre-peak at

approximately 8980 eV that confirms the presence of Cu in metallic state. The attenuation and

shift of the features at 8993 and 9002eV are expected due to the interaction with Au in the alloy131.

90

The attenuation of the feature at approximately 8997 eV indicates the presence of Cu oxidized

species presented already at room temperature. These results are in line with the XRD data,

indicating that a part of the Cu initially present in the colloidal alloy NPs is oxidized by the

contact with the support surface. A similar trend was also observed in the EXAFS oscillations (not

shown). The quantitative analysis shows the presence of Cu-Au and Cu-Cu bonds consistent with

alloy structure; however, Cu-O bonds were also observed after the deposition on both supports

(Table 5.2).

Table 5.2: First nearest-neighbor average coordination number (CN), bond lengths (R) and mean-square deviations (σ2) determined from EXAFS data collected at Cu K edge at room temperature.

Sample-condition Cu K edge

Scattering CN R (Å) σ2(Å2) R-factor

Cu foil Cu-Cu 12 2,547 0,009 0,006

Cu-Cu 6 3,605 0,009

CuO Cu-O 2 1,955 0,007 0,022

Cu-O 2 1,965 0,007

Cu-O 2 2,788 0,008

Cu-Cu 4 2,906 0,008

Au0.60Cu0.40-coloidal Cu-Cu 2,55 2,679 0,014 0,021

Cu-Au 6,08 2,682 0,014

Au0.60Cu0.40/Al2O3 Fresh Cu-O 0,97 1,943 0,007 0,030

Cu-O 0,97 1,953 0,007

Cu-Cu 2,10 2,692 0,013

Cu-Au 4,54 2,641 0,013

After activation Cu-O 1,50 1,957 0,007 0,007

Cu-O 1,50 1,967 0,007

Cu-Cu 0,35 2,913 0,013

After 1st CO Oxidation reaction Cu-O 1,39 1,953 0,007 0,007

91

Cu-O 1,39 1,963 0,007

Cu-Cu 0,22 2,987 0,013

After reducing treatment Cu-O 0,56 1,924 0,007 0,019

Cu-O 0,56 1,934 0,007

Cu-Cu 1,38 2,633 0,013

Cu-Au 4,38 2,659 0,013

Au0.60Cu0.40/SiO2

Fresh Cu-O 0,47 1,950 0,007 0,012

Cu-O 0,47 1,960 0,007

Cu-Cu 1,38 2,701 0,013

Cu-Au 5,97 2,678 0,013

After activation Cu-O 1,68 1,959 0,007 0,004

Cu-O 1,68 1,956 0,007

Cu-Cu 2,27 2,857 0,013

After 1st CO Oxidation reaction Cu-O 1,51 1,948 0,007 0,004

Cu-O 1,51 1,958 0,007

Cu-Cu 1,87 2,857 0,013

After reducing treatment Cu-Cu 3,07 2,537 0,013 0,010

Cu-Au 5,11 2,686 0,013

During activation (Figure 5.5), a clear evolution of the Cu environment can be seen. After

200°C, there is a significant modification on the XANES spectra at 8980 and 8997 eV, indicating

that a major oxidation takes place. This result also agrees with XRD data that shows that the

dealloying process is completed at this temperature. At the end of the activation process, the

XANES spectra for both catalysts resemble the CuO standard confirming that the Cu species are

totally oxidized Interesting, EXAFS analysis confirm the presence of Cu-O bonds for both

catalysts but differences were found in the Cu-Cu second shell. This contribution is much smaller

in the case of Au0.60Cu0.40/Al2O3, suggesting that CuOx species are more dispersed in the Al2O3.

Figure 5.7 shows the AuL3-edge measurements during activation for the Au0.60Cu0.40/SiO2

catalyst a metallic profile is observed during the entire process; however, it is possible to note a

92

modification on the pre-edge region regarding the dealloying of AuCu during the activation.

Around 11930 eV, it is observed a characteristic feature of the Au1-xCux alloy, related to the d-

electron transfer from Au to Cu131. The evolution to metallic Au can be easily followed during

activation.

0,0

0,2

0,4

0,6

0,8

1,0

11800 11900 12000 12100

ActivationAu L3

Temperature (°C)

350250

150

Nor

mal

ized

abs

orpt

ion

Energy (eV)

50

Au0.60Cu0.40/SiO2

11920 11930 11940 11950 119600,60

0,65

0,70

0,75

0,80

0,85

0,90

0,95

1,00

Room temperature 150 °C 250 °C 350 °C Au metallic (Standard)

Nom

aliz

ed a

bsor

ptio

n

Energy (eV)

Figure 5.6: XANES at AuL3-edge during the activation (step#1) of Au0.60Cu0.40/SiO2.

The differences between Au0.60Cu0.40/Al2O3 and Au0.60Cu0.40/SiO2 activations, in particular

the higher dispersion of the CuOx species found in the case of Au0.60Cu0.40/Al2O3were confirmed

by TEM-EDS mapping (Figure 5.7). By comparing images acquired at the end of the activation

process, it is possible to observe that while the Cu species are well spread on the Al2O3 surface, in

the case of SiO2 these species remain localized nearest to the Au NPs. These results clearly show

that the dealloying mechanism that takes place during the activation is related to the metal-support

interaction.

93

(a)

(b)

Figure 5.7: TEM-EDS images after activation of (a) Au0.60Cu0.40/Al2O3 and (b) Au0.60Cu0.40/SiO2.

5.3.3 In situ measurements during the CO oxidation

In situ XRD and XANES analysis were also carried out under CO oxidation reaction. The

initial catalyst is formed by Au NPs and CuOx species dispersed on the support. Only slightly

modifications could be found for both catalysts. By XRD, it was possible to note a small shift to

lower 2θ at higher temperatures (Figure C7). More interesting, after the cooling phase the (111)

peak position is exactly the same as the initial one, indicating no irreversible changes in the Au

phase took place during the reaction. XANES measurements at the Cu K-edge also show a slight

modification on the pre-edge peak and the white line during the heating, for both supports (Figure

C8). After cooling, the XANES profiles are similar to the initial ones, indicating that the CuOx

species do not change significantly during the reaction. The EXAFS data before and after CO

oxidation reaction confirms that the reaction atmosphere does not affect the nature of the Cu

species.

94

5.3.4 In situ measurements during the reduction

A similar set of in situ measurements were also performed under reductive pre-treatment,

following the CO oxidation reaction (Step#3). Figure 5.8a-d shows the XANES spectra at Cu K-

edge acquired during the reduction of Au0.60Cu0.40/Al2O3 and Au0.60Cu0.40/SiO2. After CO

oxidation, CuOx species remain dispersed on the support. By increasing the temperature under

reducing conditions, it is possible to note significant modifications on the white line corresponding

to copper reduction. These modifications are intensified after 200 °C and, at 350 °C, the XANES

spectra indicates the presence of both Cu0 and CuOx species. . XANES at Au L3-edge measured

for Au0.60Cu0.40/SiO2 (Figure 5.8e,f) revealed that a re-alloying process takes place with the

formation of AuCu NPs.

8900 9000 9100 9200

0,00,20,40,60,81,01,21,4

Energy (eV)

Nor

mal

ized

abs

orpt

ion

Cu KReduction

Au0.60Cu0.40/Al2O3

50150

250350

Tempe

ratur

e (°C

)

(a)

8980 9000 9020

Nom

aliz

ed a

bsor

ptio

n

Energy (eV)

Cu0

200 °C

CuOCu2O

350 °C

250 °C

150 °C

RT

(b)

95

8900 9000 9100 9200

0,00,20,40,60,81,01,21,4

Energy (eV)

Nor

mal

ized

abs

orpt

ion

Cu KReduction

Au0.60Cu0.40/SiO2 350250

15050

Tempe

ratur

e (°C

)

(c)

8980 9000 9020

Nor

mal

ized

abs

orpt

ion

Energy (eV)

Cu0

200 °C

CuOCu2O

350 °C

250 °C

150 °C

RT

(d)

0,0

0,2

0,4

0,6

0,8

1,0

11800 11900 12000 12100

Au0.60Cu0.40/SiO2

Energy (eV)

Nor

mal

ized

abs

orpt

ion

Au L3Reduction

50150

250350

Temperature (°C)

(e)

11921 11928 119350,4

0,5

0,6

0,7

0,8

0,9

Au0

RT 150 °C 200 °C 250 °C 350 °C 350 °C_ after 30 min

Nor

mal

ized

abs

orpt

ion

Energy (eV)

(f)

Figure 5.8: In situ XANES during reduction (step#3) of (a) and (b) Au0.60Cu0.40/Al2O3 and (c) and

(d) Au0,60Cu0,40/SiO2 at Cu K-edge. (d) and (f) shown the XANES at AuL3-edge during the

reduction of Au0.60Cu0.40/SiO2.

EXAFS data collected at the Cu K-edge (not shown) confirm a very interesting trend

comparing the systems supported on silica versus alumina. For Au0.60Cu0.40/SiO2 it is possible to

observe the appearance of Au-Cu bonds, indicating the once the Cu is reduced part of it migrates

diffuse into the Au NPs, forming and alloy. In addition, it is also observed the formation of shorter

96

Cu-Cu interactions characteristic of monometallic Cu, suggesting the formation of a second

population formed by Cu NPs or small metallic Cu domains. Moreover, it is also possible to note

the absence of Cu-O bonds, indicating the reduction of all CuOx species. On the other hand, for

Au0.60Cu0.40/Al2O3, the Cu-O contribution was still present, indicating that part of Cu species

remain oxidized after the reductive step. The formation of reduced Cu in an ordered segregated

phase was not observed and the analysis of the AuCu CNs suggests that this sample has a higher

reincorporation of Cu atoms into the alloy, as observed by the decrease in the number of Au

neighbours seen by Cu (NCuAu) when compared to the reduced Au0.60Cu0.40/SiO2 (see Table 5.2).

The differences between the interaction of metal phase and Al2O3 or SiO2 are also

confirmed by XRD data. ForAu0.60Cu0.40/Al2O3 (Figure 5.9a) it is possible to see, besides the

original Au domain, the appearance of a new metallic domain related to a copper-riched phase

after 200 °C (see figure S9a). Both domains are enriched by Cu by increasing the temperature and,

at the end of reduction process, the domains #1 and #2 are centered at 2θ=38.9° and 40.1°

respectively, corresponding to a richer Cu phase with 33.1% of Cu and another one with 14.2% of

Cu, according to the Vegard´s Law.

For Au0.60Cu0.40/SiO2it is possible to see the Cu incorporation but without a clear

appearance of a new domain (Figure 5.9b).At the beginning of the reduction process, only one

peak at 2θ= 38.3 ° is observed, related to Au NPs while at the end the final peak correspond to an

alloy with 11% of Cu content (see figure C9b).

30 35 40 45 50

350°C325°C300°C290°C250°C

220°C200°C

150°C100°CIn

tens

ity (A

rb. U

nits

)

2

25°C

(a)

30 35 40 45 50

Inte

nsity

(Arb

. Uni

ts)

2

350°C325°C300°C290°C250°C

220°C200°C

150°C100°C

25°C

(b)

Figure 5.9: XRD patterns of (a) Au0.60Cu0.40/Al2O3 and (b) Au0.60Cu0.40/SiO2 during the Step#3

(reductive). The black dashed lines indicate the initial position of (111).

97

The differences of Au0.60Cu0.40/Al2O3 and Au0.60Cu0.40/SiO2 after reduction are also

confirmed by TEM-EDS mapping, indicating a higher dispersion of Cu species of

Au0.60Cu0.40/Al2O3 then Au0.60Cu0.40/SiO2.

(a)

(b)

Figure 5.10: TEM-EDS images after the reduction of (a) Au0.60Cu0.40/Al2O3 and (b)

Au0.60Cu0.40/SiO2.

5.4 Conclusion

The redox pre-treatments performed for Au1-xCux supported on SiO2 and Al2O3 directly

impact in the catalytic performance. The oxidative pre-treatment induces the dealloying of the

Au1-xCux nanoparticles resulting in Au nanoparticles and CuOx phase dispersed on the support;

this process and the characteristics of the CuOx phase are support-depended. This final

configuration is stable under CO oxidation. When the reductive pre-treatment is performed, the

CuOx species are reduced and part of the copper migrates forming the alloy nanoparticles, but with

lower copper contend. This work clearly highlights the crucial role played by the support on the

stability of the alloy phase, deeply impacting in the final performance of the catalyst.

98

Chapter 6

Global discussion and conclusions

In this work it was possible to demonstrate the potentialities of the bimetallic colloidal

NPs synthesis with a good control of properties targeting catalytic applications.

Firstly it was studied the formation of colloidal AuCu NPs. In the Chapter 2, it was

possible to shown the synthesis of homogeneous spherical NPs and understand the effect of

the temperature in the final composition. By changing the final synthesis temperature, 225,

260 and 280 °C, it was possible to synthesize particles with the compositions of Au0.75Cu0.25,

Au0.60Cu0.40 and Au0.50Cu0.50 respectively, with average size of 14 nm. Using different

characterization techniques to evaluate aliquots taken at different stages of reaction, it was

possible to infer about the formation mechanism of AuCu NPs, confirming the presence of an

intermediate Cu-rich phase that is digested during the synthesis to form the alloy. Another

important step, described in the Chapter 3, was given by in situ XAFS measurements, which

enabled monitoring the evolution of the metallic precursors in real-time during the synthesis,

using a special reactor developed by our group. The results supported the solid-state diffusion

mechanism for the AuCu alloy formation.

In the Chapter 4, Au0.75Cu0.25 NPs were supported on silica and exposed to redox

pretreatments and CO oxidation reaction cycles. The pretreatments performed before the

catalytic oxidation of CO showed direct influence on the catalytic performance. After the

activation step, the CO conversion is very low but increases considerably after a reductive

pretreatment. After that, the catalyst becomes highly stable. This behavior can be related to

the removal of Cl residues, detected on the catalysts, by the reductive pretreatment and/or the

formation of a new chemical environment generating stable species on the surface that can

improve the CO conversion. The most interesting point in this stage was that, after alternate

oxidative/reductive pretreatments the CO conversion is totally stable, showing that the

material reaches a high activeequilibrium state.

99

After the promising results obtained for Au0.75Cu0.25 supported on silica, tests were

carried out to verify the stability and catalytic performance of different compositions of Au1-

xCux supported on silica and alumina. The Chapter 5 extensively explored in situ

characterization techniques to clarify the processes involved in redox pretreatments and CO

oxidation reaction cycles. By combining in situ X-ray absorption and diffraction, it was

observed a dealloying process under oxidative atmosphere and re-alloying under reductive

conditions. The AuCu alloying/dealloying process is directly influenced by the support,

generating different species on silica or alumina that directly affect the performance of the

catalyst.

This doctoral thesis explored different aspects of the use of the AuCu system in

catalysis, from the alloying mechanisms that take place in colloidal systems to its use in CO

oxidation reaction, highlighting that the species formed during pretreatments and the nature of

the support direct impact in the catalytic performance of the material.

Further studies may be carried out by changing the particle size of AuCu, using

different supports and also on the application of AuCu NPs in other reactions where a well-

defined metal-support interface is required.

100

Bibliography

(1) Li, Y.; Somorjai, G. A. Nanoscale Advances in Catalysis and Energy Applications.

Nano Lett.2010, 10, 2289–2295.

(2) Daniel, M. C. M.; Astruc, D. Gold Nanoparticles: Assembly, Supramolecular

Chemistry, Quantum-Size Related Properties and Applications toward Biology,

Catalysis and Nanotechnology,. Chem. Rev.2004, 104, 293–346.

(3) Carbone, L.; Cozzoli, P. D. Colloidal Heterostructured Nanocrystals: Synthesis and

Growth Mechanisms. Nano Today2010, 5, 449–493.

(4) Bonnemann, H.; Richards, R. M. Nanoscopic Metal Particles - Synthetic Methods and

Potential Applications. Eur. J. Inorg. Chem.2001, 2455–2480.

(5) Ghosh Chaudhuri, R.; Paria, S. Core/shell Nanoparticles: Classes, Properties, Synthesis

Mechanisms, Characterization, and Applications. Chem. Rev.2012, 112, 2373–2433.

(6) Thanh, N. T. K.; Maclean, N.; Mahiddine, S. Mechanisms of Nucleation and Growth of

Nanoparticles in Solution. Chem. Rev.2014, 3.

(7) Doyle, H.; Betley, T. A. Colloidal Synthesis of Nanocrystals and Nanocrystal

Superlattices This Paper Provides an Overview of the. 2001, 45, 47–56.

(8) Sneed, B. T.; Young, A. P.; Tsung, C.-K. Building up Strain in Colloidal Metal

Nanoparticle Catalysts. Nanoscale2015, 7, 12248–12265.

(9) Jia, C.-J.; Schüth, F. Colloidal Metal Nanoparticles as a Component of Designed

Catalyst. Phys. Chem. Chem. Phys.2011, 13, 2457–2487.

(10) Manna, L.; Kudera, S. Colloidal Foundations of Nanoscience; Berti, D.; Palazzo, G.,

Eds.; First edit.; Elsevier, 2014.

(11) Faraday, M. Experimental Relations of Gold (and Other Metals) to Light. Philos.

Trans. R. Soc. London1857, 147, 145.

(12) Yin, Y.; Alivisatos, a P. Colloidal Nanocrystal Synthesis and the Organic-Inorganic

Interface. Nature2005, 437, 664–670.

(13) Sugimoto, T. Preparation of Monodispersed Colloidal Particles. Adv. Colloid Interface

Sci.1987, 28, 65–108.

(14) Turkevich, J.; Stevenson, P. C.; Hillier, J. A Study of the Nucleation and Growth

101

Processes in the Synthesis of Colloidal Gold. Discuss. Faraday Soc.1951, 55, 55–75.

(15) Vreeland, E. C.; Watt, J.; Schober, G. B.; Hance, B. G.; Austin, M. J.; Price, A. D.;

Fellows, B. D.; Monson, T. C.; Hudak, N. S.; Maldonado-Camargo, L.; et al. Enhanced

Nanoparticle Size Control by Extending LaMer’s Mechanism. Chem. Mater.2015, 27,

6059−6066.

(16) LaMer, V.; Dinegar, R. Theory, Production and Mechanism of Formation of

Monodispersed Hydrosols. J. Am. Chem. Soc.1950, 72, 4847–4854.

(17) Romo-Herrera, J. M.; Alvarez-Puebla, R. A.; Liz-Marzán, L. M. Controlled Assembly

of Plasmonic Colloidal Nanoparticle Clusters. Nanoscale2011, 3, 1304–1315.

(18) You, H.; Yang, S.; Ding, B.; Yang, H. Synthesis of Colloidal Metal and Metal Alloy

Nanoparticles for Electrochemical Energy Applications. Chem. Soc. Rev.2013, 42,

2880–2904.

(19) Somorjai, G. A.; Tao, F.; Park, J. Y. The Nanoscience Revolution: Merging of Colloid

Science, Catalysis and Nanoelectronics. Top. Catal.2008, 47, 1–14.

(20) Piella, J.; Bastús, N. G.; Puntes, V. Size-Controlled Synthesis of Sub-10-Nanometer

Citrate-Stabilized Gold Nanoparticles and Related Optical Properties. Chem.

Mater.2016, 28, 1066−1075.

(21) Koutsopoulos, S.; Johannessen, T.; Eriksen, K. M.; Fehrmann, R. Titania-Supported Pt

and Pt-Pd Nanoparticle Catalysts for the Oxidation of Sulfur Dioxide. J. Catal.2006,

238, 206–213.

(22) Wu, Y.; Cai, S.; Wang, D.; He, W.; Li, Y. Syntheses of Water-Soluble Octahedral,

Truncated Octahedral, and Cubic Pt-Ni Nanocrystals and Their Structure-Activity

Study in Model Hydrogenation Reactions. J. Am. Chem. Soc.2012, 134, 8975–8981.

(23) Wang, Y.; Zhao, Y.; Yin, J.; Liu, M.; Dong, Q.; Su, Y. Synthesis and Electrocatalytic

Alcohol Oxidation Performance of Pd-Co Bimetallic Nanoparticles Supported on

Graphene. Int. J. Hydrogen Energy2014, 39, 1325–1335.

(24) Huang, X.; Wang, X.; Tan, M.; Zou, X.; Ding, W.; Lu, X. Selective Oxidation of

Alcohols on P123-Stabilized Au-Ag Alloy Nanoparticles in Aqueous Solution with

Molecular Oxygen. Appl. Catal. A Gen. 2013, 467, 407–413.

(25) An, K.; Alayoglu, S.; Ewers, T.; Somorjai, G. A. J. Colloid Interface Sci. 2012, 373, 1–

102

13.

(26) Johnston, R. L. Metal Nanoparticles and Nanoalloys; 1st ed.; Elsevier Ltd, 2012; Vol.

3.

(27) Pendergast, D. M. Tumbaga Object from the Early Classic Period, Found at Altun Ha,

British Honduras (Belize). Science19170, 168, 116–118.

(28) Yin, F.; Wang, Z. W.; Palmer, R. E. Formation of Bimetallic Nanoalloys by Au

Coating of Size-Selected Cu Clusters. J. Nanoparticle Res.2012, 14, 1114–1124.

(29) Xu, Z.; Lai, E.; Shao-Horn, Y.; Hamad-Schifferli, K. Compositional Dependence of the

Stability of AuCu Alloy Nanoparticles. Chem. Commun. (Camb).2012, 48, 5626–5628.

(30) Ferro, R.; Saccone, A. Elements of Alloying Behaviour Systematics. In Intermetallic

Chemistry; Elsevier, 2008.

(31) Guisbiers, G.; Mejia-Rosales, S.; Khanal, S.; Ruiz-Zepeda, F.; Whetten, R. L.; José-

Yacaman, M. Gold−Copper Nano-Alloy, “Tumbaga ”, in the Era of Nano : Phase

Diagram and Segregation. Nano Lett.2014, 14, 6718–6726.

(32) Lubarda, V. A. On the Effective Lattice Parameter of Binary Alloys. 2003, 35, 53–68.

(33) Okamoto, H.; Chakrabarti, D. J.; Laughlin, D. E.; Massalski, T. B. The Au-Cu (Gold-

Copper) System. Bull. Alloy Phase Diagrams1987, 454–473.

(34) Bracey, C. L.; Ellis, P. R.; Hutchings, G. J. Application of Copper-Gold Alloys in

Catalysis: Current Status and Future Perspectives. Chem. Soc. Rev.2009, 38, 2231–

2243.

(35) Baletto, F.; Ferrando, R. Structural Properties of Nanoclusters: Energetic,

Thermodynamic, and Kinetic Effects. Rev. Mod. Phys.2005, 77, 371–423.

(36) Pauwels, B.; Van Tendeloo, G.; Zhurkin, E.; Hou, M.; Verschoren, G.; Theil Kuhn, L.;

Bouwen, W.; Lievens, P. Transmission Electron Microscopy and Monte Carlo

Simulations of Ordering in Au-Cu Clusters Produced in a Laser Vaporization Source.

Phys. Rev. B2001, 63, 1–10.

(37) Ferrando, R.; Jellinek, J.; Johnston, R. L. Nanoalloys: From Theory to Applications of

Alloy Clusters and Nanoparticles. Chem. Rev.2008, 108, 845–910.

(38) Darby, S.; Mortimer-Jones, T. V.; Johnston, R. L.; Roberts, C. Theoretical Study of

Cu–Au Nanoalloy Clusters Using a Genetic Algorithm. J. Chem. Phys.2002, 116,

103

1536.

(39) Toai, T. J.; Rossi, G.; Ferrando, R. Global Optimisation and Growth Simulation of

AuCu Clusters. Faraday Discuss.2008, 138, 49–58.

(40) Kim, D.; Resasco, J.; Yu, Y.; Asiri, A. M.; Yang, P. Synergistic Geometric and

Electronic Effects for Electrochemical Reduction of Carbon Dioxide Using Gold-

Copper Bimetallic Nanoparticles. Nat. Commun.2014, 5, 1–8.

(41) Chorkendorff, I.; Niemantsverdriet, J. W. Concepts of Modern Catalysis and Kinetics;

2003.

(42) Bell, A. T. The Impact of Nanoscience on Heterogeneous Catalysis. Science2003, 299,

1688–1691.

(43) Bamwenda, G. R.; Tsubota, S.; Nakamura, T.; Haruta, M. The Influence of the

Preparation Methods on the Catalytic Activity of Platinum and Gold Supported on TiO

2 for CO Oxidation. Catal. Letters1997, 44, 83–87.

(44) Munnik, P.; de Jongh, P. E.; de Jong, K. P. Recent Developments in the Synthesis of

Supported Catalysts. Chem. Rev.2015.

(45) Baatz, C.; Decker, N.; Prube, U. New Innovative Gold Catalysts Prepared by an

Improved Incipient Wetness Method. J. Catal.2008, 258, 165–169.

(46) Schauermann, S.; Nilius, N.; Shaikhutdinov, S.; Freund, H.-J. Nanoparticles for

Heterogeneous Catalysis: New Mechanistic Insights. Acc. Chem. Res.2013, 46, 1673–

1681.

(47) Cao, A.; Lu, R.; Veser, G. Stabilizing Metal Nanoparticles for Heterogeneous

Catalysis. Phys. Chem. Chem. Phys.2010, 12, 13499–13510.

(48) Tao, F. F.; Schneider, W. F.; Kamat, P. V. Chemical Synthesis of Nanscale

Heterogeneous Catalysis. In Heterogeneous Catalysis at the Nanoscale for Energy

Applications; 2014; Vol. 1, pp. 9–29.

(49) Sonström, P.; Bäumer, M. Supported Colloidal Nanoparticles in Heterogeneous Gas

Phase Catalysis: On the Way to Tailored Catalysts. Phys. Chem. Chem. Phys.2011, 13,

19270–19284.

(50) Pushkarev, V. V.; Zhu, Z.; An, K.; Hervier, A.; Somorjai, G. A. Monodisperse Metal

Nanoparticle Catalysts: Synthesis, Characterizations, and Molecular Studies under

104

Reaction Conditions. Top. Catal.2012, 55, 1257–1275.

(51) Haruta, M.; Kobayashi, T.; Sano, H.; Yamada, N. Novel Gold Catalysts for the

Oxidation of Carbon Monoxide at a Temperature Far below 0°C. Chem. Lett.1987,

405–408.

(52) Zafeiratos, S.; Piccinin, S.; Teschner, D. Alloys in Catalysis: Phase Separation and

Surface Segregation Phenomena in Response to the Reactive Environment. Catal. Sci.

Technol.2012, 2, 1787–1801.

(53) Xu, W.; Si, R.; Senanayake, S. D.; Llorca, J.; Idriss, H.; Stacchiola, D.; Hanson, J. C.;

Rodriguez, J. a. In Situ Studies of CeO2-Supported Pt, Ru, and Pt–Ru Alloy Catalysts

for the Water–Gas Shift Reaction: Active Phases and Reaction Intermediates. J.

Catal.2012, 291, 117–126.

(54) Zhou, S.; Jackson, G. S.; Eichhorn, B. AuPt Alloy Nanoparticles for CO-Tolerant

Hydrogen Activation: Architectural Effects in Au-Pt Bimetallic Nanocatalysts. Adv.

Funct. Mater.2007, 17, 3099–3104.

(55) Liu, X.; Wang, A.; Zhang, T.; Su, D. S.; Mou, C. Y. Au-Cu Alloy Nanoparticles

Supported on Silica Gel as Catalyst for CO Oxidation: Effects of Au/Cu Ratios. Catal.

Today2011, 160, 103–108.

(56) Liu, J. H.; Wang, A. Q.; Chi, Y. S.; Lin, H. P.; Mou, C. Y. Synergistic Effect in an Au-

Ag Alloy Nanocatalyst: CO Oxidation. J. Phys. Chem. B2005, 109, 40–43.

(57) Remediakis, I. N.; Lopez, N.; Nørskov, J. K. CO Oxidation on Gold Nanoparticles:

Theoretical Studies. Appl. Catal. A Gen.2005, 291, 13–20.

(58) Singh, A. K.; Xu, Q. Synergistic Catalysis over Bimetallic Alloy Nanoparticles.

ChemCatChem2013, 5, 652–676.

(59) Xu, J.; White, T.; Li, P.; He, C.; Yu, J.; Yuan, W.; Han, Y.-F. Biphasic Pd-Au Alloy

Catalyst for Low-Temperature CO Oxidation. J. Am. Chem. Soc.2010, 132, 10398–

10406.

(60) Liu, H.; Kozlov, A. I.; Kozlova, A. P.; Shido, T.; Asakura, K.; Iwasawa, Y. Active

Oxygen Species and Mechanism for Low-Temperature CO Oxidation Reaction on a

TiO2-Supported Au Catalyst Prepared from Au(PPh3)(NO3) and As-Precipitated

Titanium Hydroxide. J. Catal.1999, 185, 252–264.

105

(61) Molina, L. M.; Hammer, B. Some Recent Theoretical Advances in the Understanding

of the Catalytic Activity of Au. Appl. Catal. A Gen.2005, 291, 21–31.

(62) Haruta, M. Size- and Support-Dependency in the Catalysis of Gold. Catal. Today1997,

36, 153–166.

(63) Boccuzzi, F.; Chiorino, a; Manzoli, M.; Lu, P.; Akita, T.; Ichikawa, S.; Haruta, M.

Au/TiO2 Nanosized Samples: A Catalytic, TEM, and FTIR Study of the Effect of

Calcination Temperature on the CO Oxidation. J. Catal.2001, 202, 256–267.

(64) 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. Size and Support Effects for the Water-Gas Shift

Catalysis over Gold Nanoparticles Supported on Model Al2O3 and TiO2. J. Am. Chem.

Soc.2012, 134, 4700–4708.

(65) Choudhary, T. V.; Goodman, D. W. Catalytically Active Gold: The Role of Cluster

Morphology. Appl. Catal. A Gen.2005, 291, 32–36.

(66) Schubert, M. M.; Hackenberg, S.; van Veen, A. C.; Muhler, M.; Plzak, V.; Behm, R. J.

CO Oxidation over Supported Gold Catalysts—“Inert” and “Active” Support Materials

and Their Role for the Oxygen Supply during Reaction. J. Catal.2001, 197, 113–122.

(67) Klyushin, A. Y.; Arrigo, R.; Youngmi, Y.; Xie, Z.; Hävecker, M.; Bukhtiyarov, A. V.;

Prosvirin, I. P.; Bukhtiyarov, V. I.; Knop-Gericke, A.; Schlögl, R. Are Au

Nanoparticles on Oxygen-Free Supports Catalytically Active? Top. Catal.2016, 59,

469–477.

(68) Liu, X.; Wang, A.; Yang, X.; Zhang, T.; Mou, C. Y.; Su, D. S.; Li, J. Synthesis of

Thermally Stable and Highly Active Bimetallic Au-Ag Nanoparticles on Inert

Supports. Chem. Mater.2009, 21, 410–418.

(69) Mu, R.; Fu, Q.; Xu, H.; Zhang, H.; Huang, Y.; Jiang, Z.; Zhang, S.; Tan, D.; Bao, X.

Synergetic Effect of Surface and Subsurface Ni Species at Pt-Ni Bimetallic Catalysts

for CO Oxidation. J. Am. Chem. Soc.2011, 133, 1978–1986.

(70) Yen, C.; Lin, M.; Wang, A.; Chen, S.; Chen, J.; Mou, C. CO Oxidation Catalyzed by

Au− Ag Bimetallic Nanoparticles Supported in Mesoporous Silica. J. Phys. Chem.

C2009, 113, 17831–17839.

(71) Wang, A. Q.; Chang, C. M.; Mou, C. Y. Evolution of Catalytic Activity of Au-Ag

Bimetallic Nanoparticles on Mesoporous Support for CO Oxidation. J. Phys. Chem.

106

B2005, 109, 18860–18867.

(72) Liu, X.; Wang, A.; Li, L.; Zhang, T.; Mou, C. Y.; Lee, J. F. Structural Changes of Au-

Cu Bimetallic Catalysts in CO Oxidation: In Situ XRD, EPR, XANES, and FT-IR

Characterizations. J. Catal.2011, 278, 288–296.

(73) Wang, A.; Liu, X. Y.; Mou, C.-Y.; Zhang, T. Understanding the Synergistic Effects of

Gold Bimetallic Catalysts. J. Catal.2013, 308, 258–271.

(74) Zhou, S.; Varughese, B.; Eichhorn, B.; Jackson, G.; McIlwrath, K. Pt-Cu Core-Shell

and Alloy Nanoparticles for Heterogeneous NOx Reduction: Anomalous Stability and

Reactivity of a Core-Shell Nanostructure. Angew. Chemie - Int. Ed.2005, 44, 4539–

4543.

(75) Cargnello, M.; Doan-Nguyen, V. V. T.; Gordon, T. R.; Diaz, R. E.; Stach, E. a; Gorte,

R. J.; Fornasiero, P.; Murray, C. B. Control of Metal Nanocrystal Size Reveals Metal-

Support Interface Role for Ceria Catalysts. Science2013, 341, 771–773.

(76) Yen, H.; Seo, Y.; Kaliaguine, S.; Kleitz, F. On the Role of Metal-Support Interactions,

Particle Size, and Metal-Metal Synergy in CuNi Nanocatalysts for H2 Generation. ACS

Catal.2015, 2, 5505−5511.

(77) Bauer, J. C.; Mullins, D.; Li, M.; Wu, Z.; Payzant, E. A.; Overbury, S. H.; Dai, S.

Synthesis of Silica Supported AuCu Nanoparticle Catalysts and the Effects of

Pretreatment Conditions for the CO Oxidation Reaction. Phys. Chem. Chem.

Phys.2011, 13, 2571–2581.

(78) Li, X.; See, S.; Fang, S.; Teo, J.; Foo, Y. L.; Borgna, A.; Lin, M.; Zhong, Z. Activation

and Deactivation of Au−Cu/SBA-15 Catalyst for Preferential Oxidation of CO in H2-

Rich Gas. ACS Catal.2012, 2, 360–369.

(79) Sugano, Y.; Shiraishi, Y.; Tsukamoto, D.; Ichikawa, S.; Tanaka, S.; Hirai, T. Supported

Au-Cu Bimetallic Alloy Nanoparticles: An Aerobic Oxidation Catalyst with

Regenerable Activity by Visible-Light Irradiation. Angew. Chemie2013, 125, 5403–

5407.

(80) Motl, N. E.; Ewusi-Annan, E.; Sines, I. T.; Jensen, L.; Schaak, R. E. Au-Cu Alloy

Nanoparticles with Tunable Compositions and Plasmonic Properties: Experimental

Determination of Composition and Correlation with Theory. J. Phys. Chem.2010, 114,

19263–19269.

107

(81) Toshima, N.; Yonezawa, T. Bimetallic Nanoparticles—novel Materials for Chemical

and Physical Applications. New J. Chem.1998, 22, 1179–1201.

(82) Jana, S. Advances in Nanoscale Alloys and Intermetallics: Low Temperature Solution

Chemistry Synthesis and Application in Catalysis. Dalt. Trans.2015, 44, 18692–18717.

(83) Major, K. J.; De, C.; Obare, S. O. Recent Advances in the Synthesis of Plasmonic

Bimetallic Nanoparticles. Plasmonics2009, 4, 61–78.

(84) Brust, M.; Walker, M.; Bethell, D.; Schiffrin, D. J.; Whyman, R. Synthesis of Thiol-

Derivatised Gold Nanoparticles in a Two-Phase Liquid-Liquid System Mathias. Chem.

Commun.2000, 801–802.

(85) Sardar, R.; Funston, A. M.; Mulvaney, P.; Murray, R. W. Gold Nanoparticles: Past,

Present, and Future. Langmuir2009, 25, 13840–13851.

(86) DeSantis, C. J.; Weiner, R. G.; Radmilovic, A.; Bower, M. M.; Skrabalak, S. E.

Seeding Bimetallic Nanostructures as a New Class of Plasmonic Colloids. J. Phys.

Chem. Lett.2013, 4, 3072–3082.

(87) Sra, A. K.; Schaak, R. E. Synthesis of Atomically Ordered AuCu and AuCu 3

Nanocrystals from Bimetallic Nanoparticle Precursors. J. Am. Chem. Soc.2004, 6667–

6672.

(88) Barcaro, G.; Fortunelli, A.; Rossi, G.; Nita, F.; Ferrando, R. Electronic and Structural

Shell Closure in AgCu and AuCu Nanoclusters. J. Phys. Chem. B2006, 110, 23197–

23203.

(89) Alloyeau, D.; Ricolleau, C.; Mottet, C.; Oikawa, T.; Langlois, C.; Le Bouar, Y.;

Braidy, N.; Loiseau, A. Size and Shape Effects on the Order–disorder Phase Transition

in CoPt Nanoparticles. Nat. Mater.2009, 8, 940–946.

(90) Sun, S.; Murray, C. B.; Weller, D.; Folks, L.; Moser, A. Monodisperse FePt

Nanoparticles and Ferromagnetic FePt Nanocrystal Superlattices. Science (80-. ).2000,

287, 1989–1992.

(91) Kuttiyiel, K. A.; Sasaki, K.; Su, D.; Wu, L.; Zhu, Y.; Adzic, R. R. Gold-Promoted

Structurally Ordered Intermetallic Palladium Cobalt Nanoparticles for the Oxygen

Reduction Reaction. Nat. Commun.2014, 5, 1–8.

(92) Nemoshkalenko, V. V.; Chuistov, K. V.; Aleshin, V. G.; Senkevich, A. I. Changes in

108

Energy Structure of Cu3Au and CuAu3 Alloy Studied by the Method of X-Ray

Photoelectron Spectroscopy. J. Electron Spectros. Relat. Phenomena1976, 9, 169–173.

(93) Tee, S. Y.; Ye, E.; Pan, P. H.; Lee, C. J. J.; Hui, H. K.; Zhang, S.-Y.; Koh, L. D.; Dong,

Z.; Han, M.-Y. Fabrication of Bimetallic Cu/Au Nanotubes and Their Sensitive,

Selective, Reproducible and Reusable Electrochemical Sensing of Glucose.

Nanoscale2015, 7, 11190–11198.

(94) Chen, L. Y.; Fujita, T.; Ding, Y.; Chen, M. W. A Three-Dimensional Gold-Decorated

Nanoporous Copper Core-Shell Composite for Electrocatalysis and Nonenzymatic

Biosensing. Adv. Funct. Mater.2010, 20, 2279–2285.

(95) Kim, S.-I.; Eom, G.; Kang, M.; Kang, T.; Lee, H.; Hwang, A.; Yang, H.; Kim, B.

Composition-Selective Fabrication of Ordered Intermetallic Au–Cu Nanowires and

Their Application to Nano-Size Electrochemical Glucose Detection.

Nanotechnology2015, 26, 245702.

(96) Jiang, Z.; Zhang, Q.; Zong, C.; Liu, B.-J.; Ren, B.; Xie, Z.; Zheng, L. Cu–Au Alloy

Nanotubes with Five-Fold Twinned Structure and Their Application in Surface-

Enhanced Raman Scattering. J. Mater. Chem.2012, 22, 18192–18197.

(97) He, R.; Wang, Y.; Wang, X.; Wang, Z.; Liu, G.; Zhou, W.; Wen, L.; Li, Q.; Wang, X.;

Chen, X.; et al. Facile Synthesis of Pentacle Gold–copper Alloy Nanocrystals and

Their Plasmonic and Catalytic Properties. Nat. Commun.2014, 5, 1–10.

(98) Najafishirtari, S.; Brescia, R.; Guardia, P.; Marras, S.; Manna, L.; Colombo, M.

Nanoscale Transformations of Alumina-Supported AuCu Ordered Phase Nanocrystals

and Their Activity in CO Oxidation. ACS Catal.2015, 5, 2154–2163.

(99) Zhang, L.; Niu, W.; Xu, G. Synthesis and Applications of Noble Metal Nanocrystals

with High-Energy Facets. Nano Today2012, 7, 586–605.

(100) Au, L.; Lu, X.; Xia, Y. A Comparative Study of Galvanic Replacement Reactions

Involving Ag Nanocubes and AuCl(2) or AuCl(4). Adv. Mater.2008, 20, 2517–2522.

(101) Moon, G. D.; Ko, S.; Min, Y.; Zeng, J.; Xia, Y.; Jeong, U. Chemical Transformations

of Nanostructured Materials. Nano Today2011, 6, 186–203.

(102) Xia, Y.; Xiong, Y.; Lim, B.; Skrabalak, S. E. Shape-Controlled Synthesis of Metal

Nanocrystals: Simple Chemistry Meets Complex Physics? Angew. Chem. Int. Ed.

Engl.2009, 48, 60–103.

109

(103) Liu, S.; Sun, Z.; Liu, Q.; Wu, L.; Huang, Y.; Yao, T.; Zhang, J.; Hu, T.; Ge, M.; Hu, F.;

et al. Unidirectional Thermal Diff Usion in Bimetalic Cu@Au Nanoparticles. ACS

Nano2014, 8, 1886–1892.

(104) Chen, W.; Yu, R.; Li, L.; Wang, A.; Peng, Q.; Li, Y. A Seed-Based Diffusion Route to

Monodisperse Intermetallic CuAu Nanocrystals. Angew. Chemie2010, 122, 2979–

2983.

(105) Bansal, A.; Sekhon, J. S.; Verma, S. S. Scattering Efficiency and LSPR Tunability of

Bimetallic Ag, Au, and Cu Nanoparticles. Plasmonics2013, 9, 143–150.

(106) Tanuma, S.; Powell, C. J.; Penn, D. R. Calculations of Electron Inelastic Mean Free

Paths. Surf. interface Anal.2000, 21, 165–176.

(107) Schaak, R. E.; Sra, A. K.; Leonard, B. M.; Cable, R. E.; Bauer, J. C.; Han, Y.; Means,

J.; Teizer, W.; Vasquez, Y.; Funck, E. S. Metallurgy in a Beaker : Nanoparticle Toolkit

for the Rapid Low-Temperature Solution Synthesis of Functional Multimetallic Solid-

State Materials. J. Am. Chem. Soc.2005, 3506–3515.

(108) Sinha, S. K.; Srivastava, C.; Sampath, S.; Chattopadhyay, K. Tunability of

Monodispersed Intermetallic AuCu Nanoparticles through Understanding of Reaction

Pathways. RSC Adv.2015, 5, 4389–4395.

(109) Destro, P.; Cantane, D.; Honorio, G.; Costa, L.; Meira, D.; Zanchet, D. Solid-State

Diffusion Mechanism at Nanoscale : A Strategy to Synthesizing Bimetallic

Nanoparticles. Under Prep.

(110) Gauthard, F.; Epron, F.; Barbier, J. Palladium and Platinum-Based Catalysts in the

Catalytic Reduction of Nitrate in Water: Effect of Copper, Silver, or Gold Addition. J.

Catal.2003, 220, 182–191.

(111) Shore, M. S.; Wang, J.; Johnston-Peck, A. C.; Oldenburg, A. L.; Tracy, J. B. Synthesis

of Au(core)/Ag(shell) Nanoparticles and Their Conversion to AuAg Alloy

Nanoparticles. Small2011, 7, 230–234.

(112) Smetana, A. B.; Klabunde, K. J.; Sorensen, C. M.; Ponce, A. A.; Mwale, B. Low-

Temperature Metallic Alloying of Copper and Silver Nanoparticles with Gold

Nanoparticles through Digestive Ripening. J. Phys. Chem. B2006, 110, 2155–2158.

(113) Bhaskar, S. P.; Vijayan, M.; Jagirdar, B. R. Size Modulation of Colloidal Au

Nanoparticles via Digestive Ripening in Conjunction with a Solvated Metal Atom

110

Dispersion Method: An Insight into Mechanism. J. Phys. Chem. C2014, 118, 18214–

18225.

(114) Kalidindi, S. B.; Jagirdar, B. R. Synthesis of Cu@ZnO Core-Shell Nanocomposite

through Digestive Ripening of Cu and Zn Nanoparticles. J. Phys. Chem. C2008, 112,

4042–4048.

(115) Jose, D.; Jagirdar, B. R. Au@Pd Core-Shell Nanoparticles through Digestive Ripening.

J. Phys. Chem. C2008, 112, 10089–10094.

(116) Xu, Z.; Lai, E.; Shao-Horn, Y.; Hamad-Schifferli, K. Compositional Dependence of the

Stability of AuCu Alloy Nanoparticles. Chem. Commun. (Camb).2012, 48, 5626–5628.

(117) Destro, P.; Colombo, M.; Prato, M.; Brescia, R.; Manna, L.; Zanchet, D. Au1-xCux

Colloidal Nanoparticles Synthesized via One-Pot Approach: Understanding the

Temperature Effect on the Au:Cu Ratio. RSC Adv.2016, 6, 22213–22221.

(118) Cable, R. E.; Schaak, R. E.; September, R. V; Re, V.; Recei, M.; October, V. Low-

Temperature Solution Synthesis of Nanocrystalline Binary Intermetallic Compounds

Using the Polyol Process. 2005, 6835–6841.

(119) Sra, A. K.; Ewers, T. D.; Schaak, R. E. Direct Solution Synthesis of Intermetallic

AuCu and AuCu Nanocrystals and Nanowire Networks Direct Solution Synthesis of

Intermetallic AuCu and AuCu3 Nanocrystals and Nanowire Networks. Chem.

Mater.2005, 17, 758–766.

(120) Henkel, A.; Jakab, A.; Brunklaus, G.; Sönnichsen, C. Tuning Plasmonic Properties by

Alloying Copper into Gold Nanorods. J. Phys. Chem. C2009, 113, 2200–2204.

(121) Cable, Robert E Schaak, R. E. Reacting the Unreactive: A Toolbox of Low-

Temperature Solution-Mediated Reactions for the Facile Interconversion of

Nanocrystalline Intermetallic Compounds. J. Am. Chem. Soc.2006, 128, 9129–9136.

(122) Yin, J.; Shan, S.; Yang, L.; Mott, D.; Malis, O.; Petkov, V.; Cai, F.; Shan Ng, M.; Luo,

J.; Chen, B. H.; et al. Gold–Copper Nanoparticles: Nanostructural Evolution and

Bifunctional Catalytic Sites. Chem. Mater.2012, 24, 4662–4674.

(123) Frenkel, A. I.; Rodriguez, J. a.; Chen, J. G. Synchrotron Techniques for in Situ

Catalytic Studies: Capabilities, Challenges, and Opportunities. ACS Catalysis, 2012, 2,

2269–2280.

111

(124) Frenkel, A. I. Applications of Extended X-Ray Absorption Fine-Structure

Spectroscopy to Studies of Bimetallic Nanoparticle Catalysts. Chem. Soc. Rev.2012,

8163–8178.

(125) Krishna, K. S.; Navin, C. V; Biswas, S.; Singh, V.; Ham, K.; Bovenkamp, G. L.;

Theegala, C. S.; Miller, T.; Spivey, J. J.; Kumar, C. S. S. R. Milli Fluidics for Time-

Resolved Mapping of the Growth of Gold Nanostructures. J. Am. Chem. Soc.2013,

135, 5450–5456.

(126) Campi, G.; Mari, A.; Amenitsch, H.; Pifferi, A.; Cannas, C.; Suber, L. Monitoring

Early Stages of Silver Particle Formation in a Polymer Solution by in Situ and Time

Resolved Small Angle X-Ray Scattering. Nanoscale2010, 2, 2447–2455.

(127) Sarma, L. S.; Chen, C. H.; Kumar, S. M. S.; Wang, G. R.; Yen, S. C.; Liu, D. G.; Sheu,

H. S.; Yu, K. L.; Tang, M. T.; Lee, J. F.; et al. Formation of Pt-Ru Nanoparticles in

Ethylene Glycol Solution: An in Situ X-Ray Absorption Spectroscopy Study.

Langmuir2007, 23, 5802–5809.

(128) Tsai, Y. W.; Tseng, Y. L.; Sarma, L. S.; Liu, D. G.; Lee, J. F.; Hwang, B. J. Genesis of

Pt Clusters in Reverse Micelles Investigated by in Situ X-Ray Absorption

Spectroscopy. J. Phys. Chem. B2004, 108, 8148–8152.

(129) Chen, C. H.; Hwang, B. J.; Wang, G. R.; Sarma, L. S.; Tang, M. T.; Liu, D. G.; Lee, J.

F. Nucleation and Growth Mechanism of Pd/Pt Bimetallic Clusters in Sodium bis(2-

Ethylhexyl)sulfosuceinate (AOT) Reverse Micelles as Studied by in Situ X-Ray

Absorption Spectroscopy. J. Phys. Chem. B2005, 109, 21566–21575.

(130) Oyanagi, H.; Sun, Z. H.; Jiang, Y.; Uehara, M.; Nakamura, H.; Yamashita, K.; Zhang,

L.; Lee, C.; Fukano, A.; Maeda, H. In Situ XAFS Experiments Using a Microfluidic

Cell: Application to Initial Growth of CdSe Nanocrystals. J. Synchrotron Radiat.2011,

18, 272–279.

(131) Kuhn, M.; Sham, T. K. Charge Redistribution and Electronic Behavior in a Series of

Au-Cu Alloys. Phys. Rev. B1994, 49, 1647–1661.

(132) Borgohain, K.; Murase, N.; Mahamuni, S. Synthesis and Properties of Cu 2O Quantum

Particles. J. Appl. Phys.2002, 92, 1292–1297.

(133) Chang, Y.; Teo, J. J.; Zeng, H. C. Formation of Colloidal CuO Nanocrystallites and

Their Spherical Aggregation and Reductive Transformation to Hollow Cu 2O

112

Nanospheres. Langmuir2005, 21, 1074–1079.

(134) Somorjai, G. a.; Park, J. Y. Colloid Science of Metal Nanoparticle Catalysts in 2D and

3D Structures. Challenges of Nucleation, Growth, Composition, Particle Shape, Size

Control and Their Influence on Activity and Selectivity. Top. Catal.2008, 49, 126–135.

(135) Narayanan, R.; El-Sayed, M. A. Catalysis with Transition Metal Nanoparticles in

Colloidal Solution: Nanoparticle Shape Dependence and Stability. J. Phys. Chem.

B2005, 109, 12663–12676.

(136) Narayanan, R.; El-Sayed, M. A. Shape-Dependent Catalytic Activity of Platinum

Nanoparticles in Colloidal Solution. Nano Lett.2004, 4, 1343–1348.

(137) Prieto, P. J. S.; Ferreira, A. P.; Haddad, P. S.; Zanchet, D.; Bueno, J. M. C. Designing

Pt Nanoparticles Supported on CeO2-Al 2O3: Synthesis, Characterization and

Catalytic Properties in the Steam Reforming and Partial Oxidation of Methane. J.

Catal.2010, 276, 351–359.

(138) Ribeiro, R. U.; Liberatori, J. W. C.; Winnishofer, H.; Bueno, M. C.; Zanchet, D.

Colloidal Co Nanoparticles Supported on SiO 2 : Synthesis , Characterization and

Catalytic Properties for Steam Reforming of Ethanol. Appl. Catal. B Environ.2009, 91,

670–678.

(139) Mourdikoudis, S.; Liz-Marzán, L. M. Oleylamine in Nanoparticle Synthesis. Chem.

Mater.2013, 25, 1465–1476.

(140) Na, K.; Zhang, Q.; Somorjai, G. A. Colloidal Metal Nanocatalysts: Synthesis,

Characterization, and Catalytic Applications. J. Clust. Sci.2014, 25, 83–114.

(141) Li, D.; Wang, C.; Tripkovic, D.; Sun, S.; Markovic, N. M.; Stamenkovic, V. R.

Surfactant Removal for Colloidal Nanoparticles from Solution Synthesis: The Effect on

Catalytic Performance. ACS Catal.2012, 2, 1358–1362.

(142) Blavo, S. O.; Qayyum, E.; Baldyga, L. M.; Castillo, V. A.; Sanchez, M. D.;

Warrington, K.; Barakat, M. A.; Kuhn, J. N. Verification of Organic Capping Agent

Removal from Supported Colloidal Synthesized Pt Nanoparticle Catalysts. Top.

Catal.2013, 56, 1835–1842.

(143) Huang, W.; Hua, Q.; Cao, T. Influence and Removal of Capping Ligands on Catalytic

Colloidal Nanoparticles. Catal. Letters2014, 144, 1355–1369.

113

(144) Aliaga, C.; Park, J. Y.; Yamada, Y.; Lee, H. S.; Tsung, C. K.; Yang, P.; Somorjai, G.

A. Sum Frequency Generation and Catalytic Reaction Studies of the Removal of

Organic Capping Agents from Pt Nanoparticles by UV-Ozone Treatment. J. Phys.

Chem. C2009, 113, 6150–6155.

(145) Crespo-Quesada, M.; Andanson, J.-M.; Yarulin, A.; Lim, B.; Xia, Y.; Kiwi-Minsker, L.

UV-Ozone Cleaning of Supported Poly(vinylpyrrolidone)-Stabilized Palladium

Nanocubes: Effect of Stabilizer Removal on Morphology and Catalytic Behavior.

Langmuir2011, 27, 7909–7916.

(146) Elliott, E. W.; Glover, R. D.; Hutchison, J. E. Removal of Thiol Ligands from Surface-

Confined Nanoparticles without Particle Growth or Desorption. ACS Nano2015, 9,

3050–3059.

(147) Sinha, a K.; Seelan, S.; Tsubota, S.; Haruta, M. Catalysis by Gold Nanoparticles:

Epoxidation of Propene. Top. Catal.2004, 29, 95–102.

(148) Borodko, Y.; Lee, H. S.; Joo, S. H.; Zhang, Y.; Somorjai, G. Spectroscopic Study of

the Thermal Degradation of PVP-Capped Rh and Pt Nanoparticles in H 2 and O 2

Environments. 2010, 1117–1126.

(149) Zhao, Y.; Jia, L.; Medrano, J. a; Ross, J. R. H.; Le, L. Supported Pd Catalysts Prepared

via Colloidal Method: The Effect of Acids. ACS Catal. 2013, 3, 2341–2352.

(150) Cargnello, M.; Chen, C.; Diroll, B. T.; Doan-Nguyen, V. V. T.; Gorte, R. J.; Murray,

C. B. Efficient Removal of Organic Ligands from Supported Nanocrystals by Fast

Thermal Annealing Enables Catalytic Studies on Well-Defined Active Phases. J. Am.

Chem. Soc. 2015, 137, 6906–6911.

(151) Niu, Z.; Li, Y. Removal and Utilization of Capping Agents in Nanocatalysis. Chem.

Mater. 2014, 26, 72–83.

(152) Lopez-sanchez, J. A.; Dimitratos, N.; Hammond, C.; Brett, G. L.; Kesavan, L.; White,

S.; Miedziak, P.; Tiruvalam, R.; Robert, L.; Carley, A. F.; et al. Facile Removal of

Stabiliser-Ligands from Supported Gold Nanoparticles. 2011, 27–59.

(153) Bartholomew, C. H. Mechanisms of Catalyst Deactivation. Appl. Catal. A Gen. 2001,

212, 17–60.

(154) Gracia, F. J.; Miller, J. T.; Kropf, a J.; Wolf, E. E. Kinetics, FTIR, and Controlled

Atmosphere EXAFS Study of the Effect of Chlorine on Pt-Supported Catalysts during

114

Oxidation Reactions. J. Catal. 2002, 209, 341–354.

(155) Oh, H. S.; Yang, J. H.; Costello, C. K.; Wang, Y. M.; Bare, S. R.; Kung, H. H.; Kung,

M. C. Selective Catalytic Oxidation of CO: Effect of Chloride on Supported Au

Catalysts. J. Catal. 2002, 210, 375–386.

(156) Liu, X.; Wang, A.; Li, L.; Zhang, T.; Mou, C.-Y.; Lee, J.-F. Structural Changes of Au–

Cu Bimetallic Catalysts in CO Oxidation: In Situ XRD, EPR, XANES, and FT-IR

Characterizations. J. Catal. 2011, 278, 288–296.

(157) Broqvist, P.; Molina, L. M.; Grönbeck, H.; Hammer, B. Promoting and Poisoning

Effects of Na and Cl Coadsorption on CO Oxidation over MgO-Supported Au

Nanoparticles. J. Catal. 2004, 227, 217–226.

(158) Cant, N. W.; Angove, D. E.; J. Patterson, M. The Effects of Residual Chlorine on the

Behaviour of Platinum Group Metals for Oxidation of Different Hydrocarbons. Catal.

Today 1998, 44, 93–99.

(159) Ponec, V. Alloy Catalysts: The Concepts. Appl. Catal. A Gen.2001, 222, 31–45.

(160) Bertolini, J. Model Catalysis by Metals and Alloys : From Single-Crystal Surfaces to

Well-Defined Nano-Particles. 2008, 138, 84–96.

(161) Llorca, J.; Ledesma, C.; Chimentão, R. J.; Medina, F.; Sueiras, J.; Angurell, I.; Seco,

M.; Rossell, O. Propene Epoxidation over TiO 2 -Supported Au – Cu Alloy Catalysts

Prepared from Thiol-Capped Nanoparticles. 2008, 258, 187–198.

(162) Abad, A.; Almela, C.; Corma, A.; García, H. Efficient Chemoselective Alcohol

Oxidation Using Oxygen as Oxidant. Superior Performance of Gold over Palladium

Catalysts. Tetrahedron 2006, 62, 6666–6672.

(163) Prieto, P. J. S.; Ferreira, a. P.; Haddad, P. S.; Zanchet, D.; Bueno, J. M. C. Designing

Pt Nanoparticles Supported on CeO2–Al2O3: Synthesis, Characterization and Catalytic

Properties in the Steam Reforming and Partial Oxidation of Methane. J. Catal.2010,

276, 351–359.

(164) Bauer, J. C.; Mullins, D. R.; Oyola, Y.; Overbury, S. H.; Dai, S. Structure Activity

Relationships of Silica Supported AuCu and AuCuPd Alloy Catalysts for the Oxidation

of CO. Catal. Letters2013.

(165) Comotti, M.; Li, W.-C.; Spliethoff, B.; Schüth, F. Support Effect in High Activity Gold

115

Catalysts for CO Oxidation. J. Am. Chem. Soc.2005, 128, 917–924.

(166) Li, X.; Chen, Q.; McCue, I.; Snyder, J.; Crozier, P.; Erlebacher, J.; Sieradzki, K.

Dealloying of Noble-Metal Alloy Nanoparticles. Nano Lett.2014, 14, 2569–2577.

(167) Baldizzone, C.; Gan, L.; Hodnik, N.; Keeley, G. P.; Kostka, A.; Heggen, M.; Strasser,

P.; Mayrhofer, K. J. J. Stability of Dealloyed Porous Pt/Ni Nanoparticles. ACS

Catal.2015, 5000–5007.

(168) Destro, P.; Marras, S.; Manna, L.; Colombo, M.; Zanchet, D. AuCu Alloy

Nanoparticles Supported on SiO2: Impact of Redox Pretreatments in the Catalyst

Performance in CO Oxidation. Catal. Today, 2016,

http://dx.doi.org/10.1016/j.cattod.2016.08.003.

116

Appendix

Appendix A

Supporting Information of Chapter 2

8 10 12 14 16 18 20

0

20

40

60

80

100

Num

ber o

f par

ticle

sSize (nm)

100°C14.0 1.0 nm

(a)

8 10 12 14 16 18 200

20

40

60

80

100

120

Num

ber o

f par

ticle

s

Size (nm)

120°C_I12.0 1.0

(b)

117

8 10 12 14 16 18 200

20406080

100120140160180

13.0 1.0 nm

Num

ber o

f par

ticle

s

Size (nm)

120 °C_F

(c)

8 10 12 14 16 18 200

5

10

15

20

25

30

Num

ber o

f par

ticle

s

Size (nm)

225°C_I13.01.0 nm

(d)

8 10 12 14 16 18 20

0

40

80

120

160

Num

ber o

f par

itcle

s

Size (nm)

225 °C_F14.01.0 nm

(e)

118

8 10 12 14 16 18 20

0

10

20

30

40

50

60

70

Num

ber o

f par

ticle

s

Size (nm)

260°C_I13.01.0 nm

(f)

8 10 12 14 16 18 200

50

100

150

200

Num

ber o

f par

ticle

s

Size (nm)

260°C_F13.01.0 nm

(g)

8 10 12 14 16 18 200

20406080

100120

140160

Num

ber o

f par

ticle

s

Size (nm)

280°C_I14.01.0 nm

(h)

119

8 10 12 14 16 18 20

0

20

40

60

80

100

Num

ber o

f par

ticle

s

Size (nm)

280°C_F14.01.0 nm

(i)

A1: TEM images during the synthesis at the points (a) 100°C, (b) 120°C_I, (c) 120°C_F, (d)

225°C_I, (e) 225_F, (f) 260_I, (g) 260_F, (h) 280_I and (i) 280_F

120

(a)

(b)

A2:HAADF-STEM imaging and STEM-EDS elemental analysis for the two samples obtained

at (a) 280_I and (b) 280_F

121

Appendix B

Supporting information of Chapter 3

B.1 Reactor design

A schematic representation of the home-made reactor for in situ XAFS measurements is

shown at Figure B1a. The reactor consist a round bottom flask with outer joint (1) and two

“arms” (2), all made in glass. On the two arms are two pistons (3) made in Teflon®, which are

fixed to the reactor by two devices (4). The X-rays cross the sample through the two pistons.

On the top of the piston there is a Kapton® window (5), an o-ring (6) and a piece to keep

everything together (7) even under harsh reaction conditions (250 0C and organic solvents).

The pistons allow adjusting the optical path, from about 0.3 mm to 10 mm according to the

sample concentration. Stirring is carried out by a stir plate. The heating is made by an

aluminum heating block (8) that matches perfectly to the reactor. Two cartridge resistances

(9) are coupled to the temperature controller. The temperature is monitored by a

thermocouple. The temperature can reach up to 250 0C without damaging the Kapton®

window. It the heating block is fixed to the beamline bench, facilitating the alignment of the

reactor. The reactor is coupled to a reflux system and connected to a nitrogen trap and to the

exhaustion system. Reaction can be conduct under inert gas by bubbling gas through the

reflux system. Figure B1b shows a picture of the set-up.

122

Figure B1: (a) Schematic representation of the reactor (see text for details), (b). Photo of the

reactor (1- reaction vessel, 2 – Teflon pistons, 3 - reaction solution, 4 - X ray beam). The read

arrows show the X ray beam pathway.

Table B1.Size evolution of the NPs evaluated by TEM, by taking aliquots at a given

temperature.

Temperature (°C) Number of particles Diameter(nm)

70 768 11.2±1.2

90 1470 14.8±1.2

120_initial 1611 12.8±1.1

120_final 479 13.2±1.2

160 110 14.5±0.9

200 150 13.6±0.9

225_initial 476 15.3±1.3

225_final 110 13.8±1.8

After purification 492 14.3±1.3

123

Figure B2.(a) STEM image and (b) elemental line-scan EDS-STEM for abimetallicNP.

Figure B3.Ex situ XANES spectra of standards at (a) Au L3-edge (HAuCl4 and Au foil) and

(b) Cu K-edge (Cu foil,Cu(acac)2, and Cu2O).

8950 8975 9000 9025 9050

Nor

mal

ized

abs

orpt

ion

Energy (eV)

bare Cu foils Cu(acac)2

Cu2O

(b)

124

Figure B4:.XANES spectra obtained at the Au L3-edge (a) and Cu K-edge (b). Panel (a)

AuCu NPs dispersionat 225 ºC and Au foil and panel (b) AuCu NPs dispersionat 225 ºC

andstandards (Cu foil,Cu(acac)2, and Cu2O).

125

Figure B5.Representative TEM images obtained by taking aliquots at 70, 90, 120, 160, 200,

and 225 ºC during the AuCu NPs synthesis.

126

Appendix C

Supporting information of Chapter 5

TableC1: %Cu obtained by ICP for theAu1-xCux nanoparticles studied in this work (colloidal

and supported) and total metal loading for the supported catalysts, .

Sample name molar Cu% wt % metal load

(Au+Cu)

Au0.75Cu0.25 25.57±0.08 --

Au0.60Cu0.40 41.35±0.04 --

Au0.50Cu0.50 48.79±0.07 --

Au0.75Cu0.25/Al2O3 23.08±0.09 1.87±0.03

Au0.60Cu0.40/Al2O3 40.16±0.22 2.01±0.19

Au0.50Cu0.50/Al2O3 45.47±0.25 1.72±0.11

Au0.75Cu0.25/SiO2 25.2 ± 0.5 1.87 ± 0.07

Au0.60Cu0.40/SiO2 34.2 ± 0.2 1.92 ± 0.01

Au0.50Cu0.50/SiO2 44.81 ± 0.2 1.75 ± 0.09

127

(a) (b) (c)

(d)

(e)

(f)

Figure C1. TEM imagesof (a) Au0.60Cu0.40 (b) Au0.60Cu0.40/Al2O3, (c) Au0.60Cu0.40/ SiO2 (d)

Au0.50Cu0.50 (e) Au0.50Cu0.50/Al2O3and (f) Au0.50Cu0.50/SiO2.

0 50 100 150 200 250 3000

102030405060708090

100 Run #2 Run #6

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.75Cu0.25/Al2O3

(a)

0 50 100 150 200 250 3000

102030405060708090

100 Run #2 Run #6 Run #10

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.60Cu0.40/Al2O3

(b)

0 50 100 150 200 250 3000

102030405060708090

100 Run #2 Run #6 Run #10

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.50Cu0.50/Al2O3

(c)

0 50 100 150 200 250 3000

102030405060708090

100 Run #4 Run #8

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.75Cu0.25/Al2O3

(d)

0 50 100 150 200 250 3000

102030405060708090

100 Run #4 Run #8 Run #12

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.60Cu0.40/Al2O3

(e)

0 50 100 150 200 250 3000

102030405060708090

100 Run #4 Run #8 Run #12

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.50Cu0.50/Al2O3

(f)

Figure C2:Catalytic activity in CO oxidation(Runs #2, #6 and #10) after oxidative

pretreatments for (a) Au0.75Cu0.25/Al2O3, (b) Au0.60Cu0.40/Al2O3 and (c) Au0.50Cu0.50/Al2O3, and

after reductive pretreatments (Runs #4, #8 and #12) for (d) Au0.75Cu0.25/Al2O3, (e)

Au0.60Cu0.40/Al2O3 and (f) Au0.50Cu0.50/Al2O3.

128

0 50 100 150 200 250 3000

20

40

60

80

100 Run #2 Run #6 Run #10

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.60Cu0.40/SiO2

(a)

0 50 100 150 200 250 3000

20

40

60

80

100 Run #2 Run #6 Run #10

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.60Cu0.40 /SiO2

(b)

0 50 100 150 200 250 3000

20

40

60

80

100

Au0.50Cu0.50/SiO2

Run #2 Run #6 Run #10

CO

Con

vers

ion

(%)

Temperature (°C)

(c)

0 50 100 150 200 250 3000

20

40

60

80

100 Run #4 Run #8 Run #12

CO

Con

vers

ion

(%)

Temperature (°C)

Au0.75Cu0.25/SiO2

(d)

0 50 100 150 200 250 3000

20

40

60

80

100

Au0.60Cu0.40/SiO2

Run #4 Run #8 Run #12

CO

Con

vers

ion

(%)

Temperatura (°C)

(e)

0 50 100 150 200 250 3000

20

40

60

80

100

Au0.50Cu0.50/SiO2

Run #4 Run #8 Run #12

CO

Con

vers

ion

(%)

Temperature (°C)

(f)

Figure C3:Catalytic activity in CO oxidation(Runs #2, #6 and #10) after oxidative

pretreatments for (a) Au0.75Cu0.25/SiO2, (b) Au0.60Cu0.40/SiO2 e (c) Au0.50Cu0.50/SiO2, and after

reductive pretreatments (Runs #4, #8 and #12) for (d) Au0.75Cu0.25/SiO2, (e) Au0.60Cu0.40/SiO2 e

(f) Au0.50Cu0.50/SiO2. In the case of Runs #2 of SiO2 supported samples the lower CO

conversion can be related to the presence of chloride species on the support from the colloidal

samples. Details about this trend were discussed by some of us on Destroet. al. (CatTodpaper)

129

30 35 40 45 50 55 60

0

20

40

60

80

100

Inte

nsity

(Arb

. Uni

ts)

2 (°)

Signal difference

Al2O3

Au1-xCux/Al2O3

Au1-xCux/Al2O3

(a)

30 35 40 45 50 55 60

20

40

60

80

Inte

nsity

(Arb

. uni

ts)

2

Au1-xCux/SiO2

Au1-x

Cux/SiO

2

SiO2

Signal difference

(b)

Figure C4: XRD patterns ofAu0.70Cu0.30/Al2O3, Al2O3 and the difference between

Au0.70Cu0.30/Al2O3 and Al2O3 (b)XRD patterns ofAu0.70Cu0.30/SiO2, SiO2 and the difference

between Au0.70Cu0.30/SiO2 and SiO2.

34 36 38 40 42

Inte

nsity

(Arb

. Uni

ts)

2 (°)

350°C

325°C

300°C

290°C

220°C

250°C

200°C150°C

100°C

25°C

(a)

34 36 38 40 422

Inte

nsity

(Arb

. Uni

ts.)

350°C

325°C

300°C

290°C

220°C

250°C

200°C

150°C100°C

25°C

(b)

130

50 100 150 200 250 300 3500

10

20

30

40%

Cu

Temperature (°C)

%Cu domain #1%Cu domain #2

(c)

50 100 150 200 250 300 3500

10

20

30

40 %Cu domain#1%Cu domain#2

% C

u

Temperature (°C)

(d)

Figure C5: XRD patterns ofAu0.70Cu0.30/Al2O3 in the range of 2θ=34-43° at specific

temperatures during the activation and best fits of patterns at 25, 100 and 150 °C using two

Gaussians. (b)XRD patterns ofAu0.70Cu0.30/SiO2 in the range of 2θ=34-43° at specific

temperatures during the activation and best fits of patterns at 25, 100 and 150 °C using two

Gaussians. (c) Quantification of Cu related to domain #1, domain #2 and the global

composition of Cu based on the relative areas.

34 36 38 40 42 44 46

10

20

30

40

Inte

nsity

(Arb

. Uni

ts)

2

Original signal Domain #1 Domain #2 Total area

131

Figure C6: Details about the compositions calculation. The %Cu related to the domain#1 and

domain #2 are obtained by lattice parameter values related to the Vegard´s Law. The areas

related to the domains #1 and #2 are obtained by the fitting of two Gaussians using the

software Origin 8.1.

34 36 38 40 42

Coo

ling

RT

100°C

200°C

300°C

250°C

225°C

200°C

150°C

100°C

Inte

nsity

(Arb

. Uni

ts)

2

RT

Hea

ting

AuCu/Al2O3

(a)

34 36 38 40 42

AuCu/SiO2

Coo

ling

RT

100°C

200°C

300°C

250°C

225°C

200°C

150°C100°C

Inte

nsity

(Arb

. Uni

ts)

2

RT

Hea

ting

(b)

Figure C7: XRD patterns of (a) Au0.60Cu0.60/Al2O3 and (b) Au0.60Cu0.40/SiO2 in in the range of

2θ=34-43° at specific temperatures during the CO Oxidation (Step#2)

8980 9000 9020 90400,0

0,2

0,4

0,6

0,8

1,0

1,2

1,4

1,6

Nor

mal

ized

abs

orpt

ion

Energy (eV)

RT 150 °C 250 °C 300 °C

Cu KCO_OX (Heating)

Au0.60Cu0.40/Al2O3

(a)

8980 9000 9020 90400,0

0,2

0,4

0,6

0,8

1,0

1,2

1,4

1,6

Nor

mal

ized

abs

orpt

ion

Energy (eV)

300 °C 150 °C 70°C RT

Cu KCO_OX (Cooling)

Au0.60Cu0.40/Al2O3

(b)

132

8980 9000 9020 90400,0

0,2

0,4

0,6

0,8

1,0

1,2

1,4

1,6

Nor

mal

ized

abs

orpt

ion

Energy (eV)

RT 150 °C 250 °C 300 °C

CO_OX (Heating)Cu K

Au0.60Cu0.40/SiO2

(c)

8980 9000 9020 9040

0,0

0,2

0,4

0,6

0,8

1,0

1,2

1,4

1,6

Nor

mal

ized

abs

orpt

ion

Energy (eV)

300 °C 150 °C 70 °C RT

Cu KCO_OX (Cooling)

Au0.60Cu0.40/SiO2

(d)

Figure C8: (a) XANES Cu-K of Au0.60Cu0.40/Al2O3 under CO + O2 atmosphere at Step#2

during (a) heating and (b) cooling at specific temperatures. (a) XANES Cu-K of

Au0.60Cu0.40/SiO2 under CO + O2 atmosphere at Step#2 during (a) heating and (b) cooling at

specific temperatures.

34 36 38 40 42

Inte

nsity

(Arb

. Uni

ts)

2

350°C

325°C

300°C

290°C

250°C

220°C

200°C

150°C

100°C

25°C

(a)

34 36 38 40 42

Inte

nsity

(Arb

Uni

ts)

2

350°C

325°C

300°C

290°C

250°C

220°C

200°C150°C

100°C

25°C

(b)

133

50 100 150 200 250 300 350

0

5

10

15

20

25

30

35

40

Temperature (°C)

Au0.60Cu0.40/Al2O3

%C

u

%Cu domain #1%Cu domain #2

(C)

0 50 100 150 200 250 300 350 4000

5

10

15

20

25

30

35

40

% C

u

Temperature (°C)

Au0.60Cu0.40/SiO2

%Cu

(D)

Figure C9: XRD patterns ofAu0.60Cu0.40/Al2O3in the range of 2θ=34-43° at specific

temperatures during the reduction and best fits of patterns at 200, 220, 290, 300, 325 and 350

°C using two Gaussians. (b)XRD patterns ofAu0.60Cu0.40/SiO2in the range of 2θ=34-43° at

specific temperatures during the reduction and best fits of patterns at 25, 100 and 150 °C

using one Gaussian.

134

Appendix D

Permission to use paper the paper:

“Au1-xCux colloidal nanoparticles synthesized via a one-pot approach: understanding the

temperature effect on the Au:Cu ratio”

Priscila Destro, Massimo Colombo, Mirko Prato, Rosaria Brescia, Liberato Manna and

Daniela Zanchet.

Dear Priscila

The Royal Society of Chemistry (RSC) hereby grants permission for the use of your paper(s) specified below in the printed and microfilm version of your thesis. You may also make available the PDF version of your paper(s) that the RSC sent to the corresponding author(s) of your paper(s) upon publication of the paper(s) in the following ways: in your thesis via any website that your university may have for the deposition of theses, via your university’s Intranet or via your own personal website. We are however unable to grant you permission to include the PDF version of the paper(s) on its own in your institutional repository. The Royal Society of Chemistry is a signatory to the STM Guidelines on Permissions (available on request).

Please note that if the material specified below or any part of it appears with credit or acknowledgement to a third party then you must also secure permission from that third party before reproducing that material. Please ensure that the thesis states the following:

Reproduced by permission of The Royal Society of Chemistry and include a link to the paper on the Royal Society of Chemistry’s website.

Please ensure that your co-authors are aware that you are including the paper in your thesis.

Regards

Gill Cockhead

Publishing Contracts & Copyright Executive

Gill Cockhead Publishing Contracts & Copyright Executive

Royal Society of Chemistry,

Thomas Graham House, Science Park, Milton Road, Cambridge, CB4 0WF,

UKTel +44(0)1223432134

135

Appendix E

Permission to use paper the paper:

AuCu alloy nanoparticles supported on SiO2: Impact of redox pretreatments in the catalyst

performance in CO oxidation

Priscila Destro, Sergio Marras, Liberato Manna, Massimo Colombo, Daniela Zanchet

This Agreement between Priscila Destro and Elsevier consists of your license details and the terms and conditions provided by Elsevier and Copyright Clearance Center.

License Number 3975910444492

License date Oct 20, 2016

Licensed Content Publisher Elsevier

Licensed Content Publication Catalysis Today

Licensed Content Title AuCu alloy nanoparticles supported on SiO2: Impact of redox pretreatments in the catalyst performance in CO oxidation

Licensed Content Author Priscila Destro, Sergio Marras,Liberato Manna, Massimo Colombo, Daniela Zanchet

Licensed Content Date Available online 6 August 2016

Licensed Content Volume Number n/a

Licensed Content Issue Number n/a

Licensed Content Pages 1

Start Page

End Page

Type of Use Reuse in a thesis/dissertation

Portion full article

Format print

Are you the author of this Elsevier article?

Yes

Will you be translating? No

Order reference number

Title of your thesis/dissertation Gold-copper colloidal nanoparticles: Insights about

136

the alloy formation and application in CO oxidation reaction

Expected completion date Dec 2016

Estimated size (number of pages) 150

Elsevier VAT number GB 494 6272 12

137

Appendix F

DECLARAÇÃO

As cópias dos documentos de minha autoria ou de minha co-autoria, já publicados ou

submetidos para publicação em revistas científicas ou anais de congressos sujeitos a

arbitragem, que constam da minha Dissertação/Tese de Mestrado/Doutorado, intitulada

”NANOPARTÍCULAS COLOIDAIS DE COBRE-OURO: CONSIDERAÇÕES SOBRE

A FORMAÇÃO DA LIGA E APLICAÇÃO NA REAÇÃO DE OXIDAÇÃO DE CO”

não infringem os dispositivos da Lei nº 9.610/98, nem o direito autoral de qualquer editora.

Campinas, 24 de outubro de 2016

Priscila Destro

R.G. nº 43.726.551-1

Daniela Zanchet (Orientadora)

RG nº 34.443.884-3