+ All Categories
Home > Documents > ELECTROCHEMICAL PROMOTION OF NOVEL CATALYSTS …

ELECTROCHEMICAL PROMOTION OF NOVEL CATALYSTS …

Date post: 26-Nov-2021
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
272
UNIVERSIDAD DE CASTILLA-LA MANCHA FACULTAD DE CIENCIAS Y TECNOLOGÍAS QUÍMICAS DEPARTAMENTO DE INGENIERÍA QUÍMICA ELECTROCHEMICAL PROMOTION OF NOVEL CATALYSTS WITH ALKALINE CONDUCTORS FOR HYDROGEN PRODUCTION FROM METHANOL Memoria que para optar al grado de Doctor en Ingeniería Química (Doctorado Internacional) presenta JESÚS GONZÁLEZ COBOS Directores: Dr. José Luis Valverde Palomino Dr. Antonio de Lucas Consuegra Ciudad Real, 2015
Transcript

UNIVERSIDAD DE CASTILLA-LA MANCHA

FACULTAD DE CIENCIAS Y TECNOLOGÍAS QUÍMICAS

DEPARTAMENTO DE INGENIERÍA QUÍMICA

ELECTROCHEMICAL PROMOTION OF NOVEL

CATALYSTS WITH ALKALINE CONDUCTORS FOR

HYDROGEN PRODUCTION FROM METHANOL

Memoria que para optar al grado de Doctor en Ingeniería Química

(Doctorado Internacional) presenta

JESÚS GONZÁLEZ COBOS

Directores:

Dr. José Luis Valverde Palomino

Dr. Antonio de Lucas Consuegra

Ciudad Real, 2015

Nomenclature

v

Acronyms

CAD Cathodic arc deposition

CE Counter electrode

CV Cyclic voltammetry

DLC Diamond-like carbon

EBE Electron beam evaporation

EDX Energy-dispersive X-ray spectroscopy

EELS Electron energy loss spectroscopy

EPOC Electrochemical promotion of catalysis

FCC Face-centered cubic structure

FTIR Fourier transform infrared spectroscopy

GC Gas chromatograph

GLAD Glancing angle deposition

GO Graphene oxide

HRTEM High-resolution transmission electron microscopy

JCPDS Joint committee on powder diffraction standards

LSV Linear sweep voltammetry

MD Methanol decomposition

MEPR Monolithic electrochemically promoted reactor

NASICON Sodium super ionic conductor

NEMCA Non faradaic electrochemical modification of catalytic activity

OAD Oblique angle deposition

OC Open circuit

PEMFC Proton exchange membrane fuel cell

POM Partial oxidation of methanol

PVD Physical vapor deposition

RE Reference electrode

SCY Strontia-ceria-ytterbia

SEM Scanning electron microscopy

SOFC Solid oxide fuel cell

SRM Steam reforming of methanol

STEM Scanning transmission electron microscopy

TOF Turnover frequency

tpb three-phase boundaries

TPS Temperature-programmed stabilization

XPS X-ray photoelectron spectroscopy

XRD X-ray diffraction

XRF X-ray fluorescence

Nomenclature

vi

WE Working electrode

WGS Water-gas shift

YSZ Yttria-stabilized zirconia

Symbols

B Full width at half maximum (FWHM) of a XRD peak

D Metal dispersion

d Metal particle size

F Faraday constant (96485 C)

Fi Molar flow of i compound

I Current

i Current density

KW Half-width Scherrer constant

Mm Metal atomic weight

N Total surface area

NA Avogadro number (6.023 x 1023

)

NG Active surface area

n Charge of the ionic species (1 for K+)

Promotion index

r Catalytic reaction rate under promoted conditions

r0 Catalytic reaction rate under unpromoted conditions

re, rK+ Electrocatalytic reaction rate

t Time

VWR Catalyst potential

Catalyst potential under open circuit conditions

α In OAD, zenithal incident angle

β In OAD, zenithal tilt angle of deposited metal columns

γ Permanent enhancement ratio

θ Bragg angle

Promoter coverage

λ X-ray wavelength

ρ Rate enhancement ratio

ρm Metal density

σ Metal atomic surface

ϕ Work function

Λ Faradaic efficiency

Table of contents

vii

Descripción del trabajo realizado……………………………………………….. 1

A. INTRODUCCIÓN....................................................................................... 2

A.1. El hidrógeno como vector energético………………………………... 2

A.2. El metanol como fuente de hidrógeno……………………………….. 8

A.3. La promoción electroquímica de la catálisis (EPOC)……………….. 10

A.4. Nuevas tendencias y perspectivas de la promoción electroquímica…. 18

A.5. Objetivo de la tesis doctoral…………………………………………. 25

B. MÉTODOS E INSTALACIÓN EXPERIMENTAL……………………… 27

B.1. Preparación de los catalizadores electroquímicos…………………… 27

B.2. Técnicas de caracterización………………………………………….. 29

B.3. Instalación experimental……………………………………………... 31

C. RESULTADOS OBTENIDOS…………………………………………… 33

D. CONCLUSIONES Y RECOMENDACIONES…………………………... 44

E. BIBLIOGRAFÍA………………………………………………………….. 47

Abstract……………. ……………………………………………………………... 55

Chapter 1. Electrochemical promotion of Pt for H2 production from

methanol partial oxidation and steam reforming: A better performance of

pysical vapor deposited catalyst films……………………………………………... 63

1.1. INTRODUCTION………………………………………………………... 64

1.2. EXPERIMENTAL……………………………………………………….. 67

1.2.1. Preparation of the electrochemical catalysts……………………... 67

1.2.2. Characterization measurements…………………………………... 69

1.2.3. Catalytic activity measurements………………………………….. 70

1.3. RESULTS AND DISCUSSION…………………………………………. 71

1.3.1. Partial oxidation of methanol…………………………………….. 71

i) Influence of the preparation technique………………………..... 71

ii) Electrochemical promotion mechanism and parameters………. 79

iii) Influence of the reaction conditions……………………………. 83

iv) Stability study…………………………………………………… 87

1.3.2. Steam reforming of methanol…………………………………….. 90

1.4. CONCLUSIONS…………………………………………………………. 94

1.5. REFERENCES…………………………………………………………… 95

Chapter 2. Electrochemical promotion of Pt nanoparticles dispersed in a

carbon matrix for methanol conversion: Towards more competitive catalysts

of low metal loading.. ……………………………………………………………... 101

2.1. INTRODUCTION………………………………………………………... 102

2.2. EXPERIMENTAL………………………………………………………... 104

2.2.1. Preparation of the electrochemical catalyst………………………. 104

Table of contents

viii

2.2.2. Characterization measurements…………………………………... 105

2.2.3. Catalytic activity measurements………………………………….. 106

2.3. RESULTS AND DISCUSSION………………………………………... 106

2.3.1. Development of a catalyst with suitable structural and electrical

properties…………………………………………………………. 106

2.3.2. Electrochemical promotion experiments…………………………. 112

2.3.3. Comparison between Pt-DLC and Pt…………………………….. 120

2.4. CONCLUSIONS……………………………………………………….. 123

2.5. REFERENCES…………………………………………………………. 124

Chapter 3. Electrochemical activation of Au nanoparticles dispersed in

YSZ for methanol partial oxidation: Electrochemical promotion of a

non-conductive catalyst film..……………………………………………………... 131

3.1. INTRODUCTION………………………………………………………... 132

3.2. EXPERIMENTAL……………………………………………………….. 134

3.2.1. Preparation of the electrochemical catalyst………………………. 134

3.2.2. Characterization measurements…………………………………... 135

3.2.3. Catalytic activity measurements………………………………….. 136

3.3. RESULTS AND DISCUSSION…………………………………………. 136

3.3.1. Characterization of the catalyst film and blank experiments…….. 136

3.3.2. Electrochemical promotion via galvanostatic transitions………… 143

3.3.3. Electrochemical promotion via potentiostatic transitions………... 152

3.4. CONCLUSIONS…………………………………………………………. 156

3.5. REFERENCES…………………………………………………………… 157

Chapter 4. Electrochemical promotion of Cu nanocolumns in the partial

oxidation of methanol: EPOC with highly porous non-noble metal catalysts…… 163

4.1. INTRODUCTION………………………………………………………... 164

4.2. EXPERIMENTAL……………………………………………………….. 166

4.2.1. Preparation of the electrochemical catalyst………………………. 166

4.2.2. Characterization measurements…………………………………... 167

4.2.3. Catalytic activity measurements………………………………….. 168

4.3. RESULTS AND DISCUSSION…………………………………………. 168

4.3.1. Preliminary characterization of the catalyst film………………… 168

4.3.2. Electrochemical promotion experiments…………………………. 172

4.3.3. Post-reaction characterization of the catalyst film……………….. 180

4.4. CONCLUSIONS…………………………………………………………. 185

4.5. REFERENCES…………………………………………………………… 185

Table of contents

ix

Chapter 5. Electrochemical promotion of Ni in methanol conversion

reactions: Different applications of EPOC on a single catalytic system………….. 191

5.1. INTRODUCTION………………………………………………………... 192

5.2. EXPERIMENTAL……………………………………………………….. 194

5.2.1. Preparation of the electrochemical catalyst………………………. 194

5.2.2. Characterization measurements…………………………………... 195

5.2.3. Catalytic activity measurements………………………………….. 195

5.3. RESULTS AND DISCUSSION…………………………………………. 196

5.3.1. Electrochemical activation of the catalyst………………………... 196

5.3.2. Effect of the electrochemical promotion on the catalyst

oxidation state……………………………………………………. 206

5.3.3. Control of the catalyst selectivity via electrochemical

promotion………………………………………………………… 210

5.4. CONCLUSIONS…………………………………………………………. 214

5.5. REFERENCES……………………………………………………………. 215

Chapter 6. Electrochemically assisted production and storage of hydrogen:

A novel contribution of alkaline electrochemical catalysts………………………... 221

6.1. INTRODUCTION………………………………………………………... 222

6.2. EXPERIMENTAL……………………………………………………….. 224

6.2.1. Preparation of the electrochemical catalyst………………………. 224

6.2.2. Characterization measurements…………………………………... 227

6.2.3. Catalytic activity measurements………………………………….. 228

6.3. RESULTS AND DISCUSSION………………………………………….. 228

6.3.1. Preliminary experiments of H2 production and storage………….. 228

6.3.2. Influence of the applied negative polarization and the reaction

atmosphere……………………………………………………….. 234

6.3.3. Investigation of possible surface compounds by catalyst

characterization…………………………………………………... 242

6.4. CONCLUSIONS………………………………………………………….. 246

6.5. REFERENCES……………………………………………………………. 247

Chapter 7. General conclusions and recommendations……………………….. 255

7.1. CONCLUSIONS…………………………………………………………. 256

7.2. RECOMMENDATIONS………………………………………………… 258

List of publications and conferences…………………………………………….. 261

Table of contents

x

1

DESCRIPCIÓN DEL TRABAJO REALIZADO

Este trabajo forma parte de un amplio programa de investigación sobre la

aplicación de sistemas electrocatalíticos en procesos de interés energético y

medioambiental que se está desarrollando durante los últimos años en el

Departamento de Ingeniería Química de la Universidad de Castilla-La Mancha

(UCLM).

En particular, esta Tesis Doctoral tiene como objetivo el estudio del fenómeno

de la promoción electroquímica de la catálisis en la producción de hidrógeno a partir

de metanol empleando conductores alcalinos. Este trabajo ha sido financiado por el

Ministerio de Ciencia e Innovación a través del proyecto del plan Nacional CTQ

2010-16179/PQ, y por el Ministerio de Economía y Competitividad a través del

proyecto CTQ 2013-45030-R.

Esta tesis doctoral se ha realizado en colaboración con el Instituto de Ciencia

de Materiales de Madrid (CSIC), el Institut Jean Lamour de Université de Lorraine

(Francia) y el Instituto de Ciencia de Materiales de Sevilla (CSIC).

Descripción del trabajo realizado

2

A. INTRODUCCIÓN

A.1. El hidrógeno como vector energético

El consumo masivo de combustibles fósiles como el petróleo, asociado al

fuerte desarrollo tecnológico de las últimas décadas, ha dado lugar a una alarma

permanente sobre el irremediable agotamiento de las reservas de estos recursos

naturales y el grave deterioro del medioambiente que está ocasionando. Por ello la

humanidad se enfrenta actualmente a un triple desafío: satisfacer las necesidades

energéticas de la población, buscar alternativas al agotamiento de los combustibles

fósiles, y abordar seriamente la amenaza que supone el sobrecalentamiento del planeta

debido a la emisión de gases de efecto invernadero como el CO2 [1].

En este sentido, el hidrógeno se está postulando como una de las alternativas

energéticas más prometedoras, llegándose a hablar incluso de una futura “economía

del hidrógeno” [2]. Uno de los principales motivos es su gran densidad energética

gravimétrica en comparación con el resto de combustibles (Tabla A.1).

Tabla A.1. Densidad energética de diferentes combustibles (adaptado de Dutta y col. [3]).

Combustible Densidad energética (kWh kg-1

)

Hidrógeno 33,3

Gas natural licuado 15,1

Propano 13,8

Gasolina de aviación 13,0

Gasolina de automoción 12,9

Gasoil de automoción 12,7

Etanol 8,2

Carbón 7,5

Metanol 5,5

Madera seca 4,5

Además, el H2 obtenido de forma renovable es considerado mundialmente

como un vector energético limpio, puesto que el único subproducto derivado de su

combustión con oxígeno es vapor de agua [4].

Introducción

3

A.1.1. Aplicaciones del hidrógeno

i) Materia prima en la industria química.

En torno a la mitad del hidrógeno consumido anualmente en el mundo (sobre 5

x 1010

kg H2 [4]) se emplea en la elaboración de fertilizantes basados en amoníaco por

medio del proceso Haber-Bosh (reacción A.1), que típicamente tiene lugar a elevadas

presiones (200 bar) y temperaturas (400-500 ºC) con catalizadores de Fe o Ru [5].

N2 + 3H2 → 2NH3 (A.1)

Otra gran parte del H2 se consume en la industria petroquímica,

particularmente en el refinado del petróleo (proceso de hidrocraqueo catalítico) y para

reducir la cantidad de aromáticos y sobre todo de azufre en la gasolina y el gasóleo

(hidrodesulfuración). El hidrógeno también se utiliza en la elaboración de agua

oxigenada y junto con CO (gas de síntesis) en la obtención de metanol [6] y otros

muchos combustibles (proceso Fischer-Tropsch) [7].

ii) Generación de energía.

Existen principalmente dos vías para la obtención de energía a partir del H2.

Por un lado, el empleo de motores de combustión interna o turbinas de gas permite

transformar la energía química del hidrógeno en energía térmica y esta en energía

mecánica. Por otro lado, el hidrógeno se puede oxidar electroquímicamente en una

celda de combustible, en cuyo caso la energía química se transforma directamente en

energía eléctrica, evitando la limitación del ciclo de Carnot. Por ejemplo, un coche que

necesitara unos 24 kg de gasolina para recorrer 400 km, consumiría en su lugar

aproximadamente 8 kg H2 de hidrógeno en un motor de combustión interna, o 4 kg H2

en el caso de un coche eléctrico con pila de combustible [4].

iii) Otras aplicaciones.

El hidrógeno también se emplea para refrigerar motores y generadores

eléctricos y como gas protector en el método de soldadura por arco eléctrico. Debido a

su baja densidad (0,0899 kg m-3

), en el pasado incluso se utilizaba como gas de

elevación en globos aerostáticos y dirigibles.

Descripción del trabajo realizado

4

A.1.2. Tecnologías de producción de hidrógeno

El hidrógeno es el elemento más abundante del planeta, pero menos del 1 % se

encuentra en forma de gas molecular (H2) y es necesario obtenerlo a partir de agua,

hidrocarburos u otra materia orgánica por medio de muy diversos procesos. A

continuación se mencionan algunos de los más importantes.

i) Obtención de H2 a partir de agua.

Existen básicamente tres maneras de provocar la ruptura de la molécula de

agua: termólisis (por acción de la temperatura), fotoelectrólisis (por acción de la luz

solar) y electrólisis (por acción de la electricidad). Este último proceso es de los más

estudiados y se suele llevar a cabo en electrolizadores alcalinos, electrolizadores de

membranas de intercambio de protones (tipo PEM) o de óxido sólido (tipo SOEC) [8,

9].

ii) Producción de H2 mediante procesos biológicos.

El H2 se puede producir a partir de la conversión de la biomasa mediante los

procesos de gasificación, pirolisis, o hidrólisis [10]. También se puede obtener a partir

de la fotólisis directa del agua empleando microalgas verdes o cianobacterias y

procesos fermentativos, entre otros [3, 10].

iii) Procesos de conversión de hidrocarburos y compuestos oxigenados.

Actualmente la principal vía de producción de hidrógeno a nivel industrial es

el reformado de metano con vapor de agua, si bien es cierto que el gas de síntesis (CO

+ H2) obtenido se suele utilizar en las refinerías para su consumo propio. Existen tres

procesos tradicionales de producción de H2 a partir de hidrocarburos:

Reformado con vapor de agua (“Steam Reforming”, SR):

CnHm + nH2O → nCO + (n + m/2)H2 (A.2)

Esta es una reacción altamente endotérmica que se suele llevar a cabo a

temperaturas superiores 500 ºC y 1-25 atm. Los catalizadores empleados se basan en

Ni, Cu, o metales nobles como Pt o Rh [11]. Los primeros son más baratos pero son

Introducción

5

más propensos a la desactivación debido a la sinterización térmica de las partículas

metálicas o a la deposición de carbón [12]. Para evitar estos efectos sobre los

catalizadores, estos se suelen combinar formando aleaciones como Ni-Cr o Ni-Pt, o

soportar sobre una gran variedad de óxidos metálicos como SiO2, Al2O3, CeO2 o ZrO2

[11, 13]. Otra forma de modificar estos catalizadores consiste en la adición de

promotores alcalinos o alcalino térreos [14, 15].

Oxidación parcial (“Partial Oxidation”, PO):

CnHm + n/2O2 → nCO + m/2H2 (A.3)

Esta reacción se puede llevar a cabo en ausencia de quemadores externos para

el mantenimiento de la temperatura e incluso sin catalizador (a 1300-1500 ºC), aunque

este se suele utilizar para disminuir la deposición de carbón y la temperatura de trabajo

[11]. Por otra parte, el control de esta es complicado debido a la exotermicidad de la

reacción y a la formación de puntos calientes [16], que puede dar lugar a la

sinterización localizada de la fase activa y por tanto a la desactivación del catalizador.

Se suelen utilizar catalizadores similares a los del proceso de SR.

Reformado autotérmico (“Autothermal Reforming”, AR):

CnHm + n/2H2O + n/4O2 → nCO + (n/2 + m/2)H2 (A.4)

Esta reacción consiste en una combinación de las dos anteriores, y es muy

interesante desde el punto de vista del aprovechamiento energético ya que utiliza la

exotermicidad de la reacción de oxidación parcial para aportar el calor necesario a la

reacción de reformado con vapor de agua [17].

En los últimos años también se ha intensificado el estudio de algunas variantes

de estos procesos como el reformado seco de metano (con CO2) [18]. Además, los

procesos mencionados anteriormente suelen estar acompañados de otros sistemas de

reacción donde se llevan a cabo la oxidación de CO (reacción A.5) y/o la reacción de

“Water-Gas Shift” (WGS, reacción A.6) con objeto de purificar el H2 obtenido.

CO + 1/2O2 → CO2 (A.5)

CO + H2O → CO2 + H2 (A.6)

Descripción del trabajo realizado

6

Ambas reacciones son exotérmicas y esta última se suele llevar a cabo en dos

etapas consecutivas: la primera de ellas con catalizadores de Fe3O4/Cr2O3 a elevada

temperatura (300-530 ºC), y la segunda con catalizadores de Cu/ZnO/Al2O3 a

temperaturas inferiores (200-250 ºC).

Las ventajas de utilizar hidrocarburos como el CH4 como materia prima para la

obtención de H2 son su gran disponibilidad, bajo coste y amplia red de distribución.

Sin embargo, son recursos no renovables y presentan los inconvenientes asociados a

su estado gaseoso. En este sentido destaca el empleo de compuestos oxigenados

líquidos en condiciones ambientales, como el metanol y otros alcoholes. Además, al

igual que el agua, estos se pueden utilizar directamente en las celdas de combustible

[13]. Por otro lado, se puede afirmar que los procesos de obtención de hidrógeno aquí

mencionados son respetuosos con el medio ambiente, siempre y cuando en los

procesos de electrólisis se obtenga la electricidad a partir de energías renovables como

la eólica o la solar, y en los procesos de conversión catalítica se utilicen materias

primas de origen renovable como los bioalcoholes (biometanol, bioglicerol) [3].

A.1.3. Tecnologías de almacenamiento de hidrógeno

Una de las principales desventajas del hidrógeno está asociada a la dificultad

de su transporte, manipulación y almacenamiento en condiciones seguras. Presenta un

elevado calor de combustión (142 kJ g-1

), un amplio rango de inflamabilidad (4 – 75

% en aire) y una temperatura de autoignición de 560 ºC. Sin embargo, se puede

considerar seguro en recintos abiertos debido a su enorme difusividad en el aire (0,61

cm2 s

-1) y a su carácter inocuo. Por otro lado, el desarrollo de sistemas eficientes de

almacenamiento de H2 resulta de capital importancia si se obtiene por medio de

energías renovables dadas las fluctuaciones de disponibilidad que estas suponen.

Además, este aspecto es de particular importancia debido a su naturaleza gaseosa,

principalmente para su aplicación en automoción. Por ejemplo, en el caso del coche

eléctrico comentado anteriormente, 4 kg de hidrógeno almacenados en condiciones

ambientales ocuparían un volumen de unos 45 m3 [4]. En este sentido, por ejemplo, el

Departamento de Energía de EEUU (DOE) ha establecido como objetivo para el año

2015 el desarrollo de sistemas con una capacidad de almacenamiento de H2 del 9 % en

Introducción

7

peso (incluyendo válvulas, sistemas de calentamiento, enfriamiento, aislamiento, etc.)

[19].

Los sistemas de almacenamiento de hidrógeno se pueden dividir

principalmente en tres categorías:

i) Sistemas de almacenamiento convencionales.

El H2 se suele comprimir a 200 bar en cilindros de acero convencional, o hasta

450 bar si se emplean materiales reforzados con fibra de carbono y recubrimientos

inertes especiales. Los principales inconvenientes de estos sistemas son la necesidad

de un sistema de control de la presión y el riesgo inherente al proceso de compresión.

Por otro lado, el hidrógeno licuado (por debajo de -250 ºC) presenta una mayor

densidad que el comprimido, pero requiere infraestructuras adicionales para

aprovechar las cantidades de hidrógeno que se pierden por evaporación [4, 20].

ii) Almacenamiento de hidrógeno por adsorción física.

El H2 se puede adsorber en forma molecular sobre la superficie de una gran

variedad de sólidos porosos: nanotubos y nanofibras de carbono, grafito, carbones

activados, estructuras organometálicas (MOFs), zeolitas, etc. [4, 20, 21]. Estos

sistemas presentan una rápida cinética de adsorción/desorción y una gran

reversibilidad. Sin embargo, las débiles interacciones entre el H2 y los adsorbentes dan

lugar a una baja capacidad de almacenamiento de H2 (1-5 % en peso de H2). Esta

depende del área superficial del adsorbente y de las condiciones de operación,

requiriendo elevadas presiones (hasta 100 bar) y temperaturas criogénicas (-196 ºC).

iii) Almacenamiento de hidrógeno por adsorción química.

El otro mecanismo principal de captura de H2 es su disociación y la formación

de determinados compuestos como nitruros e hidruros de metales o de aleaciones

metálicas [4, 20, 21]. Estos compuestos sólidos son muy fáciles de manipular y en

estos casos la máxima cantidad de hidrógeno admitida es superior a la obtenida por

adsorción física. Por ejemplo, el MgH2 contiene idealmente un 7,6 % H2 en peso, y

otros hidruros interrmetálicos como Li3Be2H7 o BaReH9 en torno al 9 % en peso. Sin

Descripción del trabajo realizado

8

embargo, la formación de estos compuestos es lenta y su descomposición requiere

temperaturas superiores a 300 ºC. Los alanatos y borohidruros (hidruros complejos de

Al y B, respectivamente) admiten cantidades de H2 incluso mayores. Por ejemplo, el

LiBH4 contiene un 18,3 % en peso de H2, pero se descompone a temperaturas de hasta

600 ºC y presenta cierta irreversibilidad en el proceso de captura de hidrógeno [4].

A.2. El metanol como fuente de hidrógeno

El CH3OH es un producto fundamental en la industria química mundial que se

utiliza en la obtención de formaldehido, MTBE, ácido acético, metil metacrilato,

formato de metilo, etc. También está cobrando una enorme importancia como fuente

de H2, e incluso se puede utilizar directamente para generar energía mediante celdas

de combustible (“Direct Methanol Fuel Cells”, DMFC) [13]. De hecho, al igual que

sucede con el H2, se habla de una posible “economía del metanol” [22].

En esta tesis se ha escogido el metanol como fuente de H2 debido a sus

numerosas ventajas. Tiene una elevada relación H/C (4/1), la misma que el metano,

pero su naturaleza líquida en condiciones ambientales facilita mucho su transporte,

almacenamiento y manipulación. El reformado de metanol se puede llevar a cabo a

menores temperaturas (150 – 350 ºC) que muchos otros combustibles porque no

contiene enlaces C-C, no requiere procesos de desulfuración y produce una cantidad

de CO inferior al metano [13]. No es de los combustibles más tóxicos (lo es más que

el diésel y el etanol, pero menos que la gasolina, por ejemplo), ya que es fácilmente

metabolizado por el cuerpo humano y por los microorganismos presentes en el medio

ambiente. Además, el CH3OH se puede producir de forma sostenible a partir de

fuentes biológicas [23], e incluso en procesos de revalorización de CO2, vía

hidrogenación [24] o conversión fotocatalítica [25].

A.2.1. Reacciones de conversión catalítica de metanol

Las principales reacciones de producción de hidrógeno a partir de metanol son

las de descomposición (MD, reacción A.7), reformado con vapor de agua (SRM,

reacción A.8) y oxidación parcial (POM, reacción A.9).

CH3OH → 2H2 + CO (A.7)

Introducción

9

CH3OH + H2O → 3H2 + CO2 (A.8)

CH3OH + 1/2O2 → 2H2 + CO2 (A.9)

El SRM suele tener lugar a temperaturas de 250-300 ºC y presenta la mayor

proporción de hidrógeno por átomo de carbono (H/C = 3). La oxidación parcial de

metanol obtiene mayores conversiones a menores temperaturas (a partir de 120 ºC), y

por tanto se puede llevar a cabo en reactores de menor tamaño [13]. Además,

dependiendo del catalizador empleado, se pueden obtener otros subproductos a partir

de metanol, algunos de gran valor añadido como el formaldehído (H2CO) o el formato

de metilo (HCOOCH3), por medio de mecanismos como el mostrado en la Figura A.1.

Figura A.1. Esquema de reacción de la oxidación de metanol (adaptado de Tatibouët [26]).

A.2.2. Catalizadores empleados para la producción de H2

Los catalizadores utilizados en las reacciones A.7, A.8 y A.9, así como en la

síntesis de metanol (reacción inversa de A.8) y water-gas shift (reacción A.6), se

pueden clasificar en dos categorías: catalizadores de Cu o de metales del grupo VIII.

i) Catalizadores basados en Cu.

El cobre es un metal muy económico y presenta una excelente actividad en

todas estas reacciones. Por otro lado, es pirofórico cuando se expone al aire y es

susceptible de desactivarse por la deposición de carbono sobre su superficie y,

principalmente, por la sinterización de las partículas metálicas. De hecho, entre los

metales más utilizados, solo la plata es menos estable que el Cu en este sentido según

Descripción del trabajo realizado

10

la clasificación de Hughes: Ag < Cu < Au < Pd < Fe < Ni < Co < Pt < Rh < Ru < Ir <

Os < Re [12]. Por estos motivos se suelen emplear soportes óxidos como ZnO, Al2O3,

CeO2, SiO2, TiO2 o ZrO2, que actúan como promotores estructurales (mejorando la

dispersión del metal) y/o electrónicos (modificando sus propiedades catalíticas) [11,

13, 27, 28]. También se puede promocionar el Cu adicionando sales o hidróxidos de

metales alcalinos, cuyo efecto aumenta con el radio atómico (Cs > Rb > K > Na > Li)

[29-31].

ii) Catalizadores basados en metales del grupo VIII.

Aunque estos no presentan tantos problemas de estabilidad como el Cu (el Pt y

el Pd principalmente), algunos como el Ni son mucho menos selectivos a la reacción

de reformado con vapor de agua (reacción A.8) que a la descomposición de metanol

(reacción A.7) [13]. Este metal presenta además una gran tendencia a desactivarse por

la formación de depósitos de carbono derivados de la disociación de CO o CH4 [11].

Al igual que el Cu, los metales de este grupo se suelen soportar sobre óxidos metálicos

[11, 13, 28] y también se pueden promocionar con compuestos alcalinos [32-34].

A.3. La promoción electroquímica de la catálisis (EPOC)

El fenómeno de la promoción electroquímica (“Electrochemical Promotion of

Catalysis”, EPOC) se basa en el uso de la electroquímica para activar y sintonizar un

catalizador heterogéneo en una reacción espontánea, de un modo que parece obviar la

ley de Faraday [35].

A.3.1. Origen y mecanismo de la promoción electroquímica

En 1981 Vayenas [36] observó por primera vez que la actividad y selectividad

de un catalizador metálico depositado sobre un electrolito sólido (conductor iónico) se

pueden modificar electroquímicamente durante el propio transcurso de una reacción

química. Esta modificación puede llevarse a cabo de modo controlado y reversible,

mediante la aplicación de una diferencia de potencial o intensidad eléctrica entre esta

película metálica (catalizador-electrodo de trabajo) y un segundo electrodo

(contraelectrodo) depositado sobre el lado opuesto del electrolito. Así, si sobre el

electrodo de trabajo tiene lugar una reacción catalítica (Figura A.2), la aplicación de

Introducción

11

una corriente eléctrica, (I, del orden de µA) puede ocasionar un incremento en la

velocidad de reacción (Δr), que puede superar en varios órdenes de magnitud al valor

previsto por la ley de Faraday (re). Por ello este fenómeno se conoce también como

“modificación electroquímica no faradaica de la actividad catalítica” (NEMCA).

Figura A.2. Representación del efecto EPOC empleando K/β-Al2O3 como electrolito sólido.

Las especies iónicas promotoras se generan en la región llamada tpb (three-

phase boundaries), que consiste en la interfase entre el electrolito sólido, el catalizador

y la fase gas. Estas especies son acompañadas por los correspondientes iones de

compensación de carga, formando dipolos neutros que se distribuyen a lo largo de la

superficie metálica y que constituyen la llamada doble capa efectiva (Figura A.3). Esta

modifica la densidad electrónica del catalizador y por tanto su función de trabajo (Δϕ)

que se relaciona con la variación del potencial (ΔVWR) según la ecuación A.10 [37]:

Δϕ = eΔVWR (A.10)

De este modo se varía la capacidad de enlace del catalizador con cada uno de

los reactivos y por tanto la velocidad de la reacción. Para confirmar esta teoría se han

empleado diversas técnicas de caracterización tanto catalíticas (TPD, medida de la

función de trabajo), como electroquímicas (voltametría cíclica, espectroscopía de

impedancia) y de análisis de superficies (XPS, UPS, PEEM, STM), poniendo de

manifiesto que este fenómeno es análogo a la promoción química convencional en

re I < 0

Δr

Δr >> re

re = I/nF

K+

K+ K

+

K+K+

Descripción del trabajo realizado

12

catálisis heterogénea [35, 38]. Sin embargo, mientras en la promoción química los

promotores se adicionan al catalizador durante su preparación, en la promoción

electroquímica estos son iones que migran desde el electrolito sólido hacia el metal, de

modo que se puede controlar su movimiento por medio de la corriente aplicada.

Figura A.3. Formación de la doble capa efectiva en un metal depositado sobre un conductor

aniónico (a) y otro catiónico (b), y su influencia sobre la quimisorción de los reactivos.

Por ejemplo, como se observa en esta figura, una intensidad positiva causa la

migración de iones negativos hacia el metal si se utiliza un electrolito sólido aniónico

(por ejemplo YSZ), mientras que una intensidad negativa produce la migración de

iones positivos desde un conductor catiónico como el empleado en esta tesis (K-

βAl2O3). De este modo, la adición de especies promotoras electronegativas (O2-

) y el

consecuente incremento en la función de trabajo del metal, favorece la quimisorción

de moléculas donadores de electrones (D) y desfavorece la de aceptores de electrones

(A). Por el contrario, si las especies adicionadas son promotores electropositivos (K+),

Doble capaefectiva

tpb

(I > 0)

(I < 0)

a)

b)

Introducción

13

la función de trabajo disminuye (Figura A.3b) y se favorece la quimisorción de los

aceptores de electrones frente a los donadores. Por tanto, dependiendo de las

condiciones de reacción, del catalizador empleado y del carácter electronegativo o

electropositivo de los distintos adsorbatos, una determinada polarización tendrá un

efecto positivo o negativo sobre la cinética global de la reacción. Se pueden distinguir

cuatro tipos de reacciones basadas en la promoción electroquímica [39]:

Reacciones electrofóbicas: muestran un incremento de la velocidad de

reacción para valores positivos del potencial. Este tipo de comportamiento tiene lugar

cuando la cinética es de orden positivo respecto al donador de electrones y de orden

cero o negativo respecto al aceptor de electrones; es decir, el donador de electrones es

el que se encuentra más débilmente adsorbido sobre el catalizador.

Reacciones electrofílicas: muestran un incremento de la velocidad de reacción

para valores negativos del potencial. Este tipo de comportamiento tiene lugar cuando

la cinética es de orden positivo respecto al aceptor de electrones y de orden cero o

negativo respecto al donador de electrones; es decir, el aceptor de electrones es el que

se encuentra más débilmente adsorbido sobre el catalizador.

Reacciones tipo volcán: presentan un máximo en la velocidad de reacción

respecto al potencial. Este tipo de comportamiento tiene lugar cuando tanto el donador

como el aceptor de electrones se encuentran fuertemente adsorbidos sobre el

catalizador (orden cero o negativo de ambos).

Reacciones tipo volcán invertido: presentan un mínimo en la velocidad de

reacción respecto al potencial. Este tipo de comportamiento tiene lugar cuando tanto el

donador como el aceptor de electrones se encuentran débilmente adsorbidos sobre el

catalizador (orden positivo de ambos).

Durante las tres últimas décadas el fenómeno EPOC se ha demostrado en más

de 80 sistemas catalíticos diferentes de interés industrial y medioambiental, y no

parece estar limitado a ningún tipo de reacción catalítica heterogénea, catalizador

metálico o electrolito sólido [35, 38]. A modo de ejemplo, en la Tabla A.2 se muestran

algunos de los sistemas catalíticos en los que se ha aplicado este fenómeno.

Descripción del trabajo realizado

14

Tabla A.2. Ejemplos de reacciones catalíticas que se han promocionado vía electroquímica.

Reacción Catalizador Electrolito sólido Ref.

CO + 1/2O2 → CO2 Ag Na-βAl2O3 [40]

SO2 + 1/2O2 → SO3 Pt V2O5-K2S2O7 [41]

CH3OH → H2 + H2CO

CH3OH → 2H2 + CO

Pt YSZ [42]

N2O → N2 + 1/2O2

C3H6 + 9/2O2 → 3CO2 + 3H2O

Pt K-βAl2O3 [43]

C2H4 + 3O2 → 2CO2 + 2H2O IrO2 YSZ [44]

H2 + CO → CxHy + CxHyOz Pd YSZ [40]

C3H6 + 1/2O2 → C3H6O

C3H6 + 9/2O2 → 3CO2 + 3H2O

Ag YSZ [45]

2NO + 2CO → N2 + 2CO2 Cu Na-βAl2O3 [46]

2NH3 → N2 + 3H2 Fe CaZr0,9In0,1O3-a [47]

CH3OH + 1/2O2 → H2CO + H2O

CH3OH + 3/2O2 → CO2 + 2H2O

Pt YSZ [48]

C6H5CH3+ 9O2 → 7CO2+ 4H2O Ag YSZ [49]

3H2 +C6H6 → C6H12 Pt Na-βAl2O3 [50]

C7H8 + 9O2 → 7CO2 + 4H2O Ag YSZ [51]

C3H6 + 9/2O2 → 3CO2 + 3H2O

C3H6 + 9NO → 3CO2 + 9/2N2 + 3H2O

Pt Nasicon [52]

N2 + 3H2 → 2NH3 Fe SZY [53]

1-C4H8 + 1-C4H8 + H2 → C4H10 + 2-C4H8 Pd/C Nafion [54]

CH4 + H2O → 3H2 + CO

CH4 + 1/2O2 → 2H2 + CO

CH4 + 1/2H2O + 1/4O2 → 5/2H2 + CO

CO + H2O → CO2 + H2

Pt/YSZ Na-βAl2O3 [55]

CH4 + 2O2 → CO2 + 2H2O Pd/CeO2 YSZ [56]

C2H4 + H2 → C2H6 Ni CsHSO4 [57]

CO + O2 + H2 → CO2 + H2O Pt K-βAl2O3 [58]

H2 + 1/2O2 → H2O Pt/C Nafion [59]

CO2 + H2 → CO + H2O

CO2 + 4H2 → CH4 + 2H2O

2CO2 + 6H2 → C2H4 + 4H2O

Rh, Pt, Cu/TiO2 YSZ [60]

CO + H2O → CO2 + H2 Ni K-βAl2O3 [61]

Introducción

15

El creciente interés que suscita este fenómeno en la comunidad científica

internacional también queda reflejado en la Figura A.4. En esta figura se muestra el

crecimiento en el número de citas relacionadas con el fenómeno en los últimos veinte

años.

Figura A.4. Variación del número de citas relacionadas con el fenómeno EPOC en los últimos

años (Fuente: Web of Science, Mayo 2015).

A.3.2. Promoción electroquímica con conductores alcalinos

Existe una gran variedad de electrolitos sólidos, conductores de iones H+, K

+,

Na+, Li

+, Cu

+, Ag

+, O

2- o F

- [62]. Aunque los primeros trabajos de promoción

electroquímica (EPOC) se realizaron con electrolitos sólidos aniónicos (YSZ) [36], los

conductores alcalinos también se han sido objeto de numerosos estudios [63]. En la

Tabla A.3 se muestran los principales hitos que han tenido lugar con estos materiales.

Los principales electrolitos sólidos alcalinos pertenecen a las familias de β’’-alúmina

(que se van a denominar βAl2O3 de ahora en adelante) y NASICON (“Sodium Super

Ionic Conductor”, Na3Zr2Si2PO12).

Year

Cita

tio

ns

Descripción del trabajo realizado

16

Tabla A.3. Algunas de las contribuciones más relevantes en la historia del EPOC con

conductores alcalinos (adaptado de de Lucas-Consuegra [63]).

Año Contribución Referencia

1991 Primer estudio EPOC usando un conductor alcalino (Pt/Na-βAl2O3) [64]

1995 Confirmación del mecanismo de backspillover de los iones Na+

mediante XPS

[65]

1996 Demostración la promoción de Pt con Na+ mediante caracterización

STM

[66]

1997 Primer estudio de EPOC usando un conductor de K+

(Fe/K2YZr(PO4)3)

[47]

1998 Simulación Monte Carlo de la promoción con Na+ de la reacción de

CO + NO sobre Pt

[67]

1999 Primera demostración de la promoción controlada in-situ de un metal

base (Cu) con Na+ en la reacción de CO + NO

[46]

2001 Reglas de la promoción electroquímica [68]

2001 Aplicación de la espectroscopía electrónia y EPOC a la reacción de

CO + NO sobre Rh. Mecanismo de la promoción alcalina

[69]

2003 SCR de NO con propeno usando Pt/NASICON bajo condiciones de

mezcla pobre en combustible

[52]

2005 Promoción electroquímica de Rh con K+ en la síntesis de Fischer-

Tropsch a 14 bar

[70]

2007 Promoción electroquímica de Pt/K-βAl2O3 a baja temperatura (200

ºC): Efecto permanente y estudio FTIR-EDX de las fases promotoras

[71]

2008 Desarrollo de un tubo de Pt/K-βAl2O3 para el

almacenamiento/reducción de NOx asistido electroquímicamente

[72]

2010 Regeneración in-situ de un catalizador de Pt de la deposición de

carbono mediante Na-βAl2O3

[55]

2013 Estudio a escala bancada de la promoción electroquímica de la

hidrogenación de CO2 con Cu/K-βAl2O3

[73]

Vayenas realizó el primer estudio de EPOC con electrolito alcalino en 1991

[64]. Desde entonces, tanto el Na-βAl2O3 como el NASICON (ambos conductores de

Na+) se han empleado en numerosos procesos catalíticos como la oxidación de etileno,

de CO, de propano o de propileno, la reducción de NO y la hidrogenación de benceno

o de CO2 [63]. Por otro lado, el primer estudio EPOC con un conductor de K+

(K2YZr(PO4)3) data de 1997 [47], y no es de extrañar que estuviera aplicado a la

descomposición de NH3 con un catalizador de Fe, ya que esta reacción y la de síntesis

Introducción

17

de amoniaco son ejemplos clásicos de procesos químicos industriales promocionados

con potasio [74]. Más adelante, de Lucas-Consuegra introdujo el empleo de K-βAl2O3

para la promoción electroquímica de catalizadores de Pt en la oxidación de CO y de

propileno, así como en la reducción de NOx y N2O [63]. También caben destacar en

los últimos años los estudios de EPOC con este electrolito sólido en la reacción de

hidrogenación de CO2 [73, 75].

En esta tesis se ha seleccionado K-βAl2O3 como conductor debido a su

adecuación a las condiciones de reacción empleadas en esta investigación (puede

operar a partir de 150 ºC) y a los estudios previos en bibliografía que demuestran el

efecto promotor de los metales alcalinos, y en especial del potasio (favorecido por su

mayor radio atómico en comparación con el Na, por ejemplo) en las reacciones de

descomposición de metanol [29], reformado de metanol con vapor de agua [32, 33] y

oxidación parcial de metanol [31]. En todos estos estudios, el efecto electrónico del

potasio causó una mejora de la actividad catalítica atribuida al debilitamiento del

enlace C-H en los intermedios de reacción. Sin embargo, como se ha señalado

anteriormente, en este tipo de promoción alcalina (clásica) el potasio se adiciona en

forma de sal o de hidróxido durante la etapa de preparación del catalizador. En

cambio, en los estudios de EPOC el potasio se suministra al catalizador en el propio

transcurso de la reacción por vía electroquímica mediante la aplicación de una

corriente o potencial eléctricos. Esto le confiere a este tipo de sistemas una serie de

ventajas adicionales que se enumeran en la Tabla A.4.

Algunas de estas ventajas son particularmente importantes, como la capacidad

de optimizar in-situ la cantidad de promotor bajo diferentes condiciones de reacción, y

de sintonizar el catalizador para maximizar su actividad y su selectividad hacia el

producto deseado en todo momento. También resulta de especial interés la posibilidad

de prevenir la desactivación del catalizador e incluso de regenerarlo. Estas

aplicaciones del EPOC se han estudiado previamente en diferentes sistemas de

reacción, como se recoge en distintos trabajos de revisión [38, 63]. Sin embargo, en

esta tesis se han estudiado por primera vez en procesos de producción de hidrógeno a

partir de metanol. Además, algunos aspectos como el efecto de los iones promotores

Descripción del trabajo realizado

18

sobre el estado de oxidación del metal se han estudiado por primera vez con un

electrolito alcalino en el presente trabajo de investigación.

Tabla A.4. Principales ventajas de la promoción electroquímica alcalina frente a la promoción

química clásica, y sus aplicaciones.

Ventaja Aplicación

Control in-situ de la cantidad de promotor

sobre la superficie del catalizador (mediante

la aplicación de una corriente o potencial

eléctricos)

Posibilidad de estudiar de forma rápida y

sencilla el efecto de un determinado promotor

sobre un sistema catalítico, e interpretarlo en

base a las reglas EPOC. No es necesario

preparar varios catalizadores con distintas

cantidades de promotor

Adición de la cantidad de promotor óptima

bajo diferentes condiciones de reacción

Optimización in-situ de la actividad y

selectividad de un catalizador. Utilidad en

procesos no estacionarios, como en la catálisis

de automoción

Utilización de la celda de electrolito sólido

como un sensor

Posibilidad de anticipar las condiciones

óptimas que maximizan una velocidad de

reacción

Modificación in-situ del estado de oxidación

del metal

Control de la fase activa del catalizador sin

modificar las condiciones de reacción

Control y atenuación y la adsorción

competitiva de unas moléculas frente a otras

Prevención del envenenamiento de la

superficie del catalizador

Promoción in-situ de reacciones de

eliminación de especies adsorbidas (como

carbón depositado)

Regeneración de un catalizador sin necesidad

de cambiar las condiciones de reacción

Almacenamiento de compuestos sobre la

superficie del catalizador mediante la

formación de especies derivadas alcalinas

Gran utilidad en procesos cíclicos como la

captura/reducción de NOx y el

almacenamiento de gases de interés como

CO2 o H2

A.4. Nuevas tendencias y perspectivas de la promoción electroquímica

A.4.1. Desarrollo de reactores más eficientes y catalizadores dispersos y más

competitivos

A pesar de las múltiples posibilidades que ofrecen los sistemas

electrocatalíticos empleados en los estudios de promoción electroquímica, debido a

sus particulares características presentan también ciertas limitaciones que disminuyen

su aplicabilidad. Por un lado, el reactor que se emplea en este tipo de estudios debe

Introducción

19

tener un diseño especial que le permita conectar el catalizador electroquímico con el

sistema de polarización, y generalmente no está optimizado. Por otro lado, los

catalizadores empleados en estudios de promoción electroquímica suelen presentar

una baja dispersión de las partículas y una elevada carga metálica. Por todo ello, con

vistas a una posible etapa de comercialización futura del EPOC, resulta esencial

dedicar mayores esfuerzos al diseño de reactores más compactos y eficientes, así

como a minimizar el coste de los materiales [76].

En la tabla A.5 se recogen los estudios de promoción electroquímica más

relevantes realizados durante los últimos años en ambos sentidos, algunos de los

cuales ya han sido mencionados anteriormente. Se incluyen también los catalizadores

desarrollados durante la presente investigación.

Una de las primeras configuraciones para la posible aplicación práctica del

fenómeno EPOC fue desarrollada por Yiokari y col., [77] para la síntesis de amoniaco

con catalizadores de Fe comerciales. Ésta consistía en un conjunto de 24 pellets de

CaZr0,9In0,1O3-α (conductor de H+) conectados eléctricamente en paralelo. Además, este

fue también uno de los primeros estudios de promoción electroquímica realizado a

elevadas presiones (50 atm). Otro gran avance en este sentido fue el desarrollo del

reactor monolítico promocionado electroquímicamente (“Monolithic

Electrochemically Promoted Reactor”, MEPR) [78]. Este reactor fue desarrollado en

2004 tras la consolidación de la técnica de pulverización catódica (“sputtering”) como

método de deposición de películas metálicas delgadas. Este reactor permite trabajar a

una mayor escala y se puede considerar un híbrido entre uno monolítico y otro de tipo

celda de combustible de óxido sólido. De este modo se han desarrollado varias

configuraciones de MEPR basadas en 22 unidades de Rh/YSZ/Pt o Cu-TiO2/YSZ/Au

conectadas en serie para la hidrogenación de CO2 a CO y CH4 [60] y para la reducción

catalítica selectiva de NO con etileno [78-80].

Descripción del trabajo realizado

20

Tabla A.5. Estudios de EPOC más relevantes en el desarrollo de reactores más eficientes y catalizadores más competitivos.

Diseños de reactores compactos y eficientes

Configuración del reactor Catalizador electroquímico Reacción en studio Ref.

Configuración multipellet Fe/CaIn0,1Zr0,9O3-α Síntesis de NH3 [77]

Reactor monolítico Rh/YSZ/Pt Oxidación de C2H4 y reducción de NO con C2H4 en presencia de O2 [78-80]

electropromocionado (MEPR) Cu-TiO2/YSZ Hidrogenación de CO2 [60]

Membranas de fibra hueca Pt/LSCF Oxidación de C2H4 [81]

Catalizadores electroquímicos de elevada dispersión Catalizadores de metales no nobles

Catalizador a Reacción en estudio Ref. Catalizador

a Reacción en eestudio Ref.

Pt-Au/YSZ Oxidación de C2H4 [76] Ni/K-βAl2O3 WGS [61]

Pt-LSCF-GDC/GDC Oxidación de C3H8 [82] Ni/CsHSO4 Hidrogenación de C2H4 [57]

[83] Ni/YSZ (SOFC) Reformado de CH4 [84]

Ni-CNF/YSZ Hidrogenación de CO2 [83] Ni-CNF/YSZ Hidrogenación de CO2 [83]

Ru-CNF/YSZ Ni/YSZ (tubular) Hidrogenación de CO2 [85]

Pt-C/K-βAl2O3 Oxidación de CO y C3H6 [86] Ni/ K-βAl2O3 Descomposición de CH3OH Este trabajo

Reformado de CH3OH

Pt-DLC/K-βAl2O3 Oxidación parcial de CH3OH Este trabajo Oxidación parcial de CH3OH

Reformado de CH3OH Fe/K2YZr(PO4)3 Descomposición de NH3 [47]

Au-YSZ/K-βAl2O3 Oxidación parcial de CH3OH Este trabajo Fe/SZY Síntesis de NH3 [53]

Fe/CaIn0,1Zr0,9O3-α (multipellet) Síntesis de NH3 [77]

Pt/YSZ Reducción de NO con C3H6 [87] Cu/Na-βAl2O3 Reducción de NO con CO [46, 88]

Pd/YSZ (monolito) Oxidación de CH4 [89] PtCu/Nafion (PEMFC) PROX, WGS [90]

Cu-TiO2/YSZ (MEPR) Hidrogenación de CO2 [60]

Pt-YSZ/Na-βAl2O3 Reformado de CH4 [55] Cu/ K-βAl2O3 (tubular) Hidrogenación de CO2 [75]

Oxidación parcial de CH4 Cu/SZY Hidrogenación de CO2 [91]

Reformado autotérmico de CH4 Cu/ K-βAl2O3 Oxidación parcial de CH3OH Este trabajo a A menos que se especifique lo contrario, el reactor presenta una configuración de pellet sencillo.

Introducción

21

En cuanto al desarrollo de catalizadores dispersos, se pueden encontrar

estudios en los que se emplea un material adicional que puede actuar simultáneamente

como soporte de las partículas del catalizador y como material conductor eléctrico, por

ejemplo, oro [76], carbón [54], nanofibras de carbono [83] o materiales conductores

mixtos (de iones y electrones) [81, 82]. Por otro lado, en la Tabla A.5 se muestran

diversos trabajos de promoción electroquímica, en los que se emplearon metales como

Ni [57, 61, 83, 85, 92], Fe [47, 53, 77] o Cu [46, 60, 75, 90, 91], de menor coste

económico que los típicos metales nobles empleados en estudios de EPOC como Pt,

Rh o Pd. En este sentido uno de los objetivos principales de esta Tesis ha sido el

desarrollo de catalizadores más competitivos que los empleados tradicionalmente para

su activación electroquímica en procesos de producción de H2 a partir de metanol. Así,

en los dos primeros capítulos se emplearon catalizadores electroquímicos de Pt de bajo

contenido metálico preparados mediante una técnica de deposición física de vapor

(PVD). En el segundo caso las partículas de Pt estaban, además, dispersas en una

matriz de carbono, presentando diámetros de unos 3 nm. En la misma línea, en el

capítulo 3 se desarrolló una película compuesta de nanopartículas de Au inmersas en

una matriz de zirconia estabilizada con itria (YSZ), que además de dispersar las

nanopartículas metálicas contribuía a evitar su sinterización. Por último, a partir del

capítulo 4 se emplearon catalizadores de metales no nobles (Cu y Ni) preparados

mediante técnicas de PVD que igualmente consiguieron reducir el contenido metálico

por unidad de área o aumentar el área superficial de estos electrodos.

A.4.2. Aplicación de la promoción electroquímica en la producción de H2

Al margen de la necesaria mejora y optimización de los sistemas

electrocatalíticos, otra posible estrategia para impulsar la aplicación práctica de la

promoción electroquímica es explorar nuevos procesos catalíticos en los ámbitos

energético o medioambiental donde este fenómeno pueda resultar interesante. Uno de

estos sistemas podría ser la producción y almacenamiento de H2 que tiene un gran

protagonismo en la actualidad ya que se postula como el posible vector energético del

futuro [2, 3]. Sin embargo, a excepción de los trabajos de Pitselis y col. de

descomposicón de NH3 [47] y de Yentekakis y col. de reformado de metano con vapor

Descripción del trabajo realizado

22

de agua en una celda de combustible de óxido sólido [92], solo el presente grupo de

investigación ha llevado a cabo estudios EPOC enfocados a reacciones de producción

de hidrógeno: water-gas shift [61] y conversión de metano mediante diversas vías [55,

93].

Como se mencionó en la sección A.2, la producción de hidrógeno a partir de

metanol es especialmente interesante debido fundamentalmente a su elevado

contenido en H2, su naturaleza líquida y su posible obtención por medios sostenibles.

En cuanto a los procesos de conversión de metanol, hasta el inicio de la presente tesis

solo había publicados 4 estudios de promoción electroquímica, en los que se empleó

Ag [42, 94] o Pt [48, 95] como catalizador-electrodo de trabajo, YSZ como electrolito

sólido en todos los casos, y elevadas temperaturas de trabajo de hasta 500 ºC. Por

tanto, en esta tesis se ha aplicado por primera vez el fenómeno de la promoción

electroquímica a reacciones de conversión de metanol para la producción de

hidrógeno, utilizando para ello un electrolito sólido alcalino (K-βAl2O3) y

temperaturas de operación generalmente inferiores a los estudios anteriores. En la

Tabla A.6 se resumen todos los estudios de EPOC realizados hasta la fecha en los que

se han estudiado reacciones de conversión de metanol (incluyendo los de la presente

tesis).

Las principales reacciones catalíticas que tuvieron lugar en los estudios con

YSZ fueron la descomposición del metanol a CO (reacción A.7) y H2CO (reacción

A.11), la reacción de metanación (reacción A.12) [42], y la oxidación de metanol por

medio de las reacción A.13 y A.14 [48, 94, 95], obteniendo principalmente

formaldehído como producto interés.

CH3OH → H2CO + H2 (A.11)

CH3OH + H2 → CH4 + H2O (A.12)

CH3OH + 1/2O2 → H2CO + H2O (A.13)

CH3OH + 3/2O2 → CO2 + 2H2O (A.14)

Introducción

23

Tabla A.6. Estudios de promoción electroquímica aplicada a reacciones de conversión de metanol.

Reacción Catalizador Electrolito sólido Método de preparación Principales productos Efecto EPOC Ref.

MD Ag YSZ Impregnación CO, H2CO, CH4, H2, H2O Electrofílico [42]

Ni K-βAl2O3 CADa H2, CO, C Electrofílico Este trabajo

SRM Pt K-βAl2O3 CADa H2, CO, CO2 Electrofílico Este trabajo

Pt-DLC K-βAl2O3 CADa H2, CO, CO2 Electrofílico Este trabajo

Ni K-βAl2O3 CADa H2, CO, CO2 Electrofílico Este trabajo

Ni K-βAl2O3 OADa H2, CO, CO2, C Electrofílico Este trabajo

POM Pt YSZ Pasta organometálica CO2, H2CO, H2O Volcán invertido [48]

Pt YSZ Pasta organometálica CO2, H2CO, H2O, H2 Electrofóbico [95]

Ag YSZ Pasta organometálica CO2, H2CO, H2O Electrofílico [94]

Pt K-βAl2O3 Impregnación H2, CO, CO2, H2CO, H2O Electrofóbicob Este trabajo

Pt K-βAl2O3 CADa H2, CO, CO2, H2CO, H2O Electrofílico Este trabajo

Pt-DLC K-βAl2O3 CADa H2, CO, CO2, H2CO, H2O Electrofílico Este trabajo

Au-YSZ K-βAl2O3 Pulverización catódica H2, CO, CO2, HCOOCH3, H2O Electrofílico Este trabajo

Cu K-βAl2O3 OADa H2, CO, CO2, HCOOCH3, H2O Electrofílico Este trabajo

Ni K-βAl2O3 CADa H2, CO, CO2, H2CO, H2O Electrofóbico

b Este trabajo

aCAD = Deposición por arco catódico; OAD = Deposición en ángulo oblicuo. bEfecto aparente, afectado por variaciones en el estado de oxidación del catalizador o por el bloqueo de centros activos por compuestos de potasio

Descripción del trabajo realizado

24

Así pues, esta tesis doctoral se ha enfocado a la conversión de metanol para la

producción de H2 y CO2 a través de las reacciones A.8 (SRM) y A.9 (POM), aunque

en general también se ha observado una elevada formación de CO y de otros

compuestos como formaldehído o formato de metilo, dependiendo del catalizador.

A.4.3. Aplicación de la promoción electroquímica en la captura de compuestos

En los últimos años se ha dado a conocer una nueva aplicación de los sistemas

electrocatalíticos alcalinos, que consiste en la posibilidad de almacenar

electroquímicamente compuestos sobre la superficie de una película catalítica porosa.

De Lucas-Consuegra y col. demostraron por primera vez este concepto hace siete años

[72]. Para ello depositó una película de catalizador de Pt sobre un tubo de K-βAl2O3

con el cual estudió la captura/reducción de NOx. Así pues, en una primera etapa del

proceso llevada a cabo a potenciales negativos, los iones K+ desplazados desde el

electrolito sólido hasta el catalizador no solo promocionaron la reacción de oxidación

de NO, sino también almacenaron NOx sobre la superficie del catalizador en forma de

nitratos de potasio. Además, otro estudio en la misma línea de investigación [96]

demostró la posibilidad de regenerar electroquímicamente la superficie de este tipo de

catalizadores para su empleo bajo una composición de gas pobre.

Más recientemente, Ruiz y col. aplicaron este fenómeno de captura asistida por

vía electroquímica a la eliminación de CO2 [97]. Este estudio demostró que se puede

almacenar CO2 sobre la superficie de un catalizador metálico bajo determinadas

condiciones de reacción y en presencia de iones K+ (dando lugar a la formación de

carbonatos de potasio). Adicionalmente, las especies capturadas pueden ser liberadas

al invertir la polarización regenerando el catalizador y mostrando la completa

reversibilidad de esta tecnología de captura. Por tanto, se consideró interesante

explorar también la posible utilización de estos catalizadores electroquímicos

catiónicos en sistemas acoplados de producción y almacenamiento de H2.

Introducción

25

A.5. Objetivo de la tesis doctoral

El objetivo principal de esta tesis doctoral es estudiar el fenómeno de

promoción electroquímica de la catálisis en reacciones de producción de hidrógeno a

partir de metanol empleando conductores alcalinos. Para ello se ha colaborado con

distintos grupos de investigación especializados en la preparación de películas

delgadas que puedan actuar como catalizador-electrodo de trabajo. De este modo se

pretende contribuir al campo de la producción catalítica de H2 mediante una

aproximación novedosa, que prácticamente no ha sido estudiada hasta la fecha. Con

este fin se ha desarrollado un programa de trabajo que se puede dividir en las

siguientes etapas:

- Realizar una revisión bibliográfica y poner en marcha una instalación

experimental donde llevar a cabo los experimentos de promoción electroquímica

(equipos de análisis, polarización, control de flujo, reactor electroquímico, etc).

- Estudiar la promoción electroquímica aplicada a la oxidación parcial de

metanol y al reformado de metanol con vapor de agua, utilizando catalizadores de Pt

depositados sobre K-βAl2O3 (conductor de K+) mediante una técnica de deposición

física de vapor (PVD) y otras más tradicionales (impregnación).

- Desarrollar películas delgadas de catalizador que sean más competitivas que

las comúnmente empleadas en estudios de promoción electroquímica, mediante la

aplicación de una técnica PVD (deposición por arco catódico, “sputtering”, deposición

en ángulo oblicuo). Para ello se ha procurado disminuir la carga metálica y aumentar

la dispersión por medio de una matriz (DLC, YSZ) que contenga nanopartículas del

metal, así como obtener películas de metales no nobles (Cu, Ni) que presenten cierta

estabilidad.

- Adecuar el catalizador empleado en cada estudio para poder llevar a cabo la

promoción electroquímica, mediante la deposición de una película de catalizador que

presente al mismo tiempo una buena conductividad eléctrica y adecuadas propiedades

catalíticas, y si es necesario, mediante la realización de algún proceso de

acondicionamiento del catalizador o el empleo de un colector de corriente.

Descripción del trabajo realizado

26

- Evaluar el efecto los iones K+ sobre cada catalizador en las reacciones de

producción de H2 a partir de metanol y discutir los resultados en base a la teoría de la

promoción química y con el apoyo de diversas técnicas de caracterización.

- Estudiar la posibilidad de promocionar mediante iones K+ la producción y

almacenamiento simultáneos de H2 en un único sistema catalítico, y su liberación bajo

las mismas condiciones de operación, todo ello de forma controlada en base a

variaciones en la polarización del catalizador.

Métodos e instalación experimental

27

B. MÉTODOS E INSTALACIÓN EXPERIMENTAL

B.1. Preparación de los catalizadores electroquímicos

B.1.1. Deposición de pasta organometálica

Este método consiste en la deposición de una pasta orgánica del metal y su

posterior calcinación a elevada tempertaura (800 ºC) para lograr la descomposición del

componente orgánico y una adecuada adherencia entre el electrodo y el electrolito

sólido. Esta técnica es muy rápida y sencilla, y se ha utilizado en todos los capítulos

para la deposición del contraelectrodo (CE) y electrodo de referencia (RE) de Au

(inertes).

B.1.2. Impregnación

Esta técnica se ha utilizado en la preparación de uno de los dos catalizadores

de Pt empleados en el capítulo 1. Consiste en la descomposición térmica de una

solución de H2PtCl6 0,1 M en una mezcla agua-2-propanol, por medio de sucesivas

etapas de impregnación (de varios µL de esta disolución), secado y calentamiento a

450 ºC. A diferencia del método de deposición de una pasta organometálica, este

permite establecer la cantidad exacta de metal depositado. Sin embargo, ninguno de

estos dos métodos permite controlar la microestructura de la película de catalizador.

B.1.3. Deposición física de vapor (PVD)

Este es un grupo de técnicas de preparación de películas delgadas, que se

basan en el calentamiento en vacío de una placa del material a depositar (“target”)

hasta lograr una elevada presión de vapor y su posterior condensación sobre un

sustrato frío (K-βAl2O3). Las principales diferencias entre las distintas técnicas PVD

residen en la naturaleza de la fuente del vapor (sólida en estos casos) y en el método

usado para producirlo. Así pues, en esta tesis se han utilizado tres métodos de

deposición física de vapor diferentes:

i) Deposición por arco catódico (CAD).

Esta técnica utiliza un arco eléctrico para vaporizar el material del cátodo que

posteriormente se va a condensar. Los electrodos de Pt, Pt-DLC y Ni utilizados en los

Descripción del trabajo realizado

28

capítulos 1, 2 y 5, respectivamente, se han preparado mediante esta método. Esta

técnica se ha llevado a cabo en colaboración con el Dr. José Luis Endrino del

Instituto de Ciencia de Materiales de Madrid (CSIC).

ii) Pulverización catódica (comúnmente conocida como “sputtering”).

El catalizador de Au-YSZ del capítulo 3 se ha preparado por este método. En

este caso, el material sólido objetivo (Au, Zr e Y) se evapora mediante un bombardeo

con iones energéticos. Esta técnica se ha llevado a cabo en colaboración con el Dr.

David Horwat del Institut Jean Lamour de Université de Lorraine (Francia).

iii) Evaporación por bombardeo electrónico (EBE).

Esta técnica se ha utilizado para depositar películas de Cu y Ni en los capítulos

4 y 6, respectivamente. En concreto se ha aplicado una modificación de este método

denominada “deposición en ángulo oblicuo” o “deposición en ángulo rasante” (OAD o

GLAD, respectivamente), que se basa en colocar el sustrato (K-βAl2O3) formando un

ángulo rasante respecto de la dirección de propagación del vapor que se va a depositar.

De este modo, cuando el material (Cu o Ni) incide sobre el sustrato se produce el

mecanismo de deposición con efecto sombra o “shadowing”, que favorece el

crecimiento de la película de forma columnar. Controlando el ángulo de incidencia del

material y la velocidad de giro de la muestra se consiguen todo tipo de morfologías de

la película. Una de las propiedades que hace más atractiva esta técnica de deposición

es el aumento de la superficie interna y porosidad de la lámina delgada respecto a una

capa compacta del mismo material, lo que puede ser de gran interés desde el punto de

vista electrocatalítico. Esta técnica se ha llevado a cabo en colaboración con el Dr.

Agustín González-Elipe y el Dr. Víctor Joaquín Rico del Instituto de Ciencia de

Materiales de Sevilla (CSIC – Universidad de Sevilla).

Las condiciones concretas de deposición de cada película de catalizador se

detallan en los diferentes capítulos.

Métodos e instalación experimental

29

B.2. Técnicas de caracterización

B.2.1. Microscopia electrónica de barrido (SEM)

Esta técnica se ha empleado en la mayoría de los capítulos de la Tesis. Permite

la observación y análisis de la película del catalizador de forma sencilla, aportando

información sobre la textura, tamaño y forma de las partículas. Además, si se

acompaña de una espectroscopía de energía dispersiva (EDX) permite estimar la

composición química superficial de la lámina y de posibles especies depositadas.

B.2.2. Microscopía electrónica de transmisión (TEM)

Este tipo de microscopía también ofrece información estructural y

morfológica de la muestra del catalizador, y se suele utilizar cuando se requiere una

resolución mayor que con la microscopía SEM, como es el caso de los catalizadores

dispersos Pt-DLC y Au-YSZ en los capítulos 2 y 3, respectivamente. Concretamente

se ha empleado una variante denominada STEM (“scanning transmission electron

microscopy”) que es especialmente adecuada para la obtención de imágenes de

resolución atómica donde el contraste está directamente relacionado con el número

atómico, y para el acoplamiento de técnicas espectroscópicas como EDX o EELS.

B.2.3. Espectroscopía electrónica de pérdidas de energías (EELS)

Esta técnica permite medir la composición atómica de una muestra, enlaces

químicos y propiedades electrónicas de las bandas de valencia y de conducción. Con

respecto al EDX esta técnica es generalmente más adecuada para muestras con

números atómicos relativamente bajos, como el carbono. En concreto se ha utilizado

en el capítulo 2 para identificar la estructura de la matriz carbonosa donde se

encuentran dispersas las nanopartículas de Pt.

B.2.4. Espectroscopía de infrarrojos por transformada de Fourier (FTIR)

Esta técnica aporta información acerca de los modos y tipos de vibraciones

moleculares, y permite de este modo identificar grupos funcionales de moléculas

presentes en la muestra. Se ha utilizado en el capítulo 6 para determinar los tipos de

compuestos carbonosos depositados sobre el Ni bajo promoción electroquímica.

Descripción del trabajo realizado

30

B.2.5. Espectroscopía Raman

Esta técnica es complementaria a la espectroscopía infrarroja en el estudio de

las frecuencias vibracionales de las moléculas, de modo que algunas bandas pueden

observarse solo en el espectro Raman y otras en el infrarrojo. También se utilizó en el

capítulo 6 para la identificación de los compuestos carbonosos adsorbidos.

B.2.6. Espectroscopía de fotoelectrones emitidos por rayos X (XPS)

Esta técnica se suele utilizar para estimar la estequiometria, estado químico y

estructura electrónica de los elementos presentes sobre la superficie del material. Se ha

realizado en los capítulos 4 y 6.

B.2.7. Difracción de rayos X (XRD)

Esta técnica se ha llevado a cabo prácticamente en todos los capítulos, ya que

permite analizar la estructura cristalina del catalizador y obtener un tamaño

aproximado de partícula por el método de Scherrer.

B.2.8. Variación de resistencia eléctrica superficial a temperatura

programada

Esta es una técnica muy sencilla que se ha llevado a cabo con el catalizador Pt-

DLC en el capítulo 2 para obtener una medida aproximada, in-situ, de la variación de

la resistencia eléctrica de la superficie del catalizador durante un proceso realizado a

temperatura programada y así detectar posibles cambios estructurales.

B.2.9. Voltamperometría cíclica (CV)

Esta es una técnica de caracterización electroquímica que se ha utilizado en

gran parte de los capítulos para evaluar los procesos electrocatalíticos asociados a los

iones K+ que tienen lugar sobre la superficie del catalizador y determinar el rango de

potencial adecuado para llevar a cabo los experimentos de promoción electroquímica

de modo reversible y reproducible.

Las condiciones de caracterización y las características de los equipos se

detallan en cada capítulo.

Métodos e instalación experimental

31

B.3. Instalación experimental

Para llevar a cabo los experimentos de promoción electroquímica se preparó y

puso en marcha la instalación experimental que se muestra en la Figura B.1.

Figura B.1. Instalación experimental utilizada en los estudios de promoción electroquímica.

Esta instalación consta de cuatro partes diferenciadas:

i) Sistema de alimentación: Está constituido por cuatro líneas de flujo de gases,

análogas e independientes que pueden suministrar aire (fuente de O2 en la reacción de

POM), H2 (agente reductor) y N2 o Ar, según el capítulo (gas portador de la mezcla de

reacción). Una de estas líneas de gas inerte es saturada a temperatura controlada en

metanol y la otra en agua (en los experimentos de SRM).

ii) Sistema de reacción: Es un reactor de electrolito sólido de cámara sencilla,

donde se localiza el catalizador electroquímico, de modo que los tres electrodos

(trabajo, contraelectrodo, referencia) se encuentran bajo la misma mezcla de reacción.

Descripción del trabajo realizado

32

iii) Sistema de polarización: Consiste en un potenciostato/galvanostato que

permite aplicar potenciales o corrientes eléctricas sobre el catalizador electroquímico.

iv) Sistema de análisis de los productos de reacción: Consiste en un

cromatógrafo de gases situado en linea a la salida del reactor.

Esta instalación se describe detalladamente en el primer capítulo (sección

1.2.3).

Resultados obtenidos

33

C. RESULTADOS OBTENIDOS

Como se ha mencionado, esta tesis está dirigida al estudio de la promoción

electroquímica de la catálisis (EPOC) aplicada a procesos de producción de hidrógeno

empleando metanol como materia prima. Para ello se seleccionó K-βAl2O3 como

soporte del catalizador y electrolito sólido (fuente de iones promotores K+) en todos

los estudios.

En el Capítulo 1 se realizaron los primeros estudios de promoción

electroquímica en las reacciones de oxidación parcial de metanol (POM, CH3OH +

1/2O2 → 2H2 + CO2) y de reformado de metanol con vapor de agua (SRM, CH3OH +

H2O → 3H2 + CO2), empleando como catalizador metálico de partida Pt, en base a sus

buenos resultados obtenidos en bibliografía tanto en las reacciones de POM y SRM

como en otros estudios de promoción electroquímica. Se prepararon dos tipos de

películas de Pt mediantes técnicas diferentes: una mediante impregnación de una

disolución precursora de este metal (dando lugar a un contenido final de 1,1 mg Pt

cm-2

), y otra en colaboración con el Instituto de Ciencia de Materiales de Madrid

(CSIC), mediante deposición por arco catódico (CAD), que es un tipo de técnica de

deposición física de vapor (PVD). Esta permitió obtener una película densa con un

espesor de 120 nm y un contenido metálico de solo 0,23 mg Pt cm-2

.

En primer lugar se compararon los dos catalizadores de Pt/K-βAl2O3/Au en la

reacción de oxidación parcial de metanol (CH3OH/O2 = 7,2 %/4,6 %, 320 ºC) bajo

condiciones de promoción electroquímica, es decir, bajo la aplicación de diferentes

potenciales eléctricos (entre +2 y -2 V). En ambos casos, al igual que en el resto de

experimentos de este capítulo y de toda la tesis, se aplicó un potencial positivo (VWR =

+2 V) al principio y al final del experimento para asegurar la retirada de todos los

iones K+ de la película del catalizador y por tanto para definir un estado no

promocionado o de referencia. A continuación, la disminución del potencial causó la

migración de iones K+ desde el electrolito sólido hasta el catalizador y su distribución

a lo largo de la superficie (lo que se conoce como “back-spillover”), dando lugar a la

formación de una doble capa efectiva que puede modificar la quimisorción de los

distintos reactivos sobre el Pt, tal y como se explicó en la sección A.3.1.

Descripción del trabajo realizado

34

En el caso del Pt depositado por arco catódico (CAD), el suministro de iones

K+ causó un incremento de las velocidades de producción del H2 y CO2 (productos de

la oxidación parcial de metanol) así como del CO, derivado de la descomposición del

metanol (MD, CH3OH → 2H2 + CO) y del formaldehído, obtenido también como

subproducto (CH3OH + 1/2O2 → H2CO + H2O). Este efecto promotor del potasio se

puede explicar en base a una disminución de la función de trabajo del catalizador, que

fortalecería la quimisorción sobre el Pt del aceptor de electrones (O2) y debilitaría la

del donador de electrones (CH3OH). De este modo el efecto EPOC obtenido fue de

tipo electrofílico y consiguió aumentar la producción de hidrógeno y formaldehído 6 y

4 veces, respectivamente, bajo condiciones óptimas de promoción electroquímica.

Por otro lado, en el caso del Pt preparado por impregnación, el descenso en el

potencial dio lugar a una disminución brusca de la actividad catalítica del Pt, que se

atribuyó a la mayor porosidad de este electrodo con respecto al preparado por CAD, y

a la formación de una excesiva cantidad de compuestos superficiales derivados del

potasio (tipo carbonatos) que pudieron bloquear centros activos del Pt. Esta hipótesis

quedó demostrada mediante diferentes técnicas de caracterización (voltametría cíclica,

XRD, SEM-EDX), reflejando que el comportamiento EPOC de un catalizador

electroquímico catiónico no solo depende de su capacidad de quimisorción de las

distintas moléculas de reactivo sino también de su actividad electrocatalítica y su

facilidad para formar compuestos superficiales derivados del promotor.

El catalizador de Pt preparado por CAD fue, por tanto, seleccionado para la

realización de los siguientes experimentos, en los que se estudió la influencia tanto de

la temperatura como de la composicón del alimento. Se comprobó que el efecto de la

promoción electroquímica no solo mejoró la actividad catalítica del platino, sino que

también incrementó su selectividad hacia el mecanismo de la oxidación parcial frente

a la oxidación total. Esto permitió obtener simultáneamente dos productos de elevado

interés como son el H2 y el H2CO en una sola etapa de reacción. Además, se realizaron

ensayos de reproducibilidad y durabilidad que demostraron la reversibilidad y

estabilidad del efecto promocional obtenido con este tipo de catalizadores

electroquímicos. Por último se evaluó la actividad catalítica de este catalizador en

Resultados obtenidos

35

condiciones de reformado de metanol con vapor de agua (CH3OH/H2O = 4 %/4,8 %),

y aunque fue mucho menor que en las condiciones de oxidación parcial, también se

promocionó electroquímicamente con los iones K+, activando en este caso la

quimisorción del agua (aceptor de electrones) frente al metanol (donador de

electrones). Por tanto se demostró la utilidad de los métodos de deposición física de

vapor (PVD) en la obtención de películas delgadas para su uso en estudios de

promoción electroquímica con conductores alcalinos, y por tanto los catalizadores

empleados en el resto de capítulos se prepararon mediante alguna técnica PVD.

En el Capítulo 2 se llevó a cabo una variante de la técnica de CAD empleada

en el capítulo anterior para obtener un catalizador de Pt con mejores propiedades. Esta

consistió en depositar conjuntamente dos materiales diferentes. En este caso, la

película de catalizador resultante estaba formada por nanopartículas de Pt de unos 3

nm de diámetro dispersas en una matriz de carbono tipo diamante (“diamond-like

carbón”, DLC). Con un espesor (del orden de 100 nm) similar a la película de Pt puro

preparada por el mismo método (CAD), la película de Pt-DLC contenía una carga

metálica menor (solo 0,014 mg Pt cm-2

). Como principal inconveniente, este

catalizador no presentaba conductividad eléctrica tras su deposición. Sin embargo, tras

someter el catalizador Pt-DLC/K-βAl2O3/Au a un experimento inicial a temperatura

programada en atmósfera de hidrógeno, su resistencia eléctrica superficial (medida in-

situ) descendió hasta un valor de 3,5 kΩ, que permitió su polarización en los

posteriores experimentos electrocatalíticos. Esta disminución de la resistencia eléctrica

tuvo lugar bruscamente en torno a 200 ºC, lo cual se atribuyó a una transición de la

matriz de carbono (con una hibridación mayoritaria sp3) hacia una forma más similar a

la del grafito (una mayor hibridación sp2), como se verificó mediante EELS. Además,

se comprobó la estabilidad de las nanopartículas de Pt durante este proceso mediante

microscopía STEM.

La película de catalizador Pt-DLC consiguió ser promocionada mediante la

migración electroquímica de iones alcalinos K+ de forma reversible tanto en la

oxidación parcial de metanol (CH3OH/O2 = 11 %/0,9 %), obteniendo H2, CO, CO2 y

H2CO y H2O como productos de reacción, como en el reformado de metanol con

Descripción del trabajo realizado

36

vapor de agua (CH3OH/H2O = 4 %/4,8 %), produciendo en este caso H2, CO y CO2.

Al igual que en el capítulo anterior, el efecto de promoción de tipo electrofílico

obtenido en ambos casos al aplicar un potencial lo suficientemente negativo se

atribuyó a una mayor quimisorción del aceptor de electrones (O2 o H2O) frente al

donador de electrones (CH3OH). Sin embargo, en este estudio se observó un primer

incremento de la actividad catalítica, bajo condiciones de oxidación parcial, al

disminuir el potencial desde +2 V hasta tan solo +0,5 V. Este efecto sugirió la

formación de una nueva fase de promotor sobre las nanopartículas de Pt, y pareció

confirmarse mediante voltamperometría cíclica. Durante este procedimiento se

observaron varios picos catódicos y anódicos en atmósfera de POM, mientras que solo

un único pico en cada barrido en atmósfera de SRM, lo cual pareció indicar que

dependiendo de las condiciones de reacción y del potencial aplicado los iones K+

pueden formar diferentes fases promotoras sobre el catalizador.

Por último, comparando en condiciones de SRM los dos catalizadores

preparados por arco catódico (Pt y Pt-DLC), se pudo comprobar que las

nanopartículas dispersas de Pt presentaron una mayor actividad catalítica y un

incremento de esta (vía EPOC) de una magnitud similar al obtenido con el catalizador

denso de Pt. Este estudio abre, por tanto, una nueva posibilidad para la aplicación de la

promoción electroquímica en catalizadores dispersos de muy bajo contenido metálico

y soportados sobre un material no conductor iónico.

En la misma línea de empleo de películas de catalizadores dispersos, en el

Capítulo 3 se desarrolló un catalizador Au-YSZ/K-βAl2O3/Au que se basaba en

nanopartículas de Au soportadas en una matriz de zirconia estabilizada con itria

(YSZ), preparado en esta ocasión mediante pulverización catódica (“sputtering”). Este

electrodo fue depositado en colaboración con el Institut Jean Lamour de Université de

Lorraine (Francia). Concretamente se depositó mediante el “co-sputtering” simultáneo

de Au y Y-Zr en atmósfera de Ar y O2. La película de Au-YSZ se caracterizó

mediante diferentes técnicas (XRD, STEM, voltamperometría cíclica) y se evaluó su

actividad en la oxidación parcial catalítica de metanol (CH3OH/O2 = 5,9 %/0,43 %). A

diferencia del contraelectrodo y el electrodo de referencia de Au depositados en todos

Resultados obtenidos

37

los capítulos mediante la calcinación a elevada temperatura de una pasta orgánica de

este metal, el catalizador-electrodo de trabajo formado por nanopartículas de Au

dispersas en YSZ mostró cierta actividad catalítica y, en particular, una selectividad

muy elevada hacia formato de metilo (HCOOCH3).

Debido a que el Au-YSZ no presentaba conductividad eléctrica, se utilizó una

malla de Ag como colector eléctrico y se consiguió polarizar y promocionar

electroquímicamente el catalizador con iones alcalinos. Se llevó a cabo una serie de

experimentos galvanostáticos en los cuales, tras limpiar la superficie del catalizador de

iones K+ por medio de una polarización positiva (VWR = +2 V), se aplicó una

intensidad de corriente negativa de unos pocos µA, que suponía la migración

electroquímica de iones K+ desde el electrolito sólido hacia la película de catalizador

con un flujo constante determinado por la ley de Faraday, r = I/nF. Se realizaron

distintos experimentos en los que se varió esta intensidad negativa, y también otros en

los que la promoción electroquímica se llevó a cabo de modo potenciostático

(aplicando potenciales eléctricos constantes) como en los capítulos anteriores. En

estos últimos experimentos también se podía controlar la cantidad de promotor

suministrada, si bien en estos casos el flujo de K+ no era constante y estaba

determinado por el valor de intensidad obtenido en cada momento. Así pues, en todos

los experimentos realizados en las mismas condiciones de reacción se obtuvo la

misma cantidad óptima de promotor y las mismas velocidades máximas de producción

de los distintos compuestos, demostrando que la magnitud del efecto de promoción

electroquímica con electrolitos sólidos alcalinos no depende de la intensidad eléctrica

aplicada (es decir, del flujo de promotor) ni del modo de operación (potenciostático o

galvanostático), sino de la cantidad total de promotor adicionada.

Además, al interrumpir la corriente eléctrica en estos experimentos

(condiciones de circuito abierto), se observó un mantenimiento de las distintas

velocidades de producción obtenidas durante la polarización previa negativa,

reflejando una gran estabilidad de las especies promotoras en las condiciones de

reacción estudiadas y por tanto un efecto EPOC permanente. Solo aplicando de nuevo

un potencial positivo (VWR = +2 V) al final de cada experimento se conseguía eliminar

Descripción del trabajo realizado

38

estas especies del catalizador y restablecer la actividad catalítica inicial (no

promocionada). En definitiva, no solo se consiguió promocionar electroquímicamente

una película que no era conductora eléctrica, sino que además una primera

polarización podía ser suficiente para mantener la modificación de la actividad

catalítica durante un cierto periodo de tiempo. Por otro lado, se comprobó la

estabilidad de la película de Au-YSZ a lo largo de todos estos experimentos,

demostrando la enorme importancia del soporte de YSZ para dispersar nanopartículas

de metal y evitar su sinterización.

En el Capítulo 4 se aplicó por primera vez el fenómeno de la promoción

electroquímica a la reacción de oxidación parcial de metanol, empleando un metal no

noble (Cu) como catalizador. En esta ocasión, la película de catalizador-electrodo de

trabajo se depositó sobre el soporte de K-βAl2O3 en colaboración con el Instituto de

Ciencia de Materiales de Sevilla (CSIC - Universidad de Sevilla), mediante una

técnica de deposición física de vapor muy innovadora denominada “Deposición en

ángulo oblicuo” o “Deposición en ángulo rasante” (OAD o GLAD). Esta se basa en

hacer incidir el flujo evaporado del material a depositar (Cu) con un determinado

ángulo (α) sobre el sustrato (K-βAl2O3). De este modo, durante el crecimiento de la

película se produce un “efecto sombra” que causa la formación de nanocolumnas

inclinadas del metal y por tanto la obtención de una elevada porosidad y área

superficial. El ángulo de deposición α se varió durante el proceso de deposición, de

modo que en una primera etapa (α = 0º) se forzó el crecimiento de una película con

una buena conductividad eléctrica formada por nanocolumnas Cu perpendiculares al

electrolito sólido con una estructura compacta, y en una segunda etapa (α = 80º) se

formó una película mucho más porosa (en torno al 50 %) encima de la anterior,

mediante la deposición de nanocolumnas inclinadas de Cu. El objetivo de este

procedimiento fue conseguir un electrodo de trabajo de Cu que en su conjunto

presentara buenas propiedades tanto catalíticas como electrocatalíticas.

Mediante la realización de voltamperometrías cíclicas con el catalizador de

Cu/K-βAl2O3/Au se observó que un rango de potencial entre +1 y -1 era suficiente

para formar y descomponer las especies promotoras derivadas del potasio de forma

Resultados obtenidos

39

reversible y reproducible. Por tanto se realizaron transiciones potenciostáticas entre

estos dos valores en los experimentos de promoción electroquímica. La película de

cobre demostró ser activa en la oxidación parcial de metanol (CH3OH/O2 = 4,4 %/0,3

%) obteniendo H2 y CO2, además de H2O y HCOOCH3, y la velocidad de producción

de todos estos productos se vio incrementada al disminuir el potencial aplicado y

dopar electroquímicamente la superficie del catalizador con iones K+. Al igual que en

los capítulos anteriores, este efecto promocional electrofílico se explicó en base a un

fortalecimiento de la quimisorción de oxígeno que favorecería la deshidrogenación del

metanol y de las especies metóxido intermedias (la cual se suele considerar la etapa

limitante de este tipo de reacciones).

Tras los experimentos de promoción electroquímica el catalizador se sometió a

un acondicionamiento bajo condiciones de POM y una polarización negativa para

establecer un estado promocionado, y se caracterizó mediante diferentes técnicas ex-

situ (XRD, SEM-EDX, XPS). De este modo se confirmó que bajo condiciones de

promoción electroquímica se formaron especies superficiales derivadas de los iones

K+, probablemente carbonatos u óxidos. También se detectó la presencia de cierta

cantidad de carbono superficial (aunque no se observó una apreciable desactivación

durante los ensayos catalíticos) y de compuestos de nitrógeno que probablemente se

formaron bajo el efecto promocional del potasio a partir del N2 e H2 presentes en la

artmósfera de reacción.

Cabe destacar que a pesar de presentar esta película de Cu una actividad

catalítica menor que los catalizadores de Pt de capítulos anteriores, se obtuvo un

efecto promocional de magnitud similar. Esto, por un lado, puso de manifiesto el gran

interés que puede tener en este tipo de estudios la técnica de deposición en ángulo

oblicuo para la obtención de películas porosas, y por otro lado, motivó a continuar

estudiando el fenómeno EPOC con catalizadores no nobles.

En el Capítulo 5 se preparó un catalizador de Ni/K-βAl2O3/Au por arco

catódico de modo similar a la película fina de Pt del capítulo 1 y se aplicó el fenómeno

de la promoción electroquímica con iones K+ bajo diferentes condiciones de reacción:

descomposición de metanol (MD, CH3OH = 4,4 %), reformado de metanol con vapor

Descripción del trabajo realizado

40

de agua (SRM, CH3OH/H2O = 4,4 %/5,2 %) y oxidación parcial de metanol (POM,

CH3OH/O2 = 4,4 %/0,33 %). El efecto de los iones K+, unido a las características

particulares del Ni (una gran tendencia a la deposición de carbón y a la oxidación), dio

lugar a comportamientos catalíticos muy diferentes en las tres atmósferas de reacción

y posibilitó, en definitiva, el estudio de tres aplicaciones fundamentales del fenómeno

EPOC: la activación del catalizador, la modificación de la selectividad catalítica y la

oxidación parcial del catalizador.

Bajo condiciones de MD, la promoción del catalizador de Ni mediante la

migración electroquímica de iones K+ por medio de transiciones potenciostáticas

favoreció la deshidrogenación del metanol y de las especies intermedias para obtener

H2 y CO. En estas condiciones también tuvo lugar una fuerte desactivación por

deposición de carbono que además se acrecentó por efecto de los iones K+ a

potenciales demasiado negativos. Por otro lado, bajo condiciones de SRM la actividad

catalítica del Ni fue mayor y se obtuvo también CO2, aunque los principales productos

siguieron siendo H2 y CO. El suministro de iones K+ dio lugar a un incremento en la

velocidad de producción de los tres compuestos, lo cual se atribuyó a la activación del

agua como en los capítulos 1 y 2 (con catalizadores de Pt). Del mismo modo, este

efecto propició además una atenuación de la desactivación del Ni por deposición de

carbono bajo esta atmósfera de reacción.

Bajo condiciones de POM el Ni no promocionado (VWR = +2 V) presentó una

producción de H2 y CO superior a los casos anteriores. Sin embargo, la disminución

en el potencial aplicado y el fortalecimiento de la quimisorción del oxígeno en este

caso causó un descenso brusco de la actividad del Ni y de la selectividad hacia H2 y

CO, al mismo tiempo que aumentó ligeramente la selectividad hacia CO2, agua y

formaldehído. Este efecto negativo de los iones K+ parecía estar en contra del orden

positivo de la reacción, observado en estado no promocionado, con respecto a la

concentración de oxígeno (para valores bajos como el empleado en los experimentos

de EPOC). Finalmente se atribuyó a un incremento de la formación de NiO derivado

de una mayor adsorción de oxígeno de acuerdo con los resultados obtenidos mediante

XRD. Cabe señalar, no obstante, que el efecto promocional observado en todos los

Resultados obtenidos

41

casos se llevó a cabo de una forma controlada (variando tan solo el potencial aplicado)

y totalmente reversible, en base a una completa reproducibilidad de la actividad

catalítica del Ni en cada imposición del estado de referencia (VWR = +2 V). De este

modo se demostró la posibilidad de controlar la oxidación parcial del catalizador por

vía electroquímica, y con ello la modificación de su selectividad hacia la formación de

distintas especies.

Por último, en el Capítulo 6 se ha estudiado la posibilidad de aplicar el

fenómeno de la promoción electroquímica conjuntamente para la producción y el

almacenamiento de hidrógeno. La idea fue desarrollar un catalizador electroquímico

alcalino que presentara un rendimiento adecuado en la producción de hidrógeno

mediante reformado de metanol con vapor de agua (SRM), y que fuese capaz de

almacenarlo y liberarlo de forma controlada, operando bajo diferentes polarizaciones y

condiciones suaves de reacción. Para ello se seleccionó Ni como fase activa debido a

la buena actividad catalítica en la reacción SRM mostrada en el capitulo anterior. Se

prepararon diferentes películas de Ni sobre K-βAl2O3 mediante el método de

deposición en ángulo oblicuo (OAD) utilizado en el capítulo 4), variando el ángulo de

incidencia (α) del Ni sobre la K-βAl2O3: En un caso (catalizador Ni 0º) el flujo de

evaporación de Ni incidió perpendicularmente sobre el electrolito sólido (α = 0º)

dando lugar a nanocolumnas compactas y verticales de Ni, y en otro caso el

catalizador (Ni 80º) se depositó con un ángulo de inclinación de α = 80º , obteniendo a

una película formada por nanocolumnas de Ni inclinadas con una porosidad del 35 %.

Se llevó a cabo una serie de experimentos que constaban de las siguientes

cuatro etapas: 1) Aplicación de un potencial positivo (VWR = +2 V) bajo condiciones

de SRM (CH3OH/H2O = 4,4 %/5,2 %) para asegurar la ausencia de iones K+ sobre el

catalizador y establecer un estado de referencia; 2) Aplicación de una polarización

negativa bajo condiciones de SRM para trasladar iones K+ a la superficie de Ni y

promocionar la producción y captura de compuestos; 3) Purga del reactor con gas

inerte bajo circuito abierto; y 4) Aplicación de una polarización positiva en atmósfera

inerte para la retirada de los iones K+ del catalizador y la liberación de los compuestos.

Descripción del trabajo realizado

42

Unos experimentos preliminares realizados con el catalizador Ni 0º ya

reflejaron una liberación de H2 junto con CO y CO2 (en cantidades mucho menores)

durante la última etapa. También permitió optimizar el intervalo de los análisis y el

procedimiento de polarización durante esta etapa. Posteriormente se estudió la

influencia de la corriente negativa aplicada durante la etapa de almacenamiento, de la

microestructura del catalizador y de la atmósfera de reacción en la capacidad de

almacenamiento de H2 del catalizador y en la velocidad de liberación.

En todos los casos los catalizadores de Ni mostraron una actividad mucho

mayor en la producción de H2 y CO (MD, CH3OH → 2H2 + CO) frente a la

producción CO2 (SRM, CH3OH + H2O → 3H2 + CO2). También se observó cierta

desactivación de los catalizadores atribuida a la deposición de especies carbonosas

sobre la superficie. Esta podía estar además favorecida bajo polarización negativa por

la presencia de iones K+ y su efecto promocional en la disociación de CO.

En una primera etapa se plantearon los distintos posibles mecanismos de

captura de H2. Posteriormente, se fueron analizando y descartando algunos de ellos en

base a los resultados obtenidos y a diversas técnicas de caracterización realizadas. La

gran cantidad liberada de H2 observada en todos los casos en comparación con el CO,

el CO2 y los iones K+ transferidos electroquímicamente sugirió que el principal

mecanismo de almacenamiento no podía ser la formación directa de compuestos de

potasio: KH, KOH o bicarbonatos de potasio, ni la acumulación de intermedios de

reacción tipo metóxido sobre la superficie del catalizador. De este modo el

almacenamiento de H2 se atribuyó principalmente a la quimisorción de átomos de

hidrógeno procedentes de la conversión catalítica del metanol sobre centros activos

del niquel, y su posible acumulación posterior, a través de un proceso de spillover,

sobre otros adsorbentes como los depósitos carbonosos formados durante la reacción.

De hecho, la hipótesis de una captura de H2 ligada en gran medida a estos compuestos

carbonosos cobró una especial relevancia a la vista de los resultados obtenidos con el

Ni 80º, donde los moles de H2 liberados superaron en un orden de magnitud a la

cantidad de Ni presente en el catalizador, reflejando el papel fundamental de la

porosidad y área superficial del electrodo.

Resultados obtenidos

43

El Ni 80º presentó una mayor actividad catalítica y una capacidad de

almacenamiento de H2 muy superior al Ni 0º (19 vs 0,5 g H2 x 100 g Ni-1

), lo cual se

atribuyó a su mayor porosidad y facilidad para formar depósitos carbonosos. Esto

mostró la gran influencia de la microestructura del catalizador sobre el proceso de

producción/captura de H2, y por tanto la enorme utilidad del método de deposición en

ángulo oblicuo (OAD) en el diseño de este tipo de catalizadores. Por otro lado, el

aumento de la intensidad negativa en la etapa de almacenamiento causó un incremento

de la cantidad de H2 capturado (hasta cierto valor de saturación), atribuido a un doble

efecto promotor de los iones K+: en la formación de compuestos carbonosos sobre el

Ni, y en la adsorción de hidrógeno sobre estos compuestos.

En línea con el mecanismo de almacenamiento de H2 propuesto, unos

experimentos adicionales donde se emplearon diferentes atmósferas de reacción en la

etapa de polarización negativa demostraron el papel clave de los compuestos

carbonosos, ya que el hidrogeno capturado a partir de una corriente de H2 (en lugar de

metanol) fue despreciable. Por otro lado, la presencia de agua no supuso ninguna

diferencia en la capacidad de captura de H2, aunque sí aceleró su liberación.

Por último, el catalizador Ni/K-βAl2O3/Au se caracterizó mediante diferentes

técnicas ex-situ (SEM-EDX, XPS, FTIR, Raman) y se identificó óxido de grafeno

como principal compuesto carbonoso localizado sobre la superficie del catalizador.

Por tanto, el catalizador electroquímico de Ni desarrollado en este trabajo permite

producir H2 y óxido de grafeno a partir de metanol, almacenar el H2 con un elevado

rendimiento por unidad de área y cantidad de metal, y liberarlo con una elevada

pureza bajo las mismas condiciones de reacción (280 ºC); todo ello controlado vía

electroquímica, mediante un soporte conductor de iones K+.

Descripción del trabajo realizado

44

D. CONCLUSIONES Y RECOMENDACIONES

De los resultados obtenidos en la presente investigación se pueden obtener las

siguientes conclusiones:

- Una película catalítica de Pt con un espesor 120 nm se ha preparado

mediante el método de deposición por arco catódico (CAD) y se ha promocionado

electroquímicamente con iones K+ en las reacciones de oxidación parcial de metanol

(POM) y reformado de metanol con vapor de agua (SRM). En cambio, otro

catalizador de Pt preparado por impregnación presentó un exceso de actividad

electrocatalítica y el bloqueo de centros activos de Pt por compuestos derivados del

potasio. Por tanto, los demás catalizadores empleados en esta tesis se han preparado

mediante el método de CAD u otra técnica de deposición física de vapor (PVD), como

“sputtering” o OAD.

- La promoción electroquímica de la catálisis (EPOC) permite mejorar in-situ

la actividad del catalizador de Pt preparado por CAD y su selectividad hacia la

oxidación parcial de metanol (frente a la oxidación total), mediante la migración

controlada de iones K+ desde un electrolito sólido. Además, el efecto EPOC observado

con este catalizador y con todos los demás empleados en la tesis demostró ser

reversible al retirar los iones K+ del catalizador (bajo un potencial positivo).

- Se ha desarrollado una película de catalizador formada por nanopartículas de

Pt (~3 nm) dispersas en una matriz de carbono (Pt-DLC), preparada por CAD. Se ha

podido promocionar electroquímicamente tras un proceso de grafitización previo y ha

mostrado diferentes efectos promocionales del K+ en función del potencial aplicado y

de la atmósfera de reacción (POM o SRM). Esta película ha presentado una mayor

actividad catalítica y un efecto EPOC similar que el catalizador de Pt puro, incluso

con peores propiedades eléctricas y una menor carga metálica (solo 0,014 mg cm-2

).

- Se ha conseguido promocionar electroquímicamente mediante un colector de

corriente una película catalítica no conductora eléctrica, compuesta por nanopartículas

de Au dispersas en zirconia estabilizada con itria (Au-YSZ) mediante “co-sputtering”,

para la oxidación selectiva de metanol. Se ha optimizado la cantidad de iones

Conclusiones y recomendaciones

45

promotores K+ suministrada a este catalizador, demostrando además que el efecto

EPOC obtenido es permanente (hasta una nueva polarización positiva) y que su

magnitud no depende de la velocidad del suministro de los iones ni del modo de

operación (potenciostático o galvanostático) sino exclusivamente de la cantidad total

suministrada.

- Es posible promocionar electroquímicamente películas de metales no nobles

como Cu en la reacción de POM con razones de incremento de la actividad catalítica

comparables a las obtenidas con el catalizador de Pt. Este catalizador de Cu se ha

preparado mediante una novedosa modificación del método de PVD que consiste en la

deposición en ángulo oblicuo (OAD) de nanocolumnas del metal, obteniendo películas

conductoras con una elevada porosidad muy útiles para aplicaciones catalíticas o

electrocatalíticas.

- Se ha demostrado la posible aplicación de la promoción electroquímica con

iones K+ para activar un catalizador metálico, modificar su selectividad, y controlar

parcialmente su estado de oxidación, empleando una misma película de Ni (praparada

por CAD). En atmósfera de metanol y de metanol+agua el suministro de iones K+

ocasiona la activación del mecanismo de deshidrogenación del metanol, y en este

último caso también una atenuación de la desactivación del catalizador por deposición

de carbono. Sin embargo, en atmósfera de metanol+oxígeno el fortalecimiento de la

quimisorción del oxígeno por efecto del K+ deriva en la formación de NiO, y

consecuentemente en la disminución de la selectividad hacia H2 y CO y el aumento de

la selectividad hacia CO2, H2O y H2CO.

- Se ha desarrollado un catalizador poroso de Ni depositado en ángulo oblicuo

que permite producir simultáneamente H2 y óxido de grafeno a partir de metanol, y

capturar parte del H2 sobre el Ni y el óxido de grafeno (que también actúa como

adsorbente) bajo la acción de iones K+. Además, este H2 también se puede liberar de

forma controlada vía electroquímica bajo las mismas condiciones de reacción (280

ºC), obteniendo valores de almacenamiento reversible de H2 muy elevados (19 g H2 x

100 g Ni-1

). Estos resultados abren la puerta a una nueva aplicación del fenómeno

EPOC en el campo de la producción y el almacenamiento de hidrógeno.

Descripción del trabajo realizado

46

Con objeto de completar los resultados obtenidos y continuar esta

investigación, se recomienda:

- Realizar estudios de promoción electroquímica de las reacciones de

conversión de metanol con otros tipos de electrolitos sólidos, catiónicos (Na-βAl2O3,

NASICON, SCY), aniónicos (YSZ) o conductores mixtos (tipo perovskita), y

comparar los resultados con los obtenidos en esta tesis con K-βAl2O3.

- Estudiar la promoción electroquímica de reacciones de producción de

hidrógeno empleando otras materias primas como etanol o bioalcoholes (mezclas

reales).

- Aplicar algunas de las técnicas de deposición de películas delgadas expuestas

en esta tesis (evaporación por bombardeo electrónico, pulverización catódica) para la

preparación de catalizadores soportados de Cu o Ni similares a los empleados en

catálisis convencional (Ni/Al2O3, Cu/Al2O3, Ni/CeO2, Cu/CeO2), u otros tipos de

catalizadores, como aleaciones (NiPt, NiCu) o configuraciones multicapa.

- Investigar otras técnicas para la preparación películas metálicas con una

elevada dispersión, como por ejemplo algún modo de prensar el catalizador en forma

de polvo sobre el electrolito sólido.

- Realizar estudios de promoción electroquímica de reacciones de conversión

de metanol en reactores de doble cámara, bien en configuración de pellet, o bien en

celda de combustible.

- Escalar los experimentos de promoción electroquímica de producción y/o

almacenamiento de hidrógeno, empleando configuraciones de mayor área de

electrodo, como las tubulares o monolíticas, que permitan obtener mayores valores de

conversión de metanol y/o de capacidad de almacenamiento de hidrógeno.

Bibliografía

47

E. BIBLIOGRAFÍA

[1] J. Rifkin, La economía del hidrógeno. La creacción de la red energética mundial y la

redistribución del poder en la tierra, Paidos Iberica, Barcelona, 2002.

[2] J.N. Armor, Catalysis and the hydrogen economy, Catalysis Letters, 101 (2005) 131-135.

[3] S. Dutta, A review on production, storage of hydrogen and its utilization as an energy

resource, Journal of Industrial and Engineering Chemistry, 20 (2014) 1148-1156.

[4] L. Schlapbach, A. Züttel, Hydrogen-storage materials for mobile applications, Nature, 414

(2001) 353-358.

[5] J.R. Jennings, Catalytic Ammonia Synthesis: Fundamentals and Practice, Plenum Press,

New York, 1991.

[6] J.P. Lange, Methanol synthesis: A short review of technology improvements, Catalysis

Today, 64 (2001) 3-8.

[7] H. Jahangiri, J. Bennett, P. Mahjoubi, K. Wilson, S. Gu, A review of advanced catalyst

development for Fischer-Tropsch synthesis of hydrocarbons from biomass derived syn-gas,

Catalysis Science and Technology, 4 (2014) 2210-2229.

[8] M. Carmo, D.L. Fritz, J. Mergel, D. Stolten, A comprehensive review on PEM water

electrolysis, International Journal of Hydrogen Energy, 38 (2013) 4901-4934.

[9] M.A. Laguna-Bercero, Recent advances in high temperature electrolysis using solid oxide

fuel cells: A review, Journal of Power Sources, 203 (2012) 4-16.

[10] J.D. Holladay, J. Hu, D.L. King, Y. Wang, An overview of hydrogen production

technologies, Catalysis Today, 139 (2009) 244-260.

[11] R.M. Navarro, M.A. Peña, J.L.G. Fierro, Hydrogen production reactions from carbon

feedstocks: Fossil fuels and biomass, Chemical Reviews, 107 (2007) 3952-3991.

[12] R. Hughes, Deactivation of catalysts, Academic Press, New York, 1984.

[13] D.R. Palo, R.A. Dagle, J.D. Holladay, Methanol steam reforming for hydrogen

production, Chemical Reviews, 107 (2007) 3992-4021.

[14] P.O. Graf, B.L. Mojet, L. Lefferts, Influence of potassium on the competition between

methane and ethane in steam reforming over Pt supported on yttrium-stabilized zirconia,

Applied Catalysis A: General, 346 (2008) 90-95.

[15] F. Frusteri, S. Freni, V. Chiodo, L. Spadaro, G. Bonura, S. Cavallaro, Potassium improved

stability of Ni/MgO in the steam reforming of ethanol for the production of hydrogen for

MCFC, Journal of Power Sources, 132 (2004) 139-144.

Descripción del trabajo realizado

48

[16] C. Song, Fuel processing for low-temperature and high-temperature fuel cells:

Challenges, and opportunities for sustainable development in the 21st century, Catalysis

Today, 77 (2002) 17-49.

[17] G. Nahar, V. Dupont, Recent advances in hydrogen production via autothermal reforming

process (ATR): A review of patents and research articles, Recent Patents on Chemical

Engineering, 6 (2013) 8-42.

[18] M. Usman, W.M.A. Wan Daud, H.F. Abbas, Dry reforming of methane: Influence of

process parameters - A review, Renewable and Sustainable Energy Reviews, 45 (2015) 710-

744.

[19] United States Department of Energy (DOE),

http://energy.gov/sites/prod/files/2014/03/f11/targets_onboard_hydro_storage_explanation.pdf,

[20] A.F. Dalebrook, W. Gan, M. Grasemann, S. Moret, G. Laurenczy, Hydrogen storage:

Beyond conventional methods, Chemical Communications, 49 (2013) 8735-8751.

[21] A.W.C. Van Den Berg, C.O. Areán, Materials for hydrogen storage: Current research

trends and perspectives, Chemical Communications, (2008) 668-681.

[22] S. Tibdewal, U. Saxena, A.V.P. Gurumoorthy, Hydrogen economy vs. Methanol economy,

International Journal of Chemical Sciences, 12 (2014) 1478-1486.

[23] M. Paleologou, T. Radiotis, L. Kouisni, N. Jemaa, T. Mahmood, T. Browne, D. Singbeil,

New and emerging biorefinery technologies and products for the canadian forest industry, J-

FOR, 1 (2011) 6-14.

[24] E.L. Fornero, P.B. Sanguineti, D.L. Chiavassa, A.L. Bonivardi, M.A. Baltanás,

Performance of ternary Cu-Ga2O3-ZrO2 catalysts in the synthesis of methanol using CO2-rich

gas mixtures, Catalysis Today, 213 (2013) 163-170.

[25] C.A. Trudewind, A. Schreiber, D. Haumann, Photocatalytic methanol and methane

production using captured CO2 from coal-fired power plants. Part i - A Life Cycle Assessment,

Journal of Cleaner Production, 70 (2014) 27-37.

[26] J.M. Tatibouët, Methanol oxidation as a catalytic surface probe, Applied Catalysis A:

General, 148 (1997) 213-252.

[27] S.T. Yong, C.W. Ooi, S.P. Chai, X.S. Wu, Review of methanol reforming-Cu-based

catalysts, surface reaction mechanisms, and reaction schemes, International Journal of

Hydrogen Energy, 38 (2013) 9541-9552.

[28] S. Sá, H. Silva, L. Brandão, J.M. Sousa, A. Mendes, Catalysts for methanol steam

reforming-A review, Applied Catalysis B: Environmental, 99 (2010) 43-57.

Bibliografía

49

[29] A. Guerrero-Ruiz, I. Rodriguez-Ramos, J.L.G. Fierro, Dehydrogenation of methanol to

methyl formate over supported copper catalysts, Applied Catalysis, 72 (1991) 119-137.

[30] K. Klier, Preparation of bifunctonal catallysts, Catalysis Today, 15 (1992) 361-382.

[31] L. Alejo, R. Lago, M.A. Peña, J.L.G. Fierro, Partial oxidation of methanol to produce

hydrogen over Cu-Zn-based catalysts, Applied Catalysis A: General, 162 (1997) 281-297.

[32] M. Kusche, F. Enzenberger, S. Bajus, H. Niedermeyer, A. Bösmann, A. Kaftan, M.

Laurin, J. Libuda, P. Wasserscheid, Enhanced activity and selectivity in catalytic methanol

steam reforming by basic alkali metal salt coatings, Angewandte Chemie - International

Edition, 52 (2013) 5028-5032.

[33] H.N. Evin, G. Jacobs, J. Ruiz-Martinez, U.M. Graham, A. Dozier, G. Thomas, B.H.

Davis, Low temperature water-gas shift/methanol steam reforming: Alkali doping to facilitate

the scission of formate and methoxy C-H bonds over Pt/ceria catalyst, Catalysis Letters, 122

(2008) 9-19.

[34] J.M. Pigos, C.J. Brooks, G. Jacobs, B.H. Davis, Low temperature water-gas shift: The

effect of alkali doping on the Csingle bondH bond of formate over Pt/ZrO2 catalysts,

Applied Catalysis A: General, 328 (2007) 14-26.

[35] C.G. Vayenas, S. Bebelis, C. Pliangos, S. Brosda, D. Tsiplakides, Electrochemical

Activation of Catalysis: Promotion, Electrochemical Promotion and Metal-Support

Interactions, Kluwer Academic Publishers/Plenum Press, New York, 2001.

[36] M. Stoukides, C.G. Vayenas, The effect of electrochemical oxygen pumping on the rate

and selectivity of ethylene oxidation on polycrystalline silver, Journal of Catalysis, 70 (1981)

137-146.

[37] C.G. Vayenas, S. Bebelis, S. Ladas, Dependence of catalytic rates on catalyst work

function, Nature, 343 (1990) 625-627.

[38] P. Vernoux, L. Lizarraga, M.N. Tsampas, F.M. Sapountzi, A. De Lucas-Consuegra, J.L.

Valverde, S. Souentie, C.G. Vayenas, D. Tsiplakides, S. Balomenou, E.A. Baranova, Ionically

conducting ceramics as active catalyst supports, Chemical Reviews, 113 (2013) 8192-8260.

[39] C. Vayenas, S. Brosda, Electrochemical promotion: Experiment, rules and mathematical

modeling, Solid State Ionics, 154-155 (2002) 243-250.

[40] C.G. Vayenas, M.M. Jaksic, S. Bebelis, S. Neophytides, Modern Aspects of

Electrochemistry, Plenum, New York, 1996.

[41] I.M. Petrushina, V.A. Bandur, F. Cappeln, N.J. Bjerrum, Electrochemical promotion of

sulfur dioxide catalytic oxidation, Journal of the Electrochemical Society, 147 (2000) 3010-

3010.

Descripción del trabajo realizado

50

[42] S. Neophytides, C.G. Vayenas, Non-faradaic electrochemical modification of catalytic

activity. 2. The case of methanol dehydrogenation and decomposition on Ag, Journal of

Catalysis, 118 (1989) 147-163.

[43] A. de Lucas-Consuegra, F. Dorado, C. Jiménez-Borja, J.L. Valverde, Influence of the

reaction conditions on the electrochemical promotion by potassium for the selective catalytic

reduction of N2O by C3H6 on platinum, Applied Catalysis B: Environmental, 78 (2008) 222-

231.

[44] E. Varkaraki, J. Nicole, E. Plattner, C. Comninellis, C.G. Vayenas, Electrochemical

promotion of IrO2 catalyst for the gas phase combustion of ethylene, Journal of Applied

Electrochemistry, 25 (1995) 978-981.

[45] M. Stoukides, C.G. Vayenas, ELECTROCATALYTIC RATE ENHANCEMENT OF

PROPYLENE EPOXIDATION ON POROUS SILVER ELECTRODES USING A ZIRCONIA

OXYGEN PUMP, 131 (1984) 839-845 jesoan.

[46] F.J. Williams, A. Palermo, M.S. Tikhov, R.M. Lambert, First Demonstration of in Situ

Electrochemical Control of a Base Metal Catalyst: Spectroscopic and Kinetic Study of the CO

+ NO Reaction over Na-Promoted Cu, Journal of Physical Chemistry B, 103 (1999) 9960-

9966.

[47] G.E. Pitselis, P.D. Petrolekas, C.G. Vayenas, Electrochemical promotion of ammonia

decomposition over Fe catalyst films interfaced with K+- & H+- conductors, Ionics, 3 (1997)

110-116.

[48] C.G. Vayenas, S. Neophytides, Non-faradaic electrochemical modification of catalytic

activity III. The case of methanol oxidation on Pt, Journal of Catalysis, 127 (1991) 645-664.

[49] N. Li, F. Gaillard, A. Boréave, Electrochemical promotion of Ag catalyst for the low

temperature combustion of toluene, Catalysis Communications, 9 (2008) 1439-1442.

[50] C.A. Cavalca, G.L. Haller, Solid electrolytes as active catalyst supports: Electrochemical

modification of benzene hydrogenation activity on Pt/β″(Na)Al2O3, Journal of Catalysis, 177

(1998) 389-395.

[51] F. Gaillard, N. Li, Electrochemical promotion of toluene combustion on an inexpensive

metallic catalyst, Catalysis Today, 146 (2009) 345-350.

[52] P. Vernoux, F. Gaillard, C. Lopez, E. Siebert, Coupling catalysis to electrochemistry: A

solution to selective reduction of nitrogen oxides in lean-burn engine exhausts?, Journal of

Catalysis, 217 (2003) 203-208.

Bibliografía

51

[53] M. Ouzounidou, A. Skodra, C. Kokkofitis, M. Stoukides, Catalytic and electrocatalytic

synthesis of NH3 in a H+ conducting cell by using an industrial Fe catalyst, Solid State Ionics,

178 (2007) 153-159.

[54] L. Ploense, M. Salazar, B. Gurau, E.S. Smotkin, Proton spillover promoted isomerization

of n-butylenes on Pd-black cathodes/nafion 117 [8], Journal of the American Chemical

Society, 119 (1997) 11550-11551.

[55] A. De Lucas-Consuegra, A. Caravaca, P.J. Martínez, J.L. Endrino, F. Dorado, J.L.

Valverde, Development of a new electrochemical catalyst with an electrochemically assisted

regeneration ability for H2 production at low temperatures, Journal of Catalysis, 274 (2010)

251-258.

[56] C. Jiménez-Borja, F. Dorado, A. de Lucas-Consuegra, J.M. García-Vargas, J.L. Valverde,

Complete oxidation of methane on Pd/YSZ and Pd/CeO2/YSZ by electrochemical promotion,

Catalysis Today, 146 (2009) 326-329.

[57] T.I. Politova, V.A. Sobyanin, V.D. Belyaev, Ethylene hydrogenation in electrochemical

cell with solid proton-conducting electrolyte, Reaction Kinetics & Catalysis Letters, 41 (1990)

321-326.

[58] A. de Lucas-Consuegra, A. Princivalle, A. Caravaca, F. Dorado, C. Guizard, J.L.

Valverde, P. Vernoux, Preferential CO oxidation in hydrogen-rich stream over an

electrochemically promoted Pt catalyst, Applied Catalysis B: Environmental, 94 (2010) 281-

287.

[59] D. Tsiplakides, S.G. Neophytides, O. Enea, M. Jaksic, C.G. Vayenas, Nonfaradaic

electrochemical modification of the catalytic activity of Pt-black electrodes deposited on nafion

117 solid polymer electrolytes, Journal of the Electrochemical Society, 144 (1997) 2072-2078.

[60] E.I. Papaioannou, S. Souentie, A. Hammad, C.G. Vayenas, Electrochemical promotion of

the CO2 hydrogenation reaction using thin Rh, Pt and Cu films in a monolithic reactor at

atmospheric pressure, Catalysis Today, 146 (2009) 336-344.

[61] A. De Lucas-Consuegra, A. Caravaca, J. González-Cobos, J.L. Valverde, F. Dorado,

Electrochemical activation of a non noble metal catalyst for the water-gas shift reaction,

Catalysis Communications, 15 (2011) 6-9.

[62] P.J. Gellings, H.J.M. Bouwmeester, The CRC Handbook of Solid State Electrochemistry,

CRC Press1997.

[63] A. de Lucas-Consuegra, New trends of Alkali Promotion in Heterogeneous Catalysis:

Electrochemical Promotion with Alkaline Ionic Conductors, Catalysis Surveys from Asia,

2015, DOI:10.1007/s10563-014-9179-6.

Descripción del trabajo realizado

52

[64] C.G. Vayenas, S. Bebelis, M. Despotopoulou, Non-faradaic electrochemical modification

of catalytic activity 4. The use of β″-Al2O3 as the solid electrolyte, Journal of Catalysis, 128

(1991) 415-435.

[65] I.R. Harkness, R.M. Lambert, Electrochemical Promotion of the No + Ethylene Reaction

over Platinum, Journal of Catalysis, 152 (1995) 211-214.

[66] M. Makri, G.G. Vayenas, S. Bebelis, K.H. Besocke, C. Cavalca, Atomic resolution STM

imaging of electrochemically controlled reversible promoter dosing of catalysts, Surface

Science, 369 (1996) 351-359.

[67] F.J. Williams, C.M. Aldao, A. Palermo, R.M. Lambert, A Monte Carlo simulation of the

NO + CO reaction on Na-promoted platinum, Surface Science, 412-413 (1998) 174-183.

[68] C.G. Vayenas, S. Brosda, C. Pliangos, Rules and mathematical modeling of

electrochemical and chemical promotion: 1. Reaction classification and promotional rules,

Journal of Catalysis, 203 (2001) 329-350.

[69] F.J. Williams, A. Palermo, M.S. Tikhov, R.M. Lambert, Mechanism of alkali promotion in

heterogeneous catalysis under realistic conditions: Application of electron spectroscopy and

electrochemical promotion to the reduction of NO by CO and by propene over rhodium,

Surface Science, 482-485 (2001) 177-182.

[70] A.J. Urquhart, F.J. Williams, R.M. Lambert, Electrochemical promotion by potassium of

Rh-catalysed fischer-tropsch synthesis at high pressure, Catalysis Letters, 103 (2005) 137-141.

[71] A. de Lucas-Consuegra, F. Dorado, J.L. Valverde, R. Karoum, P. Vernoux, Low-

temperature propene combustion over Pt/K-βAl2O3 electrochemical catalyst:

Characterization, catalytic activity measurements, and investigation of the NEMCA effect,

Journal of Catalysis, 251 (2007) 474-484.

[72] A. de Lucas-Consuegra, Á. Caravaca, P. Sánchez, F. Dorado, J.L. Valverde, A new

improvement of catalysis by solid-state electrochemistry: An electrochemically assisted NOx

storage/reduction catalyst, Journal of Catalysis, 259 (2008) 54-65.

[73] E. Ruiz, D. Cillero, P.J. Martínez, Á. Morales, G.S. Vicente, G. De Diego, J.M. Sánchez,

Bench scale study of electrochemically promoted catalytic CO2 hydrogenation to renewable

fuels, Catalysis Today, 210 (2013) 55-66.

[74] B. Lin, K. Wei, X. Ma, J. Lin, J. Ni, Study of potassium promoter effect for Ru/AC

catalysts for ammonia synthesis, Catalysis Science and Technology, 3 (2013) 1367-1374.

[75] E. Ruiz, D. Cillero, P.J. Martínez, Á. Morales, G.S. Vicente, G. De Diego, J.M. Sánchez,

Electrochemical synthesis of fuels by CO2 hydrogenation on Cu in a potassium ion conducting

membrane reactor at bench scale, Catalysis Today, 236 (2014) 108-120.

Bibliografía

53

[76] M. Marwood, C.G. Vayenas, Electrochemical promotion of a dispersed platinum catalyst,

Journal of Catalysis, 178 (1998) 429-440.

[77] C.G. Yiokari, G.E. Pitselis, D.G. Polydoros, A.D. Katsaounis, C.G. Vayenas, High-

pressure electrochemical promotion of ammonia synthesis over an industrial iron catalyst,

Journal of Physical Chemistry A, 104 (2000) 10600-10602.

[78] S. Balomenou, D. Tsiplakides, A. Katsaounis, S. Thiemann-Handler, B. Cramer, G. Foti,

C. Comninellis, C.G. Vayenas, Novel monolithic electrochemically promoted catalytic reactor

for environmentally important reactions, Applied Catalysis B: Environmental, 52 (2004) 181-

196.

[79] S.P. Balomenou, D. Tsiplakides, C.G. Vayenas, S. Poulston, V. Houel, P. Collier, A.G.

Konstandopoulos, C. Agrafiotis, Electrochemical promotion in a monolith electrochemical

plate reactor applied to simulated and real automotive pollution control, Topics in Catalysis,

44 (2007) 481-486.

[80] S. Souentie, A. Hammad, S. Brosda, G. Foti, C.G. Vayenas, Electrochemical promotion of

NO reduction by C2H4 in 10% O2 using a monolithic electropromoted reactor with Rh/YSZ/Pt

elements, Journal of Applied Electrochemistry, 38 (2008) 1159-1170.

[81] D. Poulidi, M.E. Rivas, B. Zydorczak, Z. Wu, K. Li, I.S. Metcalfe, Electrochemical

promotion of a Pt catalyst supported on La 0.6Sr 0.4Co 0.2Fe 0.8O 3 - δ hollow fibre

membranes, Solid State Ionics, 225 (2012) 382-385.

[82] A. Kambolis, L. Lizarraga, M.N. Tsampas, L. Burel, M. Rieu, J.P. Viricelle, P. Vernoux,

Electrochemical promotion of catalysis with highly dispersed Pt nanoparticles,

Electrochemistry Communications, 19 (2012) 5-8.

[83] V. Jiménez, C. Jiménez-Borja, P. Sánchez, A. Romero, E.I. Papaioannou, D. Theleritis, S.

Souentie, S. Brosda, J.L. Valverde, Electrochemical promotion of the co2 hydrogenation

reaction on composite ni or ru impregnated carbon nanofiber catalyst-electrodes deposited on

YSZ, Applied Catalysis B: Environmental, 107 (2011) 210-220.

[84] A.C. Kaloyannis, C.A. Pliangos, D.T. Tsiplakides, I.V. Yentekakis, S.G. Neophytides, S.

Bebelis, C.G. Vayenas, Electrochemical promotion of catalyst surfaces deposited on ionic and

mixed conductors, Ionics, 1 (1995) 414-420.

[85] E. Ruiz, D. Cillero, P.J. Martínez, Á. Morales, G.S. Vicente, G. De Diego, J.M. Sánchez,

Bench-scale study of electrochemically assisted catalytic CO2 hydrogenation to hydrocarbon

fuels on Pt, Ni and Pd films deposited on YSZ, Journal of CO2 Utilization, 8 (2014) 1-20.

[86] A. de Lucas-Consuegra, A. Princivalle, A. Caravaca, F. Dorado, A. Marouf, C. Guizard,

J.L. Valverde, P. Vernoux, Preparation and characterization of a low particle size Pt/C

Descripción del trabajo realizado

54

catalyst electrode for the simultaneous electrochemical promotion of CO and C3H6 oxidation,

Applied Catalysis A: General, 365 (2009) 274-280.

[87] A. Lintanf, E. Djurado, P. Vernoux, Pt/YSZ electrochemical catalysts prepared by

electrostatic spray deposition for selective catalytic reduction of NO by C3H6, Solid State

Ionics, 178 (2008) 1998-2008.

[88] R.M. Lambert, A. Palermo, F.J. Williams, M.S. Tikhov, Electrochemical promotion of

catalytic reactions using alkali ion conductors, Solid State Ionics, 136-137 (2000) 677-685.

[89] V. Roche, R. Revel, P. Vernoux, Electrochemical promotion of YSZ monolith honeycomb

for deep oxidation of methane, Catalysis Communications, 11 (2010) 1076-1080.

[90] F.M. Sapountzi, M.N. Tsampas, C.G. Vayenas, Methanol reformate treatment in a PEM

fuel cell-reactor, Catalysis Today, 127 (2007) 295-303.

[91] G. Karagiannakis, S. Zisekas, M. Stoukides, Hydrogenation of carbon dioxide on copper

in a H+ conducting membrane-reactor, Solid State Ionics, 162-163 (2003) 313-318.

[92] I.V. Yentekakis, Y. Jiang, S. Neophytides, S. Bebelis, C.G. Vayenas, Catalysis,

electrocatalysis and electrochemical promotion of the steam reforming of methane over Ni film

and Ni-YSZ cermet anodes, Ionics, 1 (1995) 491-498.

[93] A. Caravaca, A. De Lucas-Consuegra, C. Molina-Mora, J.L. Valverde, F. Dorado,

Enhanced H2 formation by electrochemical promotion in a single chamber steam electrolysis

cell, Applied Catalysis B: Environmental, 106 (2011) 54-62.

[94] J.K. Hong, I.H. Oh, S.A. Hong, W.Y. Lee, Electrochemical oxidation of methanol over a

silver electrode deposited on yttria-stabilized zirconia electrolyte, Journal of Catalysis, 163

(1996) 95-105.

[95] C.A. Cavalca, G. Larsen, C.G. Vayenas, G.L. Haller, Electrochemical modification of

CH3OH oxidation selectivity and activity on a Pt single-pellet catalytic reactor, Journal of

Physical Chemistry, 97 (1993) 6115-6119.

[96] A. de Lucas-Consuegra, A. Caravaca, M.J. Martín de Vidales, F. Dorado, S. Balomenou,

D. Tsiplakides, P. Vernoux, J.L. Valverde, An electrochemically assisted NOx

storage/reduction catalyst operating under fixed lean burn conditions, Catalysis

Communications, 11 (2009) 247-251.

[97] E. Ruiz, D. Cillero, Á. Morales, G.S. Vicente, G. De Diego, P.J. Martínez, J.M. Sánchez,

Bench scale study of electrochemically promoted CO2 capture on Pt/K-βAl2O3,

Electrochimica Acta, 112 (2013) 967-975.

55

ABSTRACT

The present work is part of a wider research program dealing with the

application of electrocatalytic systems in environmental and energy processes. It is

being conducted for the last years in the Laboratory of Catalysis and Materials at the

Department of Chemical Engineering at the University of Castilla-La Mancha

(UCLM).

In particular, this doctoral thesis aims at the study of the phenomenon of

electrochemical promotion of catalysis in hydrogen production processes from

methanol by using alkaline conductors. The present work has been financially

supported by the Spanish Government through the research projects CTQ 2010-

16179/PQ and CTQ 2013-45030-R.

This thesis has been carried out in collaboration with the Institute of Materials

Science of Madrid (CSIC), the Institute Jean Lamour at University of Lorraine

(France) and the Institute of Materials Science of Seville (CSIC-University of Seville).

Abstract

56

Hydrogen is a very important feedstock in the chemical industry and a

promising energy carrier with main application in internal combustion engines and

fuel cell technology as an alternative to the massive consumption of fossil fuels. H2

presents a high gravimetric energy density and can be considered as a clean synthetic

fuel depending on the sustainability of the energy and raw material employed for its

production. Hydrogen is currently obtained mainly via methane steam reforming.

However, the use of liquid hydrogen carriers such as methanol is acquiring increasing

interest.

H2 production from methanol is typically carried out through its

decomposition (MD, CH3OH → 2H2 + CO), steam reforming (SRM, CH3OH + H2O

→ 3H2 + CO2) or partial oxidation (POM, CH3OH + 1/2O2 → 2H2 + CO2), by using

catalysts based on Cu or Group VIII metals (Pt, Pd, Ni). A high catalytic activity, a

low CO selectivity and a good durability of the catalyst are the main targets. Other

valuable byproducts such as formaldehyde or methyl formate can also be obtained

from these processes. On the other hand, the low volumetric energy density of gaseous

H2 makes the development of efficient hydrogen storage systems to be of paramount

importance.

The coupling of catalysis and electrochemistry has been widely investigated

for the last years for H2 production through electrolysis processes. In this thesis, the

electrochemistry has been used to activate and tune different catalysts for H2

production from methanol through the Electrochemical Promotion Of Catalysis

(EPOC). This phenomenon is based on the modification of the chemisorption

properties of a metal catalyst by the electrochemical migration of promoter ions from

a solid electrolyte support (via application of an electric current or potential). Hence,

while in classical chemical promotion a specific amount of a promoter is added during

the preparation step of the catalyst, in the case of the electrochemical promotion,

promoter ions are electrochemically pumped between the metal catalyst and the solid

electrolyte in a controlled way during the reaction step. Then, the electrochemical

promotion presents several additional advantages, such as the possibility of optimizing

Abstract

57

the promoter coverage on the catalyst surface at different reaction conditions and the

in-situ enhancement of the catalytic activity and selectivity.

In this sense, previous studies have reported the interest of the phenomenon of

electrochemical promotion on the methanol decomposition and partial oxidation

reactions over Pt and Ag catalysts, by using yttria-stabilized zirconia (YSZ) as the

solid electrolyte, which is an O2-

conductor. However, these works were only focused

on formaldehyde production (not H2) and were carried out at relatively high reaction

temperatures. With respect to the application of the EPOC phenomenon to the

trapping and storage of surface compounds, previous works have demonstrated the

interest of using cationic electrochemical catalysts for NOx storage-reduction process

(NSR) and CO2 capture.

On the other hand, although a large number of studies have demonstrated the

electrochemical promotion mechanism and this has been applied in a wide variety of

catalytic reactions of industrial and environmental interest, a further technological

progress is necessary for its practical application. Nowadays, some of the main

challenges of EPOC are the development of compact and efficient reactors, and more

competitive catalysts with a higher dispersion (composed of metal nanoparticles) or

based in non noble metals such as Ni or Cu.

Hence, in view of the mentioned above, this doctoral thesis aims to the study

of the phenomenon of electrochemical promotion of catalysis (EPOC) in H2

production processes from methanol by using alkaline conductors. Furthermore, the

possibility of applying the EPOC in the field of H2 storage was also investigated.

Firstly, the effect of electrochemical promotion was studied in the H2

production via steam reforming and partial oxidation of methanol, being formaldehyde

simultaneously obtained in the latter case. For this purpose, two kinds of Pt catalyst

films were deposited on K-βAl2O3 (K+-conductor material) by different techniques:

impregnation and cathodic arc deposition (CAD). Both electrochemical catalysts were

compared under electrochemical promotion conditions for the methanol partial

oxidation reaction. It was found that in alkali-based solid electrolyte cells, an excess of

Abstract

58

electrocatalytic activity of the catalyst-working electrode (the case of the impregnated

Pt film) can be detrimental for the electropromotional effect due to the excessive

formation of promoter-derived surface compounds, which may block catalytic active

sites. On the other hand, the catalytic activity and selectivity of the thin Pt film of low

metal loading prepared by cathodic arc deposition were strongly enhanced by EPOC,

i.e., by the electrochemical transfer of K+ ions to the catalyst under negative potentials.

In this way, the Pt catalyst film prepared by this technique allowed to produce both H2

and H2CO at mild reaction conditions in a single reaction step. This electrochemical

catalyst presented much less catalytic activity in the steam reforming of methanol,

although it was also electrochemically promoted under these reaction conditions. The

K+ promotional effect was attributed to the decrease in the catalyst work function and,

hence, the strengthening of the Pt chemical bond with the electron acceptor (O2 or

H2O) against that with the electron donor (CH3OH). Furthermore, the Pt catalyst film

showed to be stable for long working times and the EPOC effect was found to be fully

reversible in all the experiments, given the good reproducibility observed in the

catalytic activity under every positive polarization (unpromoted or reference state). On

the basis of the obtained results, the catalyst films used in the following studies were

prepared by cathodic arc deposition or by other kind of physical vapor deposition

(PVD) technique, such as sputtering or oblique angle deposition (OAD).

Then, in other study, the phenomenon of electrochemical promotion (EPOC)

was applied on a catalyst film composed of Pt nanoparticles (of around 3 nm)

dispersed in a diamond-like carbon (DLC) matrix which was prepared by the cathodic

arc deposition technique. In first place, a temperature-programmed pretreatment was

carried out in order to achieve a suitable electrical conductivity for the electrochemical

promotion experiments. The decrease in the surface electrical resistance was due to

the transition of the sp3-hybridized carbon form into a more graphitic structure (sp

2-

hybridized) as confirmed by EELS. Moreover, the stability of the Pt nanoparticles was

verified by STEM. The catalytic performance of the Pt-DLC film in the methanol

partial oxidation (POM) and steam reforming (SRM) reactions for H2 production was

promoted by K+ ions electrochemically transferred from a K-βAl2O3 solid electrolyte.

Abstract

59

Hence, it was demonstrated that the EPOC phenomenon may be applied to catalysts

based on metal nanoparticles dispersed in an electronic non-ionic conductor support.

Moreover, two different electropromotional effects were found under POM conditions

depending on the applied potential, which were attributed to the formation of different

kinds of promoter phases on this catalyst. The higher catalytic activity of Pt-DLC,

compared to that of the pure Pt film, demonstrated the practical interest of this kind of

dispersed catalyst films with lower metal loading.

Other novel electrochemical catalyst based on Au nanoparticles dispersed in a

Yttria-Stabilized Zirconia (YSZ) matrix was deposited by reactive co-sputtering of

Zirconium-Yttrium and Au targets on a K-βAl2O3 solid electrolyte. The Au-YSZ film

was electrically non-conductive and a silver current collector was used to polarize the

catalyst film that would be used in the electrochemical promotion experiments. The

Au nanoparticles showed to be active in the partial oxidation of methanol with a high

selectivity toward methyl formate production. This configuration allowed to decrease

the amount of metal used in the solid electrolyte cell, and to activate a highly

dispersed Au catalyst via electrochemical promotion (EPOC), by in-situ controlling

and optimizing the supplied amount of K+ ions. A number of experiments confirmed

that the observed electropromotional effect did depend on neither the rate of K+ supply

nor the operation mode (galvanostatic or potentiostatic). It only depended on the

achieved final promoter coverage. The stability of the Au nanoparticles under the

explored conditions was also confirmed under EPOC reaction conditions.

Then, a novel Cu catalyst film was prepared by oblique angle physical vapour

deposition (OAD) on a K-βAl2O3 solid electrolyte. This technique allowed to obtain a

highly porous and electrically conductive Cu catalyst electrode composed of metal

nanocolumns, which was electrochemically promoted in the partial oxidation of

methanol (POM). The production rates of hydrogen, carbon dioxide and methyl

formate were in-situ enhanced in a reversible and reproducible way, by means of the

controlled migration of electropositive potassium ions. Moreover, the enhancement

ratios were comparable to those obtained with the Pt-DLC catalyst. Under the studied

reaction conditions, these promoter ions also formed potassium-derived surface

Abstract

60

compounds as demonstrated by post-reaction characterization analysis. Some nitrogen

functional groups and carbonaceous compounds were also detected. The obtained

results demonstrate the interest of the used catalyst-electrode preparation technique for

the electrochemical activation of non-noble metal catalyst films with high gas-exposed

surface area.

Then, three possible applications of the phenomenon of electrochemical

promotion of catalysis (EPOC) were demonstrated with a Ni catalyst for methanol

conversion processes: activation of the catalyst, modification of the catalytic

selectivity and partial oxidation of the catalyst. The Ni catalyst film was prepared by

cathodic arc deposition and electrochemically promoted by K+, upon negative

polarization, in the methanol decomposition (MD) and steam reforming (SRM)

reactions, showing an electrophilic EPOC behavior in both cases. In the presence of

water, the K+ ions promotional effect also attenuated the Ni deactivation by carbon

deposition. On the other hand, under methanol partial oxidation conditions (POM), the

K+ ions caused a sharp decrease in the catalytic selectivity toward H2 and CO while

the production rates of CO2 and H2CO slightly enhanced, due to an increase in the Ni

oxidation state by the alkali-induced O2 activation. All the potassium-derived effects

were fully reversible between the negative and positive polarizations, which showed

different interesting possibilities of the EPOC phenomena in heterogeneous catalysis.

Finally, the possibility of applying the EPOC phenomenon with alkaline

conductors in the fields of H2 production and storage was also explored. In this case, a

porous Ni catalyst film was deposited on K-βAl2O3 by the oblique angle deposition

technique. Under steam reforming conditions and negative polarization, this

electrocatalytic system allowed to produce and store H2 with a very high yield per

amount of metal (up to 19 g H2 x 100 g Ni-1

) due to the promotional effect of K+ ions.

Moreover, it was possible to release the stored, highly pure, H2 in a controlled way

under positive polarization, which represents a new application of the EPOC

phenomenon of great interest. The influence of the catalyst microstructure, the applied

negative current and the reaction atmosphere was studied. A H2 storage mechanism

was proposed on the basis of the obtained results. Although some K+-derived surface

Abstract

61

compounds such as bicarbonates were detected by post-reaction characterization of the

Ni catalyst surface, the very high observed H2 storage capacity was mainly attributed

to the chemisorption of H atoms on both Ni active sites and carbonaceous surface

compounds through a spillover process. These carbonaceous species were

simultaneously formed under K+-promoted reaction conditions and showed to be

mainly composed of graphene oxide.

Abstract

62

63

Chapter 1

ELECTROCHEMICAL PROMOTION OF Pt FOR H2

PRODUCTION FROM METHANOL PARTIAL OXIDATION

AND STEAM REFORMING:

A BETTER PERFORMANCE OF PHYSICAL VAPOR DEPOSITED

CATALYST FILMS

The effect of the electrochemical promotion (EPOC) has been studied for the

first time in the steam reforming and partial oxidation of methanol to produce H2 and

also formaldehyde in the latter case. Two Pt catalyst films deposited on K-βAl2O3 (K+-

conductor material) by different techniques, impregnation and cathodic arc

deposition, have been tested under different applied potentials and reaction

conditions. In alkali-based solid electrolyte systems, an excess of electrocatalytic

activity of the catalyst-working electrode (the case of the impregnated Pt film) can be

detrimental due to the excessive formation of promoter-derived surface compounds.

However, the catalytic activity and selectivity of the Pt film prepared by the cathodic

arc technique have been strongly enhanced by EPOC with K+ ions. This catalyst film

also showed to be stable for long working times and the promotional effect was fully

reversible in all the experiments. Hence, the cathodic arc deposition and other

physical vapor deposition techniques were applied in the following chapters.

ΔV < 0

Pt

K+ K+ K+

e-

e-

K-βAl2O3

Au

Reactants Products

SRM: CH3OH + H2O CO2 + 3H2

POM: CH3OH + 1/2O2 CO2 + 2H2

Δr

ELECTROCHEMICAL

PROMOTION (EPOC)

1.1 mg Pt cm-2

0.23 mg Pt cm-2

120 nm thick

Impregnation

Physical Vapor

Deposition (PVD)

0

0.2

0.4

0.6

0.8

1

1.2

0 50 100 150 200 250 300 350 400 450 500 550

Tiempo / min

r H C

O /

mo

l s-1

x 1

0-7

-2.5

-1.5

-0.5

0.5

1.5

VW

R / V

0

2

4

6

8

10

r H

/

mo

l s-1

x 1

0-7

-2.5

-1.5

-0.5

0.5

1.5

2.5

VW

R / V

Pt impregnado

Pt por CAD

VWR

Pt impregnado

Pt por CAD

VWR

22

7.2 % CH3OH

4.6 % O2

320 ºC

VW

R/ V

VW

R/ V

Impregnated Pt

Pt by PVD

VWR

Impregnated Pt

Pt by PVD

VWR

Time / min

rH

CO

/ m

ol s-1

x 1

0-7

rH

/ m

ol

s-1x

10

-72

2

Chapter 1

64

1.1. INTRODUCTION

Hydrogen is attracting increasing interest as a future clean energy carrier with

main application in fuel cell technology. It can be obtained by many different

processes and from a wide variety of raw materials (water, biomass or hydrocarbons),

the steam reforming of methane currently being the most used in industry [1-3].

However, the use of liquid hydrogen carriers at atmospheric pressure and room

temperature, such as methanol, is acquiring relevance due to the facilities arising from

its transport, storage and handling. Other advantages of methanol lie in its high H/C

ratio (4:1) and the absence of C-C bonds which allows its conversion at lower

temperatures (150-350 ºC) than most other fuels [4, 5]. Moreover, CH3OH can be

produced not only from fossil fuels (natural gas, coal, oil shale, tar sands, etc.), but

also from agricultural products and municipal waste, wood and varied biomass such as

landfill gas, sugar beets, driftwood or rice straw. It can even be obtained from

chemical recycling of carbon dioxide [6]. H2 production from methanol is mainly

carried out through the steam reforming (SRM, reaction 1.1) and partial oxidation

(POM, reaction 1.2).

CH3OH + H2O → 3H2 + CO2 (1.1)

CH3OH + 1/2O2 → 2H2 + CO2 (1.2)

The first reaction is endothermic and produces a higher H2/C ratio, while the

second one is exothermic and typically leads to higher reaction rates at lower

temperatures, also requiring smaller reactors [7]. This latter reaction can even be run

in an autothermal mode where the heat necessary for the reaction comes from the

reaction itself. The catalysts typically used in both processes, steam reforming and

partial oxidation of methanol, are similar as those employed in the methanol synthesis

(reverse reaction 1.1) and water-gas shift (WGS) reactions and have been reviewed by

several authors [4, 5, 8]. They can be divided into two main categories: copper-based

and group 8-10 metal-based catalysts. In particular, Pt-based catalysts have shown

very good activity for hydrogen production by both SRM [9-12] and POM [13-17].

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

65

Moreover, regarding the partial oxidation of methanol, it is also possible to obtain

formaldehyde (H2CO), through the following reaction:

CH3OH + 1/2O2 → H2CO + H2O (1.3)

H2CO is also of great industrial interest, as a precursor to urea formaldehyde

resin, melamine resin, phenol formaldehyde resin and 1,4-butanediol, among others

[18]. However, Pt catalysts may also lead to high deep oxidation rates, which would

decrease the selectivity to H2 and formaldehyde production. Then, it is important to

enhance the partial oxidation mechanism vs. the deep oxidation one for the production

of both kinds of products. Then, the performance of metal catalysts employed in the

SRM and POM reactions has shown to be promoted by different dopants such as Al

[19], Zn [20] or alkali metals [21-23].

In this sense, the coupling of catalysis and electrochemistry has also emerged

in the last 30 years through the Electrochemical Promotion Of Catalysis (EPOC). This

phenomenon, also known as Non-faradaic Electrochemical Modification of Catalytic

Activity (NEMCA effect), opens a new way to promote the catalytic

activity/selectivity of a metal catalyst under working reaction conditions [24, 25]. It is

based on the modification of the catalytic properties of a metal catalyst by the

electrochemical pumping of promoter ions from an electro-active catalyst support,

which is a solid electrolyte material [26]. In the case of an alkaline solid electrolyte (as

in the present thesis, K-βAl2O3), the application of a cathodic polarization (negative

current) between a catalyst-working electrode, which is deposited on one side of the

electrolyte, and an inert counter electrode located at the opposite side (Figure 1.1) lead

to the migration of promoter (K+) ions to the catalyst film (back-spillover). Once

located over the catalyst surface, as in chemical (classical) promotion, these ionic

species modify the chemisorption properties of the catalyst and hence the reaction rate

[27]. This phenomenon has been demonstrated by means of several characterization

techniques and has been applied in more than 80 catalytic systems with several

important technological possibilities [28-31].

Chapter 1

66

Figure 1.1. Scheme of the electrochemical cell used in the electrochemical promotion studies

with an alkaline ionic conductor solid electrolyte.

For instance, previous works have demonstrated the interest of EPOC in

enhancing the selectivity of several catalytic processes such as the selective catalytic

reduction of NOx [32, 33], CO2 hydrogenation [34], Fischer–Tropsch synthesis [35] or

CO preferential oxidation [36]. On the other hand, other studies have also reported the

application of the EPOC on the methanol decomposition and dehydrogenation on Ag

[37] and the methanol partial oxidation on Pt [38, 39] and Ag [40] by using yttria-

stabilized zirconia (YSZ) as the solid electrolyte, which is a O2-

conductor. These four

previous works were only focused on formaldehyde production (not H2 production)

and were carried out at relatively high reaction temperatures (up to 900 K). Here, a

main difference with respect to the present research lies, where all the electrochemical

promotion studies have been performed by means of akali conductor materials.

In the present work, it is proposed, for the first time, the electrochemical

promotion of the production of H2 via partial oxidation and steam reforming of

methanol, and also the simultaneous production of formaldehyde in the former case.

The effect of the catalyst film preparation technique, the reaction conditions and the

durability of the system for long working times have been investigated on a Pt/K-

βAl2O3 electrochemical catalyst. Under POM conditions, the novel proposed EPOC

Catalyst – Working

electrode

Counter

electrode

(Au)

Reference

electrode

(Au)

I Solid electrolite

(K-βAl2O3)

VWR

ΔrReactants Products

K+ K+ K+ Back-spilloverCathodic

polarization

I < 0

ELECTROCHEMICAL PROMOTION (EPOC)

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

67

configuration may have significant importance since it would allow to produce in a

single reaction step both kinds of products: H2 and H2CO, which could be easily

separated for their further use in different applications.

1.2. EXPERIMENTAL

1.2.1. Preparation of the electrochemical catalysts

The electrochemical catalyst consisted of a continuous, thin Pt film (geometric

area of 2.01 cm2) deposited on a side of a 19-mm-diameter, 1-mm-thick K-βAl2O3

(Ionotec) disc, which was used as a cationic solid electrolyte (i.e. as source of K+

promoter ions). The electrochemical catalyst configuration used in this chapter and in

the rest of the thesis can be observed in Figure 1.2.

Figure 1.2. Scheme of the Pt/K-βAl2O3/Au electrochemical catalyst.

In first place, Au counter (CE) and reference (RE) electrodes were deposited

on the other side of the electrolyte. These electrodes were deposited by applying thin

coatings of gold paste (Gwent Electronic Materials C1991025D2), followed by

calcination at 800 ºC for 2 h (heating ramp of 5 ºC min-1

). Au was chosen because of

its inertness for the methanol partial oxidation reaction (checked via blank

experiments), which makes it an adequate pseudo-reference electrode when using the

Catalyst – Working electrode (Pt)

Solid electrolyte (K-βAl2O3)

Counter electrode (Au)

Reference electrode (Au)

Chapter 1

68

single pellet cell design. Then, the active Pt catalyst film (WE), which also behaved as

a working electrode, was deposited on the other side of the electrolyte by two different

preparation techniques:

- In the first case, the Pt film was prepared by an impregnation method

described in detail elsewhere [41], consisting of successive steps of deposition and

thermal decomposition (450 ºC for 1 h, heating ramp of 5 ºC min-1

) of a H2PtCl6

precursor solution over a mixture of 1/1 H2O/2-propanol with a metal concentration of

0.1 M. The final Pt loading was 1.09 mg Pt cm-2

.

- In the second one, the Pt film was prepared by the pulse cathodic arc

technique (denoted as CAD; Cathodic Arc Deposition) [42]. The individual pulses

were 1 ms long and had a current of approximately 300 A. A total of 2800 pulses were

used for the deposition of the Pt layer. The substrate holder was at ground during the

deposition and was rotated at 2 r.p.m. This electrode preparation technique allowed

the preparation of a thin, highly dense, Pt film with 122 nm of thickness with high

adhesion to the substrate. The final Pt loading was 0.23 mg Pt cm-2

in this case.

This catalyst preparation method was carried out in collaboration with Dr.

José Luis Endrino from the Institute of Materials Science of Madrid (CSIC).

The resulting Pt/K-βAl2O3/Au electrochemical catalysts were placed into a

single chamber solid electrolyte cell reactor. A scheme of this reactor is shown in

Figure 1.3. It was made of a quartz tube with appropriate feed-through, where all the

electrodes were exposed to the same gas atmosphere.

The three electrodes (working, counter and reference) were connected to an

Autolab PGSTAT320-N potentiostat-galvanostat (Metrohm Autolab) by using gold

wires. Then, in the electrochemical promotion experiments, a catalyst potential (VWR)

was applied according to the procedure generally used in conventional three-

electrodes electrochemical cells [32, 33, 36].

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

69

Figure 1.3. Scheme of the single chamber solid electrolyte cell reactor.

Prior to the catalytic activity measurements, both catalyst films were subjected

to a 25 % H2 stream (N2 balance) while heating to 450 ºC (ramp of 5 ºC min-1

), to

ensure they were in a fully reduced state, and were kept at this temperature under

POM conditions (CH3OH/O2 = 7.2 %/4.6 %) overnight for their stabilization.

1.2.2. Characterization measurements

Cyclic voltammetry (CV) measurements were performed under reaction

conditions with the potentiostat–galvanostat in order to investigate the formation and

decomposition of promoter species. The samples were prior to the measurement kept

at +2 V for 30 min in order to define a reference state (Pt films clean of promoter

ions). Then, the potential was varied at a scan rate of 80 mV s-1

from +2 to -2 V.

Reference electrode (R)

Counter electrode (C)

Working electrode (W)

Thermocouple

InletOutlet

Cooling coil

Metal lid

Quartz tubes

Perforated alumina

tube

Gold wires

Reference electrode (R)

Counter electrode (C)

Working electrode (W)

Thermocouple

InletOutlet

Cooling coil

Metal lid

Quartz tubes

Perforated alumina

tube

Gold wires

Pt, W

Chapter 1

70

The two different Pt electrode films were characterized, after the EPOC

experiments, by X-ray diffraction (XRD) with a Philips PW 1710 instrument using Ni-

filtered Cu Ka radiation. Before the XRD analysis, the electrochemical catalysts were

kept under POM reaction conditions (CH3OH/O2 = 7.2 %/4.6 %) at 320 ºC overnight

under a potential VWR = -2 V in order to favour the formation of the promoter species.

The diffractograms were then compared with the JCPDS-ICDD references for

identification purposes of the promoter phases. The catalyst films were also

characterized after the electrochemical promotion experiments via scanning electron

microscopy (SEM) and EDX analysis using a FEI Nova NANOSEM 230 instrument.

1.2.3. Catalytic activity measurements

The catalytic activity measurements were carried out in the experimental setup

shown in Figure 1.4.

Figure 1.4. Scheme of the experimental set-up employed in the electrochemical promotion

experiments.

N2 / Ar

H2

Air / O2

N2 / Ar

FlowmeterTermometer

Mass

flowmeters

Potentiostat/

Galvanostat

Gas chromatograph

Reactor

Furnace

FIC

FIC

FIC

FIC

TIC

TIC

To vent

H2O

saturator

CH3OH

saturator

TIC

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

71

The reaction gases (Praxair, Inc.) were certified standards (99.999 % purity) of

air (source of the O2 used in the methanol partial oxidation experiments), H2

(reductant agent) and Ar (carrier gas). The gas flow rates were controlled by a set of

mass flowmeters (Bronkhorst EL-FLOW). Methanol (Panreac, 99.8 % purity) and

water (distilled and deionized, used in the methanol steam reforming experiments)

were fed by sparging Ar through thermostated saturators. All lines placed downstream

from the saturators were heated above 100 ºC to prevent condensation.

The electrochemical promotion experiments were performed at atmospheric

pressure with an overall gas flow rate of 6 NL h-1

, a temperature ranging from 250 to

320 ºC with a feed composition of CH3OH/O2 = 7.2 %/0 - 8.9 % (N2 balance) for the

POM experiments, and a temperature ranging from 320 to 450 ºC with a feed

composition of CH3OH/H2O = 4 %/4.8 % (N2 balance) for the SRM experiments.

Reactant and product gases were on-line analyzed by using a double channel gas

chromatograph (Bruker 450-GC) equipped with Hayesep and Q-Molsieve 13X

consecutive columns and a CP-Wax 52 CB column, along with thermal conductivity

and flame ionization detectors, respectively. The detected products were: H2, CO2,

CO, H2O and formaldehyde (H2CO), the latter two only under POM reaction

conditions. The error in the carbon atom balance did not exceed 5 % in any

experiment, which indicated no consistent loss of material, no significant formation of

other oxygenated species and/or coking of the catalyst-electrodes. Unless otherwise

stated, the error remained below this value in the following chapters.

1.3. RESULTS AND DISCUSSION

1.3.1. Partial oxidation of methanol

i) Influence of the preparation technique.

In order to study the influence of the Pt film preparation technique, a

potentiostatic experiment was carried out with both catalysts: the impregnated Pt film

and the Pt prepared by Cathodic Arc Deposition (CAD), under methanol partial

oxidation (POM) conditions. Figure 1.5 shows the variation of the different product

Chapter 1

72

reaction rates: H2, H2CO, CO2 and CO, as well as the overall CH3OH conversion vs.

time under different applied potentials (VWR), each for 60 min, under the following

POM reaction conditions: CH3OH/O2 = 7.2 %/4.6 % (N2 balance), 320 ºC.

Initially, the catalyst was in Open Circuit (OC) conditions, i.e., no electric

current or potential was being applied. Then, at the beginning and at the end of the

experiment, a catalyst potential VWR = +2 V was applied in order to remove all the K+

ions that could be located on the Pt surface. In this way, a clean (unpromoted) Pt

catalyst film was defined, and hence a reference state [36, 41]. Then, a significant

difference in the activity of the two electrochemical catalysts with varying the applied

potential can be observed in this figure. For the impregnated Pt film, a decrease in the

catalyst potential from +2 V led to a sharp reduction in the catalytic activity of the

system, while in the case of CAD Pt film, a decrease in the catalyst potential from +2

V led to a strong increase in the activity of the system (increasing the production rate

of all the products as well as the CH3OH conversion). Then, according to previous

studies of electrochemical promotion [27], one can apparently classify the case of the

impregnated Pt film as “electrophobic EPOC behaviour” and the case of the CAD Pt

film as “electrophilic EPOC behaviour”.

A decrease in the catalyst potential involved a decrease in the catalyst work

function due to the migration of positively charged potassium species, which would

weaken the Pt chemical bond with electron-donor adsorbates and would strengthen

those with electron acceptor ones. According to the previous work reported by

Vayenas and Neophytides on methanol partial oxidation [38], CH3OH can be

identified as the donor molecule and O2 as the acceptor one. In principle one can

explain the different observed EPOC behavior in Figure 1.5 according to the different

reaction orders of the two reaction molecules: acceptor (O2) and donor (CH3OH) for

the two catalyst films. Hence, the preparation method used for the two Pt films: i.e.,

impregnation of a Pt precursor solution or that based on the pulse cathodic arc

technique, seems to have a key role.

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

73

Figure 1.5. H2 (a), H2CO (b), CO2 (c) and CO (d) production rates and CH3OH conversion (e)

vs. time under different applied potentials. Partial oxidation conditions: CH3OH/O2 = 7.2 %/4.6

%, 320 ºC.

0

4

8

12

16

r CO

/

mo

l s-1

x 1

0-7

-2.5

-1.5

-0.5

0.5

1.5

VW

R / V

0

10

20

30

40

0 100 200 300 400 500

Time / min

CH

3O

H C

on

vers

ion

/

-2.5

-1.5

-0.5

0.5

1.5

VW

R / V

0.0

0.2

0.4

0.6

0.8

1.0

1.2

r H

CO

/ m

ol s

-1 x

10

-7

-2.5

-1.5

-0.5

0.5

1.5

VW

R / V

0.0

0.5

1.0

1.5

2.0

2.5

3.0

r CO

/ m

ol s

-1 x

10

-7

-2.5

-1.5

-0.5

0.5

1.5

VW

R / V

0

2

4

6

8

10

r H

/ m

ol s

-1 x

10

-7

-2.5

-1.5

-0.5

0.5

1.5

2.5

VW

R / V

2

2

2

%

a)

b)

c)

d)

e)

Impregnated Pt

Pt by CAD

VWR

Impregnated Pt

Pt by CAD

VWR

Impregnated Pt

Pt by CAD

VWR

Impregnated Pt

Pt by CAD

VWR

Impregnated Pt

Pt by CAD

VWR

Chapter 1

74

Previous studies have demonstrated [43] that the method of catalyst

preparation, affects the catalytic behavior due to a difference in the morphology

(surface roughness) and dispersion of the Pt particles. In addition, a work of Mutoro et

al. [44] also demonstrated important differences in the magnitude of the EPOC effect

in dense and porous Pt films prepared by two different techniques: application of Pt

metal pastes and pulse cathodic arc deposition. The higher EPOC effect observed with

the porous Pt film deposited on YSZ vs. the dense pulse layer deposited Pt film was

attributed to the higher tpb (three phase boundaries) length of the former, which

facilitated the formation of oxygen spillover promoting species. Then, very likely the

different morphology and tpb length of the two Pt catalyst films may be responsible of

the observed different EPOC behavior in Figure 1.5. In order to prove this hypothesis,

additional measurements (characterization techniques) were carried out.

Figure 1.6 shows the current variation vs. the applied potential (VWR) of the

two samples: impregnated Pt film (a) and Pt by CAD (b) during the cyclic

voltammetry (CV) between +2 and -2 V carried out under the same reaction

conditions of Figure 1.5 (CH3OH/O2 = 7.2 %/4.6 %, 320 ºC). As reported in previous

studies with alkali electro-promoted systems [45-47], the cathodic scan of the cyclic

voltammetry (negative potentials) could be linked to the migration of ions from the

solid electrolyte to the catalyst-working electrode surface with the consequent

formation of promoter-derived surface compounds after interaction with the

chemisorbed species. On the other hand, the anodic scan of the cyclic voltammetry

(positive potentials) could be linked to the decomposition of the previously formed

species and the migration of the ions back to the solid electrolyte. Thus, taking into

account that the gold counter electrode is constituted by big particles of negligible

activity, all the electrochemical processes displayed on the CV could be mainly

attributed to the formation and decomposition of surface compounds on the Pt

catalyst-working electrode [47].

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

75

Figure 1.6. Current variation vs. the applied potential (VWR) of the two samples: impregnated

Pt film (a) and Pt films prepared by cathodic arc deposition (b) during the cyclic voltammetry

between +2 and -2 V (CH3OH/O2 = 7.2 %/4.6 %, 320 ºC, scan rate = 80 mV.s-1

).

In this figure, one can also observe a much higher electrocatalytic activity

(much higher cathodic and anodic peaks) for the case of the impregnated Pt film vs.

the Pt film prepared by CAD, probably due, as commented above, to a higher tpb

length of the former (porous film calcined at high temperature 450 ºC). For the case of

the impregnated Pt film, it can be observed a strong cathodic peak below +1.3 V

(Figure 1.6a) that seemed to indicate the formation of an important amount of

potassium surface compounds, which in turn could block the Pt active sites and

explain the poisoning effect of the impregnated Pt film shown in Figure 1.5 at VWR =

-8

-6

-4

-2

0

2

4

6

8

Cu

rre

nt

/ m

A

-0.06

-0.04

-0.02

0.00

0.02

0.04

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5

Cu

rre

nt

/ m

A

VWR / V

a)

b)

Chapter 1

76

+1 V. For the case of the Pt sample by CAD (Figure 1.6b), a much lower cathodic

peak was observed at VWR = -1 V, demonstrating that such important formation of

potassium derived compounds did not occur on this sample in contrast with that

observed with the impregnated sample. The observed anodic peaks of both samples

were in good agreement with the cathodic ones and a much higher size was again

obtained for the case of the impregnated Pt sample. These observations were in good

agreement with the XRD analysis (Figure 1.7) of the Pt films after the application of

VWR = -2 V overnight under the reaction conditions in Figure 1.5 (CH3OH/O2 = 7.2

%/4.6 %, 320 ºC).

Figure 1.7. XRD spectra after EPOC experiments (CH3OH/O2 = 7.2 %/4.6 %, 320 ºC,

overnight at -2 V) of Pt films prepared by impregnation (a) and cathodic arc deposition (b)

techniques.

0

400

800

1200

1600

5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90

2Ѳ / º

Inte

nsi

ty /

co

un

ts

0

500

1000

1500

2000

2500

3000

3500

4000

Inte

nsi

ty /

co

un

ts

a)

b)

0

50

100

150

200

10 12 14 16 18 20 22 24 26 28

2Ѳ / º

Inte

nsit

y /

co

un

ts

(111)

(200)

(220)

(311)

(222)

Pt

K-βAl2O3

2K2CO3·3H2O

C2HKO4

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

77

On both samples, peaks at 2θ = 39.6º, 47.4º, 67.1º, 81.2º and 83.6º

corresponded to the reflections (111), (200), (220), (311) and (222), respectively, and

were consistent with a Pt face-centered cubic (FCC) structure (JCPDS, 04-0802). The

βAl2O3 phase corresponding to the solid electrolyte can also be clearly identified

(JCPDS, 02-0921). It can be observed that the ratio of the Pt peaks vs. solid electrolyte

peaks was much higher in the impregnated sample, which is indicative of a much

higher thickness of the impregnated Pt vs. the CAD film. However, the most

interesting part of the figure comes from the occurrence of two new diffraction peaks

in the impregnated sample (inset of Figure 1.7a) that were assigned to potassium

carbonate hydrate oxide: 2K2CO3·3H2O (JCPDS, 01-1014) and potassium hydrogen

carbonate oxide: C2HKO4 (JCPDS, 11-0286). These results would support the

formation of an excess of promoter phases under the negative potential application in

the impregnated Pt film due to its higher electrocatalytic activity confirmed by the

cyclic voltammetry in Figure 1.6, which resulted in the apparent electrophobic EPOC

behaviour shown in Figure 1.5.

Figure 1.8 shows the SEM micrographs along with its corresponding EDX

analysis (in the marked regions) of the impregnated Pt film at various levels of

magnification (after the application of VWR = -2 V overnight under the reaction

conditions considered in Figure 1.5). Two different regions are observed in all the

micrographs. According to the EDX analysis, the lighter region corresponds to Pt

while the darker region corresponds to a potassium-derived carbonaceous surface

compound (combination of carbon, potassium and oxygen) in good agreement with

the XRD analysis (Figure 1.7a). In fact, the elemental analysis shown in Figure 1.8c is

in quite accordance with the two compounds detected by the XRD analysis:

2K2CO3·3H2O and C2HKO4. On the other hand, the SEM analysis of the Pt film by

CAD after the EPOC experiments (not shown) just showed the lighter region

corresponding to Pt with a more dense structure.

Chapter 1

78

Figure 1.8. Top surface SEM micrographs of the Pt film prepared by impregnation after EPOC

experiments (CH3OH/O2 = 7.2 %/4.6 %, 320 ºC, overnight at -2 V) at 50 µm (a), 20 µm (b)

and 10 µm (c) along with the corresponding EDX analysis of the marked region.

0

10

20

30

40

50

At

(%)

C O Pt K

0

10

20

30

40

50

At

(%)

C O Pt K

0

10

20

30

40

50

60

At

(%)

C O Pt K

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

79

Hence, the impregnated Pt film showed a important ability to form and store

potassium surface compounds, which clearly blocked an important area of it. Very

likely, it led to a decrease in the Pt number of active sites under EPOC conditions (K+

pumping conditions in Figure 1.5 at VWR < +2V), which explaines the observed

electrophobic EPOC behaviour of the impregnated sample in Figure 1.5. Then,

according to these observations, it is interesting to note that the EPOC phenomena

with alkali ionic conductors does not only depend on the kinetics order of both

acceptor and donor molecules but also on the electrocatalytic activity of the catalyst

film to form and store surface compounds between the promoter ions and the

chemisorbed species. Thus, with cationic solid electrolyte systems such as the Pt/K-

βAl2O3 used here, an excess of electrocatalytic activity of the catalyst-electrode film

could be detrimental for an electropromotional effect, in contrast to the O2-

conductors

[27, 44].

ii) Electrochemical promotion mechanism and parameters.

In order to quantify the promotional effect of both catalyst films, Figure 1.9

shows the values of the electrochemical promotion parameters [27] calculated for H2

and H2CO production: ρ (reaction rate enhancement ratio) and PIK+ (promotion index)

vs. the potassium coverage (θK+) achieved during the potentiostatic transitions of

Figure 1.5 (correlated with VWR) and calculated from the integration of the current vs.

time curves via the Faraday’s law (equation 1.4):

θ =

(1.4)

where I is the current, t is the time, n is the potassium ion charge, i.e., +1, F is the

Faraday constant and NG is the number of Pt active sites. In this case, NG was

estimated from the total number of Pt sites in the sample and the dispersion values

obtained from the Pt particle size estimated by the Scherrer equation [48] from the

XRD measurements. The promotional parameters where calculated as follows:

ρ =

(1.5)

Chapter 1

80

θ (1.6)

where r and r0 are the promoted (VWR < +2 V) and unpromoted (VWR = +2 V) catalytic

reaction rates, respectively, and ∆r = r – r0 is the K+-induced change in the catalytic

reaction rate.

Figure 1.9. Variation of the promotional parameters: rate enhancement ratio (ρ) and promotion

index (PIK+) for H2 (a) and H2CO (b) vs. potassium coverage on the catalyst film (θK+) and

catalyst potential (VWR). Partial oxidation conditions: CH3OH/O2 = 7.2 %/4.6 %, 320 ºC.

0

1

2

3

4

5

6

7

ρH

-20

-10

0

10

20

30

40

PIK

+,H

0

1

2

3

4

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

θK+ / mol K+ x mol active Pt-1

ρH

C

O

-20

-10

0

10

20

PIK

+,H

CO

a)

b)

Impregnated Pt

Pt by CAD

Impregnated Pt

Pt by CAD

2

2

2

2

2 1 0 -1 -2 Pt by CAD

2 1 0 -1 -2 Impregnated Pt

VWR / V

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

81

Firstly, it can be observed that much lower potassium coverage was obtained

in the Pt film prepared by the CAD technique as compared to that of the impregnated

Pt film. In the latter case, a high potassium coverage (θK+ = 0.4) was almost obtained

upon decreasing the catalyst potential (VWR) just to +1 V. This was not the case for the

CAD-Pt film, where much lower potassium coverage was obtained under all the

different applied potentials, i.e., the highest potassium coverage of 0.3 was attained in

the last polarization of -2 V. These results, which are in good agreement with the

previous characterization measurements, support the poisoning behaviour of the

impregnated Pt film observed in Figure 1.5 when decreasing the catalyst potential to

+1 V and lower potentials. It has already been reported in several alkali

electropromoted [41, 49] and alkali promoter catalytic systems [50] that an excess of

promoter may lead to a poisoning catalytic behaviour due to the blocking of active

sites.

Hence, these results seem to indicate that such poisoning regime is easy to

achieve for the case of the impregnated film, i.e., a high potassium coverage around

θK+ = 0.4 was already obtained in the first polarization of +1V in Figure 1.5. As can be

observed in this figure, regarding the Pt impregnated film, a negative effect of

polarization was obtained for the case of H2 production (ρ < 1 and PIK+ < 0) while a

slight increase in the reaction rate was observed for the case of H2CO (up to ρ = 2.7

and PIK+ = 1.8) with a decrease of the potential from +2 V. However, in this latter case

and according to the results of Figure 1.5, a negligible production rate of H2CO was

observed due to the previously mentioned suppression of methanol reaction rate (the

conversion decreased below 5 %).

Regarding the CAD-Pt film, the effect of electrochemical pumping of

potassium ions led to a strong increase in both H2 and H2CO production rates (ρ > 1

and PIK+ > 0). In this case, the rate enhancement ratio (solid line) for both products

reflects an electrophilic EPOC behaviour. The promotion index (dashed line) had an

optimum value at a potassium coverage θK+ = 0.1. An increase in the potassium

coverage above this value (VWR < -0.5 V in Figure 1.5) did not further improve the

Chapter 1

82

catalytic activity of the system. This fact could be explained considering that the

promotional effect of the K+ ions began to be balanced with the negative effect of

blocking of Pt active sites at higher promoter coverage (θK+ > 0.1). Under optimal

potential conditions, the H2 production rate was increased more than 6 times leading to

a promotion index higher than 34, although for the case of H2CO that effect was less

pronounced (ρ~4 and PIK+~25). Hence, it is clear that the presence of K+ ions on the

CAD-Pt film led to a strong promotional effect in the partial oxidation mechanism,

enhancing in a very pronounced way the production rates of both H2 and H2CO. One

could explain such strong promotional effect attending to the mechanism of methanol

partial oxidation reaction on Pt catalysts. Previous spectroscopic studies [51]

demonstrated that methanol molecule is adsorbed on Pt as a methoxy specie by

dissociating H atom from the OH group according to reaction (8):

CH3OH → CH3O(a) + H(a) (1.7)

Then, the methoxy intermediate is converted to adsorbed formaldehyde by the

abstraction of a hydrogen atom by surface oxygen:

CH3O(a) + O(a) → H2CO(a) + OH(a) (1.8)

According to this mechanism, the observed promotional effect upon negative

polarization (in the presence of an electropositive promoter) can be attributed to an

increase in metal-oxygen (electron acceptor molecule) bonds strength, which would

enhance the formaldehyde production rate through the reaction step 1.8. Thus, more

active sites would be available for methoxy species chemisorption also increasing the

H2 production. Very likely, part of the produced H2CO can be further decomposed and

oxidized with stronger chemisorbed O(a) species leading to the production of CO and

CO2, respectively. It would also explain the observed lower promotional effect (lower

promotion index and enhancement ratio) of H2CO vs. H2 in Figure 1.9.

In short, all these results clearly demonstrate the superior performance of the

CAD-Pt film vs. the impregnated Pt film for the simultaneous production of H2 and

H2CO at the explored reaction conditions and potential range (VWR = +2, -2 V). Hence,

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

83

the Pt film by CAD was selected for further investigation in the reaction conditions of

the methanol partial oxidation reaction.

iii) Influence of the reaction conditions.

Figure 1.10 shows the steady state variation (after 1 h of polarization at each

potential) vs. the applied potential (VWR) of different reaction parameters: H2, H2CO

and CO2 selectivity as well CH3OH conversion at different temperatures (250-320 ºC)

for the following composition CH3OH/O2 = 7.2 %/4.6 %. These parameters were

calculated through the following equations:

(1.9)

(1.10)

(1.11)

(1.12)

where Fi,in and Fi,out are the molar flow rate of the i species at the inlet and at the outlet

of the reactor, respectively.

In good agreement with the previously obtained results, it can be observed for

the CAD-Pt film that a decrease in the catalyst potential and, hence, an increase in the

potassium coverage, increased the CH3OH partial oxidation activity favouring the H2

and H2CO selectivity (vs. CO2 selectivity) as well as the CH3OH conversion at all

explored reaction temperatures.

Chapter 1

84

Figure 1.10. H2 (a), H2CO (b), CO2 (c) selectivities and CH3OH conversion (d) vs. applied

catalyst potential (VWR) at different reaction temperatures with the Pt film prepared by cathodic

arc deposition. Partial oxidation conditions: CH3OH/O2 = 7.2 %/4.6 %.

20

25

30

35

40

45

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5

VWR / V

CH

3O

H C

on

vers

ion

/ %

250ºC280ºC320ºC

0

3

6

9

12

15

18

21

H2 S

ele

ctiv

ity

/ %

250ºC280ºC320ºC

75

80

85

90

95

CO

2 S

ele

ctiv

ity

/ %

250ºC280ºC320ºC

0

2

4

6

8

10

H2C

O S

ele

ctiv

ity

/ %

250ºC280ºC320ºC

a)

b)

c)

d)

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

85

On the other hand, an increase in the temperature enhanced the methanol

partial oxidation activity (toward H2 production, Figure 1.10a) under both:

unpromoted (VWR = +2V) and electro-promoted (VWR < +2V) conditions. A similar

increase in the POM activity with the reaction temperature has been reported in

previous works where Pt supported catalysts were used [13, 17]. However, in the

explored temperature range the CO2 selectivity (Figure 1.10c) was nearly 100 %

under unpromoted conditions (VWR = +2 V) and it was not greatly influenced by

temperature, in agreement with a study of McCabe and McCready [52]. Under

unpromoted conditions (+2 V), the methanol-derived carbon products (CO and

H2CO) were easily oxidized to CO2. Nevertheless, under electro-promoted conditions

the CO2 selectivity decreased and the production of less oxidized carbon products,

such as H2CO increased due to the previously mentioned enhancement in the partial

oxidation mechanism under the effect of K+.

It can also be observed that the promotional effect on the H2 production rate

(H2 selectivity) seemed to increase with the reaction temperature, which can be

attributed to an increase in the solid electrolyte ionic conductivity and, therefore, in

the amount of electrochemically transferred potassium ions at fixed potential. At

higher temperatures, the electrochemical activation occurred at higher potentials

(closer to +2 V). An increase in the promotional effect at higher temperatures was

also observed in previous EPOC studies with cationic solid electrolytes [30, 53].

However, in those studies the promotional effect usually increased up to an optimum

temperature was reached, which it was not the case in this study. On the other hand,

for the case of formaldehyde, an increase in the reaction temperature led to a decrease

in the H2CO selectivity (Figure 1.10b) probably due to the further oxidation of the

produced H2CO to CO2 [52].

Figure 1.11 shows the evolution of the production rates of H2, H2CO and CO2

as well as the CH3OH conversion with the O2/CH3OH ratio in the feed composition for

both: unpromoted (VWR = +2 V) and electropromoted (VWR < +2 V) conditions at a

fixed reaction temperature of 300 ºC.

Chapter 1

86

Figure 1.11. H2 (a), H2CO (b), CO2 (c) production rates and CH3OH conversion (d) variation

vs. O2/CH3OH ratio in the feed stream at 300 ºC for the unpromoted (+2 V) and electro-

promoted (-0.5 V) Pt film prepared by cathodic arc deposition.

0

1

2

3

4

5

6

7

8

r H

/ m

ol s

-1 x

10

-7

2 V

-0.5 V

0.0

0.4

0.8

1.2

1.6

r H

CO /

mo

l s-1

x 1

0-7

2 V

-0.5 V

0

10

20

30

40

50

60

0 0.25 0.5 0.75 1 1.25

O2/CH3OH ratio

CH

3O

H C

on

vers

ion

/ % 2 V

-0.5 V

0.0

0.5

1.0

1.5

2.0

2.5

3.0

r CO

/

mo

l s-1

x 1

0-6 2 V

-0.5 V

2

2

2

a)

b)

c)

d)

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

87

At negative polarization, a similar promotional effect on the methanol partial

oxidation activity was again found under all explored O2/CH3OH ratios. In this figure,

an optimum O2/CH3OH ratio can be observed for the partial oxidation mechanism vs.

the total oxidation, which has been typically observed in previous investigations using

Pt catalysts [17, 52]. At a low O2 concentration range, O2/CH3OH < 0.5, the partial

oxidation of methanol exhibited a positive order with respect to the O2 concentration.

In fact, the Pt catalytic activity was negligible in the absence of oxygen (methanol

decomposition conditions). A further increase in the O2 concentration enhanced the

total oxidation process and favoured the subsequent oxidation of the H2 and H2CO

obtained. In the case of O2/CH3OH > 0.5, a further increase in the O2 concentration

also led to a stronger poisoning effect on the partial oxidation mechanism under

electropromoted (VWR = -0.5 V) conditions vs. unpromoted conditions (VWR = +2 V)

due to the previously mentioned higher O2 coverage of the potassium-modified Pt

surface. For the case of the overall methanol reaction rate (partial oxidation and deep

oxidation), the methanol conversion curve exhibited positive order dependence with

respect to the oxygen concentration in the feed for the overall studied range. Hence, in

this case the increase in the O2 coverage under electropromoted conditions enhanced

the methanol reaction rate for the overall O2/CH3OH ratio, in good agreement with the

rules of the chemical and electrochemical promotion in heterogeneous catalysis [27,

54, 55].

iv) Stability study.

Figure 1.12 shows the variation in the production rate of H2 and H2CO vs. time

on stream at different repetitive applied potentials, at a temperature of 300 ºC and with

a feed composition of CH3OH/O2 = 7.2 %/4.6 %.

It can be observed a quite reproducible behaviour of both reaction rates during

the three cycles of positive and negative potentials. This completely reversible

NEMCA effect would demonstrate the stability of the solid electrolyte cell (working

electrode and solid electrolyte) under reaction conditions. The obtained negative

charged (not shown in the figure) upon the application of -1 V during the different

Chapter 1

88

cycles was the same for all of them and also agreed with the positive charge obtained

during the application of the catalyst potential of +2 V. This fact implies that the

amount of electrochemically transferred potassium ions from or to the solid electrolyte

during the negative and positive applications, respectively, was the same,

demonstrating the stability of promoting species under reaction conditions.

Figure 1.12. H2 (a) and H2CO (b) production rates variation vs. time under repetitive EPOC

transitions with the Pt film prepared by cathodic arc deposition. Partial oxidation conditions:

CH3OH/O2 = 7.2 %/4.6 %, 300 ºC.

Figure 1.13 shows the variation of the H2 and H2CO production rates for a

long, uninterrupted, negative polarization of -1 V for 8 h under the same reaction

conditions.

0

1

2

3

4

5

6

7

8

r H

/ m

ol s

-1 x

10

-7

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

3V

WR / V

0.0

0.4

0.8

1.2

1.6

0 100 200 300 400

Time / min

r H

CO

/ m

ol s

-1 x

10

-7

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

VW

R / V

2

2

2

2

rH

VWR

rH CO

VWR

a)

b)

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

89

Figure 1.13. H2 (a) and H2CO (b) production rates variation vs. time for a long EPOC

experiment during a durability test with the Pt film prepared by cathodic arc deposition. Partial

oxidation conditions: CH3OH/O2 = 7.2 %/4.6 %, 300 ºC.

The obtained results showed no lost of catalytic activity during the cathodic

polarization and, therefore, no lost of promotional phases under working conditions.

Previous studies with alkali-electropromoted systems have already demonstrated the

stability of this kind of systems even at higher working temperatures (above 350 ºC) in

a wide variety of catalytic systems such as the selective catalytic reduction of

N2O[53], the water gas-shift [56] or CO preferential oxidation [36]. Hence, the

obtained results demonstrate the interest of the proposed EPOC configuration for the

simultaneous production of H2 and H2CO in a wide variety of reaction conditions and

in a very stable manner in view of a possible practical application of this process.

0

1

2

3

4

5

6

7

8

r H

/ m

ol s

-1 x

10

-7

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

3

VW

R / V

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0 100 200 300 400 500 600

Time / min

r H

CO

/ m

ol s

-1 x

10

-7

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

VW

R / V

2

2

2

2

rH

VWR

rH CO

VWR

a)

b)

Chapter 1

90

1.3.2. Steam reforming of methanol

The Pt catalyst film prepared by CAD was also tested in the methanol steam

reforming (SRM) reaction with a feed composition of CH3OH/H2O = 4 %/4.8 % (N2

balance). H2, CO and CO2 were produced but neither H2CO nor other possible

methanol-derived compounds (such as CH4) were obtained under this reaction

atmosphere.

Figure 1.14 shows the variation of the different production rates vs. time under

open circuit and different applied potentials at 400 ºC. As in the previous experiments

under POM conditions, the application of a potential of +2 V (un-promoted state)

allowed to define a Pt catalyst film free of promoting ions and hence a reference state.

Potassium ions that probably migrated thermally to the Pt film were removed, leading

to the observed slight decrease in the catalytic activity during the first polarization.

The subsequent application of lower potentials increased the amount of

electrochemically transferred positive K+ ions to the Pt catalyst-working electrode.

Hence, the application of a catalyst potential of 0 V strongly enhanced the H2, CO and

CO2 production rates through the promotion of the methanol reforming reaction by K+

ions.

It should be noted that in these experiments the methanol conversion was kept

below 5 %, even in the promoted state. Unlike the previous POM results, the low

conversions obtained under the studied steam reforming conditions were more similar

to those typically found with this kind of electrocatalytic systems due to the low

surface area of the active catalyst film and the reactant bypass to the catalyst surface.

However, under optimum applied potential conditions (VWR = 0 V) the H2 production

rate was still increased in 5 times with respect to the un-promoted catalytic rate (VWR

= +2 V).

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

91

Figure 1.14. Influence of the applied potential (VWR) on the H2, CO2 and CO production rates

vs. time. Steam reforming conditions: CH3OH/H2O = 4 %/4.8 %, 400 ºC.

It can also be observed that the application of lower applied potentials (below

0 V) and, hence, the further increase in the promoter coverage, led to a decrease in the

catalytic activity of the system, which can be attributed to an excessive formation of

VWR

2

0

1

2

3

4

5

r H

/ m

ol s

-1 x

10

-8

-2.5

-1.5

-0.5

0.5

1.5

2.5

VW

R / V

0

3

6

9

12

r CO

/

mo

l s-1

x 1

0-9

-2.5

-1.5

-0.5

0.5

1.5

2.5

VW

R / V

0

3

6

9

12

0 100 200 300 400 500 600

Time / min

r CO

/ m

ol s

-1 x

10

-9

-2.5

-1.5

-0.5

0.5

1.5

2.5

VW

R / V

rH 2

VWR

rCO 2

rCO

VWR

2

a)

b)

c)

O.C.

O.C.

O.C.

Chapter 1

92

promoter-derived compounds [45] which may block catalytic active sites. Anyway,

these promoting species were completely removed when the potential was again fixed

at +2 V (unpromoted state), demonstrating the reversibility of the EPOC phenomena

in terms of catalytic activity. On the way of reaching the initial unpromoted values, a

strong increase in the H2, CO and CO2 production rates was firstly observed, which

was attributed to the progressive removal and decomposition of potassium-derived

carbonaceous products formed during the previous applied potentials. As reported in

previous works, potassium ions may form carbonates or bicarbonates species under

H2-, H2O-, CO- and CO2-containing atmospheres [35, 57]. This shows the ability of

this kind of systems to partially store reaction products [58], e.g., H2, which may be

also of significant interest for other applications.

Regarding the strong promotional effect previously mentioned (below 0 V),

one can rationalize the K+-induced variation in the catalytic activity on the basis of the

electrode work function modification [27, 59]. As mentioned in the previous section

under POM reaction conditions, a decrease in the catalyst potential involves a

decrease in the catalyst work function due the migration of positively charged

potassium species, thus weakening the Pt chemical bonds with the electron-donor

adsorbates (i.e., CH3OH), and strengthening those with the electron acceptor

adsorbates (i.e., H2O in this case, instead of O2). A similar promotional effect in the

presence of electro-positive ions has also been observed due to the activation of H2O

molecules in the steam reforming of methane [60] and water gas shift reaction[56].

In particular, previous studies on the mechanism of methanol reforming on Pt

[10] stated that the reaction proceeds via: (1) dissociative adsorption of methanol to

produce methoxy species and adsorbed hydrogen; (2) oxidative decomposition of the

methoxy groups by adsorbed water to liberate two hydrogen molecules and generate

surface formate; (3) water-assisted decomposition of formate to hydrogen and

unidentate carbonate; and (4) decomposition of carbonate to carbon dioxide. Then,

very likely, the water activation in the presence of K+ ions would favour the

decomposition of carbonaceous species (methoxy and formate species) increasing the

activity of the system. This explanation is also in good agreement with the observed

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

93

increase in the catalytic rate under unpromoted SRM conditions with respect to the

methanol decomposition reaction rate, i.e., in the absence of water (see, for instance,

O2/CH3OH ratio = 0, in Figure 1.11). Under methanol decomposition conditions a

negligible activity of the system was obtained, which demonstrates the positive effect

of water activation for the decomposition of the reaction intermediates and hence for

the methanol reforming catalytic process.

Figure 1.15 shows, under the same feed composition and different reaction

temperatures, the steady state variation of the H2 production rate vs. the applied

catalyst potential, and the variation of the optimum supplied negative charge which

maximized the H2 production rate in each case, along with the associated (maximum)

rate enhancement ratios, ρmax. The latter parameter was calculated by equation 1.5 and

the supplied charge was directly obtained from the integration of the current vs. time

curves.

As the reaction temperature raised, both higher promotional effects and

optimum supplied negative charges were observed. Thus, an increase of the

temperature not only favoured the methanol reforming kinetics [10] but also the

electrochemically assisted water activation. Above 400 ºC, the maximum rate

enhancement ratio reached a plateau at around 5, which was associated to the increase

in the unpromoted reaction rate (r0). These results demonstrate the interest of the

EPOC phenomena to promote to the optimal grade the performance of a metal catalyst

at varying reaction conditions (e.g., different temperatures).

Finally it is also interesting to note that the cathodic arc deposition method

allowed to prepare an electrochemical catalyst based on a very thin Pt layer (120 nm),

which has shown stable catalytic activity values for long experiments (10 h) under

POM and SRM conditions. Hence, this seems to be an appropriate method for the

application of thin metallic films for catalytic and electrocatalytic applications. Either

CAD or other kind of physical vapour deposition technique will be henceforth used

for the preparation of electrochemical catalysts.

Chapter 1

94

Figure 1.15. Influence of the applied potential (VWR) on the steady state H2 production rate (a),

and on the maximum rate enhancement ratio (ρmax) and the optimal supplied negative charge

(b) at different reaction temperatures. Steam reforming conditions: CH3OH/H2O = 4 %/4.8 %.

1.4. CONCLUSIONS

The following conclusions could be drawn from this study:

- The obtained results showed the interest of electrochemistry to improve the

activity of a metal catalyst in H2 production from methanol via partial oxidation

(POM) and steam reforming (SRM).

0

2

4

6

8

10

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5

r H

/

mo

l s-1

x 10

-8

VWR / V

450 ºC

400 ºC

360 ºC

320 ºC

0

0.04

0.08

0.12

0.16

0.2

0

1

2

3

4

5

6

300 330 360 390 420 450 480O

pti

mu

m n

ega

tive

ch

arge

/ C

ρm

ax

Temperature / ºC

2

a)

b)

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

95

- In contrast to the metal catalyst preparation method based on the

impregnation of a precursor solution, the cathodic arc deposition allowed to obtain a

very thin catalyst film (around 120 nm-thick) with a low metal loading (0.23 mg Pt

cm-2

). This kind of physical vapour deposition techniques is very useful for the

deposition of the catalysts employed in EPOC studies.

- The selectivity of the methanol partial oxidation reaction for H2 and H2CO

production can be greatly enhanced vs. the total combustion mechanism by

electrochemical promotion with K+ ions on a dense Pt film prepared by cathodic arc

deposition.

- The presence of electropositive K+ ions increased the metal bond strength

with the electron acceptor molecule, i.e., O2 in POM and H2O in SRM, favouring the

conversion of methoxy intermediates to hydrogen, and also to formaldehyde (in the

former case).

- This kind of electrochemically promoted catalysts may be of great interest

due to its ability of simultaneously producing H2 and H2CO with high yield in a single

reaction step. Its stability has also been verified for long working times.

- The EPOC behaviour of an alkali-based electrochemical catalyst (Pt/K-

βAl2O3) not only depends on the reaction conditions and the kinetic order of the

reactant molecules but also on the electrocatalytic activity of the catalyst-electrode

film (which is related to the deposition method) and its ability to form and store

promoter derived surface compounds, which may block active sites. This was not the

case of the cathodic arc deposited Pt. Hence, this technique or similar physical vapour

deposition methods were selected for subsequent experiments and chapters.

1.5. REFERENCES

[1] S. Dutta, A review on production, storage of hydrogen and its utilization as an energy

resource, Journal of Industrial and Engineering Chemistry, 20 (2014) 1148-1156.

[2] J.N. Armor, Catalysis and the hydrogen economy, Catalysis Letters, 101 (2005) 131-135.

Chapter 1

96

[3] J.D. Holladay, J. Hu, D.L. King, Y. Wang, An overview of hydrogen production

technologies, Catalysis Today, 139 (2009) 244-260.

[4] D.R. Palo, R.A. Dagle, J.D. Holladay, Methanol steam reforming for hydrogen production,

Chemical Reviews, 107 (2007) 3992-4021.

[5] R.M. Navarro, M.A. Peña, J.L.G. Fierro, Hydrogen production reactions from carbon

feedstocks: Fossil fuels and biomass, Chemical Reviews, 107 (2007) 3952-3991.

[6] E.L. Fornero, D.L. Chiavassa, A.L. Bonivardi, M.A. Baltanás, CO2 capture via catalytic

hydrogenation to methanol: Thermodynamic limit vs. 'kinetic limit, Catalysis Today, 172

(2011) 158-165.

[7] C. Cao, K.L. Hohn, Study of reaction intermediates of methanol decomposition and

catalytic partial oxidation on Pt/Al2O3, Applied Catalysis A: General, 354 (2009) 26-32.

[8] S. Sá, H. Silva, L. Brandão, J.M. Sousa, A. Mendes, Catalysts for methanol steam

reforming-A review, Applied Catalysis B: Environmental, 99 (2010) 43-57.

[9] G. Jacobs, B.H. Davis, In situ DRIFTS investigation of the steam reforming of methanol

over Pt/ceria, Applied Catalysis A: General, 285 (2005) 43-49.

[10] P. Tolmacsov, A. Gazsi, F. Solymosi, Decomposition and reforming of methanol on Pt

metals supported by carbon Norit, Applied Catalysis A: General, 362 (2009) 58-61.

[11] S.H. Ahn, O.J. Kwon, I. Choi, J.J. Kim, Synergetic effect of combined use of Cu-ZnO-

Al2O3 and Pt-Al2O3 for the steam reforming of methanol, Catalysis Communications, 10

(2009) 2018-2022.

[12] G. Kolb, S. Keller, S. Pecov, H. Pennemann, R. Zapf, Development of micro-structured

catalytic wall reactors for hydrogen production by methanol steam reforming over novel

Pt/In2O 3/Al2O3 catalysts, Chemical Engineering Transactions, 2011, pp. 133-138,

DOI:10.3303/cet1124023.

[13] A. Wan, C.t. Yeh, Ignition of methanol partial oxidation over supported platinum catalyst,

Catalysis Today, 129 (2007) 293-296.

[14] M.P. Zum Mallen, L.D. Schmidt, Oxidation of methanol over polycrystalline Rh and Pt:

Rates, OH desorption, and model, Journal of Catalysis, 161 (1996) 230-246.

[15] N. Kizhakevariam, E.M. Stuve, Promotion and poisoning of the reaction of methanol on

clean and modified platinum (100), Surface Science, 286 (1993) 246-260.

[16] B.E. Traxel, K.L. Hohn, Partial oxidation of methanol at millisecond contact times,

Applied Catalysis A: General, 244 (2003) 129-140.

[17] R. Ubago-Pérez, F. Carrasco-Marín, C. Moreno-Castilla, Methanol partial oxidation on

carbon-supported Pt and Pd catalysts, Catalysis Today, 123 (2007) 158-163.

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

97

[18] G. Reuss, W. Disteldorf, A.O. Gamer, A. Hilt, Formaldehyde, Ullmann's Encyclopedia of

Industrial Chemistry, Wiley-VCH, Weinheim, 2002.

[19] G.C. Behera, K. Parida, Selective gas phase oxidation of methanol to formaldehyde over

aluminum promoted vanadium phosphate, Chemical Engineering Journal, 180 (2012) 270-276.

[20] N. Iwasa, T. Mayanagi, W. Nomura, M. Arai, N. Takezawa, Effect of Zn addition to

supported Pd catalysts in the steam reforming of methanol, Applied Catalysis A: General, 248

(2003) 153-160.

[21] M. Kusche, F. Enzenberger, S. Bajus, H. Niedermeyer, A. Bösmann, A. Kaftan, M.

Laurin, J. Libuda, P. Wasserscheid, Enhanced activity and selectivity in catalytic methanol

steam reforming by basic alkali metal salt coatings, Angewandte Chemie - International

Edition, 52 (2013) 5028-5032.

[22] H.N. Evin, G. Jacobs, J. Ruiz-Martinez, U.M. Graham, A. Dozier, G. Thomas, B.H.

Davis, Low temperature water-gas shift/methanol steam reforming: Alkali doping to facilitate

the scission of formate and methoxy C-H bonds over Pt/ceria catalyst, Catalysis Letters, 122

(2008) 9-19.

[23] L. Alejo, R. Lago, M.A. Peña, J.L.G. Fierro, Partial oxidation of methanol to produce

hydrogen over Cu-Zn-based catalysts, Applied Catalysis A: General, 162 (1997) 281-297.

[24] M. Stoukides, C.G. Vayenas, The effect of electrochemical oxygen pumping on the rate

and selectivity of ethylene oxidation on polycrystalline silver, Journal of Catalysis, 70 (1981)

137-146.

[25] C.G. Vayenas, S. Bebelis, S. Ladas, Dependence of catalytic rates on catalyst work

function, Nature, 343 (1990) 625-627.

[26] S. Souentie, L. Lizarraga, E.I. Papaioannou, C.G. Vayenas, P. Vernoux, Permanent

electrochemical promotion of C3H8 oxidation over thin sputtered Pt films, Electrochemistry

Communications, 12 (2010) 1133-1135.

[27] C.G. Vayenas, S. Bebelis, C. Pliangos, S. Brosda, D. Tsiplakides, Electrochemical

Activation of Catalysis: Promotion, Electrochemical Promotion and Metal-Support

Interactions, Kluwer Academic Publishers/Plenum Press, New York, 2001.

[28] A. Katsaounis, Recent developments and trends in the electrochemical promotion of

catalysis (EPOC), Journal of Applied Electrochemistry, 40 (2010) 885-902.

[29] D. Tsiplakides, S. Balomenou, Electrochemical promoted catalysis: Towards practical

utilization, Chemical Industry and Chemical Engineering Quarterly, 14 (2008) 97-105.

Chapter 1

98

[30] A. de Lucas-Consuegra, New trends of Alkali Promotion in Heterogeneous Catalysis:

Electrochemical Promotion with Alkaline Ionic Conductors, Catalysis Surveys from Asia,

2015, DOI:10.1007/s10563-014-9179-6.

[31] P. Vernoux, L. Lizarraga, M.N. Tsampas, F.M. Sapountzi, A. De Lucas-Consuegra, J.L.

Valverde, S. Souentie, C.G. Vayenas, D. Tsiplakides, S. Balomenou, E.A. Baranova, Ionically

conducting ceramics as active catalyst supports, Chemical Reviews, 113 (2013) 8192-8260.

[32] P. Vernoux, F. Gaillard, C. Lopez, E. Siebert, Coupling catalysis to electrochemistry: A

solution to selective reduction of nitrogen oxides in lean-burn engine exhausts?, Journal of

Catalysis, 217 (2003) 203-208.

[33] F.J. Williams, N. Macleod, M.S. Tikhov, R.M. Lambert, Electrochemical promotion of

bimetallic Rh-Ag/YSZ catalysts for the reduction of NO under lean burn conditions,

Electrochimica Acta, 47 (2002) 1259-1265.

[34] E.I. Papaioannou, S. Souentie, A. Hammad, C.G. Vayenas, Electrochemical promotion of

the CO2 hydrogenation reaction using thin Rh, Pt and Cu films in a monolithic reactor at

atmospheric pressure, Catalysis Today, 146 (2009) 336-344.

[35] A.J. Urquhart, F.J. Williams, R.M. Lambert, Electrochemical promotion by potassium of

Rh-catalysed fischer-tropsch synthesis at high pressure, Catalysis Letters, 103 (2005) 137-141.

[36] A. de Lucas-Consuegra, A. Princivalle, A. Caravaca, F. Dorado, C. Guizard, J.L.

Valverde, P. Vernoux, Preferential CO oxidation in hydrogen-rich stream over an

electrochemically promoted Pt catalyst, Applied Catalysis B: Environmental, 94 (2010) 281-

287.

[37] S. Neophytides, C.G. Vayenas, Non-faradaic electrochemical modification of catalytic

activity. 2. The case of methanol dehydrogenation and decomposition on Ag, Journal of

Catalysis, 118 (1989) 147-163.

[38] C.G. Vayenas, S. Neophytides, Non-faradaic electrochemical modification of catalytic

activity III. The case of methanol oxidation on Pt, Journal of Catalysis, 127 (1991) 645-664.

[39] C.A. Cavalca, G. Larsen, C.G. Vayenas, G.L. Haller, Electrochemical modification of

CH3OH oxidation selectivity and activity on a Pt single-pellet catalytic reactor, Journal of

Physical Chemistry, 97 (1993) 6115-6119.

[40] J.K. Hong, I.H. Oh, S.A. Hong, W.Y. Lee, Electrochemical oxidation of methanol over a

silver electrode deposited on yttria-stabilized zirconia electrolyte, Journal of Catalysis, 163

(1996) 95-105.

Electrochemical promotion of Pt for H2 production from methanol partial oxidation and steam

reforming

99

[41] F. Dorado, A. de Lucas-Consuegra, P. Vernoux, J.L. Valverde, Electrochemical

promotion of platinum impregnated catalyst for the selective catalytic reduction of NO by

propene in presence of oxygen, Applied Catalysis B: Environmental, 73 (2007) 42-50.

[42] A. Anders, Cathodic Arcs: From Fractal Spots to Energetic Condensation. Springer Series

on Atomic, Optical, and Plasma Physics, Springer-Verlag, New York, 2008.

[43] E. Seker, E. Gulari, Improved N2 selectivity for platinum on alumina prepared by sol-gel

technique in the reduction of NOx by propene, Journal of Catalysis, 179 (1998) 339-342.

[44] E. Mutoro, C. Koutsodontis, B. Luerssen, S. Brosda, C.G. Vayenas, J. Janek,

Electrochemical promotion of Pt(111)/YSZ(111) and Pt-FeOx/YSZ(111) thin catalyst films:

Electrocatalytic, catalytic and morphological studies, Applied Catalysis B: Environmental,

100 (2010) 328-337.

[45] P. Vernoux, F. Gaillard, C. Lopez, E. Siebert, In-situ electrochemical control of the

catalytic activity of platinum for the propene oxidation, Solid State Ionics, 175 (2004) 609-613.

[46] N. Kotsionopoulos, S. Bebelis, In situ electrochemical modification of catalytic activity

for propane combustion of Pt/β-Al2O3 catalyst-electrodes, Topics in Catalysis, 44 (2007) 379-

389.

[47] N. Kotsionopoulos, S. Bebelis, Electrochemical characterization of the Pt/β″-Al 2O3

system under conditions of in situ electrochemical modification of catalytic activity for

propane combustion, Journal of Applied Electrochemistry, 40 (2010) 1883-1891.

[48] H.P. Klug, L.E. Alexander, X-Ray Diffraction Procedures for Polycrystalline and

Amorphous Materials, Wiley, New York, 1974.

[49] F.J. Williams, M.S. Tikhov, A. Palermo, N. Macleod, R.M. Lambert, Electrochemical

promotion of rhodium-catalyzed NO reduction by CO and by propene in the presence of

oxygen, Journal of Physical Chemistry B, 105 (2002) 2800-2808.

[50] I.V. Yentekakis, R.M. Lambert, M. Konsolakis, V. Kiousis, The effect of sodium on the

Pd-catalyzed reduction of NO by methane, Applied Catalysis B: Environmental, 18 (1998)

293-305.

[51] A. Peremans, F. Maseri, J. Darville, J.M. Gilles, Interaction of methanol with a

polycrystalline platinum surface studied by infrared reflection absorption spectroscopy,

Surface Science, 227 (1990) 73-78.

[52] R.W. McCabe, D.F. McCready, Kinetics and reaction pathways of methanol oxidation on

platinum, Journal of Physical Chemistry, 90 (1986) 1428-1435.

Chapter 1

100

[53] A. de Lucas-Consuegra, F. Dorado, C. Jiménez-Borja, J.L. Valverde, Influence of the

reaction conditions on the electrochemical promotion by potassium for the selective catalytic

reduction of N2O by C3H6 on platinum, Applied Catalysis B: Environmental, 78 (2008) 222-

231.

[54] C. Vayenas, S. Brosda, Electrochemical promotion: Experiment, rules and mathematical

modeling, Solid State Ionics, 154-155 (2002) 243-250.

[55] S. Brosda, C.G. Vayenas, J. Wei, Rules of chemical promotion, Applied Catalysis B:

Environmental, 68 (2006) 109-124.

[56] A. De Lucas-Consuegra, A. Caravaca, J. González-Cobos, J.L. Valverde, F. Dorado,

Electrochemical activation of a non noble metal catalyst for the water-gas shift reaction,

Catalysis Communications, 15 (2011) 6-9.

[57] A. de Lucas-Consuegra, F. Dorado, J.L. Valverde, R. Karoum, P. Vernoux, Low-

temperature propene combustion over Pt/K-βAl2O3 electrochemical catalyst:

Characterization, catalytic activity measurements, and investigation of the NEMCA effect,

Journal of Catalysis, 251 (2007) 474-484.

[58] A. de Lucas-Consuegra, A. Caravaca, F. Dorado, J.L. Valverde, Pt/K-βAl2O3 solid

electrolyte cell as a "smart electrochemical catalyst" for the effective removal of NOx under

wet reaction conditions, Catalysis Today, 146 (2009) 330-335.

[59] A. Nakos, S. Souentie, A. Katsaounis, Electrochemical promotion of methane oxidation

on Rh/YSZ, Applied Catalysis B: Environmental, 101 (2010) 31-37.

[60] A. De Lucas-Consuegra, A. Caravaca, P.J. Martínez, J.L. Endrino, F. Dorado, J.L.

Valverde, Development of a new electrochemical catalyst with an electrochemically assisted

regeneration ability for H2 production at low temperatures, Journal of Catalysis, 274 (2010)

251-258.

101

Chapter 2

ELECTROCHEMICAL PROMOTION OF Pt

NANOPARTICLES DISPERSED IN A CARBON MATRIX

FOR METHANOL CONVERSION:

TOWARDS MORE COMPETITIVE CATALYSTS OF LOW METAL

LOADING

This study reports the electrochemical promotion (EPOC) of Pt nanoparticles

(of around 3 nm) dispersed in a diamond-like carbon (DLC) matrix by means of the

cathodic arc deposition technique. After a temperature-programmed pretreatment, the

catalyst film achieved a suitable electrical conductivity due to the transition of the sp3-

hybridized carbon form into a more graphitic structure (sp2-hybridized). The catalytic

performance of the Pt-DLC film in the methanol partial oxidation (POM) and steam

reforming (SRM) reactions for H2 production has been promoted by K+ ions

electrochemically transferred from a K-βAl2O3 solid electrolyte. Moreover, two

different electropromotional effects have been found under POM conditions

depending on the applied potential, which can be attributed to the formation of

different kinds of promoter phases on this catalyst. The higher catalytic activity of Pt-

DLC, compared to that of the pure Pt film, demonstrates the practical interest of this

kind of dispersed catalyst films with very a low metal loading.

0

1

2

3

4

5

6

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5

r H

/ m

ol s

-1x

10

-8

VWR / V

Pt 320 ºC (0.46 mg)

Pt 360 ºC (0.46 mg)

Pt/DLC 320 ºC (0.03 mg)

Pt/DLC 360 ºC (0.03 mg)

2

2 ρH ,max = 3.37

ρH ,max = 3.54

2

ρH ,max = 1.86

2

ρH ,max = 1.53

2

ΔV < 0

Pt-DLC

K+ K+ K+

e-

e-

K-βAl2O3

Au

Δr

Catalyst film (Pt-DLC)

Dr

K+ ionsΔV Solid electrolyte (K-bAl2O3)

Counter/Reference (Au)

CH3OH

H2, CO,CO2, (H2CO)

10 nm

CH3OH/H2O = 4 %/4.8 %

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

0

1

2

3

4

5

6

0 100 200 300 400 500 600 700 800

VW

R/

V

r H

/ m

ol s

-1x

10

-8

Time / min

250 ºC

320 ºC

360 ºC

VWR

2

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

0

0.5

1

1.5

2

2.5

3

0 100 200 300 400 500 600 700 800

VW

R/

V

r H

/ m

ol s

-1cm

-2x

10

-8

Time / min

250 ºC

320 ºC

360 ºC

VWR

CH3OH/H2O = 4 %/4.8 %

2

CH3OH,

O2 / H2O

H2, CO2,

CO, (H2CO)

4 % CH3OH

4.8 % H2O

Chapter 2

102

2.1. INTRODUCTION

As already pointed out, the present research is mainly focused on the

application of the phenomenon of Electrochemical Promotion Of Catalysis (EPOC)

[1] to H2 production catalytic processes by using CH3OH as the raw material. Indeed,

methanol is one of the most important feedstock in the chemical industry, and it is also

a very promising liquid energy carrier that can be used as a fuel in Direct Methanol

Fuel Cells (DMFC) [2, 3] or as a hydrogen source through the steam reforming (SRM)

and partial oxidation (POM) reactions [3-5]. As already mentioned, the main

advantages of using methanol for H2 production are the following: high energy

density, easy availability, and safe storage and transport conditions.

In the alkali electrochemical promotion experiments carried out in the first

chapter, Pt was selected as the active phase since this metal is one of the most

commonly employed catalysts in both SRM [6-9] and POM [10-14]. Hence, the EPOC

phenomenon showed to improve the catalytic behaviour of Pt by means of the

controlled electrochemical pumping of K+ promoter ions from a K-βAl2O3 solid

electrolyte (which also acted as catalyst support). Concretely, H2 production from

SRM and POM was enhanced up to 5 and 6 times through the electrochemical

promotion of a fairly dense Pt catalyst prepared by Cathodic Arc Deposition (CAD).

However, in view of a possible practical application of this kind of systems, novel

catalyst films with higher metal particle dispersions and with a more competitive cost

should be investigated [15, 16]. The use of conductive catalyst films of pure metal

implies high costs of these electro-catalytic configurations making it difficult to

compete with conventional supported catalysts (with dispersions typically above 10

%). This point is essential for a further commercialization step of the EPOC

phenomena for the H2 production and for other catalytic systems. In this sense, recent

efforts have been done in previous studies by using a second conductive material that

can simultaneously act as both catalyst particles support and electronic conductor

material: gold [16], graphite [17], carbon [18], carbon nanofibers [19] or mixed ionic

electronic composite materials [20, 21].

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

103

In this work, we propose for the first time in literature the use of a novel Pt-

diamond like carbon (Pt-DLC) catalyst film prepared by the technique of cathodic arc

deposition. DLC is the denomination of an amorphous form of carbon in which the

bonding is composed of a mixture of diamond-type (sp3) and graphite-type (sp

2)

hybridization, exhibiting properties intermediate between them [22]. DLC presents

very unique characteristics to be used as catalyst particles support in an electrode film

such as: large potential window, high strength, thermal conductivity, low cost and

chemical inertness [23]. As expected for diamond, DLC is stable to high temperatures

(800 °C) and hence it could be used in a large number of catalytic processes.

Moreover, its properties can be further modified (improved) by the introduction of

metallic elements. In this way, the metal is dispersed on the diamond-like carbon. At

the same time, this addition should both decrease the internal stress and increase its

electrical conductivity [24, 25]. These kinds of composites (metal-DLC) are acquiring

great importance due to their potential applications, including microelectrodes in

electrochemical analysis and field emission properties [26, 27]. Pt-DLC films have

already been prepared for other applications in several previous studies by pulsed laser

deposition [24] and sputter deposition [25, 28-30]. However, to the best of our

knowledge there are no previous studies on the use of this kind of composites in

catalysis.

The aim of this study is to electrochemically promote a novel, thin, Pt-DLC

catalyst film of low metal loading where the Pt nanoparticles dispersed in the DLC

matrix could be electrochemically promoted. Thus, a Pt-DLC/K-βAl2O3

electrochemical catalyst has been prepared, characterized and tested for the H2

production via POM and SRM. The effect of the electrochemical pumping of

potassium ions on the active Pt-DLC catalyst film has been investigated under a wide

range of reaction conditions and explained according to the EPOC rules [1]. Finally,

the obtained EPOC behaviour of the Pt-DLC film has been compared with that shown

in the first chapter with a pure dense Pt film prepared by the same technique (cathodic

arc deposition) in order to evaluate the benefit of the Pt-DLC system proposed.

Chapter 2

104

2.2. EXPERIMENTAL

2.2.1. Preparation of the electrochemical catalyst

The electrochemical catalyst consisted of a continuous, thin film (geometric

area of 2.01 cm2) composed of Pt nanoparticles dispersed in a diamond-like carbon

matrix (Pt-DLC), which was deposited on a side of a 19-mm-diameter, 1-mm-thick K-

βAl2O3 (Ionotec) disc. As in the first chapter, inert Au counter (CE) and reference

(RE) electrodes were firstly deposited on the other side of the electrolyte by the

application of coatings of gold paste (Gwent Electronic Materials C1991025D2)

followed by calcination at 800 ºC for 2 h (heating ramp of 5 ºC min-1

). Then, the active

Pt-DLC catalyst film, which also behaved as a working electrode (WE), was deposited

on the other side of the electrolyte by the filter cathodic arc deposition technique [31].

The deposition system consisted of two cathodes contained in a mini-gun source

designed to operate in pulsed mode [32]. The pulses used in the production of carbon

and platinum plasmas were 200 A and had a duration of 1 ms. Plasma produced by the

source was injected into a 90-degree filter contained in a cross-sectional chamber to

remove the majority of macroparticles which were created during the cathodic arc

process. During the deposition, the substrates were biased with negative 500 V pulses

that were 1 µs long and had a duty cycle of 10 %. The pulse repetition rate was set at 2

pulses per second and the number of pulses per deposition was 25000. During the

deposition, the substrate holder was rotated at a speed of 2 revolutions per minute. The

residual gas pressure was around 10-4

Pa. Hence, this technique allowed the

preparation of a thin (150 nm-thick), highly dense, Pt-DLC film which showed good

adhesion to the substrate. The final Pt loading was 0.014 mg Pt cm-2

.

This catalyst preparation method was carried out in collaboration with Dr.

José Luis Endrino from the Institute of Materials Science of Madrid (CSIC).

The resulting Pt-DLC/K-βAl2O3/Au electrochemical catalyst was placed into

the single chamber solid electrolyte cell reactor. Since the obtained catalyst film was

electrically nonconductive, it was subjected, prior to the catalytic activity

measurements, to a 25 % H2 stream (N2 balance) with an overall flow rate of 3.6 L h-1

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

105

under temperature-programmed conditions (3.5 ºC min-1

) from room temperature to

320 ºC, in an effort to decrease its electrical resistance while keeping platinum in its

reduced state, as will be explained in the discussion section. All the electrodes were

connected to the potentiostat-galvanostat (PGSTAT320-N, Metrohom Autolab) as

reported in the first chapter, but in this case a second electric contact was also added to

the catalyst-working electrode and connected to a digital multimeter, in order to

monitor the in-plane electrical resistance of the Pt-DLC under working conditions.

2.2.2. Characterization measurements

The deposited Pt-DLC catalyst film was first characterized via scanning

electron microscopy (SEM) and EDX analysis using a FEI Nova NANOSEM 230

instrument, in order to investigate the morphology of the Pt-DLC film and identify and

approximately quantify the amount of the different elements (Pt and C). The

Temperature Programmed Stabilization process (TPS) prior to the catalytic

experiments was in-situ monitored by recording the variation of the in-plane electrical

resistance between the two contacts (1 cm separated) located on the Pt-DLC film vs.

temperature.

The cross-sectional properties of the Pt-DLC catalyst film were studied before

and after the H2 stabilization treatment by transmission electron microscopy (TEM).

Cross-section TEM lamellas of samples were fabricated in a FEI Helios 600 Nanolab

Dual-Beam system. The specimens were pre-thinned with an acceleration voltage of

30 kV of the ion column and a final thinning was performed at 5 kV to minimize any

possible preparation artifacts. TEM work was carried out using a probe corrected FEI

Titan 60-300. Scanning Transmission Electron Microscopy (STEM) imaging in bright

field (BF), and high-angle annular dark field (HAADF) modes combined with

Electron Energy Loss Spectroscopy (EELS) were also used to characterize the

morphology, chemical composition and electronic structure of the samples. Cyclic

voltammetry (CV) measurements were also in-situ performed with the potentiostat-

galvanostat under SRM and POM reaction conditions at a fixed temperature (360 ºC),

to investigate the formation and decomposition of promoter-derived species. Before

Chapter 2

106

performing these cyclic voltammetries, the sample was kept at +2 V for 30 min in

order to define a reference state (a Pt-DLC film clean of promoter ions). Then, the

potential was varied at a sweep rate of 5 mV s-1

from +2 to -2 V.

2.2.3. Catalytic activity measurements

The catalytic activity measurements were carried out in the experimental setup

described in section 1.2.3. The electrochemical promotion experiments were

performed at atmospheric pressure with an overall gas flow rate of 6 NL h-1

, a

temperature ranging from 250 to 320 ºC with a feed composition of CH3OH/O2 = 11

%/0.9 % (N2 balance) for the POM experiments, and a temperature ranging from 250

to 360 ºC with a feed composition of CH3OH/H2O = 4 %/4.8 % (N2 balance) for the

SRM experiments. H2, CO2, CO, H2O and H2CO were the only detected products, the

latter two appearing only under POM atmosphere. The error in the carbon atom

balance did not exceed 5 % in any experiment.

2.3. RESULTS AND DISCUSSION

2.3.1. Development of a catalyst with suitable structural and electrical properties

Figure 2.1 shows the top surface SEM micrograph of the prepared Pt-DLC

catalyst film along with the elemental EDX analysis.

Figure 2.1. SEM micrograph of the Pt-DLC catalyst film (a) and its corresponding EDX

analysis (b).

2.6330.5Pt

97.3769.5C

At / %Wt / %Element

2.6330.5Pt

97.3769.5C

At / %Wt / %Element

2.6330.5Pt

97.3769.5C

At / %Wt / %Element

2.6330.5Pt

97.3769.5C

At / %Wt / %Element

a) b)

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

107

It can be observed that the obtained catalyst film was dense and exhibited a

very smooth surface in agreement with a previous work related to the preparation of

Pt-DLC films by sputtering [25]. The EDX analysis confirmed the low metal loading

(30 wt. % Pt) on the electrode, being carbon (70 wt. %) the other component. Hence,

no impurities of other elements due to possible contamination during the preparation

step in the chamber were detected in the EDX analysis of the film. It should also be

mentioned that due to the very low thickness of the catalyst layer (150 nm) and to the

surface character of the system, it is very likely that most of the deposited Pt

nanoparticles may have access to the gas phase reactants.

Figure 2.2 shows the variation of the planar electrical resistance of the Pt-DLC

catalyst surface vs. temperature during the TPS experiment (3.5 ºC/min) under a 25 %

H2 stream (N2 balance). The inset of the figure zooms the final variation of the

electrical resistance at the end of the experiment.

Figure 2.2. In-plane electrical resistance variation (R) during the Temperature Programmed

Stabilization (TPS). Heating ramp of 3.5 °C/min under a 25 % H2 stream (N2 balanced).

0

5

10

15

20

25

30

35

40

45

0 10 20 30 40 50 60 70 80 90 100

Time / min

In-p

lan

e e

lect

rica

l re

sist

ance

/ Ω

x 1

06

0

50

100

150

200

250

300

350

Tem

pe

ratu

re /

ºC

Resistance

Temperature

3

4

5

6

7

8

9

80 85 90 95 100Time / min

Re

sist

ance

/

275

285

295

305

315

325

Tem

pe

ratu

re /

ºC

K-βAl2O3

Pt-DLC

Au

R

K-βAl2O3

Pt-DLC

Au

R

Chapter 2

108

It can be observed that at the beginning of the TPS experiment, the deposited

Pt-DLC catalyst film showed a high value of electrical resistance (42 x 106 Ω). It

seems to indicate that the as-prepared DLC film was on its most sp3-hybridized form

(similar to diamond). Above 200 ºC, the electrical resistance of the catalyst film

sharply decreased and finally stabilized at 3.5 kΩ at 315 ºC (zoom in Figure 2.2) due

to the transition toward a structure more similar to that of graphite (sp2-hybridized), as

will be confirmed below. These results are consistent with those obtained in previous

works [33-35] with DLC films prepared by filtered cathodic vacuum arc, in which a

transition temperature from a highly tetrahedral structure to that of a graphitic nature

took place abruptly around 200 ºC. The final achieved electrical resistance of the Pt-

DLC film (3.5 kΩ) was enough to electrochemically promote the active Pt particles as

will be also demonstrated later. This electrical resistance value was almost constant

during all the catalytic EPOC experiment, which also shows the stability of the

deposited Pt-DLC film on its most sp2-hybridized structure.

A microstructural characterization by HAADF-STEM in cross-sections of the

Pt and the Pt-DLC films, the latter both before and after the TPS treatment, is shown

in Figure 2.3. The pure Pt film employed in the first chapter (Figure 2.3a) shows a

highly dense crystalline platinum with a grain size of the order of the film thickness

(100 nm). The HAADF images of the composite Pt-DLC film before and after the

TPS (Figures 2.3b and 2.3c, respectively) indicate the formation of a nanocomposite

multilayer with different metal contents. Multilayer growth was due to the use of a

high energetic deposition technique such as a filter cathodic arc, where the high

kinetic energy of carbon and platinum ions sculpted layers with high and low metal

contents. HAADF images were characterized by presenting Z contrast, in which

brighter areas were composed by heavier atoms. On the other hand, darker regions

were composed by lighter materials. BF imaging contrast was reversed with respect to

HAADF. Therefore, the difference in contrast between the layers would correspond to

differences in the composition of the Pt-DLC thin film. The dark zones corresponded

to carbon-rich layers and the light ones to platinum-rich ones. It is interesting to note

that the same layered structure was maintained after the TPS treatment (Figure 2.3c),

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

109

which would indicate that no platinum segregation or diffusion occurred as a result of

the thermal stabilization process. This fact could indirectly indicate that the increase in

the conductivity was mostly due to a modification of the carbon matrix.

Figure 2.3. HAADF-STEM images of a pure Pt film (a), as-deposited Pt-DLC film (b) and Pt-

DLC film after TPS (c).

a)

b)

c)Fig. 2.4. g,h

20 nm

20 nm

50 nm

Fig. 2.4. e,f

Fig. 2.4. c,d

Fig. 2.4. a,b

Chapter 2

110

In order to gain in-depth structural characterization information of the

composite sample after TPS, Figure 2.4 shows both BF and HAADF images of the

sample taken in the regions indicated in Figure 2.3c.

Figure 2.4. BF-STEM images of bottom (a), middle (c,e), and top (g) of Pt-DLC film after

TPS, along with their corresponding Z-contrast HAADF-STEM images (b,d,f,h).

The morphology of the sample would correspond to that of a network of small

Pt nanoparticles of around 3 nm (bright in HAADF) surrounding carbon-rich areas

g)

e)

h)

a)

c) d)

f)

b)

10 nm

10 nm

10 nm

10 nm

f)

g)

e)

h)

a)

c) d)

f)

b)

10 nm

10 nm

10 nm

10 nm

f)

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

111

(dark in HAADF). Assuming this Pt particle size, a number of Pt active sites of 5.74 x

10-8

mol has been estimated from the total number of Pt deposited in the sample (1.44

x 10-7

mol) and hence an approximate dispersion value of around 40 %. The size of the

carbon-rich areas varied along the growth direction of the film.

As shown in Figures 2.4a and 2.4b, a network of small Pt particles embedded

in a DLC matrix started to grow quickly on the top of a very thin amorphous carbon

layer which firstly nucleated on the substrate. However, after 30 nm, a thin 10 nm

platinum-poor layer was formed due to formation of stronger carbon bonds, which

slowed down the incorporation of larger amounts of platinum. It is worth noting, as

can be observed in Figure 2.4e and 2.4f (with higher magnification), that a similar

carbon rich-layer was formed again after some deposition time. The intermediate part

of the sample was mostly composed of a continuous network of extremely small

platinum grains, 1-3 nm, distributed in the hydrogen-free carbon matrix, which

appears to percolate (Figures 2.4c and 2.4d). TEM micrographs from the top of the

TPS-treated sample are shown in Figures 2.4g and 2.4h. It is particularly interesting to

note that the size of the platinum grains at the top of the samples was of the same

order of magnitude as that of platinum grains at the bottom, which would imply that

there was no agglomeration of platinum grains or metal surface segregation. This

underscores the usability of the cathodic arc deposition method in synthesizing

optimum active platinum sites for catalytic purposes.

Given that the TPS treatment did not result in any significant microstructural

change of the nanocomposite Pt-DLC, EEL spectra of carbon were collected in the

different areas shown in the micrographs of Figure 2.4. Figure 2.5 shows the C K-edge

spectra of the samples before and after TPS treatment (denoted as As-dep and TPS,

respectively) along the standard ones corresponding to amorphous carbon and Highly

Ordered Pyrolytic Graphite (HOPG carbon).

Chapter 2

112

Figure 2.5. EEL spectra (1-4) acquired from top to bottom for the same regions shown in

Figure 4 (a-b,c-d,e-f,g-h) before and after TPS. Graphitized carbon and amorphous carbon are

shown for comparison purposes.

These spectra indicated a graphitization of the carbon-rich matrix all along the

film, as it can be observed by the increase of the π* peak in those layers after TPS

treatment. Therefore, the previously observed decrease in the surface electrical

resistance of the film after the TPS can be related to the conversion of sp3-bonds in the

amorphous carbon matrix to sp2 hybridization with higher conductivity.

2.3.2. Electrochemical promotion experiments

In order to study the EPOC phenomena of the Pt-DLC catalyst film in the

hydrogen production from methanol, several potentiostatic experiments were carried

out under different reaction atmospheres (POM and SRM) and different reaction

temperatures.

280 290 300 310 320

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Graphitized carbon

Amorphous carbon

As-dep

TpS

(4)

(3)

(2)

Energy Loss (eV)

No

rma

lize

d C

ou

nts

1s to

eV(1)

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

113

Figure 2.6 shows the variation of the different product formation rates (H2,

CO2, CO and H2CO) vs. time at different applied potentials and reaction temperatures

under POM conditions (CH3OH/O2 = 11 %/0.9 %, N2 balance). In agreement with

previous EPOC studies carried out with cationic electrochemical catalysts [36, 37], the

application of a high positive potential of 2 V (unpromoted state) at both the

beginning and the end of each experiment made it possible to define a clean catalyst-

working electrode film free of promoter ions and hence a reference state. The

subsequent decrease in the applied catalyst potential (VWR < +2 V) increased the

amount of positive K+ ions electrochemically transferred to the Pt catalyst-working

electrode film, according to the obtained negative currents (not shown here). It

allowed to obtain two different kinds of electrochemically promoted states: promoted

state I (VWR = +0.5 V) with an increase in the H2 and CO2 production rates (a slight

increase was also observed in CO production at 320 ºC) and promoted state II (VWR =

-0.5 V) with an increase in the H2, CO and H2CO production rates. At this point, it

should be mentioned that both promotional states were reproducible and appeared at

the three explored reaction temperatures. Moreover, in each promoted state,

reproducible optimum values of the product formation rates were attained, as typically

observed for alkali electropromotion reactions [38, 39], due to the occurrence of an

optimum promoter coverage on the catalyst surface (an excess of alkali promoter

coverage led to a poisoning effect).

The observed modifications of the catalytic activity vs. the applied potential

could be explained in terms of work function modifications as a consequence of the

back-spillover of electropositive ions from the solid electrolyte (K-βAl2O3) to the

catalyst-working electrode film (Pt-DLC) [40]. According to the current rules of

electrochemical promotion [1], the migration of K+ ions onto the catalyst-working

electrode (Pt-DLC surface) weakened the platinum chemical bond with the electron

donor adsorbates and strengthened that with electron acceptors. These donors and

acceptors have been identified as methanol and oxygen, respectively, under methanol

partial oxidation conditions [41, 42].

Chapter 2

114

Figure 2.6. H2 (a), CO2 (b), CO (c) and H2CO (d) production rates vs. time at different applied

potentials (VWR) and temperatures. Partial oxidation conditions: CH3OH/O2 = 11 %/0.9 %.

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

0

1

2

3

4

5

6

VW

R/

V

r CO

/

mo

l s-1

x 1

0-7

250 ºC

280 ºC

320 ºC

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

0

1

2

3

4

5

VW

R/

V

r CO

/ m

ol

s-1x

10

-7

250 ºC

280 ºC

320 ºC

z

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

0

0.5

1

1.5

2

2.5

3

VW

R/

V

r H

/ m

ol s

-1x

10

-7

250 ºC 280 ºC

320 ºC

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

0

0.5

1

1.5

0 100 200 300 400 500 600 700 800 900

VW

R/

V

r H

CO

/ m

ol s

-1x

10

-7

Time / min

250 ºC

280 ºC

320 ºC

b)

c)

a)

d)

VWR

VWR

VWR

VWR

Un-promoted

state

Promoted

state I

Un-promoted

state

Promoted

state II

2

2

2

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

115

As stated in the first chapter for a dense Pt film, the observed increase in H2,

CO and H2CO production rates (promoted state II) could be explained by the strong

promotional effect of K+

ions in the methanol partial oxidation mechanism, which was

in turn attributed to an increase in metal-oxygen bonds at a catalyst potential around

VWR < 0 V. Herein, this promotional effect could also be associated with the formation

of potassium carbonates and bicarbonates species on the catalyst film and allowed to

produce H2 and H2CO in a single reaction step. However, a new promotional effect

(promoted state I) appeared in the positive potential region (VWR = +0.5 V), which

increased the H2 and CO2 production rates. It should be noted that at this applied

potential, the overpotential (typically denoted as η in literature), was already negative

since the open circuit potential value (o

WRV ), was found in all experiments to be higher

than +0.5 V.

This new electro-promotional state, which was not reported before, could be

influenced by both the presence of Pt nanoparticles on the Pt-DLC catalyst film and

the formation of new promotional phases that may catalyze parallel processes. Hence,

we suggest that the formation of potassium oxides and superoxides species at VWR =

+0.5 V (promoted state I) may be also responsible for a parallel promotional effect on

the following reactions: water gas shift (with the H2O formed from the reaction

between H2 and O2) and CO oxidation reaction. It would lead to the observed increase

in the CO2 production rate instead of the CO one. This effect was probably hindered at

high temperature (320 ºC), where some CO was also produced at the promoted state I

(Figure 2.6c). In this case, it is probably that there was not enough O2 in the gas phase

to oxidize the higher amount of CO produced via the promotional effect on the partial

oxidation of methanol. Previous EPOC studies have already shown a promotional

effect of K+ ions on the water gas shift (WGS) [43] and CO oxidation [44] reactions.

In addition, the presence of this kind of K2O-derived promotional phases has been

reported in several studies with Pt/K-βAl2O3 systems [45, 46]. Previous works have

also demonstrated that the presence of smaller Pt particle sizes enhanced the formation

of K2O layers over platinum exposed to oxygen [47-49], which may explain the

Chapter 2

116

presence of this new promotional state on Pt-DLC, not observed on the less dispersed

pure Pt film. It was also shown in other works that the addition of alkali oxides, like

K2O, to supported Pt catalysts may significantly enhance the CO oxidation and WGS

reactions [50].

Hence, under the studied reaction conditions, the application of different

potentials led to the occurrence of different kinds of promoter phases, which made it

possible to control and optimize the rate of the different possible catalytic reactions

and, hence, the formation of different products by tuning a heterogeneous catalyst. On

the other hand, it is interesting to note that both kinds of promoter phases can be

completely removed from the catalyst surface (returning the K+

ions to the solid

electrolyte) by the final application of VWR = +2 V, leading to a completely reversible

EPOC effect. This demonstrates the stability of the Pt-DLC catalyst film after the

long-term potentiostatic experiments. Regarding the influence of raising the

temperature, the production rates of H2, CO2 and CO increased, as expected from

POM kinetics [14]. However, in the case of formaldehyde (Figure 2.6d), the reaction

rate reached an optimum value at 280 ºC, due to its further oxidation to CO2 at higher

temperatures [51].

In order to quantify the magnitude of the observed EPOC effect, the rate

enhancement ratio, ρ, can be calculated at each different applied potential, through the

following equation:

ρ

(2.1)

where r and r0 are the promoted (VWR < +2 V) and unpromoted (VWR = +2 V) catalytic

reaction rates, respectively. Hence, in the case of H2 production, a maximum value of

this parameter, max,2H , of 2.5 was obtained at 280 ºC under the application of the

optimum potential, VWR = -1 V.

Figure 2.7 shows the variation of the different products reaction rates: H2, CO2

and CO vs. time at different applied potentials and temperatures under SRM

conditions (CH3OH/H2O = 4 %/4.8 %, N2 balance).

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

117

Figure 2.7. H2 (a), CO2 (b) and CO (c) production rates vs. time at different applied potentials

(VWR) and temperatures. Steam reforming conditions: CH3OH/H2O = 4 %/4.8 %.

In this case, a decrease in the applied potential below +0.5 V led to an

important increase in the production rate of H2, CO2 and CO. This demonstrates that

the presence of K+ ions on the Pt-DLC catalyst film strongly promoted the methanol

reforming activity on the Pt particles. A decrease in the catalyst potential involved a

Chapter 2

118

decrease in the catalyst work function due to the migration of positively charged

potassium ions which weakened the chemical bond between Pt and the electron-donor

adsorbates (i.e., CH3OH) and strengthened that between this metal and the electron

acceptor ones (i.e., H2O, in this case). According to the reaction mechanism reported

in previous studies [6], based on a dissociative pathway of methanol, the activation of

water molecules through electrochemical promotion (at negative potentials) would

favour the decomposition of the reaction intermediates and hence the methanol

reforming catalytic process. Moreover, it is clear that the SRM kinetics strongly

increased with the reaction temperature in agreement with previous SRM works using

Pt-based catalysts [7]. The H2 production enhancement was increased at higher

reaction temperatures whereas the maximum promotional effect, max,2H = 3.4, was

achieved at 360 ºC and VWR = -2 V.

The observed EPOC effect was reproducible at all reaction temperatures, since

the initial un-promoted values of all products reaction rates were reached again after

the final application of +2 V (after 15 h working at each temperature). This fact

demonstrates again the stability of the Pt-DLC catalyst film. The optimal promotional

rates for POM and SRM corresponded to the electrochemical supply of 1.38 x 10-6

and

5.14 x 10-7

mol K+, respectively, calculated from the integration of the current vs. time

curves. In addition, it should be noted that in both: POM (Figure 2.6) and SRM

(Figure 2.7) under optimum electropromoted conditions, the corresponding values of

turnover frequencies of methanol consumption (TOF) (8.45 and 0.51 s-1

, respectively)

were shown to be of the same order as those obtained with conventional catalysts such

as as Cu/ZnO- [52], Cu/CeO2- [53], PdZn- [54] and VO-based [55] catalysts under

similar reaction conditions.

However, unlike the experiments under POM conditions, a single promotional

effect was found (at VWR ≤ +0.5 V) for the SRM process. In the latter case, and

analogously to the SRM results obtained in the first chapter with a pure dense Pt

catalyst film, a unique kind of promotional phase would be formed on the Pt-DLC

catalyst surface, very likely potassium carbonates or bicarbonates species. This

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

119

difference between both reaction atmospheres could be explained considering the

absence of O2 in the gas phase under SRM conditions, in contrast with POM where

both O2 and H2O were present on the gas phase. Thus, depending on both the reaction

atmosphere and the applied catalyst potential, several kinds of promoter phases, which

would affect the catalytic rates in different ways, could be formed via EPOC.

The different promotional effects, when the Pt-DLC catalyst was used for

POM and SRM processes, were characterized by cyclic voltammetry. Figure 2.8

shows the current variation vs. the applied potential of the catalyst film during a cyclic

voltammetry performed between +2 and -2 V (scan rate of 5 mV s-1

) at 360 ºC, with

two different feed compositions: CH3OH/O2 = 11 %/0.9 % (POM) and CH3OH/H2O =

4 %/4.8 % (SRM).

Figure 2.8. Current variation vs. the applied potential (VWR) of the Pt-DLC film during the

cyclic voltammetry between +2 and -2 V under POM (CH3OH/O2 = 11 %/0.9 %) and SRM

(CH3OH/O2 = 4 %/4.8 %) conditions. T = 360 ºC, scan rate = 5 mV s-1

.

-8

-6

-4

-2

0

2

4

6

8

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5

Cu

rre

nt

/ µ

A

VWR / V

SRM

POM

Chapter 2

120

As already reported in previous EPOC studies with K-βAl2O3-based

electrolytes [46, 56], the presence of different kinds of potassium promoter species

(formation and decomposition) can be followed by this technique. The cathodic peaks

shown in this figure could be linked with the formation of promoter-derived

compounds caused by the migration of K+ ions toward the Pt-DLC layer. Otherwise,

the anodic peaks could be associated to the decomposition of these formed species and

the spillover of the promoter ions back to the solid electrolyte [57, 58]. Hence, the

occurrence of different cathodic peaks (at least two) and the other two well defined

anodic peaks under POM conditions revealed the formation and decomposition of

different kinds of promoter phases under this reaction atmosphere, which could

explain the presence of different promoted states in Figure 2.6. On the other hand,

under SRM conditions just one cathodic peak with the corresponding anodic one can

be clearly observed. It could be attributed to the formation and decomposition of a

unique kind of promoter phase under this reaction atmosphere, in agreement with the

observed promotional results in Figure 2.7.

It should also be mentioned that the maximum electrical power, calculated

from the integration of the current vs VWR curve during the cathodic scan, used for

activating the catalyst was very low, i.e., 3.78 x 10-6

and 3.85 x 10-6

W cm-2

under

POM and SRM conditions, respectively. Hence, this would preserve the operation life

of the electrochemical cell and lead to a very low electrical energy consumption, i.e.,

the amount of electrochemically transferred promoter ions is very low.

2.3.3. Comparison between Pt-DLC and Pt

Finally, the catalytic performance obtained under SRM conditions with the

novel Pt-DLC catalyst electrode film was compared with that obtained in the first

chapter with the pure Pt film prepared according to the procedure described in section

1.2.1. Both catalysts were synthesized by cathodic arc deposition and resulted in

similar geometric properties (2.01 cm2, thickness of 120-150 nm). Figure 2.9 shows

the steady-state H2 production rate under SRM conditions (CH3OH/H2O = 4 %/4.8 %,

N2 balance) vs. the applied potential for both electrochemical catalysts at two different

reaction temperatures (320 and 360 ºC).

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

121

Figure 2.9. H2 production rate vs. applied catalyst potential (VWR) and maximum H2 rate

enhancement ratios (ρH2,max) at different reaction temperatures with the Pt and Pt-DLC films.

Steam reforming conditions: CH3OH/H2O = 4 %/4.8 %.

A much higher catalytic activity of the Pt-DLC catalyst film than that of the

pure dense Pt film was observed. This fact could be related to the higher dispersion of

the former catalyst film. In addition, a much stronger promotional effect of the Pt-

DLC catalyst film at each reaction temperature was also noted. In contrast, the

maximum rate enhancement ratio values ( max,2H ) for both electrochemical catalysts

were very similar at each temperature, due to the higher unpromoted catalytic rate of

the Pt-DLC film.

Figure 2.10 compares the current vs. time curves obtained with the two

catalyst-working electrodes at a given applied catalyst potential e.g., VWR = -0.5 V, for

the SRM reaction at 360 ºC.

0

1

2

3

4

5

6

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5

r H

/ m

ol s

-1x

10-8

VWR / V

Pt 320 ºC (0.46 mg Pt)

Pt 360 ºC (0.46 mg Pt)

Pt/DLC 320 ºC (0.03 mg Pt)

Pt/DLC 360 ºC (0.03 mg Pt)

2

2 ρH ,max = 3.37

ρH ,max = 3.54

2

ρH ,max = 1.86

2

ρH ,max = 1.53

2

Chapter 2

122

Figure 2.10. Current vs. time curves during the potentiostatic imposition of -0.5 V with the Pt

(a) and Pt-DLC (b) films. Steam reforming conditions: CH3OH/H2O = 4 %/4.8 %, 360 ºC.

Lower current values of the Pt-DLC catalyst film vs. Pt were found, which can

be explained by both the higher planar electrical resistance of the Pt-DLC catalyst

electrode (3.5 kΩ) vs. the pure Pt film (2 Ω) and the negligible ionic conductivity of

the DLC matrix. This led to lower rates of K+ ions supply to the catalyst film, since the

obtained current for Pt-DLC approximates more slowly to a negligible value. In fact

after 60 min of polarization, the Pt-DLC catalyst film was still increasing its potassium

coverage, in contrast with the pure Pt film which achieved a negligible current value

after the first minute of polarization, as commonly observed in other EPOC studies

[44, 56].

a)

b)

-0.10

-0.08

-0.06

-0.04

-0.02

0.00

Cu

rre

nt

/ m

APt

-0.01

-0.008

-0.006

-0.004

-0.002

0 5 10 15 20 25 30 35 40 45 50 55 60

Cu

rre

nt

/ m

A

Time / min

Pt-DLC

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

123

Thus, it seems clear that the higher electrical resistance and the absence of

ionic conductivity of the Pt-DLC catalyst film made the electrochemical supply of K+

ions to the Pt particles more difficult. However, the obtained EPOC catalytic results

showed higher electropromoted H2 production rates on the Pt-DLC catalyst film,

which presented much lower metal loading (0.014 mg Pt cm-2

) vs. the pure dense Pt

film (0.23 mg Pt cm-2

), thus demonstrating the interest of this novel kind of catalyst -

electrode materials. Finally it should also be noted that the electrical energy

consumption in EPOC systems is very low and is typically negligible vs. other

operational costs [1, 59]. This is because a low amount of electrical energy is required,

just to supply a small amount of promoter to the catalyst surface. For instance, the

overall supplied integral energy in the whole polarization at -0.5 V during 60 min

(Figure 2.10) was 3.11 x 10-3

J cm-2

for the Pt-DLC film and 6.32 x 10-3

J cm-2

for the

Pt film, which are a very low values.

2.4. CONCLUSIONS

The following conclusions could be drawn from this study:

- The technique of the cathodic arc deposition allowed to prepare a novel

catalyst film based on Pt nanoparticles (of around 3 nm) dispersed in a diamond-like

carbon (DLC) matrix with suitable properties for catalytic and electrocatalytic

purposes.

- The thermal treatment of the Pt-DLC catalyst film made it possible to

achieve a suitable surface electrical conductivity due to the transition of the sp3-

hybridized carbon form into a more graphitic structure (sp2-hybridized), as verified by

cross-section STEM and EELS characterization.

- The Pt-DLC catalyst film was successfully electrochemically promoted (by

potassium ions) for the hydrogen production from methanol via partial oxidation and

steam reforming reactions. All the observed electropromotional effects were

reproducible and reversible, the Pt-DLC catalyst film showing a good stability. Hence,

Chapter 2

124

it was demonstrated the possibility of electrochemically promoting catalyst particles

dispersed on an electronic non-ionic conductor support.

- Two different electropromotional effects were observed under methanol

partial oxidation conditions, which were attributed to the formation of different kinds

of promoter phases depending on the applied potential.

- The catalytic activity of the Pt-DLC film was higher than that observed for

the pure Pt film prepared by the same technique due to the lower Pt particle size in the

former. This kind of catalyst films based on metal nanoparticles dispersed in a DLC

material of low metal loading open new possibilities for the further development of

the EPOC effect in view of its possible practical application.

2.5. REFERENCES

[1] C.G. Vayenas, S. Bebelis, C. Pliangos, S. Brosda, D. Tsiplakides, Electrochemical

Activation of Catalysis: Promotion, Electrochemical Promotion and Metal-Support

Interactions, Kluwer Academic Publishers/Plenum Press, New York, 2001.

[2] S. Sundarrajan, S.I. Allakhverdiev, S. Ramakrishna, Progress and perspectives in micro

direct methanol fuel cell, International Journal of Hydrogen Energy, 37 (2012) 8765-8786.

[3] D.R. Palo, R.A. Dagle, J.D. Holladay, Methanol steam reforming for hydrogen production,

Chemical Reviews, 107 (2007) 3992-4021.

[4] R.M. Navarro, M.A. Peña, J.L.G. Fierro, Hydrogen production reactions from carbon

feedstocks: Fossil fuels and biomass, Chemical Reviews, 107 (2007) 3952-3991.

[5] S. Sá, H. Silva, L. Brandão, J.M. Sousa, A. Mendes, Catalysts for methanol steam

reforming-A review, Applied Catalysis B: Environmental, 99 (2010) 43-57.

[6] G. Jacobs, B.H. Davis, In situ DRIFTS investigation of the steam reforming of methanol

over Pt/ceria, Applied Catalysis A: General, 285 (2005) 43-49.

[7] P. Tolmacsov, A. Gazsi, F. Solymosi, Decomposition and reforming of methanol on Pt

metals supported by carbon Norit, Applied Catalysis A: General, 362 (2009) 58-61.

[8] S.H. Ahn, O.J. Kwon, I. Choi, J.J. Kim, Synergetic effect of combined use of Cu-ZnO-

Al2O3 and Pt-Al2O3 for the steam reforming of methanol, Catalysis Communications, 10

(2009) 2018-2022.

[9] G. Kolb, S. Keller, S. Pecov, H. Pennemann, R. Zapf, Development of micro-structured

catalytic wall reactors for hydrogen production by methanol steam reforming over novel

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

125

Pt/In2O 3/Al2O3 catalysts, Chemical Engineering Transactions, 2011, pp. 133-138,

DOI:10.3303/cet1124023.

[10] A. Peremans, F. Maseri, J. Darville, J.M. Gilles, Interaction of methanol with a

polycrystalline platinum surface studied by infrared reflection absorption spectroscopy,

Surface Science, 227 (1990) 73-78.

[11] M.P. Zum Mallen, L.D. Schmidt, Oxidation of methanol over polycrystalline Rh and Pt:

Rates, OH desorption, and model, Journal of Catalysis, 161 (1996) 230-246.

[12] B.E. Traxel, K.L. Hohn, Partial oxidation of methanol at millisecond contact times,

Applied Catalysis A: General, 244 (2003) 129-140.

[13] R. Ubago-Pérez, F. Carrasco-Marín, C. Moreno-Castilla, Methanol partial oxidation on

carbon-supported Pt and Pd catalysts, Catalysis Today, 123 (2007) 158-163.

[14] A. Wan, C.t. Yeh, Ignition of methanol partial oxidation over supported platinum catalyst,

Catalysis Today, 129 (2007) 293-296.

[15] C. Xia, M. Hugentobler, L. Yongdan, C. Comninellis, W. Harbich, Quantifying

electrochemical promotion of induced bipolar Pt particles supported on YSZ, Electrochemistry

Communications, 12 (2010) 1551-1554.

[16] M. Marwood, C.G. Vayenas, Electrochemical promotion of a dispersed platinum catalyst,

Journal of Catalysis, 178 (1998) 429-440.

[17] S.G. Neophytides, D. Tsiplakides, P. Stonehart, M.M. Jaksic, C.G. Vayenas,

Electrochemical enhancement of a catalytic reaction in aqueous solution, Nature, 370 (1994)

45-47.

[18] L. Ploense, M. Salazar, B. Gurau, E.S. Smotkin, Proton spillover promoted isomerization

of n-butylenes on Pd-black cathodes/nafion 117 [8], Journal of the American Chemical

Society, 119 (1997) 11550-11551.

[19] V. Jiménez, C. Jiménez-Borja, P. Sánchez, A. Romero, E.I. Papaioannou, D. Theleritis, S.

Souentie, S. Brosda, J.L. Valverde, Electrochemical promotion of the co2 hydrogenation

reaction on composite ni or ru impregnated carbon nanofiber catalyst-electrodes deposited on

YSZ, Applied Catalysis B: Environmental, 107 (2011) 210-220.

[20] D. Poulidi, M.E. Rivas, B. Zydorczak, Z. Wu, K. Li, I.S. Metcalfe, Electrochemical

promotion of a Pt catalyst supported on La 0.6Sr 0.4Co 0.2Fe 0.8O 3 - δ hollow fibre

membranes, Solid State Ionics, 225 (2012) 382-385.

Chapter 2

126

[21] A. Kambolis, L. Lizarraga, M.N. Tsampas, L. Burel, M. Rieu, J.P. Viricelle, P. Vernoux,

Electrochemical promotion of catalysis with highly dispersed Pt nanoparticles,

Electrochemistry Communications, 19 (2012) 5-8.

[22] A. Grill, Diamond-like carbon: State of the art, Diamond and Related Materials, 8 (1999)

428-434.

[23] J. Robertson, AMORPHOUS CARBON, Advances in Physics, 35 (1986) 317-374.

[24] N. Menegazzo, C. Jin, R.J. Narayan, B. Mizaikoff, Compositional and electrochemical

characterization of noble metal-diamondlike carbon nanocomposite thin films, Langmuir, 23

(2007) 6812-6818.

[25] J.M. Ting, H. Lee, DLC composite thin films by sputter deposition, Diamond and Related

Materials, 11 (2002) 1119-1123.

[26] G.C. Fiaccabrino, X.M. Tang, N. Skinner, N.F. De Rooij, M. Koudelka-Hep,

Interdigitated microelectrode arrays based on sputtered carbon thin-films, Sensors and

Actuators, B: Chemical, 35 (1996) 247-254.

[27] G.Y. Chen, J.S. Chen, Z. Sun, Y.J. Li, S.P. Lau, B.K. Tay, J.W. Chai, Field emission

properties and surface structure of nickel containing amorphous carbon, Applied Surface

Science, 180 (2001) 185-190.

[28] H. Lee, J.M. Ting, Platinum-containing diamond-like carbon thin films, Journal of the

American Ceramic Society, 87 (2004) 2183-2186.

[29] Y.V. Pleskov, Y.E. Evstefeeva, A.M. Baranov, Threshold effect of admixtures of platinum

on the electrochemical activity of amorphous diamond-like carbon thin films, Diamond and

Related Materials, 11 (2002) 1518-1522.

[30] K.I. Schiffmann, M. Fryda, G. Goerigk, R. Lauer, P. Hinze, A. Bulack, Sizes and

distances of metal clusters in Au-, Pt-, W- and Fe-containing diamond-like carbon hard

coatings: a comparative study by small angle X-ray scattering, wide angle X-ray diffraction,

transmission electron microscopy and scanning tunneling microscopy, Thin Solid Films, 347

(1999) 60-71.

[31] A. Anders, Cathodic Arcs: From Fractal Spots to Energetic Condensation. Springer Series

on Atomic, Optical, and Plasma Physics, Springer-Verlag, New York, 2008.

[32] A. Anders, I.G. Brown, R.A. MacGill, M.R. Dickinson, 'Triggerless' triggering of vacuum

arcs, Journal of Physics D: Applied Physics, 31 (1998) 584-587.

[33] S.R.P. Silva, X. Shi, B.K. Tay, H.S. Tan, H.J. Scheibe, M. Chhowalla, W.I. Milne, The

structure of tetrahedral amorphous carbon thin films, Thin Solid Films, 290-291 (1996) 317-

322.

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

127

[34] M. Chhowalla, J. Robertson, C.W. Chen, S.R.P. Silva, C.A. Davis, G.A.J. Amaratunga,

W.I. Milne, Influence of ion energy and substrate temperature on the optical and electronic

properties of tetrahedral amorphous carbon (ta-C) films, Journal of Applied Physics, 81

(1997) 139-145.

[35] J.L. Endrino, D. Horwat, R. Gago, J. Andersson, Y.S. Liu, J. Guo, A. Anders, Electronic

structure and conductivity of nanocomposite metal (Au, Ag, Cu, Mo)-containing amorphous

carbon films, Solid State Sciences, 11 (2009) 1742-1746.

[36] A. de Lucas-Consuegra, F. Dorado, J.L. Valverde, R. Karoum, P. Vernoux,

Electrochemical activation of Pt catalyst by potassium for low temperature CO deep oxidation,

Catalysis Communications, 9 (2008) 17-20.

[37] P. Vernoux, F. Gaillard, C. Lopez, E. Siebert, Coupling catalysis to electrochemistry: A

solution to selective reduction of nitrogen oxides in lean-burn engine exhausts?, Journal of

Catalysis, 217 (2003) 203-208.

[38] F.J. Williams, M.S. Tikhov, A. Palermo, N. Macleod, R.M. Lambert, Electrochemical

promotion of rhodium-catalyzed NO reduction by CO and by propene in the presence of

oxygen, Journal of Physical Chemistry B, 105 (2002) 2800-2808.

[39] F. Dorado, A. de Lucas-Consuegra, P. Vernoux, J.L. Valverde, Electrochemical

promotion of platinum impregnated catalyst for the selective catalytic reduction of NO by

propene in presence of oxygen, Applied Catalysis B: Environmental, 73 (2007) 42-50.

[40] C.G. Vayenas, S. Bebelis, S. Ladas, Dependence of catalytic rates on catalyst work

function, Nature, 343 (1990) 625-627.

[41] C.G. Vayenas, S. Neophytides, Non-faradaic electrochemical modification of catalytic

activity III. The case of methanol oxidation on Pt, Journal of Catalysis, 127 (1991) 645-664.

[42] C.A. Cavalca, G. Larsen, C.G. Vayenas, G.L. Haller, Electrochemical modification of

CH3OH oxidation selectivity and activity on a Pt single-pellet catalytic reactor, Journal of

Physical Chemistry, 97 (1993) 6115-6119.

[43] A. De Lucas-Consuegra, A. Caravaca, J. González-Cobos, J.L. Valverde, F. Dorado,

Electrochemical activation of a non noble metal catalyst for the water-gas shift reaction,

Catalysis Communications, 15 (2011) 6-9.

[44] A. de Lucas-Consuegra, A. Princivalle, A. Caravaca, F. Dorado, C. Guizard, J.L.

Valverde, P. Vernoux, Preferential CO oxidation in hydrogen-rich stream over an

electrochemically promoted Pt catalyst, Applied Catalysis B: Environmental, 94 (2010) 281-

287.

Chapter 2

128

[45] A. Palermo, A. Husain, M.S. Tikhov, R.M. Lambert, Ag-catalysed epoxidation of propene

and ethene: An investigation using electrochemical promotion of the effects of alkali, NOx, and

chlorine, Journal of Catalysis, 207 (2002) 331-340.

[46] A. de Lucas-Consuegra, F. Dorado, J.L. Valverde, R. Karoum, P. Vernoux, Low-

temperature propene combustion over Pt/K-βAl2O3 electrochemical catalyst:

Characterization, catalytic activity measurements, and investigation of the NEMCA effect,

Journal of Catalysis, 251 (2007) 474-484.

[47] E.L. Garfunkel, G.A. Somorjai, Potassium and potassium oxide monolayers on the

platinum (111) and stepped (755) crystal surfaces: A LEED, AES, and TDS study, Surface

Science, 115 (1982) 441-454.

[48] J.H. Vleeming, B.F.M. Kuster, G.B. Marin, Effect of platinum particle size and catalyst

support on the platinum catalyzed selective oxidation of carbohydrates, Catalysis Letters, 46

(1997) 187-194.

[49] T. Frelink, W. Visscher, J.A.R. van Veen, On the role of Ru and Sn as promotors of

methanol electro-oxidation over Pt, Surface Science, 335 (1995) 353-360.

[50] C.H. Lee, Y.W. Chen, Effect of Basic Additives on Pt/Al2O3 for CO and Propylene

Oxidation under Oxygen-Deficient Conditions, Industrial and Engineering Chemistry

Research, 36 (1997) 1498-1506.

[51] R.W. McCabe, D.F. McCready, Kinetics and reaction pathways of methanol oxidation on

platinum, Journal of Physical Chemistry, 90 (1986) 1428-1435.

[52] L. Alejo, R. Lago, M.A. Peña, J.L.G. Fierro, Partial oxidation of methanol to produce

hydrogen over Cu-Zn-based catalysts, Applied Catalysis A: General, 162 (1997) 281-297.

[53] S.C. Yang, W.N. Su, S.D. Lin, J. Rick, B.J. Hwang, Preparation of highly dispersed

catalytic Cu from rod-like CuO-CeO 2 mixed metal oxides: Suitable for applications in high

performance methanol steam reforming, Catalysis Science and Technology, 2 (2012) 807-812.

[54] B. Halevi, E.J. Peterson, A. Roy, A. Delariva, E. Jeroro, F. Gao, Y. Wang, J.M. Vohs, B.

Kiefer, E. Kunkes, M. Hävecker, M. Behrens, R. Schlögl, A.K. Datye, Catalytic reactivity of

face centered cubic PdZn α for the steam reforming of methanol, Journal of Catalysis, 291

(2012) 44-54.

[55] G. Deo, I.E. Wachs, Reactivity of Supported Vanadium Oxide Catalysts: The Partial

Oxidation of Methanol, Journal of Catalysis, 146 (1994) 323-334.

[56] A. de Lucas-Consuegra, Á. Caravaca, P. Sánchez, F. Dorado, J.L. Valverde, A new

improvement of catalysis by solid-state electrochemistry: An electrochemically assisted NOx

storage/reduction catalyst, Journal of Catalysis, 259 (2008) 54-65.

Electrochemical promotion of Pt nanoparticles dispersed in a carbon matrix for methanol

conversion

129

[57] P. Vernoux, F. Gaillard, C. Lopez, E. Siebert, In-situ electrochemical control of the

catalytic activity of platinum for the propene oxidation, Solid State Ionics, 175 (2004) 609-613.

[58] N. Kotsionopoulos, S. Bebelis, Electrochemical characterization of the Pt/β″-Al 2O3

system under conditions of in situ electrochemical modification of catalytic activity for

propane combustion, Journal of Applied Electrochemistry, 40 (2010) 1883-1891.

[59] P. Vernoux, L. Lizarraga, M.N. Tsampas, F.M. Sapountzi, A. De Lucas-Consuegra, J.L.

Valverde, S. Souentie, C.G. Vayenas, D. Tsiplakides, S. Balomenou, E.A. Baranova, Ionically

conducting ceramics as active catalyst supports, Chemical Reviews, 113 (2013) 8192-8260.

Chapter 2

130

131

Chapter 3

ELECTROCHEMICAL ACTIVATION OF Au

NANOPARTICLES DISPERSED IN YSZ FOR METHANOL

PARTIAL OXIDATION:

ELECTROCHEMICAL PROMOTION OF A NON-CONDUCTIVE

CATALYST FILM

An electrochemical catalyst based on Au nanoparticles dispersed in a Yttria-

Stabilized Zirconia (YSZ) matrix has been deposited by reactive co-sputtering of

Zirconium-Yttrium and Au targets on a K-βAl2O3 solid electrolyte. The Au/YSZ

catalyst film thus obtained has shown to be active in the partial oxidation of methanol

with a high selectivity toward methyl formate production. This configuration allowed

to decrease the amount of metal used in the solid electrolyte cell, and activate a highly

dispersed Au catalyst via electrochemical promotion (EPOC), by in-situ controlling

and optimizing the supplied amount of K+ ions. A number of experiments have

confirmed that the observed electropromotional effect did not depend on the rate of K+

supply or on the operation mode (galvanostatic or potentiostatic) and only depended

on the promoter coverage. The stability of the Au nanoparticles under the explored

conditions has also been verified.

-1

0

1

2

3

0

1

2

3

4

5

0 25 50 75 100 125 150 175

VW

R/

V

r H/

mo

l s-1

x 1

0-8

Time / min

0 2 41 3K+ supply / mol K+ x 10-7

VWR = 2 VVWR = 2 V I = 0 AI = -10 µA

ρH

2

0

3

6

9

12

2

ΔV < 0Au-YSZ

K+ K+ K+

e-

e-

K-βAl2O3

Au

Δr

CH3OH,

O2

H2, CO2,

HCOOCH3, H2O

5.9 % CH3OH

0.43 % O2

280 ºC

Optimum

Ag current

colector

Chapter 3

132

3.1. INTRODUCTION

Alkali promoters play a key role in heterogeneous catalysis and are necessary

for the development of commercial catalysts for industrial applications [1, 2]. They are

commonly employed to modify the activity and selectivity of metal catalysts. In this

sense, the electrochemical promotion of catalysis (EPOC) is a straightforward

technique to study the effect of a certain electronic promoter on a reaction rate [3]. In

the previous chapters, Pt-based catalysts were observed to be promoted by alkali (K+)

ions in H2 production reactions from methanol via EPOC, i.e., through the

electrochemical pumping of these promoter ions from a solid electrolyte support (K-

βAl2O3). Hence, while the classical alkali promotion is carried out by adding a specific

amount of a promoter during the preparation step of the catalyst [1], in the case of the

electrochemical promotion, the electrically induced back-spillover of the promoter

species enables the in-situ control and enhancement of the performance of the catalyst

film during the reaction step [4].

In the present chapter, the alkali electrochemical promotion has been applied

on a gold-based catalyst for the methanol partial oxidation, which represented a great

challenge since Au is not among the most typical catalysts. In fact, counter and

reference electrodes made of gold have been selected for the EPOC experiments

herein performed given their proved inertness under the studied reaction conditions,

which is highly favoured by the deposition method employed in this case (calcination

of an Au organometallic paste at high temperature). Only a few works can be found in

literature related to electrochemical promotion with Au catalysts: in solid oxide fuel

cell configuration for the oxidation of CO [5] and CH4 [6] with O2-

promoter ions, in

the liquid-phase SO2 oxidation with molten V2O5-K2S2O7 catalyst [7] and in the

preferential oxidation of CO in a low temperature Proton Exchange Membrane (PEM)

fuel cell reactor promoted by H+ ions and O2 crossover through the membrane [8].

Certainly, until a few decades ago, gold has been generally considered

catalytically inactive due to its difficulty to both adsorb and dissociate molecular

oxygen [9] and to remain in a high dispersion state because of its low melting point as

compared to other noble metals such as Pt or Pd [10, 11]. However, several studies

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

133

have shown good catalytic properties of unsupported nanoporous gold prepared by

leaching Ag from an Au-Ag alloy [12, 13], gold nanoparticles supported on highly

porous carbon materials [14] and, especially, on certain metal oxides (e.g. TiO2, CeO2,

Fe2O3, ZnO, or ZrO2) [10, 11, 15]. Numerous applications of gold-based catalysts in

the field of environmental catalysis have been recently reviewed [10]. In particular,

gold catalysts have demonstrated high activity in CO oxidation [11, 12, 16-19] even at

temperatures below 0 ºC [11, 12], which is of special interest in view of selective CO

sensing [16], CO removal from air at room temperature [11, 12], and preferential

oxidation of CO in H2 rich streams [11, 17]. Gold nanoparticles supported on oxide

supports have also been tested in other catalytic reactions such as CO2 hydrogenation

[20], NOx reduction [21], steam reforming of methanol [22] and ethanol [23], water

gas shift [22], and oxidation of methane [24], alkenes [25], alcohols both in liquid [26]

and in gas phase [19, 23, 27-33] and many other organic compounds [34].

In the case of the methanol partial oxidation processes [13, 19, 28-33], it

should be noted the complexity of the reaction pathways that may take place. A large

number of side reactions and intermediate species are generally involved, making

possible to obtain a wide variety of compounds. Along with the main oxidation

products, i.e., H2, CO2, CO and H2O, not only formaldehyde (H2CO) can be co-

produced [32, 33], as in the previous chapters, but also CH4 [29-32], dimethyl ether

(CH3OCH3) [33], dimethoxymethane (CH3O)2CH2 [33] or methyl formate HCOOCH3

[30-33]. In this sense, Au-based catalysts may acquire prominence against Pt or Pd,

due to the capability of Au to reach a very high selectivity in the methanol partial

oxidation toward some of these products of well-known interest in chemical

manufacturing, such as formaldehyde [19] and methyl formate [13]. The influence of

the gold particle size and the oxide support on the catalyst performance for the

methanol oxidation has been widely studied [19, 28-33] as well as the promoting

effect of a second active metallic component [28, 31, 32].

In this work, we report for the first time the electrochemical activation of Au

nanoparticles dispersed in a YSZ (Yttria-Stabilized Zirconia) matrix with

Chapter 3

134

electropositive alkali ions. This novel catalyst film (Au-YSZ) has been deposited on a

K-βAl2O3 solid electrolyte by sputter deposition and electrochemically promoted by

controlling and optimizing the K+ promoter supply. The proposed configuration

allowed both to electrochemically activate highly dispersed Au nanoparticles and to

decrease the amount of metal used in the solid electrolyte cell. The Au-YSZ/K-βAl2O3

electrochemical catalyst has been characterized and tested under methanol partial

oxidation conditions (POM), also checking its reproducibility and stability.

3.2. EXPERIMENTAL

3.2.1. Preparation of the electrochemical catalyst

The electrochemical catalyst consisted of a thin continuous gold/yttria-

stabilized zirconia (Au-YSZ) film (geometric area of 3.14 cm2) composed of fine gold

nanoparticles dispersed in a YSZ matrix, which was deposited by magnetron

sputtering on a 20-mm-diameter, 1-mm-thick K-βAl2O3 (Ionotec) disc. Firstly, inert

Au counter (CE) and reference (RR) electrodes were deposited on one of the sides of

the solid electrolyte as in the previous chapters, i.e., by applying thin coatings of gold

paste (Gwent Electronic Materials C1991025D2) followed by calcination at 800 ºC for

2 h (heating ramp of 5 ºC/min). Then, the Au-YSZ nanocomposite catalyst film, which

also behaved as a working electrode (WE), was deposited by reactive co-sputtering of

Zr-Y and Au targets in an argon-oxygen gas mixture ensuring the complete oxidation

of the Zr and Y atoms in a nanocrystalline, nearly amorphous, YSZ matrix. The

deposition time was set such that the resulting Au-YSZ film thickness was around 170

nm, as measured by STEM imaging (not shown). The oxygen gas flow rate and the

current applied to the targets were fixed such that the volume fraction of gold particles

into the composite layer was 11 vol. %, which corresponded to a gold content of 7.5

at. % or a Au nanoparticles loading of 40 g Au cm-2

(2.03 x 10-7

mol Au cm-2

). More

details on the experimental procedure and principle of segregation of metallic gold

into the YSZ matrix during the formation of the film can be found in previous studies

[35]. For characterization purposes, the same layer was prepared on glass substrates.

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

135

This catalyst preparation method was carried out in collaboration with Dr.

David Horwat from the Institute Jean Lamour at University of Lorraine (France).

The obtained Au-YSZ/K-βAl2O3/Au electrochemical catalyst was placed into a

single chamber solid electrolyte cell reactor and the three electrodes (working, counter

and reference) were connected to the potentiostat-galvanostat (PGSTAT320-N,

Metrohom Autolab). However, since the Au-YSZ nanocomposite was electrically

non-conductive from its preparation, a silver mesh (Fuel Cell Materials, AG-M40-

100) was used in this study which acted as current collector in order to carry out the

EPOC experiments. Moreover, the results reported here were obtained in both

operation modes: galvanostatic and potentiostatic, i.e. by setting the electrical current

(from 0 to -20 µA) or potential (from +2 to -2 V), respectively, according to the

procedure generally used in conventional three-electrodes electrochemical cells [3].

3.2.2. Characterization measurements

X-Ray Diffraction (XRD) was performed on the as-deposited Au-YSZ

nanocomposite layer with a AXS Brücker D8 advance diffractometer using a Cu

anode (Cu Kα radiation, λ = 1.54 Å). The thermal stability of this layer was verified

by the same technique after a long catalytic test of 12 h at 320 ºC under the studied

POM reaction conditions. Transmission Electron Microscopy (TEM) investigation and

High-Resolution TEM (HRTEM) imaging of the tested sample after all the catalytic

experiments were carried out with a JEOL ARM 200F TEM/STEM– Cold FEG (point

resolution: 0.19 nm) fitted with a GIF Quatum ER. Cyclic voltammetry (CV)

measurements were also performed prior to the EPOC experiments, in order to

evaluate the polarization of the solid electrolyte cell under POM reaction conditions.

Before the cyclic voltammetry experiment, the sample was kept at +2 V for 30 min in

order to define a reference state (Au-YSZ film clean of K+ promoter ions). Then, the

potential was cyclically varied between +2 and -2 V at a scan rate of 80 mV s-1

.

Chapter 3

136

3.2.3. Catalytic activity measurements

The activity measurements were carried out in an experimental setup used

described in section 1.2.3. Methanol partial oxidation (POM) experiments were

performed at atmospheric pressure with an overall gas flow rate of 6 NL h-1

,

temperatures up to 300 ºC and a feed composition of CH3OH/O2 = 5.9 %/0.43 % (Ar

balance). The reaction products detected in this study were: H2, CO2, H2O, HCOOCH3

(methyl formate) and traces of dimethoxymethane and other unidentified compound

which could be dimethyl ether. Without considering the two latter compounds, the

error in the carbon atom balance did not exceed 1 %, which indicated no consistent

loss of material and no significant coking of the catalyst electrode. Blank experiments

were also performed under the same reaction conditions without any catalyst and only

with the Ag current collector in order to determine its negligible catalytic contribution.

3.3. RESULTS AND DISCUSSION

3.3.1. Characterization of the catalyst film and blank experiments

Figure 3.1 shows the X-ray diffractograms of the fresh and the aged (after 24

hours at 320°C under reaction conditions) nanocomposite layer deposited on glass.

Figure 3.1. XRD diffractograms of the Au-YSZ nanocomposite layer as deposited and after

catalytic test under methanol partial oxidation conditions (CH3OH/O2 = 5.9 %/0.43 %, 320 ºC).

30 40 50 60 70 80

22

0-A

u

20

0-A

u

After 320°C

Inte

nsity (

arb

. u

nits)

2 (Degree)

As deposited

11

1-A

u

Inte

nsi

ty /

a.u

.

2θ / Degree

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

137

Both diffractogramms exhibit the characteristic diffraction peaks (111), (200)

and (220) at 2θ = 38.2º, 44.4º and 64.5º, respectively, corresponding to nanocrystalline

gold (JCPDS, 01-1172). Nevertheless, the diffracted intensity was slightly increased

after exposure to reaction conditions. The average size of the nanoparticles was

estimated from the width of the (111) diffraction peak using the Scherrer formula [36].

An average particle size of 3.3 nm was obtained in the fresh state, and slightly rose up

to 4.3 nm after exposure to reaction conditions, which indicated a minor diffusion of

the gold atoms and a good thermal stability of the nanocomposite layer. No diffraction

peaks corresponding to gold oxides (AuO or Au2O3) were detected. No diffraction

lines of Yttria-Stabilized Zirconia (YSZ) were identified in the X-ray diffractogramms

due to the ultrafine grained, nearly amorphous, nature of the YSZ matrix, which was

confirmed by HRTEM (not shown here). The XRD signal of YSZ could only be

detected after long time-annealing at 500°C, which is far above the temperatures

experienced by the samples in this study. Analyses of the films deposited on the K-

Al2O3 led to similar conclusions, therefore indicating a limited influence of the

substrate on the thermal stability of the nanocomposite layer.

Further details of the aged gold particles after exposure to reactions conditions

were obtained using TEM and HRTEM. Figure 3.2 shows bright field images of

selected zones in the Au-YSZ catalyst film. Figures 3.2a and 3.2b show extended and

local views of the typical distribution of gold nanoparticles in the core of the coatings

and at the surface exposed to the reaction mixture, respectively. The YSZ matrix

appears in clear while dark zones correspond to individual Au nanoparticles dispersed

in it. The particles were round shaped with a typical size distribution ranging between

2 and 10 nm. Nevertheless, it must be pointed out that thermally assisted diffusion

occurred upon exposure to the electron beam during experiments, which triggered a

slight increase of the particle diameter. This way, the size of the particles was slightly

over-estimated, being the measurements in accordance with the XRD data. This fact

would demonstrate the good stability of the gold nanoparticles after exposure to

reaction conditions.

Chapter 3

138

Figure 3.2. Bright field micrographs of Au-YSZ nanocomposite film after catalytic tests under

methanol partial oxidation conditions (CH3OH/O2 = 5.9 %/0.43 %, 320 ºC): Typical

microstructure in the core through the film thickness and at the film surface (a, b) and at the

substrate/film interface (c). HRTEM micrograph of a selected region in the same sample shows

the atom scale arrangement of the nanoparticles (d).

Larger particles, with diameters ranging between 15 and 50 nm, but mostly

located at the substrate/film interface were also found (Figure 3.2c). The presence of

larger size gold crystallites after catalytic tests was also observed in previous studies

of methanol oxidation over supported gold catalysts due to aggregation of smaller Au

particles [19, 30, 33]. Atom scale resolution HRTEM and Fourier filter transform

c)

a) b)

d)

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

139

(FFT) of the images (not shown) confirmed the presence of particles consisting of face

centered cubic (FCC) metallic assembly of gold atoms. For instance, the (111) atomic

arrangement of FCC gold was observed for particles in Figure 3.2d.

Figure 3.3 shows the steady-state variation at different reaction temperatures

of the production rates (hydrogen, carbon dioxide and methyl formate) obtained with

the Au-YSZ catalyst film as well as the blank results in the partial oxidation of

methanol (feed composition of CH3OH/O2 = 5.9 %/0.43 %, Ar balance). In the case of

the electrocatalytic system (current collector + Au-YSZ catalyst), the catalytic activity

was measured at each temperature under open circuit conditions, i.e. without applying

any electric current or potential, after the imposition of a catalyst potential (VWR) of +2

V for 30 minutes. In agreement with the previous chapters and other studies with K-

βAl2O3-based electrolytes [37, 38], the purpose of this preliminary positive

polarization was to remove from the Au-YSZ catalyst film the K+ ions that could have

thermally migrated from the solid electrolyte. This thereby defined a reference state

(unpromoted state).

Both the reactor blank and the silver mesh showed a negligible activity under

the studied reaction conditions as compared to that of the electrochemical catalyst. On

the other hand, it was confirmed that the Au-YSZ catalyst film presented certain

catalytic activity in the methanol partial oxidation under unpromoted conditions.

While large Au particles and bulk gold are generally considered to be inert, catalysts

based on gold nanoparticles supported on different metal oxides have been reported by

several authors to be catalytically active in this reaction [19, 29-33]. In fact, the

smaller the Au particle size, the higher the methanol partial oxidation rate as well as

the H2 selectivity vs. CO or CH4 [30-32]. In particular, gold nanoparticles dispersed on

YSZ have already been tested in the CO oxidation reaction [18] but, to the best of our

knowledge, this is the first time that the catalytic activity of a Au-YSZ composite film

is studied in the methanol partial oxidation.

Chapter 3

140

Figure 3.3. Steady-state H2 (a), CO2 (b) and methyl formate (HCOOCH3) (c) production rates

at different reaction temperatures during the blank experiments and the catalytic tests using the

Au-YSZ composite film under open circuit conditions (after application of VWR = +2 V for 30

min, i.e., un-promoted state). CH3OH/O2 = 5.9 %/0.43 %.

The H2 and CO2 production rates observed in this figure were expected from

the main POM reaction (reaction 3.1).

0

0.1

0.2

0.3

0.4

0.5

0.6

r H/

mo

l s-1

x 1

0-8 Reactor blank

Current collector

Current collector + Au-YSZ catalyst

a)

2

0

0.5

1

1.5

2

r CO

/ m

ol

s-1x

10

-8

Reactor blank

Current collector

Current collector + Au-YSZ catalyst

b)

2

0

0.5

1

1.5

2

2.5

3

240 260 280 300 320

r HC

OO

CH

/ m

ol s

-1x

10

-8

Temperature / ºC

Reactor blank

Current collector

Current collector + Au-YSZ catalyst

c)

3

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

141

CH3OH + 0.5O2 → 2H2 + CO2 (3.1)

However, the most abundant product was methyl formate (HCOOCH3), with a

selectivity of 78 % (with respect to carbon), calculated as C C 3,out

C 2,out ,out , where C C 3,out and C 2,out are the molar flow rates

of methyl formate and carbon dioxide, respectively, at the outlet of the reactor. There

are many possible intermediate and side reactions involved in the methanol oxidation

as very recently summarized [33]. Methanol dissociation resulting in methoxy species

and hydroxyl group is generally considered the first step of the sequence. Then,

formaldehyde, formic acid and carbon dioxide may be produced, in this order, from

successive oxidation reactions, but also dimethyl ether, dimethoxymethane, methyl

formate and carbon monoxide may be obtained from the dehydrogenation of the

former compounds. Hence, the methyl formate could be formed from the following

consecutive steps [39]:

CH3OH + 0.5O2 → 2CO + H2O (3.2)

H2CO + 0.5O2 → HCOOH (3.3)

HCOOH + CH3 → C C 3 + H2O (3.4)

Under lack of surface oxygen, HCOOCH3 production on supported gold

catalysts has been postulated as a direct dehydrogenation of methanol (reaction 3.5)

[30, 31], and formaldehyde dimerization (reaction 3.6) has also be considered [40].

2CH3 ↔ C C 3 + 2H2 (3.5)

2H2C → C C 3 (3.6)

In any case, the weak adsorption of the methanol partial oxidation products on

Au, as compared to other noble metals, makes the gold-based catalysts suitable for the

low temperature selective oxidation of methanol to oxygenated compounds other than

CO2, such as HCOOCH3 [13]. Moreover, methyl formate is of interest in the chemical

industry as a solvent and a precursor in the synthesis of formic acid, formamide and

dimethylformamide, among other organic compounds [41].

Chapter 3

142

Prior to the catalytic activity measurements under electrochemical promotion

conditions, the Au-YSZ/K-βAl2O3/Au electrochemical catalyst was in-situ

characterized by cyclic voltammetry under the same POM reaction conditions at 280

ºC, by recording the current (I) variation with the applied potential (VWR) between +2

and -2 V (scan rate of 80 mV s-1

) (Figure 3.4). Previously, the electrochemical catalyst

was kept at +2 V for 30 minutes in order to define a Au-YSZ film free of potassium

ions (reference state).

Figure 3.4. Current variation (I) vs. the applied potential (VWR) of the electrochemical catalyst

during the cyclic voltammetry between +2 and -2 V (scan rate = 80 mV s-1

) under methanol

partial oxidation conditions (CH3OH/O2 = 5.9 %/0.43 %, 280 ºC). The first three cycles are

depicted, in order from lighter to darker.

It can be observed that the cyclic voltammetry was reversible and almost

completely reproducible from the second cycle at the explored potential range. It has

been widely established that the cathodic peaks observed in the CV carried out with

alkali ionic conductors can be associated with the formation of surface compounds

-2

-1.6

-1.2

-0.8

-0.4

0

0.4

0.8

-2.5 -1.5 -0.5 0.5 1.5 2.5

I / A

x 1

0-3

VWR / V

1st cycle

2nd cycle3rd cycle

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

143

derived from the interaction among the different reactants adsorbed on the catalyst-

working electrode and the alkali ions coming from the solid electrolyte [37, 38, 42,

43]. No electrocatalytic processes have been considered on the gold counter electrode

since it was constituted by large particles of negligible adsorption activity [37, 38, 42,

43]. In the present case, due to the migration of K+ ions to the Au-YSZ film and the

presence of O2, CO2 and H2O in the reaction atmosphere, K-derived surface

compounds such as potassium oxides/superoxides [37, 44] or carbonates/bicarbonates

[37, 45] may be formed. Then, the peaks observed during the anodic scan could be

linked to the decomposition of these promoter-derived species and the migration of the

K+ ions back to the solid electrolyte.

Hence two conclusions can be clearly obtained from this voltammogram:

- The applied potential of +2 V was high enough to electrochemically

decompose the promoter surface compounds, and this potential can be defined as a

reference un-promoted state for the further EPOC experiments.

- The reproducibility of the voltamogramm after the 2nd

cycle revealed a good

stability of the Au/YSZ catalyst during the polarizations under reaction conditions.

The reproducibility of the Au catalytic activity will be also confirmed in view of the

following experiments.

3.3.2. Electrochemical promotion via galvanostatic transitions

In order to study the EPOC phenomenon on the Au-YSZ catalyst film, several

closed and open circuit transitions were carried out under the following methanol

partial oxidation (POM) conditions: a feed composition of CH3OH/O2 = 5.9 %/0.43 %

(Ar balance) and a temperature of 280 ºC. The unpromoted catalyst was found to be

inactive below this temperature as shown on Figure 3.3.

Figure 3.5 shows the response vs. time in both the hydrogen, carbon dioxide

and methyl formate production rates and the catalyst potential (VWR) to the following

step changes: a positive potential (VWR = +2 V) for 32 min, a negative current (I = -5

Chapter 3

144

µA) for 65 min, open circuit (I = 0) for 37 min and again a positive potential (VWR =

+2 V) for 37 min.

Figure 3.5. Variation of H2 (a), CO2 and HCOOCH3 (b) production rates as well as the

corresponding enhancement ratios (ρi) and catalyst potential (VWR) (a) vs. time upon different

potentiostatic and galvanostatic transitions. CH3OH/O2 = 5.9 %/0.43 %, 280 ºC.

Electrochemical promotion under the imposition of I = -5 µA.

In order to quantify the magnitude of the electropromotional effect, the rate

enhancement ratio of each compound (ρ) is also shown on the figure, calculated by the

following equation:

=

(3.7)

-1

0

1

2

3

0

1

2

3

4

5

VW

R/

V

r H/

mo

l s-1

x 1

0-8

0 1 20.5 1.5Promoter supply / mol K+ x 10-7

VWR = 2 VVWR = 2 V I = 0 AI = -5 µAρ

H2

0

3

6

9

12

2a)

0

6

12

0

1

2

3

4

0 25 50 75 100 125 150 175

r HC

OO

CH

/

mo

l s-1

x 1

0-8

r CO

/ m

ol

s-1x

10

-8

Time / min

3 ρH

CO

OC

H3

0

2

4

6

2

b)

2

1

0

ρC

O2

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

145

where r and r0 are the promoted (VWR < +2 V) and unpromoted (VWR = +2 V) catalytic

reaction rates, respectively. Moreover, the inserted abscissa along the negative

polarization indicates the accumulated promoter supply, which varies linearly vs. time

in the case of galvanostatic experiments and is calculated from the following equation:

=

(3.8)

where n is the potassium ion charge in this case (+1), F is the Faraday constant (96485

C) and t is the polarization time.

The behaviour of the electrochemical catalyst has been evaluated upon the

different polarization regimes. Firstly, the application of a positive potential VWR = +2

V ensured the removal of K+ ions previously migrated to the catalyst surface by

thermal migration or as a consequence of the foregoing experiments [37, 38], and

resulted in a Au catalytic activity practically coincident with that observed at the same

temperature under open circuit conditions in Figure 3.3.

Then, at t = 32 min, a negative current I = -5 µA (current density, i = 1.6 µA

cm-2

) was applied between the Au-YSZ catalyst-working electrode and the Au counter

electrode, i.e. potassium promoter ions were electrochemically pumped from the solid

electrolyte to the Au/YSZ catalyst surface at a rate, = I/F = 5.18 x 10-11

mol K+ s

-1

according to the Faraday equation. The increase in the accumulated potassium supply

was found to enhance the overall methanol conversion (not shown) and the production

rates of hydrogen (Figure 3.5a), carbon dioxide and methyl formate (Figure 3.5b).

Selectivities toward H2 and HCOOCH3 were in turn enhanced vs. those toward CO2

and H2O. At the same time, the applied negative current caused a decrease in the

catalyst potential measured between the working and reference electrodes, VWR

(Figure 3.5a), as typically observed in studies with alkali solid electrolytes [38, 43,

46]. The trend of the catalytic activity instead depends on the EPOC behaviour of the

system. According to the theory of electrochemical promotion [3], the back-spillover

of potassium ions onto the Au-YSZ catalyst film under applied negative currents led

to the decrease in the catalyst potential and work function. It strengthened the Au

Chapter 3

146

chemical bond with the electron acceptor adsorbates (oxygen) and weakened that with

the electron donors (methanol) [47]. In fact, Broqvist et al. [48] already reported the

interest of the alkali compounds as “oxygen attractors” for significantly improving the

activity of CO oxidation on a Au-based catalyst. It should be noted that the methanol

conversion obtained in this work did not exceed the 10 % probably due to the low

surface area of the active catalyst film and reactants bypass. However, very

interestingly, Au nanoparticles dispersed in a YSZ matrix were electrochemically

promoted by K+ ions under methanol partial oxidation conditions, and the obtained

EPOC effect was electrophilic as also observed in chapters 1 and 2 for Pt-based

catalysts.

Other studies have also successfully promoted metal catalysts dispersed on

YSZ by both alkali electropromotion [49] and alkali conventional promotion [50]. In

the present work, maximum production rates of 3.7 x 10-8

mol H2 s-1

, 4.3 x 10-8

mol

CO2 s-1

and 1.5 x 10-7

mol HCOOCH3 s-1

were achieved at the end of the negative

polarization, which corresponded to enhancement ratios (ρ) of 8.9, 2.5 and 5.3,

respectively, with respect to the production rates obtained with the unpromoted Au-

YSZ. Such an improvement in the catalytic activity was caused by an accumulated

promoter supply of 2.02 x 10-7

mol K+ (equation 3.8). Hence, the corresponding

optimum potassium coverage, , of 0.32 can be estimated by considering the total

number of deposited Au sites, N = 6.38 x 10-7

mol Au, by the following equation:

=

(3.9)

It is also interesting to note that upon current interruption (open circuit

conditions, at t = 97 min) a permanent EPOC phenomena was observed probably due

to the remaining of the promoter phases on the catalyst surface under open circuit

conditions, i.e., all the products reaction rates were kept at their electro-promoted

values. In addition, a very slight increase in the catalyst potential close to the open

circuit potential value typically observed in this work ( ≈ -0.4 V) was detected.

This kind of irreversible promotional effect (Permanent EPOC or Permanent

NEMCA) has been reported elsewhere for alkali-electropromoted systems [37, 38,

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

147

46]. It has been mainly attributed to the high stability under reaction conditions of

certain promoter species formed on the catalyst surface, like potassium oxides or

superoxides [37, 38]. It is very important to highlight the practical application of the

observed irreversible EPOC behaviour: not only a very low current (a few µA) is

enough to enhance the catalytic performance of the gold nanoparticles by several

times, but also a first polarization step may be already sufficient to maintain this

catalytic modification for a certain time, with the consequent energy saving.

Finally, the imposition of a potential of +2 V again at t = 134 min restored the

initial production rates values within a few minutes, leading to a completely reversible

electropromotional effect. It was due to the decomposition of the promoter phases and

the electrochemical pumping of the K+ ions back to the solid electrolyte in good

agreement with the cyclic voltammetry. This reversible effect also sustains the

stability of the Au nanoparticles supported on the YSZ under EPOC conditions.

Hence, it is clear that the use of the YSZ support allowed to disperse the metal

nanoparticles, thus avoiding the metal sintering effect, very typical on this kind of

catalytic system.

Figures 3.6a and 3.6b show the results obtained in two additional experiments

under the same POM reaction conditions, which consisted of the same potentiostatic

and galvanostatic transitions than the previous one (Figure 3.5), but applying higher

currents during the negative polarization step: I = -10 µA (i = 3.2 µA cm-2

) in Figures

3.6a-1 and 3.6a-2, and -20 µA (i = 6.4 µA cm-2

) in Figures 3.6b-1 and 3.6b-2. The

application of a potential of +2 V at the beginning and the end of each experiment

defined again a Au-YSZ free of K-derived species (unpromoted state). In this way, all

the reaction rates remained almost at the same values upon every positive polarization,

confirming the complete reversibility of the promotional effect.

Chapter 3

148

Figure 3.6. Variation of H2, CO2 and HCOOCH3 production rates as well as the corresponding

rate enhancement ratios (ρ) and catalyst potential (VWR) vs. time upon different potentiostatic

and galvanostatic transitions. CH3OH/O2 = 5.9 %/0.43 %, 280 ºC. Electrochemical promotion

under the imposition of I = -10 µA (a-1,a-2), and -20 µA (b-1,b-2).

As also observed in the first experiment during the negative polarization, the

migration of potassium ions to the catalyst surface at a rates, = I/F = 1.04 x 10-10

mol K+ s

-1 (Figures group 3.6a) and 2.07 x 10

-10 mol K s

-1 (Figures group 3.6b) sharply

enhanced the gold catalytic activity. However, it can be found under the polarizations

at -10 µA, and more clearly, at -20 µA, that all the production rates decreased from a

certain accumulated amount of supplied promoter, which was not achieved upon the

imposition of -5 µA. This poisoning effect could be attributed to the excessive

formation of alkali-derived surface compounds and the concomitant blocking of active

sites, as widely observed with alkali-based promoted systems [44, 50-52]. Moreover,

the catalyst potential measured after 65 min of negative polarization decreased from -

0.55 V (Figure 3.5a) to -0.60 V (Figure 3.6a-1) and -0.74 V (Figure 3.6b-1), due to the

increase in the maximum potassium coverage achieved in each experiment.

-1

0

1

2

3

0

1

2

3

4

VW

R/

V

r H/

mo

l s-1

x 1

0-8

0 4 82 6Promoter supply / mol K+ x 10-7

VWR = 2 VVWR = 2 V I = 0 AI = -20 µA

2

b-1)

ρH

2

0

4

6

8

10

2

0

6

12

0

1

2

3

0 25 50 75 100 125 150 175

r HC

OO

CH

/

mo

l s-1

x 1

0-8

r CO

/ m

ol

s-1x

10

-8

Time / min

3 ρH

CO

OC

H3

0

2

4

6

2

b-2)

2

1

0

ρC

O2

-1

0

1

2

3

0

1

2

3

4

5

VW

R/

V

r H/

mo

l s-1

x 1

0-8

0 2 41 3Promoter supply / mol K+ x 10-7

VWR = 2 VVWR = 2 V I = 0 AI = -10 µA

ρH

2

0

3

6

9

12

2

a-1)

0

6

12

0

1

2

3

4

0 25 50 75 100 125 150 175

r HC

OO

CH

/

mo

l s-1

x 1

0-8

r CO

/ m

ol

s-1x

10

-8

Time / min

3 ρH

CO

OC

H3

0

2

4

6

2

a-2)

2

1

0

ρC

O2

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

149

A permanent EPOC was also observed in these experiments upon the

subsequent current interruption, although it was inevitably affected by the mentioned

poisoning effect. For instance, in the case of the hydrogen production, permanent rate

enhancement ratios ( ) of 8.7, 6.4 and 3.2 can be observed at the end of the open

circuit imposition in Figures 3.5a, 3.6a-1 and 3.6b-1, respectively. This parameter has

been calculated using the following equation:

=

(3.10)

where rper is the permanent catalytic production rate following the promoted state.

Nevertheless, the optimum amount of supplied K+ ions which maximized the catalytic

activity was found to be the same in each experiment, around 3 x 10-7

mol K+ (which

corresponded to a potassium coverage, , of about 0.5). In addition, the

maximum values of the reaction rates obtained were also practically the same, i.e., the

H2, CO2 and HCOOCH3 production rates were in all cases enhanced 9.2, 2.6 and 5.5

times, respectively, under optimum electrochemically promoted conditions. Then, it is

interesting to note that the applied current had no influence on the behaviour and

magnitude of the electropromotional effect, but only on the speed of the promotion

process. Finally, the stability of the Au/YSZ catalyst film in the different experiments

was demonstrated on the basis of the reproducible catalytic behaviour.

Besides the reaction rate enhancement ratio (ρ), other typical parameter used

in literature to describe the magnitude of the electrochemical promotion effect the

Faradaic efficiency (Λ). This parameter is calculated as follows:

Λ =

(3.11)

where Δr = r – r0 is the K+-induced change in the catalytic reaction rate and n is the

number of electrons involved in the corresponding electrocatalytic reaction. Under the

studied POM reaction conditions at 280 ºC, the increase in the oxygen catalytic

consumption rate (not shown) caused by the optimum promoter coverage was around

2.1 x 10-7

mol O s-1

, which was 4075 times larger than the rate of K+ supply ( =

Chapter 3

150

I/F) to the catalyst film under the application of -5 µA ( = 5.18 x 10-11

mol K+ s

-1).

The corresponding Faradaic efficiency (Λ) would be -8150, i.e. by considering the

equivalent oxygen electrocatalytic reaction (n = 2 in equation 3.11). However, in the

present study the source of promoter ions is not any O2-

conductor material such as

YSZ, but K-βAl2O3 (a cationic conductor). Thus, the Faradaic efficiency must be

considered with caution since the alkali ions cannot interact with the gas reactants at

the same extent [53]. Hence, as stated in Chapter 1, the most significant parameters

from a catalytic point of view to characterize the alkali-based promoted systems are

the rate enhancement ratio (ρ, equation 3.7) and the promotion index ( ) [3], the

latter being defined according to the following equation:

=

θ (3.12)

Figure 3.7b depicts the promotion index ( ) for the maximum production

rates of the different compounds (H2, CO2 and HCOOCH3) obtained in the previous

figures under optimum promoted conditions, and Figure 3.7a shows the corresponding

faradaic efficiency values related to the rate of K+ promoter supply (n = 1 in equation

3.11). This latter parameter has been denoted as and has been expressed in units

of mol product/mol K+. In other words, in both Figure 3.7a and 3.7b, three points are

depicted which correspond with the maximum catalytic rates observed in the three

previous experiments (Figures 3.5, 3.6a and 3.6b).

Firstly, a clear EPOC effect can be observed in all the experiments ( > 0)

with an electrophilic behaviour ( < -1), since as previously discussed the

application of negative currents and the consequent back-spillover of K+ promoter ions

increased the Au catalytic activity (I < 0, Δr > 0). In Figure 3.7a, methyl formate

presented the highest values, being the most abundant product obtained under the

studied reaction conditions. This parameter represents the ratio between the mol of

product and the mol of K+ electrochemically supplied. In this sense, under the

application of I = -5 µA, each mol K+ migrated to the Au-YSZ catalyst film allowed

the simultaneous production of up to more than 600 mol H2, 500 mol CO2 and 2300

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

151

mol HCOOCH3. Similar maximum production rates were achieved in all the

experiments despite increasing the rate of promoter supply. Consequently, the faradic

efficiency values decreased as higher negative current was applied.

Figure 3.7. Influence of the applied negative current on the optimum faradaic efficiency

relative to the rate of K+ supply ( ) (a) and the optimum promotion index ( ) (b)

of the different products. Data obtained from the results shown in Figures 3.5 and 3.6.

On the other hand, from Figure 3.7b, it can be drawn that hydrogen showed the

highest promotion index values. In fact, the H2 production rate presented the most

pronounced increase with respect to the unpromoted rate, as also reflected in the

highest ρ values shown in the previous figures. Higher optimum promotion indexes

( ) were found for all the products under the imposition of the lowest negative

current, I = -5 µA, because these data were obtained from the maximum production

0

5

10

15

20

25

-25-20-15-10-50

PI K

,o

pt

I / A x 10-6

+

H2

CO2

HCOOCH3

b)

-2500

-2000

-1500

-1000

-500

0

ΛK

,o

pt/

mo

l p

rod

uct

mo

l K

+ -

1+

H2

CO2

HCOOCH3

a)

Chapter 3

152

rates observed at the end of this polarization step (Figure 3.5), which did not

correspond with the actual optimum electropromoted conditions. With such a low rate

of K+ supply in the first experiment (5.18 x 10

-11 mol K s

-1), the polarization time

should have been extended for another 35 min to achieve the optimum potassium

coverage (around 0.5). In this sense, the slight difference between the depicted

values obtained at I = -10 µA and -20 µA could be attributed to the non-

complete coincidence in the respective potassium coverages, which was likely due to

slight differences in the gas chromatograph analysis time. However, fixed promotion

index values of around 16, 3 and 9 for hydrogen, carbon dioxide and methyl formate,

respectively, could be considered, regardless the magnitude of applied negative

current, under optimum electrochemically promoted conditions ( ≈ 0.5).

Hence, in contrast to the parameter, it was demonstrated the independence of the

promotion index with respect to the rate of K+ ions supply, being possible to induce

the maximum improvement in the Au-YSZ catalytic activity under given reaction

conditions as long as a specific potassium coverage is reached.

3.3.3. Electrochemical promotion via potentiostatic transitions

Other kind of electrochemical promotion experiment under the same methanol

partial oxidation conditions consisted of successive potentiostatic transitions, similar

to the experiments carried out in the two first chapters. Figure 3.8 shows the dynamic

response in the H2 (a), CO2 (b) and HCOOCH3 (c) production rates to step changes in

the applied catalyst potential (VWR) at 280 ºC. A positive polarization at VWR = +2 V

before each different potential allowed starting from a reference (unpromoted) state in

all cases. The subsequent imposition of VWR = 0.5 V did not involve any modification

in the catalytic activity since it still remained above the open circuit potential value.

Then, the negative polarization at VWR = -0.5 V caused a sharp increase of all the

production rates as a result of the back-spillover of the promoter ions from the K-

βAl2O3 electrolyte to the Au-YSZ catalyst-working electrode (VWR ≤ -0.5 V, Δr > 0).

As noted in the previous galvanostatic experiments, the observed EPOC effect could

be explained in terms of the work function variation and the consequent change in

chemisorptive bond strengths on the catalyst surface [3].

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

153

Figure 3.8. H2 (a), CO2 (b) and HCOOCH3 (c) production rates vs. time under different

applied potentials (VWR) between 2 and -1 V. CH3OH/O2 = 5.9 %/0.43 %, 280 ºC.

Very interestingly, under this negative potential (VWR = -0.5 V) the steady-

state H2, CO2 and HCOOCH3 production rates were close to those obtained under

optimum electropromoted conditions in previous section, with enhancement ratios (ρ)

-2

-1

0

1

2

3

0

1

2

3

4

VW

R/

V

r H/

mo

l s-1

x 1

0-8

a)2

rH VWR

2

-2

-1

0

1

2

0

1

2

3

4

VW

R/

V

r CO

/

mo

l s-1

x 1

0-8

2rCO VWR

b)

2

-2

-1

0

1

2

0

6

12

0 100 200 300 400

VW

R/

V

r HC

OO

CH

/ m

ol s

-1x

10

-8

Time / min

3rHCOOCH VWRc)

3

Chapter 3

154

of around 8, 2 and 5, respectively. The associated amount of migrated promoter at the

end of this polarization was 3.58 x 10-7

mol K+, which was calculated from the

integration of the current vs. time curve (not shown here) via equation 3.8. It

corresponded to a potassium coverage equal to 0.56 (equation 3.9), which was

slightly higher than the optimum value found in previous experiments (around 0.5).

This fact could indicate some possible blocking of Au active sites by an excess of K-

derived surface compounds [44, 45] and explain the mildly lower catalytic activity

obtained under these promoted conditions compared to the maximum achieved in

Figures 3.6a and 3.6b. This poisoning effect was more clearly observed under the

application of VWR = -1 V. However, all the promoter species formed during the

negative polarizations were decomposed upon each successive imposition of VWR = +2

V, again proving, in good agreement with the Au-YSZ characterization results, the

reversibility and reproducibility of the EPOC phenomenon and the stability of the

YSZ-supported Au nanoparticles. Thus, in this experiment the electrochemical

catalyst definitely showed a similar qualitative and quantitative behaviour as that

observed in previous figures under applied negative currents.

Finally, the production rate of the different compounds was compared at

different temperatures with the same feed composition (CH3OH/O2 = 5.9 %/0.43 %)

under both unpromoted and promoted conditions. Figure 3.9 shows the variation of the

steady-state production rates and the accumulated promoter supply vs. the applied

catalyst potential at 280 and 300 ºC. The corresponding rate enhancement ratios (ρopt)

and promotion indexes ( ), calculated by equations 3.7 and 3.12 respectively,

are also indicated for the optimum applied potential in each case. Starting from the

reference positive polarization, VWR = +2 V (un-promoted state), the decrease in the

applied potential led to the electrochemical pumping of K+ ions from the solid

electrolyte to the Au-YSZ catalyst film, according to the obtained negative currents

(see inset of Figure 3.9d for the application of VWR = -0.5 V). As typically observed on

K-βAl2O3-based electrocatalytic systems under potentiostatic transitions [38], the

resultant negative currents were more pronounced at the beginning of the polarization

and then decreased to negligible residual values of the order of -1 µA.

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

155

Figure 3.9. Steady-state H2 (a), CO2 (b) and HCOOCH3 (c) production rates and accumulated

promoter supply (d) vs. applied potential (VWR) as well as the optimum production rates

enhancement ratios (ρi,opt) and promotion indexes ( ) at 280 and 300 ºC. CH3OH/O2 =

5.9 %/0.43 %. Inset of Figure 3.9d shows the current vs. time curve during the potentiostatic

imposition of -0.5 V.

All the production rates ( ,

and ) were increased by the

potassium promotional effect at both studied temperatures, showing an electrophilic

EPOC behavior (VWR ≤ -0.5 V, Δr > 0, ρ > 1 and > 0). Moreover, in both cases

the maximum catalytic activity was found at the same optimum applied potential,

VWR,opt = -0.5 V. This optimum potential corresponded with the migration of 3.58 x

10-7

and 4.74 x 10-7

mol K+ and potassium coverages of 0.56 and 0.74, at 280 and 300

ºC, respectively. The higher the reaction temperature, the higher the solid electrolyte

ionic conductivity and, therefore, the higher the amount of electrochemically

transferred K+ ions at fixed potential. It can be observed an increase in the catalytic

activity of the Au-YSZ catalyst film with the temperature under unpromoted

conditions (positive potentials) as typically observed in POM studies with supported

0

0.5

1

1.5

2

-1.5 -0.5 0.5 1.5 2.5

Acc

um

ula

ted

pro

mo

ter

sup

ply

/ m

ol K

+x

10-6

VWR / V

280 ºC

300 ºC

d)

-1

-0.6

-0.2

0.2

0 20 40 60

I / A

x 1

0-4

Time / min

280 ºC

300 ºC

-0.5 V

0

0.5

1

1.5

2

2.5

3

3.5

r H

/ m

ol s

-1x

10

-8280 ºC

300 ºC2

a) ρH ,opt = 7.9 PIK ,H ,opt = 12.32 2

+

ρH ,opt = 3.6 PIK ,H ,opt = 3.52 2

+

0

4

8

12

16

r HC

OO

CH

/

mo

l s-1

x 1

0-8

280 ºC

300 ºC

3

c) ρHCOOCH ,opt = 5.1 PIK ,HCOOCH ,opt = 7.33 3

+

ρHCOOCH ,opt = 2.7 PIK ,HCOOCH ,opt = 2.33 3

+

1

1.5

2

2.5

3

3.5

4

-1.5 -0.5 0.5 1.5 2.5

r CO

/

mo

l s-1

x 1

0-8

VWR / V

280 ºC

300 ºC

2

b)ρCO ,opt = 1.9 PIK ,CO ,opt = 1.2

2 2+

ρCO ,opt = 2.3 PIK ,CO ,opt = 2.32 2

+

Chapter 3

156

Au catalysts [19, 29-33]. However, under electrochemical promotion (VWR = -0.5 and

-1 V) the overall methanol conversion decreased at 300 ºC, very likely due to the

poisoning effect of the excessive promoter-derived species. In fact, the value of

potassium coverage, = 0.74, obtained at this temperature under the application of

-0.5 V was already far above the optimum value found in previous experiments

performed at 280 ºC ( ≈ 0.5). Furthermore, in good agreement with several

studies [13, 30, 31, 33], an increase in the temperature favored the formation of CO2

and hindered the HCOOCH3 selectivity due to the further oxidation of the latter.

As already remarked, the in-situ modification of the activity and selectivity of

a metallic catalysts by the controlled pumping of alkali ions (via EPOC) is of

paramount interest. This feature has been studied for different catalytic reactions such

as the epoxidation of alkenes [44], the Fischer-Tropsch synthesis [45], the selective

catalytic reduction of nitrogen oxides [51-53], and, in the first chapters, for H2 and

formaldehyde production from methanol partial oxidation. In the present work, gold

nanoparticles dispersed in a YSZ matrix were catalytically activated by K+ ions for the

production of hydrogen and methyl formate. Hence, two products of great interest can

be produced simultaneously and selectively by means of the electrochemical

promotion.

3.4. CONCLUSIONS

The following conclusions could be drawn from this study:

- An electrochemical catalyst based on Au nanoparticles dispersed in a YSZ

matrix was prepared by reactive co-sputtering of Zr-Y and Au targets. Similar to the

cathodic arc deposition in the previous chapter, this technique allowed depositing a

thin catalyst film (~170 nm) with very low particle size (~3 nm) and metal loading

(only 40 µg Au cm-2

).

- The obtained Au-YSZ catalyst showed to be active in the partial oxidation of

methanol with a very high selectivity toward methyl formate production, and it was

polarized by using a current collector, even without providing any electrical

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

157

conductivity. Hence, the Au-YSZ was successfully promoted by electrochemical

pumping potassium ions under negative polarization.

- The stability of the Au/YSZ catalyst film was verified by different

characterization techniques, demonstrating the interest of the catalyst preparation

technique and the use of the YSZ support to disperse metal nanoparticles by avoiding

metal sintering.

- Under given reaction conditions, the magnitude of the electrochemical

promotion effect with alkali solid electrolyte did not depend on the applied current, i.e.

the rate of promoter supply, but on the promoter coverage. The H2, CO2 and

HCOOCH3 production rates were enhanced, regardless the operation mode

(galvanostatic or potentiostatic), more than 9, 2 and 5 times, respectively, at 280 ºC

upon an optimum potassium coverage of around 0.5.

- A permanent electropromoted effect was found upon negative current

interruption, showing the stability of the formed promoter species under the studied

reaction conditions, which may be of great interest for practical applications.

3.5. REFERENCES

[1] W.D. Mross, ALKALI DOPING IN HETEROGENEOUS CATALYSIS, Catalysis reviews, 25

(1983) 591-637.

[2] R.J. Farrauto, C.H. Bartholomew, Fundamentals of industrial catalytic processes, Chapman

& Hall, London, 1997.

[3] C.G. Vayenas, S. Bebelis, C. Pliangos, S. Brosda, D. Tsiplakides, Electrochemical

Activation of Catalysis: Promotion, Electrochemical Promotion and Metal-Support

Interactions, Kluwer Academic Publishers/Plenum Press, New York, 2001.

[4] A. de Lucas-Consuegra, New trends of Alkali Promotion in Heterogeneous Catalysis:

Electrochemical Promotion with Alkaline Ionic Conductors, Catalysis Surveys from Asia,

2015, DOI:10.1007/s10563-014-9179-6.

[5] O.A. Mar'ina, V.A. Sobyanin, The effect of electrochemical oxygen pumping on the rate of

CO oxidation on Au electrode-catalyst, Catalysis Letters, 13 (1992) 61-69.

Chapter 3

158

[6] O.A. Mar'ina, V.A. Sobyanin, V.D. Belyaev, V.N. Parmon, The effect of electrochemical

oxygen pumping on catalytic properties of Ag and Au electrodes at gas-phase oxidation of

CH4, Catalysis Today, 13 (1992) 567-570.

[7] I.M. Petrushina, V.A. Bandur, F. Cappeln, N.J. Bjerrum, Electrochemical promotion of

sulfur dioxide catalytic oxidation, Journal of the Electrochemical Society, 147 (2000) 3010-

3010.

[8] F.M. Sapountzi, M.N. Tsampas, C.G. Vayenas, Electrochemical promotion of CO

conversion to CO2 in PEM fuel cell PROX reactor, Catalysis Today, 146 (2009) 319-325.

[9] X. Deng, B.K. Min, A. Guloy, C.M. Friend, Enhancement of O2 dissociation on Au(111) by

adsorbed oxygen: Implications for oxidation catalysis, Journal of the American Chemical

Society, 127 (2005) 9267-9270.

[10] G. Avgouropoulos, T. Tabakova, J.J. Spivey, Environmental Catalysis over Gold-Based

Materials, The Royal Society of Chemistry, Cambridge, 2013.

[11] M. Haruta, Gold as a novel catalyst in the 21st century: Preparation, working mechanism

and applications, Gold Bulletin, 37 (2004) 27-36.

[12] V. Zielasek, B. Jürgens, C. Schulz, J. Biener, M.M. Biener, A.V. Hamza, M. Bäumer,

Gold catalysts: Nanoporous gold foams, Angewandte Chemie - International Edition, 45

(2006) 8241-8244.

[13] A. Wittstock, V. Zielasek, J. Biener, C.M. Friend, M. Bäumer, Nanoporous gold catalysts

for selective gas-phase oxidative coupling of methanol at low temperature, Science, 327 (2010)

319-322.

[14] S. Gil, M. Marchena, C.M. Fernández, L. Sánchez-Silva, A. Romero, J.L. Valverde,

Catalytic oxidation of crude glycerol using catalysts based on Au supported on carbonaceous

materials, Applied Catalysis A: General, 450 (2013) 189-203.

[15] G.C. Bond, D.T. Thompson, Gold-catalysed oxidation of carbon monoxide, Gold Bulletin,

33 (2000) 41-50.

[16] T. Kobayashi, M. Haruta, S. Tsubota, H. Sano, B. Delmon, Thin films of supported gold

catalysts for CO detection, Sensors and Actuators: B. Chemical, 1 (1990) 222-225.

[17] S.S. Pansare, A. Sirijaruphan, J.G. Goodwin Jr, Au-catalyzed selective oxidation of CO: A

steady-state isotopic transient kinetic study, Journal of Catalysis, 234 (2005) 151-160.

[18] H.A.E. Dole, J.M. Kim, L. Lizarraga, P. Vernoux, E.A. Baranova, Catalytic CO oxidation

over Au nanoparticles supported on yttria-stabilized zirconia, ECS Transactions, 2012, pp.

265-274, DOI:10.1149/1.3701316.

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

159

[19] A.N. Pestryakov, V.V. Lunin, N. Bogdanchikova, O.N. Temkin, E. Smolentseva, Active

states of gold in small and big metal particles in CO and methanol selective oxidation, Fuel,

110 (2013) 48-53.

[20] H. Sakurai, M. Haruta, Synergism in methanol synthesis from carbon dioxide over gold

catalysts supported on metal oxides, Catalysis Today, 29 (1996) 361-365.

[21] M.C. Kung, K.A. Bethke, J. Yan, J.H. Lee, H.H. Kung, Catalysts for lean NOx reduction:

Structure-property relationship, Applied Surface Science, 121-122 (1997) 261-266.

[22] N. Yi, R. Si, H. Saltsburg, M. Flytzani-Stephanopoulos, Active gold species on cerium

oxide nanoshapes for methanol steam reforming and the water gas shift reactions, Energy and

Environmental Science, 3 (2010) 831-837.

[23] A. Ciftci, D.A.J.M. Ligthart, P. Pastorino, E.J.M. Hensen, Nanostructured ceria supported

Pt and Au catalysts for the reactions of ethanol and formic acid, Applied Catalysis B:

Environmental, 130-131 (2013) 325-335.

[24] G. Walther, L. Cervera-Gontard, U.J. Quaade, S. Horch, Low temperature methane

oxidation on differently supported 2 nm Au nanoparticles, Gold Bulletin, 42 (2009) 13-19.

[25] M.D. Hughes, Y.J. Xu, P. Jenkins, P. McMorn, P. Landon, D.I. Enache, A.F. Carley, G.A.

Attard, G.J. Hutchings, F. King, E.H. Stitt, P. Johnston, K. Griffin, C.J. Kiely, Tunable gold

catalysts for selective hydrocarbon oxidation under mild conditions, Nature, 437 (2005) 1132-

1135.

[26] A. Abad, P. Concepción, A. Corma, H. García, A collaborative effect between gold and a

support induces the selective oxidation of alcohols, Angewandte Chemie - International

Edition, 44 (2005) 4066-4069.

[27] P.Y. Sheng, G.A. Bowmaker, H. Idriss, The Reactions of Ethanol over Au/CeO2, Applied

Catalysis A: General, 261 (2004) 171-181.

[28] H.C. Yang, F.W. Chang, L.S. Roselin, Hydrogen production by partial oxidation of

methanol over Au/CuO/ZnO catalysts, Journal of Molecular Catalysis A: Chemical, 276 (2007)

184-190.

[29] B.P.C. Hereijgers, T.M. Eggenhuisen, K.P. De Jong, H. Talsma, A.M.J. Van Der Eerden,

A.M. Beale, B.M. Weckhuysen, Understanding the promotion effect of lanthanum oxide on

gold-based catalysts in the partial oxidation of methanol by in situ XAFS and DSC studies,

Journal of Physical Chemistry C, 115 (2011) 15545-15554.

Chapter 3

160

[30] F.W. Chang, S.C. Lai, L.S. Roselin, Hydrogen production by partial oxidation of

methanol over ZnO-promoted Au/Al2O3 catalysts, Journal of Molecular Catalysis A:

Chemical, 282 (2008) 129-135.

[31] W. Jianxin, L. Laitao, A comparative study of partial oxidation of methanol over zinc

oxide supported metallic catalysts, Catalysis Letters, 126 (2008) 325-332.

[32] T.C. Ou, F.W. Chang, L.S. Roselin, Production of hydrogen via partial oxidation of

methanol over bimetallic Au-Cu/TiO2 catalysts, Journal of Molecular Catalysis A: Chemical,

293 (2008) 8-16.

[33] K. Kähler, M.C. Holz, M. Rohe, A.C. Van Veen, M. Muhler, Methanol oxidation as probe

reaction for active sites in Au/ZnO and Au/TiO2 catalysts, Journal of Catalysis, 299 (2013)

162-170.

[34] S. Scirè, L.F. Liotta, Supported gold catalysts for the total oxidation of volatile organic

compounds, Applied Catalysis B: Environmental, 125 (2012) 222-246.

[35] R. Catrin, T. Gries, B. Raillard, F. Mücklich, S. Migot, D. Horwat, Influence of laser

interference patterning on microstructure and friction behavior of gold/yttria-stabilized

zirconia nanocomposite thin films, Journal of Materials Research, 27 (2012) 879-885.

[36] H.P. Klug, L.E. Alexander, X-Ray Diffraction Procedures for Polycrystalline and

Amorphous Materials, Wiley, New York, 1974.

[37] A. de Lucas-Consuegra, F. Dorado, J.L. Valverde, R. Karoum, P. Vernoux, Low-

temperature propene combustion over Pt/K-βAl2O3 electrochemical catalyst:

Characterization, catalytic activity measurements, and investigation of the NEMCA effect,

Journal of Catalysis, 251 (2007) 474-484.

[38] A. de Lucas-Consuegra, A. Princivalle, A. Caravaca, F. Dorado, C. Guizard, J.L.

Valverde, P. Vernoux, Preferential CO oxidation in hydrogen-rich stream over an

electrochemically promoted Pt catalyst, Applied Catalysis B: Environmental, 94 (2010) 281-

287.

[39] J.M. Tatibouët, Methanol oxidation as a catalytic surface probe, Applied Catalysis A:

General, 148 (1997) 213-252.

[40] N.W. Cant, S.P. Tonner, D.L. Trimm, M.S. Wainwright, Isotopic labeling studies of the

mechanism of dehydrogenation of methanol to methyl formate over copper-based catalysts,

Journal of Catalysis, 91 (1985) 197-207.

[41] J.S. Lee, J.C. Kim, Y.G. Kim, Methyl formate as a new building block in C1 chemistry,

Applied Catalysis, 57 (1990) 1-30.

Electrochemical activation of Au nanoparticles dispersed in YSZ for methanol partial

oxidation

161

[42] N. Kotsionopoulos, S. Bebelis, Electrochemical characterization of the Pt/β″-Al 2O3

system under conditions of in situ electrochemical modification of catalytic activity for

propane combustion, Journal of Applied Electrochemistry, 40 (2010) 1883-1891.

[43] N. Kotsionopoulos, S. Bebelis, In situ electrochemical modification of catalytic activity

for propane combustion of Pt/β-Al2O3 catalyst-electrodes, Topics in Catalysis, 44 (2007) 379-

389.

[44] A. Palermo, A. Husain, M.S. Tikhov, R.M. Lambert, Ag-catalysed epoxidation of propene

and ethene: An investigation using electrochemical promotion of the effects of alkali, NOx, and

chlorine, Journal of Catalysis, 207 (2002) 331-340.

[45] A.J. Urquhart, J.M. Keel, F.J. Williams, R.M. Lambert, Electrochemical Promotion by

Potassium of Rhodium-Catalyzed Fischer-Tropsch Synthesis: XP Spectroscopy and Reaction

Studies, Journal of Physical Chemistry B, 107 (2003) 10591-10597.

[46] C.G. Vayenas, S. Bebelis, M. Despotopoulou, Non-faradaic electrochemical modification

of catalytic activity 4. The use of β″-Al2O3 as the solid electrolyte, Journal of Catalysis, 128

(1991) 415-435.

[47] C.G. Vayenas, S. Neophytides, Non-faradaic electrochemical modification of catalytic

activity III. The case of methanol oxidation on Pt, Journal of Catalysis, 127 (1991) 645-664.

[48] P. Broqvist, L.M. Molina, H. Grönbeck, B. Hammer, Promoting and poisoning effects of

Na and Cl coadsorption on CO oxidation over MgO-supported Au nanoparticles, Journal of

Catalysis, 227 (2004) 217-226.

[49] A. De Lucas-Consuegra, A. Caravaca, P.J. Martínez, J.L. Endrino, F. Dorado, J.L.

Valverde, Development of a new electrochemical catalyst with an electrochemically assisted

regeneration ability for H2 production at low temperatures, Journal of Catalysis, 274 (2010)

251-258.

[50] I.V. Yentekakis, R.M. Lambert, M.S. Tikhov, M. Konsolakis, V. Kiousis, Promotion by

sodium in emission control catalysis: A kinetic and spectroscopic study of the Pd-catalyzed

reduction of NO by propene, Journal of Catalysis, 176 (1998) 82-92.

[51] F. Dorado, A. de Lucas-Consuegra, C. Jiménez, J.L. Valverde, Influence of the reaction

temperature on the electrochemical promoted catalytic behaviour of platinum impregnated

catalysts for the reduction of nitrogen oxides under lean burn conditions, Applied Catalysis A:

General, 321 (2007) 86-92.

Chapter 3

162

[52] I.V. Yentekakis, A. Palermo, N.C. Filkin, M.S. Tikhov, R.M. Lambert, In situ

electrochemical promotion by sodium of the platinum-catalyzed reduction of NO by propene,

Journal of Physical Chemistry B, 101 (1997) 3759-3768.

[53] P. Vernoux, F. Gaillard, C. Lopez, E. Siebert, Coupling catalysis to electrochemistry: A

solution to selective reduction of nitrogen oxides in lean-burn engine exhausts?, Journal of

Catalysis, 217 (2003) 203-208.

163

Chapter 4

ELECTROCHEMICAL PROMOTION OF Cu

NANOCOLUMNS IN THE PARTIAL OXIDATION OF

METHANOL:

EPOC WITH HIGHLY POROUS NON-NOBLE METAL

CATALYSTS

A novel Cu catalyst film has been prepared by oblique angle physical vapour

deposition (OAD) on a K-βAl2O3 solid electrolyte (alkaline ionic conductor). This

technique has allowed to obtain a highly porous and electrically conductive Cu

catalyst electrode which has been electrochemically promoted in the partial oxidation

of methanol (POM). The production rates of hydrogen, carbon dioxide and methyl

formate have been in-situ enhanced in a reversible and reproducible way, by means of

the controlled migration of electropositive potassium ions. Under the studied reaction

conditions these promoter ions formed potassium-derived surface compounds as

demonstrated by post-reaction characterization analysis. Some nitrogen functional

groups and carbonaceous compounds were also detected. The obtained results

demonstrate the interest of the used catalyst-electrode preparation technique for the

electrochemical activation of non-noble metal catalyst films.

0

2

4

6

8

10

-1.5

-1

-0.5

0

0.5

1

1.5

0 100 200 300 400 500 600

r /

mo

l s-1

x 1

0-8

VW

R/

V

Time / min

VWR

rH2rCO2

rHCOOCH3

Δr

CH3OH,

O2

H2, CO2,

HCOOCH3, H2O

4.4% CH3OH0.3% O2

320 ºC

2 µm

Nanocolumnar Cu catalyst film

Cu

K+ K+ K+ K-βAl2O3

Au

ΔV < 0

e-

e-

Chapter 4

164

4.1. INTRODUCTION

The phenomenon of Electrochemical Promotion of Catalysis (EPOC) or

NEMCA effect has been applied with a large variety of catalysts and solid electrolytes

(cationic, anionic or mixed conductors) and in a wide variety of catalytic reactions of

industrial and environmental interest [1]. However, as already pointed out, a further

technological progress is necessary for its practical application. Nowadays, some of

the main challenges of EPOC are the design of scaled-up reactors and the material cost

minimization. With respect to the former, increasingly more compact and efficient

reactors designs can be found in literature based, for example, on multi-pellet

configurations [2], hollow fibre membranes [3] or a sophisticated monolithic

electropromoted reactor (MEPR) [4-7]. Regarding the second outlook, several EPOC

studies have dealt with the improvement of catalytic materials and the development of

electrodes consisting of catalysts highly dispersed on gold [8], mixed ionic-electronic

conductors [9], different kinds of carbonaceous materials [10, 11] or yttria-stabilized

zirconia (YSZ) [12-14]. Other good examples have been reported in chapters 2 and 3

of this manuscript, where Pt and Au nanoparticles dispersed in carbon and YSZ

phases, respectively, were successfully electrochemically promoted for methanol

conversion with high selectivity toward H2 and methyl formate, respectively. A

complementary strategy on cost minimization would be the use of catalysts with a

higher extent of base metals. Some EPOC studies have already employed Ni- [10, 15-

18], Fe- [2, 19, 20] or Cu-based [7, 21-25] catalyst films in different reactions. In this

sense, the present chapter and the following have been focused on the electrochemical

promotion of non-noble metals for H2 production processes from CH3OH.

On the other hand, numerous properties of these metallic catalyst films such as

their porosity, surface area or metal dispersion, which are determining factors in the

performance of the catalytic reactions, strongly depend on the synthesis method.

Different techniques have been used in EPOC studies for the preparation of the active

catalyst film, some of which have been mentioned throughout the previous chapters.

Very common procedures for ease of application are, for instance, the deposition and

calcination of an organometallic paste [2, 14, 17, 19], the impregnation of a metal

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

165

precursor solution [8, 9, 15] or the spray deposition in a carbon matrix [11]. Other less

conventional techniques which enable a better control of the microstructure of the film

include Physical Vapour Deposition (PVD) methods such as Pulsed Laser Deposition

(PLD) [26], sputtering [4, 6, 7, 21, 22] or Cathodic Arc Deposition (CAD), the two

latter already employed in chapters 3 and 2, respectively.

In the present work, a modification of the classical PVD evaporation technique

called Oblique or Glancing Angle Deposition (OAD or GLAD) for the preparation of

the film is proposed. By this method, the substrate is placed in an oblique or glancing

angle configuration with respect to the evaporated flux of deposition material to

enhance the shadowing effects during the film growth [27]. As a result, highly porous

films formed by tilted nanocolumns are obtained, which are characterized by a high

gas-exposed surface area and a controlled microstructure depending on the vapour

incident angle, the rotation speed of the substrate, and the temperature during the

deposition, among other factors. The open porosity of these films is very useful to

prepare host materials for devices with optical, photonic or magnetic applications [28,

29] or for the development of gas and liquid sensors [27]. They have also been used

for different energy related applications including photoelectrochemical cells [27],

solid-state hydrogen storage [30] and electrocatalysis for oxygen reduction in polymer

electrolyte membrane fuel cells [31] and for the 2-propanol oxidation in alkaline direct

alcohol fuel cells [32].

Hence, a novel Cu electrode prepared by OAD with both suitable catalytic and

electrical properties was employed in the partial oxidation of methanol (POM) by

electrochemical promotion. As already mentioned, this exothermic reaction is

especially interesting for on board H2 production due to the liquid nature of methanol,

its high H/C ratio and its availability from a wide variety of sources. The selection of a

Cu catalyst is justified by its lower cost in comparison with noble metals such as Pt or

Pd and by its proven excellent catalytic activity for the POM reaction [33-37]. For this

choice, we have also considered previous works showing the suitability of Cu-based

catalysts to study fundamental aspects of the EPOC phenomenon [38] and its previous

application in different catalytic reactions such as NO reduction by CO [21, 22]

Chapter 4

166

preferential oxidation of CO [23] and CO2 hydrogenation [7, 24]. However, to the best

of our knowledge, this is the first time that this non-noble metal catalyst is

electrochemically promoted in a H2 production reaction.

4.2. EXPERIMENTAL

4.2.1. Preparation of the electrochemical catalyst

The electrochemical catalyst consisted of a continuous thin Cu layer

(geometric area of 2.01 cm2), which also behaved as working electrode (WE).

Similarly to the preceding chapters, a 19-mm-diameter, 1-mm-thick K-βAl2O3

(Ionotec) pellet was used as cationic solid electrolyte (i.e. as source of K+ promoter

ions), inert Au counter (CE) and reference (RE) electrodes were prepared from an

oganometalic paste (Fuel Cell Materials 233001). The active Cu catalyst-working

electrode film was deposited by evaporation of a Cu target (Goodfellow, 99.9999 %

purity) under vacuum conditions by bombardment with a high kinetic energy (< 5

keV) and intensity (150 mA) electron beam. The vapor was condensed onto the

surface of the substrate (K-βAl2O3) at room temperature in two steps by varying the

zenithal angle, α, formed between the perpendicular to the substrate surface and the

evaporation direction. Firstly, in order to provide a good electrical conductivity to the

catalyst, a fairly compact Cu layer was grown at α = 0º on the K-βAl2O3 pellet. The

microstructure of this compact film consisted of adjacent straight nanocolumns

perpendicular to the substrate (final inclination angle, β = 0º) and an approximate

thickness of 0.8 µm. Subsequently, a porous Cu layer with a tilted columnar

microstructure was deposited in a similar way as the former but fixing the zenithal

evaporation angle at α = 80º. This top layer presented a porosity of around 50 %

associated with large void spaces between separated nanocolumns tilted by an angle β

≈ 60º with respect to the perpendicular to the substrate. It was expected that the high

porosity of this outer layer favours the catalytic activity of the copper film. The total

thickness of the obtained Cu catalyst-working electrode was 1.6 µm approximately,

and the final metal loading was 1.08 mg Cu cm-2

(1.69 x 10-5

mol Cu cm-2

). Thus, this

novel preparation technique allowed to prepare a suitable copper electrode combining

both good electrical and catalytic properties, as shown below. Similar layers were also

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

167

deposited at different evaporation angles on silicon supports for characterization

purposes. More details about the experimental setup employed in the thin catalyst film

preparation can be found elsewhere [39, 40].

This catalyst preparation method was carried out in collaboration with Dr.

Agustín R. González-Elipe and Dr. Víctor J. Rico from the Institute of Materials

Science of Seville (CSIC – University of Seville).

The resulting Cu/K-βAl2O3/Au electrochemical catalyst was placed into the

single chamber solid electrolyte cell reactor and, although the Cu catalyst film

presented good electrical conductivity after its deposition (surface electrical resistance

of around 8 Ω), prior to the catalytic activity measurements it was subjected to a 25 %

H2 stream (Ar balance) while heating to 280 ºC (ramp of 5 ºC/min) to ensure its

complete reduction. The electrochemical promotion (EPOC) experiments were carried

out in the potentiostatic operation mode, i.e., by varying the applied catalyst potential

measured between the working and reference electrodes (VWR) according to the

procedure generally used in conventional three-electrodes electrochemical cells [41].

4.2.2. Characterization measurements

The deposited porous Cu layer was firstly characterized by X-Ray Diffraction

(XRD) analysis using a Philips PW-1710 instrument with Cu Kα radiation (λ = 1.5404

Å). The surface and in-depth microstructure of the Cu films deposited on silicon

substrates were examined by plan-view and cross-sectional scanning electron

microscopy (SEM) analysis using a Hitachi S4800 field emission microscope operated

at 2 keV. Given the possibility of Cu oxidation under reaction conditions and thus

electrical conductivity loss, the in-plane surface electrical resistance was continuously

measured by connecting the catalyst film to an additional gold wire and a digital

multimeterv (as in chapter 2 during the temperature-programmed experiment). Cyclic

voltammetry (CV) was also carried out at a scan rate of 5 mV s-1

with the potentiostat-

galvanostat (PGSTAT320-N, Metrohom Autolab) before the EPOC experiments.

Moreover, the electrochemical catalyst was characterized after the

electrochemical promotion experiments. Previously, the Cu catalyst film was exposed

Chapter 4

168

to the reaction mixture (CH3OH/O2 = 4.4 %/0.3 %) at 320 ºC while applying a

negative potential, VWR = -0.5 V in order to electrochemically supply K+ from the K-

βAl2O3 pellet to the Cu electrode. After 1 hour, the electrochemical catalyst was

cooled down to room temperature under reaction atmosphere. The applied potential

was interrupted (open circuit conditions) at 100 ºC. The aim of this procedure was to

“freeze out” the catalyst surface conditions pertaining to its electropromoted state.

Then the electrochemical catalyst was transferred to different characterization

equipments. In first place, the used catalyst film was characterized by XRD to

examine the final crystalline structure. SEM was performed along with Energy-

Dispersive X-ray spectroscopy (EDX) elemental mapping analysis by means of a

Bruker X-Flash Detector 4010 to identify the surface segregation of different

elements. X-ray photoelectron spectra (XPS) were also recorded with a PHOIBOS-

100 spectrometer with Delay Line Detector (DLD) from SPECS, which worked in the

constant pass energy mode fixed at 30 eV. Monochromatic Mg Kα radiation was used

as the excitation source and the binding energy (BE) scale of the spectra was

referenced to the C 1s of graphitic carbon taken at 284.6 eV.

4.2.3. Catalytic activity measurements

The catalytic activity measurements were carried out in the experimental setup

described in section 1.2.3. Methanol partial oxidation (POM) experiments were

performed at atmospheric pressure with an overall gas flow rate of 6 NL h-1

,

temperatures from 280 ºC to 360 ºC and a feed composition of CH3OH/O2 = 4.4 %/0.3

% (Ar balance). The detected products were: H2, CO2, H2O and HCOOCH3 (methyl

formate). The error in the carbon atom balance did not exceed 5 %, indicating no

consistent loss of material and no significant formation of other oxygenated species.

4.3. RESULTS AND DISCUSSION

4.3.1. Preliminary characterization of the catalyst film

Figure 4.1 shows the X-ray diffractogram of an oblique angle deposited copper

film on a silicon substrate at α = 80º. This diffractogram evidences that the as-

deposited Cu catalyst film exhibited a well-defined face-centered cubic (FCC)

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

169

crystalline structure (JCPDS, 89-2838) depicting the (111), (200), (220), (311) and

(222) diffraction peaks of copper at 2θ = 43.2º, 50.4º, 74.1º, 89.9º and 95.1º,

respectively. A very intense Si(400) peak was also found at 2θ = 69.1º (JCPDS, 35-

1241), which corresponded to the substrate.

Figure 4.1. XRD spectrum of the Cu film as-deposited by OAD on a silicon substrate.

The microstructure of the Cu layers with different porosities prepared by

either normal or oblique evaporation (i.e., OAD) on silicon substrates was

characterized by SEM. Figures 4.2a and 4.2b show the cross-section and plan-view

images, respectively, of the rather compact layer prepared at α = 0º and reveal that it

was formed by perpendicular thin and close-packed nanocolumns as expected for

metal thin films deposited under the conditions of the zone I of the zone structure

model [42]. This first compact Cu layer constituted the inner layer of the copper

electrode and provided a very good electrical conductivity (6 – 8 Ω) as well as a high

mechanical stability to the electrocatalytic cell. Nevertheless, as observed in the top

view of Figure 4.2b, this film presents some diffusion surfaces at the interfacial

0 10 20 30 40 50 60 70 80 90 100

Inte

nsi

ty /

a.u

.

2θ / º

Cu

Si

40 50 60 70 80 90 100

Inte

nsi

ty /

a.u

.

2θ / º

(200)

(220)

(222)(311)

(111)

Chapter 4

170

boundaries of Cu columns which may allow the migration of potassium ions through

itself to the gas-exposed Cu catalyst surface.

Figure 4.2. Cross-sectional and top view SEM images of the Cu films as-prepared on silicon

substrate at a normal geometry (a and b, respectively) and at an angle of deposition of 80 º (c

and d, respectively). These layers constituted the inner and outer Cu layers at the

electrochemical catalyst, respectively.

On the other hand, Figures 4.2c and 4.2d show the microstructure of the

copper layer prepared by OAD at α = 80º. A thickness of around 0.8 µm and a

microstructure consisting of tilted and separated nanocolumns with an inclination β =

60º can be deduced from the cross section image. The open porosity of this layer,

which was in practice extended from the interface with the dense copper underlayer up

to the external surface of the catalyst, provided a free access and intimate contact with

the reaction gases. Although not all columns presented exactly the same height, a quite

uniform nanocolumn cross-sectional diameter of around 120 nm and a column density

of the order of 109 columns per cm

2 can be also deduced from these images. Similar

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

171

microstructures obtained by OAD have been reported for a large variety of materials

(see, for example, previous works on Co [43], Pt-Ni [31], TiO2 [28, 29, 39] or Ta2O5

[29, 40] films).

Hence, the OAD technique showed its capability to deposit metallic films with

a suitable porous microstructure for catalytic and electrocatalytic purposes. In

particular, an electrode film was herein developed consisting of a copper compact-

porous bilayer (which will be henceforth treated as a single film) that was utilized as

an electrochemical catalyst in the POM reaction.

Prior to the EPOC experiments, the Cu catalyst film was also

electrochemically characterized. Figure 4.3 shows the current (I) variation with the

applied potential (VWR) under the studied POM reaction conditions at a quite slow

scan rate of 5 mV/s.

Figure 4.3. Current (I) variation vs. the applied potential (VWR) during the cyclic voltammetry

between +1 and -1 V (scan rate = 5 mV/s). CH3OH/O2 = 4.4 %/0.3 %, 320 ºC.

This voltammogram was found to be very reproducible from the second cycle.

Other important feature is the appearance of clear cathodic and anodic peaks in the

studied potential range, which can be linked to the formation and decomposition of

-9

-7

-5

-3

-1

1

3

5

7

-1.2 -0.8 -0.4 0 0.4 0.8 1.2

I / A

x 1

0-5

VWR / V

Chapter 4

172

promoter-derived surface compounds, respectively. The area of the two peaks from

their respective baselines showed to be quite similar, with a difference of near 10 %.

Moreover, this similarity in both cathodic and anodic operation was even higher by

considering the positive and negative I vs time curves obtained during the EPOC

experiments as will be observed below. This means that a positive potential of 1 V led

to the decomposition of practically all the promoter compounds previously formed on

the catalyst surface during the cathodic scan. One can find in literature a wide range of

positive unpromoted potentials employed in alkali-based electropromoted systems,

from +0.1 V [21] or +0.7 V [38] up to +3 V [11] or +4 V [24]. Hence, although in the

previous chapters the reference (un-promoted) state was set at VWR = +2 V, in the

present study a reference potential of +1 V was selected, since this was the lowest

possible value to avoid the deterioration of the catalyst film and/or the solid

electrolyte, while providing a completely reversible electrocatalytic behavior (Figure

4.3).

4.3.2. Electrochemical promotion experiments

The Cu/K-βAl2O3/Au solid electrolyte cell was tested in the methanol partial

oxidation reaction with a feed composition of CH3OH/O2 = 4.4 %/0.3 % (Ar balance)

under electrochemical promotion conditions.

Figure 4.4 shows the dynamic response of the reaction rate of the detected

products, i.e., hydrogen, carbon dioxide and methyl formate, as well as the current (I)

vs. time curves obtained upon the application of different catalyst potentials (VWR)

between +1 and -1 V at a temperature of 320 ºC. Each polarization step was performed

for 1 hour to reach a steady-state rate value. Similarly to the previous chapters and to

other EPOC studies based on K+-conductor materials [11, 15, 19, 44, 45], the

application of a positive potential (VWR = +1 V) at the beginning of the experiment led

to the removal of all the potassium promoter possibly located on the surface of the

catalyst/working electrode. Hence, this potential was defined as the reference state and

its imposition before each different polarization allowed to start from the unpromoted

state in all cases and to check the reversibility and reproducibility of the promotional

effect.

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

173

Figure 4.4. Response of H2, CO2 and HCOOCH3 production rates (a) and obtained electric

current (I) (b) vs. time to step changes in the applied catalyst potential (VWR). CH3OH/O2 = 4.4

%/0.3 %, 320 ºC.

In Figure 4.4a, it can be observed that, under unpromoted conditions (VWR =

+1 V), the Cu catalyst film already showed certain catalytic activity for the production

of hydrogen and carbon dioxide, as well as also some water (not depicted), from the

methanol partial oxidation reaction. CO2 was the main product obtained in this

reference (unpromoted) state. Neither carbon monoxide nor formaldehyde were

detected. However, an appreciable amount of methyl formate (HCOOCH3) was

0

2

4

6

8

10

-1.5

-1

-0.5

0

0.5

1

1.5

r /

mo

l s-1

x 1

0-8

VW

R/

V

VWR

rH2rCO2

rHCOOCH3

-4

-3

-2

-1

0

1

2

-1.5

-1

-0.5

0

0.5

1

0 100 200 300 400 500 600I /

A x

10

-4

VW

R/

V

Time / min

VWR I

a)

b)

Chapter 4

174

produced, as in chapter 3, although here the HCOOCH3 selectivity was appreciably

lower. As previously mentioned, this compound is a very interesting by-product with

many industrial applications; for example, as a precursor in the synthesis of formic

acid, formamide and dimethylformamide. The decrease in the applied potential to VWR

= +0.5 V caused only a slight increase in H2 and CO2 production rates. This potential

was close to the open circuit value ( ), which oscillated in the range of 0.3-0.5 V

throughout the entire chapter (under different reaction temperatures). Hence, this first

decrease in the applied potential did not constitute a clear negative overpotential, thus

not a significant transfer of potassium species.

However, under the application of VWR = 0 V, the reaction rate of all products

(H2, CO2 and HCOOCH3) sharply increased (electrochemically promoted state), being

H2 the main reaction product (9.27 x 10-8

mol H2 s-1

). It should be noted that, in this

case, the applied potential was already lower than the open circuit potential value,

leading to the observed negative currents (Figure 4.4b), i.e., under these conditions, K+

ions were electrochemically pumped from the K-βAl2O3 pellet to the Cu catalyst film.

An increase in the alkali concentration on the catalyst surface upon decreasing the

applied potential has been previously demonstrated, for instance, by Prof. Lambert and

co-workers by means of in-situ XP spectroscopy studies on Cu [21, 22, 38], Pt [46]

and Rh [47]. Hence, the observed modification of the Cu catalytic activity with the

applied potential can be attributed to a decrease in the catalyst work function as a

consequence of the migration of electropositive ions and to the subsequent

modification of the chemisorptive properties of the Cu during the reaction [38].

Regarding the reaction mechanism, it is widely accepted that the first reaction

step in the methanol oxidation processes is the formation of methoxy species from

methanol (reactions 4.1 and 4.2), which is favoured by the presence of surface oxygen

(reactions 4.3 and 4.4) through a dissociative adsorption mechanism [34-36].

CH3OH → CH3OH(a) (4.1)

CH3OH(a) → CH3O(a) + H(a) (4.2)

O2 → 2O(a) (4.3)

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

175

CH3OH + O(a) → CH3O(a) + OH(a) (4.4)

Then, several consecutive reactions may take place leading to the production

of CO2 (reactions 4.5-4.7), H2 (reaction 4.8) and H2O (reaction 4.9) [36].

CH3O(a) + O(a) → H2CO(a) + OH(a) (4.5)

H2CO(a) + O(a) → HCOO(a) + H(a) (4.6)

HCOO(a) → H(a) + CO2 (4.7)

2H(a) → H2 (4.8)

H(a) + OH(a) → H2O (4.9)

On the other hand, the formation of methyl formate can be explained through

different reaction pathways and intermediate species reported elsewhere [34, 37, 48,

49] for the methanol oxidation over Cu catalysts. For instance, the following reaction

mechanism based on the occurrence of methoxy species could be proposed [49]:

CH3O(a) + O(a) → HCOO(a) + H2 (4.10)

CH3O(a) + HCOO(a) → HCOOCH3(a) + O(a) (4.11)

HCOOCH3(a) → HCOOCH3 (4.12)

According to the theory of the electrochemical promotion [41], under

electropromoted conditions (VWR ≤ 0 V in this work) the back-spillover of K+ ions

from the solid electrolyte onto the catalyst/working electrode would strengthen the

chemical bond between copper and electron-acceptor adsorbants, while the bonding

with the electron-donor ones should be weaken. Under POM conditions, these

molecules have been previously identified as O2 and CH3OH, respectively. Hence, the

enhancement of the Cu catalytic activity observed under negative polarization (Figure

4.4a) is in good agreement with POM studies [35] where a positive order of the

reaction rate was found with respect to the oxygen partial pressure (for low values, as

in this case). Likewise, an increase in the methanol reaction rate with O2 concentration

was found in previous EPOC studies with different electrochemical catalysts [50, 51],

and also in Figure 1.7 (chapter 1) until certain oxygen concentration was exceeded. A

Chapter 4

176

beneficial effect of alkali addition upon cathodic polarization was also observed over

Cu catalysts in other studies on NO reduction by CO [21, 22] and, very recently, on

CO2 hydrogenation under certain reaction conditions [24]. In the present study, the

adsorbed oxygen would play an important role by favouring the O-H bond rupture in

reaction 4.4. Nevertheless, the rate-limiting step in this kind of processes is usually

considered the methoxy decomposition (reactions 4.5 and 4.10) [35, 36]. The

electrophilic EPOC effect could have also favoured the C-H bond cleavage in the

methanol and intermediates decomposition reactions [50, 51].

Figure 4.4a also shows that all the products reaction rates (H2, CO2 and

HCOOCH3) were restored to their initial values upon every positive polarization at

VWR = +1 V. This demonstrates the complete reversibility and reproducibility of the

electropromoted effect, as well as the high stability of the Cu electrode along the

whole experiment. Moreover, optimal promotional conditions were found at VWR = 0

V. A further decrease of the applied potential down to -0.5 and -1 V led again to an

electro-promotional effect. In this case, a less pronounced increase of all the reaction

rates with respect to those found under unpromoted conditions occurred, which can be

attributed to a detrimental effect of an excess of alkali promoter and the concomitant

formation of alkali-derived surface compounds that may partially block catalytic

active sites. This is a common effect observed both in EPOC [15, 19, 44, 45] and in

conventional alkali chemical promotion studies [52, 53].

Regarding the possible oxidation of copper during this experiment, the in-

plane surface electrical resistance was continuously measured leading to values lower

than 10 Ω throughout the entire study. As it will be later discussed, these low

resistance values would indicate that the catalyst/working electrode mainly remained

on its metallic state during the reaction. It should also be mentioned that the methanol

conversion obtained in this work (not depicted) did not exceed the 5 %, which was

probably influenced by the characteristics of the catalyst, in film form, and thus a still

low geometric area. Nevertheless, it is important to stress that both catalytic activity

and selectivity of the Cu film in the POM reaction can be in-situ improved under

polarization conditions and tuned depending on the applied potential (via

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

177

electrochemical promotion). This feature of the EPOC phenomenon has been shown in

the previous chapters with alkali-promoted noble metals catalysts. In the present

study, the Cu-catalyzed H2 production was enhanced up to three times at 320 ºC, the

H2 selectivity being increased from 27.4 % (unpromoted conditions, VWR = +1 V) to

39.1 % (optimum electropromoted conditions, VWR = 0 V). These values were

calculated as H2,out CH3OH,in CH3OH,out , where Fi,in and Fi,out are the

molar flow rate of the i species at the inlet and at the outlet of the reactor, respectively.

Moreover, from the integration of the current vs. time curves (Figure 4.4b), the

transferred amount of promoter, and hence the promoter coverage (potassium to

copper ratio, θK+) can be estimated at each potential through the following equations:

=

(4.13)

=

(4.14)

where n is +1 in this case, F is the Faraday constant (96485 C), t is the polarization

time and N is the total amount of Cu deposited (3.4 x 10-5

mol Cu). θK+ was estimated

in this way due to the difficulty of calculating the Cu active sites in this

electrocatalytic system. The potassium coverage varied from 1 x 10-3

(at VWR = +0.5

V) to 8.5 x 10-3

(at VWR = -1 V). Hence, as in the previous chapters, the EPOC allowed

to determine the optimum promoter coverage (θK+,opt = 6.3 x 10-3

, obtained at VWR = 0

V) which led to the maximum reaction rate (9.27 x 10-8

mol H2 s-1

).

The influence of the reaction temperature on the copper catalytic activity was

studied at a fixed feed composition (CH3OH/O2 = 4.4 %/0.3 %) and different applied

potentials. Figure 4.5 shows the variation of the steady-state reaction rates vs. the

applied catalyst potential at different reaction temperatures between 280 and 360 ºC.

In order to evaluate the magnitude of the electropromotional effect, the maximum rate

enhancement ratios (ρmax) for the optimum applied potential are also included in the

figure. This parameter was calculated, as in previous chapters, as follows:

=

(4.15)

Chapter 4

178

where r and r0 are the promoted (VWR < +1 V) and unpromoted (VWR = +1 V) catalytic

reaction rates.

Figure 4.5. Steady-state H2 (a), CO2 (b) and HCOOCH3 (c) production rates vs. the applied

potential (VWR) as well as the corresponding maximum production rates enhancement ratios

(ρmax) at different reaction temperatures. CH3OH/O2 = 4.4 %/0.3 %.

0

3

6

9

12

15

r H/

mo

l s-1

x 1

0-8280 ºC

320 ºC

360 ºC

2

ρH ,max = 1.582

ρH ,max = 2.632

ρH ,max = 2.242

2

4

6

8

10

12

r CO

/ m

ol s

-1x

10-8

280 ºC

320 ºC

360 ºC

2

ρCO ,max = 1.582

ρCO ,max = 1.582

ρCO ,max = 1.552

0

0.5

1

1.5

2

-1.2 -0.8 -0.4 0 0.4 0.8 1.2

r HC

OO

CH

/ m

ol s

-1x

10-8

VWR / V

280 ºC

320 ºC

360 ºC

3

ρHCOOCH ,max = 1.913

ρHCOOCH ,max = 2.703

ρHCOOCH ,max = 1.413

a)

b)

c)

Increasing amount of transferred K+

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

179

Starting from the reference unpromoted state (VWR = +1 V) and according to

the previous considerations, the decrease in the applied potential caused the

electrochemical pumping of K+ ions from the solid electrolyte to the Cu catalyst film.

As a result, at the three studied temperatures, all the products reaction rates (H2, CO2,

HCOOCH3) increased because of the potassium promotional effect, evidencing an

electrophilic EPOC behavior (Δr > 0, ρ > 1, upon negative polarization) [41]. It should

be noted that at every reaction temperature, the same operational procedure was

carried out than in Figure 4.4, i.e., a potential of +1 V was applied at the end of every

experiment and before each different applied potential. In this way, both the

reversibility and the reproducibility of the electropromotional effect were verified at

all the reaction temperatures, as well as the stability of the Cu catalytic activity

throughout the entire study.

Under unpromoted conditions (VWR = +1 V), the Cu catalytic activity

increased with the reaction temperature in good agreement with the numerous POM

studies over conventional Cu supported catalysts (e.g., Cu/ZnO [35], Cu/ZnO/Al2O3

[37, 48] or Cu/ZrO2 [33]). As an exception, a decrease in the methyl formate

production rate was found at the highest temperature (360 ºC), which can be attributed

to its further oxidation to CO2 as reported in other works with Cu-based catalysts [37,

48, 49]. Meanwhile, under optimum electropromoted conditions (VWR = 0 V), an

increase in both the H2 and HCOOCH3 production rate enhancement ratios (ρ) was

found when increasing the temperature from 280 to 320 ºC, due to an increase in both

the copper catalytic activity and the solid electrolyte ionic conductivity at fixed

potential. The less pronounced EPOC effect found at 360 ºC should be likely related

with the higher unpromoted reaction rate (r0). In line with the results shown in Figure

4.4, a possible poisoning effect by potassium-derived surface compounds could also

take place under the application of the lowest applied potentials (VWR = -0.5 and -1 V).

The POM results obtained at different temperatures showed that the catalytic

performance of this novel Cu film prepared by OAD can be enhanced in a controlled

and reversible way via alkaline electrochemical promotion. Under optimal reaction

conditions (320 ºC, VWR = 0 V), the hydrogen and methyl formate production rates

Chapter 4

180

increased by 2.63 and 2.70 times, respectively. It is clear that the catalytic activity of

this Cu catalyst is far below that of the Pt catalysts. However, it must be stressed that

the maximum enhancement of the H2 production rate obtained here ( )

was higher, for instance, than that previously obtained in chapter 2

with an electrochemical catalyst based on platinum nanoparticles (0.014 mg

Pt/cm2) and similar methanol partial oxidation conditions (320 ºC and O2/CH3OH ratio

= 0.08). These results show that, under the explored reaction conditions, a non-noble

metal catalyst such as Cu can be electrochemically activated in a similar or even more

pronounced way than a noble metal catalyst such as Pt. In this way, one can suggest

that EPOC phenomena could be used to increase the catalytic activity of a non-noble

metal up to the typically higher activity of a noble catalyst, for example, in H2

production processes.

4.3.3. Post-reaction characterization of the catalyst film

It should be remarked that, during the conditioning of the catalyst prior to the

post-reaction characterization (320 ºC, POM conditions, VWR = -0.5 V), the obtained

reaction rates were very similar to those obtained during the first experiment

performed at 320 ºC (Figure 4.4). This verifies the good reproducibility of Cu catalytic

activity even after heating up to 360 ºC (for the study of the temperature influence).

The XRD spectrum of the used catalyst film after the EPOC experiments is

shown in Figure 4.6. This diffractogram shows again the five diffraction peaks

corresponding to the typical face-centered cubic structure (JCPDS, 89-2838) of

metallic copper. An average Cu particle size (d) of 43 nm can be estimated from the

width of the main diffraction peak, Cu(111), by means of the Scherrer formula. Peaks

associated with the K-βAl2O3 pellet (JCPDS, 21-618) can also be observed, but no

diffraction peaks related to potassium compounds like carbonates or bicarbonates were

found. Neither peaks associated with copper oxides like CuO nor Cu2O were detected.

However, the minority presence of these compounds cannot be ruled out because

some overlapping may take place with the peaks of alumina, for instance, in the case

of the CuO peak at 2θ = 35.4º (JCPDS, 80-1917).

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

181

Figure 4.6. XRD diffractogram of the electrochemical catalyst after the EPOC experiments

(320 ºC, CH3OH/O2 = 4.4 %/0.3 %, VWR = -0.5 V for 1 h).

Figure 4.7 shows the SEM micrograph of a selected region of the catalyst-

electrode where highly grown surface crystals were observed. Its corresponding

elemental mapping and spectra by EDX are also reported.

In Figures 4.7b, large concentrations of potassium (in blue) were found on

certain areas of the Cu surface (in green). This seems to indicate that some potassium-

derived surface compounds were formed on the Cu surface during the EPOC

experiments and might block Cu active sites causing the decrease in the reaction rates

observed in Figures 4.4 and 4.5 under the application of high negative potentials. As

suggested in chapter 1, the most feasible K-derived surface compounds under our

working conditions would be potassium oxides and/or carbonates formed by reaction

with O2, H2O and carbon-containing molecules in the reaction mixture.

0 10 20 30 40 50 60 70 80 90 100

Inte

nsi

ty /

a.u

.

2θ / º

(111)

Cu

K-βAl2O3

(200)

(220)

(222)(311)

Chapter 4

182

Figure 4.7. Top view SEM image of a selected area of the electrochemical catalyst (a) after the

EPOC experiments (320 ºC, CH3OH/O2 = 4.4 %/0.3 %, VWR = -0.5 V for 1 h), along with the

corresponding elemental mapping (b1 and b2) of Cu (green), K (blue) and N (red) and the

EDX spectra from different regions (c1 and c2).

These potassium oxide or carbonate compounds were previously found over

other metal catalysts such as Rh [54], Pt [45] or Ag [44], and may be obtained, for

instance, through the following electrocatalytic reactions:

2K+ + ½O2 + 2e

- → K2O (4.16)

2K+ + ½O2 + CO2 + 2e

- → K2CO3 (4.17)

Although the potassium distribution along the catalyst thickness was not

studied, it was important to demonstrate that K+ ions were able to migrate through the

catalyst film and reach the outermost surface. The presence of potassium compounds

was also revealed by EDX analysis taken in different areas of the micrograph. Figure

4.7c(2) shows that the K and O signals were much higher in the spectra taken in the

agglomerates than those from the flat region (Figure 4.7c(1)). It is also worth noting

0 1 2 3 4 5 6 7 8 9 10

Inte

nsi

ty /

a.u

.

Energy / keV

0 1 2 3 4 5 6 7 8 9 10

Inte

nsi

ty /

a.u

.

Energy / keV

3 µma)

b1) b2)

c2)c1)

Cu

OK

Cu

K

Cu

O

K

Cu

K

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

183

that nitrogen (red points) was found uniformly distributed on the film surface. This

elemental nitrogen can only proceed from the fed air stream (source of the oxygen

reactant), and its presence on the catalyst surface needed to be confirmed by XPS

analysis.

Figure 4.8 shows all the obtained XPS spectra. In this way, the chemical state

and composition of the outer surface layers of the Cu catalyst were studied.

Figure 4.8. XPS spectra of the electrochemical catalyst after the EPOC experiments (320 ºC,

CH3OH/O2 = 4.4 %/0.3 %, VWR = -0.5 V for 1 h). Spectra taken at different binding energy

regions to show the presence of the different elements present on the surface of the sample (a).

The corresponding surface atomic composition expressed in percentages (b).

a)

b)

0

10

20

30

40

50

O C K Al N Cu Na

At

/ %

945 940 935 930 925

Counts

/ a

.u.

Binding energy (eV)

Cu 2p3/2

540 535 530 525

Binding energy (eV)

O 1s

405 400 395 390

Binding energy (eV)

N 1s

345 340 335 330

Counts

/ a

.u.

Binding energy (eV)

Cu LVV

300 295 290 285 280

Binding energy (eV)

C 1s + K 2p

85 80 75 70 65

Binding energy (eV)

Cu 3p + Al 2p

Chapter 4

184

The Cu 2p photoemission (BE = 932.4 eV) and Cu LVV Auger (kinetic energy

915.14 eV) spectra obtained (Figure 4.8a) can be used to determine the Auger

parameter of copper (AP = BE photoemission peak + kinetic energy Auger peak)

which presented a value of 1847.54 eV, typical of Cu+ [55] (neither Cu

0 nor Cu

2+).

This finding would agree with some studies where the Cu active state in the methanol

partial oxidation reaction is considered Cu+ [48]. However, it should be noted that the

XPS characterization is only sensitive to a surface thickness of a few nanometers and

that, therefore, the Cu2O observed at the outermost catalyst layer could also be

attributed, at least in part, to the exposure of the sample to the air atmosphere during

its transfer from the reaction cell to the characterization equipment. An oxidation

limited to the outermost surface layers would be consistent with the very low surface

electrical resistance (below 10 ohms) and the XRD analysis of the utilized catalyst,

which would remain mainly as metallic copper under the studied reaction conditions.

On the other hand, the N 1s peak observed at 398 eV demonstrated the

presence of some reduced nitrogen on the catalyst surface, probably derived from a

certain K-promoted interaction between nitrogen from air and produced hydrogen. It is

well known that ammonia synthesis is typically promoted by potassium [53].

Furthermore, the ammonia synthesis reaction has been promoted by H+ ions in

previous EPOC studies with Fe catalysts under both high pressure [2] and atmospheric

pressure [20]. The C 1s signal in the spectrum of the used sample also shows the

presence of surface carbon mostly in the form of graphitic or hydrogenated carbon.

This indicates that a certain coking could have taken place at the highest reaction

temperatures. However, at least part of the O, C and K amounts observed in this figure

can be associated to promoter derived species as previously mentioned. The

aluminium also found in this XPS spectrum can be attributed to some uncovered

regions of the K-βAl2O3 electrolyte which might have appeared due to small scratches

caused by the contact with the gold wires. A quantitative estimation of the elemental

concentration in the examined outer layer of the electrochemical catalyst can be

observed in Figure 4.8b.

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

185

4.4. CONCLUSIONS

The following conclusions can be drawn from this study:

- The Oblique Angle Deposition (OAD) technique is a straightforward method

to prepare a kind of novel Cu catalytic films with a tuneable microstructure.

Combining the physical vapour deposition at normal and oblique angles for the

fabrication of a bilayer film yielded a Cu catalyst film with both a high surface

electrical conductivity and a high porosity (around 50 %), very suitable for catalytic

and electrocatalytic applications.

- A non-noble catalyst was electrochemically activated in the partial oxidation

of methanol. The catalytic activity of the Cu electrode was enhanced by

electrochemically supplying of K+ ions in a controlled, reversible and reproducible

way. The observed promotional effect can be explained according to the rules of

chemical/electrochemical promotion.

- Under electrochemical promotion conditions, the formation of potassium-

derived surface compounds, likely potassium carbonates or oxides, was confirmed by

post-reaction EDX and XPS characterization.

- Adsorbed nitrogen functional groups seemed to be formed under the

potassium promotional effect due to the presence of N2 and H2 in the reaction

atmosphere. Some graphitic or hydrogenated carbon was also deposited during the

EPOC experiments.

4.5. REFERENCES

[1] P. Vernoux, L. Lizarraga, M.N. Tsampas, F.M. Sapountzi, A. De Lucas-Consuegra, J.L.

Valverde, S. Souentie, C.G. Vayenas, D. Tsiplakides, S. Balomenou, E.A. Baranova, Ionically

conducting ceramics as active catalyst supports, Chemical Reviews, 113 (2013) 8192-8260.

[2] C.G. Yiokari, G.E. Pitselis, D.G. Polydoros, A.D. Katsaounis, C.G. Vayenas, High-

pressure electrochemical promotion of ammonia synthesis over an industrial iron catalyst,

Journal of Physical Chemistry A, 104 (2000) 10600-10602.

Chapter 4

186

[3] D. Poulidi, M.E. Rivas, B. Zydorczak, Z. Wu, K. Li, I.S. Metcalfe, Electrochemical

promotion of a Pt catalyst supported on La 0.6Sr 0.4Co 0.2Fe 0.8O 3 - δ hollow fibre

membranes, Solid State Ionics, 225 (2012) 382-385.

[4] S. Balomenou, D. Tsiplakides, A. Katsaounis, S. Thiemann-Handler, B. Cramer, G. Foti, C.

Comninellis, C.G. Vayenas, Novel monolithic electrochemically promoted catalytic reactor for

environmentally important reactions, Applied Catalysis B: Environmental, 52 (2004) 181-196.

[5] S.P. Balomenou, D. Tsiplakides, C.G. Vayenas, S. Poulston, V. Houel, P. Collier, A.G.

Konstandopoulos, C. Agrafiotis, Electrochemical promotion in a monolith electrochemical

plate reactor applied to simulated and real automotive pollution control, Topics in Catalysis,

44 (2007) 481-486.

[6] S. Souentie, A. Hammad, S. Brosda, G. Foti, C.G. Vayenas, Electrochemical promotion of

NO reduction by C2H4 in 10% O2 using a monolithic electropromoted reactor with Rh/YSZ/Pt

elements, Journal of Applied Electrochemistry, 38 (2008) 1159-1170.

[7] E.I. Papaioannou, S. Souentie, A. Hammad, C.G. Vayenas, Electrochemical promotion of

the CO2 hydrogenation reaction using thin Rh, Pt and Cu films in a monolithic reactor at

atmospheric pressure, Catalysis Today, 146 (2009) 336-344.

[8] M. Marwood, C.G. Vayenas, Electrochemical promotion of a dispersed platinum catalyst,

Journal of Catalysis, 178 (1998) 429-440.

[9] A. Kambolis, L. Lizarraga, M.N. Tsampas, L. Burel, M. Rieu, J.P. Viricelle, P. Vernoux,

Electrochemical promotion of catalysis with highly dispersed Pt nanoparticles,

Electrochemistry Communications, 19 (2012) 5-8.

[10] V. Jiménez, C. Jiménez-Borja, P. Sánchez, A. Romero, E.I. Papaioannou, D. Theleritis, S.

Souentie, S. Brosda, J.L. Valverde, Electrochemical promotion of the co2 hydrogenation

reaction on composite ni or ru impregnated carbon nanofiber catalyst-electrodes deposited on

YSZ, Applied Catalysis B: Environmental, 107 (2011) 210-220.

[11] A. de Lucas-Consuegra, A. Princivalle, A. Caravaca, F. Dorado, A. Marouf, C. Guizard,

J.L. Valverde, P. Vernoux, Preparation and characterization of a low particle size Pt/C

catalyst electrode for the simultaneous electrochemical promotion of CO and C3H6 oxidation,

Applied Catalysis A: General, 365 (2009) 274-280.

[12] A. Lintanf, E. Djurado, P. Vernoux, Pt/YSZ electrochemical catalysts prepared by

electrostatic spray deposition for selective catalytic reduction of NO by C3H6, Solid State

Ionics, 178 (2008) 1998-2008.

[13] V. Roche, R. Revel, P. Vernoux, Electrochemical promotion of YSZ monolith honeycomb

for deep oxidation of methane, Catalysis Communications, 11 (2010) 1076-1080.

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

187

[14] A. De Lucas-Consuegra, A. Caravaca, P.J. Martínez, J.L. Endrino, F. Dorado, J.L.

Valverde, Development of a new electrochemical catalyst with an electrochemically assisted

regeneration ability for H2 production at low temperatures, Journal of Catalysis, 274 (2010)

251-258.

[15] A. De Lucas-Consuegra, A. Caravaca, J. González-Cobos, J.L. Valverde, F. Dorado,

Electrochemical activation of a non noble metal catalyst for the water-gas shift reaction,

Catalysis Communications, 15 (2011) 6-9.

[16] T.I. Politova, V.A. Sobyanin, V.D. Belyaev, Ethylene hydrogenation in electrochemical

cell with solid proton-conducting electrolyte, Reaction Kinetics & Catalysis Letters, 41 (1990)

321-326.

[17] I.V. Yentekakis, Y. Jiang, S. Neophytides, S. Bebelis, C.G. Vayenas, Catalysis,

electrocatalysis and electrochemical promotion of the steam reforming of methane over Ni film

and Ni-YSZ cermet anodes, Ionics, 1 (1995) 491-498.

[18] E. Ruiz, D. Cillero, P.J. Martínez, Á. Morales, G.S. Vicente, G. De Diego, J.M. Sánchez,

Bench-scale study of electrochemically assisted catalytic CO2 hydrogenation to hydrocarbon

fuels on Pt, Ni and Pd films deposited on YSZ, Journal of CO2 Utilization, 8 (2014) 1-20.

[19] G.E. Pitselis, P.D. Petrolekas, C.G. Vayenas, Electrochemical promotion of ammonia

decomposition over Fe catalyst films interfaced with K+- & H+- conductors, Ionics, 3 (1997)

110-116.

[20] M. Ouzounidou, A. Skodra, C. Kokkofitis, M. Stoukides, Catalytic and electrocatalytic

synthesis of NH3 in a H+ conducting cell by using an industrial Fe catalyst, Solid State Ionics,

178 (2007) 153-159.

[21] F.J. Williams, A. Palermo, M.S. Tikhov, R.M. Lambert, First Demonstration of in Situ

Electrochemical Control of a Base Metal Catalyst: Spectroscopic and Kinetic Study of the CO

+ NO Reaction over Na-Promoted Cu, Journal of Physical Chemistry B, 103 (1999) 9960-

9966.

[22] R.M. Lambert, F. Williams, A. Palermo, M.S. Tikhov, Modelling alkali promotion in

heterogeneous catalysis: in situ electrochemical control of catalytic reactions, Topics in

Catalysis, 13 (2000) 91-98.

[23] F.M. Sapountzi, M.N. Tsampas, C.G. Vayenas, Methanol reformate treatment in a PEM

fuel cell-reactor, Catalysis Today, 127 (2007) 295-303.

[24] E. Ruiz, D. Cillero, P.J. Martínez, Á. Morales, G.S. Vicente, G. De Diego, J.M. Sánchez,

Electrochemical synthesis of fuels by CO2 hydrogenation on Cu in a potassium ion conducting

membrane reactor at bench scale, Catalysis Today, 236 (2014) 108-120.

Chapter 4

188

[25] G. Karagiannakis, S. Zisekas, M. Stoukides, Hydrogenation of carbon dioxide on copper

in a H+ conducting membrane-reactor, Solid State Ionics, 162-163 (2003) 313-318.

[26] E. Mutoro, C. Koutsodontis, B. Luerssen, S. Brosda, C.G. Vayenas, J. Janek,

Electrochemical promotion of Pt(111)/YSZ(111) and Pt-FeOx/YSZ(111) thin catalyst films:

Electrocatalytic, catalytic and morphological studies, Applied Catalysis B: Environmental,

100 (2010) 328-337.

[27] M. Matthew, M.T. Hawkeye, M.J.B. Taschuk, Glancing Angle Deposition of Thin Films:

Engineering the Nanoscale, Wiley2014.

[28] M. Oliveros, L. González-García, V. Mugnaini, F. Yubero, N. Roques, J. Veciana, A.R.

González-Elipe, C. Rovira, Novel guests for porous columnar thin films: The switchable

perchlorinated trityl radical derivatives, Langmuir, 27 (2011) 5098-5106.

[29] J.R. Sánchez-Valencia, A. Borrás, A. Barranco, V.J. Rico, J.P. Espinos, A.R. González-

Elipe, Preillumination of TiO2 and Ta2O5 photoactive thin films as a tool to tailor the

synthesis of composite materials, Langmuir, 24 (2008) 9460-9469.

[30] Y. He, J. Fan, Y. Zhao, The role of differently distributed vanadium nanocatalyst in the

hydrogen storage of magnesium nanostructures, International Journal of Hydrogen Energy, 35

(2010) 4162-4170.

[31] N.N. Kariuki, W.J. Khudhayer, T. Karabacak, D.J. Myers, GLAD Pt-Ni alloy nanorods for

oxygen reduction reaction, ACS Catalysis, 3 (2013) 3123-3132.

[32] S.A. Francis, R.T. Tucker, M.J. Brett, S.H. Bergens, Structural and activity comparison of

self-limiting versus traditional Pt electro-depositions on nanopillar Ni films, Journal of Power

Sources, 222 (2013) 533-541.

[33] H. Chen, A. Yin, X. Guo, W.L. Dai, K.N. Fan, Sodium hydroxide-sodium oxalate-assisted

co-precipitation of highly active and stable Cu/ZrO2 catalyst in the partial oxidation of

methanol to hydrogen, Catalysis Letters, 131 (2009) 632-642.

[34] S.T. Yong, C.W. Ooi, S.P. Chai, X.S. Wu, Review of methanol reforming-Cu-based

catalysts, surface reaction mechanisms, and reaction schemes, International Journal of

Hydrogen Energy, 38 (2013) 9541-9552.

[35] L.A. Espinosa, R.M. Lago, M.A. Peña, J.L.G. Fierro, Mechanistic aspects of hydrogen

production by partial oxidation of methanol over Cu/ZnO catalysts, Topics in Catalysis, 22

(2003) 245-251.

[36] S. Sakong, A. Gross, Total oxidation of methanol on Cu(110): A density functional theory

study, Journal of Physical Chemistry A, 111 (2007) 8814-8822.

Electrochemical promotion of Cu nanocolumns in the partial oxidation of methanol

189

[37] R.M. Navarro, M.A. Peña, J.L.G. Fierro, Production of hydrogen by partial oxidation of

methanol over a Cu/ZnO/Al2O3 catalyst: Influence of the initial state of the catalyst on the

start-up behaviour of the reformer, Journal of Catalysis, 212 (2002) 112-118.

[38] F.J. Williams, A. Palermo, M.S. Tikhov, R.M. Lambert, The Origin of Electrochemical

Promotion in Heterogeneous Catalysis: Photoelectron Spectroscopy of Solid State

Electrochemical Cells, Journal of Physical Chemistry B, 104 (2000) 615-621.

[39] V. Rico, P. Romero, J.L. Hueso, J.P. Espinós, A.R. González-Elipe, Wetting angles and

photocatalytic activities of illuminated TiO2 thin films, Catalysis Today, 143 (2009) 347-354.

[40] V. Rico, A. Borrás, F. Yubero, J.P. Espinós, F. Frutos, A.R. González-Elipe, Wetting

angles on illuminated ta2o5 thin films with controlled nanostructure, Journal of Physical

Chemistry C, 113 (2009) 3775-3784.

[41] C.G. Vayenas, S. Bebelis, C. Pliangos, S. Brosda, D. Tsiplakides, Electrochemical

Activation of Catalysis: Promotion, Electrochemical Promotion and Metal-Support

Interactions, Kluwer Academic Publishers/Plenum Press, New York, 2001.

[42] J.A. Thornton, INFLUENCE OF APPARATUS GEOMETRY AND DEPOSITION

CONDITIONS ON THE STRUCTURE AND TOPOGRAPHY OF THICK SPUTTERED

COATINGS, J Vac Sci Technol, 11 (1974) 666-670.

[43] D.X. Ye, Y.P. Zhao, G.R. Yang, Y.G. Zhao, G.C. Wang, T.M. Lu, Manipulating the

column tilt angles of nanocolumnar films by glancing-angle deposition, Nanotechnology, 13

(2002) 615-618.

[44] A. Palermo, A. Husain, M.S. Tikhov, R.M. Lambert, Ag-catalysed epoxidation of propene

and ethene: An investigation using electrochemical promotion of the effects of alkali, NOx, and

chlorine, Journal of Catalysis, 207 (2002) 331-340.

[45] A. de Lucas-Consuegra, F. Dorado, J.L. Valverde, R. Karoum, P. Vernoux, Low-

temperature propene combustion over Pt/K-βAl2O3 electrochemical catalyst:

Characterization, catalytic activity measurements, and investigation of the NEMCA effect,

Journal of Catalysis, 251 (2007) 474-484.

[46] I.V. Yentekakis, A. Palermo, N.C. Filkin, M.S. Tikhov, R.M. Lambert, In situ

electrochemical promotion by sodium of the platinum-catalyzed reduction of NO by propene,

Journal of Physical Chemistry B, 101 (1997) 3759-3768.

[47] F.J. Williams, A. Palermo, M.S. Tikhov, R.M. Lambert, Electrochemical promotion by

sodium of the rhodium-catalyzed NO + CO reaction, Journal of Physical Chemistry B, 104

(2000) 11883-11890.

Chapter 4

190

[48] Y. Choi, H.G. Stenger, Fuel cell grade hydrogen from methanol on a commercial

Cu/ZnO/Al2O3 catalyst, Applied Catalysis B: Environmental, 38 (2002) 259-269.

[49] V.A. Matyshak, O.N. Sil'Chenkova, I.T. Ismailov, V.F. Tret'Yakov, Properties of surface

compounds in methanol conversion on copper-containing catalysts based on CeO2 according

to in situ IR-spectroscopic data, Kinetics and Catalysis, 50 (2009) 784-792.

[50] C.G. Vayenas, S. Neophytides, Non-faradaic electrochemical modification of catalytic

activity III. The case of methanol oxidation on Pt, Journal of Catalysis, 127 (1991) 645-664.

[51] J.K. Hong, I.H. Oh, S.A. Hong, W.Y. Lee, Electrochemical oxidation of methanol over a

silver electrode deposited on yttria-stabilized zirconia electrolyte, Journal of Catalysis, 163

(1996) 95-105.

[52] I.V. Yentekakis, M. Konsolakis, R.M. Lambert, N. MacLeod, L. Nalbantian,

Extraordinarily effective promotion by sodium in emission control catalysis: NO reduction by

propene over Na-promoted Pt/γ-al2O3, Applied Catalysis B: Environmental, 22 (1999) 123-

133.

[53] B. Lin, K. Wei, X. Ma, J. Lin, J. Ni, Study of potassium promoter effect for Ru/AC

catalysts for ammonia synthesis, Catalysis Science and Technology, 3 (2013) 1367-1374.

[54] A.J. Urquhart, J.M. Keel, F.J. Williams, R.M. Lambert, Electrochemical Promotion by

Potassium of Rhodium-Catalyzed Fischer-Tropsch Synthesis: XP Spectroscopy and Reaction

Studies, Journal of Physical Chemistry B, 107 (2003) 10591-10597.

[55] J.P. Espinós, J. Morales, A. Barranco, A. Caballero, J.P. Holgado, A.R. González-Elipe,

Interface effects for Cu, CuO, and Cu2O deposited on SiO2 and ZrO2. XPS determination of

the valence state of copper in Cu/SiO2 and Cu/ZrO2 catalysts, Journal of Physical Chemistry

B, 106 (2002) 6921-6929.

191

Chapter 5

ELECTROCHEMICAL PROMOTION OF Ni IN

METHANOL CONVERSION REACTIONS:

DIFFERENT APPLICATIONS OF EPOC ON A SINGLE

CATALYTIC SYSTEM

In this work, three possible applications of the phenomenon of electrochemical

promotion of catalysis (EPOC) are demonstrated for methanol conversion processes:

activation of the catalytic activity, modification of the catalytic selectivity and partial

oxidation of the catalyst. A Ni catalyst film prepared by cathodic arc deposition has

been electrochemically promoted by K+, upon negative polarization, in the methanol

decomposition (MD) and steam reforming (SRM) reactions, showing an electrophilic

EPOC behavior. On the other hand, under methanol partial oxidation conditions

(POM), the K+ ions caused the decrease in the catalytic selectivity toward H2 and CO

while favoured that toward CO2 and H2CO, due to an increase in the Ni oxidation

state by the alkali-induced O2 activation. All the potassium-derived effects have been

reversible between the negative and positive polarizations, which shows different

interesting possibilities of the EPOC phenomena in heterogeneous catalysis.

Methanoldecomposition

CHXO

Steam reformingof methanol

H2O CH3OH

Partial oxidation of methanol

O2 CH3OHA A AD D

ΔV < 0

Ni

K+ K+ K+

Au

e-

e-

ELECTOPROMOTEDSTATE

CH3OH CO + 2H2

CO(a) C(a) + O(a)

CH3OH CO + 2H2

CH3OH + H2O CO2 + 3H2

C(a) + H2O CO + H2

CH3OH CO + 2H2

CH3OH + 1/2O2 CO2 + 2H2

CH3OH + 3/2O2 CO2 + 2H2O

CH3OH + 1/2O2 H2CO + H2O

xNi + yO(a) NixOy

A D

A

D

Acceptor

Donor

Activation of thecatalyst

Partial oxidationof the catalyst

Modification of selectivity

Activation of thecatalyst

Modification of selectivity

Chapter 5

192

5.1. INTRODUCTION

Based on the results presented in the previous chapters, one can appreciate

some of the main possibilities offered by the phenomenon of electrochemical

promotion (EPOC), in particular in H2 production processes from methanol. One can

also asseverate that the difference between electrochemical and chemical promotion is

operational and not functional, as Prof. Vayenas and co-workers stated in a seminal

book about electrochemical activation of catalysis through promotion, electrochemical

promotion and metal-support interactions (MSI) [1]. There is a very rich literature on

the similarities and differences between these three processes [1-5]. All of them are

based on the addition of promoting species to the catalytic active surface, which

modify its chemisorptive properties and hence its activity and/or its selectivity.

However, in the case of the EPOC phenomenon, these promoter ions are

electrochemically pumped between the metallic catalyst and a solid electrolyte

material (which acts as catalyst support) by means of the external imposition of an

electrical current or potential.

Then, the electrochemical promotion presents several additional advantages

against the classical promotion [3, 6]. One of the greatest is related to the first EPOC

studies with yttria-stabilized zirconia (YSZ) electrolytes, and lies on its capability to

continuously supply promoters on the catalyst surface, which is of special interest in

the case of short-lived effective promoters such as Oδ-

[1]. The electrochemical

promotion can be also carried out in a controlled, reversible way and, very

interestingly, in the course of the catalytic reaction, which has been pointed out

throughout the entire manuscript (especially in chapter 3). Thus, it is possible to

optimize the promoter coverage on the catalyst surface at different reaction conditions,

which is of great practical relevance in non-stationary processes like automotive

catalysis. This fact also enables a fast and direct evaluation of the effect of a specific

promoter on a catalytic system. Furthermore, the EPOC phenomenon is not only

aimed to enhance the catalytic activity and selectivity (as in all the previous chapters),

but also may affect the oxidation state of the catalyst [7-11] and can be focused to

other applications such as the prevention of poisoning effects and catalyst regeneration

Electrochemical promotion of Ni in methanol conversion reactions

193

[7, 12]. A large number of examples on the different EPOC applications can be found

in exhaustive reviews [3, 5, 6].

In the case of the methanol conversion reactions, before this research started

up, only four studies [13-16] dealt with the application of the phenomenon of the

electrochemical promotion, specifically using YSZ (O2-

conductor) as solid electrolyte

and Pt [14, 15] or Ag [13, 16] catalyst films. The main catalytic reactions that took

place in these systems were the methanol decomposition to CO (reaction 5.1) and

H2CO (reactions 5.2) along with some methanation (reaction 5.3) [13], and the

methanol oxidation through reactions 5.4 and 5.5 [14-16], thus obtaining

formaldehyde as product of highest interest.

CH3OH → 2H2 + CO (5.1)

CH3OH → H2CO + H2 (5.2)

CH3OH + H2 → CH4 + H2O (5.3)

CH3OH + 1/2O2 → H2CO + H2O (5.4)

CH3OH + 3/2O2 → CO2 + 2H2O (5.5)

Herein, in chapters 1 and 2, the alkali electrochemical promotion was applied

on Pt-based catalysts focused on the production of H2 via methanol steam reforming

(SRM, reaction 5.6) and partial oxidation (POM, reaction 5.7), although formaldehyde

was also obtained through reaction (5.4).

CH3OH + H2O → 3H2 + CO2 (5.6)

CH3OH + 1/2O2 → 2H2 + CO2 (5.7)

CO was also produced in these chapters at large extent, which was not

surprising given the preference of Group VIII metals for the methanol decomposition

(MD, reaction 5.1). However, in chapters 3 and 4, where Au- and Cu-based catalysts

were employed, neither CO nor H2CO were detected. In these cases, besides methanol

partial oxidation (reaction 5.7), methyl formate production also took place, which is

typically postulated by some of the following reactions:

Chapter 5

194

CH3OH + 1/2O2 → 1/2HCOOCH3 + H2O (5.8)

CH3OH → 1/2HCOOCH3 + H2 (5.9)

In the present study, the EPOC phenomenon has been applied in this kind of

reactions by using Ni as the active phase, which is indeed commonly employed

particularly in the decomposition and steam reforming of methanol [17-21]. Reactions

5.1, 5.2, 5.6 and 5.7 are of special interest from an energetic and environmental point

of view, as promising on-board H2 production processes from a sustainable liquid

source such as CH3OH. However, important drawbacks are typically found in this

kind of reactions with Group 8-11 metals, and especially with non noble ones such as

Cu or Ni, which are related to the thermal sintering, the deactivation by coke

deposition and undesirable changes in the oxidation state of the catalyst [22, 23].

Thus, these metals are usually supported on certain oxide supports such as CeO2,

ZrO2, Al2O3, ZnO, SiO2 or TiO2 [22-24], which may act as structural and/or electronic

promoters to improve the catalytic activity, decrease the CO selectivity and increase

the durability of the catalyst. The decomposition, steam reforming and partial

oxidation of methanol, as well as the water-gas shift and methanol synthesis reactions,

were also found to be promoted by alkali dopants, typically added as oxides or

hydroxides [25-29]. In the present work, the modifications of the activity and the

selectivity of a Ni catalyst via electrochemical promotion with K+ ions have been

studied under different reaction atmospheres (MD, SRM and POM). This way, three

different possibilities of the EPOC effect with Ni catalyst have been evaluated:

electrochemical activation of catalytic activity, alteration of the catalyst oxidation state

and electrochemical modification of the catalytic selectivity.

5.2. EXPERIMENTAL

5.2.1. Preparation of the electrochemical catalyst

The electrochemical catalyst consisted of a continuous thin Ni film (geometric

area of 2.84 cm2), which also behaved as a working electrode (WE), deposited on a

19-mm-diameter, 1-mm-thick K-βAl2O3 (Ionotec) pellet. Au counter (CE) and

reference (RE) electrodes were prepared by calcination of an organometallic paste

Electrochemical promotion of Ni in methanol conversion reactions

195

(Gwent Electronic Materials). Then, the active Ni catalyst-working electrode was

deposited similarly to Pt in chapter 1, i.e., by the filter cathodic arc deposition (CAD)

technique [30]. This physical vapour deposition technique allowed the preparation of a

thin Ni film, 150 nm thick, with a high adhesion to the substrate and a good surface

electrical conductivity, which are necessary for the electrochemical promotion

(EPOC) studies. The final metal loading was 0.13 mg Ni cm-2

, and the obtained Ni/K-

βAl2O3/Au catalyst was placed into the single chamber solid electrolyte cell reactor.

This catalyst preparation method was carried out in collaboration with Dr.

José Luis Endrino from the Institute of Materials Science of Madrid (CSIC).

5.2.2. Characterization measurements

The Ni catalyst film was characterized by X-Ray Diffraction (XRD) analysis

using a Philips PW-1710 instrument with Cu Kα radiation (λ = 1.5404 Å) after

exposing to different reaction mixtures and applied potentials. To this end, the

following steps were carried out before each XRD analysis in order to keep the Ni

unpromoted/promoted states as unaltered as possible: 1) Every time the Ni catalyst

film was introduced into the reactor, a temperature programmed reduction (TPR) was

firstly performed under a 25 % H2 stream (Ar balance) from room temperature to 400

ºC (ramp of 5 ºC min-1

) in order to ensure the Ni fully reduced state. 2) A temperature

of 360 ºC was fixed and the catalyst was subjected for 1 hour to the reaction

atmosphere and the applied potential under study. 3) The reactor temperature was

decreased to 100 ºC by keeping constant the rest of reaction conditions. 4) The catalyst

was finally cooled down to room temperature under Ar atmosphere and open circuit

conditions. Then it was transferred to the characterization equipment under inert

conditions (N2 atmosphere).

5.2.3. Catalytic activity measurements

The catalytic activity measurements were performed in the experimental setup

described in section 1.2.3. Different methanol conversion reactions, i.e.,

decomposition (MD), partial oxidation (POM) and steam reforming (SRM) were

studied by carrying out experiments at atmospheric pressure with an overall gas flow

Chapter 5

196

rate of 6 NL h-1

, a maximum temperature of 360 ºC and feed compositions of CH3OH

= 4.4 % for MD reaction, CH3OH/O2 = 4.4 %/0.33 % for POM reaction and

CH3OH/H2O = 4.4 %/5.2 % for SRM reaction (Ar balance). The detected products

were: H2, CO2, CO, H2O and H2CO (the latter two mainly under POM conditions).

5.3. RESULTS AND DISCUSSION

5.3.1. Electrochemical activation of the catalyst

First, the activity of the Ni catalyst film was tested at 360 ºC in the methanol

decomposition reaction, i.e., by feeding the reactor with a 4.4 % CH3OH stream (Ar

balance). H2, CO, and trace amounts of CO2 were detected in the reactor outlet. Figure

5.1 shows the variation in the production rates vs. time obtained under such reaction

conditions and different applied catalyst potentials (VWR) from +2 to -2 V.

Figure 5.1. Response of H2, CO and CO2 production rates vs. time to step changes in the

applied catalyst potential (VWR) under methanol decomposition conditions: CH3OH = 4.4 %,

360 ºC.

-2.5

-1.5

-0.5

0.5

1.5

2.5

0

0.5

1

1.5

2

2.5

3

3.5

4

0 100 200 300 400 500

VW

R/

V

r /

mo

l s-1

x 1

0-7

Time / min

rCO

VWR

rH2

Electrochemical promotion of Ni in methanol conversion reactions

197

As discussed in previous chapters and typically stated at electrochemical

catalysts based on alkaline ionic conductors [6], such as K-βAl2O3, the application of a

positive current (I) or potential (VWR) led to the removal from the catalyst-working

electrode of the positive promoter ions (K+) that may have previously migrated

(thermally) or as a consequence of the foregoing polarizations. On the contrary, the

decrease in the applied current or potential below I < 0 or VWR < (open circuit

potential value), respectively, caused the migration of alkaline promoter ions from the

solid electrolyte to the catalyst surface (back-spillover). Hence, at the beginning and at

the end of each experiment, a catalyst potential of VWR = +2 V was imposed in order

to define a clean Ni catalyst film, free of promoter ions and thus a reference

(unpromoted) state. In this way, it was also possible to check the reversibility of the

electrochemical promotion effect. Under these unpromoted conditions (at +2 V and

still at +0.5 V) hydrogen and carbon monoxide were produced, through reaction 5.1, at

rates of around 7 x 10-8

mol H2 s-1

and 2 x 10-8

mol CO s-1

. No other reaction products

were detected. It can be observed, however, that these production rates were

progressively decreasing with time, and that the obtained H2 to CO ratio was far above

the stoichiometric value expected from reaction 5.1 (H2/CO = 2). This was likely due

to the deposition of carbonaceous species on the catalyst surface and the concomitant

poisoning effect. In fact, Ni is among the most susceptible metals to this kind of

deactivation derived from the following CO dissociation reaction [31, 32]:

CO(a) → C(a) + O(a) (5.10)

Then, a decrease in the applied potential to VWR = 0 caused a very sharp

increase in both H2 and CO production rates (electropromoted state), up to around 3.5

x 10-7

mol H2 s-1

and 1.5 x 10-7

mol CO s-1

, respectively, reaching a H2/CO ratio closer

to 2, i.e., the stoichiometric value from reaction 5.1. As in the previous EPOC study

on methanol decomposition with a Ag catalyst [13], this modification in the catalytic

activity upon decreasing the applied potential can be explained on the basis of a

decrease in the catalyst work function which led to strong modification in the catalyst

chemisorption properties. In the mentioned work, this effect was caused by the

removal of electronegative O2-

ions from the catalyst surface by means of an yttria-

Chapter 5

198

stabilized zirconia (YSZ) solid electrolyte. In the present case, electropositive K+ ions

were electrochemically pumped from the K-βAl2O3 to the Ni catalyst at this applied

potential, according to the obtained negative currents (not shown here). Consequently,

in both cases, the lower catalyst work function would strengthen the chemical bond

between the metal and the electron acceptor adsorbate. This can be explained through

two approaches: the strengthening of the direct (electrostatic) attractive interactions

between the electron acceptor adsorbant and the electric field created by the

coadsorbed electropositive promoter, and the increase in the Fermi level with the

concomitant favouring of back-donation of binding electrons from the metal [33]. In

this point, it is important to analyze the role of the different adsorbates and

intermediate present in the reaction under study (reaction 5.1). For Group 10 metals

such as Ni, it is generally accepted that the MD reaction involves the adsorption of

CH3OH on the catalyst surface (reaction 5.11) and its successive dehydrogenation

(reactions 5.12-5.14), yielding CO (reaction 5.15) and H2 (reaction 5.16) [18].

CH3OH → CH3OH(a) (5.11)

CH3OH(a) → CH3O(a) + H(a) (5.12)

CH3O(a) → H2CO(a) + H(a) (5.13)

H2CO(a) → CO(a) + 2H(a) (5.14)

CO(a) → CO (5.15)

2H(a) → H2 (5.16)

Hydrogen abstraction from methoxy (reaction 5.13) is typically considered the

rate-limiting step [19, 34]. Hence, one could expect the methoxy group (CH3O) and

the successive intermediate species (CHxO, x = 0-2) to retain an electron acceptor

character versus the intramolecularly bonded H [8, 13]. This would explain the drastic

improvement in the methanol dehydrogenation rate, yielding H2 and CO, upon the

decrease in the applied potential and the consequent increase in the potassium

coverage. From Figure 5.1 it is also noteworthy that more negative potentials led to a

strong decrease in the Ni catalytic activity to H2 and CO production values close to

Electrochemical promotion of Ni in methanol conversion reactions

199

those initially obtained. This can be attributed to an increasing activation of the C-O

bond cleavage by K+ ions [14, 35], thus favouring the carbon deposition (reaction

5.10). Indeed, under negative applied potentials, a much abrupt slope can be observed

as compared to positive polarizations, which verifies the stronger deactivation process.

A similar experiment was carried out by changing the reaction atmosphere to

steam reforming conditions. Figure 5.2 shows the results obtained at 360 ºC with a

feed composition of CH3OH/H2O = 4.4 %/5.2 % upon the same potentiostatic

transitions used in Figure 5.1. As in the former case, the application of positive

potentials led to establish an unpromoted Ni catalyst state, i.e. with no potassium

species on the surface. Then, the catalyst was activated under the application of VWR =

0 V and lower potentials, due to the back-spillover of K+ promoter ions to the catalyst

surface and the concomitant decrease in the Ni work function as discussed above.

Figure 5.2. Response of H2, CO and CO2 production rates vs. time to step changes in the

applied catalyst potential (VWR) under methanol steam reforming conditions: CH3OH/H2O =

4.4 %/5.2 %, 360 ºC.

-2.5

-1.5

-0.5

0.5

1.5

2.5

0

1

2

3

4

5

6

7

0 100 200 300 400 500

VW

R/

V

r /

mo

l s-1

x 10

-7

Time / min

rCO

VWR

rH2

rCO2

Chapter 5

200

Under SRM conditions, H2 and CO were still the main obtained compounds

(likely from reaction 5.1) but an appreciable amount of CO2 was also produced

(reaction 5.6). The steam reforming of methanol can be globally considered as the sum

of the methanol decomposition and the water-gas shift reactions [23]. However,

unalloyed Group VIII metals in general, and Ni in particular, are especially active for

the methanol decomposition to CO rather than the steam reforming reaction [17, 20,

21]. A higher methanol conversion can be appreciated from this figure in comparison

with the methanol decomposition results (Figure 5.1), which is in good agreement

with other comparative studies [36]. This can be attributed to the key role of H2O in

assisting the decomposition of methoxy groups and subsequent intermediate species

involved in the MD reaction mechanism previously described [37]. Hence, one could

consider water as an electron acceptor adsorbate (vs. methanol) as also deduced from

other EPOC studies [12, 38]. It is then expected that a K+-derived decrease in the

catalyst work function upon negative polarizations favoured the water activation and

thus the production rate of H2, CO and CO2, in agreement with the SRM results

obtained with Pt catalysts in chapters 1 and 2.

On the other hand, although some coke deposition likely took place under

SRM conditions, a much lower catalyst deactivation was observed as compared to the

MD results. In fact, it is well known the beneficial effect of water on this kind of

catalytic systems through the performance of carbon gasification (reaction 5.17) [36,

39].

C(a) + H2O → CO + H2 (5.17)

The electrochemical activation of H2O by K+ ions would also favour this

reaction under negative applied potentials, as can be observed in Figure 5.2. The

deviation in the carbon atom balance obtained from the gaseous products decreased

approximately from 18 % at +2 V (unpromoted conditions) to 2 % at -0.5 V (promoted

conditions), which demonstrates that the H2O chemisorption enhancement via EPOC

attenuated the deactivation process.

Electrochemical promotion of Ni in methanol conversion reactions

201

The variation of the Ni catalytic activity with the applied potential was also

studied under partial oxidation conditions (CH3OH/O2 = 4.4 %/0.33 %) at the same

temperature (360 ºC), as shown in Figure 5.3. In this case and depending on the

catalyst potential, not only H2, CO and CO2 were produced but also H2CO and H2O

from reactions 5.4 and 5.5, respectively.

Figure 5.3. Response of H2, CO, CO2 and H2CO production rates vs. time to step changes in

the applied catalyst potential (VWR) under methanol partial oxidation conditions: CH3OH/O2 =

4.4 %/0.33 %, 360 ºC.

Under unpromoted conditions, H2 and CO were again the main reaction

products with a H2/CO ratio of around 2, which confirmed the excellent

dehydrogenation activity of nickel (reaction 5. 1) and the fairly complete absence of

coke deposition under this reaction atmosphere (no deactivation was observed during

the application of +2 V). This latter fact explains the much higher catalytic activity of

the unpromoted Ni catalyst as compared to the previous experiments, which was also

likely influenced by the oxygen participation in the stepwise hydrogen abstraction

-2.5

-1.5

-0.5

0.5

1.5

2.5

0

4

8

12

16

0 100 200 300 400 500

VW

R/

V

r /

mo

l s-1

x 10

-7

Time / min

rCO

VWR

rH2

rCO2

rH CO2

Chapter 5

202

from methanol (reaction 5.12) and methoxy species (reaction 5.13) [40, 41]. However,

the most interesting feature of this figure is the extremely sharp decrease in the Ni

catalytic activity observed upon decreasing the applied potential from VWR = +0.5 V,

which led to almost negligible H2 and CO production from reactions 5.1 and 5.7, while

some CO2, H2CO and H2O were obtained through reactions 5.4 and 5.5. One could

state that the back-spillover of K+ ions presented a strong negative effect in the

methanol partial oxidation for H2 production, as opposed to the enhancement effect

observed in the previous alkaline electrochemical promotion studies performed under

POM conditions with Pt-, Au- and Cu-based catalysts. Hence, in contrast with the two

previous experiments under MD and SRM conditions, a detrimental effect of negative

polarization was clearly observed for H2 and CO production. It could be attributed to

the presence of O2 in the gas phase which may lead to some modification in the Ni

catalyst oxidation state as will be discussed below. Anyway, it is clear that the

electrically-induced promotional (or poisoning) effects obtained under the different

reaction atmospheres (MD, SRM and POM) were completely reversible since, under

the final application of VWR = +2 V, all the reaction rates reached the same initial

values (Figures 5.1, 5.2 and 5.3, respectively).

The magnitude of the promotion effect obtained under the different methanol

conversion reactions was evaluated by means of two EPOC parameters commonly

used in this kind of studies: the reaction rate enhancement ratio (ρ) and the promotion

index (PIK+) defined by the following equations:

=

(5.18)

=

(5.19)

where r and r0 are the promoted (VWR < +2 V) and un-promoted (VWR = +2 V)

catalytic production rates, respectively, ∆r = r – r0 is the K+-induced change in the

catalytic reaction rate and K+ is the potassium coverage, which was calculated at each

applied catalyst potential (VWR) from the integration of the obtained current (I) vs.

time (t) curve through the Faraday’s law:

Electrochemical promotion of Ni in methanol conversion reactions

203

=

(5.20)

where n is the potassium ion charge, i.e., +1, F is the Faraday constant (96485 C), and

NG is the active surface area, 2.34 x 10-7

mol Ni. This value was estimated from the

total amount of Ni deposited (2.28 x 10-6

mol Ni cm-2

) and an approximate catalyst

dispersion value (D) of 3.62 %, which was calculated from the following semi

empirical equation [42]:

ρ (5.21)

where Mm is the nickel atomic weight, ρm is the Ni density (in g m-3

), NA is the

Avogadro number, d is an average Ni particle diameter, 337 Å, as estimated from the

XRD measurements, and is the metal atomic surface, 5.38 x 10-20

m2 atom

-1, as

obtained for Ni(111) [43], since this was found to be the main Ni crystallographic

direction (as will be observed below).

Figure 5.4a depicts the steady-state H2, CO, CO2 and H2CO production rate

enhancement ratios (ρ) obtained in the three studied atmospheres (MD, POM and

SRM) upon the different applied catalysts potentials. Figure 5.4b shows the

corresponding promotional index values (PIK+,i).

As was discussed in previous chapters, the ρ parameter represents the factor by

which the production rate of one compound has increased, under certain reaction

conditions, at each applied potential with respect to its unpromoted value (i.e. at VWR

= +2 V). On the other hand, the PIK+ provides a measure of the K+-induced

improvement in the catalytic activity normalized by the promoter concentration

created on the Ni surface. Hence, the dashed lines in figures groups a and b represent

the neutral values of these two parameters (ρi = 1 and PIK+,i = 0, respectively), i.e., the

un-promoted state.

Chapter 5

204

Figure 5.4. Effect of the applied potential (VWR) on the rate enhancement ratio (a1-a4) and

promotion index (b1-b4) values obtained for H2, CO, CO2 and H2CO at 360 ºC under the same

methanol decomposition (MD), steam reforming (SRM) and partial oxidation (POM)

conditions than Figures 5.1-5.3.

0

1

2

3

4

5

HMD

SRM

POM

0

2

4

6

-2.5 -1.5 -0.5 0.5 1.5 2.5

ρC

O

VWR / V

MD

SRM

POM

0.5

1

1.5

2

2.5

3

ρC

O

SRM

POM

0

5

10

15

20

25

-2.5 -1.5 -0.5 0.5 1.5 2.5

ρH

C

O

VWR / V

POM

2 22

a1)

a2)

a3)

a4)

-2

2

6

10

14

18

PI K

,C

O

MD

SRM

POM

-10

0

10

20

30

40

PI K

,H

MD

SRM

POM

-20

0

20

40

-2.5 -1.9 -1.3 -0.7 -0.1 0.5 1.1

PI K

,C

O

VWR / V

MD

SRM

POM

5

10

15

20

-2.5 -1.9 -1.3 -0.7 -0.1 0.5 1.1

PI K

,H

C

O

VWR / V

POM

2 22

++

++

b1)

b2)

b3)

b4)

Electrochemical promotion of Ni in methanol conversion reactions

205

It is clear that a decrease in the applied catalyst potential, under methanol

decomposition and steam reforming conditions, caused an enhancement in the

production of H2, CO and CO2. Thus, according to the rules of electrochemical

promotion [33], both MD (reaction 5.1) and SRM (reaction 5.6) are considered

electrophilic reactions under the studied conditions (upon ΔV < 0, ρ > 1 and PIK+ > 0).

The slight decrease in the H2 and CO production rates at VWR = +0.5 V (still

unpromoted state) and the progressive drop observed at very negative potentials can

be attributed to the concurrence of the CO dissociation on the catalyst surface

(reaction 5.10) as discussed above. Due to this deactivation under MD conditions, and

hence to the lower unpromoted reaction rates (vs. SRM), H2 and CO obtained via

methanol decomposition presented much higher ρ and PIK+ values, also in comparison

with the previous SRM studies with Pt and Pt-DLC catalyst films prepared by the

same deposition method.

On the other hand, it can also be observed the strong negative effect of K+ ions

on the H2 and CO production rates under POM conditions (upon ΔV < 0, ρ < 1 and

PIK+ < 0). However, CO2 and especially formaldehyde production showed favourable

EPOC parameters, which could not be well appreciated from Figure 5.3. When

reaction rates such as those of H2, CO and CO2 in the studied POM conditions barely

vary upon further decreasing the applied potential from a certain value (VWR = 0 V), it

is worth noting that the continuing rise in the K+ coverage makes the PIK+ values

approach zero. Hence, in such cases, this parameter actually provides useful

information only at low K+ coverages. Although the global partial oxidation reaction

presented apparent electrophobic behaviour, the Ni catalyst seemed then to be partially

activated for CO2 and H2CO production upon negative polarization. One can find in

literature other EPOC studies where Ni-based catalysts were activated in different

reactions such as water-gas shift [38], C2H4 hydrogenation [44], CH4 steam reforming

[45] or CO2 hydrogenation [35, 46]. In the present work, a Ni catalyst has been clearly

electrochemically promoted, for the first time, in the methanol decomposition and

steam reforming reactions for H2 production. However, given the unexpected sharp

drop in the Ni catalytic activity observed in the methanol partial oxidation under K+-

Chapter 5

206

promoted conditions, as compared with the results previously obtained with Pt, Au

and Cu catalysts, additional experiments were carried out in this reaction atmosphere.

5.3.2. Effect of the electrochemical promotion on the catalyst oxidation state

Figure 5.5 shows the influence of the oxygen partial pressure ( ) on steady-

state reaction rates at the same temperature and feed methanol concentration than

those used in preceding experiments, upon the imposition of two different potentials:

+2 V (unpromoted state) and -1 V (electropromoted state). The dashed line depicts the

partial oxidation conditions studied in Figure 5.3.

In the overall studied range of feed oxygen concentration (0.33 – 1.53 % O2),

the same promotional behaviour was found as in the previous potentiostatic transition

experiment, i.e., H2 and CO production rates sharply decreased upon negative

polarization, while the formation of carbon dioxide and especially formaldehyde was

enhanced. Under unpromoted conditions, both CO2 and H2CO production showed a

positive order with respect to oxygen partial pressure, which agrees with the

experimental results previously discussed, i.e., the strengthening of the metal-oxygen

chemisorptive bond via K+ backspillover was derived in increasing both CO2 and

H2CO production rates, thus according to the rules of chemical and electrochemical

promotion [1]. In the same sense, one could then expect a zero to negative order of H2

and CO production with respect to oxygen partial pressure in view of the previous

EPOC experimental results. Nevertheless, it is striking that both formation rates were

favoured by increasing the oxygen partial pressure up to around 0.8 kPa. This fact

seemed then to disagree with the poisoning effect observed under negative

polarizations in H2 and CO production with a 0.33 % O2 feed composition (Figure

5.3).

Electrochemical promotion of Ni in methanol conversion reactions

207

Figure 5.5. Effect of O2 partial pressure on H2 (a), CO (b), CO2 (c) and H2CO (d) production

under unpromoted (+2 V) and electropromoted (-1 V) state. CH3OH = 4.4 %, 360 ºC.

0

5

10

15

20

r H/

mo

l s-1

x 1

0-7

+2 V

-1 V

0

2

4

6

8r C

O/

mo

l s-1

x 1

0-7

+2 V

-1 V

0

0.5

1

1.5

2

r CO

/ m

ol s

-1x

10

-7

+2 V

-1 V

0

0.1

0.2

0.3

0 0.4 0.8 1.2 1.6

r HC

O/

mo

l s-1

x 1

0-7

PO / kPa

+2 V

-1 V

22

2

2

a)

b)

c)

d)

Chapter 5

208

In order to clarify this behaviour, the Ni catalyst film was characterized by

XRD after exposing the catalyst to both steam reforming and partial oxidation

conditions (with two different feed oxygen concentrations in the latter case), under

both unpromoted (VWR = +2 V) and electropromoted (VWR = -1 V) conditions. The

obtained diffractograms are shown in Figure 5.6.

Figure 5.6. XRD diffractograms of the Ni catalyst film after exposure for 1 hour to the same

methanol partial oxidation (a-c) and steam reforming (d,e) conditions than Figures 5.2 and 5.3,

under unpromoted (+2 V) and electropromoted (-1 V) state. Figure 5.6f shows the pattern of β-

Al2O3 solid electrolyte.

In all cases, the three main diffraction peaks of metallic nickel, (111), (200)

and (220), can be clearly observed at 2 = 44.5º, 51.8º and 76.4º, respectively,

Inte

nsi

ty /

a.u

.

Ni/K-βAl2O3

after POM 0.33% O2

-1 V

b) NiNiONi2O3

NiO2

25 35 45 55 65 75

Inte

nsi

ty /

a.u

.

2 θ / º

Ni/K-βAl2O3

after POM 1.53% O2

+2 V

c) Ni

Inte

nsi

ty /

a.u

.

Ni/K-βAl2O3

after POM 0.33% O2

+2 V

a) Ni

Inte

nsi

ty /

a.u

.

Ni/K-βAl2O3

after SRM+2V

d) Ni

25 35 45 55 65 75

Inte

nsi

ty /

a.u

.

2 Ѳ / º

βAl2O3f)

Inte

nsi

ty /

a.u

.

Ni/K-βAl2O3

after SRM-1V

e) Ni

Electrochemical promotion of Ni in methanol conversion reactions

209

exhibiting a face-centered cubic (FCC) crystalline structure (JCPDS, 87-0712). The

spectrum of the K-βAl2O3 solid electrolyte (Figure 5.6f) was also present in all the

diffractograms. This signal was even higher than that of Ni, probably due to the very

low thickness of the catalyst-working electrode (150 nm). No diffraction peaks

associated with nickel oxides were found under methanol steam reforming atmosphere

(Figures 5.6d and 5.6e) and neither under methanol partial oxidation conditions as

long as a positive potential was applied (Figures 5.6a and 5.6c), i.e., regardless of the

gas phase oxygen concentration in the overall range studied (see Figure 5.5).

However, very interestingly, upon negative polarization under POM atmosphere

(Figure 5.6b), the main diffraction peaks of NiO appeared at 2 = 37.26º, 43.29º,

62.89º and 79.43º (JCPDS, 78-0643). In addition, other peak at 2 = 39.13º and other

two peaks at 2 = 58.66º and 78.06º were also found, which could be attributed to the

presence of Ni2O3 (JCPDS, 14-0481) and NiO2 (JCPDS, 85-1977), respectively. From

this analysis, one cannot confirm the complete absence of oxidized Ni species under

other reaction conditions, but their presence was much more pronounced when the

catalyst film was subjected to POM atmosphere and negative applied potentials.

It becomes clear that the K+-promoted strengthening of the nickel-oxygen

chemisorptive bond under POM reaction conditions favoured the formation of

oxidized Ni species on the catalyst surface. Although this kind of effect is not very

commonly found in EPOC studies, some authors reported the oxidation or reduction

of Ru- [8], Rh- [9-11] and Pd-based [7] catalysts through the strengthening (upon

negative polarization) or weakening (upon positive polarization) of oxygen

chemisorption, respectively. While all the mentioned EPOC studies employed YSZ as

solid electrolyte, the oxidation state of a catalyst has been modified by alkali ions, for

the first time, in the present work. Then, this increase in the Ni oxidation state under

O2 atmosphere and negative polarization would explain the modification of the

reaction rates observed in Figure 5.3, since NiO is typically shown to be active in the

deep oxidation of methanol (CO2 and H2O production through reaction 5.5) instead of

the decomposition or partial oxidation reactions for CO and H2 production (reactions

5.1 and 5.7) [17, 47]. In addition it has been shown that NiO may present certain

Chapter 5

210

selectivity toward formaldehyde production [48, 49]. NiO has already been detected in

previous studies with Ni catalysts under methanol partial oxidation conditions, for

instance, by XRD [17] or XPS [47] analysis. Thus, the most intriguing results obtained

in this work demonstrate that Ni was not significantly oxidized by increasing the feed

O2 concentration up to 1.53 % under VWR = +2 V (Figure 5.6c), but it was partially

oxidized under 0.33 % O2 by applying –1 V (Figure 5.6b). Moreover, in view of

Figure 5.3, the reversibility of the phenomenon is clear, i.e., the activity was again

recovered after the final application of +2 V and, very likely, the reduced state of the

Ni catalyst.

Hence, the results obtained in the present work with a very thin Ni film

prepared by the cathodic arc deposition technique show the ability of the EPOC

phenomenon for the in-situ, reversible, partial modification of the catalyst oxidation

state, by varying only the K+ coverage rather than the oxygen partial pressure.

Moreover, on the basis of the XRD spectra obtained after the catalyst

exposition to the different reaction atmospheres, an approximate Ni particle size was

calculated in each case from the width of the main diffraction peak, Ni(111), by means

of the Scherrer equation:

Å

(5.22)

where λ is the X-ray wavelength (in Å), B is the full width at half maximum (FWHM)

of the diffraction peak (in radians) corrected for instrumental broadening, is the

Bragg angle and KW is the half-width Scherrer constant which can be considered 0.9

for this kind of catalyst [50]. Ni particle sizes between 31.5 and 35 nm were estimated

in all cases, which also denotes the high stability of this Ni catalyst film under the

studied reaction conditions.

5.3.3. Control of the catalyst selectivity via electrochemical promotion

The variation in the catalyst selectivity with the applied potential and the

influence of the temperature were evaluated under POM conditions (CH3OH/O2 = 4.4

%/0.3 %), that was the reaction atmosphere which allowed to obtain a wider variety of

Electrochemical promotion of Ni in methanol conversion reactions

211

compounds. Hence, Figure 5.7 shows the steady-state Ni selectivity toward H2, CO,

CO2 and H2CO under different applied catalyst potentials and reaction temperatures

between 280 and 360 ºC. As in previous chapters, the products selectivities were

calculated through the following equations:

(5.23)

(5.24)

(5.25)

(5.26)

where Fi,in and Fi,out are the molar flow rate of the i species at the inlet and outlet of the

reactor, respectively.

Figure 5.7. Steady-state H2 (a), CO (b), CO2 (c) and H2CO (d) selectivity vs. applied potential

(VWR) at different reaction temperatures. Methanol partial oxidation conditions: CH3OH/O2 =

4.4 %/0.33 %.

20

40

60

80

100

H2

Sele

ctiv

ity

/ %

280 ºC

320 ºC

360 ºC

0

20

40

60

80

-1.5 -0.5 0.5 1.5 2.5

CO

Sel

ecti

vity

/ %

VWR / V

280 ºC

320 ºC

360 ºC

0

15

30

45

60

75

CO

2Se

lect

ivit

y /

%

280 ºC

320 ºC

360 ºC

0

20

40

60

-1.5 -0.5 0.5 1.5 2.5

H2C

O S

elec

tivi

ty /

%

VWR / V

280 ºC

320 ºC

360 ºC

a)

b)

c)

d)

Increasing θK+ Increasing θK+

Chapter 5

212

This figure clearly reflects the strong detrimental effect of the K+-induced

formation of NiO on the H2 and CO production from methanol under negative

potentials, as well as also the concomitant increase in the CO2 and H2CO selectivities.

As expected, under unpromoted conditions (VWR = +2 V), the methanol conversion

(not shown here) increased with the reaction temperature. The Ni activity in methanol

dehydrogenation for H2 and CO production (reaction 5.1) was also favoured by the

temperature increase, as typically reported [17, 20, 21], while formaldehyde selectivity

gradually decreased, probably due to its decomposition to CO and/or oxidation to CO2

[40, 48]. However, the most pronounced change in Ni catalytic performance was

derived from the K+ back-spillover and the likely increase in the NiO formation (as

discussed above). For instance, at 280 ºC, H2 and CO selectivity decreased from

around 80 % (under unpromoted conditions) to 20 and 0 % (under electropromoted

conditions), respectively, while formaldehyde selectivity increased from 0 % to 70 %.

As already pointed out, the possibility of modifying the selectivity of the catalyst

toward different products via in-situ alkali promotion is one of the main applications

of the EPOC phenomenon and it has been studied in other reactions such as CO2

hydrogenation [51], selective catalytic reduction of NOx [52, 53] or Fischer-Tropsch

synthesis [54]. In this sense, the large variety of products and intermediate species

involved in the methanol conversion processes, along with the ability of EPOC to

partially modify the catalyst oxidation state, offer a wide range of possibilities with

focus, for instance, on fine chemistry.

As a summary, Figure 5.8 shows the main catalytic reactions found to take

place under the different studied reaction atmospheres (MD, SRM and POM) and both

unpromoted (Figure 5.8a) and electropromoted (Figure 5.8b) conditions. In all cases,

as discussed in the first section, the decrease in the applied catalyst potential and thus

the increase in the potassium coverage on the Ni catalyst surface led to an

strengthening of the chemisorptive bond between the metal and the electron acceptor

adsorbates, which were methoxy and other CHXO–like intermediate species in the

methanol decomposition (MD), H2O in the methanol steam reforming (SRM) and O2

in the methanol partial oxidation (POM).

Electrochemical promotion of Ni in methanol conversion reactions

213

Figure 5.8. Schematic representation of the main catalytic reactions taking place under positive

(a) and negative (b) polarization.

In the cases of MD and SRM, this K+ back-spillover caused the activation of

the methanol dehydrogenation mechanism. As a consequence, the Ni catalytic activity

for H2 and CO production via MD and also for H2 and CO2 production via SRM was

drastically enhanced showing a clear electrophilic behaviour, similarly to the previous

EPOC studies on methanol decomposition [13] and steam reforming (chapters 1 and

2) with noble metal catalysts. Under SRM atmosphere, the activation of water also

favoured the removal of carbonaceous species previously deposited on the Ni catalyst

MD

(4.4 % CH3OH)

CHXO

SRM

(4.4 % CH3OH, 5.2 % H2O)

H2O CH3OH

POM

(4.4 % CH3OH, 0.33 % O2)

O2 CH3OH

A

A

A

D

D

ΔV > 0

Ni (WE)

K+ K+ K+

Au (CE)

e-

e-

UNPROMOTED STATE

ΔV < 0

Ni (WE)

K+ K+ K+

Au (CE)

e-

e-

ELECTOPROMOTED STATE

CH3OH CO + 2H2

CO(a) C(a) + O(a)

CH3OH CO + 2H2

CO(a) C(a) + O(a)

CH3OH CO + 2H2

CH3OH + H2O CO2 + 3H2

C(a) + H2O CO + H2

CH3OH CO + 2H2

CO(a) C(a) + O(a)

CH3OH + H2O CO2 + 3H2

C(a) + H2O CO + H2

CH3OH CO + 2H2

CH3OH + 1/2O2 CO2 + 2H2

CH3OH + 3/2O2 CO2 + 2H2O

CH3OH CO + 2H2

CH3OH + 1/2O2 CO2 + 2H2

CH3OH + 3/2O2 CO2 + 2H2O

CH3OH + 1/2O2 H2CO + H2O

xNi + yO(a) NixOy

A D A D

A

D

Electron acceptor

Electron donor

a) b)

Chapter 5

214

surface under positive applied potentials (decreasing the deactivation of the catalyst).

Then, this kind of electrocatalytic systems could also be useful to obtain highly pure

H2 or syngas depending on the catalyst promotion state. However, under the

application of negative potentials and the presence of oxygen in the reaction mixture

(POM conditions), H2 and CO production was sharply decreased while CO2 and H2CO

production rate was slightly increased. This was attributed to a K+-promoted increase

in the nickel oxidation state as confirmed by XRD analysis.

Similar promotional effects can be observed in related studies of classical

chemical promotion. For instance, the addition of different alkali compounds such as

Na, K or Cs already showed to promote the methanol decomposition [25] and steam

reforming [26] reactions by weakening the C-H bond in the methoxy and formate

intermediate species. Alkali and alkaline earth oxides have also been employed to

improve the catalyst resistance to carbon deposition via water activation [39].

Moreover, in the same sense, many metal oxides such as CeO2 or ZnO are commonly

used as catalyst supports to modify its activity and selectivity in methanol conversion

processes as well as to increase its stability, involving in many cases some transition

between the catalyst oxidation states [3]. It should be then remarked that, in the

present work, the possible potassium promotional effects have been in-situ studied for

a single catalytic system under different reaction atmospheres, by only varying the

applied electric potential. In this way, the promoter coverage was exhaustively

controlled and the consequent modifications of the Ni catalytic performance were

carried out in a fully reversible way, which shows a great versatility and practical

utility of the phenomenon of electrochemical promotion.

5.4. CONCLUSIONS

The following conclusions can be drawn from this study:

- The phenomenon of electrochemical promotion of catalysis (EPOC) was

applied in different methanol conversion reactions for H2 production by using a Ni

catalyst film prepared by the cathodic arc deposition technique. The catalytic activity

Electrochemical promotion of Ni in methanol conversion reactions

215

and selectivity were in-situ modified by electrochemically pumping K+ ions from a

solid electrolyte support upon negative polarization.

- Under methanol decomposition conditions, the electrochemical promotion of

the Ni catalyst by K+ ions enhanced the dehydrogenation of methanol and intermediate

species to H2 and CO, although a strong deactivation by carbon deposition also took

place.

- Under methanol steam reforming conditions, the electrically-induced back-

spillover of K+ promoter ions favoured the H2, CO, and CO2 production, as well as

also the carbon gasification, via water activation.

- Under methanol partial oxidation conditions, the strengthening of oxygen

chemisorption caused a dramatic decrease in both the Ni catalytic activity and the

selectivity toward H2 and CO, probably due to an increase in the formation of surface

NiO, as observed in the XRD analysis. This shows the possibility of in-situ modifying

the catalyst oxidation state via electrochemical promotion with alkali ions.

- All the potassium-derived promotion effects were carried out in a controlled

way, by varying the applied electric potential, and showed to be fully reversible,

demonstrating the great interest of the EPOC phenomenon for many different practical

applications.

5.5. REFERENCES

[1] C.G. Vayenas, S. Bebelis, C. Pliangos, S. Brosda, D. Tsiplakides, Electrochemical

Activation of Catalysis: Promotion, Electrochemical Promotion and Metal-Support

Interactions, Kluwer Academic Publishers/Plenum Press, New York, 2001.

[2] A. Wieckowski, E.R. Savinova, C.G. Vayenas, Catalysis and Electrocatalysis at

Nanoparticles, Marcel Dekker, New York, 2003.

[3] P. Vernoux, L. Lizarraga, M.N. Tsampas, F.M. Sapountzi, A. De Lucas-Consuegra, J.L.

Valverde, S. Souentie, C.G. Vayenas, D. Tsiplakides, S. Balomenou, E.A. Baranova, Ionically

conducting ceramics as active catalyst supports, Chemical Reviews, 113 (2013) 8192-8260.

[4] C.G. Vayenas, Promotion, electrochemical promotion and metal-support interactions:

Their common features, Catalysis Letters, 143 (2013) 1085-1097.

Chapter 5

216

[5] A. Katsaounis, Recent developments and trends in the electrochemical promotion of

catalysis (EPOC), Journal of Applied Electrochemistry, 40 (2010) 885-902.

[6] A. de Lucas-Consuegra, New trends of Alkali Promotion in Heterogeneous Catalysis:

Electrochemical Promotion with Alkaline Ionic Conductors, Catalysis Surveys from Asia,

2015, DOI:10.1007/s10563-014-9179-6.

[7] C. Jiménez-Borja, F. Dorado, A. De L.-Consuegra, J.M. G.-Vargas, J.L. Valverde,

Electrochemical promotion of CH4 combustion over a Pd/CeO 2-YSZ catalyst, Fuel Cells, 11

(2011) 131-139.

[8] D. Theleritis, S. Souentie, A. Siokou, A. Katsaounis, C.G. Vayenas, Hydrogenation of CO

2 over Ru/YSZ electropromoted catalysts, ACS Catalysis, 2 (2012) 770-780.

[9] E.A. Baranova, A. Thursfield, S. Brosda, G. Fóti, C. Comninellis, C.G. Vayenas,

Electrochemical promotion of ethylene oxidation over Rh catalyst thin films sputtered on YSZ

and TiO2/YSZ supports, Journal of the Electrochemical Society, 152 (2005) E40-E49.

[10] N. Kotsionopoulos, S. Bebelis, Electrochemical promotion of the oxidation of propane on

Pt/YSZ and Rh/YSZ catalyst-electrodes, Journal of Applied Electrochemistry, 35 (2005) 1253-

1264.

[11] A. Nakos, S. Souentie, A. Katsaounis, Electrochemical promotion of methane oxidation

on Rh/YSZ, Applied Catalysis B: Environmental, 101 (2010) 31-37.

[12] A. De Lucas-Consuegra, A. Caravaca, P.J. Martínez, J.L. Endrino, F. Dorado, J.L.

Valverde, Development of a new electrochemical catalyst with an electrochemically assisted

regeneration ability for H2 production at low temperatures, Journal of Catalysis, 274 (2010)

251-258.

[13] S. Neophytides, C.G. Vayenas, Non-faradaic electrochemical modification of catalytic

activity. 2. The case of methanol dehydrogenation and decomposition on Ag, Journal of

Catalysis, 118 (1989) 147-163.

[14] C.G. Vayenas, S. Neophytides, Non-faradaic electrochemical modification of catalytic

activity III. The case of methanol oxidation on Pt, Journal of Catalysis, 127 (1991) 645-664.

[15] C.A. Cavalca, G. Larsen, C.G. Vayenas, G.L. Haller, Electrochemical modification of

CH3OH oxidation selectivity and activity on a Pt single-pellet catalytic reactor, Journal of

Physical Chemistry, 97 (1993) 6115-6119.

[16] J.K. Hong, I.H. Oh, S.A. Hong, W.Y. Lee, Electrochemical oxidation of methanol over a

silver electrode deposited on yttria-stabilized zirconia electrolyte, Journal of Catalysis, 163

(1996) 95-105.

Electrochemical promotion of Ni in methanol conversion reactions

217

[17] N. Iwasa, M. Yoshikawa, W. Nomura, M. Arai, Transformation of methanol in the

presence of steam and oxygen over ZnO-supported transition metal catalysts under stream

reforming conditions, Applied Catalysis A: General, 292 (2005) 215-222.

[18] M.S. Hegde, Electron spectroscopic studies of the adsorption and decomposition of

methanol on transition metals. A review, Proceedings of the Indian Academy of Sciences -

Chemical Sciences, 93 (1984) 373-387.

[19] G.C. Wang, Y.H. Zhou, Y. Morikawa, J. Nakamura, Z.S. Cai, X.Z. Zhao, Kinetic

mechanism of methanol decomposition on Ni(111) surface: A theoretical study, Journal of

Physical Chemistry B, 109 (2005) 12431-12442.

[20] B.S. Chen, J.L. Falconer, Alcohol Decomposition by Reverse Spillover, Journal of

Catalysis, 144 (1993) 214-226.

[21] N. Takezawa, N. Iwasa, Steam reforming and dehydrogenation of methanol: Difference in

the catalytic functions of copper and group VIII metals, Catalysis Today, 36 (1997) 45-56.

[22] R.M. Navarro, M.A. Peña, J.L.G. Fierro, Hydrogen production reactions from carbon

feedstocks: Fossil fuels and biomass, Chemical Reviews, 107 (2007) 3952-3991.

[23] D.R. Palo, R.A. Dagle, J.D. Holladay, Methanol steam reforming for hydrogen

production, Chemical Reviews, 107 (2007) 3992-4021.

[24] S. Sá, H. Silva, L. Brandão, J.M. Sousa, A. Mendes, Catalysts for methanol steam

reforming-A review, Applied Catalysis B: Environmental, 99 (2010) 43-57.

[25] A. Guerrero-Ruiz, I. Rodriguez-Ramos, J.L.G. Fierro, Dehydrogenation of methanol to

methyl formate over supported copper catalysts, Applied Catalysis, 72 (1991) 119-137.

[26] H.N. Evin, G. Jacobs, J. Ruiz-Martinez, U.M. Graham, A. Dozier, G. Thomas, B.H.

Davis, Low temperature water-gas shift/methanol steam reforming: Alkali doping to facilitate

the scission of formate and methoxy C-H bonds over Pt/ceria catalyst, Catalysis Letters, 122

(2008) 9-19.

[27] K. Klier, Preparation of bifunctonal catallysts, Catalysis Today, 15 (1992) 361-382.

[28] J.M. Pigos, C.J. Brooks, G. Jacobs, B.H. Davis, Low temperature water-gas shift: The

effect of alkali doping on the Csingle bondH bond of formate over Pt/ZrO2 catalysts,

Applied Catalysis A: General, 328 (2007) 14-26.

[29] L. Alejo, R. Lago, M.A. Peña, J.L.G. Fierro, Partial oxidation of methanol to produce

hydrogen over Cu-Zn-based catalysts, Applied Catalysis A: General, 162 (1997) 281-297.

[30] A. Anders, Cathodic Arcs: From Fractal Spots to Energetic Condensation. Springer Series

on Atomic, Optical, and Plasma Physics, Springer-Verlag, New York, 2008.

Chapter 5

218

[31] P.B. Tøttrup, Kinetics of decomposition of carbon monoxide on a supported nickel

catalyst, Journal of Catalysis, 42 (1976) 29-36.

[32] Z. Murayama, I. Kojima, E. Miyazaki, I. Yasumori, Dissociation of carbon monoxide on

stepped nickel surfaces, Surface Science, 118 (1982) L281-L285.

[33] C.G. Vayenas, S. Brosda, Electron Donation-Backdonation and the Rules of Catalytic

Promotion, Top. Catal., 57 (2014) 1287-1301.

[34] Y. Liu, T. Hayakawa, T. Ishii, M. Kumagai, H. Yasuda, K. Suzuki, S. Hamakawa, K.

Murata, Methanol decomposition to synthesis gas at low temperature over palladium

supported on ceria-zirconia solid solutions, Applied Catalysis A: General, 210 (2001) 301-

314.

[35] E. Ruiz, D. Cillero, P.J. Martínez, Á. Morales, G.S. Vicente, G. De Diego, J.M. Sánchez,

Bench-scale study of electrochemically assisted catalytic CO2 hydrogenation to hydrocarbon

fuels on Pt, Ni and Pd films deposited on YSZ, Journal of CO2 Utilization, 8 (2014) 1-20.

[36] W. Cao, G. Chen, S. Li, Q. Yuan, Methanol-steam reforming over a ZnO-Cr2O3/CeO2-

ZrO2/Al2O3 catalyst, Chemical Engineering Journal, 119 (2006) 93-98.

[37] G. Jacobs, B.H. Davis, In situ DRIFTS investigation of the steam reforming of methanol

over Pt/ceria, Applied Catalysis A: General, 285 (2005) 43-49.

[38] A. De Lucas-Consuegra, A. Caravaca, J. González-Cobos, J.L. Valverde, F. Dorado,

Electrochemical activation of a non noble metal catalyst for the water-gas shift reaction,

Catalysis Communications, 15 (2011) 6-9.

[39] I. Alstrup, B.S. Clausen, C. Olsen, R.H.H. Smits, J.R. Rostrup-Nielsen, Promotion of

steam reforming catalysts, Studies in Surface Science and Catalysis, 1998, pp. 5-14,

[40] K. Kähler, M.C. Holz, M. Rohe, A.C. Van Veen, M. Muhler, Methanol oxidation as probe

reaction for active sites in Au/ZnO and Au/TiO2 catalysts, Journal of Catalysis, 299 (2013)

162-170.

[41] L.A. Espinosa, R.M. Lago, M.A. Peña, J.L.G. Fierro, Mechanistic aspects of hydrogen

production by partial oxidation of methanol over Cu/ZnO catalysts, Topics in Catalysis, 22

(2003) 245-251.

[42] J.R. Anderson, Structure of Metallic Catalysts, Academic Press1975.

[43] Y. Hoshino, S. Matsumoto, Y. Kido, Ultrathin Ni layers grown epitaxially on SiC(0001)

at room temperature, Physical Review B - Condensed Matter and Materials Physics, 69 (2004)

155303-155301-155303-155307.

Electrochemical promotion of Ni in methanol conversion reactions

219

[44] T.I. Politova, V.A. Sobyanin, V.D. Belyaev, Ethylene hydrogenation in electrochemical

cell with solid proton-conducting electrolyte, Reaction Kinetics & Catalysis Letters, 41 (1990)

321-326.

[45] I.V. Yentekakis, Y. Jiang, S. Neophytides, S. Bebelis, C.G. Vayenas, Catalysis,

electrocatalysis and electrochemical promotion of the steam reforming of methane over Ni film

and Ni-YSZ cermet anodes, Ionics, 1 (1995) 491-498.

[46] V. Jiménez, C. Jiménez-Borja, P. Sánchez, A. Romero, E.I. Papaioannou, D. Theleritis, S.

Souentie, S. Brosda, J.L. Valverde, Electrochemical promotion of the co2 hydrogenation

reaction on composite ni or ru impregnated carbon nanofiber catalyst-electrodes deposited on

YSZ, Applied Catalysis B: Environmental, 107 (2011) 210-220.

[47] H. Iwai, T. Umeki, M. Yokomatsu, C. Egawa, Methanol partial oxidation on Cu-Zn thin

films grown on Ni(1 0 0) surface, Surface Science, 602 (2008) 2541-2546.

[48] S. Damyanova, M.L. Cubeiro, J.L.G. Fierro, Acid-redox properties of titania-supported

12-molybdophosphates for methanol oxidation, Journal of Molecular Catalysis A: Chemical,

142 (1999) 85-100.

[49] J. Estellé, J.E. Sueiras, Methanol oxidation on semiconducting oxides, Reaction Kinetics

& Catalysis Letters, 51 (1993) 119-124.

[50] J.I. Langford, A.J.C. Wilson, Scherrer after Sixty Years: A Survey and Some New Results

in the Determination of Crystallite Size, J. Appl. Cryst., 11 (1978) 102-113.

[51] E. Ruiz, D. Cillero, P.J. Martínez, Á. Morales, G.S. Vicente, G. De Diego, J.M. Sánchez,

Bench scale study of electrochemically promoted catalytic CO2 hydrogenation to renewable

fuels, Catalysis Today, 210 (2013) 55-66.

[52] F. Dorado, A. de Lucas-Consuegra, P. Vernoux, J.L. Valverde, Electrochemical

promotion of platinum impregnated catalyst for the selective catalytic reduction of NO by

propene in presence of oxygen, Applied Catalysis B: Environmental, 73 (2007) 42-50.

[53] A. Palermo, R.M. Lambert, I.R. Harkness, I.V. Yentekakis, O. Mar'ina, C.G. Vayenas,

Electrochemical promotion by Na of the platinum-catalyzed reaction between CO and NO,

Journal of Catalysis, 161 (1996) 471-479.

[54] A.J. Urquhart, F.J. Williams, R.M. Lambert, Electrochemical promotion by potassium of

Rh-catalysed fischer-tropsch synthesis at high pressure, Catalysis Letters, 103 (2005) 137-141.

Chapter 5

220

221

Chapter 6

ELECTROCHEMICALLY ASSISTED PRODUCTION AND

STORAGE OF HYDROGEN

A NOVEL CONTRIBUTION OF ALKALINE ELECTROCHEMICAL

CATALYSTS

In this last chapter, the possibility of applying the electrochemical promotion

(EPOC) with alkaline conductors in the fields of H2 production and storage was

investigated. For this purpose, Ni catalyst films were deposited on K-βAl2O3 by the

oblique angle deposition technique. Under methanol steam reforming conditions (280

ºC) and negative polarization, this electrocatalytic system allowed to produce and

store H2 with a very high yield per amount of metal (up to 19 g H2 x 100 g Ni-1

, on a

Ni catalyst film with a porosity of 35 %) due to the promotional effect of K+ ions.

Then, it was also possible to release the captured H2 under positive polarization. This

represents a new application of the EPOC of great interest for the in-situ production,

storage and release of H2 in a controlled way. The influence of the catalyst

microstructure, the applied negative current and the reaction atmosphere was studied,

and a H2 storage mechanism was proposed with the support of different

characterization techniques. It was mainly based on the spillover of H atoms on

grapheme oxide which was simultaneously formed under electropromoted conditions.

e-

Ni

K+ K+ K+ K-βAl2O3

Au

ΔV < 0

e-K+

K+K+

K+ K+

K+K+

K+

K+ K+ K+K+ K+

HH CO

K+

K+

K+

HH

H

HHH

H

K+

H

H

COH2

CH3OH

KOH

KH

K-C-O-H

K-C-O-H

GONi Ni

Ni

Graphene oxide (GO)

K+-assisted

H2 production

K+-assisted

GO formation

K+-assisted

H adsorption

on GO

K+

Chapter 6

222

6.1 INTRODUCTION

As mentioned in previous chapters, hydrogen is a feedstock of paramount

importance in the chemical industry (around 5 x 1010

kg consumed per year worldwide

[1]). H2 is also a very promising energy carrier with application in internal combustion

engines and fuel cell technology. One of its main advantages is its high gravimetric

energy density, although its volumetric energy density is very low. It is the most

abundant element on Earth but less than 1 % is present as molecular hydrogen. In this

thesis, the electrochemical activation of different catalysts for its catalytic production

from methanol has been considered, but H2 can also be obtained from other fossil

fuels, water electrolysis or photoelectrolysis processes [2-4]. Hence, hydrogen can be

considered as a clean synthetic fuel if it is finally burnt with oxygen (in this case the

only exhaust gas is water). In addition, an improvement from the environmental point

of view would be attained if renewable energies are employed to split the water, or

sustainable sources such as bioalcohols are used in the catalytic conversion processes.

Moreover, despite its bad reputation, H2 is safe in open enclosures due to its high

volatility and non-toxicity. However, its storage remains a problem for stationary and

specially transportation applications.

Hydrogen is conventionally stored on its pure form either as a compressed gas

or as a cryogenic liquid. In the first case, H2 cylinders are typically filled at 200 bar (as

those employed in some sections of this manuscript), although pressures up to 450 bar

can be reached if carbon-fibre-reinforced composite materials are used. In the second

case, liquid hydrogen presents a higher density than the compressed one, but it

requires specialized infrastructure to avoid gaseous H2 losses [1, 5]. Hence, these

traditional storage methods suffer from difficult operational conditions and sizable

safety risks. Some chemical compounds like methanol, formic acid or ammonia can

also be themselves considered as H2 storage media, since these molecules can be

decomposed (dehydrogenated) likely in a catalytic process [5, 6]. This storage method

would be only reversible in the case that the dehydrogenated compound is collected

and re-hydrogenated, being totally subjected to the availability of the raw material.

Other method is, for example, the encapsulation of the H2 gas inside a guest (solid)

Electrochemically assisted production and storage of hydrogen

223

structure, from which hydrogen can be released by a pressure and temperature swing

[5, 7]. However, most of the H2 storage systems can be divided into two main

categories: physisorption of molecular H2 and chemisorptions of atomic hydrogen.

Since H2 physisorption is a non-activated process, fast kinetics and

reversibility are expected. However, the interaction between H2 and the adsorbant

(dominated by weak van der Waals forces) usually leads to unspecific (non localized)

adsorption, thus, significant hydrogen storage is only possible under cryogenic

temperatures. H2 physisorption on all kinds of carbon materials (carbon nanotubes,

carbon fibers, graphite, graphite oxide, activated carbons, etc) as well as on metal-

organic frameworks (MOFs), covalent organic frameworks (COFs), different kinds of

porous polymers and zeolites has been widely studied [1, 5-8]. In all these studies, a

maximum H2 storage capacity of about 1.5 wt. % is generally achieved at room

temperature while values ranging from 5 to 6 wt. % can be obtained at -196 ºC. In

these processes, great efforts are being carried out to develop adsorbents with ultra

high surface area.

On the other hand, atomic hydrogen chemisorption is usually followed by

chemical compounds formation leading to H2 storage as metal hydrides or nitrides.

Many metals and alloys are capable of reversibly adsorbing large amounts of

hydrogen by this mechanism. Charging can be performed using molecular gas (which

needs to be dissociated before adsorption) or H atoms from an electrolyte (where the

main drawbacks are the covering of the adsorbent by water or H+ species and the harsh

acidic environment) [9]. For instance, alloys derived from LaNi5 or CoNi5 present

some promising properties such as fast and reversible sorption with small hysteresis,

good cycling life, pressure and temperature conditions inside the operational window

for hydrogen PEMFC (25-100 ºC, 1-10 atm) and good volumetric hydrogen density.

However, their gravimetric capacities are very low (e.g., 0.9 wt. % for LaNi5H6, 1.1

wt. % for CoNi5H4). A higher hydrogen storage capacity is obtained with other

hydrides such as Mg2NiH4 (3.6 wt. %), MgH2 (7.6 wt. %) or LiBH4 (18 wt. %), but in

these cases the hydride formation is much slower and high temperatures (up to 600 ºC)

may be required for its decomposition, leading to problems of storage reversibility [1,

Chapter 6

224

5, 7]. In this kind of systems, the main strategies for the enhancement of the storage

performance consist of the decrease in the metal grain size, the increase in the defects

concentration, and the incorporation of additional materials on which H atoms may

spill over, thus increasing the global storage capacity [8, 10-13].

The aim of this chapter is to explore the possibility of applying the EPOC

phenomena in the field of H2 production and storage. Previous works have already

demonstrated the use of cationic electrochemical catalysts for NOx storage-reduction

process (NSR) [14, 15] and CO2 capture [16]. Thus, in the present work, a novel

electrocatalytic system is developed for the electrochemically assisted production and

storage of H2 from methanol-water streams. The idea is to prepare an electrochemical

catalyst of suitable performance for H2 production from methanol steam reforming

(SRM), which is capable of operating upon different electrochemical polarization for

the controllable H2 storage and release, under mild reaction conditions. For that

purpose, a porous Ni catalyst film was deposited on a K-βAl2O3 solid electrolyte. Ni

was selected as the active phase because of its good catalytic activity under methanol

steam reforming conditions (proven in the previous chapter). The Ni catalyst film was

prepared by an oblique angle deposition procedure, which is similar to that carried out

in chapter 4, in order to achieve a Ni catalyst film with suitable porosity and high

catalyst-electrode surface area. The influence of the applied electrical current, the

reaction atmosphere and the microstructure of the catalyst film on the H2 storage

capacity was studied. In addition, different ex-situ characterization techniques were

performed: SEM, XPS, Raman Spectroscopy and FTIR analysis in order to identify

the nature of the different H2 storage surface compounds formed under the different

polarization conditions.

6.2. EXPERIMENTAL

6.2.1. Preparation of the electrochemical catalyst

The electrochemical catalyst consisted of a continuous thin Ni film (geometric

area of 2.01 cm2), which also behaved as working electrode (WE). Similarly to the

previous chapters, a 19-mm-diameter, 1-mm-thick K-βAl2O3 (Ionotec) pellet was used

Electrochemically assisted production and storage of hydrogen

225

as the cationic solid electrolyte (i.e. as source of K+ promoter ions), and inert Au

counter (CE) and reference (RE) electrodes were prepared from an oganometalic paste

(Fuel Cell Materials 233001). Two different active Ni catalyst films were deposited by

oblique angle deposition (OAD). In both cases, a Ni target pellets (Goodfellow,

99.9999 % purity) were evaporated under vacuum conditions by bombardment with a

high kinetic energy (< 5 keV) and intensity (150 mA) electron beam. Then, the vapour

was condensed onto the surface of the substrate (K-βAl2O3) at room temperature. Two

different zenithal angles, α, formed between the perpendicular to the substrate surface

and the evaporation direction, were used to prepare two different samples. The first

catalyst (called as Ni 0º) was grown at α = 0º and consisted of a fairly compact Ni film

with a thickness of around 1.3 µm, as obtained from the cross-sectional SEM image

(not shown). The second Ni catalyst (called as Ni 80º) was deposited at a zenithal

evaporation angle of α = 80º to enhance the shadowing effects during the film growth

(as in chapter 4) [17-19], leading to a 0.6 µm-thick film composed of tilted Ni

nanocolumns. The final metal loading was 1.09 mg Ni cm-2

in the case of Ni 0º and

0.36 mg Ni cm-2

in the case of Ni 80º, as determined by X-ray fluorescence (XRF) in a

Siemens SRS 3000 sequential spectrophotometer with a rhodium tube as the radiation

source. Hence, a relative film porosity of around 35 % was estimated in the latter case

on the basis of the thickness values obtained from cross-sectional micrographs (1.3

and 0.6 µm for the catalysts Ni 0º and Ni 80º, respectively). Similar films were also

prepared on (polished) silicon substrates for comparison purposes.

This catalyst preparation method was carried out in collaboration with Dr.

Agustín R. González-Elipe and Dr. Víctor J. Rico from the Institute of Materials

Science of Seville (CSIC – University of Seville).

Figure 6.1 shows the top-view SEM images of the nickel catalyst films as-

deposited by normal (Ni 0º) and oblique (Ni 80º) glancing angle deposition on silicon

substrate for a reference study (Figures 6.1a and 6.1b) and on the K-βAl2O3 pellet

(Figures 6.1c and 6.1d) which will be used for the subsequent electrochemical

experiments below.

Chapter 6

226

Figure 6.1. Top view SEM images of the Ni films deposited at a normal geometry (Ni 0º) and

at a deposition angle of 80º (Ni 80º) on silicon (a and b) and K-βAl2O3 (c and d) substrates.

It can be observed that normal physical vapour deposition of Ni (i.e., at an

incident angle of α = 0º) led to fairly compact films composed of perpendicular, thin,

close-packed nanocolumns. On the other hand, the oblique angle deposited (OAD)

nickel at α = 80º resulted in separated groups of tilted nanocolumns, thus providing a

greater porosity (35 %), as can be observed in Figures 6.1b and 6.1d for the silicon

and K-βAl2O3 substrates, respectively. SEM images of the sample deposited on silicon

(Figures 6.1a and 6.1b) exhibited a very ordered and homogeneous distribution,

similar to oblique angle deposited films on silicon shown in previous studies based on

Co [20], Ti [21, 22], Pt-Ni [23], V-Mg [21, 22], TiO2 [18, 24] or Ta2O5 [19, 24]. The

Ni film prepared on the K-βAl2O3 pellet (Figures 6.1c and 6.1d) presented a much

more irregular microstructure, which was probably influence by the higher surface

roughness of this substrate and the different interaction between the Ni atoms on

silicon or K-βAl2O3 substrate. A more pronounced shadowing effect was also obtained

on K-βAl2O3 during the oblique angle deposition, giving rise to agglomerations of Ni

1 µm 1 µm

5 µm 5 µm

a) b)

c) d)

Ni 0º Ni 80º

Electrochemically assisted production and storage of hydrogen

227

columns separated by large areas of alumina which were left in view (Figure 6.1d).

Anyway, as measured by XRF, the same amount of Ni was deposited on both kinds of

substrates. Thus, it is clear that the porosity, the morphology and the gas-exposed

metal surface area of the catalyst films can be severely modified by varying the

zenithal angle during their deposition.

Both obtained Ni catalyst-working electrodes (Ni 0º and Ni 80º) were

electrically conductive. The resulting Ni/K-βAl2O3/Au electrochemical catalysts were

placed into the single chamber solid electrolyte cell reactor and the three electrodes

(working, counter and reference) were connected to the potentiostat-galvanostat as

done in the previous chapters.

6.2.2. Characterization measurements

The surface microstructure of the Ni films deposited on silicon and K-βAl2O3

substrates was examined by scanning electron microscopy (SEM) analysis using a

Hitachi S4800 field emission microscope operated at 2 keV. In order to investigate the

composition of the catalyst surface during the H2 storage process and identify the

possible surface compounds on it, the catalyst film Ni 80º was characterized after the

catalytic activity measurements. Firstly, it was exposed to methanol steam reforming

conditions (CH3OH/H2O = 4.4 %/5.2 %, Ar balance, 6 NL h-1

) at 280 ºC while a

positive potential of VWR = +2 V was applied for 1 hour, in order to define a Ni

surface free of K+ ions and hence a cleaned (reference) catalyst state. A subsequent

negative potential of VWR = -1 V was applied to electrochemically transfer K+ ions

from the K-βAl2O3 solid electrolyte to the catalyst film, which favored the H2

production/storage process as will be explained below. After 1 hour under these

reaction conditions, the electrochemical catalyst was cooled down and the applied

potential was interrupted (open circuit conditions) at 100 ºC. Then, the catalyst was

transferred to different characterization equipments.

Besides SEM, X-ray photoelectron spectroscopy (XPS) was performed with a

PHOIBOS-100 spectrometer with Delay Line Detector (DLD) from SPECS, which

worked in the constant pass energy mode fixed at 30 eV. Monochromatic Mg Kα

Chapter 6

228

radiation was used as excitation source and the binding energy (BE) scale of the

spectra referenced to the C 1s of graphitic carbon taken at 284.6 eV. The study of the

binding of the surface compounds was performed through specular reflectance FTIR

spectroscopy using a Jasco FT/IR-6200 spectrometer. All spectra were typically

obtained using 500 scans with a resolution of 4 cm-1

. To remove the background, the

signal obtained from a gold substrate was subtracted. Raman spectra were also

recorded with a HORIBA HR-800-UV microscope. For these measurements, a green

laser (532.14 nm) working at 600 lines per mm and a 100x objective were used.

6.2.3. Catalytic activity measurements

The catalytic activity measurements were carried out in the experimental setup

described in Chapter 1, section 1.2.3. A temperature of 280 ºC was fixed in the whole

study and the feeding stream was varied along the experiments. In general, the H2

production and storage was performed under methanol steam reforming (SRM)

conditions (CH3OH/H2O = 4.4 %/5.2 %, Ar balance, 6 NL h-1

), while the

electrochemical decomposition of the stored compounds was carried out in Ar stream.

In this way, released H2, CO and CO2 molecules were clearly detected in the outlet

stream. The different stages that comprised each experiment will be described in detail

in the next section.

6.3. RESULTS AND DISCUSSION

6.3.1. Preliminary experiments of H2 production and storage

Different systematic electrocatalytic experiments were carried out with both

kinds of catalyst films (Ni 0º and Ni 80º) in order to study the possibility of producing

and storing H2 at 280 ºC. Each experiment consisted of 4 consecutive steps, as

follows:

1) The Ni catalyst was subjected to methanol steam reforming (SRM)

conditions (CH3OH/H2O = 4.4 %/5.2 %, Ar balance, 6 NL h-1

) and to a positive

applied potential of VWR = 2 V for 1 hour in order to remove all the positive promoter

ions (K+) that may have remained on the catalyst/working electrode. In this way, a

reference cleaned Ni catalyst surface was established.

Electrochemically assisted production and storage of hydrogen

229

2) A constant negative current, I, was applied for 45 min to electrochemically

transfer K+ ions at a constant rate, = I/F (F being the Faraday constant), from the

K-βAl2O3 solid electrolyte to the Ni catalyst. Once located on the Ni catalyst, these

potassium ions may likely migrate over the entire gas-exposed Ni catalyst-surface

and/or react with coadsorbed species [25]. Under the studied reaction conditions, as

analysed in previous chapters, the back-spillover of K+ ions could modify the catalytic

properties of Ni in the methanol steam reforming or in some side reaction due to the

promotional effect of the potassium ions. In addition, a purely electrocatalytic reaction

between potassium ions and adsorbed reactants or products would lead to the

formation of different surface compounds, which may contain hydrogen and be

decomposed by inverting the polarization [14-16, 26].

3) After adsorption, the reactor was purged with Ar and the system was kept

under open circuit conditions.

4) Finally, a positive potential was applied (still under Ar atmosphere) to

decompose the possible formed potassium-derived surface compounds and to return

the K+ ions back to the solid electrolyte. The experiment ended when gas compounds

stopped of being released from the catalyst surface.

Figure 6.2 shows the time variation of the outlet molar flow rates of the

different detected products, the catalyst potential (VWR) and the current (I) during one

of these experiments performed at 280 ºC with the catalyst Ni 0º. After one hour under

methanol steam reforming conditions and an applied catalyst potential of +2 V (not

depicted), the time (t) was set at 0 min and the unpromoted H2, CO and CO2

production rates reached the following values: 2.55 x 10-8

mol H2 s-1

, 5.8 x 10-9

mol

CO s-1

and 8 x 10-10

mol CO2 s-1

, respectively. According to these values, the methanol

decomposition reaction (MD, reaction 6.1) seemed to prevail over the steam reforming

reaction (SRM, reaction 6.2) as commonly found on supported Ni catalysts [27-29].

CH3OH 2H2 + CO (6.1)

CH3OH + H2O 3H2 + CO2 (6.2)

Chapter 6

230

Figure 6.2. Time variation of H2, CO and CO2 molar flow rates (a), current (I) and catalyst

potential (VWR) (b) with catalyst film Ni 0º during the following steps at 280ºC: H2

production/storage, CH3OH/H2O = 4.4 %/5.2 %, I = -20 µA; Cleaning, 100 % Ar, OC; H2

release, 100 % Ar, VWR = +2 V.

Then, the imposition of a constant negative current of -20 µA under the same

reaction conditions led to the alkali ions supply to the catalyst film (at a rate of I/F =

2.1 x 10-10

mol K+ s

-1, according to Faraday’s law). After 45 min, a total number of 5.6

x 10-7

mol K+ was electrochemically migrated and the catalyst potential decreased to

around -0.7 V. During this step, the hydrogen obtained with respect to the CO and

CO2 widely exceeded the stoichiometric value according to the previous reactions, and

0

1

2

3

r /

mo

l s-1

x 10

-8

-1

0

1

2

-0.1

0.3

0.7

1.1

0 10 20 30 40 50 60 70

VW

R/

V

I / A

x 1

0-3

Time / min

-20 µA4.4 % CH3OH + 5.2 % H2O

+2 VAr

OCAr

rH2

rCO

rCO2

VWR

I

a)

b)

Electrochemically assisted production and storage of hydrogen

231

both H2 and CO production rates decreased with time on stream to 1.92 x 10-8

mol H2

s-1

, 3.1 x 10-9

mol CO s-1

. This was attributed to the deactivation of the Ni catalyst as a

consequence of the occurrence of carbon deposits derived from the dehydrogenation

of methyl groups or, most likely, from CO dissociation [30, 31], which was probably

favoured by the K+ promotional effect. In fact, Ni-based catalysts are typically prone

to the formation of different forms of carbon, that may poison the catalytic activity,

from a wide variety of molecules such as: methane [32-34], ethylene [35], methanol

[36, 37] or other alcohols [38, 39]. At first glance one cannot appreciate any apparent

promotional effect of potassium on the Ni catalytic activity, probably hindered by the

previously mentioned poisoning effect. This kind of Ni catalyst films thus behaved

very differently from those previously prepared by CAD and electrochemically

promoted in the methanol steam reforming reaction in chapters 1, 2 and 5.

Once the steam reforming reaction was interrupted and all the lines were fully

cleaned by Ar, it can be observed that the final application of VWR = +2 V (which

caused the migration of the K+ ions back to the solid electrolyte according to the

obtained positive current), led to certain H2, CO and CO2 production. Due to the

absence of any possible gaseous molecules in this step (Ar atmosphere), these peaks

could only be attributed to the desorption of the previously stored surface species.

Hence, potassium ions very likely participated in the promotion/formation of such

storing phases under the previous methanol steam reforming conditions and the

negative current imposition. Moreover, the K+ ions not only presented a promotional

effect on the H2 capture mechanism, but also on its release. The subsequent removal of

potassium ions under Ar atmosphere by the positive polarization caused the

decomposition and the release of such stored phases leading to the observed presence

of H2, CO and CO2 in the gas phase. From the integration of each peak, total amounts

of 4.68 x 10-6

mol H2, 4.8 x 10-7

mol CO and 8 x 10-8

mol CO2 were released during

the last step of the experiment under anodic polarization.

The physisorption of H2 molecules on the Ni catalyst was fully dismissed

because of the employed operation conditions (1 atm, 280 ºC). Hence, one can suggest

different possibilities for explaining the above results through hydrogen chemisoprtion

Chapter 6

232

mechanisms. Firstly, H atoms adsorbed on Ni active sites could lead to the formation

of nickel hydrides [1, 7]. On the other hand, the H adatoms could also migrate (via

spillover) from the Ni active sites and adsorb on the surface of a secondary material

[10, 11], e.g., carbonaceous surface species formed during the SRM reaction. Another

plausible explanation could be the H2 storage in the form of some methanol-derived

intermediate species on the Ni active sites, such as methoxy groups (CH3O). In fact,

potassium has shown to stabilize this kind of intermediate species on catalytically

active sites [40-42]. Moreover, one cannot rule out the electrocatalytic reaction

between K+ ions and the H2O, H2, CO2 and CO molecules present in the reaction

atmosphere to form potassium bicarbonates species [26, 43] as a source of stored H2

stored molecules. Finally, one could also suggest that K+ ions may react with H2 or

H2O adsorbed molecules to form potassium hydride or hydroxide [44, 45].

Looking at the ratio of the different desorbed molecules during the last step of

the experiment, the formation of either methoxy or bicarbonate species cannot be

considered as the main H2 storage mechanism because much higher CO and CO2

formation rates, according to stoichiometric factors, should then be observed during

the final positive polarization. Furthermore, the number of K+ ions transferred to the

solid electrolyte during the imposition of -20 µA for 45 min (5.6 x 10-7

mol K+) was

one order of magnitude lower than the amount of released H2 (4.68 x 10-6

mol H2) to

the gas phase. Hence, potassium hydride or hydroxide can also be diminished as the

main source of the H2 storage. Then, it seems to indicate that the main plausible

mechanism of H2 storage was the hydrogen chemisorption on either Ni actives sites or

other surface compounds (e.g., carbonaceous deposits).

Before going deeper into the mechanism of H2 storage, a second experiment

was carry out in order to confirm the reproducibility of this process (Figure 6.3). The

only modification in this experiment with respect to that reflected in Figure 6.2 was

the application of a linear sweep voltammetry (LSV) in the final release step. This

linear voltammetry was performed from -0.6 to +2 V at a scan rate of 1 mV s-1

under

Ar atmosphere (instead of the imposition of a fixed VWR = +2 V).

Electrochemically assisted production and storage of hydrogen

233

Figure 6.3. Variation of H2, CO and CO2 molar flow rates (a), current (I) and catalyst potential

(VWR) (b) with catalyst film Ni 0º during the following steps at 280ºC: H2 production/storage,

CH3OH/H2O = 4.4 %/5.2 %, I = -20 µA; Cleaning, 100 % Ar, OC; H2 release, 100 % Ar, LSV

from -0.6 V to +2 V (1 mV s-1

).

It should be mentioned that almost the same catalytic activity and Ni

deactivation trend were observed during the imposition of -20 µA for 45 min in both

Figures 6.2 and 6.3. Very similar amounts of transferred K+ ions and released H2

molecules were also obtained during the LSV, which confirms the good

reproducibility of both catalytic and electrocatalytic activity of Ni under the studied

reaction conditions. The only difference with respect to the previous experiment was

0

1

2

3

r /

mo

l s-1

x 10

-8

-1

0

1

2

-3

-1

1

3

5

0 20 40 60 80 100

VW

R/

V

I / A

x 1

0-5

Time / min

-20 µA4.4 % CH3OH + 5.2 % H2O

From -0.6 V to +2 VAr

OCAr

rH2

rCO

rCO2

VWR

I

a)

b)

Chapter 6

234

that, as the potential was linearly increased, the removal of K+ ions (according to the I

vs. time curve) and the H2 release took place in a gradual way. Hence, this operational

procedure was selected for further experiments since it allowed a better quantification

of the H2 release through the periodic analysis of the GC.

Moreover, the release peaks in both figures (6.2 and 6.3) seemed to behave in

a similar way than the positive current vs. time peaks (which were related to the

respective transfer of K+ ions back to the solid electrolyte). This means that the K

+

ions had a direct implication on both the H2 storage and its release.

6.3.2. Influence of the applied negative polarization and the reaction atmosphere

Different experiments were performed at 280 ºC with both kinds of Ni

catalysts, i.e., the normal (Ni 0º) and oblique (Ni 80º) angle deposited films, by

varying in each experiment the negative current applied during the H2 storage step.

Figures 6.4a and 6.4b show the time variation of the H2 molar flow rate and the

current (I) obtained for both catalysts during the release step (positive linear sweep

voltammetry under Ar stream) after imposition of three different negative currents

during the storage step (under a feed composition of CH3OH/H2O = 4.4 %/5.2 %).

It should be noted that under negative polarization and reaction conditions (not

shown) a higher catalytic activity and a more pronounced deactivation were observed

on the oblique angle deposited Ni if compared to those of catalyst Ni 0º, which is in

agreement with the higher porosity of the former (35 %). This different microstructure

was probably also a determining factor in the larger amount of released H2 observed

for the case of catalyst Ni 80º. From the integration of the H2 release peaks one can

calculate a maximum value of 6.99 x 10-5

mol H2 stored after 45 min at -30 µA for

catalyst Ni 80º (Figure 6.4b-1) vs. 5.3 x 10-6

mol H2 under analogous conditions for

catalyst Ni 0º (Figure 6.4a-1). Hence, one can observe an increase of one order of

magnitude in the H2 storage capacity of the more porous catalyst film (Ni 80 º). In

addition, the H2 release from the catalyst Ni 80º was extended in time (some hours). A

potential of VWR = +2 V was thus applied at the end of the linear sweep voltammetry

in this case to complete the depletion of H2.

Electrochemically assisted production and storage of hydrogen

235

Figure 6.4. Variation of H2 molar flow rate and current (I) with catalyst films Ni 0º (a-1 and a-

2) and Ni 80º (b-1 and b-2) at 280 ºC during the H2 release step (100 % Ar, LSV from -0.6 V to

+2 V, 1 mV s-1

) after the imposition of different negative currents for 45 min during the H2

production/storage step (CH3OH/H2O = 4.4 %/5.2 %).

As in the previous figures (6.2 and 6.3), in all these experiments a much lower

amount of transferred K+ ions was obtained if compared to that of released H2, and

both CO and CO2 releases (not depicted) were again almost negligible. Apart from

certain possible formation of adsorbed methoxy species, potassium hydride, hydroxide

or bicarbonates, the main storage mechanism seemed to be the chemisorption of

atomic hydrogen on Ni and carbonaceous deposits. At this point, it is important to

remark that the H to Ni ratio obtained with catalyst Ni 80º was quite high (up to 11,

after 45 min at -30 µA), thus suggesting that the worthy H2 storage results were rather

due to the spillover of H atoms on carbon compounds formed under SRM conditions.

Furthermore, one can appreciate from this figure a very strong effect of the

transfer of K+ ions during the storage step (negative current supply) on the process

performance. The application of more negative currents during the H2

0

2

4

6

8

10

12

r H

/ m

ol s

-1x

10-9

-1

0

1

2

-1

2

5

8

0 10 20 30 40

VW

R/

V

I / A

x 1

0-5

Time / min

After -10 µA

After -20 µA

After -30 µA

I after -10 µA

I after -20 µA

I after -30 µA

a-1)

2

VWR

-1

0

1

2

-1

2

5

8

0 50 100 150 200 250

VW

R/

V

I / A

x 1

0-5

Time / min

0

3

6

9

12

15

18

21

r H

/ m

ol s

-1x

10-9

After -10 µA

After -20 µA

After -30 µA

I after -10 µA

I after -20 µA

I after -30 µA

b-1)

2

VWR

a-2) b-2)

Ni 0º Ni 80º

Chapter 6

236

production/storage step caused a remarkable subsequent release of H2 up to certain

saturation value (similar hydrogen storage results were obtained after the application

of either -20 µA or -30 µA). As expected, higher positive current values were also

observed during the linear sweep voltammetries (Figures 6.4a-2 and 6.4b-2) after

higher negative polarizations. This information has been summarized in Figure 6.5. It

shows the results obtained with both Ni catalysts (Ni 0º and Ni 80º) as a function of

the negative current applied during the H2 production/storage step, in terms of total

hydrogen released (in units of g H2 x 100 g Ni-1

) during the subsequent LSV.

Figure 6.5. Influence of the negative current (I) applied for 45 min during the H2

production/storage step (CH3OH/H2O = 4.4 %/5.2 %) on the H2 storage capacity of catalyst

films Ni 0º and Ni 80º at 280 ºC, measured during the subsequent H2 release steps (100 % Ar,

LSV from -0.6 V to +2 V, 1 mV s-1

).

From both Figures 6.4 and 6.5, one can state that there is a clear enhancement

effect of K+ ions on the H2 storage capacity of the Ni catalysts. As discussed in the

previous chapter, the presence of alkali (K+) ions on Ni likely activated the CH3OH

adsorption and successive dehydrogenation toward H2 and CO (reactions 6.3-6.8).

CH3OH → CH3OH(a) (6.3)

CH3OH(a) → CH3O(a) + H(a) (6.4)

CH3O(a) → H2CO(a) + H(a) (6.5)

0

3

6

9

12

15

18

21

After -10 µA After -20 µA After -30 µA

Tota

l H

2re

leas

ed

/ g

H2

x 10

0 g

Ni-1

Ni 0º

Ni 80º

Electrochemically assisted production and storage of hydrogen

237

H2CO(a) → CO(a) + 2H(a) (6.6)

2H(a) → H2 (6.7)

CO(a) → CO (6.8)

Then, potassium ions would also favor the subsequent dissociation of CO

molecules and hence the carbon deposition (reaction 6.9), as typically reported in

other EPOC [46, 47] and classical alkali promotion [48, 49] studies.

CO(a) → C(a) + O(a) (6.9)

K+ ions could also stabilize the hydrogen adsorbed on these carbon compounds

as observed in other studies where alkali ions were intercalated into graphitic layers

[50-52]. This would explain the similarity observed in all the experiments, especially

with catalyst Ni 0º, between the H2 release and the positive current peaks.

This way, the increase in the H2 storage capacity of the Ni catalyst films upon

increasing the applied negative current under SRM conditions could be attributed to a

triple promoting effect of potassium ions: 1) the promotion of the methanol

dehydrogenation reaction, 2) the increase in the formation of surface carbon

compounds which acted as hydrogen adsorbents and 3) the hydrogen chemisorption

and stabilization on the carbonaceous stored compounds. This storage mechanism

would also explain the slower H2 release rate observed in Figure 6.4b-1 after the

application of more negative currents. A reverse spillover of the H adatoms toward the

Ni active sites would be necessary before these H atoms can be recombined and

desorbed to the gas phase as H2 molecules [53]. This H back-spillover would be

slowed since the increase in the K+-assisted formation of carbonaceous compounds

would make the H adsorption to take place on less accessible sites.

In contrast to previous studies where NOx [14, 15] or CO2 [16] were captured

by using K+-conductor solid electrolytes through the formation of some nitrates or

carbonates, respectively, in the present study K+ ions should barely be bonded with the

stored compound (H2). Instead, these K+ ions could be involved in the formation of an

adsorbent carbonaceous material and the subsequent H2 adsorption.

Chapter 6

238

Regarding the influence of the Ni catalyst microstructure on the H2 storage

capacity, Figure 6.5 depicts a great difference between both catalyst films. Whilst

catalyst Ni 0º stored/released up to 0.5 g H2 x 100 g Ni-1

, catalyst Ni 80º reached a

maximum H2 storage capacity of around 19 g H2 x 100 g Ni-1

. This increase was

attributed to the higher porosity of the latter (35 %) and its paramount role in the

formation of larger carbon deposits on which hydrogen could be adsorbed. The

potential interest of the oblique angle deposited catalysts for the H2 storage in form of

MgH2 was previously reported for V-decorated and V-doped Mg nanostructures [21,

22]. In these studies the H2 adsorption and desorption temperatures generally

decreased upon increasing the zenithal deposition angle [22], and the theoretical

hydrogen storage capacity for MgH2 (7.6 wt. %) was reached at around 280 ºC [21].

Herein, the deposition of a Ni catalyst film at an incident angle of α = 80º

allowed to obtain apparent H2 storage capacities (up to 19 wt. %) which were excellent

as compared to those shown, for instance, with Mg2NiH4 (3.6 wt. %), other hydrides

of Ni-containing alloys working at lower operation temperatures such as CoNi5H4 or

LaNi5H6 (less than 2 wt. %), and even the best-performing hydrides, i.e., the

borohydrides (around 18 wt. %), which typically require decomposition temperatures

up to 600 ºC [1, 5, 7]. However, this comparison must be carried out with caution. The

weight percentage values of stored H2 shown in Figure 6.5 were related to the amount

of Ni deposited, since this parameter was quantified by X-ray fluorescence

spectroscopy and remained, regardless the operation conditions, unalterable.

In order to clarify the mechanism of the adsorbed species responsible for the

hydrogen trapping, additional experiments were carried out with catalyst Ni 80º, in

which H2 was stored under the application of a potential of VWR = -1 V for 1 hour and

different feed compositions: methanol steam reforming conditions (4.4 % CH3OH, 5.2

% H2O), methanol decomposition conditions (only 4.4 % CH3OH, i.e., in absence of

water) and a 2.3 % H2 stream, all of them at 280 ºC and with an overall flow rate of 6

L h-1

(Ar balance). Figure 6.6 shows the variation in the released H2 flow rate and the

current during the last step of these experiments, consisting of a positive LSV from -1

to +2 V (scan rate = 1 mV s-1

) and maintenance of the latter potential.

Electrochemically assisted production and storage of hydrogen

239

Figure 6.6. Variation of H2 molar flow rate (a) and current (I) (b) with catalyst film Ni 80º at

280 ºC during the H2 release step (100 % Ar, LSV from -1 V to +2 V, 1 mV s-1

) following the

H2 production/storage step (VWR = -1 V) under different atmospheres.

Regarding the LSV performed after the negative polarization under the 2.3 %

H2 stream, a small peak can be detected in Figure 6.6a that corresponded to 4 x 10-7

mol H2 released. This peak was quite synchronized with the positive current peak

(Figure 6.6b), which corresponded to an amount of transferred K+ ions of the same

order of magnitude (5 x 10-7

mol K+), thus suggesting that some H2 was stored during

the negative polarization step by interacting with the K+ ions. However, the H2 release

peak detected in this experiment was still negligible compared to those obtained after

the negative polarization step carried out in CH3OH or CH3OH+H2O atmospheres.

-1.5

-0.5

0.5

1.5

-1

2

5

0 50 100 150 200 250 300 350

VW

R/

V

I / A

x 1

0-5

Time / min

-1.5

-0.5

0.5

1.5

2.5

0

2

4

6

8

10

VW

R/

V

r H

/ m

ol s

-1x

10

-9

a)

b)

After H2

After CH3OHAfter CH3OH + H2O

VWR

After H2

After CH3OHAfter CH3OH + H2O

VWR

2

Chapter 6

240

These findings are in good agreement with the proposed mechanism of H2 storage

based on a chemical adsorption of H atoms on carbon compounds, which were formed

during the SRM reaction under the effect of K+ supply. Furthermore, the total amount

of obtained H2 was very similar in both the absence and the presence of water (6.98 x

10-5

and 6.82 x 10-5

mol H2, respectively), the release process being faster in the later

case. Thus one could state that both the hydrogen and the carbonaceous adsorbent

were derived from the methanol itself, whereas water would facilitate the H spillover

process through the formation of surface hydroxyl groups [10, 53, 54].

The spillover-assisted hydrogen storage has been reported in literature for

many different carbon-based or metal oxide adsorbents which acted as catalyst support

[8, 10, 11, 13]. On the other hand, hydrogen is typically supplied either from an

aqueous solution (electrochemically, i.e., as H+) [9, 55-58] or from the gas phase (as

H2) [53, 59-68]. Group VIII metals such as Ni [67, 68], Pt [53, 59-63] or Pd [62-66]

are commonly employed catalysts in the latter case due to their good activity in

hydrogen dissociation. In the present work, both hydrogen and carbon-based adsorbent

were obtained from the very process of methanol conversion in a single reaction step

on Ni catalysts, which was assisted by the promotional effect of K+ ions. Hence, the

coupling of solid-state electrochemistry and catalysis, along with the employment of

this novel kind of oblique angle deposited catalyst films with high porosity, allowed to

obtain vey high yields of H2 storage per unit area and amount of metal. This would

represent an alternative to the H2 capture classically performed in metal hydrides with

high decomposition temperatures such as MgH2, alanates or borohydrides.

Summarizing, the proposed Ni/K-βAl2O3 electrocatalytic system shows the

potential application of EPOC for the in-situ controlled storage and release of

hydrogen under mild reaction conditions (280 ºC). The obtained highly pure hydrogen

could be then conditioned and used by either fuel cells or other catalytic processes.

The oblique angle deposited Ni catalyst film (Ni 80º), which showed a

superior performance for H2 production and storage, was characterized by several

techniques. Figure 6.7 shows SEM images of a selected area from catalyst Ni 80º,

after being exposed to the application of VWR = -1 V for 1 hour under SRM conditions.

Electrochemically assisted production and storage of hydrogen

241

Figure 6.7. Top view SEM image of a selected area (a) of catalyst Ni 80º after the H2

production/storage experiments and the final application of VWR = -1 V for 1 h (280 ºC, 4.4 %

CH3OH, 5.2 % H2O), along with different magnifications (b and c).

Nickel nanocolumns can be observed in these micrographs, grouped and

distributed on the K-βAl2O3 support in a quite similar way than those observed in the

10 µm

a)

5 µm

b)

2 µm

c)

Chapter 6

242

fresh sample (see Figure 6.1d). This seems to indicate that the Ni catalyst film did not

suffer significant modifications during the H2 production and storage experiments,

demonstrating the good stability of the oblique angle deposited (OAD) catalyst films.

Then, additional characterization measurements were carried out to identify

the deposited carbonaceous molecules and the storage process that takes place.

6.3.3. Investigation of possible surface compounds by catalyst characterization

Figure 6.8 shows other SEM micrographs of catalyst film Ni 80º after

exposure to methanol steam reforming conditions (CH3OH/H2O = 4.4 %/5.2 %) and

the application of VWR = -1 V.

Figure 6.8. Top view SEM images of selected areas of catalyst Ni 80º after the H2

production/storage experiments and the final application of VWR = -1 V for 1 h (280 ºC, 4.4 %

CH3OH, 5.2 % H2O).

5 µm

b)

10 µm

a)

Electrochemically assisted production and storage of hydrogen

243

In these SEM images one can observe black regions of pillared fragments

corresponding to carbonaceous deposits covering the Ni surface. These compounds

will be later identified mainly as graphene oxide (GO), which could fit this two-

dimensional appearance layerwise. In this micrographs, single thin GO sheets cannot

be appreciated, as those observed in other studies where graphene oxide was obtained

from graphite oxidation and exfoliation [69, 70]. Nevertheless, highly concentrated

graphene oxide typically exhibits a wrinkle and overlapped structure similar to that

shown in this figure [70, 71].

Figure 6.9 shows the XPS spectra obtained from catalyst film Ni 80º after the

exposition to the mentioned reaction conditions, along with a quantitative estimation

of the elemental concentration, relative to the main component, i.e., Ni, in the

examined outer layer.

Figure 6.9. XPS spectra (a) of catalyst Ni 80º after the H2 production/storage experiments and

the final application of VWR = -1 V for 1 h (280 ºC, 4.4 % CH3OH, 5.2 % H2O), and the

corresponding surface atomic composition relative to Ni (b).

80 70 60

Co

un

ts / a

.u.

Binding energy / eV

880 870 860 850 840

Co

un

ts / a

.u.

Binding energy / eV

540 535 530 525

Co

un

ts / a

.u.

Binding energy / eV

420 410 400 390C

ou

nts

/ a

.u.

Binding energy / eV

290 280 270

Co

un

ts / a

.u.

Binding energy / eV

O 1sNi 2p

Al 2p + Ni 3p

Ni LMM

C 1s + K 2p

at. K/Ni at. O/Ni at. C/Ni at. Al/Ni1.26 9.16 4.34 0.09

a)

b)

Chapter 6

244

The peak observed at 284.7 eV corresponded to C 1s and the K 2p region

presented two contributions at 292.4 eV (K 2p3/2) and 295 eV (K 2p1/2). Interestingly,

other peak was found in the intermediate region (at 289.4 eV), which was assigned to

the presence of potassium carbonates on the catalyst surface [72, 73]. These species

would be formed under negative polarization due to the interaction between the K+

ions and the molecules present in the reaction mixture as discussed in the previous

section. Their decomposition would probably lead to some CO and CO2 release during

the positive polarization step (see Figure 6.2 or 6.3). On the other hand, the Ni outer

surface showed to be at least partially oxidized, since the peaks observed at 531 eV (O

1s region), at 855.1 eV and 860.9 eV (Ni 2p region) can be attributed to the presence

of Ni(OH)2 or Ni2O3 [74, 75]. Moreover, as also observed in chapter 4, a high amount

of carbon was found on the Ni electrochemical catalyst, which covered the catalyst

surface (C/Ni atomic ratio of 4.34). This finding would agree with the H2 storage

mechanism through H spillover on carbonaceous deposits proposed.

Figure 6.10 shows the RAMAN spectrum of catalyst film Ni 80º exposed to

the same reaction conditions as those of previous figures.

Figure 6.10. Raman spectrum of catalyst Ni 80º after the H2 production/storage experiments

and the final application of VWR = -1 V for 1 h (280 ºC, 4.4 % CH3OH, 5.2 % H2O).

800 1600 2400 3200 4000

Inte

nsity /

a.u

.

Raman shift / cm-1

Electrochemically assisted production and storage of hydrogen

245

A narrow peak at 1072 cm-1

can be assigned to the ν(CO32-

) symmetric stretch

[76, 77], in good agreement with the carbonate species previously detected by XPS.

However, the most relevant finding was the presence of strong disordered (D) and

graphitic (G) bands at 1343 cm-1

and 1597 cm-1

, respectively, which are characteristic

of graphene oxide (GO) [59, 78, 79], along with less intense 2D bands at 2500-3200

cm-1

. The ratio of intensities of the D and G peaks (ID/IG) was lower than the unity,

which suggested a quite ordered crystalline structure.

Finally, Figure 6.11 shows the corresponding FTIR spectrum of the catalyst

film Ni 80º in the range from 750 to 1900 cm-1

. No clear bands were found at higher

wavenumbers, associated with ν(CH3) or ν(OH) stretching vibrations (not shown).

Figure 6.11. FTIR spectrum of catalyst Ni 80º after the H2 production/storage experiments and

the final application of VWR = -1 V for 1 h (280 ºC, 4.4 % CH3OH, 5.2 % H2O).

The following vibrational modes were identified in this spectrum: C=O stretch

at 1650 cm-1

, adsorbed water and skeletal vibrations of unoxidized graphitic domains

(C=C) at 1627 cm-1

, COO- symmetric stretch at 1366 cm

-1, epoxide stretch at 970 cm

-1

and C-H out-of-plane wag at 831 cm-1

. The presence of these functional groups would

be in good agreement with the formation of grapheme oxide, as also observed in FTIR

spectra reported elsewhere [79, 80]. Indeed, according to the well-accepted Lerf-

1800 1600 1400 1200 1000 800

Ab

so

rba

nce

/ a

.u.

Wavenumbers / cm-1

Chapter 6

246

Klinowski model of graphene oxide [81], this compound consists of aromatic units of

variable size which are separated from each other by aliphatic 6-membered rings,

where the basal plane contains atomic oxygen in the form of hydroxyl (C-OH) or

epoxide (C-O-C) groups. The edges of graphite and graphene oxide sheets and

oxidized carbons can also contain other oxygen species such as C-OH, carbonyl

(C=O) and carboxyl (O=C-OH) groups.

This way, the presence of graphene oxide on the catalyst surface was

confirmed by both Raman and infrared spectroscopy. Some concurrence of potassium-

derived species such as bicarbonates cannot be ruled out. However, according to the

analysis carried out, graphene oxide can be postulated as the main adsorbent involved

in the H2 storage mechanism discussed in the previous section. It is then supposed that

the hydroxyl and oxygen groups present on the graphene oxide favoured the spillover

of H atoms [13, 53, 54], which would specially adsorb at the edge sites of the layers

and in the vicinity of these functional groups [12, 13, 61, 82].

6.4. CONCLUSIONS

The following conclusions can be drawn from this study:

- The phenomenon of electrochemical promotion (EPOC) was for the first time

applied in the H2 production and storage processes, by using oblique angle deposited

Ni catalyst films and a K+-conductor support.

- H2 was simultaneously produced from methanol steam reforming (SRM) and

stored under the promotional effect of K+ ions (cathodic polarization). Then, hydrogen

was released with high purity upon removing the K+ ions from the catalyst film

(anodic polarization).

- A higher H2 release was obtained by increasing both the negative current

applied during the H2 production/storage step and the porosity of the Ni catalyst film

(which was related to the incident angle during the catalyst deposition). The presence

of methanol in the reaction atmosphere showed to be needed to obtain an appreciable

amount of released H2. On the other hand, in view of the amounts of transferred K+

ions, and those of released H2, CO and CO2 during the different experiments, the

Electrochemically assisted production and storage of hydrogen

247

direct formation of potassium-derived compounds (hydride, hydroxide, bicarbonates)

or the stabilization of some reaction intermediate (like methoxy) were not considered

as the main sources of the observed released H2.

- Regarding the obtained catalytic results and different ex-situ characterization

measurements, a H2 storage mechanism was proposed based on the hydrogen

chemisorption on both Ni active sites and graphene oxide, which was deposited under

SRM electropromoted conditions, through a spillover process. Hence, a triple

promoting effect of K+ ions would take place: 1) promotion of the methanol

dehydrogenation, 2) promotion of the formation of graphene oxide, and 3) promotion

of the hydrogen chemisorption on graphene oxide.

- A very high H2 storage capacity per amount of metal was obtained (19 g H2 x

100 g Ni-1

with the oblique angle deposited Ni catalyst film), which shows a very

promising application of this kind of electrocatalytic systems and the EPOC

phenomenon.

6.5. REFERENCES

[1] L. Schlapbach, A. Züttel, Hydrogen-storage materials for mobile applications, Nature, 414

(2001) 353-358.

[2] S. Dutta, A review on production, storage of hydrogen and its utilization as an energy

resource, Journal of Industrial and Engineering Chemistry, 20 (2014) 1148-1156.

[3] J.N. Armor, Catalysis and the hydrogen economy, Catalysis Letters, 101 (2005) 131-135.

[4] J.D. Holladay, J. Hu, D.L. King, Y. Wang, An overview of hydrogen production

technologies, Catalysis Today, 139 (2009) 244-260.

[5] A.F. Dalebrook, W. Gan, M. Grasemann, S. Moret, G. Laurenczy, Hydrogen storage:

Beyond conventional methods, Chemical Communications, 49 (2013) 8735-8751.

[6] P. Makowski, A. Thomas, P. Kuhn, F. Goettmann, Organic materials for hydrogen storage

applications: From physisorption on organic solids to chemisorption in organic molecules,

Energy and Environmental Science, 2 (2009) 480-490.

[7] A.W.C. Van Den Berg, C.O. Areán, Materials for hydrogen storage: Current research

trends and perspectives, Chemical Communications, (2008) 668-681.

Chapter 6

248

[8] L. Wang, R.T. Yang, Hydrogen storage on carbon-based adsorbents and storage at

ambient temperature by hydrogen spillover, Catalysis Reviews - Science and Engineering, 52

(2010) 411-461.

[9] S. Mukherjee, B. Ramalingam, S. Gangopadhyay, Hydrogen spillover at sub-2 nm Pt

nanoparticles by electrochemical hydrogen loading, Journal of Materials Chemistry A, 2

(2014) 3954-3960.

[10] W.C. Conner Jr, J.L. Falconer, Spillover in heterogeneous catalysis, Chemical Reviews,

95 (1995) 759-788.

[11] R. Prins, Hydrogen spillover. Facts and fiction, Chemical Reviews, 112 (2012) 2714-

2738.

[12] S. Gadipelli, Z.X. Guo, Graphene-based materials: Synthesis and gas sorption, storage

and separation, Progress in Materials Science, 69 (2015) 1-60.

[13] G.M. Psofogiannakis, G.E. Froudakis, Fundamental studies and perceptions on the

spillover mechanism for hydrogen storage, Chemical Communications, 47 (2011) 7933-7943.

[14] A. de Lucas-Consuegra, Á. Caravaca, P. Sánchez, F. Dorado, J.L. Valverde, A new

improvement of catalysis by solid-state electrochemistry: An electrochemically assisted NOx

storage/reduction catalyst, Journal of Catalysis, 259 (2008) 54-65.

[15] A. de Lucas-Consuegra, A. Caravaca, M.J. Martín de Vidales, F. Dorado, S. Balomenou,

D. Tsiplakides, P. Vernoux, J.L. Valverde, An electrochemically assisted NOx

storage/reduction catalyst operating under fixed lean burn conditions, Catalysis

Communications, 11 (2009) 247-251.

[16] E. Ruiz, D. Cillero, Á. Morales, G.S. Vicente, G. De Diego, P.J. Martínez, J.M. Sánchez,

Bench scale study of electrochemically promoted CO2 capture on Pt/K-βAl2O3,

Electrochimica Acta, 112 (2013) 967-975.

[17] M. Matthew, M.T. Hawkeye, M.J.B. Taschuk, Glancing Angle Deposition of Thin Films:

Engineering the Nanoscale, Wiley2014.

[18] V. Rico, P. Romero, J.L. Hueso, J.P. Espinós, A.R. González-Elipe, Wetting angles and

photocatalytic activities of illuminated TiO2 thin films, Catalysis Today, 143 (2009) 347-354.

[19] V. Rico, A. Borrás, F. Yubero, J.P. Espinós, F. Frutos, A.R. González-Elipe, Wetting

angles on illuminated ta2o5 thin films with controlled nanostructure, Journal of Physical

Chemistry C, 113 (2009) 3775-3784.

[20] D.X. Ye, Y.P. Zhao, G.R. Yang, Y.G. Zhao, G.C. Wang, T.M. Lu, Manipulating the

column tilt angles of nanocolumnar films by glancing-angle deposition, Nanotechnology, 13

(2002) 615-618.

Electrochemically assisted production and storage of hydrogen

249

[21] Y. He, Y. Zhao, Hydrogen storage and cycling properties of a vanadium decorated Mg

nanoblade array on a Ti coated Si substrate, Nanotechnology, 20 (2009).

[22] Y. He, J. Fan, Y. Zhao, The role of differently distributed vanadium nanocatalyst in the

hydrogen storage of magnesium nanostructures, International Journal of Hydrogen Energy, 35

(2010) 4162-4170.

[23] N.N. Kariuki, W.J. Khudhayer, T. Karabacak, D.J. Myers, GLAD Pt-Ni alloy nanorods for

oxygen reduction reaction, ACS Catalysis, 3 (2013) 3123-3132.

[24] J.R. Sánchez-Valencia, A. Borrás, A. Barranco, V.J. Rico, J.P. Espinos, A.R. González-

Elipe, Preillumination of TiO2 and Ta2O5 photoactive thin films as a tool to tailor the

synthesis of composite materials, Langmuir, 24 (2008) 9460-9469.

[25] C.G. Vayenas, S. Bebelis, C. Pliangos, S. Brosda, D. Tsiplakides, Electrochemical

Activation of Catalysis: Promotion, Electrochemical Promotion and Metal-Support

Interactions, Kluwer Academic Publishers/Plenum Press, New York, 2001.

[26] A. de Lucas-Consuegra, F. Dorado, J.L. Valverde, R. Karoum, P. Vernoux, Low-

temperature propene combustion over Pt/K-βAl2O3 electrochemical catalyst:

Characterization, catalytic activity measurements, and investigation of the NEMCA effect,

Journal of Catalysis, 251 (2007) 474-484.

[27] N. Iwasa, M. Yoshikawa, W. Nomura, M. Arai, Transformation of methanol in the

presence of steam and oxygen over ZnO-supported transition metal catalysts under stream

reforming conditions, Applied Catalysis A: General, 292 (2005) 215-222.

[28] B.S. Chen, J.L. Falconer, Alcohol Decomposition by Reverse Spillover, Journal of

Catalysis, 144 (1993) 214-226.

[29] N. Takezawa, N. Iwasa, Steam reforming and dehydrogenation of methanol: Difference in

the catalytic functions of copper and group VIII metals, Catalysis Today, 36 (1997) 45-56.

[30] M.T. Tavares, I. Alstrup, C.A. Bernardo, J.R. Rostrup-Nielsen, Carbon formation and CO

methanation on silica-supported nickel and nickel-copper catalysts in CO + H2 mixtures,

Journal of Catalysis, 158 (1996) 402-410.

[31] Z. Murayama, I. Kojima, E. Miyazaki, I. Yasumori, Dissociation of carbon monoxide on

stepped nickel surfaces, Surface Science, 118 (1982) L281-L285.

[32] A. Serrano-Lotina, L. Daza, Long-term stability test of Ni-based catalyst in carbon dioxide

reforming of methane, Applied Catalysis A: General, 474 (2014) 107-113.

[33] J.L. Pinilla, I. Suelves, M.J. Lázaro, R. Moliner, J.M. Palacios, Parametric study of the

decomposition of methane using a NiCu/Al 2O3 catalyst in a fluidized bed reactor,

International Journal of Hydrogen Energy, 35 (2010) 9801-9809.

Chapter 6

250

[34] H.F. Abbas, W.M.A. Wan Daud, Hydrogen production by methane decomposition: A

review, International Journal of Hydrogen Energy, 35 (2010) 1160-1190.

[35] V. Jiménez, A. Nieto-Márquez, J.A. Díaz, R. Romero, P. Sánchez, J.L. Valverde, A.

Romero, Pilot plant scale study of the influence of the operating conditions in the production

of carbon nanofibers, Industrial and Engineering Chemistry Research, 48 (2009) 8407-8417.

[36] P. Wei, W. Xia, J.Z. Li, H. Long, J. Chen, T. Li, M. Fan, Single-phase Ni3Sn alloy alkali-

leached for hydrogen production from methanol decomposition, Renewable Energy, 78 (2015)

357-363.

[37] J.H. Jang, Y. Xu, D.H. Chun, M. Demura, D.M. Wee, T. Hirano, Effects of steam addition

on the spontaneous activation in Ni3Al foil catalysts during methanol decomposition, Journal

of Molecular Catalysis A: Chemical, 307 (2009) 21-28.

[38] G. Wu, C. Zhang, S. Li, Z. Han, T. Wang, X. Ma, J. Gong, Hydrogen production via

glycerol steam reforming over Ni/Al 2O3: Influence of nickel precursors, ACS Sustainable

Chemistry and Engineering, 1 (2013) 1052-1062.

[39] R. Carrera Cerritos, R. Fuentes Ramírez, A.F. Aguilera Alvarado, J.M. Martínez Rosales,

T. Viveros García, I.R. Galindo Esquivel, Steam reforming of ethanol over Ni/Al2O3-La 2O3

catalysts synthesized by sol-gel, Industrial and Engineering Chemistry Research, 50 (2011)

2576-2584.

[40] J. Raskó, J. Bontovics, F. Solymosi, FTIR study of the interaction of methanol with clean

and potassium-doped Pd SiO2 catalysts, Journal of Catalysis, 146 (1994) 22-33.

[41] J.T. Miller, B.L. Meyers, F.S. Modica, G.S. Lane, M. Vaarkamp, D.C. Koningsberger,

Hydrogen Temperature-Programmed Desorption (H2 TPD) of Supported Platinum Catalysts,

Journal of Catalysis, 143 (1993) 395-408.

[42] F. Solymosi, A. Berkó, Z. Tóth, Adsorption and dissociation of CH3OH on clean and K-

promoted Pd(100) surfaces, Surface Science, 285 (1993) 197-208.

[43] P. Vernoux, F. Gaillard, C. Lopez, E. Siebert, In-situ electrochemical control of the

catalytic activity of platinum for the propene oxidation, Solid State Ionics, 175 (2004) 609-613.

[44] V.V. Simonyan, J.K. Johnson, Hydrogen storage in carbon nanotubes and graphitic

nanofibers, Journal of Alloys and Compounds, 330-332 (2002) 659-665.

[45] R.T. Yang, Hydrogen storage by alkali-doped carbon nanotubes-revisited, Carbon, 38

(2000) 623-626.

[46] C.G. Vayenas, S. Neophytides, Non-faradaic electrochemical modification of catalytic

activity III. The case of methanol oxidation on Pt, Journal of Catalysis, 127 (1991) 645-664.

Electrochemically assisted production and storage of hydrogen

251

[47] E. Ruiz, D. Cillero, P.J. Martínez, Á. Morales, G.S. Vicente, G. De Diego, J.M. Sánchez,

Bench-scale study of electrochemically assisted catalytic CO2 hydrogenation to hydrocarbon

fuels on Pt, Ni and Pd films deposited on YSZ, Journal of CO2 Utilization, 8 (2014) 1-20.

[48] J.E. Crowell, G.A. Somorjai, The effect of potassium on the chemisorption of carbon

monoxide on the Rh(111) crystal face, Applied Surface Science, 19 (1984) 73-91.

[49] R.W. Joyner, R.A.v. Santen, Elementary Reaction Steps in Heterogeneous Catalysis,

Springer, Netherlands, 2012.

[50] H. Zabel, S.A. Solin, Graphite Intercalation Compounds II: Transport and Electronic

Properties, Springer-Verlag, Berlin Heidelberg, 2013.

[51] W.Q. Deng, X. Xu, W.A. Goddard, New alkali doped pillared carbon materials designed

to achieve practical reversible hydrogen storage for transportation, Physical Review Letters,

92 (2004) 166103-166101.

[52] M. Inagaki, O. Tanaike, Determining factors for the intercalation into carbon materials

from organic solutions, Carbon, 39 (2001) 1083-1090.

[53] W.J. Ambs, M.M. Mitchell Jr, Hydrogen spillover on platinum-alumina, effect of water,

Journal of Catalysis, 82 (1983) 226-229.

[54] Q. Li, A.D. Lueking, Effect of surface oxygen groups and water on hydrogen spillover in

pt-doped activated carbon, Journal of Physical Chemistry C, 115 (2011) 4273-4282.

[55] R. Krishna, E. Titus, L.C. Costa, J.C.J.M.D.S. Menezes, M.R.P. Correia, S. Pinto, J.

Ventura, J.P. Araújo, J.A.S. Cavaleiro, J.J.A. Gracio, Facile synthesis of hydrogenated reduced

graphene oxide via hydrogen spillover mechanism, Journal of Materials Chemistry, 22 (2012)

10457-10459.

[56] Z. Dongping, J. Velmurugan, M.V. Mirkin, Adsorption/desorption of hydrogen on Pt

nanoelectrodes: Evidence of surface diffusion and spillover, Journal of the American Chemical

Society, 131 (2009) 14756-14760.

[57] D.V. Esposito, I. Levin, T.P. Moffat, A.A. Talin, H2 evolution at Si-based metal-

insulator-semiconductor photoelectrodes enhanced by inversion channel charge collection and

H spillover, Nature Materials, 12 (2013) 562-568.

[58] S. Sata, M.I. Awad, M.S. El-Deab, T. Okajima, T. Ohsaka, Hydrogen spillover

phenomenon: Enhanced reversible hydrogen adsorption/desorption at Ta2O5-coated Pt

electrode in acidic media, Electrochimica Acta, 55 (2010) 3528-3536.

[59] V.H. Pham, T.T. Dang, K. Singh, S.H. Hur, E.W. Shin, J.S. Kim, M.A. Lee, S.H. Baeck,

J.S. Chung, A catalytic and efficient route for reduction of graphene oxide by hydrogen

spillover, Journal of Materials Chemistry A, 1 (2013) 1070-1077.

Chapter 6

252

[60] F.H. Yang, A.J. Lachawiec Jr, R.T. Yang, Adsorption of spillover hydrogen atoms on

single-wall carbon nanotubes, Journal of Physical Chemistry B, 110 (2006) 6236-6244.

[61] P.C.H. Mitchell, A.J. Ramirez-Cuesta, S.F. Parker, J. Tomkinson, D. Thompsett,

Hydrogen spillover on carbon-supported metal catalysts studied by inelastic neutron

scattering. Surface vibrational states and hydrogen riding modes, Journal of Physical

Chemistry B, 107 (2003) 6838-6845.

[62] A.D. Lueking, R.T. Yang, Hydrogen spillover to enhance hydrogen storage - Study of the

effect of carbon physicochemical properties, Applied Catalysis A: General, 265 (2004) 259-

268.

[63] C.C. Huang, N.W. Pu, C.A. Wang, J.C. Huang, Y. Sung, M.D. Ger, Hydrogen storage in

graphene decorated with Pd and Pt nano-particles using an electroless deposition technique,

Separation and Purification Technology, 82 (2011) 210-215.

[64] A.J. Lachawiec Jr, G. Qi, R.T. Yang, Hydrogen storage in nanostructured carbons by

spillover: Bridge-building enhancement, Langmuir, 21 (2005) 11418-11424.

[65] E. Yoo, L. Gao, T. Komatsu, N. Yagai, K. Arai, T. Yamazaki, K. Matsuishi, T.

Matsumoto, J. Nakamura, Atomic hydrogen storage in carbon nanotubes promoted by metal

catalysts, Journal of Physical Chemistry B, 108 (2004) 18903-18907.

[66] L. Wang, F.H. Yang, R.T. Yang, M.A. Miller, Effect of surface oxygen groups in carbons

on hydrogen storage by spillover, Industrial and Engineering Chemistry Research, 48 (2009)

2920-2926.

[67] L. Zubizarreta, J.A. Menéndez, J.J. Pis, A. Arenillas, Improving hydrogen storage in Ni-

doped carbon nanospheres, International Journal of Hydrogen Energy, 34 (2009) 3070-3076.

[68] K.Y. Lin, W.T. Tsai, T.J. Yang, Effect of Ni nanoparticle distribution on hydrogen uptake

in carbon nanotubes, Journal of Power Sources, 196 (2011) 3389-3394.

[69] J. Kim, L.J. Cote, J. Huang, Two dimensional soft material: New faces of graphene oxide,

Accounts of Chemical Research, 45 (2012) 1356-1364.

[70] G. Sobon, J. Sotor, J. Jagiello, R. Kozinski, M. Zdrojek, M. Holdynski, P. Paletko, J.

Boguslawski, L. Lipinska, K.M. Abramski, Graphene Oxide vs. Reduced Graphene Oxide as

saturable absorbers for Er-doped passively mode-locked fiber laser, Optics Express, 20 (2012)

19463-19473.

[71] Y. Shao, J. Wang, M. Engelhard, C. Wang, Y. Lin, Facile and controllable

electrochemical reduction of graphene oxide and its applications, Journal of Materials

Chemistry, 20 (2010) 743-748.

Electrochemically assisted production and storage of hydrogen

253

[72] A. Iordan, M.I. Zaki, C. Kappenstein, Interfacial chemistry in the preparation of catalytic

potassium-modified aluminas, Journal of the Chemical Society, Faraday Transactions, 89

(1993) 2527-2536.

[73] M.E. Gálvez, S. Ascaso, P. Stelmachowski, P. Legutko, A. Kotarba, R. Moliner, M.J.

Lázaro, Influence of the surface potassium species in Fe-K/Al2O3 catalysts on the soot

oxidation activity in the presence of NOx, Applied Catalysis B: Environmental, 152-153 (2014)

88-98.

[74] A.P. Grosvenor, M.C. Biesinger, R.S.C. Smart, N.S. McIntyre, New interpretations of

XPS spectra of nickel metal and oxides, Surface Science, 600 (2006) 1771-1779.

[75] K.S. Kim, N. Winograd, X-ray photoelectron spectroscopic studies of nickel-oxygen

surfaces using oxygen and argon ion-bombardment, Surface Science, 43 (1974) 625-643.

[76] Y. Morizet, M. Paris, F. Gaillard, B. Scaillet, Carbon dioxide in silica-undersaturated

melt. Part I: The effect of mixed alkalis (K and Na) on CO2 solubility and speciation,

Geochimica et Cosmochimica Acta, 141 (2014) 45-61.

[77] C. Xu, R. Reed, J.P. Gorski, Y. Wang, M.P. Walker, The distribution of carbonate in

enamel and its correlation with structure and mechanical properties, Journal of Materials

Science, 47 (2012) 8035-8043.

[78] S.C. Ray, S.K. Bhunia, A. Saha, N.R. Jana, Graphene oxide (GO)/reduced-GO and their

composite with conducting polymer nanostructure thin films for non-volatile memory device,

Microelectronic Engineering, 146 (2015) 48-52.

[79] C.Y. Ho, C.C. Liang, H.W. Wang, Investigation of low thermal reduction of graphene

oxide for dye-sensitized solar cell counter electrode, Colloids and Surfaces A:

Physicochemical and Engineering Aspects, 481 (2015) 222-228.

[80] C. Galande, A.D. Mohite, A.V. Naumov, W. Gao, L. Ci, A. Ajayan, H. Gao, A.

Srivastava, R. Bruce Weisman, P.M. Ajayan, Quasi-molecular fluorescence from graphene

oxide, Scientific Reports, 1 (2011).

[81] A. Lerf, H. He, M. Forster, J. Klinowski, Structure of graphite oxide revisited, Journal of

Physical Chemistry B, 102 (1998) 4477-4482.

[82] M. Kayanuma, U. Nagashima, H. Nishihara, T. Kyotani, H. Ogawa, Adsorption and

diffusion of atomic hydrogen on a curved surface of microporous carbon: A theoretical study,

Chemical Physics Letters, 495 (2010) 251-255.

Chapter 6

254

255

Chapter 7

GENERAL CONCLUSIONS AND RECOMMENDATIONS

This chapter lists the main conclusions obtained from the research carried out

in this doctoral thesis. In addition, some recommendations are suggested to be taken

into account in further studies.

Chapter 7

256

7.1. CONCLUSIONS

The results obtained from the present research work support the following

main conclusions:

- A 120 nm-thick Pt catalyst film was prepared by the cathodic arc deposition

(CAD) technique on K-βAl2O3 (a K+ conductor material) and was electrochemically

promoted in the methanol partial oxidation (POM) and steam reforming (SRM)

reactions. On the other hand, a Pt catalyst film prepared by impregnation presented an

excess of electrocatalytic activity and a poisoning effect due to potassium-derived

compounds which blocked Pt active sites. Thus, the catalyst films henceforth used in

this thesis were prepared through cathodic arc deposition or a similar physical vapor

deposition (PVD), such as sputtering or oblique angle deposition (OAD).

- The electrochemical promotion of catalysis (EPOC) allowed in-situ

enhancing the catalytic activity of the Pt film prepared by CAD and its selectivity

towards methanol partial oxidation against total oxidation mechanism, through the

controlled migration of K+ ions (under negative polarization). Moreover, the EPOC

effect observed on this catalyst and the others showed to be reversible upon removal

of these ions from the catalyst (under positive polarization).

- A novel catalyst film composed of Pt nanoparticles (of around 3 nm)

dispersed in a carbon matrix (Pt-DLC) was developed by CAD. This film was

successfully electrochemically promoted after a graphitization process and different

K+-derived promotional effects were found depending on the applied potential and the

reaction atmosphere (POM or SRM). The Pt-DLC film showed a higher catalytic

activity and a similar enhancement ratio than the pure Pt catalyst, even with worse

electric properties and much lower amount of metal (only 0.014 mg cm-2

). These

results demonstrate the possibility of electrochemically promoting catalyst

nanoparticles dispersed in an electronic non-ionic conductor support.

- It was possible to electrochemically activate a non-conductive catalyst film

formed by Au nanoparticles dispersed in a YSZ matrix by the co-sputtering of Au, Y

and Zr targets. For this purpose a silver current collector was used on the Au-YSZ

General conclusions and recommendations

257

catalyst film which was tested in the methanol partial oxidation reaction. The amount

of supplied K+ promoter ions was optimized and a permanent EPOC effect was found.

This study also demonstrated that the magnitude of the promotional effect obtained

with alkaline solid electrolytes does depend on neither the rate of promoter supply nor

the operational mode (either potentiostatic or galvanostatic) but on the final promoter

coverage achieved.

- A novel catalyst film configuration was developed to conduct

electrochemical promotion experiments, which consists of tilted nanocolumns of the

active metal grown on the support through an oblique angle deposition method. In this

way, a Cu catalyst film was prepared with a high porosity and gas-exposed surface

area, which results of great interest for catalytic and electrocatalytic applications. This

non noble metal catalyst was electrochemically promoted with K+ ions in the reaction

of partial oxidation of methanol and showed catalytic activity enhancement ratios of a

similar magnitude than those previously obtained with Pt catalysts.

- It was demonstrated the possible application of the EPOC phenomenon with

K+ ions for the activation of a metal catalyst, the modification of its catalytic

selectivity and the controlled variation of its oxidation state, by using a single Ni

catalyst film prepared by CAD. Under methanol decomposition and steam reforming

conditions, the K+ ions migration caused the activation of the methanol

dehydrogenation mechanism, and in the latter case, the K+ effect also attenuated the Ni

deactivation by carbon deposition. Under methanol partial oxidation conditions, the

K+-assisted strengthening of the oxygen chemisorption led to an increase in NiO

formation. As a consequence, the catalyst selectivity toward H2 and CO decreased

while that toward CO2, H2O and H2CO slightly increased.

- A novel porous Ni catalyst film prepared by the oblique angle deposition

technique with high porosity allowed to simultaneously produce H2 and graphene

oxide (GO) from methanol and to store part of the H2 on the GO (which acted as

adsorbent) under the promotional effect of K+ ions. Moreover, it was possible to

release the stored H2 in a controlled way, via electrochemical transfer of the promoter

ions under mild reaction conditions (280 ºC). The high obtained values of H2 storage

Chapter 7

258

capacity (19 g H2 x 100 g Ni-1

) open a new interesting possibility for the practical

application of the EPOC phenomenon in the fields of hydrogen production and

storage.

7.2. RECOMMENDATIONS

The following proposals can be stated in order to complete and extend this

research work:

- To perform EPOC studies in the methanol conversion reactions by using

other kinds of solid electrolytes, either cationic (Na-βAl2O3, NASICON, SCY), or

anionic (YSZ) or mixed ionic electronic conductors (perovskite-like materials) and

comparing the results with those herein obtained with K-βAl2O3.

- To study the influence of the phenomenon of electrochemical promotion on

hydrogen production processes from other materials such as ethanol or bioalcohols

(real mixtures).

- To apply some of the thin film deposition techniques reported in this thesis

(OAD, sputtering) for the preparation of either supported Cu or Ni catalysts similar to

those employed in classical catalysis studies (Ni/Al2O3, Cu/Al2O3, Ni/CeO2, Cu/CeO2)

or other kinds of configurations such as bimetallic (NiPt, NiCu) or multilayered

catalysts.

- To investigate other techniques for the preparation of metal catalyst films

with high dispersion; for instance, by pressing some catalyst powder on the solid

electrolyte or by using some kind of binder, in such a way that a suitable adhesion to

the substrate is achieved.

- To develop electrochemical promotion studies of methanol conversion

reactions by using some kind of double chamber solid electrolyte membrane reactor,

in such a way that some product of interest (e.g. H2) is obtained separately, and/or

some reactant (e.g. O2) is supplied from one chamber to the other (where the reaction

takes place).

General conclusions and recommendations

259

- To scale up the electrochemical promotion experiments for H2 production

and/or storage by using configurations with higher surface area, such as tubular or

monolithic catalysts, which would allow to obtain higher methanol conversion values

and/or higher H2 storage capacities.

Chapter 7

260

List of publications and conferences

261

The research results obtained in this doctoral thesis have led to the following

publications in international refereed journals:

1. Electrochemical activation of the catalytic methanol reforming reaction for H2

production. A. de Lucas-Consuegra, J. González-Cobos, Y. García-Rodríguez,

J.L. Endrino, J.L. Valverde. Electrochemistry Communications, 19 (2012) 55-

58.

2. Enhancing the catalytic activity and selectivity of the partial oxidation of methanol

by electrochemical promotion. A. de Lucas-Consuegra, J. González-Cobos, Y.

García-Rodríguez, A. Mosquera, J.L. Endrino, J.L. Valverde. Journal of

Catalysis, 293 (2012) 149-157.

3. Electrochemical promotion of Pt nanoparticles dispersed on a diamond-like

carbon matrix: A novel electrocatalytic system for H2 production. A. de Lucas-

Consuegra, J. González-Cobos, V. Carcelén, C. Magén, J.L. Endrino, J.L.

Valverde. Journal of Catalysis, 307 (2013) 18-26.

4. Electrochemical activation of Au nanoparticles for the selective partial oxidation

of metanol. J. González-Cobos, D. Horwat, J. Ghanbaja, J.L. Valverde, A. de

Lucas-Consuegra. Journal of Catalysis, 317 (2014) 293-302.

5. Electrochemical activation of an oblique angle deposited Cu catalyst film for H2

production. J. González-Cobos, V.J. Rico, A.R. González-Elipe, J.L. Valverde, A.

de Lucas-Consuegra. Catalysis Science and Technology, 5 (2015) 2203-2214.

6. Practical applications of the electrochemical promotion of catalysis in methanol

conversion processes. J. González-Cobos, J.L. Valverde, A. de Lucas-Consuegra.

Submitted to Topics in Catalysis.

List of publications and conferences

262

In addition, results obtained in this doctoral thesis have been presented in

the following international and national conferences:

- Keynotes in international conferences:

1. Coupling catalysis and electrochemistry for H2 production from alcohols. A.

de Lucas-Consuegra, J.L. Endrino, J. González-Cobos, D. López, J.A. Díaz,

J.L. Valverde. ANQUE’s International Congress of Chemical Engineering

(ICCE). Sevilla (Spain), June 2012.

- Oral presentations in international conferences:

1. Electrochemical promotion of partial oxidation of methanol for the production

of H2 and H2CO. A. de Lucas-Consuegra, J. González-Cobos, Y. García-

Rodríguez, A. Caravaca, J.L. Endrino, J.L. Valverde. 7th International

Conference on Environmental Catalysis (ICEC). Lyon (France), September

2012.

2. H2 production from methanol by electrochemical promotion of metallic

catalysts using cationic solid electrolytes. J. González-Cobos, A. de Lucas-

Consuegra, A. Nieto-Prado, J.L. Endrino, J.L. Valverde. 11th International

Conference on Catalysis in Membrane Reactors (ICCMR). Porto (Portugal),

July 2013.

3. Electrochemical promotion of hydrogen production reactions with alkali ionic

conductors. A. de Lucas-Consuegra, A. Caravaca, J. González-Cobos, F.

Dorado, J.L. Valverde. International Workshop on Ionically Conducting

Ceramics for Catalysis (IC3). Lyon (France), September 2013.

4. Hydrogen production from methanol by electrochemical promotion of Pt

catalysts. J. González-Cobos, C. Jiménez-Borja, N. Gutiérrez, A. de Lucas-

Consuegra, J.L. Valverde. 2013 Energy and Environment Knowledge Week

(E2KW), Toledo (Spain), November 2013.

List of publications and conferences

263

5. Electrochemical promotion of novel electrodes for the catalytic partial

oxidation of methanol. A. de Lucas-Consuegra, J. González-Cobos, D. López-

Pedrajas, D. Horwat, J. Ghanbaja, J.L. Valverde. 65th Annual Meeting of the

International Society of Electrochemistry (ISE). Lausanne (Switzerland),

August - September 2014.

6. Recent contributions of electrochemical promotion of catalysis for H2

production. A. de Lucas-Consuegra, J. González-Cobos, J.L. Valverde.

Accepted for oral presentation at the Hydrogen Power Theoretical and

Engineering Solutions International Symposium (HYPOTHESIS). Toledo

(Spain), September 2015.

7. Hydrogen production and storage by coupling of catalysis and

electrochemistry. J. González-Cobos, V.J. Rico, A.R. González-Elipe, J.L.

Valverde, A. de Lucas-Consuegra. Accepted for oral presentation at the 66th

Annual Meeting of the International Society of Electrochemistry (ISE). Taipei

(Taiwan), October 2015.

- Oral presentations in national conferences:

1. Promoción electroquímica para la producción de hidrógeno a partir de

alcoholes. J. González-Cobos. VI Jornadas de la Ciencia Joven. Ciudad Real

(Spain), May - June 2012.

- Posters in international conferences:

1. Electrochemical promotion of H2 production from methanol on novel metal

catalyst films. A. de Lucas-Consuegra, J. González-Cobos, V. Carcelén, C.

Magén, J.L. Endrino, J.L. Valverde. 11th European Congress on Catalysis

(EuropaCat). Lyon (France), September 2013.

2. Novel electrocatalytic systems for hydrogen production. J. González-Cobos,

N. Gutiérrez, C. Jiménez-Borja, A. de Lucas-Consuegra, J.L. Valverde. 64th

Annual Meeting of the International Society of Electrochemistry (ISE).

Santiago de Queretaro (Mexico), September 2013.

List of publications and conferences

264

- Posters in national conferences:

1. Producción de hidrógeno a partir de metanol mediante promoción

electroquímica. J. González Cobos, J.L. Valverde, A. de Lucas Consuegra. II

Jornadas Doctorales de Castilla-La Mancha. Toledo (Spain), November 2012.

2. Promoción electroquímica de nuevos catalizadores nanoestructurados para la

producción de H2. J. González-Cobos, A. de Lucas-Consuegra, A. Nieto-

Prado, N. Gutiérrez, D. Horwat, F. Soldera, J.L. Valverde. Reunión de la

Sociedad Española de la Catálisis (SECAT). Sevilla (Spain), June 2013.

3. Oxidación parcial de metanol mediante promoción electroquímica de

nanopartículas de oro. J. González-Cobos, D. López-Pedrajas, D. Horwat,

J.L. Valverde, A. de Lucas-Consuegra. I Encuentro de Jóvenes Investigadores

de la SECAT, Málaga (Spain), June 2014.

4. Producción de hidrógeno mediante catalizadores dispersos promocionados

electroquímicamente. J. González-Cobos, J.L. Valverde, A. de Lucas-

Consuegra. IV Jornadas Doctorales de Castilla-La Mancha. Cuenca (Spain),

October 2014.


Recommended