You are on page 1of 318

UNIVERSIDAD DE CASTILLA-LA MANCHA

FACULTAD DE CIENCIAS Y TECNOLOGÍAS QUÍMICAS


DEPARTAMENTO DE INGENIERÍA QUÍMICA

CO2 HYDROGENATION TO METHANOL


AT ATMOSPHERIC PRESSURE

Memoria que para optar al grado de Doctor en Ingeniería Química


presenta

JAVIER DÍEZ RAMÍREZ

Directores:
Dr. Fernando Dorado Fernández
Dra. Paula Sánchez Paredes
Composición del tribunal:
Dr. Michalis Konsolakis
Dr. Antonio de Lucas Consuegra
Dr. Raúl Sanz Martín
Profesores que han emitido informes favorable de la tesis:
Dr. Michalis Konsolakis
Dr. George Marnellos
D. Fernando Dorado Fernández, Catedrático de Ingeniería Química de la
Universidad de Castilla-La Mancha, y Dª. Paula Sánchez Paredes, Catedrática
de Ingeniería Química de la Universidad de Castilla- La Mancha,

CERTIFICAN: Que el presente trabajo de investigación titulado: “CO2


hydrogenation to methanol at atmoshperic pressure”, constituye la memoria que
presenta D. Javier Díez Ramírez para aspirar al grado de Doctor en Ingeniería
Química y que ha sido realizada en los laboratorios del Departamento de
Ingeniería Química de la Universidad de Castilla-La Mancha bajo su supervisión.

Y para que conste a efectos oportunos, firman el presente certificado

En Ciudad Real a de de 2018

Fernando Dorado Fernández Paula Sánchez Paredes


TABLE OF CONTENTS

ABSTRACT…………………………………………………………………....….1

INTRODUCTION.…………....……………………………………………….....13

CHAPTER 1. Influence of the preparation method of Pd/ZnO


catalysts………………………………………........................................................47

CHAPTER 2. Effect of carbon nanofibers supports on palladium/zinc


catalysts……………………………………………………………………….73

CHAPTER 3. Influence of the calcination, reduction and metal loading on Cu/ZnO


catalysts…………………………………………………………….................111

CHAPTER 4. Effect of support nature on cobalt catalysts……………………...147

CHAPTER 5. Optimization of the Pd/Cu ratio in Pd-Cu-Zn/SiC catalyst……..181

CHAPTER 6. Kinetics of the hydrogenation of CO2 to methanol at atmospheric


pressure using a Pd-Cu-Zn/SiC catalyst……………………………………...219

CHAPTER 7. Electrochemical promotion and characterization of PdZn alloy


catalysts with K and Na ionic conductors for pure gaseous CO2
hydrogenation………………………………………………………………..251

CONCLUSIONS……………………………………………………………....285

FUTURE WORK………………………………………………………………289

APPENDIX: EXPERIMENTAL SETUPS………………………………………..…293

ABOUT THE AUTHOR………………………………………………………...301


ÍNDICE

RESUMEN……………………………………………………………………….1

INTRODUCCIÓN…………………………………………………………….....13

CAPÍTULO 1. Influencia del método de preparación del catalizador


Pd/ZnO.………………………………………........................................................47

CAPÍTULO 2. Efecto de las nanofibras de carbono como soporte en los


catalizadores de Pd/Zn....……………….…………………………………….73

CAPÍTULO 3. Influencia de la calcinación, redución y contenido metálico en los


catalizadores de Cu/ZnO….……………………………………….................111

CAPÍTULO 4. Efecto del soporte en los catalizadores de cobalto…….……...147

CAPÍTULO 5. Optimización de la relación molar Pd/Cu en los catalizadores de


Pd-Cu-Zn/SiC………………………………………………………………...181

CAPÍTULO 6. Cinética de la reacción de hidrogenación de CO2 hacia metanol a


presión atmosférica usando el catalizador de Pd-Cu-Zn/SiC………………...219

CAPÍTULO 7. Promoción electroquímica y caracterización de la aleación PdZn


utilizando electrolitos sólidos conductores de iones Na y K en la hidrogenación
gaseosa de CO2……………………………………………………………...251

CONCLUSIONES..…………………………………………………………....285

RECOMENDACIONES…..……………………………………………………289

ANEXO: INSTALACIONES EXPERIMENTALES…..…………………………..…293

SOBRE EL AUTOR…………………………………………………………......301

2
RESUMEN
MARCO DE INICIO DE LA TESIS

DESARROLLO DE LA INVESTIGACIÓN
RESUMEN

-Marco de inicio de la tesis


Esta tesis surge tras la concesión del proyecto europeo de investigación
“CO2 and H2O toward methanol at atmospheric pressure in co-ionic electrochemical
membrane reactors “GREEN MEOH”, financiado por la red CAPITA ERA-NET del
VII Programa Marco de la Unión Europea (ACE.07.016), al departamento de
Ingeniería Química de la Universidad de Castilla-La Mancha (UCLM). La tesis
nace siendo el eje motor del proyecto, buscando desarrollar los objetivos
principales del mismo:

Primer objetivo: Búsqueda de un catalizador activo y selectivo hacia


metanol en la reacción de hidrogenación de CO2 a altas temperaturas de
operación. Comprende a su vez las siguientes tareas:

1.1. Uso de distintos metales y soportes.


1.2. Optimización de las condiciones de operación.
1.3. Desarrollo de un modelo cinético con el catalizador
elegido.

Segundo objetivo: Deposición del catalizador en el cátodo de un reactor


electro-catalítico de doble cámara con una membrana co-iónica. Las tareas que
incluye son:

2.1. Diseño del reactor electro-catalítico de doble cámara.

2.2. Deposición efectiva del catalizador en el cátodo del sistema.

2.3. Experimentos electro-catalíticos con membrana co-iónica.

La idea conceptual del proyecto puede verse en la Figura 1: tras la


aplicación de un potencial determinado, en el ánodo se producirá la oxidación
del agua, dejando libres los protones que migrarán hacia el cátodo donde,
gracias al catalizador escogido, reaccionarán con el dióxido de carbono dando
lugar a metanol (CH3OH). Al mismo tiempo la migración de iones O2- desde el

1
RESUMEN

cátodo activará el enlace C=O de las especies COx adsorbidas, acelerando la


síntesis de metanol.

2
Catalizador del electrodo catódico

Conductor co-iónico

Catalizador del electrodo anódico

Figura 1. Esquema del sistema electrocatalítico del proyecto “GREEN MeOH”.

A pesar de llegar hasta el último objetivo del proyecto, este no pudo ser
desarrollado en su totalidad, por la escasa conductividad iónica de las
membranas de electrolito sólido comerciales actuales a temperaturas alrededor
de 300 ºC, temperatura a la cual la termodinámica aún permite obtener metanol,
así como los problemas operativos del sistema electro-catalítico. Sin embargo, el
resto de objetivos fueron satisfactoriamente desarrollados, quedando a la
espera de las futuras mejoras que se produzcan en los sistemas electro-catalíticos
y las membranas de electrolito sólido comerciales, mejoras que quedan lejos del
objetivo de esta tesis doctoral.

-Desarrollo de la investigación

La presente tesis doctoral cuenta con siete capítulos a lo largo de los


cuales se desarrollan los objetivos anteriormente mencionados.

Los cinco primeros capítulos detallan el proceso de selección del


catalizador elegido como el mejor en términos de selectividad y actividad hacia
metanol. El sexto está dedicado al modelo cinético desarrollado para el
catalizador elegido. Por último, el séptimo capítulo se ocupa de explicar la
inmersión en la parte electro-catalítica, la preparación de los electrodos y los
experimentos electro-catalíticos en reactores de cámara simple.

2
RESUMEN

A continuación, se detallan cada uno de ellos:

El capítulo 1: “INFLUENCE OF THE PREPARATION METHOD OF Pd/ZnO


CATALYSTS” abre la investigación con un catalizador base de Pd/ZnO. En primer
lugar, se estudió la influencia de la temperatura de reducción (Figura 2). Se
comprobó que tiene un papel determinante en la síntesis de metanol debido a la
formación de un número mayor de centros activos de la aleación PdZn y menor
cantidad de paladio metálico.

5.0
Velocidad de formación metanol

4.5 N-10-F-500
4.0 N-10-F-425
(mol min g )

3.5 N-10-F-350
-1 -1

3.0
2.5
2.0
1.5
1.0
0.5
0.0
150 175 200 225 250 275 300
Temperatura (ºC)

Figura 2. Resultados catalíticos comparando distintas temperaturas de reducción.

Se continuó estudiando el efecto de la etapa de calcinación, ensayando


una calcinación rápida y una lenta, lo cual estaba directamente relacionado con
el tamaño final de las partículas de PdZn. Partículas más grandes provocaban
una menor conversión de CO2 pero una mayor selectividad y actividad hacia
metanol. Como se verá en el capítulo cinco, se determinó que la selectividad y
actividad aumentan, no tanto por el aumento del tamaño de las partículas de
aleación PdZn, sino por la disminución de partículas metálicas de paladio que se
incorporaban a la estructura de las partículas de aleación (haciendo que estas
aumentasen su tamaño). En la tercera etapa se modificó el contenido metálico. A
pesar de la mejora en términos de actividad y selectividad que confería un
3
RESUMEN

mayor porcentaje de paladio en la muestra, no se conseguía un aumento


proporcional de estas variables al del aumento del porcentaje de paladio. Por
último, se estudió la influencia del precursor de Pd utilizado: nitrato de paladio
y nitrato de tetramin paladio. Se concluyó con que el segundo influía en el
proceso de formación de las partículas de aleación de PdZn, dando lugar a
partículas con una estructura cristalina amorfa e inactiva en la síntesis de metanol,
por lo que fue descartado en beneficio del primero.

En el capítulo 2: “EFFECT OF CARBON NANOFIBERS SUPPORTS ON


PALLADIUM/ZINC CATALYSTS” se continúa utilizando el mismo metal activo que
en capítulo anterior pero modificando el soporte. Se utilizaron nanofibras de
carbono (fishbone y platelet) debido, por un lado, a la amplia experiencia del
grupo de Catálisis y Materiales en la síntesis, caracterización y uso de las mismas,
y por otro, por el aumento de conductividad eléctrica que pueden conferir al
catalizador depositado en el cátodo del sistema electro-catalítico. Así, se estudió
la influencia de las nanofibras sobre la formación de la aleación PdZn, donde se
observó que el propio carbón de las nanofibras actuaba como agente reductor
durante la etapa de calcinación en N2, desembocando en la formación de centros
activos PdZn sin necesidad de una etapa de reducción. Sin embargo, una
reducción posterior aumentaba el número de sitios activos de aleación,
incrementando la producción y selectividad de metanol. La influencia de la
relación molar Pd/Zn también fue estudiada para las nanofibras tipo fishbone.
Menores relaciones molares Pd/Zn producían un desplazamiento de la curva de
formación de metanol hacia temperaturas más altas. Además, se obtenían
partículas de mayor tamaño, que resultan ser más selectivas y activas hacia
metanol. Por último, se estudió la influencia de los dos tipos de nanofibras,
observando, mediante la técnica de difracción de rayos X, que una nueva
aleación PdZn con ratio 48,5:51,5 se había formado en las nanofibras platelet.
Esta nueva aleación, con un ratio ligeramente distinto al convencional 50:50,
confería al catalizador de PdZn soportado sobre nanofibras platelet una mayor
4
RESUMEN

adsorción de reactivos, mayor interacción metal-soporte y mayor actividad y


selectividad hacia metanol, todo ello debido a que estas nanofibras presentan
una distinta orientación de las láminas de grafeno (Figura 3).

a) b)

10 nm 10 nm

Figura 3. Orientación de las láminas de grafeno: (a) fishbone and (b) platelet supports.

En el capítulo 3: “INFLUENCE OF THE CALCINATION, REDUCTION AND


METAL LOADING ON Cu/ZnO CATALYSTS” se estudia el cobre como metal
activo en la síntesis de metanol. Debido a su bajo precio en comparación con el
paladio y a otras propiedades características, el cobre es el metal usado en la
síntesis industrial de metanol a partir de CO y CO2. Sin embargo, hay pocos
estudios que muestren en detalle la caracterización del catalizador base de
Cu/ZnO. Es por ello que surge la motivación de este capítulo. En primer lugar, se
analizó la influencia de la temperatura de calcinación, que afecta al tamaño de
las partículas. Al contrario de lo que ocurría con las partículas de aleación PdZn,
aquí un mayor tamaño no conllevaba una mayor actividad y selectividad hacia
metanol. En el caso de los catalizadores de Cu/ZnO el parámetro relevante es
la relación de intensidad entre los picos del cobre (1 1 1) y (2 0 0). Esta relación
está a su vez conectada con la morfología de los cristales de cobre. Así, los
mejores catalizadores en términos de actividad y selectividad hacia metanol
fueron aquellos que mayor relación I (111)/ I (200) presentaron. El estudio de la
temperatura de reducción mostró que temperaturas bajas (~150 ºC) provocaban
una reducción incompleta de los óxidos de cobre y temperaturas altas (> 400
ºC) daban lugar a la formación de una aleación CuZn que mejoraba la
5
RESUMEN

selectividad del catalizador a altas temperaturas. Por último, el estudio de


comparación entre distintos catalizadores con un porcentaje másico de cobre
diferente ofreció resultados coherentes con la teoría del mecanismo de doble
sitio activo, donde el H2 es adsorbido en los sitios activos de ZnO y el CO2 en los
de cobre. Así, se observa que el mecanismo es favorecido a mayor cercanía de
sitios activos y, por lo tanto, a mayor contenido metálico. Sin embargo,
porcentajes de cobre demasiado elevados (40%) ya no mejoran de forma
significativa la formación de metanol, lo que indica que hay un óptimo en el
porcentaje de cobre utilizado.

En el capítulo 4: “EFFECT OF SUPPORT NATURE ON COBALT CATALYSTS”


se estudia el cobalto como posible candidato en la síntesis de metanol. Los
resultados obtenidos muestran que el producto mayoritario es el metano, siendo
la selectividad para productos de mayor cadena carbonada, entre los que se
encuentra el metanol, menor del 5%. Aun así, en este capítulo se detalla la
caracterización y comparación de cuatro soportes (ZnO, CeO2, Gd2O3 y ZrO2)
en la síntesis de metano, componente primordial del gas natural, a partir de CO2.

En base a los resultados para los mejores catalizadores obtenidos hasta


el momento (Figura 4) donde se observa la superioridad del paladio frente al
cobre en términos de actividad hacia metanol, se decidió abordar una
optimización del catalizador base (Pd/ZnO), debido al alto precio del paladio,
introduciendo cobre en la composición del catalizador.

6
RESUMEN

100

Selectividad metanol (%)


a) N-10-F-500 (1º Capítulo)
90
PdZnO0.13/Plat (2º Capítulo)
80
20CuZnO-350-200 (3º Capítulo)
70
60
50
40
30
20
10
0
0 2 4 6 8 10
Conversión de CO2(%)
10
Velocidad de formación metanol

9 b) N-10-F-500 (1º Capítulo)


8 PdZnO0.13/Plat (2º Capítulo)
20CuZnO-350-200 (3º Capítulo)
(mol min g )

7
-1 -1

6 Co/ZnO (4º Capítulo)


5
4
3
2
1
0
150 175 200 225 250 275 300
Temperatura (ºC)

Figura 4. Resultados comparativos de los mejores catalizadores utilizados en los cuatro primeros
capítulos.

Así surge el capítulo 5: “OPTIMIZATION OF THE Pd/Cu RATIO IN Pd-


Cu-Zn/SiC CATALYST”. En este capítulo, cuyos resultados son de gran relevancia
en esta tesis doctoral, se sintetizó un catalizador cuyos valores de actividad a
altas temperaturas se acercan a los valores máximos que termodinámicamente
se pueden obtener, como se muestra en la Figura 5.

7
RESUMEN

Actividad metanol (mol min )


25

-1
Datos termodinámicos
20
Datos experimentales
15

10

0
140 160 180 200 220 240 260 280 300 320
Temperatura (ºC)

Figura 5. Comparación entre los valores del equilibrio termodinámico y los resultados experimentales
del catalizador 37.5PdCuZn/SiC.

Este catalizador se obtuvo tras estudiar distintos porcentajes molares


Pd:Cu, siendo la combinación 37,5% Pd, 12,5% Cu y 50% Zn la elegida por su
alta actividad y selectividad hacia metanol. Se compararon los resultados con
los catalizadores bimetálicos de PdCu, PdZn y CuZn, demostrando el efecto
sinérgico que la combinación de los tres metales tiene en la síntesis de metanol.
El carburo de silicio, utilizado como soporte, fue elegido debido a que es
químicamente inerte, lo que permite centrar el estudio en la interacción metal-
metal y no metal-soporte. Así, con este estudio se logró una combinación de sitios
activos no publicados hasta el momento: PdZn y PdCu. En los catalizadores
convencionales de paladio, estudiados en los capítulos 1 y 2, pequeñas partículas
de paladio permanecían en la estructura a pesar de disminuir la relación molar
Pd:Zn al mínimo, las cuales aumentaban la conversión de CO2 hacia CO, un
producto no deseado. La incorporación de cobre en el catalizador produce la
unión de estas partículas de paladio al cobre añadido, formándose la aleación
PdCu, que aunque es inactiva hacia la síntesis de metanol reduce
considerablemente la actividad total hacia CO. Por lo tanto, en los novedosos

8
RESUMEN

catalizadores trimetálicos de este capítulo, el cobre actúa como inhibidor del CO,
formando junto a las pequeñas partículas de paladio la aleación PdCu y
favoreciendo la dispersión de las partículas de aleación PdZn, haciendo estas
más activas hacia la síntesis de metanol. Se consigue así el catalizador más activo
y selectivo hacia metanol de este trabajo (Figura 6).

Selectividad metanol (%)


100 a) N-10-F-500 (1º Capítulo)
90 PdZnO0.13/Plat (2º Capítulo)
80 20CuZnO-350-200 (3º Capítulo)
70 37.5PdCuZn/SiC (5ºCapítulo)
60
50
40
30
20
10
0
0 2 4 6 8 10
Conversión de CO2(%)
10
Velocidad de formación metanol

b) N-10-F-500 (1º Capítulo)


PdZnO0.13/Plat (2º Capítulo)
8 20CuZnO-350-200 (3º Capítulo)
(mol min g )
-1 -1

Co/ZnO (4º Capítulo)


6 37.5PdCuZn/SiC (5º Capítulo)

0
150 175 200 225 250 275 300
Temperatura (ºC)

Figura 6. Resultados comparativos de los mejores catalizadores utilizados en cada capítulo.

El capítulo 6: “KINETICS OF THE HYDROGENATION OF CO2 TO


METHANOL AT ATMOSPHERIC PRESSURE USING A Pd-Cu-Zn/SiC CATALYST”
muestra el estudio y modelo cinético desarrollado para el catalizador
37.5PdZnCu/SiC. Se llevó a cabo un análisis de sensibilidad para evaluar el
efecto de las condiciones de reacción (la temperatura, la relación volumétrica
9
RESUMEN

CO2/H2 y la presencia de productos en la corriente de entrada) sobre la


actividad catalítica. Los resultados de este estudio se usaron para desarrollar
tres modelos cinéticos tipo Langmuir-Hinshelwood en los que se modificó el
término de adsorción (adsorción competitiva vs. mecanismo de doble sitio activo
vs. mecanismo de tres sitios activos). El tercer modelo, de tres sitios activos, fue el
que estadísticamente resultó ser el más apropiado, con una menor suma residual
de cuadrados, y el único que cumplió todas las restricciones de parámetros
utilizadas. Este modelo no solo reproduce los resultados experimentales, sino que
también apoya la teoría presentada en el capítulo anterior, que establece que
la síntesis de metanol y CO se produce en los sitios activos de PdZn y PdCu,
respectivamente, mientras que en el ZnO se produce la adsorción del hidrógeno.

El capítulo 7: “ELECTROCHEMICAL PROMOTION AND


CHARACTERIZATION OF PdZn ALLOY CATALYSTS WITH K AND Na IONIC
CONDUCTORS FOR PURE GASEOUS CO2 HYDROGENATION” muestra el paso
a la parte electro-catalítica del proyecto. En él se utiliza por primera vez la
aleación PdZn como cátodo de un sistema electrocatalítico para realizar un
estudio de promoción electroquímica de la catálisis (Non-Faradaic Electrochemical
Modification of Catalytic Activity, NEMCA). Se detalla así la preparación de los
electrodos de trabajo y los experimentos de promoción electroquímica sobre dos
electrolitos sólidos de β-Al2O3, uno conductor de iones K+ y otro de iones Na+.
Se estudió la influencia de los distintos iones en la promoción electroquímica
obtenida y la relación Pd/Zn en la preparación del electrodo. Por primera vez
en este tipo de estudios, se obtuvieron únicamente como productos CO y CH3OH.
Aunque los resultados de promoción logrados, en los que se observa que a mayor
cantidad de iones potasio o sodio migrados hacia el cátodo se produce la
inhibición de metanol y la producción de mayores cantidades de CO, no son los
deseados, este estudio sirve como base para mostrar la preparación del
electrodo de aleación PdZn y, en un futuro, poder aplicarlo a un sistema de doble
cámara con una membrana protónica (Figura 7) o una membrana co-iónica que
10
RESUMEN

sea capaz de conducir iones a temperaturas relativamente bajas (±350ºC)


donde la síntesis de metanol no se ve aún inhibida termodinámicamente.

CO2 CH3OH

CO2 CH3OH

H2O O2 PdZnCu

V H+ H+
é
Pd/Pt

H2O O2

Figura 7. Esquema de un sistema electrocatalítico de doble cámara con una membrana protónica.

11
INTRODUCCIÓN
A.1.PROBLEMÁTICA ENERGÉTICA Y AMBIENTAL

A.2. EL DIÓXIDO DE CARBONO

A.3. EL METANOL

BIBLIOGRAFÍA
INTRODUCCIÓN

A.1. Problemática energética y ambiental

El último informe publicado por la Organización de las Naciones Unidas


prevé que la población mundial actual de 7.600 millones de personas, alcance
los 8.600 millones para el año 2030 y los 11.200 millones para el 2100 [1].

En un mundo donde hoy día el 17% de la población global (1,2 billones


de personas) no tienen acceso a la electricidad y un 38% (2,7 billones de
personas) utilizan algún tipo de biomasa tradicional (Ej.: madera) para cocinar
[2], parece evidente que hay que buscar alternativas energéticas para
abastecer a todas esas nuevas personas que poblarán la Tierra.

En esta demanda energética creciente, uno de los principales problemas


al que nos enfrentamos es la emisión de gases de efecto invernadero (vapor de
agua, dióxido de carbono, metano, óxido de nitrógeno y ozono), que producen
el archiconocido efecto invernadero intensificado y la subida de temperaturas.
Entre los gases culpables de este fenómeno, el dióxido de carbono es el que
despierta un mayor interés, debido a su incremento del 30% en los últimos 50
años [3].

Es por ello que las acciones políticas van dirigidas a paliar o disminuir
estas emisiones; así en octubre del 2014 los dirigentes de la Unión Europea
acordaron en el Marco sobre clima y energía para 2030 [4] alcanzar los
siguientes objetivos:

-Al menos un 40% de reducción de las emisiones de gases de efecto


invernadero (en relación con los niveles de 1990).

-Al menos un 27% de cuota de energías renovables.

-Al menos un 27% de mejora de la eficiencia energética.

15
INTRODUCCIÓN

En diciembre de 2015, 195 países firmaron el primer acuerdo vinculante


mundial sobre el clima en la Conferencia de París sobre el Clima (COP21) [5].
Los Gobiernos acordaron:

-El objetivo a largo plazo de mantener el aumento de la temperatura


media mundial muy por debajo de 2 °C sobre los niveles preindustriales.

-Limitar el aumento a 1,5 °C, lo que reducirá considerablemente los


riesgos y el impacto del cambio climático.

-Que las emisiones globales alcancen su nivel máximo cuanto antes, si bien
reconocen que en los países en desarrollo el proceso será más largo.

-Aplicar después rápidas reducciones basadas en los mejores criterios


científicos disponibles.

Sin embargo, tal como muestra la Figura 8 a pesar de las nuevas políticas
implementadas para alcanzar estos objetivos, los compromisos de emisión de los
cuatro principales emisores no dejarían espacio para que el resto de países del
mundo emitieran CO2 más allá de 2037, si se quisiese llegar a un aumento medio
de 2ºC sobre los niveles preindustriales.

Figura 8. Pronóstico de las emisiones de CO2 según la Agencia Internacional de Energía [6].

16
INTRODUCCIÓN

Todos estos datos no hacen más que indicar que es necesario desarrollar
e investigar en nuevos procesos con los que disminuir las emisiones a la atmósfera
de CO2.

A.2. El dióxido de carbono


El dióxido de carbono (CO2), también conocido como óxido de carbono
(IV), gas carbónico y anhídrido carbónico (aunque los dos últimos cada vez son
menos utilizados), se caracteriza por ser inodoro, incoloro, no inflamable, más
pesado que el aire y ligeramente ácido. Dicha molécula está formada por dos
átomos de oxígeno y uno de carbono, tal como representa la Figura 9. Tiene un
peso molecular de 44,01 g/mol y su geometría es lineal y simétrica, lo cual hace
que sea una molecular apolar a pesar de tener enlaces polares.

O C O

Figura 9. Estructura de Lewis de la molécula de CO2.

Aunque a temperatura y presión ambiental se encuentra en estado


gaseoso, es posible obtenerlo en sus diferentes estados modificando las
condiciones ambientales: como sólido (nieve carbónica), líquido o en estado
supercrítico, a temperaturas por encima del punto crítico (T=31,1ºC y P=73 atm),
tal como se esquematiza la Figura 10.

Tal como se ha mencionado anteriormente, el creciente aumento del CO2


en la atmósfera ha hecho que el efecto invernadero, cuya función normal es la
de mantener a la Tierra con una temperatura media de 15ºC, se intensifique, con
el consecuente aumento de la temperatura global del planeta.

17
INTRODUCCIÓN

Figura 10. Diagrama de fases del CO2 en función de la presión y la temperatura.

Por ello y para evitar las emisiones de grandes cantidades de este


contaminante se han desarrollado un conjunto de tecnologías que reciben el
nombre de Captura, Almacenamiento y Uso de carbono (CAC)

A.2.1. Métodos de captura

Tal como muestra la Figura 11 se pueden distinguir distintos sistemas de


captura de CO2:

N2
O2
Carbón
PPost-combustión Gas Calor y electricidad Separación
de CO2 CO2
Biomasa
Aire
Carbón CO2
Aire/O2
Biomasa
vapor

Pre-combustión N2 O2 Compresión
Reformador Calor y
Gasificación de CO2
+ sep. de electricidad
Gas, petróleo
CO2 Y deshidratación
Aire

Carbón CO2
Oxicombustión Gas Calor y electricidad
Biomasa
O2
N2
Separación
Aire de Aire

Figura 11. Sistemas de captura de CO2 [7].

18
INTRODUCCIÓN

- Post-combustión: el objetivo es separar el CO2 que se encuentra


diluido en el resto de componentes de un gas de combustión, obtenido
al quemar con aire un combustible fósil o cualquier otro basado en
carbono. Los procesos de captura, generalmente realizados con
absorbentes químicos (base alcalina como una amina, MEA) se utilizan
mayoritariamente en procesos de combustión con aire (centrales
térmicas, cementeras, refinerías, cerámicas, etc). Su principal
problemática es lo diluido que está el CO2 en la corriente efluente
de la combustión, con muy alto flujo y sin presión.

- Pre-combustión: el combustible se transforma en una mezcla gaseosa


de H2, que puede ser separado y utilizado como combustible, y CO2.
Los pasos de conversión del combustible son más complejos: una etapa
previa de formación de gas de síntesis, luego la reacción de gas-
agua o “Water Gas Shift” y una posterior separación del CO2 con
absorción química, física, adsorción o membranas. Debido a esto, no
es la tecnología más adecuada para aplicar a la mayoría de las
plantas de generación eléctrica ya existentes. Sin embargo, es la
tecnología que la industria química utiliza desde hace décadas para
producir por ejemplo: H2, CO2 o NH3.

- Oxicombustión: consiste en la combustión del combustible en presencia


de oxígeno puro en lugar de aire, lo que incrementa la concentración
de CO2 en el gas efluente y facilita su separación final antes del
almacenamiento respecto al sistema de “post-combustión”. Su
principal problemática es el incremento de la inversión para producir
oxígeno puro y el encarecimiento de los equipos adaptados a esta
combustión.

A.3.2. Transporte
Los métodos principales para transportar el CO2 son:
19
INTRODUCCIÓN

- Transporte continuo mediante gaseoductos. Es el método ideal


siempre que exista la infraestructura correspondiente. Se puede
realizar mediante alguna de estas alternativas:

o Transporte en forma gaseosa a baja presión (P<45 atm)

o Transporte en fase densa (P>80 atm)

o Transporte en estado líquido.

- Transporte discontinuo en forma de gas licuado, mediante buques,


camiones o vagones que lo transportan en cisternas isotérmicas a
temperaturas y presiones inferiores a las ambientales (-20ºC y 2
MPa).

A.2.3. Almacenamiento

En muchas ocasiones, en lugar de ser valorizado, el CO2 es almacenado


y para ello las técnicas más comunes utilizadas son:

- Almacenamiento geológico: se consigue mediante la inyección del CO2


en forma condensada en una formación rocosa subterránea, siendo
candidatos a este tipo de almacenamiento los yacimientos de petróleo y
gas, formaciones salinas profundas y capas de carbón inexplotables. Una
vez inyectado, el CO2 es retenido mediante una capa impermeable de
pizarra o rocas arcillosas.

- Almacenamiento oceánico: consiste en la inyección de forma directa del


CO2 en los fondos marinos, pasando al ciclo global del carbono una vez
que se haya disuelto. Este tipo de almacenamiento sigue en fase
experimental [8].

- Carbonatación mineral: se refiere a la fijación del CO2 mediante su


reacción con óxidos alcalinos y alcalinotérreos, como óxido de magnesio
y óxido de calcio, de modo que se produce la formación de carbonatos.

20
INTRODUCCIÓN

A.2.4. Aplicaciones

Tal como indica la Figura 12 las aplicaciones del CO2 son muy variadas:

Agente inerte
para el envasado Carbonatos Urea
de alimentos Mejora la Metanol (aditivo)
inorgánicos
Sistemas de recuperación de
Policarbonatos Ácido salicílico
refrigeración petróleo Comercial
Extintores Horticultura

Hielo seco USO DIRECTO CO2 USO INDIRECTO

Pretratamiento de aguas
residuales Disolvente Bajo investigación
Sistemas
de soldadura Bebidas con gas
Carbonato de Ácido acético
Precipitación del dimetilo
carbonato de calcio Metanol Gas de síntesis
(CO2 + H2) (Reformado seco)

Procesos catalíticos

Figura 12. Usos directos e indirectos comunes del dióxido de carbono [9].

Este trabajo se centrará en el estudio de uno de los procesos catalíticos


que pueden ser utilizados para valorizar el CO2 mediante la reacción de
hidrogenación de CO2 hacia metanol.

A.3. Metanol

El compuesto químico metanol, también conocido como alcohol metílico, se


presenta a temperatura ambiente como un líquido ligero (de baja densidad),
incoloro, inflamable y tóxico que contiene menos carbono y más hidrógeno que
cualquier otro combustible líquido (Figura 13).

H O
C H
H H

Figura 13. Estructura de Lewis de la molécula de CH3OH.

21
INTRODUCCIÓN

El metanol cuenta con una serie de ventajas [10]:

-Tiene una alta densidad energética por unidad de masa y volumen


(22,034 MJ/kg y 15,6 MJ/L) para su utilización como combustible.

- Es sencillo de transportar y almacenar. A diferencia del H2 podría ser


utilizado en la infraestructura energética existente hoy día.

- No es necesaria una modificación de los motores actuales para su


integración como combustible, puede incluso ser mezclado con gasolina
[11] o usado en celdas de combustibles [12].

- No produce un gran impacto en el medio ambiente durante su


producción y uso como materia prima para obtener compuestos orgánicos
y como intermedio de varios productos químicos, tal como muestra la
Figura 14.

Metilaminas
Cloro-metanos
MTO/MTP
Disolventes
Otros/DMT
Formaldehído
Ácido acético
MTBE
MMA
Mezcla con gasolina
Biodiesel
DME

Figura 14. Porcentaje del uso final mundial de metanol 2015 [13].

22
INTRODUCCIÓN

En la actualidad, la demanda mundial de metanol es de


aproximadamente 90 millones de toneladas métricas al año [13]. La industria
del metanol es una de las más dinámicas y vibrantes del mundo, ya que es una
molécula básica que produce químicos que están presentes en nuestra vida
cotidiana. Desde el bloque básico de pinturas, disolventes y plásticos, a
aplicaciones innovadoras en energía, combustible para el transporte y la
construcción de celdas de combustible químico, el metanol es un producto clave y
una parte integral de nuestra economía global. Como se observa en la Figura
15, la industria del metanol se extiende por todo el mundo, con una producción
en mayoritatia en Asia.
Millones de toneladas métricas

América del Norte América del Sur Europa Occidental Europa Central CEI & Estados Bálticos
África Oriente Medio Subcontinente Indio Sudeste Asiático Noreste Asiático

Figura 15. Demanda mundial de metanol por regiones [13].

A.3.1. Producción industrial de metanol actual

El primer proceso comercial para la producción de CH3OH mediante la


destilación destructiva de la madera ya estaba en funcionamiento en el siglo XIX
[14]. La primera planta industrial dedicada a la producción de metanol a partir
de gas de síntesis (H2 y CO) fue construida por BASF en 1923. Este proceso
utilizaba catalizadores de ZnO o Cr2O3 a temperaturas de 300ºC y 200 atm. El
proceso es conocido como síntesis de metanol a alta presión [15].

23
INTRODUCCIÓN

En la actualidad se utiliza el proceso conocido como síntesis de metanol a


baja presión (<100 atm) usando catalizadores de Cu/ZnO/Al2O3 a 250ºC y 60
atm [14]. Entre los procesos industriales existentes para la síntesis de metanol, los
más usados son los desarrollados por las firmas comerciales Lurgi Corp. e
Imperial Chemical Insdustries Ltd. (ICI). Ambos procesos constan de tres etapas
principales (Figura 16):

Figura 16. Planta de producción de metanol.

- Reformado de la alimentación, sea ésta gas natural, mezcla de


hidrocarburos líquidos o carbón. Se realiza en dos etapas, un
reformado con vapor (800ºC y 40 atm) y una oxidación parcial
utilizando oxígeno previamente separado del aire (950ºC para el
gas natural y temperaturas de 1400-1500ºC y 55-60 atm para el
carbón o hidrocarburos líquidos.

- La síntesis de metanol producida a temperaturas de entre 240-270ºC


y 70-100 atm consiste en las siguientes reacciones:

24
INTRODUCCIÓN

CO + 2H2 ⇆ CH3 OH ∆H25°C = –90,77 kJ mol-1 (1)

CO2 + 3H2 ⇆ CH3 OH + H2 O ∆H25°C = –49.5 kJ mol-1 (2)

CO2 + H2 ⇆ CO + H2 O ∆H25°C = 41 kJ mol-1 (3)

Comercialmente se utilizan dos tipos de reactores:

o Lurgi: utiliza un reactor tubular cuyos tubos están llenos de


catalizador y son enfriados exteriormente por agua en
ebullición.

o ICI: consiste en un reactor de lecho fluidizado en el cuál el gas


de síntesis ingresa por la base y el metanol se obtiene por la
cabeza, manteniendo así el catalizador fluidizado en su
interior.

- Destilación del metanol para su purificación y separación del resto


de compuestos.

A pesar de su producción industrial, todavía existe controversia en muchos


factores importantes relacionados con los catalizadores, la cinética y el
mecanismo de la reacción de la síntesis comercial de metanol. Una de las
preguntas principales es cuál es la fuente principal de metanol, ¿el CO o el CO2?
[16, 17].

A.3.2. Síntesis de metanol a partir de la reacción de hidrogenación de CO2

La conversión de CO2 en metanol usando energía que no provenga de


combustibles fósiles ha sido sugerida como una de las vías de almacenamiento
energético capaz de resolver dos de los problemas principales: el calentamiento
global y la crisis energética de los combustibles fósiles [10].

Es así como surge el uso de esta reacción (Ec.2) utilizando las energías
renovables para la producción de hidrógeno. Entre ellas, la energía solar parte
como favorita, ya que según el Departamento de Energía de Estados Unidos, si
25
INTRODUCCIÓN

tan solo la radiación solar de un 1% de la superficie de la Tierra fuera convertida


en energía con un 10% de eficiencia, se obtendría alrededor de 105 TW de
energía o lo que es lo mismo a diez veces la energía mundial estimada que será
necesaria en el año 2050 [18].

El primer hito importante en el desarrollo de esta reacción se logró


cuando Lurgi AG y Sud-Chemie sintetizaron juntos un catalizador muy activo y
selectivo hacia metanol a partir de CO2 y H2 a 260ºC. La actividad de este
catalizador disminuyó aproximadamente a la misma velocidad que la actividad
del catalizador comercial [19]. El segundo hito fue el desarrollo de la primera
planta piloto para la producción de metanol (50 kg/h) a partir de CO2 e H2 que
se construyó en Japón usando un catalizador de Cu/ZnO modificado con SiO2.
Reciclando el alimento y con una velocidad espacial de 600 g L-1 h-1 se obtuvo
un 99,9% de selectividad hacia metanol durante 8000 horas de operación a
250ºC y 50 atm [20]. Otras plantas piloto han sido también construidas en Japón
por Mitsui Chemicals [21], en ella utilizan la energía solar para producir la
electrolisis del agua, o en Iceland por Carbon Recycling International [22] donde
utilizan energía geotérmica local para suplir sus requerimientos energéticos.

A.3.3. Estado del arte de los catalizadores usados en la hidrogenación


catalítica convencional del CO2

Una amplia variedad de catalizadores heterogéneos han sido evaluados


en la reacción de CO2 e H2 hacia metanol. Estos pueden ser clasificados en tres
categorías:

- Catalizadores basados en cobre: aunque el cobre solo no es eficiente


en la síntesis de metanol a partir de CO2 [23], su combinación con
ZnO sigue siendo el foco de estudios recientes [24,25] ya que el óxido
de zinc puede mejorar la dispersión y la estabilización de las
partículas de cobre [26]. Otros soportes como Al2O3 [27, 28] o ZrO2
[29, 30] han sido combinados con el ZnO o utilizados como soporte
26
INTRODUCCIÓN

principal con el afán de mejorar los resultados catalíticos del


catalizador base Cu/ZnO. Igualmente, promotores como el Ga2O3 o
el Cr2O3 [31] o el Au [32] han sido utilizados para modificar el estado
electrónico del cobre entre Cu0, Cu+ y Cu2+. Por último, el cobre
también se ha probado formando bimetálicos: PdCu [33] o el NiCu
[34].

- Catalizadores basados en metales nobles: el paladio es el metal


noble más usado debido a su alta actividad y selectividad hacia
metanol en la reacción de hidrogenación de CO2. Bahruji y col. [35]
han investigado recientemente la relación entre estructura y actividad
para el catalizador de Pd/ZnO preparado mediante diferentes
métodos. Sus resultados mostraban como a altas temperaturas de
reducción se formaba la aleación PdZn que reducía la producción de
CO mediante la reacción inversa a la reacción de gas de agua (Ec.3).
Además, partículas pequeñas de aleación PdZn y una menor área
superficial de paladio metálico incrementaban la producción de
metanol. Otros metales nobles como el Au [36] o la Ag [37] han sido
también probados con resultados satisfactorios en la síntesis de
metanol a partir de CO2 e H2.

- Otros metales: en los últimos años se han desarrollado nuevas


estructuras catalíticas para la síntesis de metanol. Es el caso del
In2O3/ZrO2, el catalizador muestra una excelente selectividad hacia
metanol (~100%) y una estabilidad de 1000h a 50 atm y 300ºC
[38]. En el caso de la combinación ZnO-ZrO2, se genera metanol con
una selectividad de 91% y es capaz de mantener la actividad más
de 550 horas de reacción a 50 atm y 315ºC, incluso en la presencia
de sulfuro (50 ppm), que es un conocido veneno de muchos
catalizadores [39].

27
INTRODUCCIÓN

Los estudios mencionados hasta ahora han sido llevados a cabo a


presiones mayores a la atmosférica, generalmente se encuentran entre las 10-
100 atm. El uso de altas presiones en un sistema encarece el proceso global, por
lo que algunos autores han decidido investigar esta misma reacción a presión
atmosférica con resultados generales de bajas conversiones y selectividades
medias (Tabla 1).
Tabla 1. Comparación de catalizadores en la reacción de hidrogenación de CO2 hacia metanol a
presión atmosférica.

Catalizador Pd/ZnO Cu/ZnO Cu/ZnO Cu/ZnO Au/CeO2 Au/ZnO


Contenido metálico (%) 10 10 50 60 1 1
Temperatura de reacción
190 190 167 165 225 225
(ºC)
CO2/H2 1/9 1/9 1/9 1/9 1/9 1/9
W/F (g min cm3) - - 0,005 0,005 - -
Velocidad de formación
de metanol 1,25 1,74 4,21 4,17 2,46 1,53
(μmol min-1 g-1)
Selectividad hacia
65,1 30,4 67,2 57,4 62,2 88,9
metanol (%)
Referencia [40] [40] [41] [42] [36] [36]

A.3.4. Estado del arte de la hidrogenación gaseosa de CO2 mediante


electrocatálisis (Efecto EPOC)

Antes de abordar los últimos estudios realizados en la hidrogenación de


CO2 en electrocatálisis estudiando el efecto de promoción electroquímica
(Electrochemical promotion of catalysis, EPOC), es necesario entender el
funcionamiento y estructura de este tipo de reactores electrocatalíticos.

A.3.4.1. Reactores de membrana de electrolito sólido (SEMRs)

Los reactores de membrana de electrolito sólido (Solid Electrolyte


Membrane reactor, SEMRs) son un tipo de reactores compuestos por un electrolito
sólido cerámico y dos electrodos porosos depositados a ambos lados del mismo.
Los electrolitos sólidos son materiales que deben ser químicamente estables, tener
una elevada conductividad iónica y ser impermeables a cualquier otra especie

28
INTRODUCCIÓN

no cargada eléctricamente [43]. Las membranas de electrolito sólido se pueden


clasificar en: membranas de conductores mixtos de iones-electrones (Mixed Ion-
Electron Conductors, MIEC) y membranas de electrolitos sólidos puros. Mientras
que las primeras se caracterizan por tener valores comparables de
conductividad iónica-electrónica, en las segundas la conductividad electrónica es
al menos dos órdenes de magnitud inferior a la conductividad iónica. Es por ello
que los electrolitos sólidos puros requieren un circuito externo para el transporte
de los electrones. Con este tipo de configuraciones se puede controlar de un
modo eficiente el transporte de iones hacia el electrodo que se desee mediante
la aplicación de potenciales eléctricos o intensidades en un sentido o en otro [43-
45].

-Tipos y modos de operación.

En función de cómo estén expuestos los electrodos a la atmósfera de


reacción, los SEMRs se pueden clasificar en reactores de cámara sencilla y
reactores de doble cámara. Los reactores de cámara sencilla se caracterizan por
tener ambos electrodos (ánodo y cátodo) expuestos a la misma atmósfera de
reacción. Por el contrario, los reactores de doble cámara se caracterizan por
tener los electrodos expuestos a diferentes mezclas de reacción.

Una de las principales ventajas de los reactores de cámara sencilla


respecto a los de doble cámara es su mayor simplicidad, así como la mayor
facilidad para ser implementados en sistemas en los que el soporte catalítico se
sustituye por el electrolito sólido. Por el contrario, los reactores de doble cámara
ofrecen la posibilidad de separar los productos obtenidos al mismo tiempo que
son formados, incrementando de este modo la economía del proceso.

Los SEMRs pueden funcionar de acuerdo a los siguientes modos de


operación [44]:

- Condiciones de circuito abierto (OCC, Open Circuit Conditions). En este


modo de operación no se aplica corriente eléctrica a través del electrolito
29
INTRODUCCIÓN

sólido, siendo la diferencia de potencial químico la verdadera fuerza


impulsora de la celda.
- Condiciones de circuito cerrado. (CCC, Closed Circuit Conditions). En estas
condiciones se aplica una corriente eléctrica que permite transferir los
iones a través del electrolito sólido, de un electrodo a otro, y reaccionar
así con los reactivos presentes en la fase gas. De este modo, el reactor
podría trabajar como una celda de combustible si el objetivo es obtener
energía.

Por otro lado, si el objetivo fuera obtener un determinado producto de


reacción, se podría aplicar cierta corriente eléctrica (equivalente al flujo
de iones transferidos a través del electrolito sólido) en la dirección
deseada, con el objetivo de favorecer la reacción de dichos iones y de
los reactivos presentes en la atmósfera de reacción.

-Electrolitos sólidos

Se conocen numerosos electrolitos sólidos que se suelen clasificar de


acuerdo al ion móvil que se desplaza a su través. A día de hoy se han descubierto
conductores de O2-, F-, H+, K+, Na+, Cu+, Ag+ and Li+ [46].

En este trabajo nos centraremos en los electrolitos solidos basados en β”-


alúmina. Este tipo de materiales presentan conductividad de iones Na+ y K+ a
baja temperatura (> 150 ºC). Cabe destacar que los materiales de esta familia
permiten la promoción electroquímica (Figura 17) sobre superficies metálicas
[47].

Reactantes Productos

Catalizador/
Electrodo de trabajo

I<0 K+ K+
K+
Contraelectrodo Electrodo de
referencia

Figura 17. Representación del efecto EPOC empleando K-βAl2O3 como electrolito sólido.
30
INTRODUCCIÓN

A.3.4.2. Promoción electroquímica de la catálisis

El fenómeno de promoción electroquímica de la catálisis (Electrochemical


Promotion Of Catalysis, EPOC) o efecto NEMCA (Non- Faradaic Electrochemical
Modification of Catalytic Activity) ha sido definido como “una herramienta
importante de la electroquímica que permite alterar de manera pronunciada,
reversible y predecible la actividad y selectividad catalítica de catalizadores
conductores soportados sobre electrolitos sólidos” [47]. En este sentido, se puede
afirmar que es un tipo especial de catálisis en el que las propiedades catalíticas
se modifican mediante la aplicación de un campo eléctrico.

Este fenómeno fue descubierto en el año 1981, por el grupo del profesor
Vayenas, quienes observaron que la actividad y la selectividad de un
catalizador depositado sobre un electrolito sólido podía ser electroquímicamente
modificada in-situ durante el propio proceso de reacción [48]. Esto ocurre cuando
se aplica un voltaje o una intensidad eléctrica entre un electrodo metálico (que
actúa como catalizador y electrodo de trabajo) depositado a un lado de un
electrolito sólido, y un segundo electrodo (contra-electrodo) depositado en el
lado opuesto de dicho electrolito. De este modo, si sobre el electrodo de trabajo
está teniendo lugar una reacción catalítica heterogénea, la aplicación de
corriente eléctrica puede provocar un incremento pronunciado de la velocidad
de reacción, pudiendo llegar a ser de 10 a 105 veces el valor que predice la
ley de Faraday. Este hecho motivó que este fenómeno también sea conocido
como modificación electroquímica no faradaica de la actividad catalítica (efecto
NEMCA en sus siglas en inglés).

El origen de este fenómeno ha sido atribuido al movimiento de especies


promotoras controlado electroquímicamente mediante la aplicación de un
potencial [47]. Estas especies son generadas en la región conocida como tbp
(three-phase bounderie; la interfase entre el electrolito sólido, catalizador-
electrodo de trabajo y la fase gas).

31
INTRODUCCIÓN

Durante el proceso de migración, el movimiento de estos iones está


acompañado por el correspondiente ion de compensación de carga, lo que da
lugar a dipolos neutros superficiales. Estos dipolos se distribuyen a lo largo de
toda la superficie metálica, constituyendo lo que se conoce como doble capa
efectiva. La formación de la doble capa efectiva produce una modificación de
la función de trabajo del metal, es decir, una variación de la densidad
electrónica, modificando de este modo su capacidad de enlace con cada una de
las moléculas de reactivo tal y como se muestra en la siguiente ecuación:

ΔΦ=eΔUWR (4)

donde ΔΦ es la variación de la función de trabajo, e es la carga del


electrón y ΔUWR es la modificación de la diferencia de potencial. Así pues,
cambiando la función de trabajo de la película metálica se puede alterar la
capacidad de enlace con cada una de las moléculas de reactivo, lo que se
traduce en una modificación del comportamiento catalítico del metal, que
dependerá de la naturaleza del promotor [49]. De este modo, si la especie
promotora es electronegativa (O2-) se produce un incremento en la función de
trabajo, lo que favorece la quimisorción de adsorbatos donadores de electrones
y se desfavorece la de aceptores de electrones. Por el contrario, una disminución
de la función de trabajo mediante la adición de promotores electropositivos
(Na+) favorece la quimisorción de especies aceptoras de electrones.

Esta teoría, que ha sido demostrada mediante una gran variedad de


técnicas de caracterización tanto catalíticas, electrocatalíticas y de análisis de
superficies (XPS, TPD, PEEM, STM, voltamperometrías cíclicas) [47], ha puesto de
manifiesto que este fenómeno es análogo al de la promoción química
convencional de la catálisis heterogénea. Sin embargo, el fenómeno EPOC cuenta
con la ventaja adicional de poder controlar, de un modo preciso y reversible, la
cantidad de promotor en el catalizador durante el propio proceso de reacción.

32
INTRODUCCIÓN

-Tipos de reacción

Se pueden distinguir cuatro tipos de reacciones basadas en el fenómeno


de promoción electroquímica teniendo en cuenta las interacciones atractivas o
repulsivas entre los promotores y adsorbatos, [47]:

- Reacciones electrofóbicas: son aquellas reacciones que muestran un


incremento de la velocidad de reacción para valores positivos de
potencial. Este tipo de comportamiento tiene lugar cuando la cinética es
de orden positivo en el donador de electrones y de orden cero o negativo
en el aceptor de electrones, es decir, el donador de electrones es el que
se encuentra más débilmente adsorbido sobre el catalizador.

- Reacciones electrofílicas: son aquellas reacciones que muestran un


incremento de la velocidad de reacción para valores negativos del
potencial. Este tipo de reacciones se dan cuando la cinética es de orden
positivo en el aceptor de electrones y de orden cero o negativo en el
donador de electrones, es decir, el aceptor de electrones se muestra más
débilmente adsorbido sobre el catalizador.

- Reacciones tipo volcán: estas reacciones presentan un máximo local de


la velocidad de reacción respecto al potencial aplicado. Este
comportamiento tiene lugar cuando tanto el donador como el aceptor de
electrones se encuentran fuertemente adsorbidos sobre el catalizador.

- Reacciones tipo volcán invertido: son aquellas que presentan un mínimo


en la velocidad de reacción respecto al potencial aplicado. En este caso,
este tipo de comportamiento tiene lugar cuando el donador y el aceptor
de electrones se encuentran débilmente adsorbidos sobre el catalizador.

33
INTRODUCCIÓN

A.3.4.3. Investigaciones en la hidrogenación de CO2 en reactores de cámara


simple

En la Tabla 2 se muestra un resumen de todos los estudios publicados


sobre la hidrogenación de CO2 en reactores de cámara simple. Como se puede
observar, en la mayoría de los casos los dos únicos productos son CO y CH4,
excepto en los estudios realizados a escala bancada con un área de electrodo
de trabajo mayor. El efecto obtenido, en general, es un efecto electrofílico para
el CO, cuya velocidad de reacción se ve aumentada al disminuir el potencial, y
un efecto electrofóbico hacia el CH4, cuya producción se ve desfavorecida a
potenciales menores debido a una menor quimisorción de H2 en el catalizador,
como consecuencia de la migración de los iones K+ hacia el electrodo de trabajo
y catalizador.

34
Tabla 2. Estudios EPOC en la reacción de hidrogenación de CO 2.

Catalyst film Main


Solid T CO2 H2 FT EPOC
Catalyst preparation reaction Reference
electrolyte (ºC) (%) (%) (N L h-1) behaviour
technique products

CO, CH4, Inverted


Rh, Pt, Cu YSZ Sputtering 220-380 1 5,6 60 [50]
C2H4 volcano

YSZ, Organometallic Inverted


Pd 533-605 72,8 22,7 2,3-18 CO [51]
Na-βAl2O3 paste volcano

Electrophobic
Organometallic (CH4)
Rh YSZ 346-477 5,5 63 0,9-3 CO, CH4 [52]
paste Electrophibic
(CO)

35
Ni, Ru YSZ Impregnation 200-440 1 5,6 6 CO, CH4 Electrophibic [53]

Electrophobic
(CH4)
Ru YSZ Impregnation 200-300 3 30 12 CO, CH4 [54]
Electrophibic
(CO)

Electrophobic
CO, CH4, (CH4)
CH3OH, Electrophibic
Pt K-βAl2O3 Dip-coating 400 19-47,5 47,5-76 90-522 (CO). The [55]
C2H5OH,
C2+C3 others vary
with reaction
conditions
INTRODUCCIÓN
Continuación Tabla 2. Estudios EPOC en la reacción de hidrogenación de CO2.

Catalyst film Main


Solid T CO2 H2 FT EPOC
Catalyst preparation reaction Reference
electrolyte (ºC) (%) (%) (N L h-1) behaviour
technique products
INTRODUCCIÓN

CH3OH, It varies with


Electroless
Cu K-βAl2O3 200-400 19-31 63,3-76 90-522 C2H5OH, reaction [56]
deposition
C2H6O conditions

CO, CH4,
Organometallic It varies with
CH3OH,
Pt, Ni, Pd YSZ paste/ 225-450 19-47,5 47,5-76 90 reaction [57]
C2H6, C3H6,
Electroless conditions
C2H6O

36
Electrophobic
YSZ (CH4)
Ru Impregnation 200-340 0,025-2 1-15 6 CO, CH4 [58]
K-βAl2O3 Electrophibic
(CO)

Organometallic 0,6-
Cu SZY paste/ Cu 550-750 0,3-4,9 12-18 CO Electrophibic [59]
2,24
powder

Electrophobic
(CH4)
Ru K-βAl2O3 Impregnation 280-420 3 30 6 CO, CH4 [60]
Electrophibic
(CO)

Organometallic
Pt YSZ 650-800 1-10 1-10 1,2 CO Electrophibic [61]
paste
Continuación Tabla 2. (Tabla 1) Estudios EPOC en la reacción de hidrogenación de CO 2.

Catalyst film Main


Solid T CO2 H2 FT EPOC
Catalyst preparation reaction Reference
electrolyte (ºC) (%) (%) (N L h-1) behaviour
technique products

Electrophobic
Organimetallic (CH4)
Ni K-βAl2O3 240-300 1,5-10 3-30 1,2-12 CO, CH4 [62]
paste Electrophibic
(CO)

37
Electrophobic
(CH4)
Ru BZY Impregnation 350-450 0,25-2,3 1-15 12-24 CO, CH4 [63]
Electrophibic
(CO)
INTRODUCCIÓN
INTRODUCCIÓN

Bibliografía
[1] WorldPopulation 2017. United Nations. Department of Economic and Social
Affairs.
https://esa.un.org/unpd/wpp/Publications/Files/WPP2017_Wallchart.pdf

[2] Energy Access Outlook 2017 from Poverty to Prosperity. International Energy
Agency.
https://www.iea.org/publications/freepublications/publication/WEO2017Spe
cialReport_EnergyAccessOutlook.pdf

[3] P.T. Ed. Dlugokencky, NOAA/ESRL (www.esrl.noaa.gov/gmd/ccgg/trends/).

[4] Clima y energía 2020-2030: Marco de actuación.


https://www.eixoatlantico.com/images/informes/documentos/energia2030.pdf

[5] The Road from Paris: assessing the implications of the Paris Agreement and
accompanying the proposal for a Council decision on the signing, on behalf of
the European Union, of the Paris agreement adopted under the United Nations
Framework Convention on Climate Change http://eur-lex.europa.eu/legal-
content/EN/TXT/?uri=CELEX:52016DC0110

[6] Energy, Climate Change & Environment 2016 Insights. International Energy
Agency.
http://www.iea.org/publications/freepublications/publication/ECCE2016.pdf

[7] IPCC Special Report on Carbon Dioxide Capture and Storage 2005.
https://www.ipcc.ch/pdf/special-reports/srccs/srccs_wholereport.pdf

[8] Reith, F.; Keller, D. P.; Oschlies, A. Revisiting ocean carbon sequestration by
direct injection: A global carbon budget perspective. Earth System Dynamics 7
(2016) 797-812.

[9] Raudaskoski, R.; Turpeinen, E.; Lenkkeri, R.; Pongrácz, E.; Keiski, R. L. Catalytic
activation of CO2: Use of secondary CO2 for the production of synthesis gas and

38
INTRODUCCIÓN

for methanol synthesis over copper-based zirconia-containing catalysts. Catalysis


Today, 144 (2009) 318-323.

[10] Ganesh, I. Conversion of carbon dioxide into methanol – a potential liquid


fuel: Fundamental challenges and opportunities (a review). Renewable and
Sustainable Energy Reviews, 31 (2014) 221-257.

[11] Olah, G. A. Beyond oil and gas: the methanol economy. Angewandte Chemie
International Edition, 44 (2005) 2636-2639.

[12] Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M.; Joshi, J.B. Catalytic carbon
dioxide hydrogenation to methanol : a review of recent studies. Chemical
Engineering Research and Design, 92 (2014) 2557-2567.

[13] HIS Chemical Bulletin. Insights. The changing face of the global methanol
industry: http://www.methanol.org/wp-content/uploads/2016/07/IHS-
ChemicalBulletin-Issue3-Alvarado-Jun16.pdf

[14] McKetta, J. J. Encyclopedia of chemical processing and design, vol. 29


Marcel Dekker, New York, 1988.

[15] Lee, S. Methanol synthesis technology, CRC Press, The United States 1990
p.236

[16] Skrzypek, J.; Lachowska, M.; Grzesik, M.; Sloczynski, J.; Nowak, P.
Thermodynamics and kinetics of low pressure methanol synthesis, The Chemical
Engineering Journal and the Biochemical Engineering Journal, 58 (1995) 101-
108.

[17] Ostrovskii, V. E. Mechanisms of methanol synthesis from hydrogen and carbon


oxides at Cu-Zn-containing catalysts in the context of some fundamental problems
of heterogeneous catalysis, Catalysis Today, 77 (2002) 141-160.

[18] Centi, G.; Perathoner, S. CO2-based energy vectors for the storage of solar
energy. Greenhouse Gas Science and Technology, 1 (2011) 21-35.

39
INTRODUCCIÓN

[19] Goehna, H.; Koenig, P. Producing methanol from CO2, Chemical Technology,
6 (1994) 36-39.

[20] Saito, M. R&D activities in Japan on methanol synthesis from CO2 and H2,
Catalysis Surveys from Japan, 2 (1998) 175-184.

[21] Tremblay, J.F. CO2 as feedstock. Mitsui will make methanol from the
greenhouse gas. Chemical Engineering, 86 (2008) 13

[22] Shulenberger, A. M.; Jonsson, F. R.; Ingolfsson, O.; Tran, K. C. Process for
producing liquid fuel from carbon dioxide and water, US Patent Appl.
0244208A1 2007

[23] Toyir, J.; De la Piscina, P. R.; Fierro, J. L. G.; Homs, N. Highly effective
conversión of CO2 to methanol over supported and promoted copper-based
catalysts: influence of support and promoter. Applied Catalysis B: Environmental,
29 (2001) 207-215.

[24] Karelovic, A.; Ruiz, P. The role of copper particle size in low pressure
methanol synthesis via CO2 hydrogenation over Cu/ZnO catalysts. Catalysis
Science & Technology, 5 (2015) 869.

[25] Li, W.; Lu, P.; Xu, D.; Tao, K. CO2 hydrogenation to methanol over Cu/ZnO
catalysts synthesized via a facile solid-phase grinding process using oxalic acid.
Korean Journal of Chemical Engineering, 35 (2018) 110-117

[26] Brown, N. J.; Weiner, J.; Hellgardt, K.; Shaffer, M. S. P.; Williams, K.
Phosphinate stabilized ZnO and Cu colloidal nanocatalysts for CO2
hydrogenation to methanol. Chemical Communications, 49 (2013) 11074-11076.

[27] Bahmani, M.; Vasheghani Farahani, B.; Sahebdelfar, S. Preparation of high


performance nano-sized Cu/ZnO/Al2O3 methanol synthesis catalyst via aluminum
hydrous oxide sol. Applied Catalysis-A , 520 (2016) 178.

40
INTRODUCCIÓN

[28] Behrens, M.; Kibner, S.; Girsgdies, F.; Kasatkin, I.; Hermerschmidt, F.; Mette,
K.; Ruland, H.; Muhler, M.; Schlögl, R. Knowledge-based development of a
nitrate-free synthesis route for Cu/ZnO Mmethanol synthesis catalysts via formate
precursors. Chemical Communications, 47 (2011) 1701.

[29] Ladera, R.; Pérez-Alonso, F. J.; González-Carballo, J.M.; Ojeda, M.; Rojas,
S.; Fierro, J. L. G. Catalytic valorization of CO2 via methanol synthesis with Ga-
promoted Cu-ZnO-ZrO2 catalyts. Applied Catalysis B: Environmental, 142 (2013)
241.

[30] Bonura, G.; Arena, F.; Mezzatesta, G.; Cannilla, C.; Spadaro, L.; Frusteri, F.
Role of the ceria promoter and carrier on the functionality of Cu-based catalysts
in the CO2 to methanol hydrogenation reaction. Catalysis Today, 171 (2011)
251.

[31] Toyir, J.; Miloua, R.; Elkadri, N. E.; Nawdali, M.; Toufik, H.; Miloua, F.; Saito,
M. Sustainable process for the production of methanol from CO2 and H2 using
Cu/ZnO-based multicomponent catalyst. Physics Procedia, 2 (2009) 1075-1079.

[32] Martin, O.; Mondelli, C.; Curulla-Ferré, D.; Drouilly, C.; Hauert, R.; Pérez-
Ramírez, J. Zinc-rich copper catalysts promoted by gold for methanol synthesis.
ACS Catalysis, 5 (2015) 5607-5616.

[33] Jiang, X.; Koizumi, N.; Guo, X.; Song, C. Bimetallic Pd-Cu catalysts for
selective CO2 hydrogenation to methanol. Applied Catalysis B-Environmental,
170-171 (2015) 173-185.

[34] Zhao, F.; Gong, M.; Zhang, Y.; Li, J. The performance and structural study of
CuNi alloy catalysts for methanol synthesis. Journal of Porous Materials, 23
(2016) 733-740.

[35] Bahruji, H.; Bowker, M.; Hutchings, G.; Dimitratos, N.; Wells, P.; Gibson, E.;
Jones, W.; Brookes, C.; Morgan, D.; Lalev, G. Pd/ZnO catalysts for direct CO2
hydrogenation to methanol. Journal of catalysis, 343 (2016) 133-146.
41
INTRODUCCIÓN

[36] Vourros, A.; Garagounis, I.; Kyriakou, V.; Carabineiro, S. A. C.; Maldonado-
Hódar, F. J.; Marnellos, G. E.; Konsolakis, M. Carbon hydrogenation over
supported Au nanoparticles: effect of the support, Journal of CO2 Utilization, 19
(2017) 247-256.

[37] R. Grabowski, J. Sloczynski, M. Sliwa, D. Mucha, R. P. Socha, Influence of


polymorphic ZrO2 phases and the silver electronic state on the activity of
Ag/ZrO2 catalysts in the hydrogenation of CO2 to methanol. ACS Catalysis, 1
(2011) 266-278.

[38] Martin, O.; Martín, A. J.; Mondelli, C.; Mitchell, S.; Segawa, T. F.; Hauert, R.;
Drouilly, C.; Curulla-Ferré, D.; Pérez-Ramírez, J. Indium oxide as a superior
catalyst for methanol synthesis by CO2 hydrogenation. Angewandte Chemie
International Edition, 55 (2016) 6261-6265.

[39] Jacoby, M. Catalysis: Bimetallic catalyst selectively converts CO2 to


methanol. Chemical and Engineering News, 95 (2017) 9

[40] Iwasa, N.; Suzuki, H.; Terashita, M.; Arai, M.; Takezawa, N. Methanol
synthesis from CO2 under atmospheric pressure over supported Pd catalysts.
Catalysis Letters, 96 (2004) 75-78.

[41] Fujita, S. I.; Moribe, S.; Kanamori, Y.; Kakudate, M.; Takezawa, N.
Preparation of a Coprecipitated Cu/ZnO Catalyst for the Methanol Synthesis
from CO2 - Effects of the Calcination and Reduction Conditions on the Catalytic
Performance. Applied Catalysis A, 207 (2001) 121.

[42] Fujita, S. I.; Kanamori, Y.; Satriyo, A. M.; Takezawa, N. Methanol Synthesis
from CO2 over Cu/ZnO Catalysts Prepared from Various Coprecipitated
Precursors. Catalysis Today, 45 (1998) 241.

[43] Ni, M.; Leung, M. K. H.; Leung D. Y. C. Technological development of


hydrogen production by solid oxide electrolyzer cell (SOEC) International Journal
of Hydrogen Energy, 33 (2008) 2337-2354.
42
INTRODUCCIÓN

[44] Garagounis, I; Kyriakou, V; Anagnostou, C.; Bourganis, V.; Papachristou, I.;


Stoukides, M. Solid electrolytes: applications in heterogeneous catalysis and
chemical cogeneration. Industrial and Engineering Chemistry Research, 50 (2011)
431-472.

[45] Stoukides, M. Methane conversion to C2 hydrocarbons in solid electrolyte


membrane reactors. Research on Chemical Intermediates, 32 (2006) 187-204.

[46] Gellings, P. J.; Bouwmeester, H. J. The CRC handbook of solid state


electrochemistry, 1997.

[47] Vayenas, C. G.; Bebelis, S.; Pliangos, C.; Brosda, S.; Tsiplakides, D.
Electrochemical activation of catalysis: promotion, electrochemical promotion, and
metal-support interactions, 2001.

[48] Stoukides, M.; Vayenas, C.G. The effect of electrochemical oxygen pumping
on the rate and selectivity of ethylene on polycrystalline silver. Journal of
Catalysis, 70 (1981) 137-146.

[49] Vayenas, C. G.; Bebelis, S.; Ladas, S. Dependence of catalytic rates on


catalyst work function. Nature, 343 (1990) 625

[50] Papaioannou, E. I.; Souentie, S.; Hammad, A.; Vayenas, C. G. 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.

[51] Bebelis, S.; Karasali, H.; Vayenas, C. G. Electrochemical promotion of the


CO2 hydrogenation on Pd/YSZ and Pd/β″-Al2O3 catalyst-electrodes. Solid State
Ionics, 179 (2008) 1391-1395.

[52] Bebelis, S.; Karasali, H.; Vayenas, C.G. Electrochemical promotion of CO2
hydrogenation on Rh/YSZ electrodes. Journal of Applied Electrochemistry, 38
(2008) 1127-1133.

43
INTRODUCCIÓN

[53] Jiménez, V.; Jiménez-Borja, C.; Sánchez, P.; Romero, A.; Papaioannou, E. I.;
Theleritis, D.; Souentie, S.; Brosda, S.; Valverde, J. L. 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.

[54] Theleritis, D.; Souentie, S.; Siokou, A.; Katsaounis, A.; Vayenas, C. G.
Hydrogenation of CO2 over Ru/YSZ Electropromoted Catalysts. ACS Catalysis,
2 (2012) 770-780.

[55] Ruiz, E.; Cillero, D.; Martínez, P. J.; Morales, Á.; Vicente, G. S.; De Diego,
G.; Sánchez J. M. Bench scale study of electrochemically promoted catalytic CO2
hydrogenation to renewable fuels. Catalysis Today, 210 (2013) 55-66.

[56] Ruiz, E.; Cillero, D.; Martínez, P.J.; Morales, Á.; Vicente, G.S.; De Diego, G.;
Sánchez, J. M. 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.

[57] Ruiz, E.; Cillero, D.; Martínez, P. J.; Morales, Á.; Vicente, G. S.; De Diego, G.;
Sánchez, J.M. 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.

[58] Theleritis, D.; Makri, M.; Souentie, S.; Caravaca, A.; Katsaounis, A.; Vayenas,
C. G. Comparative study of the electrochemical promotion of CO2 hydrogenation
over Ru supported catalysts using electronegative and electropositive
promoters. ChemElectroChem, 1 (2014) 254-262.

[59] Karagiannakis, G.; Zisekas, S.; Stoukides, M. Hydrogenation of carbon


dioxide on copper in a H+ conducting membrane-reactor, Solid State Ionics, 162-
163 (2003) 313-318.

44
INTRODUCCIÓN

[60] Makri, M.; Katsaounis, A.; Vayenas, C. G. Electrochemical promotion of CO2


hydrogenation on Ru catalyst–electrodes supported on a K–β″–Al2O3 solid
electrolyte. Electrochimica Acta 179 (2015) 556-564.

[61] Pekridis, G.; Kalimeri, K.; Kaklidis, N.; Vakouftsi, E.; Iliopoulou, E. F.;
Athanasiou, C.;. Marnellos, G. E. Study of the reverse water gas shift (RWGS)
reaction over Pt in a solid oxide fuel cell (SOFC) operating under open and
closed-circuit conditions. Catalysis Today, 127 (2007) 337.

[62] Gutiérrez-Guerra, N.; González-Cobos, J.; Serrano-Ruiz, J. C.; Valverde, J.


L.; De Lucas-Consuegra, A. Electrochemical Activation of Ni Catalysts with
Potassium Ionic Conductors for CO2 Hydrogenation, Topics in Catalysis, 58 (2015)
1256-1269.

[63] Kalaitzidou, I.; Katsaounis, A.; Norby, T.; Vayenas, C. G. Electrochemical


promotion of the hydrogenation of CO2 on Ru deposited on a BZY proton
conductor, Journal of Catalysis, 331 (2015) 98-109.

45
CHAPTER 1
INFLUENCE OF THE PREPARATION

METHOD OF Pd/ZnO CATALYSTS


CHAPTER 1

Abstract

The aim of the work described here was to evaluate the catalytic
performance of palladium catalysts supported on zinc oxide (Pd/ZnO) in the
hydrogenation of CO2 to obtain methanol at atmospheric pressure. The influence
of the reduction temperature, calcination conditions, metal loading and Pd
precursor on the catalytic performance was studied.

Influence of

Reduction Calcination

5.0
5.0
4.5 N-10-F-500 N-10-F-500
4.5
Methanol formation rate

N-10-F-425

Methanol formation rate


4.0 4.0 N-10-S-500
N-10-F-350
(mol min g )

3.5

(mol min g )
-1 -1

3.5

-1 -1
3.0 3.0
2.5 2.5
2.0 2.0
1.5 1.5
1.0 1.0
0.5
0.5
0.0
0.0
150 175 200 225 250 275 300 150 175 200 225 250 275 300 325
Temperature (ºC) Temperature (ºC)

Pd loading Pd precursor

7 5.0
N-10-F-500 4.5 N-10-F-773
Methanol formation rate

6
-1 -1formation rate

N-18-F-500 4.0 T-10-F-773


(mol min g )

(mol min g )

5
-1 -1

-1 -1

3.5
2.25
4 3.0
2.00
2.5
CH4 formation rate

3 1.75
(mol min g )

2.0
Methanol

1.50
2 1.5
1.25
1.0
1 1.00
0.5
0 0.75
0.0
150 175 200 225 250 275 300 0.50 125 150 175 200 225 250 275 300 325
0.25
Temperature (ºC) Temperature (ºC)
0.00
400 425 450 475 500 525 550 575 600
Temperature (K)

10
9
0.25
CH4 formation rate

8
(mol min g )

CH4 formation rate


-1 -1

0.20 7
g )
-1 -1

6
0.15 5
(mol min

4
0.10
49 3
2
0.05
1
0.00 0
150 175 200 225 250 275 300 400 425 450 475 500 525 550 575

Temperature (ºC) Temperature (K)


CHAPTER 1

1. Introduction

The increase in atmospheric contamination and the desire to reduce the


use of fossil fuels in recent decades have led to a search for alternative fuels to
meet energy demands [1].

Hydrogen has been considered as a possible alternative fuel in recent


years. However, the difficulties associated with the storage of hydrogen, along
with the absence of energy distribution infrastructure and its low energy density
for most practical applications, have severely limited the use of hydrogen [2].
However, it can be used as a promising energy carrier, as it could be transformed
into other compounds. Methanol represents one way to store hydrogen in the
liquid state and this overcomes the problems outlined above. Methanol is
frequently used as solvent and feedstock for chemicals production. Furthermore,
it could be used like an alternative fuel in the energy distribution infrastructure
that exists nowadays, or it can also be blended with gasoline [3]. Methanol is
therefore one of the alternative fuels that is widely studied by the scientific
community [4].

At present, most commercial methanol is produced from synthetic gas or


syngas (H2 and CO), which is obtained by catalytic reforming of fossil fuels [3].
Nevertheless, the real scientific target is to replace carbon monoxide with carbon
dioxide, as shown in Eq. (1.1), and to obtain good conversion and selectivity
towards methanol [5].

CO2 + 3H2 ⇆ CH3 OH + H2 O ∆H25°C = –49.5 kJ∙mol-1 (1.1)

On the one hand, the use of carbon dioxide for the production of methanol
is one of the ways in which value can be added to this pollutant, which is closely
associated with the greenhouse effect. Carbon dioxide would be captured by
different methods and this would decrease the levels of this pollutant in the
atmosphere [6]. On the other hand, the use of hydrogen in that reaction limits the

50
CHAPTER 1

economy of the process, but it could change if the hydrogen is produced by (for
instance) the hydrolysis of water, which is an abundant component in nature, and
the energy required for the hydrolysis is obtained from renewable sources like
solar or wind power [7]. Accordingly, the methanol produced could be considered
as a ‘green fuel’ because it would be obtained from renewable energy and the
net balance of carbon dioxide in the atmosphere is not increased if the methanol
production utilizes more CO2 than that produced in the manufacture of H2 [8].

Despite the fact that methanol synthesis has been widely studied, in most
references in the literature this reaction is carried out at high pressure [8–11].
Studies carried out at atmospheric pressure are scarcer [12, 13] and in-depth
studies of the preparation steps for the catalysts employed for the reaction have
not been described. In the work reported here, the hydrogenation of CO2 to
methanol over Pd/ZnO catalysts at atmospheric pressure was carried out. This
catalyst was selected because it is well known to give good performance in this
reaction. The influence of the calcination and reduction steps, the metal loading
and the Pd salt precursor on the catalytic activity were studied.

2. Experimental

2.1. Catalyst preparation

Catalysts were prepared by the wet impregnation method using ZnO


(Panreac) as the support and palladium (II) nitrate [Pd(NO3)2.xH2O, Aldrich] or
tetraamminepalladium (II) nitrate [Pd(NH3)4(NO3)2, Aldrich] as precursors, with
the appropriate quantities added to an aqueous solution in order to obtain
catalysts with a Pd load of 10 or 15 wt%. After impregnation, the catalysts were
dried at 120ºC overnight. Two different methods were used for the calcination
step. The ‘slow calcination’ was carried out in a tubular reactor (75 cm length
divided at the middle into two parts with two different diameters, 1.5 and 0.7
cm, respectively) under an air stream (333 mL/min). The temperature was first
increased to 180ºC at a heating rate of 1.3 ºC/min and then held at this
51
CHAPTER 1

temperature for 2 h before being increased again up to 500ºC, at the same


heating rate, and then held for a further 2 h. The ‘fast calcination’ was carried
out in an oven under static air at 500ºC for 3 h at a heating rate of 5 ºC/min.
Finally, a reduction step was carried out prior to reaction and this is discussed in
Sect. 2.3. The catalysts were denoted as X-number1-Ynumber2, where X indicates
the precursor palladium (II) nitrate (N) or tetraamminepalladium (II) nitrate (T),
number1 the metal loading, Y the calcination method, fast (F) or slow (S), and
number2 the reduction temperature. For ease of reference the preparation
conditions and nomenclature of the samples are summarized in Table 1.1.
Table 1.1. Nomenclature and preparation method of the samples.

Metal
Reduction
Precursor loading Calcination step Nomenclature
temperature
(%)
350 ºC N-10-F-350
Fast Calcination 425 ºC N-10-F-425
10.9
Palladium (II) nitrate 500 ºC N-10-F-500
Slow Calcination 500 ºC N-10-S-500
18.1 Fast Calcination 500 ºC N-18-F-500
Tetraamminepalladium (II)
9.9 Fast Calcination 500 ºC T-10-F-500
nitrate

2.2. Catalyst characterization

Pd metal loading was determined by atomic absorption (AA)


spectrophotometry on a SPECTRA 220FS analyser. Samples (ca. 0.5 g) were
treated with 2 mL HCl, 3 mL HF and 2 mL H2O2 followed by microwave digestion
(250ºC). Surface area/porosity measurements were carried out using a
QUADRASORB 3SI sorptometer apparatus with N2 as the sorbate at -196ºC. The
samples were outgassed at 250ºC under vacuum (5x10-3 Torr) for 12 h prior to
analysis. Specific surface areas were determined by the multi-point BET method.
Specific total pore volume was evaluated from N2 uptake at a relative pressure
of P/Po = 0.99. Temperature-programmed reduction (TPR) experiments were
conducted in a commercial Micromeritics AutoChem 2950 HP unit with TCD
detection. Samples (ca. 0.15 g) were loaded into a U-shaped tube and ramped
52
CHAPTER 1

from room temperature to 700ºC (10 ºC/min), using a reducing gas mixture of
17.5 % v/v H2/Ar (60 cm3/min). The same unit was used to measure the metal
dispersion. Catalysts (150 mg) were typically purged with dry helium at 100 ºC
for 1 h and then reduced in 17.5 % v/v H2/Ar (25 cm3/min) at 500 ºC for 3 h.
Subsequently, the sample was cooled down to room temperature and exposed
to pulses of H2 until five consecutive pulses yielded identical signal areas. XRD
analyses were conducted with a Philips X’Pert instrument using nickel-filtered Cu-
Ka radiation. Samples were scanned at a rate of 0.02º step-1 over the range
5º≤ 2θ ≤ 90º (scan time = 2 s step-1). Transmission electron microscopy (TEM)
analyses were conducted on a JEOL JEM-4000EX unit with an accelerating
voltage of 400 kV. Samples were prepared by ultrasonic dispersion in acetone
with a drop of the resulting suspension evaporated onto a holey carbon-
supported grid. The instrument was equipped with an energy dispersive X-ray
spectroscopy (EDS) unit. Saturation was assumed to be complete after three
successive peaks showed the same peak areas. Thermogravimetric analysis was
carried out on a TGA apparatus (TGA-DSC 1, Mettler Toledo). The sample was
heated from ambient to 727ºC at 10 ºC/min under a reactive atmosphere of 21
% of oxygen and 79 % nitrogen. The TG curve represents the evolution of the
mass as a function of temperature.

2.3. Catalyst activity measurement

Experiments were carried out in a tubular quartz reactor (45 cm length


and 1 cm diameter). The catalyst, with a particle size in the range 250–500 lm
and without dilution, was placed on a fritted quartz plate located at the end of
the reactor. The temperature of the catalyst was measured with a K-type
thermocouple (Thermocoax) placed inside the inner quartz tube. The entire
reactor was placed in a furnace (Lenton) equipped with a temperature-
programmed system. Reaction gases were Praxair certified standards of CO2
(99.999 % purity), H2 (99.999 % purity) and N2 (99.999 % purity). The gas

53
CHAPTER 1

flows were controlled by a set of calibrated mass flowmeters (Brooks 5850 E and
5850 S).

Prior to the reaction, catalysts (0.8 g) were reduced in situ in a hydrogen


stream (10 vol %) diluted with nitrogen at a flow rate of 25 cm3/min. The
temperature was increased at a heating rate of 1.3 ºC/min up to different final
values, as shown in Table 1.1. The reduction step was finished once the final
reduction temperature was reached. The reaction was then carried out at
atmospheric pressure in the temperature range 150–300 ºC. The total flow rate,
which was a CO2/H2 mixture (CO2/H2 = 1/9), was maintained at 100 cm3/min.
Gas effluents were monitored by a micro gas chromatograph (Varian CP-4900).

3. Results and Discussion

3.1. Influence of the reduction temperature

The XRD patterns for different samples, before and after reduction, are
shown in Fig. 1.1. Prior to reduction (Fig. 1.1a) the main diffraction peaks
corresponded to ZnO and PdO. After reduction (Fig. 1.1b), the PdO peaks
disappeared and new diffraction peaks (around 2h = 41.22º and 44.14º) were
found and these are assigned to PdZn alloys. It can be observed that higher
reduction temperatures led to a higher intensity for these new peaks, indicating
that the formation of crystalline PdZn alloys is favored by higher reduction
temperatures. Methanol production takes place on PdZn alloy sites, whereas CO
is selectively produced on metallic Pd [12]. The methanol, CO and CH4 activities
at different temperatures are represented in Fig. 1.2. The catalytic results were
consistent with the XRD patterns. CH4 and CO activity increased when the catalyst
was reduced at a lower temperature (more metallic Pd and less PdZn alloys),
whereas methanol formation was favored by a higher reduction temperature. The
selectivity to methanol at the same CO2 conversion level (~4 %) was 15.7, 8.3
and 6.9 % for the N-10-F-500, N-10-F-425 and N-10-F-350 samples,
respectively. It shows that a major quantity of PdZn alloy provides a higher
54
CHAPTER 1

selectivity. It should be noted that, even though the catalytic experiments were
carried out under atmospheric pressure, the results are far away from the
thermodynamic equilibrium values, showing that there are not thermodynamic
limitations (see Table 1.2). The thermodynamic equilibrium values were calculated
using a flowsheet simulator (Aspen HYSYS V8.4 licensed by Aspen Technology,
Inc.). Peng Robinson was used as the equation of state and the reactor modeling
was based on a Gibbs reactor. The conditions used for the simulation (flow rate,
CO2/H2 ratio) were the same as in the experimental reactor.

a) b)
^

Intensity (a.u.)
Intensity (a.u.)

^
^ ^ N-10-F-500
^ ^

^ N-10-F-425
^
^
^ N-10-F-350
º ^
º º º º^ ^

20 30 40 50 60 70 80 90 40 41 42 43 44 45
2(º)
2(º)

N-10-F-500 N-10-F-500
c) N-10-S-500 dd))
N-18-F-500
Intensity (a.u.)

Intensity (a.u.)

40 41 42 43 44 45 40 41 42 43 44 45
2(º) 2(º)

Figure 1.1. XRD profiles of (a) N-10-F catalyst before reduction, where (cap symbol) denotes reflection
of ZnO and (degree symbol) denotes reflection of PdO. (b) N-10-F-350, N-10-F-425 and N-10-F-500.
(c) N-10-F -500 and N-10-S-500. d) N-10-F-500 and N-15-F-500 catalysts.

55
CHAPTER 1

5.0
4.5 a) N-10-F-500

Methanol formation rate


4.0 N-10-F-425

(mol min g )
N-10-F-350

-1 -1
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
b)
70
CO formation rate

60
(mol min g )
-1 -1

50
40
30
20
10
0
0.40 c)
0.35
CH4 formation rate
(mol min g )
-1 -1

0.30
0.25
0.20
0.15
0.10
0.05
0.00
150 175 200 225 250 275 300
Temperature (ºC)

Figure 1.2.Catalytic activity for N-10-F-350, N-10-F-425 y N-10-F-500. Formation rates of (a)
methanol (b) CO and (c) CH4. Reaction conditions: CO2/H2=1/9 and W/F=0.008 g min/cm3.

Table 1.2. Comparison between thermodynamic equilibrium and experimental values for the catalyst
N-10-F-500.
Thermodynamic Experimental
Temperature (ºC)
(µmol/min) (µmol/min)

150 22.24 0.10


175 14.72 0.21
200 10.09 0.55
225 7.14 1.22
250 5.19 1.58
275 3.87 0.85

56
CHAPTER 1

3.2. Influence of the calcination conditions

The catalyst N-10 was calcined using two different methods, slow and
fast, as described in Sect. 2.1. It should be noted that the calcination temperature
was chosen based on thermogravimetric (TG) analysis of the N-10 sample (Fig.
1.3). The marked weight loss at around 127–227 ºC is due to the removal of the
water absorbed during the impregnation step and to decomposition of the nitrate
precursor. The sample is stable in the range 427–727ºC.

Derivate weight loss (wt. %/ºC) x 10


100.0
2
97.5
95.0 1
92.5
% Weight

0
90.0
87.5 -1

85.0 -2
82.5
-3
80.0

4
77.5 -4
100 200 300 400 500 600 700
Temperature (ºC)

Figure 1.3. Thermogravimetric profile of N-10 catalyst.

Therefore, for the slow method, an initial step was carried out at 180ºC
to allow a gradual loss of water and decomposition of the metal precursor, and
a second step was then performed at 500ºC, i.e., the maximum reduction
temperature at which the sample is still stable.

The XRD profiles of samples obtained by the two calcination methods (Fig.
1.1c) were not significantly different. However, interestingly, the catalytic results
(Fig. 1.4) showed very clear differences between the two samples. The CO2
conversion was higher for the sample N-10-S-500, although consequently the
selectivity to methanol decreased. At the same CO2 conversion level (~7.8 %),
57
CHAPTER 1

the selectivity to methanol was of 0.95 and 0.77 % for the N-10-F-500 and N-
10-S-500 samples, respectively. Therefore, the methanol formation rate was
higher for the sample obtained by fast calcination, whereas the CO and CH4
activities were lower. Thus, PdZn alloy formation is not the only parameter that
affects the catalytic performance.
5.0
4.5 a) N-10-F-500
Methanol formation rate

4.0 N-10-S-500
3.5
(mol min g )
-1 -1

3.0
2.5
2.0
1.5
1.0
0.5

100 b)
90
CO formation rate

80
(mol min g )
-1 -1

70
60
50
40
30
20
10

2.25 c)
2.00
CH4 formation rate

1.75
(mol min g )
-1 -1

1.50
1.25
1.00
0.75
0.50
0.25
0.00
150 175 200 225 250 275 300 325
Temperature (ºC)

Figure 1.4. Catalytic activity for N-10-F-500 and N-10-S-500. Formation rates of (a) methanol (b) CO
and (c) CH4. Reaction conditions: CO2/H2= 1/9 and W/F=0.008 g min/cm3.

In an effort to understand this result, the catalysts were characterized in


more depth. The surface properties of the materials are given in Table 1.3.
58
CHAPTER 1

Surface area and pore volume are slightly higher for the sample obtained by
slow calcination, probably due to a different morphology created during the
calcination step. This change in the support morphology could be due to the faster
kinetic of the support degradation during the fast calcination. This degradation is
caused by the metal precursor, which can dissolve the ZnO (due the acidity of the
metal nitrate aqueous solution [14]). It would explain the different ZnO crystal
size for the N-10-F-500 and N-10-S-500 samples (98 and 189 nm, respectively).
Table 1.3. Main physical properties of the catalysts.

N-10-F N-10-S N-18-F T-10-F

Surface Area (m2/g) 8.6 9.5 7.1 7.5

Total pore volume (cm3/g) x 102 3.4 5.5 3.6 4.1

Dispersion (%) 1.47 1.59 0.82 0.34

Particle diameter PdZn from XRD (nm) 60.8 50.3 77.1 33.4

Particle diameter Pd0 from XRD (nm) - - - 63.9

Particle diameter ZnO from XRD (nm) 98 189 - -

Particle diameter from TEM (nm) 22.1 14.9 28.4 71.6

Moreover, metal dispersion is somewhat higher for the sample obtained


by slow calcination. These results are consistent with the TEM pictures (Fig. 1.5a,
b). The smallest particles found in the TEM images correspond to metallic Pd,
whereas the largest particles are assigned to PdZn alloys. It must be noted that
the TEM particle sizes shown in Table 1.3 were measured by taking into account
all Pd particles, i.e., without distinguishing between metallic Pd and PdZn alloys.
As in previous studies [15], the Pd was found to be dispersed on the ZnO matrix
with an average particle size of around 4 nm. It should be noted that the peaks
corresponding to metallic Pd did not appear in the XRD profiles for two reasons:
(i) the amount of Pd0 is less than 1 wt% and (ii) these particles are smaller than
2 nm, which is beyond the detection limit of XRD.

59
CHAPTER 1

a)
Region A

f)
27.5
25.0 N-10-F-500
N-10-S-500
22.5
20.0

Frequency (%)
17.5
15.0
12.5
b) 10.0
7.5
5.0
2.5
0.0
Region B 5-7.49 7.5-9.99 10-12.49 12.5-14.99 15-17.49 17.5-19.99 20-22.49 22.5-25 >25

Particle Size (nm)

g)
22.5
N-10-F-500
c) 20.0 N-18-F-500
17.5
Frequency (%)

15.0
12.5
10.0
PdZn Alloy 7.5
5.0
2.5
0.0
5-7.49 7.5-9.99 10-12.49 12.5-14.99 15-17.49 17.5-19.99 20-22.49 22.5-25 >25

d) Particle Size (nm)

h)

Region C 22.5
T-10-F-500
20.0
17.5
Frequency (%)

15.0
12.5
e) 10.0
7.5
5.0
2.5
0.0
15-19.99 25-29.99 35-39.99 45-49.99 55-59.99 65-69.99 75-79.99 85-89.99

Region D
Particle Size (nm)

Figure 1.5. High resolution TEM and metal particle distribution on (a) N-10-F-500 (b) N-10-S-500 (c)
N-18-F-500 and (d,e) T-10-F-500. (f) N-10-F-500 and N-10-S-500. (g) N-10-F-500 and N-18-F-500. (h)
T-10-F-500.

60
CHAPTER 1

Energy dispersive X-ray microanalysis (Table 1.4) was carried out to


demonstrate that the large particles correspond to PdZn alloys. Analysis in region
A detected Pd and Zn in similar quantities and this is related to the presence of
PdZn alloys. Region B showed only Zn and this was therefore assigned to the ZnO
support.
Table 1.4. Energy dispersive X-ray microanalysis results of region A, B, C and D from Fig.5.

Weight (%)
Element Region A Region B Region C Region D
Zn 57.70 99.7 47.54 31.97
Pd 42.30 0.3 52.46 68.03

The TEM images showed that smaller particles were present in the
catalyst calcined by the slow method. This means that there was more metallic Pd,
which is consistent with the higher CO and methane production observed.
Moreover, the N-10-S-500 catalyst showed in XRD and TEM analysis smaller
particle size for the PdZn alloy particles that has previously been correlated with
higher CO selectivity in the steam reforming of methanol [16, 17]. The metal
particle distribution (Fig. 1.5f) was also in agreement with this result. Both
catalysts showed a Gaussian particle distribution, but for the slow calcination
method this was shifted to the left. Therefore, the Pd particles in this catalyst,
related to the smaller particles in the Gaussian distribution, were found in major
proportion than those obtained for the fast calcination. In addition, the particles
related to the formation of PdZn alloys (bigger particles) and, consequently, to
the higher methanol production rate were detected in a lesser proportion when
the slow calcination method was employed. For instance, for the N-10-F-500
sample, the 8.6 % of the particles were larger than 25 nm. This effect could be
due to an easier mobility of the palladium cations. When the metal is loaded, it
is surrounded by H2O ligands coming from the aqueous solution. These ligands
are removed during the calcination step, and logically the faster heating rate,
the easier the ligands removal. When the metal becomes ‘‘bared’’, it can migrate
61
CHAPTER 1

to more stable positions [18]. It could explain the formation of bigger particles
of PdZn alloy.

TPR profiles obtained from room temperature to 700 ºC for these two
catalysts are shown in Fig. 1.6a. The two profiles are quite similar. The inverse
peak found at around 47 ºC is assigned to PdHx decomposition to give metallic
Pd. According to the literature [19, 20], when hydrogen is fed over the catalyst
at room temperature, hydrogen consumption occurs rapidly and PdO is partially
converted to PdHx [15]. This peak in hydrogen consumption was not observed in
our profiles, as hydrogen is in contact with the catalyst before the TPR experiment
starts. Interpretation of the peaks above 327ºC is not trivial, although they are
commonly related to crystalline PdZn alloy formation [15, 19, 20]. The two peaks
at around 327 and 477 ºC could be assigned to the different changes in the
Pd/ZnO crystal structure during the reduction process [21], where the Pd/ZnO
structure is modified in different steps as follows: PdO/ZnO → Pd/ZnO →
PdZnO1–x/ZnO → amorphous PdZn alloy/ZnO → crystalline PdZn alloy/ZnO.

a) b) c)
TCD Signal (a.u.)

N-10-F-500 N-10-F-500 N-10-F-500


N-10-S-500 N-18-F-500 T-10-F-500

0 100 200 300 400 500 600 700 100 200 300 400 500 600 700 100 200 300 400 500 600 700
Temperature (ºC)

Figure 1.6. TPR profiles. Comparison between (a) N-10-F and N-10-S. (b) N-10-F and N-18-F.
(c) N-10-F and T-10-F.

3.3. Influence of the metal loading

In order to study the influence of the metal loading, two samples (N-10-
F-500 and N-18-F-500) with different metal contents were compared. The

62
CHAPTER 1

catalytic activities obtained for these samples are shown in Fig. 1.7. As expected,
the methanol formation rate increased with the metal loading as there was more
Pd available to form PdZn alloys. This result is consistent with the XRD profiles
obtained for these samples (Fig. 1.1d), where the main peaks corresponding to
PdZn alloys were much more intense for the sample with a higher metal loading.
Moreover, an increase in the metal loading improved the selectivity to methanol,
as shown in Fig. 1.7d. A methanol selectivity of 100 % was obtained for the
catalyst with a metal content of 18 % at 150 ºC.

Figure 1.7. Catalytic activity for N-10-F-500 and N-18-F-773. Formation rates of (a) methanol (b) CO
and (c) CH4 (d) methanol selectivity and (e) methanol activity calculated per gram of Pd metal on the
catalyst for the different metal loadings. Reaction conditions: CO 2/H2=1/9 and W/F=0.008 g min/cm3.

The main physical properties of these catalysts can be compared from the
data in Table 1.3. The N-10-F-500 catalyst surface area was higher than that of
the ZnO support (7.0 m2 g-1). However, the surface area was lower for the N-
18-F-500 sample. Similar behavior has been described in the literature [14]. As

63
CHAPTER 1

commented above, it is caused by dissolution of the support (ZnO) during the


impregnation of Pd, so that the textural properties (porosity and crystal structure
of ZnO) are modified. When the metal content was increased, the Pd was
distributed in such a way that larger particles or alloys were formed, thus
increasing the volume of the particles and reducing the surface area of the
catalyst. This situation was confirmed by the lower metal dispersion found in the
catalyst with a higher metal loading. The TEM pictures and the corresponding
metal particle distribution were also in agreement with this situation. Large PdZn
alloy particles were found in the N-18-F-500 sample (Fig. 1.5c). The percentage
of particles greater than 25 nm for this catalyst was 10.7 %, whereas the value
for the N-10-F-500 sample was 8.6 %.

TPR profiles for these samples are shown in Fig. 1.6b. As mentioned in the
previous section, the catalyst with an 18 % metal loading had a lower quantity
of metallic palladium, which is related to the inverse peak at 52ºC, and more
crystalline PdZn alloys related to the peak at 477ºC. The low quantity of
available metallic Pd would explain the poor methanation activity, whereas the
higher quantity of PdZn alloys is responsible for the increase in methanol
production. However, it should be noted that even though there were more PdZn
alloy particles for the catalysts with a higher metal content, they were also larger
and a proportion of the Pd was inaccessible as it was trapped inside these
particles. Thus, if the methanol formation rate is calculated in terms of TOF
(turnover frequency) to take into account the available Pd sites, better results
were obtained for the N-10-F-500 catalyst (Fig. 1.7e), indicating that the loss of
available Pd sites is not compensated by the extra metal load.

3.4. Influence of the precursor

The performance levels of two catalysts (N-10-F-500 and T-10-F-500)


prepared with different precursors were compared. The catalytic results are

64
CHAPTER 1

represented in Fig. 1.8. As observed, the activity for the T-10-F-500 sample was
almost negligible.

5.0
a)

Methanol formation rate


4.5 N-10-F-500
4.0 T-10-F-500

(mol min g )
-1 -1
3.5
3.0
2.5
2.0
1.5
1.0
0.5
180
b)
160
140
CO formation rate
(mol min g )
-1 -1

120
100
80
60
40
20

9 c)
CH4 formation rate

8
(mol min g )

7
-1

6
-1

5
4
3
2
1
0
150 175 200 225 250 275 300
Temperature (ºC)

Figure 1.8. Catalytic activity for N-10-F-500 and T-10-F-500. Formation rates of (a) methanol (b) CO
and (c) CH4. Reaction conditions: CO2/H2= 1/9 and W/F=0.008 g min/cm3.

The XRD profiles for the two catalysts before and after the reduction step
are shown in Fig. 1.9. There are some interesting differences between the two
precursors. Before reduction of the N-10-F-500 sample, only the peaks assigned
to the support and the palladium oxide were observed. However, for the T-10-

65
CHAPTER 1

F-500 sample the peaks assigned to metallic palladium were also observed. For
this catalyst, the NH3 formed by degradation of the salt during the calcination
step was able to reduce the palladium, as it is well known [18]. After reduction,
the peaks assigned to PdZn alloys were observed for both samples, but they
were rather less intense for the T-10-F-500 catalyst, for which the signal
corresponding to metallic palladium was still observed.

Figure 1.9. XRD profile (a) before reduction and (b) after reduction of N-10-F-500 and T-10-F-500,
where (cap symbol) denotes reflection of ZnO, (degree symbol) denotes reflection of PdO, (plus sign)
denotes reflection of PdZn alloys and (asterisk) denotes reflection of metallic palladium.

66
CHAPTER 1

The TEM pictures for the latter sample are shown in Fig. 1.5d, e. It can be
clearly observed that the morphology of the particles changes completely when
compared to the other catalysts (much larger particles were detected). Energy
dispersive X-ray microanalysis was carried out to measure the element contents
in the two characteristic regions (C and D) for the T-10-F-500 sample (Table 1.4).
The larger particles would correspond to PdZn alloys (region C, where the metal
contents of Zn and Pd are very close) and to metallic palladium deposited on the
ZnO matrix (region D, where the palladium content is higher than that of zinc).
This situation is also consistent with the XRD profile of this catalyst, in which the
peaks for both species’ were detected. The formation of metallic Pd during the
calcination step and its subsequent sintering during the reduction step would
explain the formation of the large PdO particles. In any case, the metal particle
distribution for the T-10-F-500 sample (Fig. 1.5h) showed a high degree of
heterogeneity and particles of various different sizes were found. An average
particle size obtained from the XRD patterns and TEM images is given in Table
1.3. In agreement with the previous results, the metallic Pd particles showed a
much higher value than those determined for the other catalysts.

The TPR profile for this sample is shown in Fig. 1.6c. This profile is
completely different to those obtained for the other samples. There is a large
inverse peak and this confirms the high quantity of metallic palladium. After this,
only a single peak at around 377 ºC is observed. The second peak found for the
other catalysts at 477 ºC was not detected for this sample, which suggests that
the evolution from PdO/ZnO to a crystalline PdZn alloy/ZnO is incomplete and
that very little of the latter species is formed.

Therefore, for the T-10-F-500 there are smaller PdZn alloy particles and
most of these are amorphous and inactive to methanol production, meaning that
methanol is not obtained. In addition, the presence of larger metallic Pd particles
leads to higher CO and CH4 production.

67
CHAPTER 1

4. Conclusions

The following conclusions can be drawn from this study:

- The reduction temperature has a marked effect on the distribution of the species
that are formed. A higher reduction temperature leads to the formation of more
PdZn alloy particles. These alloy particles are directly related to a major
conversion towards methanol. Hence, an increase in the reduction temperature
improves the reaction rate and selectivity to methanol.

- The calcination conditions are important to control how particles are formed.
Smaller particles of metallic palladium were found after a slow calcination. These
particles are related to CO production and they decrease methanol selectivity
and production.

- A higher metal loading leads to higher methanol production as more PdZn alloy
particles are formed. However, the TOF is lower as these new particles are
bigger and the dispersion is poor.

- The precursor used to load the Pd has a marked influence on the final catalyst
structure. When tetraamminepalladium (II) nitrate was used, the ammonia formed
by salt degradation reduced the metal during the calcination step. Hence, during
the reduction step there is strong metal sintering. Numerous metallic Pd particles,
which have a large size, are obtained and the formation of PdZn alloy particles
is hindered. As a consequence, methanol production is suppressed.

References

[1] Guo, K. W. Current issues of fossil fuels and their future prospects, Fossil fuels:
sources, environmental concerns and waste management practices. Nova Science
Publishers, New York, 2013, 85-99

[2] Stranhan, D. Whatever happened to the hydrogen economy? New Scientist,


200 (2008) 40-43.

68
CHAPTER 1

[3] Olah, G. A. Beyond oil and gas: the methanol economy. Angewandte Chemie.
44 (2005) 2636-2639.

[4] Bockris, J. O. Hydrogen no longer a high cost solution to global warming. New
ideas. International Journal of Hydrogen Energy, 33 (2008) 2129-2131

[5] Ganesh, I. Conversion of carbón dioxide into methanol- a potential liquid fuel:
fundamental challenges and opportunities (a review). Renewable and sustainable
energy reviews, 31 (2014) 221-257.

[6] Spigarelli, B. P.; Kawatra, S. K. Opportunities and challenges in carbón


dioxide capture. Journal of CO2 utilization, 1 (2013) 69-87.

[7] Raudaskoski, R.; Turpeinen, E.; Lenkkeri, R.; Pongrácz, E.; Keiski, R. L. Catalytic
activation of CO2: use of secondary CO2 for the production of synthesis gas and
for methanol synthesis over copper-based zirconia-containing catalysts, Catalysis
Today, 144 (2009) 318-323.

[8] Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M. ; Joshi, J. B. Catalytic carbon


dioxide hydrogenation to methanol : a review of recent studies. Chemical
Engineering Research and Design, 92 (2014) 2557-2567.

[9] Ban, H.; Li, C.; Asami, K.; Fujimoto, K. Influence of rare-earth elements (La, Ce,
Nd and Pr) on the performance of Cu/Zn/Zr catalyst for CH3OH synthesis from
CO2. Catalysis Communications, 54 (2014) 50-54.

[10] Liang, X. L.; Xie, J. R. ; Liu, Z. M. A novel Pd-decorated carbon nanotubes-


promoted Pd-ZnO catalyst for CO2 hydrogenation to methanol. Catalysis Letters,
145 (2015) 1138-1147.

[11] Lei, H.; Nie, R.; Wu, G.; Hou, Z. Hydrogenation of CO2 to CH3OH over
Cu/ZnO catalysts with different ZnO morphology. Fuel, 154 (2015) 161-166.

69
CHAPTER 1

[12] Iwasa, N.; Suzuki, H.; Terashita, M.; Arai, M.; Takewaza, N. Methanol
synthesis from CO2 under atmospheric pressure over supported Pd catalysts,
Catalysis Letters, 96 (2004) 75-78

[13] Maniecki, T. P.; Mierczynski, P.; Maniukiewicz, W.; Bawolak, K.; Gebauer,
D.; Jozwiak, W. K. Bimetallic Au-Cu, Ag-Cu/CrAl3O6 catalysts for methanol
synthesis, Catalysis Letters, 130 (2009) 481-488.

[14] Chin, Y. H.; Wang, Y.; Dagle, R. A.; Li, X. S. Methanol steam reforming over
Pd/ZnO: Catalyst preparation and pretreatment studies. Fuel Processing, 83
(2003) 193-201.

[15] Chin, Y. H.; Dagle, R.; Hu, J.; Dohnalkova, A. C.; Wang, Y. Steam reforming
of methanol over highly active Pd/ZnO catalysts. Catalysis Today, 77 (2002) 79-
88.

[16] Zhang, H.; Sun, J.; Dagle, V. L.; Halevi, B.; Datye, A. K.; Wang, Y. Influence
of ZnO facets on Pd/ZnO catalysts for methanol steam reforming. ACS Catalysis,
4 (2014) 2379-2386.

[17] Dagle, R. A.; Chin, Y. H.; Wang, Y. The effects of PdZn crystallite size on
methanol steam reforming. Topics in catalysis, 46 (2007) 358-362.

[18] Sachtler W. M. H.; Zhang, Z. Zeolite-supported transition metals catalysts.


Advances in Catalysis, 39 (1993) 129-220.

[19] Iwasa, N.; Mayanagi, T.; Ogawa, N.; Sakata, K.; Takewaza, N. New
catalytic functions of Pd-Zn, Pd-Ga, Pd-In, Pt-Zn, Pt-Ga and Pt-In alloys in the
conversions of methanol. Catalysis Letters, 54 (1998) 119-123.

[20] Iwasa, N.; Masuda, S.; Ogawa, N.; Takezawa, N. Steam reforming of
methanol over Pd/ZnO: Effect of the formation of PdZn alloys upon the reaction.
Applied Catalysis A: General. 125 (1995) 145-157.

70
CHAPTER 1

[21] Wang, Y.; Zhang, J.; Xu, H. Interaction between Pd and ZnO during reduction
of Pd/ZnO catalyst for steam reforming of methanol to hydrogen. Chinese
Journal of Catalysis, 27 (2006) 217-222.

71
CHAPTER 2
EFFECT OF CARBON NANOFIBERS
SUPPORTS ON Pd/Zn CATALYSTS
CHAPTER 2

Abstract

Palladium/zinc catalysts supported on carbon nanofibers (CNFs) have


been used to study the catalytic performance in the hydrogenation of CO 2 to
obtain methanol at atmospheric pressure. The carbon nanofiber support has an
influence on the nature of the PdZn alloy formed. The effect of the Pd/Zn molar
ratio on the PdZn alloy particle size was analyzed. Lower Pd/Zn molar ratio
leads to higher PdZn alloy particle size, which was associated with higher
selectivity toward methanol. The influence of the type of nanofiber (platelet or
fishbone) on the catalytic behavior was also studied and compared with that of
a conventional Pd/ZnO catalyst. The palladium/zinc catalyst supported on
platelet nanofiber was considered to be a good candidate for the hydrogenation
of carbon dioxide to methanol.

Influence of

the carbon the Pd/Zn the orientation of


nanofibers support molar ratio graphene sheets

Molar
C H ratio
Pd/Zn Fishbone Platelet
O OH O OH
PdZn alloy
on the nature of the
particle size
PdZn alloy formed

75
CHAPTER 2

1. Introduction
World CO2 emissions from fuel combustion increased by 51.5% from
1990 to 2012 [1]. In addition, projections confirm that CO2 emissions from the
power sector will rise from 13.2 Gt (where Gt denotes gigatonnes) in 2012 to
15.4 Gt in 2040, retaining a share of ∼40% of global emissions over this period
[2]. Based on this information, it is important to focus our efforts on reducing the
levels of emissions. One way to mitigate the increase in CO2 in the atmosphere is
to exploit carbon dioxide to obtain products of value.

The hydrogenation of carbon dioxide has been assessed as one of the


reactions to produce added-value products such as hydrocarbons or alcohols [3].
In this reaction, hydrogen attacks the nonreactive CO2, as shown in the general
reaction described in Eq (2.1):
y
xCO2 + (2x-z+ 2) H2 →Cx Hy Oz +(2x-z)H2 O (2.1)

The use of hydrogen in this reaction limits the economy of the process;
however, this aspect could be changed if the hydrogen was obtained from
renewable sources [4]. One of the most important products obtained from the
aforementioned reaction is methanol (when x = 1, y = 4, and z = 1), as shown in
Eq (2.2):

CO2 + 3H2 ⇆CH3 OH+ H2 O ∆H25°C =-49.5 kJ/mol (2.2)

Methanol is frequently used as a solvent and a feedstock for the


production of chemicals. Furthermore, methanol could be used as an alternative
fuel in the energy distribution infrastructure that exists today, or it could be
blended with gasoline [5]. Moreover, methanol can be considered as a “green
fuel” if the CO2 net balance in the atmosphere does not increase when the
production of the methanol uses more CO2 than is produced in the manufacture
of H2 [6].

76
CHAPTER 2

The main byproducts obtained in the reaction to produce methanol are


carbon monoxide and methane, as shown in Eqs (2.3) and (2.4). One of the current
scientific targets is to obtain a catalyst that provides a high conversion and
selectivity toward methanol [7].

CO2 +H2 ⇆CO+ H2 O ∆H25°C =41kJ/mol (2.3)

CO2 +4H2 ⇆CH4 + 2H2 O ∆H25°C =-165 kJ/mol (2.4)

However, most of the methanol synthesized at present is produced at high


pressure and syngas is employed as the reactant. The real goal of this work is
the use of carbon dioxide and the transformation of this pollutant into methanol
at atmospheric pressure. Several studies [8-12] have been carried out on the
hydrogenation of CO2 to give methanol at atmospheric pressure. The two most
widely used active metals, because of their well-known properties and their high
efficacy in this reaction, are copper and palladium. In both cases, one of the most
commonly used supports used is ZnO. In the case of palladium, the support
interacts with the metal to form PdZn alloys and this behavior has been widely
studied in the steam reforming of methanol to hydrogen [13-18].

In the work described here, we combine the advantages offered by PdZn


alloys, in terms of the selectivity and activity in the formation of methanol from
the hydrogenation of CO2 and the advantageous characteristics of nanofibers,
which are used as the support. Carbon nanofibers (CNFs) are based on ordered
parallel graphene layers that are arranged in a specific conformation. These
materials have special properties (high mechanical strength, high surface area,
low internal masstransfer resistance, and surface defects for holding catalyst
particles) that make them useful in numerous applications, including
heterogeneous catalysis, where carbon materials generally have been widely
applied [19, 20].

77
CHAPTER 2

Many papers have been published on the use of carbon materials, carbon
nanotubes (CNTs) [21-26], and nanofibers [27] in the hydrogenation of CO2 to
methanol, but all of these references described the use of high pressures and
other active metals.

Two types of nanofibers were used in the study reported here, namely
fishbone and platelet, which were synthesized at 600 and 450 °C, respectively.
It has been reported [28] that the synthesis temperature has an influence of the
properties of CNFs. The nanofibers used in this work were reported previously in
other papers for the Fischer−Tropsch synthesis [29, 30] and the methanation of
CO and CO2 [31].

In the study described here, the influence of the support on the PdZn alloy
formed was studied and compared with a reference Pd/ZnO catalyst. Moreover,
catalysts with different Pd/Zn molar ratios were used for the hydrogenation of
CO2 to methanol.

This work is the first step in a project that is focused on the synthesis of
methanol with a feed of CO2 and H2O at atmospheric pressure and using a co-
ionic electrochemical membrane reactor. In this step, a catalyst that has a high
activity and selectivity toward methanol at higher temperatures is desired. Higher
temperatures lead to better ionic conductivity, which, in turn, favors the
electrochemical reactor performance. Nanofibers play a fundamental role in this
work, because of their excellent conductive properties.

2. Experimental

2.1. Support/Catalyst preparation

Different catalysts were used in the different sections of this work. The
main catalysts were prepared using carbon nanofibers as support. CNFs with
different crystalline structures were prepared by the catalytic decomposition of
ethylene over a Ni/SiO2 catalyst, at 600 °C in the case of fishbone-type

78
CHAPTER 2

nanofiber and at 450 °C for platelet-type nanofiber, according to a literature


procedure [28]. The carbon nanofibers were subsequently dissolved in
hydrofluoric acid (HF, 70%) for 15 h with vigorous stirring, in order to remove
any particles of the Ni/SiO2 catalyst and recover the carbon material. After this
treatment, the nanofibers were filtered off, washed, and dried at 110 °C for 12
h.

Hydrogenation catalysts were prepared by the wet impregnation


method. First, the support was placed in a glass vessel and kept under vacuum at
room temperature for 2 h to remove water and other impurities adsorbed on the
structure. Second, a solution of palladium (II) nitrate (Pd (NO3)2.xH2O, Aldrich)
and zinc nitrate hexahydrate (Zn (NO3)2.6H2O, Panreac) was poured over the
sample, with the appropriate quantities to obtain catalysts with a Pd load of 10
wt % and different Pd/Zn molar ratios. Third, the solvent was removed under
vacuum at 90 °C for 2 h. After impregnation, the catalysts were dried at 120 °C
overnight.

The calcination was carried out inside the reactor at 500 °C under a N2
atmosphere to prevent gasification of the CNFs. Prior to the reaction, catalysts
were reduced in situ in a hydrogen stream (10 vol %) diluted with nitrogen at a
flow rate of 25 cm3 min−1 at 500 °C with a heating rate of 1.3 °C min−1.

Five catalysts were prepared using nanofibers as support. Three


additional catalysts were prepared using the same methodology, in order to
compare them with those supported on nanofibers: (i) palladium on zinc oxide
[ZnO, Panreac] as support; (ii) copper(II) nitrate 3-hydrate [Cu(NO3)2.3H2O,
Panreac] on fishbone nanofibers; and (iii) palladium and zinc on aluminum oxide
[Al2O3, Alfa Aesar].

The catalysts were denoted as XY/Z, where X indicates the metal(s) used
as the active phase, Y the Pd/Zn molar ratio where appropriate, and Z the

79
CHAPTER 2

support used (abbreviated as Fish (fishbone), Plat (platelet), ZnO (zinc oxide),
and Al2O3 (aluminum oxide)).

2.2. Support/Catalyst characterization

Palladium and zinc metal loadings were determined by atomic absorption


(AA) spectrophotometry on a Spectra Model 220FS analyzer. Samples (ca. 0.5
g) were treated with 2 mL of HCl, 3 mL of HF, and 2 mL of H2O2, followed by
microwave digestion (250 °C). Surface area/porosity measurements were
carried out using a Quadrasorb Model 3SI sorptometer apparatus with N2 as the
sorbate at −196 °C. The samples were outgassed at 250 °C under vacuum (5 ×
10−3 Torr) for 12 h prior to analysis. Specific surface areas were determined by
the multipoint BET method. Specific total pore volume was evaluated from N2
uptake at a relative pressure of P/P0 = 0.99. Temperature-programmed
reduction (TPR) experiments were conducted in a commercial Micromeritics
AutoChem 2950 HP unit with TCD detection. Samples (ca. 0.15 g) were loaded
into a Ushaped tube and ramped from room temperature to 900 °C (10 °C
min−1), using a reducing gas mixture of 17.5 vol % H2/Ar (60 cm3 min−1).
Temperature-programmed decomposition (He TPD, H2 TPD, and CO2 TPD)
analyses were conducted in the same unit. He TPD analysis was carried out using
the same procedure as that explained for the TPR, but with the gas changed to
helium (99.999% purity, 60 cm3 min−1). In the case of H2 TPD, prior to the analysis,
the sample was prereduced in situ by a gas mixture of 17.5 vol % H2/Ar (30 cm3
min−1) at 500°C with a heating rate of 2.8 °C min−1, and then flushed by an
argon stream (99.999% purity, 30 cm3 min−1) at 500 °C for 30 min to clean the
surface, followed by cooling to 160 °C, switching to a H2 (99.999% purity)
stream for hydrogen adsorption at 160 °C for 30 min and subsequently at room
temperature for 480 min, and then flushing with the argon stream at room
temperature until a stable baseline was observed. The TPD measurement was
then conducted from room temperature to 900 °C with a heating rate of 10 °C

80
CHAPTER 2

min−1. CO2 TPD was conducted as explained for the H2 TPD; however, after
cleaning the surface, it is followed by cooling to 50 °C, switching to a CO2
(99.999% purity) stream for carbon dioxide adsorption at 50 °C for 30 min, and
then flushing with the argon stream at room temperature until a stable baseline
was observed. The TPD measurement was then conducted from room temperature
to 900 °C with a heating rate of 10 °C min−1. The crystallinity of CNFs, the mean
crystal size, and the Pd and Zn species were determined by X-ray diffraction
(XRD) analyses. The XRD experiments were conducted with a Philips X’Pert
instrument using nickel-filtered Cu Kα radiation. Samples were scanned at a rate
of 0.02° step−1 over a range of 5° ≤ 2θ ≤ 90° (scan time = 2 s step−1). To
complete these measurements, transmission electron microscopy (TEM) analyses
were carried out on a JEOL Model JEM-4000EX unit with an accelerating voltage
of 400 kV. Samples were prepared by ultrasonic dispersion in acetone with a
drop of the resulting suspension evaporated onto a holey carbon-supported grid.
The external morphology of the different catalyst was evaluated using a Phenom
Pro X scanning electron microscopy (SEM) system. This instrument was equipped
with an energydispersive X-ray spectroscopy (EDX) analyzer to determine the
average composition of the samples. Raman spectra of the catalysts were
recorded in a SENTERRA Raman spectrometer with 600 lines per mm grating and
a laser wavelength of 532 nm at a very low laser power level (<1 mW) to avoid
any heating effect.

2.3. Catalyst activity measurement

Catalytic performance tests were carried out in a tubular quartz reactor


(45 cm length and 1 cm diameter). The catalyst (0.8 g), which had a particle size
in the range of 250−500 μm and was not diluted, was placed on a fritted quartz
plate located at the end of the reactor. The temperature of the catalyst was
measured with a Type K thermocouple (Thermocoax) placed inside the inner
quartz tube. The entire reactor was placed in a furnace (Lenton) equipped with
81
CHAPTER 2

a temperature-programmed system. Reaction gases were Praxair certified


standards of CO2 (99.999% purity), H2 (99.999% purity), and N2 (99.999%
purity). The gas flows were controlled by a set of calibrated mass flowmeters
(Brooks, Models 5850 E and 5850 S).

The hydrogenation of CO2 was carried out at atmospheric pressure in the


temperature range of 150−300 °C. The total flow rate, which was a CO2/H2
mixture (CO2/H2 = 1/9), was maintained at 100 cm3 min−1. Gas effluents were
monitored with a micro gas chromatograph (Varian CP-4900), which contained a
PoraPLOT Q column and a molecular sieve column, each of which was connected
to a thermal conductivity detection (TCD) system.

3. Results and discussion

3.1. Influence of the carbon nanofiber (CNF) support on the PdZn alloys

The XRD patterns for the samples PdZn0.13/Fish and Pd/Fish before the
reduction step are shown in Fig.2.1a and 2.1b, respectively. These samples were
selected as references, but it should be noted that the same behavior was found
for all of the PdZn/nanofiber catalysts. After the calcination step, the main
diffraction peaks that one would expect to find should correspond to ZnO and
PdO. However, these peaks were not observed at all. The main peaks found
correspond to (a) metallic palladium (Joint Committee on Powder Diffraction
Standards (JCPDS) File No. 87-0645) for Pd/Fish, and (b) PdZn alloy with a
Pd/Zn ratio of 1:1 (JCPDS File No. 06-0620) for PdZn0.13/Fish. In an effort to
understand this result, the catalysts and the support were characterized in greater
depth.

CNFs are well-known to have oxygen functional groups on their surface


[32−34]. Despite the fact that the support was not previously chemically
activated with an activating agent (typically alkaline/earth-alkaline hydroxides,
as well as different acids or chlorides) to incorporate oxygen groups on its

82
CHAPTER 2

surface, a He TPD from room temperature to 900 °C was performed to


investigate the nature of the possible surface groups (Fig. 2.2). The peaks found
at the maximum temperature in the range of 70−627 °C correspond to the
decomposition of (mainly) carboxylic and lactonic groups to CO2, and the peak
at 700− 900 °C is attributed in the literature to decomposition, mainly to CO, of
phenolic, carbonyl, anhydride, ether, and quinone groups [34]. Since a reducing
agent (CO) is not likely to be formed at the calcination temperature of 500 °C,
and according to recent studies [35, 36], it is believed that carbon itself can act
as a reducing agent.

CNF Fishbone support Al2O3 support


a) +
c)
+ ZnO + ZnO
* Metallic palladium
º PdO
º PdZn alloy
+ º
+ + +
Intensity (a.u.)

+
+ +
º + + +º + + +
* º º + +º +º + º + º º + º +

* * CNF Fishbone support


b) d)
* Metallic copper
^ Cu2O
* *
* *
*
^ ^
*

20 30 40 50 60 70 80 90 10 20 30 40 50 60 70 80 90
2(º) 2(º)

Figure 2.1. XRD profiles before the reduction step of (a) PdZn0.13/Fish, (b) Pd/Fish, (c) Pd/Al2O3 and
(d) Cu/Fish.

1000
900
800
Intensity (a.u.)

700
Temperature (ºC)

600
500
400
300
200
100
0
0 10 20 30 40 50 60 70 80 90 100
Time (min)

Figure 2.2. Temperature-programmed decomposition in helium (He TPD) of the fishbone carbon
nanofiber (CNF) support.

83
CHAPTER 2

In an effort to confirm this theory, two further catalysts were prepared to


investigate why the metal was reduced during the calcination step. A
PdZn0.13/Al2O3 catalyst was prepared in an effort to ascertain whether the
PdZn alloy is formed on the support during the calcination process. The XRD
pattern of this sample (Fig. 2.1c) showed that the PdZn alloy had not been
formed. A Cu/Fish catalyst was also prepared in order to determine whether the
same behavior occurs with other metals (Fig. 2.1d). The main peaks correspond
to metallic copper (JCPDS File No. 85-1326), which indicates that copper had
indeed been reduced. This result suggests that the carbon nanofibers are the only
species responsible for the formation of PdZn and metallic palladium.

Since the PdZn alloy is formed without a reduction step, and the PdZn
alloy is the active phase for the production of methanol, the PdZn0.13/Fish prior
to reduction was used as a reference to determine whether the reduction step is
necessary in the process. The characterization of this catalyst will be described in
the next section, along with the corresponding catalytic results for the
hydrogenation of carbon dioxide.

3.2. Influence of the Pd/Zn molar ratio

Four catalysts were prepared in order to study the influence that the
Pd/Zn molar ratio has on the hydrogenation of carbon dioxide. The XRD results
for the different catalysts are shown in Fig. 2.3, including those for the Pd/Fish
used as a reference, to confirm the PdZn alloy behavior for methanol production.
The decrease in the Pd/Zn molar ratio means that the main peaks are due to
PdZn alloy in the case of PdZn0.75/Fish (Fig. 2.3c) and ZnO in the case of
PdZn0.13/Fish (Fig. 2.3a). Furthermore, the diffractograms of the PdZn0.13/Fish
sample before and after reduction overlap each other and they appear to be
very similar, in relative terms. It can be seen from the enlargement that the
intensities of the main peaks for the PdZn alloy (JCPDS File No. 06-620) are
higher in the case of the reduced catalyst. A higher intensity in the XRD peaks is

84
CHAPTER 2

due to a larger amount of PdZn alloy in the catalyst and this, in turn, leads to a
higher methanol production [8].

CNF Fishbone support Reduced

Intensity (a.u.)
+ ZnO Not reduced
* Metallic palladium
º PdZn alloy
+
+ + º 39 40 41 42 43 44 45 46
+ 2(º) + +
Intensity (a.u.) a) º + + +
º º + º +º+
º
+
++ º + + +
b) º º +++ º + º+ +
º
º
c) + ++ + º+ + º ++ º +º +
*
* *
d) * *

20 30 40 50 60 70 80 90
2(º)

Figure 2.3. XRD profiles of (a) PdZn0.13/Fish before and after reduction, (b) PdZn0.25/Fish, (c)
PdZn0.75/Fish, and (d) Pd/Fish after reduction.

This situation is consistent with the catalytic results (Fig. 2.4), which show
that the reduced catalyst has higher activity and selectivity toward methanol.
Therefore, the nonreduced PdZn0.13/Fish was discarded. A reduction step is still
necessary in order to obtain higher methanol formation rates. It is in agreement
with the different changes in the Pd/ZnO crystal structure during the reduction
process [15], where it is modified in different steps as follows: PdO/ZnO →
Pd/ZnO → PdZnO1−x/ZnO → amorphous PdZn alloy/ZnO → crystalline PdZn
alloy/ZnO. Hence, the nonreduced PdZn0.13/Fish sample would not have
completed all these steps during the autoreduction and, consequently, it is still
necessary an additional reduction under H2 atmosphere. Moreover, Pd/Fish gave
rise to the lowest methanol formation rate and the highest methane formation
rate, because metallic palladium (JCPDS File No. 87-0645) leads to the
85
175 Pd/Fish

CO formation rate (m


150
125
100
75
CHAPTER 2 50
25
0
formation of carbon monoxide and methane
15 [37].
c) The Pd/Zn
PdZn0.13/Fish molar ratio also
not reduced

CH4 formation rate (mol min g )


-1
14 PdZn0.13/Fish reduced
13
modifies the catalytic performance. The methanol formation rate curve is

-1
PdZn0.25/Fish
12
PdZn0.75/Fish 11
displaced to higher temperatures when the Pd/Zn molar
Pd/Fish
ratio is decreased, but 10
9
8
the maximum of this curve does not change. The formation of byproducts (CO and 7
6
5
CH4) is favored by a higher Pd/Zn molar ratio and, as a consequence, the 4
3
2
selectivity to methanol is decreased. 1
0
150 175 200 225 250 275 300
7.0 Temperature (ºC)
Methanol formation rate (mol min g )
-1

6.5 a) PdZn0.13/Fish not reduced


-1

6.0 PdZn0.13/Fish reduced


5.5 PdZn0.25/Fish
5.0 PdZn0.75/Fish
4.5 Pd/Fish
4.0
3.5
3.0 d) PdZn0.13/Fish not reduced
50
2.5 PdZn0.13/Fish reduced
2.0 45
PdZn0.25/Fish
CO2 conversion (%)

1.5 40 PdZn0.75/Fish
1.0 35 Pd/Fish
0.5
0.0 30

250 b) PdZn0.13/Fish not reduced 25


CO formation rate (mol min g )
-1

PdZn0.13/Fish reduced 20
225
-1

PdZn0.25/Fish
200 15
PdZn0.75/Fish
175 Pd/Fish 10

150 5

125 0
45 e) PdZn0.13/Fish not reduced
100
PdZn0.13/Fish reduced
75 40
CH3OH selectivity (%)

PdZn0.25/Fish
50 35 PdZn0.75/Fish
Pd/Fish
25 30
0 25
15 c) PdZn0.13/Fish not reduced
CH4 formation rate (mol min g )
-1

14 PdZn0.13/Fish reduced 20
13
-1

PdZn0.25/Fish 15
12
11 PdZn0.75/Fish
10
10 Pd/Fish
9 5
8 0
7 150 175 200 225 250 275 300
6
5 Temperature (ºC)
4
3
2
1
0
150 175 200 225 250 275 300

Temperature (ºC)

Figure 2.4. Catalytic activity for the catalyst with different Pd/Zn molar ratios. Formation rates of (a)
methanol (b) CO and (c) CH4, (d) CO2 conversion and (e) methanol selectivity. Reaction conditions:
CO2/H2= 1/9 and W/F=0.008 g min/cm3.

50 d) PdZn0.13/Fish not reduced


PdZn0.13/Fish reduced
45
PdZn0.25/Fish
86
CO2 conversion (%)

40 PdZn0.75/Fish
35 Pd/Fish
30
25
20
15
10
5
0
45 e) PdZn0.13/Fish not reduced
PdZn0.13/Fish reduced
40
CH3OH selectivity (%)

PdZn0.25/Fish
35 PdZn0.75/Fish
Pd/Fish
30
25
20
15
10
5
CHAPTER 2

The results in Table 2.1 show that a lower Pd/Zn molar ratio leads to a
higher PdZn alloy particle size and to a lower surface area, total pore volume,
and average pore radius in the support. The presence of larger species on the
same surface can lead to blockage of the pores and this would lower the free
surface area. Higher Pd/Zn molar ratios lead to smaller particles, which are more
active and lead to higher CO2 conversion and lower selectivity toward methanol
at higher temperatures. This situation is consistent with the TEM analysis carried
out on the catalysts PdZn0.13/Fish (Fig. 2.5a) and PdZn0.75/Fish (Fig. 2.5b). It
can be appreciated that, in the former case, the PdZn alloy is supported on
carbon nanofibers and zinc oxide. As a consequence, the real support is a
combination of these two compounds. For the PdZn0.75/Fish sample (Fig. 2.5b),
all of the images indicate that the alloy is supported on nanofibers. The particle
size distribution is represented in Fig. 2.5d, where more than 500 particles were
measured. Both of the catalysts show a Gaussian particle distribution. The particle
distribution shows that the PdZn0.75/Fish sample has a much higher proportion
of smaller particles. Smaller particles imply a higher number of active sites and,
consequently, a higher conversion, which, at the same time, involves a lower
selectivity at higher temperatures. Interestingly, the maximum methanol formation
rate is approximately the same for the three PdZn/Fish samples. It suggests that,
from a certain number of PdZn alloy particles, the maximum methanol production
is reached, regardless of further changes on the Pd/Zn molar ratio.

CO2 TPD (Fig. 2.6a) and H2 TPD (Fig. 2.6b) were carried out in order to
study the adsorption behavior of these reactants on the catalysts. Desorption
quantities are listed in Table 2.2. In both cases, the catalyst exhibited two
desorption peaks. The first peak, at ∼50 °C, is attributed to the loss of weakly
adsorbed CO2 and H2; the second peak, which is observed in the temperature
range of 500−900 °C, is probably due to a combination of different peaks. It
has been reported in the literature that the strength of the CO2 binding is related
to the calcination temperature [27].
87
Table 2.1. Main physical properties of the Fishbone carbon nanofibers catalysts.
CHAPTER 2

Fishbone PdZn0.13/Fish not


PdZn0.13/Fish PdZn0.25/Fish PdZn0.75/Fish Pd/Fish
support reduced

Pd loading (%) - 12.98 12.98 12.25 11.25 15.4

Zn loading (%) - 45.72 45.72 30.22 9.68 -

Surface areaa (m2g-1) 116.58 56.66 53.39 82.77 103.0 97.9

Total pore volumeb (x 102 cm3 g-1) 61.45 24.05 20.15 40.58 52.0 34.9

88
Average pore radiusb (Å) 105.4 84.88 75.48 98.04 100.9 71.4

Particle diameter PdZn from XRD - - 56.1 54.3 26.8 43.6


before reaction (nm)

Particle diameter PdZn from XRD - - 62.3 55.9 31.4 45.2


after reaction (nm)

a Measurement error ~5 m2/g

b Measurement error of < 5%


CHAPTER 2

a)
d)
32.5
30.0 PdZn0.75/Fish
27.5 PdZn0.13/Fish
25.0

Frequency (%)
22.5
20.0
17.5
15.0
12.5
10.0
7.5
5.0
b) 2.5
0.0
5-7.49 10-12.49 15-17.49 20-22.49 > 25
Particle Size (nm)
e)
24
22 PdZn0.13/Fish
20 PdZn0.13/Pla
18 PdZnO

Frequency (%)
16
14
c) 12
10
8
6
4
2
0
5-7.49 10-12.49 15-17.49 20-22.49 > 25
Particle Size (nm)

f) g)

h) i)

Figure 2.5. High-resolution TEM images of (a) PdZn0.13/Fish, (b) PdZn0.75/Fish, and (c)
PdZn0.13/Plat. Metal particle distribution of (d) PdZn0.75/Fish and PdZn0.13/Fish and (e)
PdZn0.13/Fish, PdZn0.13/Plat, and Pd/ZnO. High-resolution TEM images of (f) fishbone support and
(g) platelet support. SEM images of (h) PdZn0.13/Fish and (i) PdZn0.13/Plat.
89
CHAPTER 2

1000
a)
900

Intensity (a.u.)
800
700

Temperature (ºC)
Intensity (a.u.)
600
0 2 4 6 8 10 500
Time (min) PdZn0.75/Fish 400
PdZn0.25/Fish
300
PdZn0.13/Fish
200
100
0
0 10 20 30 40 50 60 70 80 90 100
Time (min)
1000
b)
900
Intensity (a.u.)

800
700

Temperature (ºC)
Intensity (a.u.)

600
0 1 2 3 4 5 6 7 8 9 10
500
Time (min)
PdZn0.75/Fish 400
PdZn0.25/Fish 300
PdZn0.13/Fish
200
100
0
0 10 20 30 40 50 60 70 80 90 100
Time (min)

0.050 1.7
c)
1.6
0.045
CO2 adsorbed (mmol/g)

0.0166T 1.5
CO2 adsorbed = 0.0008e
H2 adsorbed (mmol/g)

0.040 1.4
R² = 0.9999
0.035 1.3

1.2
0.030
1.1
0.012T
0.025 H2 adsorbed= 0.0793e 1.0
R² = 0.9964 0.9
0.020
0.8
195 200 205 210 215 220 225 230 235 240 245 250 255
Temperature (ºC)

Figure 2.6 (a) CO2 TPD, (b) H2 TPD, and (c) exponential relationship between the reactants adsorbed
and the temperature at which the highest methanol formation rate was obtained for the catalysts with
different Pd/Zn molar ratios.

90
CHAPTER 2

Table 2.2. Total amounts of H2 and CO2 released in CO2 and H2 TPDs.

Catalyst CO2 (mmol/g) H2 (mmol/g)

PdZn0.13/Fish 0.047±0.003 1.603±0.034

PdZn0.25/Fish 0.031±0.002 1.151±0.027

PdZn0.75/Fish 0.021±0.002 0.880±0.025

In this work, however, the strength of the CO2 and H2 binding is related
to the Pd/Zn molar ratio. Although this adsorption does not have any influence on
the maximum methanol formation rate, there is a link between CO2 and H2
adsorption and the temperature at which the highest formation rate toward
methanol was obtained. This situation can clearly be seen in Fig. 6c, in which an
exponential link between reactant adsorption and the temperature with the
highest methanol formation rate can be observed. This behavior is probably due
to the equilibrium (Eq 2.2). If the maximum methanol formation rate is the same
but the concentration of the reactants on the surface of the catalyst increases due
to the effect of better adsorption, then that maximum should be obtained at a
higher temperature. An in-depth study using kinetic expressions would be
necessary to confirm this point, but an approximate link has been demonstrated.

TPR profiles are shown in Figure 2.7. The TPR of the sample Pd/Fish was
carried out in order to determine the differences that exist when Zn is absent from
the catalyst. The high inverse peak found at ∼50 °C is assigned to PdHx
decomposition to give metallic palladium. According to literature data [16, 18],
PdHx is formed rapidly when hydrogen is fed over the catalyst at room
temperature, with PdO converted to PdHx. In the profiles obtained in this work,
this H2 consumption peak is not observed, because the hydrogen is in contact with
the catalyst at room temperature before the TPR experiment starts. For the other
91
CHAPTER 2

catalysts, the inverse peak was observed to a lesser extent, which indicates that
the metallic palladium is present in all of the catalysts but at lower levels. The
large peak at ∼600 °C is associated with an onset of low-temperature
gasification of the support, which is catalyzed in the presence of metallic
palladium [38]. For the catalysts with Zn, the broad H2 consumption found in the
range of 400−600 °C can be attributed to the different changes that occur
during the formation of the PdZn alloys [15, 37], and the peaks at higher
temperatures can be assigned to the gasification of nanofibers and the different
degrees of carbon surface oxidation [39].

The Pd/Zn molar ratio that was selected as being the most appropriate
for the hydrogenation of carbon dioxide in this work was 0.13. The
PdZn0.13/Fish catalyst is able to work at higher temperatures with higher
selectivity toward methanol. As explained in the Introduction, one future objective
of the project is the deposition of the catalyst onto the cathode of an
electrochemical reactor, in which ions would move more easily at high
temperatures. As a consequence, PdZn0.13/Fish was selected for the next stage
of the study.

1000
900
800
700
Temperature (ºC)
Intensity (a.u.)

Pd/Fish
600
500

PdZn0.75/Fish 400
PdZn0.25/Fish 300
200
PdZn0.13/Fish
100
0
0 10 20 30 40 50 60 70 80 90 100 110 120
Time (min)

Figure 2.7. Temperature-programmed reduction (TPR) profiles. Comparison between the catalysts with
different Pd/Zn molar ratios.
92
CHAPTER 2

3.3. Influence of the different nanofiber supports

Two types of nanofibers were compared in this work: fishbone and


platelet. The only difference between these materials was the synthesis
temperature: 600 °C for fishbone and 450 °C for platelet. A Pd/ZnO catalyst
was also prepared as a reference sample in order to compare it with the
nanofiber-based catalysts. The reference catalyst was prepared using the best
preparation conditions identified in a previous study [37].

The XRD patterns of the platelet nanofiber-based catalysts before


reduction (not shown here) showed the same trend as found for the fishbone
nanofibers, where the carbon itself acted as a reducing agent.

The XRD patterns after reduction for the nanofiber-based catalysts are
shown in Fig. 2.8. The patterns are fairly similar but it can be seen from the
enlargement that the intensities of the main peaks for the PdZn alloy (JCPDS File
No. 87-0645) are higher for the PdZn0.13/Fish catalyst. This finding suggests a
higher level of crystalline PdZn alloy in this sample, which could be due to the
influence that the support has on the PdZn alloy formation. This influence is also
shown by the slight displacement to higher angles 2θ (deg) of the PdZn alloy
peaks for the PdZn0.13/Plat sample. It means that contraction between lattices
of the PdZn alloy occurs. It indicates that a change in the lattice parameters of
the PdZn alloy tetragonal system (JCPDS File No. 87-0645) has occurred. These
lattice parameters were calculated based on Bragg’s law (Eq. 2.5) and the
formula for the tetragonal system crystal structure (Eq. 2.6):

n λ=2 d sin θ (2.5)


a
d= (2.6)
a2
√h2 +k2 +l2 ( 2 )
c

where n is a positive integer, λ is the wavelength of the incident wave, d is the interplanar distance, θ is the
scattering angle, a is the lattice spacing of the cubic crystal and h, k, l are the Miller indices of the Bragg
plane.
93
CHAPTER 2

The lattice parameters changed from a = 0.410 nm and c = 0.335, for


the PdZn0.13/Fish catalyst, to a = 0.409 and c = 0.333, for the PdZn0.13/Plat
sample. According to Kazuki et al. [40], this contraction between lattices is related
to a slight decrease in the Pd/Zn ratio of the alloy. Therefore, the alloy formed
in the PdZn0.13/Plat catalyst has a higher amount of zinc in its structure.
Preliminary calculations of the percentage of zinc incorporated to the structure
were made using Vergard’s law, which has been used for other solid mixtures,
such as Ptmetal alloys [41, 42]. Note that, taking into account the limitations of
this law [43], this number is just a rough approximation. According to these
calculations, the new PdZn alloy formed in the PdZn0.13/Plat catalyst would be
Pd48.5Zn51.5.

PdZn0.13/Plat
PdZn0.13/Fish
Intensity (a.u.)

* CNF +
+ ZnO
Intensity (a.u.)

º PdZn alloy
º
+ 40 41 42 43 44 45 46
+ + 2(º)
+
+
º + + +
º
* º º + +º + +

20 30 40 50 60 70 80 90
2(º)

Figure 2.8. XRD profiles after the reduction step of PdZn0.13/Fish and PdZn0.13/Plat.

The main physical properties of the catalysts are provided in Table 2.3.
A lower PdZn alloy particle size was found for the platelet-based catalysts,
probably because of the higher surface area of this support. It could be expected
that this sample would provide higher conversions. However, the catalytic results
(Fig. 2.9 and Table 2.4) did not show a clear relationship between the PdZn alloy
94
CHAPTER 2

particle size and the catalyst performance. It seems clear that the different
supports play an important role in the interaction with the PdZn alloy, considered
as the active phase in the production of methanol. The CO2 conversion, methanol
selectivity, and methanol yield are shown in Table 2.4. The highlighted values
show that the Pd/ ZnO catalyst works better at lower temperatures, while the
PdZn0.13/Plat catalyst works better at higher temperatures. Additional
characterization was carried out in an effort to understand the differences
between the different supports.
Table 2.3. Main physical properties of the catalysts with different supports.

Catalysts PdZn0.13/Fish PdZn0.13/Plat PdZnO

Pd loading (wt.%) 12.98 10.92 10.90

Zn loading (wt.%) 45.72 40.20 -

Surface Area (m2g-1)a 53.39 72.20 8.60

Total pore volume (cm3g-1) x 102 b 20.15 24.91 3.40

Average pore radius (Å) b 75.48 68.99 79.10

Particle diameter PdZn from XRD before 56.1 38.6 60.8


reaction(nm)

Particle diameter PdZn from XRD after 62.3 42.3 -


reaction(nm)

Average interlayer spacing d002 (nm) 0.3406 0.3437 -

Number of grapheme planes in the crystallites 29.2 21.6 -


npgc

Average crystal domain size along a direction 9.95 7.41 -


perpendicular to the basal planes LC (nm)

H2 consumption TPR analysis (µmol/g) 139 741 208

Supports Fishbone Platelet ZnO

Surface Area (m2g-1)a 116.58 150.94 7.23


a Measuring errors ~5m2/g.
b Measuring errors < 5%.
c npg = LC/d002

95
CHAPTER 2

Methanol formation rate (mol min g )


-1
a) PdZn0.13/Fish

-1
5 PdZn0.13/Plat
Pd/ZnO
4

0
b) PdZn0.13/Fish
CO formation rate (mol min g )

90
-1

PdZn0.13/Plat
-1

80
Pd/ZnO
70
60
50
40
30
20
10
0
PdZn0.13/Fish
CH4 formation rate (mol min g )

0.35 c)
-1

PdZn0.13/Plat
-1

0.30 Pd/ZnO
0.25

0.20

0.15

0.10

0.05

0.00
150 175 200 225 250 275 300
Temperature (ºC)

Figure 2.9. Catalytic activity for the catalyst with different supports. Formation rates of (a) methanol
(b) CO and (c) CH4. Reaction conditions: CO2/H2= 1/9 and W/F=0.008 g min/cm3.

96
Table 2.4. Conversion, methanol selectivity and methanol yield comparison between the different supports.

PdZn0.13/Fish PdZn0.13/Plat PdZnO

CO2 CH3OH CO2 CH3OH CO2 CH3OH


Temperature CH3OH CH3OH CH3OH
conversion, selectivity, conversion, selectivity, conversion, selectivity,
(ºC) yield (%) yield (%) yield (%)
X (%) S (%) X (%) S (%) X (%) S (%)

150 0.20 8.14 1.66 0 0 0 0.05 43.23 2.35

175 0.61 6.44 3.95 0.08 33.95 2.77 0.21 27.80 5.75

97
200 1.51 8.54 12.85 0.28 25.22 7.05 0.52 25.60 13.40

225 2.73 12.52 34.17 0.66 22.63 15.01 1.27 22.81 28.98

250 4.58 7.34 33.63 1.68 20.95 35.16 2.45 15.79 38.67

275 8.90 1.77 15.72 3.29 12.14 39.99 4.40 4.74 20.85

300 17.25 0.33 5.76 6.04 3.93 23.73 7.95 0.98 7.79
CHAPTER 2
CHAPTER 2

As mentioned above, the difference between fishbone and platelet


nanofibers is the synthesis temperature. It is well known that an increase in the
reaction temperature leads to more crystalline structures [44]. The graphitic
character of the supports can be evaluated by XRD. The data for the interlayer
spacing (d002), average crystalline parameter (LC), and average number of
planes of graphite crystals (npg), which are listed in Table 2.3, provide a measure
of the structural order of the materials. The order increased with both decreasing
values of d002 and increasing values of LC and npg. Therefore, the structural order
is fishbone > platelet. A more disordered structural nature means that the
materials have a higher surface area (and certainly a higher surface area was
found for the more-disordered platelet nanofibers), more defects, exposed edge
planes and surface C−H/O−H groups. This structure could play an important role
in the PdZn alloy deposition and, consequently, on the catalytic activity. Structural
features were further assessed by Raman spectroscopy (see Fig. 2.10). Raman
spectra of the nanofibers supports exhibited two peaks, habitually denoted as
D- and G-bands, at ca. 1354 and 1600 cm−1, respectively. The D-band has been
attributed to the presence of defects and/or curvature in the carbon structure
[45, 46] while the G-band is associated with well-ordered structures [47, 48].
Thus, the relative intensities of D- and G-bands (ID/IG) can be used as an index
to assess graphitic character. The following structural order was found: fishbone
(ID/IG = 1.02) > platelet (ID/IG = 1.18). In addition, the differences found in the
values of intensity of the D-band mean that platelet and fishbone supports could
have different curvatures. It has been reported in the literature [49] that the
interaction of the transition metals with the nanofibers is dependent on their
curvature. Thus, this observation is more proof that the interaction with the PdZn
alloy is different for both supports.

TEM images at a resolution of 100 nm seem to show, at first glance, that


the PdZn0.13/Plat (Fig. 2.5c) and PdZn0.13/Fish (Fig. 2.5a) catalysts are quite
similar. However, the structures of these two materials are quite different, as
98
CHAPTER 2

shown in Fig. 2.5f and 2.5g. The platelet structure has hexagonal planes
perpendicular to the fiber axis, and the graphene layers of the fishbone structure
terminate on the surface with a defined inclination angle. Some previous papers
have shown that this orientation of graphene sheets can significantly influence the
catalytic behavior and the selectivity toward different compounds [50−53]. The
metal particle distribution (Fig. 2.5e) shows some characteristics that are consistent
with the PdZn alloy particle size obtained by XRD. The Gaussian distribution of
these particles is shifted to lower particle sizes in the order PdZn0.13/Plat <
PdZn0.13/Fish ≃ Pd/ZnO. SEM images for the PdZn0.13/Fish catalyst (Fig. 2.5h)
and for the PdZn0.13/Plat sample (Fig. 2.5i) show that the external morphology
of both catalysts is quite different. EDX analysis confirmed that the PdZn0.13/Fish
catalyst showed a higher percentage of palladium. It could be due to the position
of the PdZn alloy particles. These particles could be hidden in the pores of the
structure of the PdZn0.13/Plat catalyst, but are located on the surface for the
PdZn0.13/Fish catalyst. Such a phenomenon would have an effect on the catalytic
behavior, with the PdZn0.13/Fish sample being more active for the CO2
hydrogenation and, therefore, less selective toward methanol, because of the
greater level of PdZn alloy particles exposed.

PdZn0.13/Fish
Intensity (a.u.)

PdZn0.13/Plat

1100 1200 1300 1400 1500 1600 1700 1800


-1
Raman Shift (cm )

Figure 2.10. Raman spectra for the PdZn0.13/Fish (ID/IG = 1.02) and PdZn0.13/Plat (ID/IG = 1.18).
99
CHAPTER 2

The CO2 and H2 TPD profiles are shown in Fig. 2.11. The same exponential
relationship found in the previous section, but with an additional point, was again
observed, which confirmed the correlation between the amount of reactant
adsorbed and the temperature with the highest methanol production. The
nanofiber TPD profiles showed the same peaks as in the previous study. The
PdZn0.13/Plat catalyst gave the highest values for CO2 and H2 adsorptions. This
result is understandable, bearing in mind the previous discussion on the catalyst
structure. It is interesting to note that the Pd/ZnO sample did not adsorb any
reactant. Therefore, the adsorption of compounds is clearly influenced by the
support. In other carbon materials, such as carbon nanotubes (CNTs) [22, 25] it
has been observed that the H2 species adsorbed on the catalyst could generate
surface microenvironments with high stationary state concentrations of H-
adspecies on the catalyst. These active H-adspecies could be readily transferred
to active sites via hydrogen spillover, thus increasing the specific reaction rate of
the hydrogenation. In our case, such an increase in the reaction rate was not
observed but, as explained previously, it increases the temperature at which the
highest methanol rate is obtained (due to the displacement of the equilibrium).

TPR analysis (Fig. 2.12) showed the same peaks that were explained in
the previous section. The two peaks observed at ∼300−500 °C for the Pd/ZnO
catalyst are shifted to higher temperatures for the nanofiber-based samples. The
PdZn0.13/Plat catalyst gave rise to the largest peak located at the highest
temperature (∼600 °C). The peak displacement at higher temperatures can be
attributed to greater metal−support interaction, which is influenced by the
orientation of graphene sheets [54, 55].

On balance, we suggest that the combination of a different structural


nature with a different graphene sheets orientation, the different ratio Pd/Zn in
the alloy, higher reactant adsorption on the surface and a major metal−support

100
CHAPTER 2

interaction can explain the higher activity and selectivity toward methanol of the
PdZn0.13/Plat catalyst.
1000
a)
900

Intensity (a.u.)
800
700

Temperature (ºC)
Intensity (a.u.)
600
0 1 2 3 4 5 6 7 8 9 10 500
Time (min)
400
PdZn0.13/Plat
Pd/ZnO 300
PdZn0.13/Fish 200
100
0
0 10 20 30 40 50 60 70 80 90 100
Time (min) 1000
b)
900
Intensity (a.u.)

800
700

Temperature (ºC)
Intensity (a.u.)

600
0 1 2 3 4 5 6 7 8 9 10
500
Time (min)
400
PdZn0.13/Plat
Pd/ZnO 300
PdZn0.13/Fish 200
100
0
0 10 20 30 40 50 60 70 80 90 100
Time (min)
3.5
0.09 c) PdZn0.13/Plat
0.0187T
CO2 adsorbed = 0.0005e 3.0
0.08
CO2 adsorbed (mmol/g)

R² = 0.9912
H2 adsorbed (mmol/g)

0.07 0.0168T 2.5


H2 adsorbed= 0.028e
0.06 R² = 0.9474 2.0
0.05
PdZn0.13/Fish 1.5
0.04
PdZn0.75/Fish
0.03 1.0
PdZn0.25/Fish

0.02
0.5
200 210 220 230 240 250 260 270 280
Temperature (ºC)

Figure 2.11. (a) CO2 TPD and (b) H2 TPD. (c) Exponential link between the reactants adsorbed and the
temperature at which the highest methanol formation rate was obtained for the catalysts with
different supports.
101
CHAPTER 2

1000
900
PdZn0.13/Plat 800
700

Temperature (ºC)
Intensity (a.u.)
600
PdZn0.13/Fish 500
400
300
Pd/ZnO 200
100
0
0 10 20 30 40 50 60 70 80 90 100
Time (min)

Figure 2.12. Temperature-programmed reduction (TPR) profiles. Comparison between the catalysts
with different supports.
Methanol formation rate (mol min g )
-1

2.9 3.40
-1

2.8
3.35
2.7
3.30 CO2 Conversion (%)
2.6
2.5 3.25
2.4
3.20
2.3
2.2 3.15
2.1 3.10
2.0
3.05
1.9
1.8 3.00
0 5 10 15 20 25 30 35 40 45 50
Time (h)

Figure 2.13. Catalytic stability for the PdZn0.13/Plat catalyst. Reaction conditions: 275 °C, CO2/H2 =
1/9 and W/F = 0.008 g min cm−3.

102
CHAPTER 2

Finally, in order to check the stability of this catalyst, an additional


experiment of 48 h was carried out with the same reaction conditions (CO2/H2 =
1/9 and W/F = 0.008 g min cm−3) at the temperature where more methanol was
obtained (275 °C). The results (Fig. 2.13) show a high stability for the
PdZn0.13/Plat catalyst. The CO2 conversion was kept constant (relative change
of −1.4%), not increasing the undesirable products and obtaining similar values
of methanol formation rate (relative change of 2.6%). As a consequence of its
properties and stability, this is a promising catalyst to carry out this reaction in an
electrochemical reactor.

4. Conclusions

The following conclusions can be drawn from this study:

- The PdZn catalysts supported on fishbone and platelet nanofibers can be


autoreduced by the supports in an inert atmosphere of N2 at 500 °C. However,
the catalysts that were reduced after that point showed a better performance.

- A decrease in the Pd/Zn molar ratio leads to displacement of the methanol


formation rate curve toward higher temperatures. An increase in the Pd/Zn molar
ratio leads to the formation of smaller particles, which are more active and,
therefore, less selective to methanol at higher temperatures.

- Based on the equilibrium reaction, an exponential relation was found between


the amount of reactant (CO2 and H2) adsorbed on the catalyst and the
temperature at which the higher methanol formation rate is obtained. This relation
was tested for the fishbone nanofiber-based catalysts and was confirmed with
the platelet-based samples.

- Based on XRD analysis, an alloy with a different Pd/Zn ratio (Pd48.5Zn51.5) was
found in the platelet support.

- The higher activity and selectivity toward methanol at higher temperature for
the PdZn0.13/Plat catalyst was attributed to different factors: a higher reactant
103
CHAPTER 2

adsorption, a different ratio Pd/Zn in the alloy, and a greater metal−support


interaction, because of the different graphene sheet orientations.

- The PdZn0.13/Plat has demonstrated good stability in the production of


methanol in a 48 h experiment.

- The platelet nanofiber-based catalyst has properties that are good enough for
it to be used as the cathode in an electrochemical reaction for methanol synthesis.

References

[1] IEA Statistics. CO2 Emissions from Fuel Combustion, 2014 Edition; International
Energy Agency: Paris, 2014.

[2] World Energy Outlook 2014; International Energy Agency: Paris, 2014.

[3] Saeidi, S.; Amin, N. A. S.; Rahimpour, M. R. Hydrogenation of CO2 to value-


added products-A review and potential future developments. Journal of CO2
Utilization, 5 (2014) 66−81.

[4] Raudaskoski, R.; Turpeinen, E.; Lenkkeri, R.; Pongrácz, E.; Keiski, R. L. Catalytic
activation of CO2: Use of a secondary CO2 for the production of synthesis gas
and for methanol synthesis over copper-based zirconia-containing catalysts.
Catalysis Today, 144 (2009) 318−323.

[5] Olah, G. A. Beyond Oil and Gas: The Methanol Economy. Angewandte
Chemie International Edition, 44 (2005) 2636−2639.

[6] Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M.; Joshi, J. B. Catalytic carbon
dioxide hydrogenation to methanol: A review of recent studies. Chemical
Engineering Research and Design, 92 (2014) 2557−2567.

[7] Ganesh, I. Conversion of carbon dioxide into methanol-A potential liquid fuel:
Fundamental challenges and opportunities (a review). Renewable Sustainable
Energy Review, 31 (2014) 221−257.

104
CHAPTER 2

[8] Iwasa, N.; Suzuki, H.; Terashita, M.; Arai, M.; Takezawa, N. Methanol synthesis
from CO2 under atmospheric pressure over supported Pd catalysts. Catalysis
Letters, 96 (2004) 75−78.

[9] Fujita, S.i.; Usui, M.; Ohara, E.; Takezawa, N. Methanol synthesis from carbon
dioxide at atmospheric pressure over Cu/ZnO catalyst. Role of methoxide species
formed on ZnO support. Catalysis Letters, 13 (1992) 349−358.

[10] Fujita, S.; Usui, M.; Hanada, T.; Takezawa, N. Methanol synthesis from
CO2−H2 and from CO−H2 under atmospheric pressure over Pd and Cu catalysts.
Reaction Kinetics and Catalysis Letters, 56 (1995) 15−19.

[11] Fujita, S. I.; Kanamori, Y.; Satriyo, A. M.; Takezawa, N. Methanol synthesis
from CO2 over Cu/ZnO catalysts prepared from various coprecipitated
precursors. Catalysis Today, 45 (1998) 241−244.

[12] Maniecki, T. P.; Mierczynski, P.; Maniukiewicz, W.; Bawolak, K.; Gebauer,
D.; Jozwiak, W. K. Bimetallic Au-Cu, Ag-Cu/CrAl3O6 catalysts for methanol
synthesis. Catalysis Letters, 130 (2009) 481−488.

[13] Chin, Y. H.; Wang, Y.; Dagle, R. A.; Shari Li, X. S. Methanol steam reforming
over Pd/ZnO: Catalyst preparation and pretreatment studies. Fuel Processing
Technology, 83 (2003) 193−201.

[14] Chin, Y. H.; Dagle, R.; Hu, J.; Dohnalkova, A. C.; Wang, Y. Steam reforming
of methanol over highly active Pd/ZnO catalyst. Catalysis Today, 77 (2002)
79−88.

[15] Wang, Y.; Zhang, J.; Xu, H. Interaction between Pd and ZnO during
Reduction of Pd/ZnO Catalyst for Steam Reforming of Methanol to Hydrogen.
Chinese Journal of Catalysis, 27 (2006) 217−222.

105
CHAPTER 2

[16] Iwasa, N.; Mayanagi, T.; Ogawa, N.; Sakata, K.; Takezawa, N. New
catalytic functions of Pd-Zn, Pd-Ga, Pd-In, Pt-Zn, Pt-Ga and PtIn alloys in the
conversions of methanol. Catalysis Letters, 54 (1998) 119−123.

[17] Zhang, H.; Sun, J.; Dagle, V. L.; Halevi, B.; Datye, A. K.; Wang, Y. Influence
of ZnO facets on Pd/ZnO catalysts for methanol steam reforming. ACS Catalysis,
4 (2014) 2379−2386.

[18] Iwasa, N.; Masuda, S.; Ogawa, N.; Takezawa, N. Steam reforming of
methanol over Pd/ZnO: Effect of the formation of PdZn alloys upon the reaction.
Applied Catalysis-A, 125 (1995) 145−157.

[19] Su, D. S.; Perathoner, S.; Centi, G. Nanocarbons for the development of
advanced catalysts. Chemical Reviews 113 (2013) 5782-5816.

[20] Xiong, H.; Jewell, L. L.; Coville, N. J. Shaped Carbons as Supports for the
Catalytic Conversion of Syngas to Clean Fuels. ACS Catalysis, 5 (2015)
2640−2658.

[21] Liang, X. L.; Xie, J. R.; Liu, Z. M. A Novel Pd-decorated Carbon Nanotubes-
Promoted Pd-ZnO Catalyst for CO2 Hydrogenation to Methanol. Catalysis Letters
145 (2015) 1138−1147.

[22] Liang, X. L.; Dong, X.; Lin, G. D.; Zhang, H. B. Carbon nanotube-supported
Pd-ZnO catalyst for hydrogenation of CO2 to methanol. Applied Catalysis-B, 88
(2009) 315−322.

[23] Zhang, H. B.; Liang, X. L.; Dong, X.; Li, H. Y.; Lin, G. D. Multiwalled carbon
nanotubes as a novel promoter of catalysts for CO/CO2 hydrogenation to
alcohols. Catalysis Survive Asia, 13 (2009) 41−58.

[24] Dong, X.; Zhang, H. B.; Lin, G. D.; Yuan, Y. Z.; Tsai, K. R. Highly active CNT-
promoted Cu-ZnO−Al2O3 catalyst for methanol synthesis from H2/CO/CO2.
Catalysis Letters, 85 (2003) 237−246.
106
CHAPTER 2

[25] Kong, H.; Li, H. Y.; Lin, G. D.; Zhang, H. B. Pd-decorated CNT-promoted Pd-
Ga2O3 catalyst for hydrogenation of CO2 to methanol. Catalysis Letters, 141
(2011) 886−894.

[26] Wang, G.; Chen, L.; Sun, Y.; Wu, J.; Fu, M.; Ye, D. Carbon dioxide
hydrogenation to methanol over Cu/ZrO2/CNTs: effect of carbon surface
chemistry. RSC Advances, 5 (2015) 45320−45330.

[27] Ud Din, I.; Shaharun, M. S.; Subbarao, D.; Naeem, A. Synthesis,


characterization and activity pattern of carbon nanofibers based
copper/zirconia catalysts for carbon dioxide hydrogenation to methanol:
Influence of calcination temperature. Journal of Power Sources, 274 (2015)
619−628.

[28] Jiménez, V.; Nieto-Márquez, A.; Díaz, J. A.; Romero, R.; Sánchez, P.;
Valverde, J. L.; Romero, A. Pilot plant scale study of the influence of the operating
conditions in the production of carbon nanofibers. Industrial & Engineering
Chemistry Research, 48 (2009) 8407−8417.

[29] Díaz, J. A.; Akhavan, H.; Romero, A.; Garcia-Minguillan, A. M.; Romero, R.;
Giroir-Fendler, A.; Valverde, J. L. Cobalt and iron supported on carbon
nanofibers as catalysts for Fischer−Tropsch synthesis. Fuel Processing Technology,
128 (2014) 417−424.

[30] Díaz, J. A.; Martínez-Fernández, M.; Romero, A.; Valverde, J. L. Synthesis of


carbon nanofibers supported cobalt catalysts for Fischer− Tropsch process. Fuel,
111 (2013) 422−429.

[31] Jiménez, V.; Sánchez, P.; Panagiotopoulou, P.; Valverde, J. L.; Romero, A.
Methanation of CO, CO2 and selective methanation of CO, in mixtures of CO
and CO2, over ruthenium carbon nanofibers catalysts. Applied Catalysis-A, 390
(2010) 35−44.

107
CHAPTER 2

[32] Nieto-Márquez, A.; Jiménez, V.; Raboso, A. M.; Gil, S.; Romero, A.; Valverde,
J. L. Influence of the chemical activation of carbon nanofibers on their use as
catalyst support. Applied Catalysis-A, 393 (2011) 78−87.

[33] Aksoylu, A. E.; Freitas, M. M. A.; Figueiredo, J. L. Pt-Sn catalysts supported


on activated carbon. I. The effects of support modification and impregnation
strategy. Applied Catalysis-A, 192 (2000) 29−42.

[34] Figueiredo, J. L.; Pereira, M. F. R.; Freitas, M. M. A.; Órfão, J. J. M.


Modification of the surface chemistry of activated carbons. Carbon, 37 (1999)
1379−1389.

[35] Xiong, H.; Moyo, M.; Rayner, M. K.; Jewell, L. L.; Billing, D. G.; Coville, N. J.
Autoreduction and Catalytic Performance of a Cobalt Fischer−Tropsch Synthesis
Catalyst Supported on Nitrogen-Doped Carbon Spheres. ChemCatChem, 2
(2010) 514−518.

[36] Xiong, H.; Motchelaho, M. A. M.; Moyo, M.; Jewell, L. L.; Coville, N. J.
Influence of supports on catalytic behavior of nickel catalysts in carbon dioxide
reforming of toluene as a model compound of tar from biomass gasification.
Journal of Catalysis, 278 (2011) 26−40.

[37] Díez-Ramírez, J.; Valverde, J. L.; Sánchez, P.; Dorado, F. CO2 Hydrogenation
to methanol at atmospheric pressure: influence of the preparation method of
Pd/ZnO catalysts. Catalysis Letters 146 (2016) 373.

[38] Leino, A. R.; Mohl, M.; Kukkola, J.; Mäki-Arvela, P.; Kokkonen, T.; Shchukarev,
A.; Kordas, K. Low-temperature catalytic oxidation of multi-walled carbon
nanotubes. Carbon, 57 (2013) 99−107.

[39] Román-Martínez, M. C.; Cazorla-Amorós, D.; Linares-Solano, A.; de Lecea,


C.S.-M. TPD and TPR characterization of carbonaceous supports and Pt/C
catalysts. Carbon, 31 (1993) 895−902.

108
CHAPTER 2

[40] Nozawa, K.; Endo, N.; Kameoka, S.; Tsai, A. P.; Ishii, Y. Catalytic properties
dominated by electronic structures in PdZn, NiZn, and PtZn intermetallic
compounds. Journal of the Physical Society of Japan, 80 (2011) 064801-
064814.

[41] Antolini, E.; Cardellini, F. Formation of carbon supported PtRu alloys: An XRD
analysis. Journal of Alloys and Compounds, 315 (2001) 118−122.

[42] Hyun, K.; Lee, J. H.; Yoon, C. W.; Kwon, Y. The effect of platinum based
bimetallic electrocatalysts on oxygen reduction reaction of proton exchange
membrane fuel cells. International Journal of Electrochemical Science, 8 (2013)
11752−11767.

[43] Jacob, K. T.; Raj, S.; Rannesh, L. Vergard’s law: A fundamental relation or
an approximation? International Journal of Materials Research, 98 (2007)
776−779.

[44] Gil, S.; Muñoz, L.; Sánchez-Silva, L.; Romero, A.; Valverde, J. L. Synthesis and
characterization of Au supported on carbonaceous material-based catalysts for
the selective oxidation of glycerol. Chemical Engineering Journal, 172 (2011)
418−429.

[45] Choi, S.; Park, K. H.; Lee, S.; Koh, K. H. Raman spectra of nano-structured
carbon films synthesized using ammonia-containing feed gas. Journal of Applied
Physics, 92 (2002) 4007−4011.

[46] Dresselhaus, M. S.; Dresselhaus, R.; Saito, R.; Jorio, A. Raman spectroscopy
of carbon nanotubes. Physics Reports, 409 (2005) 47−99.

[47] Tuinstra, F.; Koenig, J. L. Raman spectrum of graphite. Journal of Chemical


Physics, 53 (1970) 1126.

109
CHAPTER 2

[48] Nemanich, R. J.; Solin, S. A. First- and second-order Raman scattering from
finite-size crystals of graphite. Physical Review B- Condensed Matter and
Materials, 20 (1979) 392−401.

[49] Menon, M.; Andriotis, A. N.; Froudakis, G. E. Curvature dependence of the


metal catalyst atom interaction with carbon nanotubes walls. Chemical Physics
Letters, 320 (2000) 425−434.

[50] Plomp, A.; Schubert, T.; Storr, U.; de Jong, K.; Bitter, J. Reducibility of
Platinum Supported on Nanostructured Carbons. Topics Catalysis, 52 (2009)
424−430.

[51] Chambers, A.; Nemes, T.; Rodriguez, N. M.; Baker, R. T. K. Catalytic Behavior
of Graphite Nanofiber Supported Nickel Particles. 1. Comparison with Other
Support Media. Journal of Physical Chemistry B, 102 (1998) 2251−2258.

[52] Chesnokov, V. V.; Prosvirin, I. P.; Zaitseva, N. A.; Zaikovskii, V. I.; Molchanov,
V. V. Effect of the Structure of Carbon Nanofibers on the State of an Active
Component and on the Catalytic Properties of Pd/C Catalysts in the Selective
Hydrogenation of 1,3-Butadiene. Kinetics and Catalysis, 43 (2002) 838−846.

[53] Ochoa-Fernández, E.; Chen, D.; Yu, Z.; Tøtdal, B.; Rønning, M.; Holmen, A.
Effect of carbon nanofiber-induced microstrain on the catalytic activity of Ni
crystals. Surface Science, 554 (2004) L107−L112.

[54] Zhao, T.-J.; Chen, D.; Dai, Y.-C.; Yuan, W.-K.; Holmen, A. The effect of
graphitic platelet orientation on the properties of carbon nanofiber supported
Pd catalysts prepared by ion exchange. Topics in Catalysis, 45 (2007) 87−91.

[55] Xiong, H.; Motchelaho, M. A. M.; Moyo, M.; Jewell, L. L.; Coville, N. J. Cobalt
catalysts supported on a micro-coil carbon in Fischer−Tropsch synthesis: A
comparison with CNTs and CNFs. Catalysis Today, 214 (2013) 50−60.

110
CHAPTER 3
INFLUENCE OF THE CALCINATION,
REDUCTION AND METAL LOADING
ON Cu/ZnO CATALYSTS
CHAPTER 3

Abstract
Cu/ZnO catalysts have been widely studied for the hydrogenation of
carbon dioxide to methanol at atmospheric pressure. In the work described here,
several interesting issues are highlighted that have rarely been considered
previously. An extensive study of the influence of the calcination and reduction
temperatures and the metal loading was carried out. The best conditions found
for catalyst preparation were calcination at 350 °C and reduction at 200 °C.
The role of the different oxidation states of copper (Cu2+, Cu1+, and Cu0) was
proven in the methane and methanol formation. CuZn alloy formation was
observed when a reduction temperature of 400 °C was used. The use of this alloy
led to higher methanol selectivity at higher temperatures (>200 °C). Finally, the
metal loading study confirm the dual-site nature of the methanol synthesis
mechanism.

Calcination Reduction temperature


Metal loading (%)
temperature (ºC) (ºC)

300 350 400 600 150 200-300 400 5 10

20 40
Copper particle size CuO Cu Cu Cu-Zn alloy
CH3OH formation rate (mol min g )
CH3OH formation rate (mol min g )

-1
CH3OH formation rate (mol min g )

-1
-1

5
-1

300 150
-1
-1

1.2 5
1.2 350 200
250 4 10
400 0.9 20
0.9 300
600 400 3 40
0.6
0.6
2
0.3
0.3
1
0.0
0.0 0
150 175 200 225 250 275 300 150 175 200 225 250 275 300 150 175 200 225 250 275 300
Temperature (ºC) Temperature (ºC) Temperature (ºC)

113
CHAPTER 3

1. Introduction
Cu/ZnO catalysts for the hydrogenation of carbon dioxide to give
methanol (Eq. 1) have been widely studied [1−24]. This reaction is a route for the
valorization of CO2 [25, 26], which is well-known for its influence on the
greenhouse effect and therefore on global warming. Cu/ZnO catalysts play an
important role in methanol production because of their high activity and selectivity
toward this valuable product. Moreover, compared to catalysts based on noble
metals, such as Pd/ ZnO [27, 28], Cu/ZnO catalysts are more cost-effective, and
this allows the transfer of the process from the laboratory to the industrial scale.
In this sense, the first pilot plant for the production of methanol from CO2 and H2
was built in Japan, and a SiO2-modified Cu/ZnO catalyst was used in the
hydrogenation process [29]. SiO2 was incorporated in this catalyst in order to
improve the Cu/ZnO catalytic performance. Other compounds, such as Al2O3 [30,
31] Ga2O3, or ZrO2 [32−35] have also been incorporated in an effort to increase
the methanol production (Eq. 3.1) while simultaneously decreasing the selectivity
to the undesired byproduct carbon monoxide (Eq. 3.2):

CO2 + 3H2 ⇆ CH3 OH + H2 O ∆H25°C = –49.5 kJ/mol (3.1)

CO2 + H2 ⇆ CO + H2 O ∆H25°C = 41 kJ/mol (3.2)

Numerous studies have been published on the Cu/ZnO characteristics


involved in the synthesis of methanol, which are the same that appear in the
reverse process of steam reforming of methanol [36−38], and the role played
by each of these. However, definitive conclusions remain elusive due to the large
number of variables involved in the synthesis of methanol, and a general
consensus has not been reached to date. It has to be taken into account that the
main findings about CO2 hydrogenation have been found in catalytic runs
performed under pressure. In general, some authors explain the results by
considering the area of copper metal obtained or the interaction with the support
114
CHAPTER 3

[11, 23, 35]. The oxidation state is also a controversial issue, as the nature of the
active phase is not clear: metallic copper, a ratio of metallic Cu and Cu+, Cu-Zn
sites, or CuZn alloys are some of the theories published to date [4−6, 12, 24, 39].
In addition, the catalytic behavior of the different surfaces, e . g . , Cu (111), Cu
(100), and Cu (110) [16, 17, 21], the surface structural changes during the
methanol synthesis [18, 19], the influence of the oxygen vacancies [15], the
presence of lattice defects [40] and the different precursor structures have also
been studied [9, 10, 41]. Based on all of the variables mentioned above, some
authors have studied the different pathways for the CO and CH3OH reactions
(Eq. 3.1and 3.2) along with the different intermediate species formed and how
they influence the methanol synthesis [3, 8, 15, 20, 22, 42]. Finally, models of the
kinetics for methanol synthesis and deactivation have also been developed
[43−46].

The work described here concerns the influence that the Cu/ZnO catalyst
has on the hydrogenation of carbon dioxide to methanol at atmospheric pressure.
As stated by Fujita et al. [1], very few studies have been published in this area,
probably due to commercial interests. Fujita et al. [1, 7] studied the influence of
the heating rate in the calcination step and different reduction methods, but they
did not study the influence of the calcination and reduction temperatures. The
different conditions used for the reduction/calcination steps in a range of papers
are summarized in Table 3.1. However, the influence of the calcination or
reduction temperature on the Cu/ZnO catalyst has not been studied previously in
any depth. For this reason, the work described here is intended to clarify these
points and to contribute to a better understanding of the Cu/ZnO catalyst.

115
CHAPTER 3

Table 3.1. Summary of preparation and reaction conditions using the catalyst Cu/ZnO.
Calcination Reduction CO2/H2 Reaction Reaction
Copper loading (wt.%) W/F (g min cm-3) Reference
temperature (ºC) temperature ratio pressure temperature (ºC)
(ºC)
350 170/250 50 0.005 1/9 Patm 167 [1]
350 250 30-50-60-70 0.005 1/9 Patm 165 [2]
350 250 30 0.0188 1/9 Patm 165 [3]

116
450 270 ~67 0.005 1/3 3 MPa 240 [4]
350 250 0.5-1-3-5-8-15 0.0075-0.03 1/9 7 bar 160-225 [5]
350 250 50 0.0033 1/3 5 MPa 250 [6]
350 170/250 50 0.005 1/9 Patm 167 [7]

450 250 33-50-66-75 0.005 ~1/3 27.2 bar 250 [8]


350 250 10-30-50-70-90-100 0.001-0.005-0.01-0.02 1/9 Patm 165-190-210 [9]
CHAPTER 3

2. Experimental

2.1. Catalyst preparation

Catalysts were prepared by the impregnation method using ZnO


(Panreac, 99% of minimum purity) as the support and copper (II) nitrate trihydrate
[Cu(NO3)2.3H2O, Panreac, 99.95% purity] as the precursor. First, the support was
placed in a glass vessel and kept under vacuum at room temperature (∼25 °C)
for 2 h to remove water and other impurities adsorbed on the structure. Second,
a solution of copper nitrate in distilled water was then poured over the sample,
with the appropriate quantities to obtain catalysts with Cu loadings of 5, 10, 20,
and 40 wt %. Third, the solvent was removed under vacuum at 90 °C for 2 h.
After impregnation, the catalysts were dried at 120 °C overnight.

The calcination was carried out in a furnace Nabertherm HTC 03/15,


which is open to atmospheric air, at different temperatures: 300, 350, 400, and
600 °C. Each temperature was kept constant for 4 h, and the heating rate was
3.5 °C min−1 in all cases. This heating rate was selected based on the paper by
Fujita et al. [7], which confirmed that a slow heating rate is necessary in order to
obtain a lower CuO crystal size and a higher methanol synthesis activity.

Prior to the reaction, the catalysts were reduced in situ for 2 h in a


hydrogen stream (10 vol %) diluted with nitrogen at a flow rate of 25 cm3 min−1
at different temperatures (150, 200, 250, 300, and 400 °C) at a heating rate
of 5 °C min−1.

The catalysts were denoted as XCuZnO-Y-Z, where X indicates the metal


loading, Y the calcination temperature, and Z the reduction temperature. In an
effort to facilitate identification, the preparation conditions and the nomenclature
of the samples are summarized in Table 3.2.

117
CHAPTER 3

Table 3.2. Nomenclature and preparation conditions of the samples.

Calcination Reduction
Metal loading (%) Nomenclature
temperature (ºC) temperature (ºC)
300 200 10CuZnO-300-200
150 10CuZnO-350-150
200 10CuZnO-350-200
350 250 10CuZnO-350-250
10.8
300 10CuZnO-350-300
400 10CuZnO-350-400
400 200 10CuZnO-400-200
600 200 10CuZnO-600-200
5.6 350 200 5CuZnO-350-200
20.6 350 200 20CuZnO-350-200
43.2 350 200 40CuZnO-350-200

2.2. Support/catalyst characterization


The Cu metal loading was determined by atomic absorption (AA)
spectrophotometry on a SPECTRA 220FS analyzer. Samples (ca. 0.5 g) were
treated with 2 mL of HCl, 3 mL of HF, and 2 mL of H2O2 followed by microwave
digestion (250 °C). Surface area/ porosity measurements were carried out using
a QUADRASORB 3SI sorptometer apparatus with N2 as the sorbate at −196 °C.
The samples were outgassed at 250 °C under vacuum (5 × 10−3 Torr) for 12 h
prior to analysis. Specific surface areas were determined by the multipoint BET
method. Specific total pore volume was evaluated from N2 uptake at a relative
pressure of P/P0 = 0.99. Temperature-programmed reduction (TPR) experiments
were conducted in a commercial Micromeritics AutoChem 2950 HP unit with TCD
detection. Samples (ca. 0.15 g for the catalyst with 10 wt % of copper) were
loaded into a U-shaped tube and ramped from room temperature (∼25 °C) to
900 °C (10 °C min−1), with a reducing gas mixture of 17.5% v/v H2/Ar (60 cm3
min−1). The XRD experiments were conducted with a Philips X’Pert instrument using
nickel-filtered Cu Kα radiation. Samples were scanned at a rate of 0.02° step−1
118
CHAPTER 3

over the range 5° ≤ 2θ ≤ 90° (scan time = 2 s step−1). Transmission electron


microscopy (TEM) analyses were carried out on a JEOL JEM-4000EX unit with an
accelerating voltage of 400 kV. Samples were prepared by ultrasonic dispersion
in acetone with a drop of the resulting suspension evaporated onto a holey
carbon-supported grid.

2.3. Equation section

In this paper, the lattice parameters of the ZnO hexagonal crystalline


structure and copper cubic structure were calculated from XRD diffractograms
using Bragg’s Law (Eq. 3.3), the formula for the hexagonal crystal structure (Eq.
3.4), and the formula for the cubic crystal structure (Eq. 3.5):

n λ=2 d sin θ (3.3)

1 4 h2 +hk+k2 l2
= ( ) + c2 (3.4)
d2 3 a2

1 h2 +k2 +l2
2= (3.5)
d a2

where n is a positive integer, λ is the wavelength of the incident wave, d


is the interplanar distance, θ is the scattering angle, a and c are the lattice spacing
of the cubic and hexagonal crystal, respectively, and h, k and l are the Miller
indices of the Bragg plane.

Particle size from TEM images was also calculated. Mean copper particle
size evaluated as the surface-area weighted diameter (ds) was computed
according to Eq.3.6 where ni represents the number of particles with diameter di
(Σini ≥ 200).

∑ nd 3
d̅ s = ∑i i i2 (3.6)
i ni di

119
CHAPTER 3

2.4. Catalyst activity

Catalytic performance tests were carried out in a tubular quartz reactor


(45 cm length and 1 cm diameter). The catalyst, which had a particle size in the
range 250−500 μm and was not diluted, was placed on a fritted quartz plate
located at the end of the reactor. The amount of catalyst used in the experiments
was 0.8 g.

The temperature of the catalyst was measured with a K-type


thermocouple (Thermocoax) placed inside the inner quartz tube. The entire
reactor was placed in a furnace (Lenton) equipped with a temperature-
programed system. Reaction gases were Praxair certified standards of CO2
(99.999% purity), H2 (99.999% purity), and N2 (99.999% purity). The gas flows
were controlled by a set of calibrated mass flow meters (Brooks 5850 E and
5850 S).

The hydrogenation of CO2 was carried out at atmospheric pressure in the


temperature range 150−300 °C. The total flow rate used in the experiments,
which involved a CO2/H2 mixture (CO2/H2 = 1/9), was 100 cm3 min−1. Gas
effluents were monitored with a micro gas chromatograph (Varian CP-4900)
fitted with a PoraPLOT Q column and a molecular sieve column, each of which
was connected to a thermal conductivity detector (TCD). All the catalytic tests
were performed twice, with the relative error being lower than 5%.

3. Results and discussion


3.1. Influence of the calcination temperature
Four different calcination temperatures were tested for the catalysts with
copper loadings of 10 wt %: namely 300, 350, 400, and 600 °C. The XRD
patterns of the catalysts 10CuZnO-300 and 10CuZnO-600 after the calcination
step are shown in Fig. 3.1a. The main peaks correspond to the hexagonal crystal
structure of zinc oxide (JCPDS 80-0075) and the monoclinic system of copper

120
CHAPTER 3

oxide (JCPDS 80-1917). The diffractograms of the catalysts overlap each other,
and they appear to be very similar However, it can be seen from the enlargement
in Fig. 1a that there is a slight displacement to higher 2θ angles (deg) of the ZnO
peaks for the 10CuZnO-600 sample. This indicates that a contraction between
lattices of the ZnO occurs and that a change in the lattice parameters of the ZnO
hexagonal system (JCPDS 80-0075) has taken place. These lattice parameters
were calculated based on Bragg’s Law (Eq. 3.3) and the formula for the
hexagonal crystal structure (Eq. 3.4). The lattice parameters changed from a =
3.249 Å and c = 5.209 Å for the 10CuZnO-300 catalyst to a = 3.245 Å and c
= 5.182 Å for the 10CuZnO-600 catalyst. After the reduction step (Fig. 3.1b),
the copper oxide peaks disappeared and the diffraction peaks of the metallic
copper cubic crystal structure appeared (JCPDS 85-1326).

a) 10CuZnO-300
10CuZnO-600

^ ZnO
34.2 34.5 34.8
º CuO
^ 2(º)

^ ^ ^
Intensity (a.u.)

^ ^
^ ^
º º º º º ^ ^ º^

b) 10CuZnO-300-200
10CuZnO-600-200

^ ZnO
34.2 34.5 34.8
* Cu
^ 2 (º)
(1 1 1)

^
(2 0 0)

^
(2 2 0)

^ ^ ^
^ ^
*
* ^ ^* ^

25 30 35 40 45 50 55 60 65 70 75 80
2 (º)
Figure 3.1. XRD profiles of (a) before reduction and (b) after reduction of the catalyst calcined at
different temperatures.

121
CHAPTER 3

As shown in the enlargement in Fig. 3.1b, after the reduction step a


restructuring of the ZnO crystal structure had occurred and, at the end, both
catalysts had the same lattice parameters, a = 3.249 Å and c = 5.184 Å.
Therefore, at first glance the crystal structures could seem to be the same.
However, the main physical properties of the catalysts (Table 3.3) show that there
are numerous differences between them. The surface area and the total pore
volume are higher for the samples calcined at lower temperatures. The trend in
the copper oxide particle size is the opposite, and this is due to the water
produced during the decomposition of the precursor (copper nitrate), which
accelerates the growth of CuO crystallites [1]. During the reduction, sintering of
the Cu0 crystals occurs due to an increase in the local temperature of the CuO
crystallites due to the exothermic reaction between hydrogen and CuO [7]. On
the other hand, the decrease in the concomitant surface area favours the metal
sintering. Therefore, as shown by the results in Table 3.3, large Cu0 particle sizes
were also observed for the higher calcination temperature.

Table 3.3. Physical properties of catalyst with different calcination temperature.

10CuZnO- 10CuZnO- 10CuZnO- 10CuZnO-


Sample 300-200 350-200 400-200 600-200

Surface area (m2g-1) 6.0 6.0 5.7 4.6

Total pore volume (cm3g-1)


3.0 3.0 2.7 2.4
102

CuO particle size from XRD


45 49 53 57
(nm)

Cu0 particle size from XRD


69 72 81 136
(nm)

Cu0 particle size from TEM


- 134 - 199
(nm)

I(111)/I(200) 2.5 2.7 2.3 2.2

Temperature peaks in TPR


172/193 172/194 173/195 175/195
experiment (ºC)

122
CHAPTER 3

This situation is consistent with the TEM analysis carried out on the
10CuZnO-350-200 (Fig. 3.2a) and 10CuZnO-600-200 (Fig. 3.2c) samples. The
particle size distribution for both samples is represented in Figures 3.2b and 3.2d,
respectively, where more than 200 particles were measured. Both catalysts
showed a Gaussian particle distribution. The particle distribution shows that the
10CuZnO-600-200 has a higher proportion of larger particles. Finally, the
intensity ratio I(111)/I(200) was also calculated (Table 3.3). It has been
suggested [1, 7] that the morphology of copper crystallites depends on the
reduction method, and this in turn affects the selectivity for methanol formation. It
was pointed out [7] that the methanol selectivity increases with the I(111)/I(200)
ratio. The highest ratio was found for the sample 10CuZnO-350-200.

TPR profiles are shown in Fig. 3.3. The reduction of CuO to metallic Cu
takes place from 130 to 220 °C in two steps, as is shown by α and β. The first
peak (α) is related to the reduction of Cu2+ to Cu+, and the second (β) to the
reduction of Cu+ to Cu0 [47]. For all samples, these processes overlapped and the
second peak appeared as a shoulder on the first one. This overlap was more
pronounced at higher calcination temperatures. This finding indicates that a higher
calcination temperature makes it more difficult to reduce the catalyst. This
behaviour is due to the larger particles found at higher calcination temperatures,
which are more difficult to reduce because hydrogen cannot access the bulk easily
and mainly interacts with the surface (following a classical unreacted core model)
[48]. The slower reduction kinetic for the larger particles explains the overlap of
the peaks.

The catalytic results for these samples are represented in Fig. 3.4. The
10CuZnO-600-200 catalyst showed the lowest CO2 conversion. 10CuZnO-350-
200 and 10CuZnO-300-200 showed similar behaviour, with higher carbon
dioxide conversion than the sample calcined at 600 °C. Regarding the activity
and selectivity toward methanol, these results are related to the intensity ratio

123
CHAPTER 3

I(111)/I(200), which in turn is linked to the calcination temperature, as commented


above. Thus, the highest methanol formation rates were found for the samples
10CuZnO-350-200 and 10CuZnO-300-200 at temperatures of 225 and 240
°C, respectively. For reaction temperatures from 150 to 225 °C, the former
sample showed a higher selectivity to methanol. In contrast, lower reaction
temperatures minimize CO formation. Therefore, the calcination temperature
selected as the most appropriate was 350 °C. It should be noted that, even
though the catalytic experiments were carried out under atmospheric pressure,
the results are far away from the thermodynamic equilibrium values, showing that
there are not thermodynamic limitations (see Table 3.4). The thermodynamic
equilibrium values were calculated using a flowsheet simulator (Aspen HYSYS
V8.4 licensed by Aspen Technology, Inc.). Peng−Robinson was used as the
equation of state, and the reactor modeling was based on a Gibbs reactor. The
conditions used for the simulation (flow rate, CO2/H2 ratio) were the same as in
the experimental reactor.
Table 3.4. Comparison between thermodynamic equilibrium and experimental values for the
10CuZnO-350-200 catalyst.

Thermodynamic Experimental
Temperature (ºC)
(µmol/min) (µmol/min)

150 22.24 0.09


165 17.29 0.20
180 13.61 0.37
195 10.85 0.54
210 8.76 0.70
225 7.14 0.75
240 5.88 0.67
255 4.89 0.52
270 4.10 0.40
285 3.47 0.31
300 2.95 0.22

124
CHAPTER 3

25

a) b)
20

Frequency (%)
15

10

0
25-50 50-75 75-100 100-125 125-150 150-175 175-200

Particle size (nm)

c) d) 20

Frequency (%)
15

10

0
75-100 125-150 175-200 225-250 275-300

Particle size (nm)


45

e) f) 40

35
Frecuency (%)

30

25

20

15

10

0
4-6 6-8 8-10 >10

Particle size (nm)


70

g) h) 60

50
Frequency (%)

40

30

20

10

0
5-25 25-45 45-65 65-85 85-105

Particle size (nm)

Figure 3.2. TEM images and metal particle distribution of (a, b) 10CuZnO-350-200 (c, d) 10CuZnO-
600-200 (e, f) 10CuZnO-350-150 (g, h) 10CuZnO-350-400.

125
CHAPTER 3



10CuZnO-300


10CuZnO-350

 
10CuZnO-400


10CuZnO-600

100 200 300 400 500 600 700


Temperature (ºC)
Figure 3.3. TPR profiles. Comparison between the catalysts calcined at different temperatures.

3.2. Influence of the reduction temperature

Five different reduction temperatures (150, 200, 250, 300, and 400 °C)
were tested for the catalyst 10CuZnO calcined at the best temperature (350 °C).

An enlargement of the XRD pattern is shown in Fig. 3.5 in order to focus


attention on the changes produced by the reduction step. The diffractograms of
the catalyst before reduction (10CuZnO-350) and after reduction at 150 °C are
shown in Fig. 3.5a. These diffractograms are fairly similar. The main peaks
correspond to zinc oxide (JCPDS 80-0075) and copper oxide (JCPDS 80-1917).
However, the 10CuZnO-350 sample showed a small peak at around 43.3°, which
corresponds to metallic copper (JCPDS 85-1326). Thus, the reduction carried out
at 150 °C was not sufficient to reduce completely the CuO to Cu0 and, as a result,
only a few particles of metallic copper were formed.

126
CHAPTER 3

50
a) 10CuZnO-300-200
40 10CuZnO-350-200

CO formation rate
(mol min g )
-1 -1
10CuZnO-400-200
30 10CuZnO-600-200

20

10

CH3OH formation rate


1.2
b)
(mol min g ) 1.0
-1

0.8
-1

0.6
0.4
0.2

c) 10CuZnO-300-200
CH3OH selectivity ()

60
10CuZnO-350-200
50 10CuZnO-400-200
40 10CuZnO-600-200
30
20
10

d) 10CuZnO-300-200
CO2 conversion ()

8 10CuZnO-350-200
10CuZnO-400-200
6 10CuZnO-600-200
4

0
150 175 200 225 250 275 300
Temperature (ºC)

Figure 3.4. Catalytic activity for the catalysts calcined at different temperatures. Formation rates of (a)
CO and (b) CH3OH. (c) CH3OH selectivity and (d) CO2 conversion. Reaction conditions: CO2/H2 = 1/9
and W/F = 0.008 g min cm–3.
127
CHAPTER 3

a) 10CuZnO-350
10CuZnO-350-150

Intensity (a.u.)

(1 1 1)
38 40 42 44 46
2 (º)

b) 10CuZnO-350-200
10CuZnO-350-250
10CuZnO-350-300
Intensity (a.u.)

10CuZnO-350-400

31.6 31.8 32.0 32.2 32.4


2 (º)

c) 10CuZnO-350-200
10CuZnO-350-250
10CuZnO-350-300
Intensity (a.u.)

10CuZnO-350-400
(1 1 1)

43.0 43.2 43.4 43.6 43.8 44.0


2(º)
Figure 3.5. XRD profiles of (a) The 10CuZnO-350 sample before and after reduction at 150 ºC in an
enlargement to show the CuO (38.8º) and metallic Cu (43.3º) peaks; (b) the rest of the samples
reduced at different temperatures in an enlargement of the ZnO peak (around 31.8º) and (c) in an
enlargement of the metallic Cu (around 43.3º).

128
CHAPTER 3

The TEM image (Fig. 3.2e) also shows this behavior; particles seem not to
be completely formed and a high amount of small particles were found, as shown
by the particle distribution (Fig. 3.2f). The XRD diffractograms for the other
samples showed peaks corresponding to zinc oxide (JCPDS 80-0075) and
metallic copper (JCPDS 85-1326). As for the zinc oxide peaks observed during
the calcination step, a displacement to higher 2θ angles (deg) of the ZnO (Fig.
3.5b) and Cu0 peaks (Fig. 3.5c) was found when the reduction temperature was
increased. The 10CuZnO- 350-200 and 10CuZnO-350-250 samples were
similar, whereas the 10CuZnO-350-400 catalyst showed the most marked
change. For this latter sample, taking into account the high displacement
observed, it is reasonable to consider that a new structure had been formed. As
explained in the previous section, this displacement indicates that a contraction
between lattices of the ZnO and Cu occurs, which in turn means that both lattice
parameters have changed. These lattice parameters (Table 3.5) were calculated
based on Bragg’s Law (Eq. 3.3), the formula for the hexagonal crystal structure
for ZnO (Eq. 3.4), and the formula of the cubic crystal structure for metallic copper
(Eq. 3.5).

The physical properties of the catalysts are provided in Table 3.5. The
copper particle size increased with the reduction temperature, with the exception
of the 10CuZnO-350-400 sample. TEM results (Fig. 3.2g) are in agreement with
XRD analysis. The big particles detected in the 10CuZnO-350-200 sample seem
to have been broken into smaller ones in the 10CuZnO-350-400 sample, as also
confirmed by the particle distribution for this sample (Fig. 3.2h). These data
support the theory that a new structure had been formed. The c/a ratio was
calculated to compare the hexagonal structures formed for the ZnO compound
with the ideal structure: i.e., the hexagonal wurtzite (Wz) structure, where each
anion is surrounded by four cations at the corners of a tetrahedron and vice versa.

129
CHAPTER 3

Table 5. Physical properties of catalyst with different reduction temperature

Lattice parameters
Cu particle
size from ZnO (hexagonal) Cu
Sample I(111)/I(200) (cubic)

XRD TEM a (Å) c (Å) c/a a (Å)


(nm) (nm)

10CuZnO-350-150 26 22 - 3.253 5.210 1.602 3.622

10CuZnO-350-200 72 134 2.7 3.249 5.209 1.603 3.620

10CuZnO-350-250 76 - 2.4 3.249 5.209 1.603 3.614

10CuZnO-350-300 81 - 2.3 3.245 5.231 1.612 3.612

10CuZnO-350-400 49a 54a 2.0 3.209 5.209 1.623 3.593


a Particle size of the CuZn alloy
This ideal structure has a c/a ratio of 1.633 [49]. Therefore, on increasing
the reduction temperature, the hexagonal crystal system became more ideal.
According to different authors [4, 12, 18], the new structure formed when the
catalyst was reduced at 400 °C was a CuZn alloy. In other solid mixtures, such as
Pt-alloys [50, 51], the peaks due to the platinum are shifted to higher 2θ angles
when the other metals in the alloy are incorporated into the platinum crystal
structure. In our case, the displacement was produced in both structures (ZnO and
Cu crystal systems), which means that particles of copper moved into the
hexagonal structure of ZnO and particles of Zn were incorporated into the cubic
structure of the metallic copper. For the 10CuZnO-350-400 sample, preliminary
calculations of the percentage of zinc incorporated into the copper structure were
carried out by applying Vergard’s Law, which has been used for other solid
mixtures such as Pt-metal alloys [50, 51] and PdZn alloys [28]. It should be noted
that, taking into account the limitations of this law [52], the value obtained is only
a rough approximation. According to these calculations, the new CuZn alloy
formed would be Cu95.7Zn4.3.

130
CHAPTER 3

The catalytic results for the samples reduced at different temperatures


are shown in Fig. 3.6. The 10CuZnO-350-150 catalyst was the most active for
methane production (Fig. 3.6b). This result could be due to two different factors:
(i) 10CuZnO-350-150 has the lowest Cu particle size, which increases the carbon
dioxide conversion; and (ii), as shown by the XRD pattern, the reduction carried
out at 150 °C was not sufficient to completely reduce the CuO to Cu. The XRD
pattern after reaction of this sample (not shown) only contained peaks
corresponding to metallic copper (with a copper particle size of 57.84 nm),
whereas those corresponding to CuO had disappeared. This finding indicates that
these particles continued to be reduced during the reaction. Thus, there was a
combination of different active sites (Cu2+, Cu1+ and Cu0), and this led to higher
amounts of methane. It has to be taken into account that the reduction during the
reaction only is produced in the 10CuZnO-350-150 catalyst. The other samples
showed XRD diffractograms before and after reaction (not shown) with no
difference, so that metal reduction during reaction can be discarded. The samples
10CuZnO-350-200 and 10CuZnO-350-250 seem to show similar catalytic
behavior, although the selectivity toward methanol is higher for the former. This
behavior is consistent with the intensity ratio I(111)/I(200) (Table 3.5), for which
this sample has the highest value. The 10CuZnO-350-400 catalyst showed the
lowest CO2 conversion despite having a small particle size. This result is due to
the presence of CuZn alloy active sites. This catalyst showed the highest values
for methanol selectivity above 185 °C (Fig. 3.6d). This result is consistent with the
literature data [18, 39] and suggests that surface alloys may yield structures with
high activity toward methanol.

131
CHAPTER 3

50
a) 10CuZnO-350-150
10CuZnO-350-200
40

CO formation rate
(mol min g )
10CuZnO-350-250

-1 -1
10CuZnO-350-300
30 10CuZnO-350-400

20

10

10CuZnO-350-150
b) 10CuZnO-350-200
30 10CuZnO-350-250
CH4 formation rate
(mol min g )
-1

10CuZnO-350-300
10CuZnO-350-400
-1

20

10

1.2
c)
CH3OH formation rate
(mol min g )

1.0
-1
-1

0.8
0.6
0.4
0.2
d) 10CuZnO-350-150
CH3OH selectivity ()

60 10CuZnO-350-200
50 10CuZnO-350-250
10CuZnO-350-300
40 10CuZnO-350-400
30
20
10

14 e) 10CuZnO-350-150
CO2 conversion ()

10CuZnO-350-200
12 10CuZnO-350-250
10 10CuZnO-350-300
10CuZnO-350-400
8
6
4
2
0
150 175 200 225 250 275 300
Temperature (ºC)

Figure 3.6. Catalytic activity for the catalysts reduced at different temperatures. Formation rates of (a)
CO, (b) CH4 and (c) CH3OH. (d) CH3OH selectivity and (e) CO2 conversion. Reaction conditions:
CO2/H2 = 1/9 and W/F = 0.008 g min cm–3.
132
CHAPTER 3

It is remarkable that the lowest intensity ratio I(111)/I(200) was


determined for this catalyst, so that for this sample there is no link between
methanol selectivity and the intensity ratio. The low intensity ratio is clearly not a
drawback for performance. This parameter is mentioned in the literature [1, 7]
when the catalyst has been reduced at lower temperature and, therefore, copper
does not form an alloy with zinc. If the alloy has been formed, the relationship
between the intensity ratio I(111)/I(200) and the methanol selectivity cannot be
applied. The lower CO2 conversion and higher CH3OH selectivity for the
10CuZnO- 350-400 catalyst appear to indicate a different mechanism over
CuZn active sites than over the Cu surface is produced. To demonstrate this
assumption, a further study using DFT comparison or IR analysis in situ would be
required.

Finally, the temperature at which the highest methanol formation rate is


obtained was shifted to higher values as the reduction temperature increased
(Fig. 3.6c). The best reduction temperature was 200 °C. On applying this
reduction temperature, the maximum methanol yield was obtained at lower
reaction temperatures, where carbon monoxide formation does not take place.

3.3. Influence of the metal loading

The XRD patterns for the four samples with different Cu loadings (5, 10,
20, and 40 wt %) are shown in Fig. 3.7. The intensity of the copper peaks
increased with the metal loading, as expected, but other differences were not
observed. The main physical properties for these catalysts are listed in Table 3.6.
The surface area and the total pore volume decreased with the Cu loading. This
decrease occurs for two reasons: (i) the support is degraded by the metal
precursor, which can dissolve the ZnO (due to the acidity of the aqueous copper
nitrate solution [53]), and this degradation is more intense with higher metal
loadings; and (ii) when the metal content increased, more particles can block the
support and reduce the surface area. The particle size values are quite similar,

133
CHAPTER 3

which means that an increase in the metallic copper content does not affect the
size of the particles formed.

40CuZnO-350-200 ^ ZnO

(1 1 1)
^ * Cu
*

(2 0 0)

(2 2 0)
^ ^ ^ ^ ^
^
* * ^
^ ^ ^ ^ ^
^
(1 1 1) 20CuZnO-350-200

^
(2 0 0)
Intensity (a.u.)

^ ^

(2 2 0)
^ ^
* ^ ^
* ^ ^* ^ ^
^ 10CuZnO-350-200

^
(1 1 1)

^ ^

(2 2 0)
(2 0 0)

^ ^
^ ^
* ^
* ^* ^ ^
^ 5CuZnO-350-200

^ ^
(1 1 1)

^
(2 0 0)

^ ^
(2 2 0)

^ ^
* * ^ ^* ^ ^
25 30 35 40 45 50 55 60 65 70 75 80 85
2(º)

Figure 3.7. XRD profiles for the samples with different metal loadings after reduction.

134
Table 3.6. Physical properties of catalyst with different metal loading.

Sample ZnO support 5CuZnO-350-200 10CuZnO-350-200 20CuZnO-350-200 40CuZnO-350-200

Surface area (m2 g-1) 7.2 6.3 6.0 3.1 2.5

Total pore volume (cm3 g-1) 102 3.0 3.0 3.0 2.3 1.9

135
Cu0 particle size from XRD (nm) - 63 72 69 71

I (111) / I (200) - 2.1 2.7 2.3 2.4

Temperature peaks in TPR experiment - 168/187 173/196 173/196/213 167/196/213


(ºC)
CHAPTER 3
CHAPTER 3

The TPR profiles are shown in Fig. 3.8. TPR measurements have been
performed with a weight of sample depending on the metal loading, so that the
metal content in every measurement was always the same. The 5CuZnO-350-200
sample showed a predominant peak of reduction (called α in the figure) at 168
°C. This indicates the faster kinetic of reduction of copper particles from Cu2+ to
Cu0 when the particles are better dispersed (as shown by the XRD particle size).
At higher metal loading a third peak appears (γ) at higher temperatures (around
213 °C). It can be related to the more difficult reduction of a bulk of CuO
particles [11], which are more connected between them. It would also be in
agreement with the surface area decrease.



5CuZnO-350

 
Intensity (a.u.)

10CuZnO-350



20CuZnO-350

 
40CuZnO-350

100 200 300 400 500 600 700


Temperature (ºC)

Figure 3.8. TPR profiles. Comparison between the catalysts with different metal loadings.

According to the catalytic results (Fig. 3.9), a higher metal loading gives
rise to a higher methanol formation rate and higher CO2 conversion. Due to the
different numbers of active sites, the turnover frequency (TOF) was calculated to
normalize the results (Fig. 3.9b).

136
CHAPTER 3

5
a) 5CuZnO-350-200

CH3OH formation rate


10CuZnO-350-200
4 20CuZnO-350-200

(mol min-1 g-1)


40CuZnO-350-200
3

2
5CuZnO-350-200

TOF CH3OH (min ) 10


6 b) 10CuZnO-350-200
-1 20CuZnO-350-200
5 40CuZnO-350-200
4
3
2
1

c)
CH3OH selectivity ()

5CuZnO-350-200
100 10CuZnO-350-200
20CuZnO-350-200
80 40CuZnO-350-200

60
40
20

18 d) 5CuZnO-350-200
CO2 conversion ()

10CuZnO-350-200
15 20CuZnO-350-200
40CuZnO-350-200
12
9
6
3
0
150 175 200 225 250 275 300
Temperature (ºC)

Figure 3.9. Catalytic activity for the catalysts with different metal loadings. (a) Formation rate of
CH3OH (b) TOF CH3OH (C) CH3OH selectivity and (d) CO2 conversion. Reaction conditions: CO2/H2 =
1/9 and W/F = 0.008 g min cm–3.

137
CHAPTER 3

TOF results of the samples with 5, 10, and 20 wt % of copper, show that
an increase in the metal loading favors the synthesis of methanol and the
selectivity toward this compound. This could be explained following the theory of
the dual site mechanism for the methanol production [33, 43], where the
dissociative adsorption of the hydrogen molecule is produced on the ZnO sites
while the carbon dioxide is adsorbed on copper active sites. Therefore, the
mechanism is favored when there are more active sites together. Hence, an
increase in the metal loading favors the mechanism of the methanol synthesis
reaction. However, if the increment in the metal loading is very high, as happened
for the 40CuZnO-350-200 sample, doubling the amount of copper does not
duplicate the activity toward methanol. Thus, an optimum in the metal loading can
be reached, as also can be seen in the selectivity (Fig. 3.9c). The 20CuZnO-350-
200 sample showed the highest selectivity to methanol, although it does not have
the highest I(111)/I(200) ratio. Due to the different amounts of active sites, this
parameter does not have influence in this study compared with the influence which
is provided by the mechanism.

4. Conclusions

The following conclusions can be drawn from this study:

- A high activity and selectivity toward methanol at low temperatures (<225 °C)
was obtained when the catalyst was calcined at 350 °C. The methanol selectivity
increased with the I(111)/I(200) ratio.

- The best reduction temperature was 200 °C. The CuZnO-350-200 catalyst gave
the highest methanol yield at low temperatures, where CO formation is still
negligible (<225 °C). A reduction temperature of 150 °C was not sufficient to
reduce the catalyst completely.

- Methanol was demonstrated to be obtained on Cu0 active sites. In addition, the


combination of Cu2+, Cu1+, and Cu0 sites, found for the 10CuZnO-350-150
sample, led to high quantities of methane as product. At a reduction temperature
138
CHAPTER 3

of 400 °C, CuZn particle alloys were formed. These alloy particles led to high
selectivity at higher temperatures (>200 °C). A different mechanism over CuZn
active sites than over the Cu surface could be produced. However, to demonstrate
this assumption, a further study using DFT comparison or IR analysis in situ would
be required.

- The metal loading study is according to the theory of the dual-site nature of the
main reaction path, confirming the role of the ZnO and metallic copper in the
hydrogenation of CO2 to methanol in Cu/ZnO catalyst.

References
[1] Fujita, S. I.; Moribe, S.; Kanamori, Y.; Kakudate, M.; Takezawa, N. Preparation
of a coprecipitated Cu/ZnO catalyst for the methanol synthesis from CO2 - effects
of the calcination and reduction conditions on the catalytic performance. Applied
Catalysis A, 207 (2001) 121-128.

[2] Fujita, S. I.; Kanamori, Y.; Satriyo, A. M.; Takezawa, N. Methanol synthesis
from CO2 over Cu/ZnO catalysts prepared from various coprecipitated
precursors. Catalysis Today, 45 (1998) 241-244.

[3] Fujita, S. I.; Usui, M.; Ohara, E.; Takezawa, N. Methanol synthesis from carbon
dioxide at atmospheric pressure over Cu/ZnO catalyst. Role of methoxide species
formed on ZnO support. Catalysis Letters, 13 (1992) 349-358.

[4] Choi, Y.; Futagami, K.; Fujitani, T.; Nakamura, J. The difference in the active
sites for CO2 and CO hydrogenations on Cu/ZnO-based methanol synthesis
catalysts. Catalysis Letters, 73 (2001) 27-31.

[5] Karelovic, A.; Ruiz, P. The role of copper particle size in low pressure
methanol synthesis via CO2 hydrogenation over Cu/ZnO catalysts. Catalysis
Science and Technology, 5 (2015) 869-881.

139
CHAPTER 3

[6] Fujitani, T.; Saito, M.; Kanai, Y.; Kakumoto, T.; Watanabe, T.; Nakamura, J.;
Uchijima, T. The role of metal oxides in promoting a copper catalyst for methanol
synthesis. Catalysis Letters, 25 (1994) 271-276.

[7] Fujita, S. I.; Moribe, S.; Kanamori, Y.; Takezawa, N. Effects of the calcination
and reduction conditions on a Cu/ZnO methanol synthesis catalyst. Reaction
Kinetics and Catalysis Letters, 70 (2000) 11-16.

[8] Joo, O. S.; Jung, K. D.; Han, S. H.; Uhm, S. J.; Lee, D. K.; Ihm, S. K. Migration
and reduction of formate to form methanol on Cu/ ZnO catalysts. Applied
Catalysis-A, 135 (1996) 273-286.

[9] Shen, G. C.; Fujita, S. I.; Takezawa, N. Preparation of precursors for the
Cu/ZnO methanol synthesis catalysts by coprecipitation methods: effects of the
preparation conditions upon the structures of the precursors. Journal of Catalysis,
138 (1992) 754-758.

[10] Himelfarb, P. B.; Simmons, G. W.; Klier, K.; Herman, R. G. Precursors of the
copper-zinc oxide methanol synthesis catalysts. Journal of Catalysis 93 (1985)
442-450.

[11] Lei, H.; Nie, R.; Wu, G.; Hou, Z. Hydrogenation of CO2 to CH3OH over
Cu/ZnO Catalysts with different ZnO morphology. Fuel 154 (2015) 161-166.

[12] Nakamura, J.; Choi, Y.; Fujitani, T. On the issue of the active site and the role
of ZnO in Cu/ZnO methanol synthesis catalysts. Topics in Catalysis 22 (2003)
277-285.

[13] Choi, Y.; Futagami, K.; Fujitani, T.; Nakamura, J. Role of ZnO in Cu/ZnO
methanol synthesis catalysts - morphology effect or active site model? Applied
Catalysis-A, 208 (2001) 163-167.

140
CHAPTER 3

[14] Porosoff, M. D.; Yan, B.; Chen, J. G. Catalytic reduction of CO2 by H2 for
synthesis of CO, methanol and hydrocarbons: challenges and opportunities.
Energy Environmental Science. 9 (2016) 62-73.

[15] Ovesen, C. V.; Clausen, B. S.; Schiøtz, J.; Stoltze, P.; Topsøe, H.; Nørskov, J.
K. Kinetic implications of dynamical changes in catalyst morphology during
methanol synthesis over Cu/ZnO catalysts. Journal of Catalysis, 168 (1997) 133-
142.

[16] Yoshihara, J.; Campbell, C. T. Methanol synthesis and reverse water-gas shift
kinetics over Cu (110) model catalysts: structural sensitivity. Journal of Catalysis,
161 (1996) 776-782.

[17] Rasmussen, P. B.; Kazuta, M.; Chorkendorff, I. Synthesis of methanol from a


mixture of H2 and CO2 on Cu (100). Surface Science 318 (1994) 267-280.

[18] Topsøe, N. Y.; Topsøe, H. FTIR Studies of dynamic surface structural changes
in Cu-based methanol synthesis catalysts. Journal of Molecular Catalysis A:
Chemical. 141 (1999) 95-105.

[19] Hadden, R. A.; Sakakini, B.; Tabatabaei, J.; Waugh, K. C. Adsorption and
reaction induced morphological changes of the copper surface of a methanol
synthesis catalyst. Catalysis Letters, 44 (1997) 145-151.

[20] Fujita, S.; Usui, M.; Ito, H.; Takezawa, N. Mechanisms of methanol synthesis
from carbon dioxide and from carbon monoxide at atmospheric pressure over
Cu/ZnO. Journal of Catalysis, 157 (1995) 403-413.

[21] Nakamura, I.; Fujitani, T.; Uchijima, T.; Nakamura, J. The synthesis of methanol
and the reverse water-gas shift reaction over Zn-deposited Cu(100) and Cu(110)
surfaces: comparison with Zn/ Cu(111). Surface Science, 400 (1998) 387.

141
CHAPTER 3

[22] Fujitani, T.; Nakamura, I.; Uchijima, T.; Nakamura, J. The kinetics and
mechanism of methanol synthesis by hydrogenation of CO2 over a Zn-deposited
Cu(111) surface. Surface Science, 383 (1997) 285-298.

[23] Spencer, M. S. Role of ZnO in methanol synthesis on copper catalysts.


Catalysis Letters, 50 (1998) 37-40.

[24] Nakamura, J.; Uchijima, T.; Kanai, Y.; Fujitani, T. The role of ZnO in Cu/ZnO
methanol synthesis catalysts. Catalysis Today, 28 (1996) 223-230.

[25] Quadrelli, E. A.; Centi, G.; Duplan, J. L.; Perathoner, S. Carbon dioxide
recycling: emerging large-scale technologies with industrial potential.
ChemSusChem, 4 (2011) 1194-1215.

[26] Aresta, M.; Dibenedetto, A.; Quaranta, E. State of the art and perspectives
in catalytic processes for CO2 conversion into chemicals and fuels: the distinctive
contribution of chemical catalysis and biotechnology. Journal of Catalysis, 343
(2016) 2-45.

[27] Díez-Ramírez, J.; Valverde, J. L.; Sánchez, P.; Dorado, F. CO2 Hydrogenation
to methanol at atmospheric pressure: influence of the preparation method of
Pd/ZnO catalysts. Catalysis Letters, 146 (2016) 373-382.

[28] Díez-Ramírez, J.; Sánchez, P.; Rodríguez-Gómez, A.; Valverde, J. L.; Dorado,
F. Carbon nanofiber-based palladium/zinc catalysts for the hydrogenation of
carbon dioxide to methanol at atmospheric pressure. Industrial and Engineering
Chemistry Research, 55 (2016) 3556-3567.

[29] Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M.; Joshi, J. B. Catalytic carbon
dioxide hydrogenation to methanol: a review of recent studies. Chemical
Engineering Research and Design, 92 (2014) 2557-2567.

142
CHAPTER 3

[30] Bahmani, M.; Vasheghani Farahani, B.; Sahebdelfar, S. Preparation of high


performance nano-sized Cu/ZnO/Al2O3 methanol synthesis catalyst via aluminum
hydrous oxide sol. Applied Catalysis-A, 520 (2016) 178-187.

[31] Behrens, M.; Kißner, S.; Girsgdies, F.; Kasatkin, I.; Hermerschmidt, F.; Mette,
K.; Ruland, H.; Muhler, M.; Schlögl, R. Knowledge-based development of a
nitrate-free synthesis route for Cu/ZnO methanol synthesis catalysts via formate
precursors. Chemical Communications, 47 (2011) 1701-1703.

[32] Ladera, R.; Pérez-Alonso, F. J.; González-Carballo, J. M.; Ojeda, M.; Rojas,
S.; Fierro, J. L. G. Catalytic valorization of CO2 via methanol synthesis with Ga-
Promoted Cu-ZnO-ZrO2 catalysts. Applied Catalysis B: Environmental, 142-143
(2013) 241-248.

[33] Bonura, G.; Arena, F.; Mezzatesta, G.; Cannilla, C.; Spadaro, L.; Frusteri, F.
Role of the ceria promoter and carrier on the functionality of Cu-based catalysts
in the CO2-to-methanol hydrogenation reaction. Catalysis Today, 171 (2011)
251-256.

[34] Arena, F.; Italiano, G.; Barbera, K.; Bonura, G.; Spadaro, L.; Frusteri, F. Basic
evidences for methanol-synthesis catalyst design. Catalysis Today, 143 (2009)
143, 80-85.

[35] Bonura, G.; Cordaro, M.; Cannilla, C.; Arena, F.; Frusteri, F. The changing
nature of the active site of Cu-Zn-Zr catalysts for the CO2 hydrogenation reaction
to methanol. Applied Catalysis-B: Environmental, 152-153, (2014) 152-161.

[36] Wan, Y.; Zhou, Z.; Cheng, Z. Hydrogen production from steam reforming of
methanol over CuO/ZnO/Al2O3 catalysts: catalytic performance and kinetic
modeling. Chinese Journal of Chemical Engineering, 24 (2016) 1186-1194.

[37] Tong, W.; West, A.; Cheung, K.; Yu, K. M.; Tsang, S. C. E. Dramatic effects
of gallium promotion on methanol steam reforming Cu−ZnO catalyst for hydrogen

143
CHAPTER 3

production: formation of 5 Å copper clusters from Cu−ZnGaOx. ACS Catalysis, 3


(2013) 1231-1244.

[38] Lorenzut, B.; Montini, T.; De Rogatis, L.; Canton, P.; Benedetti, A.; Fornasiero,
P. Hydrogen production through alcohol steam reforming on Cu/ZnO-based
catalysts. Applied Catalysis-B, 101 (2011) 397-408.

[39] Batyrev, E. D.; Shiju, N. R.; Rothenberg, G. Exploring the activated state of
Cu/ZnO(0001)-Zn, a model catalyst for methanol synthesis. Journal Physical
Chemistry C, 116 (2012) 19335-19341.

[40] Kandemir, T.; Kasatkin, I.; Girgsdies, F.; Zander, S.; Kühl, S.; Tovar, M.;
Schlögl, R.; Behrens, M. Microstructural and defect analysis of metal nanoparticles
in functional catalysts by diffraction and electron microscopy: The Cu/ZnO
catalyst for methanol synthesis. Topics in Catalysis, 57 (2014) 188-206.

[41] Álvarez Galván, C.; Schumann, J.; Behrens, M.; Fierro, J. L. G.; Schlögl, R.;
Frei, E. Reverse water-gas shift reaction at the Cu/ZnO interface: influence of the
Cu/Zn ratio on structure-activity correlations. Applied Catalysis-B, 195 (2016)
104-111.

[42] Kunkes, E. L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M.
Hydrogenation of CO2 to Methanol and CO on Cu/ZnO/Al2O3: Is There a
Common Intermediate or Not? Journal of Catalysis, 328 (2015) 43-48.

[43] Lim, H. W.; Park, M. J.; Kang, S. H.; Chae, H. J.; Bae, J. W.; Jun, K. W.
Modeling of the kinetics for methanol synthesis using Cu/ ZnO/Al2O3/ZrO2
catalyst: influence of carbon dioxide during hydrogenation. Industrial and
Engineering Chemistry Research, 48 (2009) 10448-10455.

[44] Vanden Bussche, K. M.; Froment, G. F. A Steady-State kinetic model for


methanol synthesis and the water gas shift reaction on a commercial
Cu/ZnO/Al2O3 catalyst. Journal of Catalysis, 161 (1996) 1-10.
144
CHAPTER 3

[45] Graaf, G. H.; Stamhuis, E. J.; Beenackers, A. A. C. M. Kinetics of low-pressure


methanol synthesis. Chemical Engineering Science, 43 (1988) 3185-3195.

[46] Fichtl, M. B.; Schlereth, D.; Jacobsen, N.; Kasatkin, I.; Schumann, J.; Behrens,
M.; Schlögl, R.; Hinrichsen, O. Kinetics of deactivation on Cu/ZnO/Al2O3 methanol
synthesis catalysts. Applied Catalysis-A, 502 (2015) 262-270.

[47] Kargol, M.; Zajac, J.; Jones, D. J.; Rozière, J.; Steriotis, T.; Jiménez-López, A.;
Rodríguez-Castellón, E. Copper- and silver-containing monolithic silica-supported
preparations for selective propene-propane adsorption from the gas phase.
Chemistry of Materials, 17 (2005) 6117-6127.

[48] Gutiérrez-Guerra, N.; Moreno-López, L.; Serrano-Ruiz, J. C.; Valverde, J. L.;


de Lucas-Consuegra, A. Gas Phase Electrocatalytic Conversion of CO2 to syn-
fuels on Cu based catalysts-electrodes. Applied Catalysis-B, 188 (2016) 272-
282.

[49] Morkoç, H.; Özgür, Ü. Zinc oxide: fundamentals, materials and device
technology; Wiley-VCH, (2009) 1-477

[50] Hyun, K.; Lee, J. H.; Yoon, C. W.; Kwon, Y. The effect of platinum based
bimetallic electrocatalysts on oxygen reduction reaction of proton exchange
membrane fuel cells. International Journal of Electrochemical Science, 8 (2013)
11752-11767.

[51] Antolini, E.; Cardellini, F. Formation of carbon supported PtRu alloys: an XRD
analysis. Journal of Alloys and Compounds, 315 (2001) 118-122.

[52] Jacob, K. T.; Raj, S.; Rannesh, L. Vegard’s law: A fundamental relation or an
approximation? International Journal of Materials Research, 98 (2007) 776-779.

[53] Chin, Y. H.; Wang, Y.; Dagle, R. A.; Li, X. S. Methanol steam reforming over
Pd/ZnO: catalyst preparation and pretreatment studies. Fuel Processing
Technology, 83 (2003) 193-201.
145
CHAPTER 4
EFFECT OF SUPPORT NATURE ON
COBALT CATALYSTS
CHAPTER 4

Abstract
CO2 hydrogenation to value added chemicals/fuels has gained
considerable interest, in terms of sustainable energy and environmental
mitigation. In this regard, the present work aims to investigate the CO2
methanation performance of cobalt-based catalysts supported on different metal
oxides (MxOy: CeO2, ZrO2, Gd2O3, ZnO) at low temperatures (200–300 °C) and
under atmospheric pressure. The results revealed a significant impact of support
nature on the CO2 hydrogenation performance. The following order, in terms of
CH4 yield (YCH4), was recorded at 300 °C: Co/CeO2 (∼96%) > Co/ZnO
(∼54%) > Co/G2O3 (∼53%) ∼Co/ZrO2 (∼53%). On the basis of the
characterization results, the superiority of Co/CeO2 catalyst can be mainly
ascribed to its enhanced reducibility linked to Co-Ceria interactions. Moreover,
Co/CeO2 demonstrated a stable conversion/selectivity performance under
subsequent reaction cycles, in contrast to Co/ZnO, which progressively activated
under reaction conditions. The latter is related with the modifications induced in
elemental chemical states and surface composition of Co/ZnO upon pretreatment
in reaction conditions, in contrast to Co/CeO2 sample where a stable surface
performance was observed.

100
Co/ZnO2
CeO2
CO2 Conversion (%)

ZnO Co/CeO2
80
Co/ZrO2
60 Co/Gd2O3

40 ZrO2
Gd2O3
20

0
210 225 240 255 270 285 300
Temperature (ºC)

149
CHAPTER 4

1. Introduction
During the past two centuries, the world’s energy demands have
considerably increased due to the growing population and industrialization [1,
2]. Nowadays, almost 80% of these needs are met by carbon-containing fossil
fuels, such as oil, natural gas and coal leading to increasing CO2 emissions [1–3].
The dependence on fossil fuels is not expected to decline sufficiently in the
forthcoming years, while the energy demands are predicted to grow by ca. 35%
[1]. On the other hand, the intensive use of fossil fuels is considered responsible
for the increase of CO2 atmospheric concentration by more than 30% in the last
50 years [4]. Carbon dioxide constitutes the major greenhouse gas, contributing
to the Earth’s global warming effect and to the acidification of the oceans [4].

To efficiently manage the risks associated with global warming, the


anthropogenic CO2 emissions should be limited in order to meet the challenging
targets set by the global Paris Agreement and the EU 2030 Framework for
climate and energy. Three main routes could be followed toward addressing this
issue: i) CO2 emissions reduction [5], ii) Carbon Capture and Sequestration (CCS)
[3, 5-8] and iii) CO2 transformation to value added chemicals or fuels [3, 6]. The
first approach requires the development of more efficient energy conversion fossil
fuel technologies and/or the switching to renewable energy sources (RES). On the
other hand, CCS faces some major issues related mainly to the long-term stability
of storage locations [3, 5-8]. The chemical transformation of CO2 to useful
products is among the most viable strategies, since not only could sustain the
atmospheric concentration of CO2 at acceptable levels but also provides value
added chemicals/fuels, such as methane and methanol [6, 9, 10].

However, the CO2 molecule is very stable, requiring an energy


demanding, catalyst-aided, process for its activation [11, 12]. The most common
and sustainable way to activate CO2 involves its interaction with hydrogen being
generated from RES-powered water electrolysis. This particular process could

150
CHAPTER 4

also serve as a promising approach to efficiently store the excess electricity from
intermittent renewable sources (solar, wind), which in turn can be transformed via
CO2 hydrogenation to fuels, such as methane or methanol [3, 11, 13-15]. In this
regard, the CO2 hydrogenation constitutes one of the most challenging processes
in the field of catalysis [3, 11, 13-15]. Depending on the catalyst used, a variety
of products can be derived, such as methanol [13], methane (Sabatier reaction)
[14], higher hydrocarbons [15] and CO [11], according to the following reactions:

CO2(g) + H2(g) ↔ CO(g) + H2O(g) ΔH298 = 41.2 kJ/mol (4.1)

CO2(g) + 3H2(g) ↔ CH3OH(g) + H2O(g) ΔΗ298 = - 49.6 kJ/mol (4.2)


CO2(g) + 4H2(g) ↔ CH4(g) + 2H2O(g) ΔΗ298 = - 165.0 kJ/mol (4.3)

xCO2(g) + yH2(g) ↔ CxH2y-4x(g) + 2H2O(g) (4.4)

In the last few years, a wide variety of materials have been tested as
catalysts for the above reactions. Among them, Cu, Pd and Zn have been reported
as the most active catalysts for reducing CO2 to methanol and CO (Eq. (4.1),
(4.2)) [3, 13, 16-21], while Ni, Rh and Ru are the most effective for producing
hydrocarbons (Eq. (3), (4)) [22-24]. Nibased catalysts are much cheaper than Ru
and Rh, but they suffer from deactivation, due to the sintering of Ni particles and
carbon poisoning [25, 26].

On the other hand, cobalt-based catalysts are widely used in Fischer-


Tropsch synthesis, since they combine high performance for the hydrogenation of
CO to hydrocarbons with a relatively low cost [3]. However, if CO feed is
switched to CO2, then Co-based catalysts are highly active and selective toward
the formation of methane (Eq. (4.3)) [3, 27-36]. The latter has been ascribed to
the weaker adsorption of CO2 as compared to CO on the Co surface, leading to
lower C/H ratios, which favor the formation of methane instead of higher
hydrocarbons (Eq. (4.4)) [37].

151
CHAPTER 4

The conversion of CO2 and the selectivity to CH4 depends on various


parameters, involving among others, the preparation procedure [29, 38, 39], the
metal loading [30, 40, 41] and the supporting carrier [27, 30, 40]. Especially,
the latter is of crucial importance. It has been well established that the support
nature can notably affect the redox and adsorptive properties of the catalyst,
through strong metal-support interactions, with important effects on the catalytic
performance [14, 27, 32, 42-45]. Suslova et al. [27] showed that apart from the
particle size and oxidation state of Co entities, the catalytic activity and
selectivity of CO2 hydrogenation process can be notably affected by the nature
of the support. The presence or not of a reducible carrier could modify the local
surface structure of Co and consequently, its catalytic behavior [27].

In the light of the above aspects, the present work aims to systematically
explore the impact of support nature on the physicochemical properties and the
CO2 hydrogenation activity/stability of Co-based catalysts, at intermediate
temperatures (200–300 °C) and under atmospheric pressure. Four metal oxides
(CeO2, ZrO2, Gd2O3, ZnO) of different textural/redox characteristics were
employed as supporting materials. The physicochemical properties of the
catalysts were evaluated by means of XRD, BET, TPR and XPS techniques to gain
insight into possible structure-activity relationships.

2. Experimental

2.1. Catalyst preparation

Co-based catalysts were prepared through the wet impregnation method,


employing Co(NO3)2.6H2O (Sigma Aldrich, 99.999% purity) as precursor
compound [46]. For the synthesis of bare oxides (CeO2, ZrO2, Gd2O3 and ZnO),
the appropriate stoichiometric quantities of the corresponding precursor nitrates
(> 99% metal purity, Sigma Aldrich) were diluted in distilled water to prepare
an aqueous solution. The solution was stirred on a hot plate until water
evaporation, followed by drying overnight at 100 °C and calcination at 600 °C
152
CHAPTER 4

for 2 h. The resulting oxides were then added to the aqueous solution of Co
nitrate, at the appropriate concentration of the desired loading (40 wt.% Co).
This specific loading was dictated from preliminary studies (not included)
concerning the impact of metal loading (20–60 wt.%) on the catalytic activity.
The solvent was removed via heating under continuous stirring and the sample
was then dried overnight at 100 °C and calcined at 600 °C for 2 h.

2.2. Catalyst characterization

Cobalt loading was determined by atomic absorption (AA)


spectrophotometry on a SPECTRA 220FS analyser. Samples (ca. 0.5 g) were
treated with 2 mL HCl, 3 mL HF and 2 mL H2O2 followed by microwave digestion
(250 ºC). Surface area/porosity measurements were carried out using a
QUADRASORB 3SI sorptometer apparatus with N2 as the sorbate at -196 ºC.
The samples were outgassed at 250 ºC under vacuum (5 × 10−3 Torr) for 12 h
prior to analysis. Specific surface areas were determined by the multi-point BET
method. The total pore volume was evaluated from N2 uptake at a relative
pressure of P/P0= 0.99. Temperature-programmed reduction (TPR) experiments
were conducted in a commercial Micromeritics AutoChem 2950 HP unit with TCD
detection. Samples (ca. 0.15 g) were loaded into a U-shaped tube and ramped
from room temperature to 700 ºC (10 ºC min−1), using a reducing gas mixture of
17.5% v/v H2/Ar (60 cm3 min−1). The degree of reduction was calculated
according to the following equation:
mmol
Actual H2 amount ( )
g
R(%)= mmol (4.5)
Theoretical H2 amount ( )
g

where the actual amount of H2 is determined by the quantification of TPR


profiles in the low temperature range where the reduction of CoxOy is mainly
taking place, whereas the theoretical H2 corresponds to the stoichiometric amount

153
CHAPTER 4

of H2 required to completely reduce Co3O4 according to (Eq. (4.6)), assuming


negligible reduction of oxide carriers in the low temperature range [41].

Co3 O4 +4H2 →3Co+4H2 O (4.6)

XRD analyses were conducted with a Philips X’Pert instrument using nickel-
filtered Cu-Kα radiation. Samples were scanned at a rate of 0.02° step−1 over
the range of 5° ≤2θ ≤90° (scan time = 2 s step−1). Transmission electron
microscopy (TEM) analyses were conducted in a JEOL JEM-4000EX unit with an
accelerating voltage of 400 kV. Samples were prepared by ultrasonic dispersion
in acetone with a drop of the resulting suspension evaporated onto a holey
carbon-supported grid. TEM images were obtained over used samples, i.e. over
samples subjected to reaction conditions (see below). The particle size was
calculated from TEM images. The mean particle size evaluated as the surface-
area weighted diameter (ds), and was calculated according to Eq. (4.7) where ni
represents the number of particles with diameter di (Σini≥200).

∑ nd 3
d̅ s = ∑i i i2 (4.7)
i ni di

The X-ray photoelectron spectroscopy (XPS) measurements were carried


out in an ultrahigh vacuum (UHV) spectrometer described elsewhere [47]. Spectra
were recorded with a VSW Class WA hemispherical electron analyzer using a
monochromated Al Kα X-ray source (1486.6 eV). Survey and high resolution XP
spectra were recorded in constant pass energy mode (100 and 44 eV,
respectively). An electron flood gun was used to compensate for differential
surface charging of insulating samples, while the C 1s of adventitious carbon at
285 eV was used as a reference of the binding energy scale. The surface atomic
ratio was calculated using the integrated XPS peak areas, normalized to the
atomic sensitivity factors.

154
CHAPTER 4

2.3. Catalytic evaluation studies


Experiments were carried out in a tubular quartz reactor (45 cm length
and 1 cm diameter). The catalyst (0.6 g), with a particle size in the range of 250–
500 μm and without dilution, was placed on a fritted quartz plate located at the
end of the reactor. The temperature of the catalyst was measured with a K-type
thermocouple (Thermocoax) placed inside the catalyst bed. The entire reactor
was placed in a furnace (Lenton) equipped with a temperature-programmed
controller. Reaction gases were Praxair certified standards of CO2 (99.999%
purity), H2 (99.999% purity) and N2 (99.999% purity). The gas flows were
controlled by a set of calibrated mass flowmeters (Brooks 5850 E and 5850 S).

The experimental procedure followed for the activity and stability tests is
described in the following. Prior to the reaction, all samples were reduced in situ
in a hydrogen stream (10% vol) diluted with nitrogen at a flow rate of 25 cm3
min−1. The temperature was increased at a heating rate of 10 °C min−1 up to 450
°C, and kept constant for 2 h. Then the sample was purged in N2 flow and the
temperature decreased to 200 °C. The reaction was then carried out at
atmospheric pressure in the temperature range of 200–300 °C. The feed
contained a CO2/H2 mixture (CO2/H2 molar ratio= 1/9) with PCO2 = 10 kPa,
PH2= 90 kPa and a total flowrate of 75 cm3/min. Gases were monitored by a
micro gas chromatograph (Varian CP-4900). Product’s selectivity (Si) was
calculated as shown in Eq. (4.8) considering the following chemical species: CO,
CH4, C2H6, C3H8 and CH3OH. It should be noted that under the present reaction
conditions, the selectivity to higher hydrocarbons was always lower than 5%.
n ri
Si (%)= ∑ i 100 (4.8)
i ni ri

where ni is the number of carbon molecules and ri the formation rate of


product i, respectively.

155
CHAPTER 4

To gain insight into the intrinsic reactivity of Co/MxOy catalysts, the


turnover frequency (TOFi, min−1) for each product (i) was also calculated
according to the equation:
-1
ri (mol min-1 g-1
Co ) ACo (gmol )
TOFi (min )=
-1
(4.9)
DCo /100

where ri is the formation rate of product i, ACo is the atomic mass of Co


(58.93 gCo mol−1), and DCo is the Co dispersion. The dispersion of cobalt, DCo, was
calculated based on the Co particle size obtained by TEM analysis, according to
the following equation [48]:
6 ACo ns
DCo (%)= 100 (4.10)
ρCo NAv dCo

where ns is the number of Co atoms at the surface per unit area (1.51 x
1019 at m−2), ρCo is the density of cobalt (8900 x 103 g m−3), NAv is the Avogadro’s
number (6.023 x 1023 at mol−1) and dCo is the average Co particle diameter
determined by TEM.

Motivated by the progressive activation under reaction conditions and in


order to examine the stability performance, indicated samples were imposed to
successive heating/cooling cycles during reaction. In particular, after the
completion of the 1st cycle of catalytic evaluation measurements, the sample was
cooled down to room temperature under N2 flow, remaining at this stage for 12
h. Then, the catalytic behavior was again explored (2nd cycle) by increasing the
temperature up to 300 °C. The following nomenclature was used throughout the
text in relation to the state of catalysts: fresh (as prepared samples calcined at
600 °C for 2 h), reduced (fresh catalysts imposed to reduction at 450 °C for 2 h
under 10% v/v H2), used (after the completion of the 2nd cycle of catalytic
activity measurements).

156
CHAPTER 4

3. Results and discussion

3.1. Textural and structural characteristics

Table 4.1 summarizes the main textural characteristics of as prepared


(fresh) Co-based samples, in terms of surface area and total pore volume. It is
evident that the Co/CeO2 sample possesses the highest BET area (29 m2 g−1),
followed by Co/Gd2O3 (19 m2 g−1) > Co/ZrO2 (14 m2 g−1) > Co/ZnO (4.7 m2
g−1). These differences could be realized by taking into account the different
structural characteristics of the supporting carriers, as will be discussed below.
Table 4.1. Textural and structural characteristics of Co/MxOy samples.
Co/Gd2O3 Co/ZrO2 Co/CeO2 Co/ZnO
Cobalt loading (wt.%)a 37.6 41.7 42.3 39.8
Surface area (m2 g -1)b 19 14 29 4.7
Total pore volume (cm3 g-1) 0.08 0.06 0.11 0.04
Crystallite size of metallic
n.d.c - 31.3d 45.9c -34.2d 42.2c -27.4d 47.3c -40.9d
Co after reduction (nm)
Cobalt dispersion (%)e 3.2 2.9 3.6 2.4
a determined by atomic absorption (AA) spectrophotometry; b determined by the multi-point BET
method; c determined by the Scherrer equation (XRD) ; d determined by TEM ; e determined by
Eq. (4.9) using the particle size from TEM
Fig. 4.1a shows the XRD patterns of as prepared (fresh) Co-based
samples. To gain insight into the impact of reduction pretreatment, the
corresponding XRD patterns of reduced samples are comparatively shown in Fig.
4.1b. In Table 4.1, the mean crystallite size of metallic cobalt phase as
determined by the Scherrer equation and TEM images is also presented. Cobalt
particle size and dispersion values are in agreement with BET area results: the
higher the particle size, the lower the surface area.

Regarding the fresh samples, the cobalt oxide (Co3O4, JCPDS 78-1970)
was the predominant phase of cobalt entities detected in Co/CeO2 and Co/ZrO2
samples. This inorganic compound is a mixed oxide with cobalt valences of +2

157
CHAPTER 4

and +3. It generally adopts the normal spinel structure with Co2+ ions in
tetrahedral interstices and Co3+ ions in the octahedral interstices of the cubic
close-packed lattice of oxide anions [49]. However, for the Co/ZnO catalyst, the
binary spinel of Co3O4 was found. It is listed as CoCo2O4 (JCPDS 80-1545), to
distinguish among the double and triple positively charged ions at the tetrahedral
site [50]. The other peaks that appear in the diffractograms were related to the
different supports. Zinc oxide (JCPDS 80-0075) showed a hexagonal system and
cerium oxide a cubic system (JCPDS 81-0792). For zirconium oxide, there were
reflections peaks of two different systems: monoclinic system (JCPDS 07-0343)
along with the tetragonal system (JCPDS 88-1007). Finally, Co/Gd2O3 catalyst
showed an amorphous structure, hindering the precise identification of different
phases, although the presence of Co3O4 may be sensed.

a) * Co3O4 º CoCo2O4 b) + Co ^ CoO


º
Co/ZnO º º
º +
º º º º ºº º ºº º +^ + ^ ^+
Intensity (a.u.)

*
* * +
Co/CeO2 * * * *
+
* * * * * + +
*
Co/ZrO2 * * * +
+
+
* * * * ** * +
*
* *
+
Co/Gd2O3

0 10 20 30 40 50 60 70 80 30 40 50 60 70 80 90
2(º) 2(º)

Figure 4.1. XRD profiles (a) prior and (b) after the reduction for the different supported catalysts.

A comparison between the XRD patterns of fresh (Fig. 4.1a) and reduced
(Fig. 4.1b) samples reveals various new reflections at 41.78°, 44.36°, 47.30°
and 51.28°, which are attributed to the cubic metallic cobalt (JCPDS 01-1259).
The Co/ZnO catalyst showed some peaks of cobalt oxide (CoO) with cubic
structure (JCPDS 78-0431), demonstrating an incomplete reduction of cobalt
oxide under the present reduction conditions. This could be assigned to the
formation of a hardly reducible ZnCo2O4 phase, which it is possible to be formed
158
CHAPTER 4

at a moderate cobalt content [51]. This compound shows a cubic structure (JCPDS
23-1390) with peaks similar to those of CoCo2O4. Therefore, its presence cannot
be discarded. On the other hand, there were no oxide peaks in the Co/CeO2
and Co/ZrO2 samples, suggesting their complete reduction. Finally, as
commented above, Co/Gd2O3 XRD spectrum is unclear and hence it is impossible
to deny or confirm the existence of oxides or other compounds.

TEM images are shown in Fig. 4.2. All catalysts exhibit a Gaussian
distribution of particle size, with the maximum following the order of: Co/ZnO >
Co/ZrO2 > Co/Gd2O3 > Co/CeO2. Thus, the Co/ZnO sample showed the highest
particle size and Co/CeO2 exhibited the lowest, confirming the difference in the
BET area between the samples.

3.2. Redox properties

Fig. 4.3 depicts the TPR profiles of the Co/MxOy samples. In Table 4.2,
the main TPR peaks are summarized. The Co/ZrO2 catalyst exhibited only one
reduction peak at 341 °C, probably due to the superimposition of the successive
reduction processes of Co+3 to Co+2 and Co+2 to Co0. For the Co/ZnO catalyst,
two reduction peaks were observed, located, however, at higher temperatures
as compared to Co/CeO2. These peaks could be ascribed to the stepwise
reduction of cobalt oxide entities with different interaction with the support as
well as to the possible contribution of support reduction. Finally, the Co/Gd2O3
catalyst showed a wide reduction profile consisting of three broad peaks. This
can be attributed to the complex metal-support interactions and support porous
structure, resulting in differently reducible cobalt species [44, 52].

To gain insight into the impact of support nature on the reducibility of Co-
based samples, the amount of H2 consumption (mmol g−1) was also estimated from
the quantification of TPR profiles (Table 4.2). The reduction degree of Co-based
samples, i.e., the ratio of the actual H2 consumption to the theoretical amount of
H2 required for the complete reduction of Co3O4 in each sample, follows the
159
CHAPTER 4

order: Co/CeO2 (91%) > Co/Gd2O3 (87%) > Co/ZrO2 (84%) > Co/ZnO
(79%). The lowest reduction degree of Co/ZnO catalyst is in agreement with the
appearance of cobalt oxide phases in the XRD patterns of reduced samples (Fig.
4.1). As explained above, the unreduced cobalt oxide could be due to a hardly
reducible ZnCo2O4 phase formed during the calcination of the Co/ZnO catalyst.
This phase could be also responsible for the reduction temperature shift to higher
values for this catalyst [51]. This compound can be reduced in two steps, i.e.,
ZnCo2O4 to CoO followed by CoO to Co [53–55]. degree (%)
Reduction

87

84

91

79
theoretical
(mmol/g)

8.5

9.4

9.6

9.0
H2
consumption
(mmol/g)

7.4

7.9

8.7

7.1
H2

(mmol/g)a
Table 4.2. TPR characteristics of Co/MxOy samples

content
Co3O4

2.1

2.4

2.4

2.3
497

-
Main peaks (ºC)

364

362

415
-
199

341

310

367
Co/Gd2O3

Co/CeO2
Co/ZrO2

Co/ZnO
Sample

160
CHAPTER 4

45
a) b) 40

35

Frequency (%)
30

25

20

15

10

0
15 - 25 25 - 35 35 - 45 45 - 55 55 - 65

Particle size (nm)

25
c) d)
20

Frequency (%)
15

10

0
5 - 10 10 - 15 15 - 20 20 - 25 25 - 30 30 - 35 35 - 40 40 - 45

Particle size (nm)

e) f)
15
Frequency (%)

10

0
5 - 10 10 - 15 15 - 20 20 - 25 25 - 30 30 - 35 35 - 40 40 - 45 45 - 50

Particle size (nm)

g) h)
15
Frequency (%)

10

0
5 - 10 10 - 15 15 - 20 20 - 25 25 - 30 30 - 35 35 - 40 40 - 45
Particle size (nm)

Figure 4.2. TEM images and metal particle distribution of (a, b) Co/ZnO, (c, d) Co/CeO 2 , (e, f) Co/ZrO2
and (g, h) Co/Gd2O3 used samples.

161
CHAPTER 4

Co/ZrO2
Intensity (a.u.)

Co/CeO2

Co/ZnO

Co/Gd2O3

100 200 300 400 500 600 700 800 900


Temperature (ºC)

Figure 4.3. TPR profiles of Co/MxOY samples.

3.3. Catalytic evaluation studies

The impact of support nature on the CO2 hydrogenation performance of


Co/MxOy samples, in terms of CO2 conversion, selectivity and TOFCH4, is shown in
Fig. 4.4. Significant differences, both in the CO2 conversion and the selectivity to
various products, were observed amongst the Co/MxOy samples, implying the
key-role of the support on the CO2 hydrogenation activity. In particular, the
Co/CeO2 sample demonstrated the optimum performance in terms of CO2
conversion (97.0% at 300 °C), followed by Co/Gd2O3 (67.4%) > Co/ZrO2
(57.6%) > Co/ZnO (55.1%). It is worth noticing that the conversion performance
follows the same trend as the reduction degree of Co/MxOy catalysts (Table
4.2), implying a possible reducibility-activity correlation, as will be further
discussed below. Regarding the selectivity to various products, both Co/CeO2
and Co/ZnO are highly selective to CH4 (> 90%) at all temperatures examined
(200–300 °C).
162
CHAPTER 4

100

CH4 Selectivity (%) CO2 Conversion (%)


80
60
40
20

80 Co/ZrO2
Co/ZnO
60 Co/CeO2 Co/Gd2O3
40
20
CO Selectivity (%)

80
60
40
20

2.0
TOFCH4 (min )
-1

1.5

1.0

0.5

210 225 240 255 270 285 300


Temperature (ºC)

Figure 4.4. Conversion of CO2, selectivity toward CH4 and CO and TOF of CH4 for the Co/ MxOy
catalysts during CO2 hydrogenation. Reaction conditions: CO2/H2= 1/9 and W/ F = 0.008 g min cm−3

163
CHAPTER 4

On the other hand, the selectivity towards CH4 of Co/ZrO2 and


Co/Gd2O3 follows a similar sigmoidal trend; progressively increasing with the
temperature at the expense of CO. At 300 °C the following CH4 yields (the
product of CO2 conversion and methane selectivity) are obtained: Co/CeO2
(∼96%) > Co/ZnO (∼54%) > Co/G2O3 (∼53%) ∼Co/ZrO2 (∼53%). Higher
hydrocarbons (ethane and propane) as well as methanol were also observed at
the lower temperatures examined, but with selectivities lower than 5%.
Interestingly the TOF values for methane formation follow the same trend with the
achieved CH4 yields. In particular, at 300 °C, the following activity order, in terms
of TOF values were recorded: Co/CeO2 (2.0 min−1) > Co/ZnO (1.8 min−1) >
Co/Gd2O3 (1.3 min−1) ∼ Co/ZrO2 (1.3 min−1). The temperature of 300 °C was
selected to carry out a 24-h stability test for all catalysts (Fig. 4.5). All samples
exhibited stable performance, with insignificant variations in the values of CO2
conversion and CH4 selectivity.

100

100

90
CO2 Conversion (%)

CH4 Selectivity (%)


80

60
80
40
Co/ZnO Co/ZrO2
20 Co/CeO2 Co/Gd2O3 70

0
0 5 10 15 20 25
Time (h)

Figure 4.5. Catalytic stability of the Co/MxOy catalysts during CO2 hydrogenation. Reaction
conditions: 300 °C, CO2/H2 = 1/9 and W/ F =0.008 g min cm−3.

164
CHAPTER 4

Motivated by the incomplete reduction of Co/ZnO under H2 flow (Fig.


4.1, Table 4.2) as well as by its inferior conversion performance compared to
CeO2-based samples, the catalytic behavior of these two samples was further
assessed through successive heating/cooling reaction cycles (see Section 2) to
reveal the impact of pre-treatment under reaction conditions (CO2/H2 mixture)
on the methanation performance.

Fig. 4.6 shows the conversion and selectivity performance of Co/CeO2


and Co/ZnO during the two subsequent cycles.

100 100
CH4 Selectivity (%) CO2 Conversion (%)

CH4 Selectivity (%) CO2 Conversion (%)


st
Co/ZnO (1 cycle)
80 80 nd
Co/ZnO (2 cycle)
60 60
40 st 40
Co/CeO2 (1 cycle)
20 nd 20
Co/CeO2 (2 cycle)
0

98 80
96 60
94 40
92 20
CO Selectivity (%)
CO Selectivity (%)

80
1.5
60
1.0
40
0.5 20

210 225 240 255 270 285 300 210 225 240 255 270 285 300
Temperature (ºC) Temperature (ºC)

Figure 4.6. Conversion and selectivity performance of Co/CeO2 (a) and Co/ZnO (b) in two subsequent
catalytic evaluation tests: catalytic assessment upon temperature increase (1st cycle) followed by cool
down to room temperature under N2 flow (12 h), and reevaluation (2nd cycle). Reaction conditions:
CO2/H2=1/ 9 and W/F = 0.008 g min cm−3.

It is evident that both CO2 conversion and selectivity to CH4 and CO were
not altered in the case of Co/CeO2 sample, implying the establishment of a stable
performance upon the completion of the 1st cycle. In total contrast, the CO2
165
CHAPTER 4

conversion of Co/ZnO was notably increased in the 2nd cycle, indicating a


progressive activation under reaction conditions. At the same time, the selectivity
to CH4 was decreased at lower temperatures during the 2nd cycle, in contrast to
1st cycle where it remains almost stable (> 90%) in the whole temperature range.
It should be also noticed that the catalytic behavior of the used Co/ZnO is
practically the same to the one obtained with the fresh Co/Gd2O3 and Co/ZrO2
samples (Fig. 4.4); at low temperatures the CO2 conversion to CO is favored at
the expense of CH4, whereas at high temperatures the opposite is true. Thus, this
behavior seems to be predominant for the metallic cobalt active phase under
atmospheric pressure. The latter is further confirmed by the XPS findings (see
below) which imply an increased population of reduced Co species on the surface
of the used Co/ZnO catalyst.

However, a different behavior was observed for Co/CeO2 catalyst. At


low temperatures and high CO2 conversion conditions, the selectivity to CH4 is
favored. This could be understood taking into account the mechanism of the CO2
hydrogenation reaction in relation to the specific characteristics of the ceria
support. On the one hand, it is likely that CO2 hydrogenation to methane
proceeds through CO2 adsorption and dissociation into CO and O ad-species
and the subsequent hydrogenation of adsorbed CO [56]. CO2 adsorption takes
place preferably on the metal-support interface, while CO2 dissociation proceeds
on the active metal surface. On the other hand, the presence of hydroxyl (-OH)
groups on partially reduced ceria favors the formation of formate species, which
are considered as active intermediates toward methane formation [56–58]. The
latter could explain the superiority of cobalt-ceria catalyst to convert the
chemisorbed CO to methane in the temperature range explored. In a similar
manner Ni/CeO2 catalysts showed the highest CH4 yield at low temperatures
compared to α-Al2O3-, TiO2- and MgO-based catalysts, which was attributed to
the high surface coverage of ceria by CO2-derived species in conjunction to its
enhanced redox properties [59,60]. In view of the above, Ceria-based mixed
166
CHAPTER 4

oxides have been recently denoted as one of the most promising carriers for CO2
methanation due to their unique redox/surface properties [60,61].

3.4. Surface chemistry elucidation

To gain further insight into the impact of pretreatment procedure (fresh,


reduced, used) on the catalytic performance of Co/CeO2 and Co/ ZnO samples,
XPS analysis was next carried out. Valuable insights in relation to elemental
chemical states and surface composition can be obtained. Fig. 4.7 depicts the Co
2p, Zn 2p and O 1 s XPS spectra for fresh, reduced and used Co/ZnO sample.
The Co 2p spectrum of the fresh sample is characterized by a main peak at ca.
779 eV, accompanied by a low intensity satellite and a spin-orbit doublet
Co2p1/2–Co2p3/2 of about 15 eV. These features point to the formation of Co3+
species in Co3O4-like phase [41,62,63]. Reduction pretreatment results in an
increased intensity of satellite peak at ca.786 eV as well as in an upward shift
of Co 2p peaks by about 0.5 eV. These findings suggest the partial reduction of
Co3+ to Co2+ in octahedral sites [64–68]. Interestingly, after reaction conditions
(used catalysts) an even higher upward shift (by ca. 1.0 eV) is observed, implying
a more efficient reduction of cobalt oxide species. The latter is further verified
by the increase of spinorbit doublet Co2p1/2–Co2p3/2 to 15.6 eV, as compared
to 15.0 and 15.1 eV over fresh and reduced samples, respectively [69].

At the same time the Zn 2p3/2 peaks shown in Fig. 4.7b remain practically
unaffected upon the different treatments. It is known that the Zn 2p doublet is not
particularly sensitive to modifications of the chemical environment around Zn
atoms [70], and even in cases of mixed Zn1-xCoxO oxides fails to give measurable
binding energy shifts as compared to ZnO [71].

Fig. 4.7c depicts the corresponding O 1s XPS spectra for fresh, reduced
and used Co/ZnO sample. The fresh sample is characterized by two main
components at ca. 530 and 531.5 eV. The low binding energy (BE) peak (OI) at
530 eV can be ascribed to the lattice oxygen, whereas the one at high BE (OII)
167
CHAPTER 4

at 531.5 eV is related to adsorbed oxygen species, hydroxyl/carbonate groups


and oxygen vacancies [72].

1022 eV (b)
(a) Co/ZnO
Co/ZnO 780
0
Co Zn 2p3/2

795.6
Used
shake-up satellites 779.5
XPS Intensity (a.u.)

XPS Intensity (a.u.)


Used
794.6
Reduced 779

794 Reduced

Fresh

Fresh
Co 2p1/2 Co 2p3/2

804 798 792 786 780 774 1029 1026 1023 1020 1017 1014

Binding Energy (eV) Binding Energy (eV)

Co/ZnO OII OI (c)


O 1s

Used
XPS Intensity (a.u.)

Reduced

Fresh

538 536 534 532 530 528 526

Binding Energy (eV)

Figure 4.7. XPS spectra of Co/ZnO sample imposed to different pre-treatment in the Co 2p (a), Zn 2p
(b) and O 1s (c) region.

168
CHAPTER 4

The high BE band gains in intensity on reduced and especially on used


samples, implying an increase on the relative population of particularly active
surface oxygen species and/or oxygen vacancies at the expense of lattice
oxygen [69].

The present results reveal a dynamic change of Co/ZnO surface under


reduction (H2) and particularly under reaction (CO2/H2) conditions, which may be
accounted for the differences obtained between the 1st and 2nd catalytic cycle
(Fig. 4.6). No complete reduction is achieved by the H2 reduction procedure prior
the 1st cycle catalytic evaluation, in complete agreement with the XRD and TPR
results, which imply the existence of partially oxidized CoO phase after the
reduction treatment. However, a progressive reduction can be achieved under
reaction conditions (1st cycle test), resulting in a progressive activation and thus
enhanced catalytic performance during the 2nd cycle.

The corresponding XPS spectra for Co/CeO2 sample are depicted in Fig.
4.8. The Ce 3d and O 1s spectra slightly affected by the pretreatment, whereas
the Co 2p spectrum follows a similar behavior to that obtained for Co/ZnO
sample (Fig. 4.7). These findings could be considered responsible for the stable
performance of Co/CeO2 sample upon the subsequent cycles (Fig. 4.6), as further
demonstrated below.

In Table 4.3, the surface content of all chemical elements was calculated
using the corresponding core level peaks, properly normalized to the
photoemission cross section and assuming a homogeneous distribution
arrangement model. It is evident that minor changes in the surface content occur
in the Co/CeO2 samples upon the different treatments. However, the Co/O and
Co/Zn atomic ratios are considerably increased over the used Co/ZnO sample,
as compared to the reduced one, implying an enrichment of catalyst surface to
Co species. Hence, Co/ZnO pretreatment under reaction conditions results in an
increased population of reduced Co species on the catalyst surface, which could

169
CHAPTER 4

be accounted for the differences observed between the 1st and the 2nd cycle of
catalytic measurements. In view of this fact, it was revealed by means of ambient
pressure XPS that the active phase of Co3O4 oxide during the CO2 hydrogenation
to CH4 is metallic cobalt [73].

Figure 4.8. XPS spectra of Co/CeO2 sample imposed to different pre-treatment in the Co 2p (a), Ce 3d
(b) and O 1s (c) region.
170
CHAPTER 4

Table 4.3: Surface atomic ratios in Co/ZnO and Co/CeO2 samples imposed to a different pretreatment.
Sample state
Fresh Reduced Used
Atomic ratio Co/CeO2
Co/O 0.09 0.08 0.09
Ce/O 0.32 0.34 0.31
Co/Ce 0.28 0.23 0.29
Atomic ratio Co/ZnO
Co/O 0.18 0.20 0.33
Zn/O 0.18 0.31 0.27
Co/Zn 1.00 0.64 1.23

4. Conclusions

The CO2 hydrogenation process was investigated on Co/MxOy (M: Ce, Zr,
Gd and Zn) catalysts at temperatures between 200 and 300 °C and under
atmospheric pressure. The results revealed the superiority of Co/CeO2 catalyst,
which exhibited an up to 100% conversion of CO2 to methane. In terms of
methane production yield, the following order was found: Co/CeO2 > Co/ZnO
> Co/ZrO2 > Co/Gd2O3. The improved catalytic performance of Co/CeO2
sample could be mainly ascribed to its superior reducibility linked to Co-Ceria
interactions. Moreover, Co/CeO2 catalyst demonstrated a stable performance in
terms of CO2 conversion and CH4 selectivity upon subsequent reaction cycles. In
contrast, Co/ZnO was progressively activated under reaction conditions due to
the enrichment of catalyst surface to metallic cobalt.

References

[1] Ruhl, C. BP global energy outlook 2030, Vopr. Econ. 5 (2013).

[2] Kondratenko, E.V.; Mul, G.; Baltrusaitis, J.; Larrazábal, G. O.; Pérez-Ramírez,
J. Status and perspectives of CO2 conversion into fuels and chemicals by catalytic,
photocatalytic and electrocatalytic processes. Energy Environmental Science, 6
(2013) 3112-3135.

171
CHAPTER 4

[3] Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic
hydrogenation of carbon dioxide. Chemical Society Reviews, 40 (2011) 3703-
3727.

[4] P.T. Ed. Dlugokencky, NOAA/ESRL (www.esrl.noaa.gov/gmd/ccgg/trends/).

[5] Yang, H.; Xu, Z.; Fan, M.; Gupta, R.; Slimane, R. B.; Bland, A. E.; Wright, I.
Progress in carbon dioxide separation and capture: a review. Journal of
Environmental Science, 20 (2008) 14-27.

[6] Mikkelsen, M.; Jorgensen, M.; Krebs, F. C. The teraton challenge. A review of
fixation and transformation of carbon dioxide. Energy Environmental Science, 3
(2010) 43-81.

[7] Cuéllar-Franca, R. M.; Azapagic, A. Carbon capture, storage and utilisation


technologies: a critical analysis and comparison of their life cycle environmental
impacts. Journal of CO2 Utilization, 9 (2015) 82-102.

[8] Hunt, A. J.; Sin, E. H. K.; Marriott, R.; Clark, J. H. Generation, capture, and
utilization of industrial carbon dioxide. ChemSusChem 3 (2010) 306-322.

[9] Férey, G.; Serre, C.; Devic, T.; Maurin, G.; Jobic, H.; Llewellyn, P. L.; De
Weireld, G.; Vimont, A.; Daturi, M.; Chang, J. S. Why hybrid porous solids
capture greenhouse gases? Chemical Society Reviews 40 (2011) 550-562.

[10] Song, C. Global challenges and strategies for control conversion and
utilization of CO2 for sustainable development involving energy, catalysis,
adsorption and chemical processing. Catalysis Today, 115 (2006) 2-32.

[11] Xiaoding, X.; Moulijn, J. A. Mitigation of CO2 by chemical conversion:


plausible chemical reactions and promising products. Energy Fuels 10 (1996)
305-325.

172
CHAPTER 4

[12] Hu, B.; Guild, C.; Suib, S. L. Thermal, electrochemical, and photochemical
conversion of CO2 to fuels and value-added products. Journal of CO2 Utilization,
1 (2013) 18-27.

[13] Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M.; Joshi, J. B. Catalytic carbon
dioxide hydrogenation to methanol: a review of recent studies. Chemical
Engineering Research and Design, 92 (2014) 2557-2567.

[14] Su, X.; Xu, J.; Liang, B.; Duan, H.; Hou, B.; Huang, Y. Catalytic carbon dioxide
hydrogenation to methane: a review of recent studies. Journal of Energy
Chemistry, 25 (2016) 553-565.

[15] Saeidi, S.; Amin, N. A. S.; Rahimpour, M. R. Hydrogenation of CO2 to value-


added products-a review and potential future developments. Journal of CO2
Utilization, 5 (2014) 66-81.

[16] Bansode, A.; Urakawa, A. Towards full one-pass conversion of carbon


dioxide to methanol and methanol-derived products. Journal of Catalysis, 309
(2014) 66-70.

[17] Díez-Ramírez, J.; Dorado, F.; De la Osa, A.R.; Valverde, J. L.; Sánchez, P.
Hydrogenation of CO2 to methanol at atmospheric pressure over Cu/ZnO
catalysts: influence of the calcination, reduction, and metal loading. Industrial &
Engineering Chemistry Research, 56 (2017) 1979-1987.

[18] Kuhl, K. P.; Cave, E. R.; Abram, D. N.; Jaramillo, T. F. New insights into the
electrochemical reduction of carbon dioxide on metallic copper surfaces. Energy
Environmental Science, 5 (2012) 7050-7059.

[19] Waugh, K. C. Methanol synthesis. Catalysis Letters, 142 (2012) 1153-1166.

[20] Díez-Ramírez, J.; Valverde, J. L.; Sánchez, P.; Dorado, F. CO2 hydrogenation
to methanol at atmospheric pressure: influence of the preparation method of
Pd/ZnO catalysts. Catalysis Letters, 146 (2016) 373-382.

173
CHAPTER 4

[21] Díez-Ramírez, J.; Sánchez, P.; Rodríguez-Gómez, A.; Valverde, J. L.; Dorado,
F. Carbon nanofiber-based palladium/zinc catalysts for the hydrogenation of
carbon dioxide to methanol at atmospheric pressure. Industrial & Engineering
Chemistry Research 55 (2016) 3556-3567.

[22] Gao, J.; Liu, Q.; Gu, F.; Liu, B.; Zhong, Z.; Su, F. Recent advances in
methanation catalysts for the production of synthetic natural gas. RSC Advances,
5 (2015) 22759-22776.

[23] Aziz, M. A. A.; Jalil, A. A.; Triwahyono, S.; Ahmad, A. CO2 methanation over
heterogeneous catalysts: recent progress and future prospects. Green Chemistry,
17 (2015) 2647-2663.

[24] Zaǧli, E.; Falconer, J. L. Carbon dioxide adsorption and methanation on


ruthenium. Journal of Catalysis, 69 (1981) 1-8.

[25] Prairie, M. R.; Renken, A.; Highfield, J. G.; Ravindranathan Thampi, K.;
Grätzel, M. A fourier transform infrared spectroscopic study of CO2 methanation
on supported ruthenium. Journal of Catalysis, 129 (1991) 130-144.

[26] Wierzbicki, D.; Debek, R.; Motak, M.; Grzybek, T.; Gálvez, M. E.; Da Costa,
P. Novel Ni-La-hydrotalcite derived catalysts for CO2 methanation. Catalysis
Communications, 83 (2016) 5-8.

[27] Suslova, E. V.; Chernyak, S. A.; Egorov, A. V.; Savilov, S. V.; Lunin, V. V. CO2
hydrogenation over cobalt-containing catalysts. Kinetics Catalysis, 56 (2015)
646-654.

[28] Gnanamani, M. K.; Jacobs, G. ; Hamdeh, H. H.; Shafer, W. D.; Liu, F.; Hopps,
S. D.; Thomas, G. A.; Davis, B. H. Hydrogenation of carbon dioxide over Co-Fe
bimetallic catalysts. ACS Catalysis, 6 (2016) 913-927.

[29] Das, T.; Sengupta, S.; Deo, G. Effect of calcination temperature during the
synthesis of Co/Al2O3 catalyst used for the hydrogenation of CO2. Reaction
Kinetics Mechanism and Catalysis, 110 (2013) 147-162.
174
CHAPTER 4

[30] Das, T.; Deo, G. Effects of metal loading and support for supported cobalt
catalyst, Catalysis Today 198 (2012) 116-124.

[31] Bakar, W. A. W. A.; Ali, R.; Kadir, A. A. A.; Rosid, S. J. M.; Mohammad, N.
S. Catalytic methanation reaction over alumina supported cobalt oxide doped
noble metal oxides for the purification of simulated natural gas, Ranliao Huaxue
Xuebao. Journal of Fuel Chemistry and Technology, 40 (2012) 822-830.

[32] Srisawad, N.; Chaitree, W.; Mekasuwandumrong, O.; Shotipruk, A.;


Jongsomjit, B.; Panpranot, J. CO2 hydrogenation over Co/Al2O3 catalysts
prepared via a solid-state reaction of fine gibbsite and cobalt precursors.
Reaction Kinetics, Mechanisms and Catalysis, 107 (2012) 179-188.

[33] Zhou, G.; Wu, T.; Xie, H.; Zheng, X. Effects of structure on the carbon dioxide
methanation performance of Co-based catalysts. International Journal of
Hydrogen Energy, 38 (2013) 10012-10018.

[34] Razzaq, R.; Li, C.; Usman, M.; Suzuki, K.; Zhang, S. A highly active and stable
Co4N/γ-Al2O3 catalyst for CO and CO2 methanation to produce synthetic natural
gas (SNG). Chemical Engineering Journal, 262 (2015) 1090-1098.

[35] Lablokov, V.; Beaumont, S. K.; Alayoglu, S.; Pushkarev, V. V.; Specht, C.;
Gao, J.; Alivisatos, A. P.; Kruse, N.; Somorjai, G. A. Size-controlled model Co
nanoparticle catalysts for CO2 hydrogenation: synthesis, characterization, and
catalytic reactions. Nano Letters, 12 (2012) 3091-3096.

[36] Janlamool, J.; Praserthdam, P.; Jongsomjit, B. Ti-Si composite oxide-


supported cobalt catalysts for CO2 hydrogenation. Journal of Natural Gas
Chemistry, 20 (2011) 558-564.

[37] Visconti, C. G.; Lietti, L.; Tronconi, E.; Forzatti, P.; Zennaro, R.; Finocchio, E.
Fischer- Tropsch synthesis on a Co/Al2O3 catalyst with CO2 containing syngas,
Applied Catalysis A, 355 (2009) 61-68.

175
CHAPTER 4

[38] Liu, Y.; Wu, H.; Jia, L.; Fu, Z.; Chen, J.; Li, D.; Yin, D.; Sun, Y. Effect of the
calcination temperature on the catalyst performance of ZrO2-supported cobalt
for Fischer-Tropsch synthesis. Advanced Material Research, 347-353 (2012)
3788-3793.

[39] Takanabe, K.; Nagaoka, K.; Nariai, K.; Aika, K.-i. Influence of reduction
temperature on the catalytic behavior of Co/TiO2 catalysts for CH4/CO2
reforming and its relation with titania bulk crystal structure. Journal of Catalysis,
230 (2005) 75-85.

[40] Choi, J. G.; Rhee, H. K.; Moon, S.H. IR, TPR, and TPD study of supported
cobalt catalysts- effects of metal loading and support materials. Korean Journal
of Chemical Engineering, 1 (1984) 159-164.

[41] Konsolakis, M.; Sgourakis, M.; Carabineiro, S. A. C. Surface and redox


properties of cobalt-ceria binary oxides: on the effect of Co content and
pretreatment conditions. Applied Surface Science, 341 (2015) 48-54.

[42] Jongsomjit, B.; Panpranot, J.; Goodwin Jr., J. G. Co-support compound


formation in alumina-supported cobalt catalysts. Journal of Catalysis, 204 (2001)
98-109.

[43] Puskas, I.; Fleisch, T. H.; Full, P. R.; Kaduk, J. A.; Marshall, C. L.; Meyers, B. L.
Novel aspects of the physical chemistry of Co/SiO2 Fischer-Tropsch catalyst
preparations. The chemistry of cobalt silicate formation during catalyst
preparation or hydrogenation. Applied Catalysis A: General, 311 (2006) 146-
154.

[44] Chu, W.; Chernavskii, P. A.; Gengembre, L.; Pankina, G. A.; Fongarland, P.;

Khodakov, A. Y. Cobalt species in promoted cobalt alumina-supported Fischer-

Tropsch catalysts, J. Catal. 252 (2007) 215–230.

176
CHAPTER 4

[45] Konsolakis, M. The role of copper-ceria interactions in catalysis science:


recent theoretical and experimental advances. Applied Catalysis B, 198 (2016)
49-66.

[46] Al-Musa, A.; Al-Saleh, M.; Ioakeimidis, Z. C.; Ouzounidou, M.; Yentekakis, I.
V.; Konsolakis, M.; Marnellos, G. E. Hydrogen production by iso-octane steam
reforming over Cu catalysts supported on rare earth oxides (REOs). International
Journal of Hydrogen Energy, 39 (2014) 1350-1363.

[47] Luo, W.; Zafeiratos, S. Tuning morphology and redox properties of cobalt
particles supported on oxides by an in between graphene layer. Journal of
Physical Chemistry C, 120 (2016) 14130-14139.

[48] Borodziński, A.; Bonarowska, M. Relation between crystallite size and


dispersion onsupported metal catalysts. Langmuir, 13 (1997) 5613-5620.

[49] Greenwood, N. N.; Earnshaw, A. Chemistry of the Elements, 2nd edition,


(1997).

[50] Bahmani, M.; Vasheghani Farahani, B.; Sahebdelfar, S. Preparation of high


performance nano-sized Cu/ZnO/Al2O3 methanol synthesis catalyst via aluminum
hydrous oxide sol. Applied Catalysis A, 520 (2016) 178-187.

[51] Rekha, V.; Sumana, C.; Douglas, S. P.; Lingaiah, N. Understanding the role
of Co in Co-ZnO mixed oxide catalysts for the selective hydrogenolysis of
glycerol. Applied Catalysis A, 491 (2015) 156-162.

[52] Ladera, R.; Pérez-Alonso, F. J.; González-Carballo, J. M.; Ojeda, M.; Rojas,
S.; Fierro, J. L. G. Catalytic valorization of CO2 via methanol synthesis with Ga-
promoted Cu-ZnO-ZrO2 catalysts. Applied Catalysis B, 142-143 (2013) 241-
248.

[53] Zhang, Y.; Wei, D.; Hammache, S.; Goodwin Jr., J. G. Effect of water vapor
on the reduction of Ru-promoted Co/Al2O3. Journal of Catalysis, 188 (1999)
281-290.
177
CHAPTER 4

[54] Madikizela-Mnqanqeni, N. N.; Coville, N. J. The effect of cobalt and zinc


precursors on Co (10%)/Zn (x%)/TiO2 (x =0,5) Fischer-Tropsch catalysts. Journal
of Molecular Catalysis A: Chemical, 225 (2005) 137-142.

[55] Haneda, M.; Kintaichi, Y.; Bion, N.; Hamada, H. Alkali metal-doped cobalt
oxide catalysts for NO decomposition. Applied Catalysis B, 46 (2003) 473-482.

[56] Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, M.; Lefebvre, J.;

Prabhakaran, P.; Bajohr, S. Review on methanation-from fundamentals to current

Projects. Fuel, 166 (2016) 276-296.

[57] Ernst, B.; Hilaire, L.; Kiennemann, A. Effects of highly dispersed ceria addition
on reductibility, activity and hydrocarbon chain growth of a Co/SiO2 Fischer-
Tropsch catalyst. Catalysis Today 50 (1999) 413-427.

[58] Li, C.; Sakata, Y.; Arai, T.; Domen, K.; Maruya, K.; Onishi, T. Adsorption of
carbon monoxide and carbon dioxide on cerium oxide studied by Fourier-
transform infrared spectroscopy. Part 2. Formation of formate species on
partially reduced CeO2 at room temperature. Journal of the Chemical Society,
Faraday Transactions, 185 (1989) 1451-1461.

[59] Tada, S.; Shimizu, T.; Kameyama, H.; Haneda, T.; Kikuchi, R. Ni/CeO2
catalysts with high CO2 methanation activity and high CH4 selectivity at low
temperatures. International Journal of Hydrogen Energy, 37 (2012) 5527-5531.

[60] Aziz, M. A. A.; Jalil, A. A.; Triwahyono, S.; Ahmadab, A. CO2 methanation
over heterogeneous catalysts: recent progress and future prospects, Green
Chemistry 17 (2015) 2647-2663.

[61] Su, X.; Xu, J.; Liang, B.; Duan, H.; Hou, B.; Huang, Y. Catalytic carbon dioxide
hydrogenation to methane: a review of recent studies. Journal of Energy
Chemistry, 25 (2016) 553-565.

178
CHAPTER 4

[62] Konsolakis, M.; Ioakeimidis, Z. Surface/structure functionalization of copper-


based catalysts by metal-support and/or metal–metal interactions. Applied
Surface Science, 320 (2014) 244-255.

[63] Kim, J. H. X-ray photoelectron spectroscopy analysis of (Ln1-xSrx) CoO3-δ (Ln:


Pr, Nd and Sm). Applied Surface Science, 258 (2011) 350–355.

[64] Kang, M.; Song, M. W.; Lee, C. H. Catalytic carbon monoxide oxidation over
CoOx/ CeO2 composite catalysts. Applied Catalysis A, 251 (2003) 143-156.

[65] Liotta, L. F.; Di Carlo, G.; Pantaleo, G.; Venezia, A. M.; Deganello, G.
Co3O4/CeO2 composite oxides for methane emissions abatement: relationship
between Co3O4-CeO2 interaction and catalytic activity. Applied Catalysis B 66
(2006) 217-227.

[66] Gawade, P.; Bayram, B.; Alexander, A. M. C.; Ozkan, U. S. Preferential


oxidation of CO (PROX) over CoOx/CeO2 in hydrogen-rich streams: effect of
cobalt loading. Applied Catalysis B, 128 (2012) 21-30.

[67] Gómez, L. E.; Boix, A. V.; Miró, E. E. Co/ZrO2, Co/CeO2 and MnCoCe
structured catalysts for COPrOx. Catalysis Today, 216 (2013) 246-253.

[68] Bonnelle, J. P.; Grimblot, J. ; D'Huysser, A. Influence de la polarisation des


liaisons sur les spectres esca des oxydes de cobalt. Journal Electron Spectroscopy
and Related Phenomena, 7 (1975) 151-162.

[69] Tarwal, N. L.; Gurav, K. V.; Prem Kumar, T.; Jeong, Y. K.; Shim, H. S.; Kim, I.
Y.; Kim, J. H.; Jang, J. H.; Patil, P. S. Structure, X-ray photoelectron spectroscopy
and photoluminescence investigations of the spray deposited cobalt doped ZnO
thin films. Journal of Analytical and Applied Pyrolysis, 106 (2014) 26-32.

[70] Ton-That, C.; Phillips, M. R.; Foley, M.; Moody, S. J.; Stampfl, A. P. J.; Surface
electronic properties of ZnO nanoparticles, Applied Physics Letters, 92 (2008).

179
CHAPTER 4

[71] Turczyniak, S.; Luo, W.; Papaefthimiou, V.; Ramgir, N. S.; Haevecker, M.;
Machocki, A.; Zafeiratos, S. A comparative ambient pressure X-ray photoelectron
and absorption spectroscopy study of various cobalt-based catalysts in reactive
atmospheres, Topics in Catalysis, 59 (2016) 532-542.

[72] Konsolakis, M.; Ioakimidis, Z.; Kraia, T.; Marnellos, G. E. Hydrogen production
by ethanol steam reforming (ESR) over CeO2 supported transition metal (Fe, Co,
Ni, Cu) catalysts: insight into the structure-activity relationship. Catalysts, 6 (2016).

[73] Zhu, Y.; Zhang, S.; Ye, Y.; Zhang, X.; Wang, L.; Zhu, W.; Cheng, F.; Tao, F.
Catalytic conversion of carbon dioxide to methane on ruthenium-cobalt bimetallic
nanocatalysts and correlation between surface chemistry of catalysts under
reaction conditions and catalytic performances. ACS Catalysis, 2 (2012) 2403-
2408.

180
CHAPTER 5
OPTIMIZATION OF THE Pd/Cu
RATIO IN Pd-Cu-Zn/SiC CATALYST
CHAPTER 5

Abstract
PdCuZn/SiC catalysts were synthesized with different Pd:Cu:Zn molar
compositions and tested in the hydrogenation of carbon dioxide to methanol at
atmospheric pressure. Trimetallic catalysts were compared with the corresponding
bimetallic ones (PdZn/SiC, CuZn/SiC and PdCu/SiC). Catalysts were
characterized by N2 adsorption/ desorption, temperature-programed reduction
(TPR), X-ray diffraction (XRD), transmission electron microscopy (TEM), energy
dispersive X-ray spectroscopy (EDS) and X-ray photoelectron spectroscopy (XPS).
The Pd0 active sites were related to carbon monoxide formation via reverse
water-gas-shift (RWGS), whereas the PdZn alloys catalyzed methanol synthesis.
The role of copper in trimetallic catalysts was to inhibit the deposition of metallic
palladium by forming a PdCu alloy that proved to be less active to CO formation.
Moreover, the active sites of trimetallic catalysts were smaller and better
dispersed than those of the corresponding bimetallic ones, probably due to a
synergistic effect between the three metals. The catalyst with a molar composition
of 37.5:12.5:50 Pd:Cu:Zn (mol.%) was selected as the most active for the
methanol synthesis, as this sample showed the highest activity and selectivity to
methanol. The role of copper was also shown to be crucial in trimetallic catalyst
by comparing the best example with an equivalent bimetallic PdZn/SiC with a
Pd:Zn molar ratio of 37.5:62.5.
Methanol formation rate (mol min-1 g-1)

CuZn/SiC 6 CuZn/SiC
12.5 PdCuZn/SiC
PdZn/SiC
PdCu/SiC
12.5PdCuZn/SiC
4 25PdCuZn/SiC
PdZn/SiC 37.5PdCuZn/SiC 25 PdCuZn/SiC
2

PdCu/SiC 37.5 PdCuZn/SiC


0
150 175 200 225 250 275 300
Temperature (ºC)

183
CHAPTER 5

1. Introduction

Carbon dioxide emissions not only lead to dramatic environmental


problems, from their contribution to global warming [1] to acidification of the
oceans [2], but also to important economic issues [3]. The development of new
legislation concerning CO2 emissions and the challenging targets set by the global
Paris Agreement and the EU 2030 Framework for Climate and Energy are forcing
governments and companies to implement innovative solutions to prevent and
reduce CO2 emissions [4].

In this sense, the valorization of CO2 has gained attention in recent


decades [5–9] since it provides both a reduction of atmospheric emissions and
the synthesis of added-value chemicals. In this respect, the hydrogenation of CO2
(Eq. (5.1)) has been assessed as one of the potential valorization routes, from
which CO, hydrocarbons, alcohols or aldehydes can be produced [6,8].

xCO2 + (2x – z + y/2) H2 → CxHyOz + (2x – z) H2O (5.1)

Methanol is reported to be one of the most interesting products (Eq. (5.2))


because it is used as a solvent and a feedstock for the production of chemicals
[7, 10]. Furthermore, methanol could be used as an alternative fuel in the energy
distribution infrastructure that currently exists or it could be blended with gasoline
[11]. The economic limitation of this reaction is currently related to the use of high
amounts of hydrogen in the process, which could be overcome by applying
different technologies such as the electrolysis of water using renewable energies
(e.g., solar energy) [12].

CO2 + 3H2 ⇆ CH3 OH + H2 O ∆H25°C = –49.5 kJ/mol (5.2)

Due to the stable nature of the CO2 molecule, this reaction needs to
overcome a high thermodynamic barrier. CO2 activation takes place at 150–200
°C and at this temperature the reverse WGS reaction (Eq. (5.3)), which has an
adverse effect on methanol production, also occurs.

184
CHAPTER 5

CO2 + H2 ⇆ CO + H2 O ∆H25°C = 41 kJ/mol (5.3)

Therefore, the use of catalysts to control the selectivity towards the


different products is a prerequisite. Although metals such as Au [13-15], Ag [16]
or Mo [17] have been used in CO2 hydrogenation to methanol, copper- and
palladium-based catalysts have been the most widely studied because of their
well-known properties and high efficiency in this reaction [18-24]. Indeed, the
bimetallic Pd-Cu configuration has also been studied at high pressure and the
main conclusion was that the PdCu alloy formed was a crucial factor for the
bimetallic promotion [25]. On the other hand, zinc oxide (ZnO) has been reported
to enhance the catalytic performance of Cu and Pd. In the case of Pd, ZnO is
necessary to form the PdZn alloy, which has proven to be more active and
selective to methanol than the corresponding isolated metals [22, 24]. Regarding
Cu catalysts, Ahouari et al. [18] summarized the role of ZnO as follows: (i) it
promotes higher dispersion of Cu, thus preventing the agglomeration of Cu
particles, (ii) it improves the resistance of Cu particles to poisoning by feed gas
impurities, (iii) ZnO, as a basic oxide, partially neutralizes the acidity of the
catalyst and enhances CO2 adsorption on the catalyst surface and (iv) ZnO also
acts as a reservoir of atomic hydrogen and provides this gas to achieve methanol
synthesis on the Cu surface.

The combination of these three metals has been reported previously in the
literature [26-29]. Melían-Cabrera et al. [27] compared CuO/ZnO with Pd-
promoted CuO/ZnO catalysts (PdO:CuO:ZnO = 2:28:70 wt.%) under an
operating pressure of 60 bar and found a considerable improvement when Pd
was used. They suggested that this improvement was a result of an enhancement
of the spillover mechanism of H2 due to the presence of Pd, as well as a further
stabilization of Cu due to possible PdCu alloy formation. More recently,
Siriworarat et al. [29] studied the influence of Pd loading (5, 10 and 15 wt.%)
on a Pd-Cu-Zn catalyst in which Cu and Zn were settled at 25 wt.%. In this study,

185
CHAPTER 5

which was carried out at 250 °C and 25 bar, the highest performance was
observed on using the catalyst containing 15 wt.% Pd, 25 wt.% Cu and 25 wt.%
Zn, which gave methanol selectivity and space time yield of 25% and 112 g kg
cat−1 h−1, respectively. Although these papers revealed that the combination of
these metals improves the methanol yield, they only considered the role of Pd as
a promoter of Cu/Zn catalysts and did not obtain an optimum composition for the
catalyst. As Pd is reported to be active in this reaction, an in-depth study on the
Pd-Cu-Zn composition is still necessary.

Regarding supports, different materials have been used to date (Al2O3,


CeO2, SiO2, ZrO2, CNF, CNT, etc.) [22, 30–36]. However, the use of β-SiC in this
reaction has not been widely reported in the literature [37], even though it has
shown excellent properties in many reactions [38–43]. β-SiC exhibits high thermal
conductivity and mechanical strength, low specific weight and chemical inertness,
so that this support is appropriate to avoid interactions with the active phases.

In the work reported here trimetallic Pd-Cu-Zn catalysts were prepared,


characterized and tested in the synthesis of methanol from CO2 hydrogenation
at atmospheric pressure. Taking into account the secondary role of Zn in Pd/ZnO
and Cu/ZnO formulations (see above), a highly enough proportion of Zn was
deposited in each catalyst. On the other hand, as Pd and Cu are reported to be
the active phases in these formulations, the proportion of these metals was
modified in a broad range in order to get an optimal trimetallic formulation.
These catalysts were compared with the corresponding bimetallic ones (PdCu,
PdZn and CuZn). In all cases, β-SiC was used as a support. To the best of our
knowledge, there is a lack of studies that deal with this reaction at atmospheric
pressure [21, 22, 24, 44, 45].

186
CHAPTER 5

2. Experimental

2.1. Catalyst preparation

Catalysts were prepared by the impregnation method using β-SiC (SICAT


CATALYST, pellets) as support and three different nitrates as precursors of the
Cu, Pd and Zn: copper (II) nitrate trihydrate [Cu (NO3)2.3H2O, Panreac, 99.95%
purity], palladium(II) nitrate [Pd (NO3)2.xH2O, Aldrich] and zinc nitrate
hexahydrate [Zn(NO3)2.6H2O, Panreac, 99% purity].

Firstly, 5 g of the support were placed in a glass vessel and kept under
vacuum at room temperature (∼25 °C) for 2 h to remove water and other
impurities adsorbed on the structure. Secondly, an aqueous solutions of the
corresponding metal nitrates were poured over the support, using the
appropriate quantities to obtain catalysts with a total amount of 0.01 mol and
different Pd, Cu and Zn contents, as shown in Table 5.1. Thirdly, the solvent was
removed under vacuum at 90 °C for 2 h. After impregnation, the catalysts were
dried at 120 °C overnight.

The calcination was carried out at 500 °C in a Nabertherm HTC 03/15


furnace, which was open to the atmosphere. This temperature was kept constant
for 3 h and the heating rate was 5 °C min−1. Prior to the reaction, the catalysts
were reduced in situ in a 25% v/v H2/N2 stream at a flow rate of 100 Ncm3
min−1 from room temperature to 500 °C with a heating rate of 1.3 °C min−1.

The nomenclature of the catalysts is shown in Table 5.1. Trimetallic


catalysts are denoted as XPdCuZn/SiC, where X indicates the theoretical molar
percentage of metallic palladium in the sample.

187
Table 5.1. Nomenclature and main physical properties of the catalysts.

Metal loading (mol.%) Surface Total pore Average XRD- TEM-


area volume pore Particle Particle
CHAPTER 5

Palladium Copper Zinc m2 g-1) x 102 radius (nm) diameter diameter


(cm3 g-1) (nm)a (nm)
Nomenclature Theo. Exp. Theo. Exp. Theo. Exp.

SiC - - - - - - 25 14.1 11.5 - -

CuZn/SiC - - 50 54 50 46 17 10.7 12.5 52 (Cu) 40.2

PdZn/SiC 50 55 - - 50 45 29 11.7 8 22 (PdZn) 20.0

188
PdCu/SiC 50 40.7 50 59.3 - - 25 16.8 13.4 53 (PdCu) 28.5
50 (Cu)

12.5PdCuZn/SiC 12.5 16 37.5 40 50 44 21 13.4 13 26 (PdCuZn) 25.6


35 (Cu)

25PdCuZn/SiC 25 30 25 24 50 46 23 13.4 11.5 14 (PdCuZn) -

37.5PdCuZn/SiC 37.5 39.4 12.5 14.3 50 46.3 29 12.5 8.4 11 (PdCuZn) 12.5

a In brackets, the compound which it refers to.


CHAPTER 5

2.2. Support/catalyst characterization

The Cu, Pd and Zn metal loadings were determined by atomic absorption


(AA) spectrophotometry on a SPECTRA 220FS analyzer. Samples (ca. 0.5 g) were
treated with 2 mL HCl, 3 mL HF and 2 mL H2O2 followed by microwave digestion
(250 °C). Surface area/porosity measurements were carried out on a
QUADRASORB 3SI sorptometer apparatus with N2 as the sorbate at −196 °C.
The samples were outgassed at 250 °C under vacuum (5 ×10−3 Torr) for 12 h
prior to analysis. Specific surface areas were determined by the multi-point BET
method and the mesopore size distribution with the BJH method. Specific total
pore volume was evaluated from N2 uptake at a relative pressure of P/P0 =
0.99. The relative error of the all the experiments was ± 5% Temperature-
programed reduction (TPR) experiments were conducted in a commercial
Micromeritics AutoChem 2950 HP unit with TCD detection. Samples (ca. 0.15 g)
were loaded into a U-shaped tube and ramped from room temperature (∼25
°C) to 900 °C (10 °C min−1), with a reducing gas mixture of 20% v/v H2/Ar (60
cm3 min−1). XRD experiments were conducted on a Philips X’Pert instrument using
nickel-filtered Cu-Kα radiation. Samples were scanned at a rate of 0.02° step−1
over the range 5° ≤2θ≤ 90° (scan time = 2 s step−1). Transmission electron
microscopy (TEM) analyses were carried out on a JEOL JEM-4000EX unit with an
accelerating voltage of 400 kV. Samples were prepared by ultrasonic dispersion
in acetone with a drop of the resulting suspension evaporated onto a holey
carbon-supported grid. The instrument was equipped with an energy dispersive
X-ray spectroscopy (EDS) unit. Saturation was assumed to be complete after three
successive peaks showed the same peak areas. The particle size was calculated
from TEM images. The mean particle size evaluated as the surface-area weighted
diameter (ds) was computed according to:

∑ nd 3
d̅ s = ∑i i i2 (5.4)
i ni di

189
CHAPTER 5

where ni represents the number of particles with diameter di (Σini≥200).

XPS analyses were performed on an ESCAPlus Omicron spectrometer


using a monochromatized Mg source (MnKα 1253.6 eV). The high-resolution
spectra were recorded with a 40 eV pass energy, and the constant charging of
the samples was corrected by referencing all energies to the C1s peak at 284.6
eV. Curve fitting was performed using the CasaXPS software.

2.3. Catalyst activity

Catalytic performance tests were carried out in a tubular quartz reactor


(45 cm length and 1 cm diameter). The catalyst, which consisted of pellets that
were 3 mm length and 1 mm diameter, was placed on a fritted quartz plate
located at the end of the reactor. The amount of catalyst used in the experiments
was 0.8 g.

The temperature of the catalyst was measured with a K-type


thermocouple (Thermocoax) placed inside the inner quartz tube. The entire
reactor was placed in a furnace (Lenton) equipped with a
temperatureprogramed system. Reaction gases were Praxair certified standards
of CO2 (99.999% purity), H2 (99.999% purity) and N2 (99.999% purity). The
gas flows were controlled by a set of calibrated mass flowmeters (Brooks 5850
E and 5850 S).

The hydrogenation of CO2 was carried out at atmospheric pressure in the


temperature range 150–300 °C. The total flow rate used in the experiments,
which involved a CO2/H2 mixture (CO2/H2 =1/9 v/v), was 100 Ncm3 min−1. Gas
effluents were monitored with a micro gas chromatograph (Varian CP-4900)
fitted with a PoraPLOT Q column and a molecular sieve column, each of which
was connected to a thermal conductivity detector (TCD). All catalytic tests were
performed twice and the relative error was less than 5%.

Catalytic parameters were calculated as follows (Eq. (5.5)–(5.7)):


190
CHAPTER 5

F
Formation ratei (μmol∙min-1 ∙g-1 )= Catalysti weight (5.5)

Fi
Selectivityi (%)= F0 × 100 (5.6)
CO2 -FCO2

F0CO2 -FCO2
CO2 conversion (%)= × 100 (5.7)
F0CO2

where Fi represents the molar flow (μmol min−1) of the i component


(CH3OH or CO).

3. Results and discussion

3.1. Textural properties

The main textural properties of the support and the prepared catalysts
are listed in Table 5.1 and the corresponding nitrogen adsorption-desorption
isotherms are plotted in Fig. 5.1.

It can be observed that all isotherms corresponded to type II–IV according


to the IUPAC classification, and this is characteristic of macroporous and
mesoporous materials. A slight volume increase in the middle relative pressure
range (0.2–0.5) was observed, whereas a sharp increase took place at high
partial pressure (0.8–1) [46, 47]. The observed H3-type hysteresis loop (IUPAC
classification), which is characteristic of mesoporous materials, is associated with
nitrogen capillary condensation [46]. It is worth noting that the shapes of the
catalyst isotherms were almost identical to that of the β-SiC support, indicating
that the support surface structure did not suffer any significant change after metal
incorporation. The variations in the BET surface area, pore volume and average
pore radius observed in the catalysts when compared to the support are due to
the partial blockage of the β-SiC pores. This blockage is clearly observed in the
mesopore distribution of CuZn/ SiC (Fig. 5.1c), which exhibited the same shape
as the support but a lower percentage of pores for each radius. Copper particles
191
CHAPTER 5

led to blockage of the pores but the incorporation of palladium led to a decrease
in the size of these pores. This behavior can be observed in the mesopore
distributions of the trimetallic catalysts (Fig. 5.1d), where it can be seen that the
higher the quantity of palladium, the smaller the pore size. This fact suggests that
the presence of Pd led to the formation of smaller metal particles, which partially
blocked the pores and thus created smaller ones.

100 a) CuZn/SiC - A b) 12.5PdCuZn/SiC - A


CuZn/SiC - D 12.5PdCuZn/SiC - D
PdZn/SiC - A 25PdCuZn/SiC - A
80 PdZn/SiC - D 25PdCuZn/SiC - D
Volume (cc/g)

PdCu/SiC - A 37.5PdCuZn/SiC - A
60 PdCu/SiC - D 37.5PdCuZn/SiC - D
SiC - A SiC - A
40 SiC - D SiC - D

20

0
0.0 0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
Partial Pressure (P/P0) Partial Pressure (P/P0)
CuZn/SiC 12.5PdCuZn/SiC
0.20 c) d)
PdZn/SiC 25PdCuZn/SiC
PdCu/SiC 37.5PdCuZn/SiC
0.15 SiC SiC
dV/d(log r)

0.10

0.05

0.00

10 100 1000 10000 10000010 100 1000 10000 100000


Pore radius (Å) Pore radius (Å)

Figure 5.1. Nitrogen adsorption-desorption isotherms and mesopore distributions of the support and
prepared catalysts.

3.2. Metal reducibility

TPR experiments (Fig. 5.2) were conducted in order to identify an


appropriate reduction temperature for all of the catalysts. The results of the TPR
analyses for bimetallic catalysts are shown in Fig. 5.2a. Firstly, CuZn/SiC analysis
gave rise to a single broad peak at around 200 °C, which was assigned to the
combination of the reduction of Cu2+ to Cu1+ and Cu1+ to Cu0 [21]. Moreover, a
192
CHAPTER 5

small broad peak was observed between 450 °C and 680 °C, which can be
attributed to ZnO reduction. The PdCu/SiC profile showed a peak at around 50
°C with a shoulder at 90 °C, which was caused by the combination of the reduction
of PdO to metallic Pd and the reduction of copper (CuO to Cu0). Therefore, it can
be concluded that Pd catalyzed the Cu reduction and led to a shift of the
corresponding Cu reduction peak to lower temperatures. Finally, PdZn/SiC
showed a small peak at around 55 °C, which was also assigned to the reduction
of PdO to metallic palladium [48], as well as broad peaks above 400 °C (similar
to those that can be seen more clearly in trimetallic catalysts, Fig. 5.2b), which
are commonly related to crystalline PdZn alloy formation [49–51]. It is important
to note that the TPR profile of the support did not show any noticeable peak.

If the amount of hydrogen consumed for these peaks were calculated,


only a small proportion of palladium would have been reduced. According to the
literature [49, 50], hydrogen consumption occurs rapidly when hydrogen is fed
over the catalyst at room temperature and PdO is partially converted to PdHx
[48]. These palladium hydrides are then decomposed to metallic palladium [24].
A proper quantification of such changes was not possible to carry out as hydrogen
was in contact with the catalyst before the TPR data recording started, so that it
was not possible to draw any conclusion concerning this fact.

As the amount of Pd increased, the trimetallic catalyst profiles changed


from the CuZn/SiC-type to the PdZn/SiC-type. In this way, the profile for
12.5PdCuZn/SiC showed a broad peak (50 °C) with two shoulders (60 and 140
°C), which are ascribed to Pd (PdO → Pd0) and Cu (Cu2+ →Cu+ → Cu0) reduction,
respectively. As mentioned above, this finding revealed how palladium catalyzes
the reduction of copper particles and thus decreases the temperature of the
reduction peak. The 25PdCuZn/SiC profile was similar to that of the bimetallic
PdCu/SiC, with the shoulder for the trimetallic catalyst shifted to higher
temperatures due to the lower amount of Pd. Finally, the 37.5PdCuZn/SiC sample

193
CHAPTER 5

showed a pattern similar to that of PdZn/SiC. All of the trimetallic catalysts


showed different broad peaks above 400 °C and these could be related to the
formation of PdZn alloy during the reduction process [52], in a similar way to
PdZn/SiC.

CuZn/SiC
PdCu/SiC
PdZn/SiC
SiC
TCD Signal (a.u.)

0 100 200 300 400 500 600

12.5PdCuZn/SiC
25PdCuZ/SiC
37.5PdCuZn/SiC
SiC
TCD Signal (a.u.)

0 100 200 300 400 500 600

0 100 200 300 400 500 600 700


Temperature (ºC)
Figure 5.2. Temperature-programmed reduction (TPR) profiles of (a) bimetallic and (b) tri-metallic
catalysts.

194
CHAPTER 5

Bearing in mind the TPR results, 500 °C was selected as an appropriate


reduction temperature to ensure the complete reduction of all metals and the
formation of the PdZn alloy, which is an active phase for methanol formation [22,
24]. It is worth noting that the β-SiC structure remained unchanged at this
temperature, since this material did not show any appreciable peak in the
temperature range studied.

3.3. Crystalline structure and active metal phases

The X-ray diffractograms of reduced bimetallic and trimetallic catalysts


are shown in Fig. 5.3. On the one hand, the diffractograms of bimetallic catalysts
contain the main peaks related to the metallic phases or alloys formed, as well
as those attributed to the β-SiC (JPCDS 02-1050). Hence, the CuZn/SiC
diffractogram showed the metallic cubic copper crystal structure (JPCDS 85-
1326), whereas the PdZn/SiC diffractogram contained the peaks of the PdZn
alloy (JPCDS 06-0620). Finally, the PdCu/SiC diffractogram showed the peaks
related to PdCu alloy (JPCDS 48-1551) along with those of metallic copper. The
presence of ZnO in all of the catalysts prepared with zinc was expected but the
main diffraction peak at 2θ = 36.2° (JCPDS 80-0075) was hidden by those of
β-SiC.

The peak identification was more difficult in the case of trimetallic


catalysts as it was expected that some diffraction peaks would overlap. Although
it could be asserted that 12.5PdCuZn/SiC contained metallic copper particles in
its structure, the other peak at around 41° could correspond to PdZn, PdCu or
both alloys, the main diffraction peaks of which appear at similar values (2θ–
41.2° and 41.4°), respectively. The same problem was found in the
diffractograms of 25PdCuZn/SiC and 37.5PdCuZn/SiC, for which it was not
possible to identify the broad peak and its shoulder observed between 2θ = 40°
and 45°. Even if the identification were easier, the peak could shift respect its

195
CHAPTER 5

corresponding 2θ value, which would indicate the formation of alloys as observed


in other metal configurations [53, 54]. Nevertheless, it was not possible to
conclude anything from XRD results. Finally, although diffraction peaks of Pd0
were expected, the high reflections of β-SiC along with the high dispersion of this
metal made impossible any identification.

# #
# SiC ^ Cu * PdZn + PdCu
# # # #
^
^ ^
CuZn/SiC 12.5PdCuZn/SiC ^
# #
Intensity (a.u.)

* #
# #
#
*
PdZn/SiC 25PdCuZn/SiC
^ #
#
#
# # #
+
PdCu/SiC ^ 37.5PdCuZn/SiC

10 20 30 40 50 60 70 80 10 20 30 40 50 60 70 80 90
2(º) 2(º)

Figure 5.3. XRD profiles of the reduced catalysts.

The particle size diameters obtained by XRD, which were calculated from
the main diffraction peaks of the corresponding compounds, are listed in Table
5.1. The results are consistent with the N2 adsorption/ desorption results. On the
one hand, larger metal particles, which blocked the pores of the support, were
expected for CuZn/SiC and PdCu/SiC. On the other hand, smaller metal particles
were found as the palladium content increased in trimetallic catalysts. These
smaller metal particles partially blocked the β-silicon carbide pores, thus
decreasing their size. It is important to note that the average particle sizes for
25PdCuZn/SiC and 37.5PdCuZn/SiC were lower than that of PdZn/ SiC, which
suggests that a synergistic effect occurs between metals and this leads to smaller
particles.

196
CHAPTER 5

To gain an insight into the compounds formed on the surface of the


support, TEM and EDX analyses were carried out on various regions of the
reduced catalysts. As an example, representative results are shown in Fig. 5.4
and Table 5.2, respectively. All catalysts exhibited a Gaussian particle size
distribution and the mean particle sizes evaluated by TEM are consistent with
those obtained by XRD (Table 5.1). The smallest particles were found in trimetallic
37.5PdCuZn/SiC. Energy dispersive X-ray (EDX) microanalysis results revealed
that, regardless of the percentage of copper, the particles of both trimetallic
catalysts contained a high level of this metal in their structure. The differences in
metal percentages between the other metals (Pd and Zn) confirmed a higher
percentage of palladium in 37.5PdCuZn/SiC. As for PdZn/SiC, it was observed
in all cases a Pd/Zn ratio higher than 1 (1.25 in Region B), which suggested, in
agreement with XRD results, the formation of PdZn and Pd0 phases during
reduction.

Bearing in mind the characterization results obtained to this point, it is


possible to suggest a theory regarding the active phases formed in each sample
during reduction (Fig. 5.5). On the one hand, the active phases of bimetallic
catalysts should be those expected from XRD- with the exception of PdZn/SiC,
for which the smallest particles found in the particle size distribution (< 5 nm, Fig.
5.4d) are probably related to metallic palladium, since this noble metal is known
to disperse well. On the other hand, the formation of the active phases is more
complex in trimetallic catalysts. According to the representative EDX results listed
in Table 5.2, in which the atomic percentage of Cu was far higher than those of
Pd and Zn, the first metal that is deposited on the support is copper, followed by
palladium and zinc, which are deposited to form the PdZn alloy observed by XRD
(Fig. 5.3). In addition, the remaining palladium is likely to form an alloy (PdCu)
with the copper on which it is deposited. Finally, due to the higher percentage of
copper impregnated in 12.5PdCuZn/SiC, isolated copper metal particles are
deposited, as also observed by XRD (Fig. 5.3).
197
CHAPTER 5

35

a) 30 b)
25

Frequency (%)
20

15

Region A 10

0
< 10 10-20 20-30 30-40 40-50 50-60 60-70
Particle size (nm)

c) Region B 40 d)
30

Frequency (%)
20

10

0
<5 5-10 10-15 15-20 20-25 25-30 30-35 35-40
Particle size (nm)

e) 25 f)
20
Frequency (%)

15

10
Region C
5

0
5-10 15-20 25-30 35-40 45-50
Particle size (nm)

30

g) 25
h)
20
Frequency (%)

15

10

5
Region D
0
5-10 10-15 15-20 20-25 25-30 30-35 35-40 > 40
Particle size (nm)

25
i) j)
20
Frequency (%)

15
Region E
10

0
2.5-5 7.5-10 12.5-15 17.5-20
Particle size (nm)

Figure 5.4. TEM images and metal particle distribution of (a, b) CuZn/SiC; (c, d) PdZn/SiC; (e, f)
PdCu/SiC; (g, h) 12.5PdCuZn/SiC and (i, j) 37.5PdCuZn/SiC samples.
198
CHAPTER 5

Outside the regions marked in Fig. 5.4, isolated ZnO was expected to
deposit because of the higher percentage of this compound. This fact was
confirmed by XPS spectra, which also provided valuable information that
supported the theory outlined above (see hereafter).

Table 5.2. Energy dispersive X-ray microanalysis results of regions from Fig. 5.4.

Atomic %
Element Region A Region B Region C Region D Region E
Cu 78.7 - 68 74.3 62.8
Pd - 55.5 32 5.9 18.9
Zn 21.3 44.5 - 19.8 18.3

CuZn/SiC 12.5 PdCuZn/SiC

SiC

25 PdCuZn/SiC ZnO
PdZn/SiC
Cu
PdZn

PdCu/SiC 37.5 PdCuZn/SiC PdCu

Pd

Figure 5.5. Proposed theory for active phase formation during reduction.

199
CHAPTER 5

3.4. Metal surface structures

To gain a more in-depth knowledge of the metal surface structures formed


in the catalysts and detect the presence of potential alloys, the chemical state
and the relative surface abundance of copper, palladium and zinc were
evaluated by X-ray photoelectron spectroscopy (XPS). The Cu 2p, Pd 3d and Zn
2p XPS spectra for the bimetallic and trimetallic catalysts are shown in Fig. 5.6
and the binding energies and the atomic ratios of all elements are listed in Table
5.3.

The Cu 2p spectra showed a single peak at around 932 eV, which


corresponds to the 2p3/2 value of metallic copper. As expected, in trimetallic
catalysts the lower the amount of copper, the lower the intensity of the peak. On
the other hand, the low atomic ratios of Cu/Zn and Cu/Pd demonstrate that
copper was not deposited efficiently on the surface, with levels that are
practically negligible in CuZn/SiC and 37.5PdCuZn/SiC. These findings confirm
the theory proposed above (Fig. 5.5).

As far as the Pd 3d spectra are concerned, the 3d5/2 binding energies of


the peaks observed in trimetallic samples were similar to those reported in the
literature for the PdZn alloy (336 eV) [55]. The 3d5/2 peak of PdZn/SiC was
shifted to lower values and this confirmed the presence of metallic palladium
(335 eV) on the surface of the catalyst. This fact was corroborated by
quantification of the Pd/Zn atomic ratios (Table 5.3), with a value higher than
one obtained for PdZn/SiC. In contrast, the value for 37.5PdCuZn/SiC was close
to one. Finally, the PdCu/SiC sample also gave the 3d5/2 peak in the binding
energy of metallic palladium (335 eV). In contrast to PdZn, it has been reported
in previous studies [25, 56] that the Pd 3d doublet of the bulk PdCu alloy did not
show any peak shift with respect to the pure metal.

Finally, three different profiles were observed in the Zn 2p spectra.


CuZn/SiC and 12.5PdCuZn/SiC showed a peak at 1023 eV, which corresponds
200
CHAPTER 5

to the 2p3/2 level of zinc oxide. As to PdZn/SiC, this peak was shifted to lower
binding energies, thus confirming that all of the surface zinc was used to form the
PdZn alloy.
Cu0
CuZn/SiC Cu 2p

PdCu/SiC

CPS
12.5PdCuZn/SiC

37.5PdCuZn/SiC

940 935 930 925


Binding Energy (eV)

Pd 3d PdZn alloy Pd0


PdCu/SiC

PdZn/SiC
CPS

12.5PdCuZn/SiC

37.5PdCuZn/SiC

345 342 339 336 333 330


Binding Energy (eV)

Zn 2p ZnO Zn0

CuZn/SiC

PdZn/SiC
CPS

12.5PdCuZn/SiC

37PdCuZn/SiC

1050 1040 1030 1020


Binding Energy (eV)
Figure 5.6. XPS spectra of reduced catalysts in the Cu 2p, Pd 3d and Zn 2p regions.

201
CHAPTER 5

The case of 37.5PdCuZn/SiC warrants a special mention as the 2p3/2


peak maximum appeared between the values of ZnO and Zn0. In this catalyst,
although Pd and Zn also interacted to form a PdZn alloy, the initial amount of Pd
introduced was lower than that of Zn and the remaining ZnO was observed in the
spectrum. For this reason, in the interpretation of the species distribution in the
catalyst surface (Fig. 5.5) it was expected that the ZnO would be present in
trimetallic catalysts. In brief, the XPS results support the theory proposed and are
consistent with the results of the characterization techniques described above.
Table 5.3. Binding energies and surface atomic ratios for the bimetallic and tri-metallic catalysts.

Binding Energies (eV) Atomic ratio


Sample Pd 3d5/2 Cu 2p3/2 Zn 2p3/2 Pd/Zn Cu/Zn Cu/Pd
CuZn/SiC - 932.5 1023.3 - 0.09 -
PdCu/SiC 335.3 932.8 - - - 0.16
PdZn/SiC 335.1 - 1021.3 2.04 - -
12.5PdCuZn/SiC 336.1 932.6 1023.2 0.71 0.15 0.21
37.5PdCuZn/SiC 335.7 931.8 1022.8 1.18 0.07 0.06

3.5. Catalytic activity

The catalytic performance of bimetallic and trimetallic catalysts is shown


in Fig. 5.7. On the one hand, it can be observed that a higher reaction
temperature led to a higher formation rate of carbon monoxide (Fig. 5.7a). The
PdZn/SiC catalyst showed the highest CO formation rate by the reverse water-
gas-shift reaction (RWGS, eq. 5.3), closely followed by CuZn/SiC, whereas the
behavior of the bimetallic PdCu/SiC was similar to that observed for trimetallic
catalysts. The curves for the methanol formation rate were different to those for
carbon monoxide (Fig. 5.7b), with a maximum value observed between 225 and
275 °C depending on the catalyst. PdCu/SiC did not show any activity towards
methanol and this indicates that the PdCu alloy observed in this catalyst was not
active for methanol production at atmospheric pressure-in contrast to the situation

202
CHAPTER 5

reported for high pressure conditions (4.1 MPa), where the PdCu alloy was
selective to methanol [25]. The methanol formation rate for trimetallic catalysts
increased as the palladium loading increased, and the curves varied from the
CuZn/SiC type profile, whose activity to methanol was limited, to the PdZn/SiC
type profile, in which the PdZn alloy played an important role in methanol
formation [24,44]. Interestingly, the methanol formation rate for 37.5PdCuZn/SiC
was higher at 250 and 275 °C than that for PdZn/ SiC, which revealed that the
combination of Pd-Cu-Zn improved the catalytic performance towards methanol
synthesis. This attractive behavior was observed more clearly in the selectivity
towards methanol (Fig. 5.7c), where the trimetallic catalyst showed the highest
values over the whole range of temperatures studied. Indeed, the values obtained
with 12.5PdCuZn/SiC, in which only a low level of palladium was deposited, were
higher than those for PdZn/SiC and CuZn/SiC. This finding also underlines the
synergistic effect produced by the combination of Pd, Cu and Zn. Finally, the CO2
conversion curves (Fig. 5.7d) showed a similar trend to the CO formation rate
curves, since these species were the predominant products at high temperatures.

It should be noted that, even though the catalytic experiments were


carried out under atmospheric pressure, the results are far away from the
thermodynamic equilibrium values (except at 275 °C) showing that there are not
thermodynamic limitations (see Table 5.4 where the thermodynamic and
experimental results for the 37.5PdCuZn/SiC catalyst are compared). The
thermodynamic equilibrium values were calculated using a flowsheet simulator
(Aspen HYSYS V8.4 licensed by Aspen Technology,Inc.). Peng Robinson was used
as the equation of state and the reactor modeling was based on a Gibbs reactor.
The conditions used for the simulation (flow rate, CO2/H2 ratio) were the same as
in the experimental reactor.

203
CHAPTER 5

CH3OH formation rate (mol min-1 g-1) CO formation rate (mol min g )
-1 -1
60 a) CuZn/SiC
PdZn/SiC
50
PdCu/SiC
40 12.5PdCuZn/SiC
25PdCuZn/SiC
30 37.5PdCuZn/SiC
20

10

b)
5

c)
100
CH3OH Selectivity (%)

80

60

40

20

d) CuZn/SiC
10 PdZn/SiC
CO2 conversion (%)

PdCu/SiC
8
12.5PdCuZn/SiC
6
25PdCuZn/SiC
37.5PdCuZn/SiC
4

0
150 175 200 225 250 275 300
Temperature (ºC)

Figure 5.7. Catalytic activity of bimetallic and trimetallic catalysts. Formation rates of (a) CO and (b)
CH3OH. (c) CH3OH selectivity and (d) CO2 conversion. Reaction conditions: CO2/H2 = 1/9 (v/v) and
W/F=0.008 g min cm-3.
204
CHAPTER 5

Table 5.4. Comparison between thermodynamic equilibrium and experimental values for the catalyst
37.5PdCuZn/SiC.

Thermodynamic Experimental
Temperature (ºC)
(µmol/min) (µmol/min)
150 22.24 0.39
175 14.72 0.75
200 10.09 1.47
225 7.14 2.53
250 5.19 3.85
275 3.87 3.82
300 2.95 1.94

The catalytic results can be explained by the theory proposed for the
formation of particles on the catalyst surface (Fig. 5.5). The high CO formation
rate observed in PdZn/SiC was probably due to the presence of small particles
of metallic palladium, which were observed in the reduced catalyst, and this idea
is consistent with previous studies [22, 24]. On the other hand, when trimetallic
catalysts were reduced, all of the metallic palladium interacted with Zn and Cu
to form the corresponding PdZn and PdCu alloys. This new alloy (PdCu) was not
as active to carbon monoxide as metallic palladium (Fig. 5.7a) and it can
therefore be concluded that copper acted as an inhibitor of this metal in
trimetallic catalysts, which in turn led to a higher selectivity to methanol. Moreover,
it can be seen from the catalyst particle diameters (Table 5.1) that trimetallic
37.5PdCuZn/SiC average particle size was lower than that of bimetallic
PdZn/SiC. Therefore, the metal particles were better dispersed in the trimetallic
catalyst, probably due to a synergistic effect between the metals. The higher
methanol formation rate obtained on using 37.5PdCuZn/SiC could be explained
by taking into account two facts: (i) The PdZn alloy particles of this catalyst were
smaller and better dispersed than those of bimetallic PdZn/SiC, and (ii) once the
metallic palladium had been transformed into PdCu alloy, which was observed
to be less active towards CO formation than metallic palladium, the reactant
molecules (H2 and CO2) were prone to react on the PdZn active sites, thus leading
to a higher methanol formation rate.

205
CHAPTER 5

In an effort to demonstrate that the presence of copper is necessary to


obtain a catalyst that is more active and selective to methanol than the PdZn/SiC
catalyst, a new PdZn/SiC catalyst (37.5PdZn/SiC) was prepared with molar
percentages of 37.5% and 62.5% for palladium and zinc, respectively. The
results for comparison with the corresponding trimetallic catalyst
(37.5PdCuZn/SiC) are shown in Fig. 5.8. It was observed that the activities to
methanol and CO of 37.5PdZn/SiC were lower and higher, respectively, than
those of 37.5PdCuZn/SiC. These results support the situation described above in
that the activity of 37.5PdZn/SiC to CO was higher because, in the absence of
copper, particles of metallic palladium that led to CO production remained on
the catalyst. At the same time, the activity to CH3OH decreased due to the high
activity of these metallic palladium active sites, which led to competition between
them and the PdZn sites to adsorb H2 and CO2. Therefore, the contribution of
copper in trimetallic 37.5PdCuZn/SiC proved to be crucial.

Finally, the catalytic performance of 37.5PdCuZn/SiC was compared to


the results reported by Iwasa et al. [44], which to the best of our knowledge is
the only study on CO2 hydrogenation at atmospheric pressure with palladium zinc
catalysts (Table 5.5). Despite the differences between the catalysts (metal
loading, morphology) and the reaction conditions (463 vs 473 K), it can be
asserted that trimetallic 37.5PdCuZn/SiC was more active and, in particular,
selective towards methanol than Pd/ZnO or Cu/ZnO.

In conclusion, the aim of this work was achieved and the trimetallic catalyst
37.5PdCuZn/SiC proved to be more active and selective to methanol than the
bimetallic PdZn/SiC one. The synergistic effect between Pd-Cu-Zn is responsible
for this promising behavior.

206
CHAPTER 5

50

CO formation rate (mol min-1 g-1)


a) 37.5PdCuZn/SiC
40 37.5PdZn/SiC

30

20

10

CH3OH Selectivity (%) CH OH formation rate (mol min-1 g-1)


b)
5 37.5PdCuZn/SiC
37.5PdZn/SiC
4

0
3

c) 37.5PdCuZn/SiC
100
37.5PdZn/SiC
80

60

40

20

8 d) 37.5PdCuZn/SiC
37.5PdZn/SiC
CO2 conversion (%)

0
150 175 200 225 250 275 300
Temperature (ºC)

Figure 5.8. Comparison between the catalytic activity of 37.5PdCuZn/SiC and 37.5 PdZn/SiC. Formation
rates of (a) CO and (b) CH3OH. (c) CH3OH selectivity and (d) CO2 conversion. Reaction conditions:
CO2/H2 = 1/9 (v/v) and W/F = 0.008 g min cm–3

207
CHAPTER 5

Table 5.5. Comparison of the 37.5PdCuZn/SiC catalytic performance with the literature.

Methanol catalytic
Reaction conditions performance
Temperature CO2/H2 Activity SCH3OH
Sample (ºC) (v/v) (µmol∙g-1∙min-1) (%) Ref.
37.5PdCuZn/SiC 200 1/9 1.84 80.9 This work
Pd/ZnO 190 1/9 1.28 65.1 [44]
Cu/ZnO 190 1/9 1.74 30.4 [44]

4. Conclusions
The following conclusions can be drawn from this study:

- Cu, Pd and Zn exhibited a synergistic effect in trimetallic catalysts. On the one


hand, Pd interacted with Zn and Cu to form the corresponding PdZn and PdCu
alloys, which were selective towards methanol and CO, respectively. On the other
hand, the average metal particle size was generally lower than that of the
corresponding bimetallic catalysts.

- TEM and XPS analyses of trimetallic catalysts suggest that, during the reduction
process, ZnO and Cu were the first compounds deposited onto the SiC support
and the alloys of PdZn and PdCu were then formed on top.

- Cu acted as an inhibitor of metallic palladium through the formation of the PdCu


alloy, which was less active towards CO than Pd0.

- The combination of PdZn and PdCu active sites led to highly methanol-selective
trimetallic catalysts.

- An optimum trimetallic catalyst (37.5PdCuZn/SiC) was selected as the best


example as it showed the highest methanol formation rate. Hence, an optimization
of the composition of the trimetallic catalysts was performed, which was of crucial
interest because of the high price of palladium.

208
CHAPTER 5

References
[1] Harde, H. Scrutinizing the carbon cycle and CO2 residence time in the
atmosphere, Global Planet. Change, 152 (2017) 19-26.

[2] Pistevos, J. C. A.; Nagelkerken, I.; Rossi, T.; Connell, S. D. Ocean acidification
alters temperature and salinity preferences in larval fish. Oecologia, 183 (2017)
545-553.

[3] Nordhaus, W. D. Revisiting the social cost of carbon. Proceedings of the


National Academy of Science of the United States, 114 (2017) 1518-1523.

[4] Villoria-Sáez, P.; Tam, V. W. Y.; Río Merino, M. D.; Viñas Arrebola, C.; Wang,
X. Effectiveness of greenhouse-gas Emission Trading Schemes implementation: a
review on legislations. Journal of Cleaner Production, 127 (2016) 49-58.

[5] Alaba, P. A.; Abbas, A.; Daud, W. M. W. Insight into catalytic reduction of
CO2: Catalysis and reactor design. Journal of Cleaner Production, 140 (2017)
1298-1312.

[6] Saeidi, S.; Amin, N. A. S.; Rahimpour, M. R. Hydrogenation of CO2 to value-


added products - A review and potential future developments. Journal of CO2
Utilization, 5 (2014) 66-81.

[7] Ganesh, I. Conversion of carbon dioxide into methanol - A potential liquid


fuel: Fundamental challenges and opportunities (a review). Renewable &
Sustainable Energy Reviews, 31 (2014) 221-257.

[8] Olajire, A. A. Valorization of greenhouse carbon dioxide emissions into value-


added products by catalytic processes. Journal of CO2 Utilization, 3-4 (2013)
74-92.

[9] Gutiérrez-Guerra, N.; Moreno-López, L.; Serrano-Ruiz, J. C.; Valverde, J. L.;


de Lucas-Consuegra, A. Gas phase electrocatalytic conversion of CO2 to syn-fuels
on Cu based catalysts-electrodes. Applied Catalysis B, 188 (2016) 272-282.

209
CHAPTER 5

[10] Ali, K. A.; Abdullah, A. Z.; Mohamed, A. R. Recent development in catalytic


technologies for methanol synthesis from renewable sources: A critical review.
Renewable & Sustainable Energy Reviews, 44 (2015) 505-518.

[11] Olah, G. A. Beyond Oil and Gas: The Methanol Economy. Angewandte
Chemie International Edition, 44 (2005) 2636-2639.

[12] Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M.; Joshi, J. B. Catalytic carbon
dioxide hydrogenation to methanol: A review of recent studies. Chemical
Engineering Research and Design, 92 (2014) 2557-2567.

[13] Sloczynski, J.; Grabowski, R.; Kozlowska, A.; Olszewski, P.; Stoch, J.;
Skrzypek, J.; Lachowska, M. Catalytic activity of the M/(3ZnO·ZrO2) system (M
= Cu, Ag Au) in the hydrogenation of CO2 to methanol. Applied Catalysis A, 278
(2004) 11-23.

[14] Asara, G. G.; Ricart, J. M.; Rodriguez, J. A.; Illas, F. Exploring the activity of
a novel Au/TiC (001) model catalyst towards CO and CO2 hydrogenation.
Surface Science, 640 (2015) 141-149.

[15] Vourros, A.; Garagounis, I.; Kyriakou, V.; Carabineiro, S. A. C.; Maldonado-
Hódar, F. J.; Marnellos, G. E. ; Konsolakis, M. Carbon hydrogenation over
supported Au nanoparticles: effect of the support. Journal of CO2 Utilization, 19
(2017) 247-256.

[16] Grabowski, R.; Sloczynski, J.; Sliwa, M.; Mucha, D.; Socha, R. P. Influence of
polymorphic ZrO2 phases and the silver electronic state on the activity of
Ag/ZrO2 catalysts in the hydrogenation of CO2 to methanol. ACS Catalysis, 1
(2011) 266-278.

[17] Reyes, P.; Concha, I.; Pecchi, G.; Fierro, J. L. G. Changes induced by metal
oxide promoters in the performance of Rh-Mo/ZrO2 catalysts during CO and
CO2 hydrogenation. Journal of Molecular Catalysis A: Chemical, 129 (1998)
269–278.
210
CHAPTER 5

[18] Ahouari, H.; Soualah, A.; Le Valant, A.; Pinard, L.; Magnoux, P.; Pouilloux, Y.

Methanol synthesis from CO2 hydrogenation over copper based catalysts.

Reactions Kinetics Mechanisms and Catalysis, 110 (2013) 131-145.

[19] Kunkes, E. L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M.

Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2O3: Is there a common

intermediate or not? Journal of Catalysis, 328 (2015) 43-48.

[20] Wang, G.; Chen, L.; Sun, Y.; Wu, J.; Fu, M.; Ye, D. Carbon dioxide

hydrogenation to methanol over Cu/ZrO2/CNTs: effect of carbon surface

chemistry. RSC Advances, 5 (2015) 45320-45330.

[21] Díez-Ramírez, J.; Dorado, F.; de la Osa, A. R.; Valverde, J. L.; Sánchez, P.

Hydrogenation of CO2 to methanol at atmospheric pressure over Cu/ZnO

catalysts: influence of the calcination, reduction, and metal loading. Industrial &

Engineering Chemistry Research, 56 (2017) 1979-1987.

[22] Díez-Ramírez, J.; Sánchez, P.; Rodríguez-Gómez, A.; Valverde, J. L.;

Dorado, F. Carbon nanofiber-based palladium/zinc catalysts for the

hydrogenation of carbon dioxide to methanol at atmospheric pressure. Industrial

& Engineering Chemistry Research, 55 (2016) 3556-3567.

[23] Díez-Ramírez, J.; Sánchez, P.; Valverde, J. L.; Dorado, F. Electrochemical

promotion and characterization of PdZn alloy catalysts with K and Na ionic

conductors for pure gaseous CO2 hydrogenation. Journal of CO2 Utilization, 16

(2016) 375-383.

211
CHAPTER 5

[24] Díez-Ramírez, J.; Valverde, J. L.; Sánchez, P.; Dorado, F. CO2 Hydrogenation

to methanol at atmospheric pressure: influence of the preparation method of

Pd/ZnO catalysts. Catalysis Letters, 146 (2016) 373-382.

[25] Jiang, X.; Koizumi, N.; Guo, X.; Song, C. Bimetallic Pd-Cu catalysts for

selective CO2 hydrogenation to methanol. Applied Catalysis B, 170-171 (2015)

173-185.

[26] Melián-Cabrera, I.; López Granados, M.; Terreros, P.; Fierro, J. L. G. CO2

hydrogenation over Pd-modified methanol synthesis catalysts. Catalysis Today,

45 (1998) 251-256.

[27] Melián-Cabrera, I.; Granados, M. L.; Fierro, J. L. G. Effect of Pd on Cu-Zn

catalysts for the hydrogenation of CO2 to methanol: Stabilization of Cu metal

against CO2 oxidation. Catalysis Letters, 79 (2002) 165-170.

[28] Sahibzada, M.; Chadwick, D.; Metcalfe, I. S. Hydrogenation of carbon

dioxide to methanol over palladium-promoted Cu/ZnO/Al2O3 catalysts.

Catalysis Today, 26 (1996) 367-372.

[29] Siriworarat, K.; Deerattrakul, V.; Dittanet, P.; Kongkachuichay, P. Production

of methanol from carbon dioxide using palladium-copper-zinc loaded on MCM-

41: Comparison of catalysts synthesized from flame spray pyrolysis and sol-gel

method using silica source from rice husk ash. Journal of Cleaner Production, 142

(2017) 1234-1243.

212
CHAPTER 5

[30] Mierczynski, P.; Ciesielski, R.; Kedziora, A.; Shtyka, O.; Maniecki, T. P.

Methanol synthesis using copper catalysts supported on CeO2−Al2O3 mixed

oxide. Fibre Chemistry, 48 (2017) 271-275.

[31] Phan, X. K.; Walmsley, J. C.; Bakhtiary-Davijany, H.; Myrstad, R.; Pfeifer, P.;

Venvik, H.; Holmen, A. Pd/CeO2 catalysts as powder in a fixed-bed reactor and

as coating in a stacked foil microreactor for the methanol synthesis. Catalysis

Today, 273 (2016) 25-33.

[32] Phongamwong, T.; Chantaprasertporn, U.; Witoon, T.; Numpilai, T.; Poo-

arporn, Y.; Limphirat, W.; Donphai, W.; Dittanet, P.; Chareonpanich, M.;

Limtrakul, J. CO2 hydrogenation to methanol over CuO–ZnO–ZrO2–SiO2

catalysts: Effects of SiO2 contents. Chemical Engineering Journal, 316 (2017)

692-703.

[33] Larmier, K.; Liao, W. C.; Tada, S.; Lam, E.; Verel, R.; Bansode, A.; Urakawa,

A.; Comas-Vives, A.; Copéret, C. CO2-to-Methanol hydrogenation on zirconia-

supported copper nanoparticles: reaction intermediates and the role of the

metal–support interface. Angewandte Chemie International Edition, 56 (2017)

2318-2323.

[34] Dong, X.; Zhang, H. B.;. Lin, G. D; Yuan, Y. Z.; Tsai, K. R. Highly Active CNT-

Promoted Cu–ZnO–Al2O3 Catalyst for Methanol Synthesis from H2/CO/CO2.

Catalysis Letters, 85 (2003) 237-246.

213
CHAPTER 5

[35] Díez-Ramírez, J.; Sánchez, P.; Kyriakou, V.; Zafeiratos, S.; Marnellos, G. E.;

Konsolakis, M.; Dorado, F. Effect of support nature on the cobalt-catalyzed CO2

hydrogenation. Journal of CO2 Utilization, 21 (2017) 562–571.

[36] Bonura, G.; Cordaro, M.; Cannilla, C.; Arena, F.; Frusteri, F. The changing

nature of the active site of Cu-Zn-Zr catalysts for the CO2 hydrogenation reaction

to methanol. Applied Catalysis B, 152–153 (2014) 152–161.

[37] Halim, N. S. A.; Zabidi, N. A. M.; Tasfy, S. F. H.; Shaharun, M. S. Morphology

and performance of Cu/ZnO based catalyst: comparison between Al2O3 and SiC

support. AIP Conference Proceedings 2016.

doi: http://dx.doi.org/10.1063/1.4968073.

[38] García-Vargas, J. M.; Valverde, J. L.; Díez, J.; Sánchez, P.; Dorado, F.

Influence of alkaline and alkaline-earth cocations on the performance of Ni/β-

SiC catalysts in the methane tri-reforming reaction. Applied Catalysis B, 148-149

(2014) 322-329.

[39] García-Vargas, J. M.; Valverde, J. L.; Díez, J.; Sánchez, P.; Dorado, F.

Preparation of Ni-Mg/β-SiC catalysts for the methane tri-reforming: effect of the

order of metal impregnation. Applied Catalysis B, 164 (2015) 316-323.

[40] García-Vargas, J. M.; Valverde, J. L.; Díez, J.; Dorado, F.; Sánchez, P.

Catalytic and kinetic analysis of the methane tri-reforming over a Ni-Mg/β-SiC

catalyst. International Journal of Hydrogen Energy, 40 (2015) 8677-8687.

[41] Díaz, J. A.; Calvo-Serrano, M.; De La Osa, A. R.; García-Minguillán, A. M.;

Romero, A.; Giroir-Fendler, A.; Valverde, J. L. β-Silicon carbide as a catalyst


214
CHAPTER 5

support in the fischer-tropsch synthesis: Influence of the modification of the support

by a pore agent and acidic treatment. Applied Catalysis A, 475 (2014) 82-89.

[42] Zou, J.; Mu, X.; Zhao, W.; Rukundo, P.; Wang, Z. J. Improved catalytic activity

of SiC supported Ni catalysts for CO2 reforming of methane via surface

functionalizations. Catalysis Communications, 84 (2016) 116-119.

[43] Li, J.; Wang, J.; Gao, D.; Li, X.; Miao, S.; Wang, G.; Bao, X. Silicon carbide-

supported iron nanoparticles encapsulated in nitrogen-doped carbon for oxygen

reduction reaction. Catalysis Science Technology, 6 (2016) 2949-2954.

[44] Iwasa, N.; Suzuki, H.; Terashita, M.; Arai, M.; Takezawa, N. Methanol

synthesis from CO2 under atmospheric pressure over supported Pd catalysts.

Catalysis Letters, 96 (2004) 75-78.

[45] Maniecki, T. P.; Mierczynski, P.; Maniukiewicz, W.; Bawolak, K.; Gebauer,

D.; Jozwiak, W.K. Bimetallic Au-Cu, Ag-Cu/CrAl3O6 catalysts for methanol

synthesis. Catalysis Letters, 130 (2009) 481-488.

[46] Sing, K. S. W.; Everett, D. H.; Haul, R. A. W.; Moscou, L.; Pierotti, R. A.;

Rouquerol, J.; Siemieniewska, T. Reporting physisorption data gor gas/solid

systems. 1984, 567-583.

[47] Díaz, J. A.; Akhavan, H.; Romero, A.; Garcia-Minguillan, A. M.; Romero, R.;

Giroir-Fendler, A.; Valverde, J. L. Cobalt and iron supported on carbon

nanofibers as catalysts for Fischer-Tropsch synthesis. Fuel Processing Technology,

128 (2014) 417-424.

215
CHAPTER 5

[48] Chin, Y. H.; Dagle, R.; Hu, J.; Dohnalkova, A. C.; Wang, Y. Steam reforming

of methanol over highly active Pd/ZnO catalyst. Catalysis Today, 77 (2002) 79-

88.

[49] Iwasa, N.; Mayanagi, T.; Ogawa, N.; Sakata, K.; Takezawa, N. New

catalytic functions of Pd-Zn, Pd-Ga, Pd-In, Pt-Zn, Pt-Ga and Pt-in alloys in the

conversions of methanol. Catalysis Letters, 54 (1998) 119-123.

[50] Iwasa, N.; Masuda, S.; Ogawa, N.; Takezawa, N. Steam reforming of

methanol over Pd/ZnO: Effect of the formation of PdZn alloys upon the reaction.

Applied Catalysis A, 125 (1995) 145-157.

[51] Chin, Y. H.; Wang, Y.; Dagle, R. A.; Li, X. S.; Methanol steam reforming over

Pd/ZnO: Catalyst preparation and pretreatment studies. Fuel Processing

Technology, 83 (2003) 193-201.

[52] Wang, Y.; Zhang, J.; Xu, H. Interaction between Pd and ZnO during

Reduction of Pd/ZnO Catalyst for Steam Reforming of Methanol to Hydrogen.

Chinese Journal of Catalysis, 27 (2006) 217-222.

[53] Antolini, E.; Cardellini, F. Formation of carbon supported PtRu alloys: an XRD

Analysis. Journal of Alloys and Compounds, 315 (2001) 118–122.

[54] K. Hyun, J. H. Lee, C. W. Yoon, Y. Kwon, The effect of platinum based

bimetallic electrocatalysts on oxygen reduction reaction of proton exchange

membrane fuel cells. International Journal of Electrochemical Science, 8 (2013)

11752–11767.

216
CHAPTER 5

[55] Bahruji, H.; Bowker, M.; Hutchings, G.; Dimitratos, N.; Wells, P.; Gibson, E.;

Jones, W.; Brookes, C.; Morgan, D.; Lalev, G. Pd/ZnO catalysts for direct CO2

hydrogenation to methanol. Journal of Catalysis, 343 (2016) 133–146.

[56] Rochefort, A.; Abon, M.; Delichère, P.; Bertolini, J. C. Alloying effect on the

adsorption properties of Pd50Cu50{111} single crystal surface. Surface Science,

294 (1993) 43–52.

217
CHAPTER 6
KINETICS OF THE HYDROGENATION
OF CO2 TO METHANOL AT
ATMOSPHERIC PRESSURE USING A
Pd-Cu-Zn/SiC CATALYST
CHAPTER 6

Abstract
The kinetics of the hydrogenation of CO2 to methanol (MeOH) at
atmospheric pressure using a Pd-Cu-Zn/SiC catalyst has been analyzed. An initial
sensitivity study was performed in order to evaluate the effect of reaction
conditions (temperature, CO2/H2 ratio and the presence of products in the feed
stream) on the catalytic performance. The results of this study were used to
develop three Langmuir–Hinshelwood kinetic models in which the adsorption term
was modified (competitive vs two-site vs three-site adsorption mechanism). All of
the kinetic models predicted the experimental results well and the corresponding
parameters were statistically meaningful. Model discrimination revealed that the
three-site adsorption mechanism led to the lowest residual sum of squares and
was the only one that met all of the parameter constraints. The quality of this
model was evaluated by comparing the results of additional experiments with
the predicted values. The three-site adsorption mechanism agreed with the
catalytic observations reported previously, where it was observed that, in the
presence of a Pd-Cu-Zn/SiC catalyst, the synthesis of MeOH by hydrogenation
of CO2 took place on PdZn active sites, whereas the Reverse Water Gas Shift
(RWGS), which led to CO, was catalyzed by PdCu sites. The H2 dissociative
adsorption was believed to take place on ZnO.
CH3OH formation rate (μmol·min-1·g-1)

Langmuir-Hinselwood three-sites mechanism

CO2 H2 CO2 H2
CH3OH CO

CO2* CO2*
H* H* H* H*

20 Nml·min-1 CO2
Feed volume flow: 80 Nml·min-1 H2
10 Nml·min-1 N2

221
CHAPTER 6

Nomenclature
CS CO2 surface concentration
Deff Catalyst effective diffusivity [m2/s]
Dx Adsorption terms
Eax Activation energy [kJ/mol]
FPi Species molar flow
ΔH Heat of reaction [kJ/mol]
ΔH0ads Enthalpy of adsorption [kJ/mol]
k'x Reaction rate constant
Ki,S Adsorption equilibrium constant of species i in the active site s
Keqx Equilibrium constant of reaction rate
n Number of data
NW-P Weisz–Prater number
p Number of parameters
Pi Partial pressure of the component i [bar]
R Gas constant [R = 8.314 J/mol K]
r CO2 initial conversion rate [%]
rx Reaction rate [moli/gcat min]
r2 Regression coefficient
s1 PdZn active site
s2 ZnO active site
s3 PdCu active site
si Residual (Piexp – Pitheo)
ΔS0ads Entropy of adsorption [kJ/mol K]
T Temperature [K]
W Catalyst mass [g]
w Weighting factor
WRSS Weighted Residual Sum of Squares

Greek Letters

ρc Catalyst density [kg/m3]

Subscripts

exp Experimental data


i Species
theo Theoretical data
x Reactions (MeOH-CO2, RWGS, MeOH-CO)

222
CHAPTER 6

1. Introduction
Carbon dioxide is considered to be the main contributor to global
warming and it is therefore one of the most harmful pollutants to the ecosystem
[1]. At the same time, it is one of the most promising sources of carbon to produce
other compounds such as hydrocarbons, alcohols, and aldehydes, among others
[2]. In this regard, the scientific community is working on the development of
processes to transform the CO2 pollution problem into an opportunity to obtain
valuable products.

The hydrogenation of carbon dioxide to methanol (MeOH, Eq. (6.1)) is


one of the ways in which this pollutant can be valorized [3]. MeOH is frequently
used as a solvent and a feedstock to produce chemicals. MeOH could be used as
a fuel in the energy distribution infrastructure that currently exists, in a direct
MeOH fuel cell [4], or it could be blended with gasoline [5].

CO2 + 3H2 ⇆ CH3 OH + H2 O ∆H25°C = –49.5 𝑘𝐽∙𝑚𝑜𝑙 −1 (6.1)

The reverse water gas shift reaction (RWGS, Eq. (6.2)), which also
participates in the hydrogenation process, leads to carbon monoxide and this is
considered to be an undesirable by-product. In this sense, the use of catalysts in
order to improve the MeOH selectivity and the CO2 conversion is crucial.

CO2 + H2 ⇆ CO + H2 O ∆H25°C = 41 𝑘𝐽∙𝑚𝑜𝑙 −1 (6.2)

Up to date, the vast majority of the research concerning this reaction is


focused on the development of novel catalysts. In this regard, Wisaijorn et al. [6]
modified copper-based catalysts by oyster shellderived calcium oxide, and
concluded that this material can prevent sintering of Cu particles and promoted
the adsorption of CO2. Zhao et al. [7] used Atomic Layer Deposition (ALD) to
deposite Ni particles over Cu/Al2O3 catalysts, which led to the formation of CuNi
alloys which were active in this reaction. Another example is the work reported
by An et al. [8], who supported in situ Cu/ZnOx nanoparticles in Metal-Organic

223
CHAPTER 6

Frameworks (MOFs), at a reaction temperature of 250 °C under hydrogen


atmosphere. As a result, they obtained ultra-small Cu/ZnOx nanoparticles which
showed high catalytic activity and selectivity to methanol.

Several different kinetic models for the synthesis of MeOH by CO2


hydrogenation have been reported in the literature to date. Skrzypek et al. [10]
proposed a Langmuir–Hinshelwood (LH) kinetic model that considers the
competitive adsorption of species. Graaf et al. [11] supposed a dual-site
mechanism in which H2 and H2O are adsorbed by ZnO active sites and
carbonaceous species by metallic copper sites. This model has also been
proposed by several authors [13]. Finally, Park et al. [12] supposed a three-site
adsorption mechanism, where the adsorption of hydrogen occurs on ZnO whereas
CO and CO2 are adsorbed on Cu1+ and Cu0, respectively. The vast majority of
the kinetic models, such as those outlined above, have been developed using
commercial Cu/ZnO/Al2O3 catalysts and only some few ones have been
developed with other catalyst configurations, such as Cu/ZnO/Al2O3/ZrO2 [13]
or Pd-Ga2O3/Silica [14]. Moreover, only Chiavassa et al. [14] studied exclusively
the hydrogenation of CO2.

In our previous work [15], we optimized a PdCuZn/SiC catalyst by


modifying the Pd/Cu ratio. Taking into account the secondary role of Zn in this
reaction (hydrogen adsorption and dissociation), a highly enough proportion of
this metal was deposited in each catalyst. On the other hand, as Pd and Cu are
reported to be active phases in this reaction, the proportion of these metals was
modified in a broad range in order to get an optimal trimetallic formulation,
which was more active and selective than the corresponding bimetallic ones. By
doing this, it was possible to tailor the formation of PdZn and PdCu nanoparticles
which catalyzed the synthesis of methanol (Eq. (6.1)) and inhibited the RWGS (Eq.
(6.2)), respectively, thus obtaining an optimum trimetallic formulation (see below)
that maximizes the methanol formation rate. This catalyst formulation will be used
in the later stage of the project, in which an electrocatalytic reactor, operating at
224
CHAPTER 6

atmospheric pressure, will be developed for the synthesis of methanol with a feed
of CO2 and H2O. This electrocatalytic reactor will have two chambers separated
by a co-ionic ceramic conductor. Water splitting will take place in the anodic
chamber, whereas methanol synthesis will occurred in the cathodic one. By
application of an electrical current, the system will act as an electrochemical
hydrogen/oxygen pump (H+ will be pumped to the cathode and O2– will be
pumped away from the cathode). Under this mode of operation, controlled H+
fluxes to the cathodic catalyst are expected to create high surface hydrogen
activities, which in turn will force equilibrium to the production of methanol,
bypassing the conventional necessity for high pressure. At the same time O2–
pumping from the cathode is expected to activate the C=O bond of the adsorbed
COX species, so that methanol synthesis will be accelerated. This way, it is
expected to overcome the two main drawbacks of the so-called conventional
catalytic operation: the need of high pressure operation and fossil fuel-derived
hydrogen gas.

The aim of the work described here was to perform the kinetic analysis
and modeling for the synthesis of MeOH by CO2 hydrogenation at atmospheric
pressure, using the PdCuZn/SiC catalyst above mentioned, as a prior step to the
development of the electrocatalytic reactor. Firstly, a sensitivity study was carried
out by varying the composition of the feed stream and the temperature. Secondly,
three different Langmuir-Hinshelwood kinetic models were developed as a
function of the adsorption mechanism. Finally, the proposed models were
compared and a proper model discrimination was carried out.

2. Experimental

2.1. Catalyst preparation and characterization

A PdCuZn/SiC catalyst with a molar percentage of 37.5% Pd, 12.5% Cu


and 50% Zn (0.01 total moles) was used in this work. All of the information
related to the preparation and characterization (N2 Adsorption/Desorption, XRD,
225
CHAPTER 6

TPR, TEM and XPS) was published in a previous paper by our group [15]. This tri-
metallic catalyst showed better catalytic performance in the hydrogenation of
CO2 to MeOH than the bi-metallic PdZn/SiC and CuZn/SiC counterparts.

2.2. Catalyst activity measurements

Catalytic tests were carried out in a tubular quartz reactor (45 cm length
and 1 cm diameter). The catalyst, which consisted of pellets that were 3mm in
length and 1mm in diameter, was placed on a fritted quartz plate located at the
end of the reactor. The amount of catalyst used in the experiments was 0.8 g.

The temperature of the catalyst was measured with a K-type


thermocouple (Thermocoax) placed inside the inner quartz tube. The entire
reactor was placed in a furnace (Lenton) equipped with a
temperatureprogramed system. Reaction gases were Praxair certified standards
of CO2 (99.999% purity), H2 (99.999% purity), N2 (99.999% purity), CH3OH
(0.5% diluted in N2) and CO (99.999% purity). The gas flows were controlled
by a set of calibrated mass flowmeters (Brooks 5850 E and 5850 S).

The hydrogenation of CO2 was carried out at atmospheric pressure. The


feed composition was different in each experiment, keeping the total flow at 110
Nml min−1, using N2 as a balance (GHSV=6600 h−1). The specific compositions
used in all experiments are listed in Table 6.1. Experiments with CO2 and H2 in
the feed stream were carried out at 473, 498, 523 and 548 K. The rest of the
experiments, in which CO or MeOH were added to the feed, were evaluated at
498 and 523K.

Gas effluents were monitored with a micro gas chromatograph (Varian


CP-4900) fitted with a PoraPLOT Q column and a molecular sieve column, each
of which was connected to a thermal conductivity detector (TCD).

226
CHAPTER 6

Table 6.1. List of experiments (species volumen flow in Nml min-1)

Experiment CO2/H2 CO2 H2 CH3OH CO N2


ratio
1 0.30 25 82.5 0 0 2.5
2 0.30 25 82.5 0 0.5 2
3 0.33 25 75 3 0 7
4 0.33 25 75 0 0.5 9.5
5 0.33 25 75 0 0 10
6 0.19 17.5 90 0 0.5 2
7 0.19 17.5 90 0 0 2.5
8 0.21 17.5 82.5 3 0 7
9 0.21 17.5 82.5 0 0.5 9.5
10 0.21 17.5 82.5 0 0 10
11 0.23 17.5 75 3 0 14.5
12 0.23 17.5 75 0 0.5 17
13 0.23 17.5 75 0 0 17.5
14 0.11 10 90 3 0 7
15 0.11 10 90 0 0.5 9.5
16 0.11 10 90 0 0 10
17 0.12 10 82.5 3 0 14.5
18 0.12 10 82.5 0 0.5 17
19 0.12 10 82.5 0 0 17.5
20 0.13 10 75 3 0 22
Extra experiments (for comparative purposes)
21 0.25 20 80 0 0 10
22 0.15 13 85 0 0 12

Catalytic parameters were calculated as follows (Eq. (6.3)–(6.5)):

Fi -F0i
Formation ratei (μmol∙min-1 ∙g-1 )= (6.3)
W

Fi -F0i
Selectivityi (%)= F0 x 100 (6.4)
CO2 -FCO2

F0CO2 -FCO2
CO2 conversion (%)= x 100 (6.5)
F0CO2

where Fi and Fºi represent the outlet and inlet molar flow (μmol∙min–1) of
the i component (CH3OH or CO) and W refers to the catalyst mass (g).
227
CHAPTER 6

3. Results and discussion

3.1. Catalyst activity

3.1.1. Influence of the temperature

The catalytic results for all the experiments in which only CO2 and H2 were
fed are represented in Fig. 6.1. The results are arranged from highest to lowest
CO2/H2 ratio in the feed. The MeOH synthesis reaction is exothermic (Eq. (6.1))
and, as a consequence, MeOH production is favored at low temperatures, as can
be seen in the selectivity graph (Fig. 6.1b; note that the temperature axis is in
reverse order). However, due to the high stability of the CO2 molecule, high
temperatures are required to activate it (Fig. 6.1a). For this reason, the methanol
formation rate curve is bell-shaped with a maximum at 523 K (Fig. 6.1c). The
MeOH selectivity decreased at higher temperatures as a consequence of the
formation of CO (Fig. 6.1d), the formation of which by the RWGS is favored at
high temperatures (Eq. (6.2)).

3.1.2. Influence of the CO2/H2 ratio

The influence of the CO2/H2 ratio on the CO2 conversion is represented


in Fig. 6.1a. Three different regions, which are related to CO2 volume flows of
25 Nml min−1 (Exp. 5 and 1), 17.5 Nml min−1 (Exp. 13, 10 and 7) and 10 Nml
min−1 (Exp. 19 and 16), can be identified. In each region, the variation of H2
volume flow did not affect CO2 conversion. On the other hand, the differences in
CO2 conversion between the regions are significant: the lower the CO2 volume
flow in the feed stream, the higher the CO2 conversion, which can be explained
considering thermodynamics of reaction (1): by using excess H2 the equilibrium is
pushed to the products side.

228
CHAPTER 6

a) 473 K
3.0 498 K
523 K
2.5 548 K

CO2 conversion (%)


2.0

1.5

1.0

0.5

0.0
0 1
0.33 2
0.30 3
0.23 4 5
0.19 6 7
0.11 8
0.21 0.12
Ratio CO2/H2
100
b) 473 K
90 498 K
80 523 K
548 K

CH3OH selectivity (%)


70
60
50
40
30
20
10
0
0 1
0.33 2
0.30 3
0.23 4 5
0.19 6 7
0.11 8
0.21 0.12
Ratio CO2/H2
7
c) 473 K
CH3OH formation rate (mol min-1 g-1)

6 498 K
523 K
5 548 K

0
0 1
0.33 2
0.30 3
0.23 4 5
0.19 6 7
0.11 8
0.21 0.12
Ratio CO2/H2

20 d) 473 K
CO formation rate (mol min-1 g-1)

498 K
18
523 K
16 548 K
14
12
10
8
6
4
2
0
0 1
0.33 2
0.30 3
0.23 4 5
0.19 6 7
0.11 8
0.21 0.12
Ratio CO2/H2

Figure 6.1. Catalytic results of all experiments with only CO2 and H2 in the feed: a) CO2 conversion, b)
CH3OH selectivity, c) CH3OH formation rate and d) CO formation rate.
229
CHAPTER 6

The influence of the CO2/H2 ratio on methanol selectivity is shown in Fig.


6.1b. In general, higher MeOH selectivity (Fig. 6.1b) and lower CO production
(Fig. 6.1d) were obtained at lower CO2/H2 ratios, as also observed in a previous
study [16]. For the same CO2 volume flow in the regions mentioned above, a
higher H2 volume flow gives rise to higher methanol selectivity. This trend is also
consistent with Fig. 1c: the highest methanol formation rates were obtained in the
experiments in which the highest volume flow of H2 was used at constant CO2
values. According to stoichiometry of reactions (6.1) and (6.2), a higher hydrogen
concentration favored methanol formation against RWGS.

The selection of the appropriate CO2/H2 ratio to carry out the CO2
hydrogenation reaction depends on the requirements of the system in which the
MeOH synthesis is carried out. For instance, if high conversions are needed, it
would be better to work at high hydrogen concentrations, which is required in
electrochemical reactors where conversion is usually poor [16]. A decrease in the
CO2/H2 ratio leads to an increase in the CO2 conversion and MeOH selectivity.
It should be considered that the use of hydrogen limits the economic benefits of
the methanol production [3, 5] and, given this limitation, higher ratios (e.g. 1/3,
1/4) are used [17-19].

3.1.3. Influence of the presence of CO and CH3OH in the feed

Experiments with small amounts of MeOH and CO in the feed stream were
carried out in order to evaluate the role of the products in the reaction mechanism.
The catalytic results of the experiments with only CO2 and H2 in the feed stream
at 498 and 523 K are represented in Fig. 6.2 along with those in which CO or
MeOH was also included. Although marked differences in the results are not
observed, some interesting points should be highlighted. As expected, the
presence of both products led to a decrease in the CO2 conversion (Fig. 6.2a)
since they shifted the equilibrium reactions (Eqs. (6.1) and (6.2)) to the left-hand
side.

230
CHAPTER 6

2.0 a) Only CO2 & H2


With CO
With CH3OH

CO2 conversion (%)


1.5

1.0

0.5

0.0 13 1211
0 1 10 92 8 191817
3 131211
4 10 95 8 191817
6 Exp.number
7

498 K 523 K

70 b) Only CO2 & H2


With CO
60
With CH3OH
CH3OH selectivity (%) 50
40
30
20
10
0 13 1211
0 1 10 92 8 191817
3 131211
4 10 95 8 191817
6 Exp.number
7

498 K 523 K
5
CH3OH formation rate (mol min-1 g-1)

c) Only CO2 & H2


With CO
4 With CH3OH

0 13 1211
0 1 10 92 8 191817
3 131211
4 10 95 8 191817
6 Exp.number
7

498 K 523 K
9
CO formation rate (mol min-1 g-1)

d) Only CO2 & H2


8
With CO
7 With CH3OH
6
5
4
3
2
1
0 13 1211
0 1 10 92 8 191817
3 131211
4 10 95 8 191817
6 Exp.number
7

498 K 523 K

Figure 6.2. Influence of the presence of CO and CH3OH in the feed on the catalytic results: a) CO2
conversion, b) CH3OH selectivity, c) CH3OH formation rate and d) CO formation rate.

231
CHAPTER 6

Regarding methanol selectivity (Fig. 6.2b), a general decrease in the


experiments with CO and MeOH was also observed – except for experiments
12 and 18, which showed a slight increase. Finally, the effect on the
corresponding formation rates (Fig. 6.2c and d) warrants a special mention. On
the one hand, the aforementioned equilibrium shift explained the fact that, when
MeOH or CO were introduced in the feed, the corresponding formation rates
decreased. On the other hand, the addition of MeOH did not practically affect
the CO formation rate, whereas the addition of CO led to a decrease in the
formation of MeOH. This behavior can be explained by considering the species
adsorption equilibria. In agreement with the results shown in Fig. 6.2c and d,
MeOH adsorption on the catalyst active sites was practically negligible whereas
CO adsorption was considerable, since it also had a negative effect on the MeOH
formation rate. This adsorption effect was confirmed by the fact that, in
experiments with CO, higher reaction temperatures led to a less marked decrease
in the MeOH formation rate, as a consequence of the weaker CO adsorption at
higher reaction temperatures.

3.2. Kinetic model

3.2.1. Evaluation of diffusion limitations

Potential external diffusion limitations were evaluated using a


semiempirical method based on the following criterion:

Pi -Pis
| | <0.1 (6.6)
Pi

where Pi is the partial pressure of the component i in the feed and Pis is
the corresponding surface concentration. Pis was calculated using an iterative
procedure proposed by Froment and Bischoff [20]: firstly, Pis was supposed to be
equal to Pi, then this value was used to estimate the mass transfer coefficient (kg),
which finally was used to calculate the new value of Pis. The procedure was
repeated until no change of Pis was observed. As an example, the value obtained
232
CHAPTER 6

in Eq. (6.6) using the results for the CO2 conversion rate at 523 K was 0.0001,
which means that external diffusion limitations could be ruled out.

Regarding internal diffusion limitations, the Weisz–Prater criterion was


used. The Weisz–Prater number, NW-P, should meet the following inequation:

r ρC R2C
NW-P = ≤1 (6.7)
CS Deff

where r is the CO2 initial conversion rate, ρc the catalyst density (800
kg/m3), RC the catalyst pellet volume/surface area ratio, CS the CO2 surface
concentration (equal to the bulk concentration since external diffusion limitations
were not observed) and Deff the catalyst effective diffusivity. As an example, the
results for the reaction at 523 K led to an NW-P value of 7 x 10−16, which shows
that internal diffusion limitations can be ruled out.

Finally, a stability test was carried out in order to evaluate potential


catalyst deactivation. In this test four reaction cycles between 423 and 523 K
were performed and the results were compared, using a mixture of 100 Nml
min−1 of CO2 and H2 with a volume ratio 1/9 (Fig. 6.3). Each temperature was
kept until steady state was attained. Once a cycle was completed, the catalyst
was kept under N2 atmosphere at 423 K overnight, prior to carry out the next
one. As can be observed, all cycles overlapped and this indicates that the catalyst
was stable under the conditions studied in this work and catalyst deactivation was
not included in the kinetic models proposed thereafter.

3.2.2. Kinetic analysis

Firstly, it was evaluated whether the adsorption, chemical surface reaction


or desorption phenomena was the rate determining step (r.d.s.) of the process.
This was achieved by carrying out reactions with different CO2/H2 ratios (v/v)
and evaluating the influence on the initial reaction rate (Fig. 6.4). It was observed
that the reaction rate reached a maximum at a CO2/H2≈0.36 and then showed
233
CHAPTER 6

a slight decrease. According to the literature [21], only those reactions in which
the chemical reaction is the r.d.s. can show this profile, where the reactants ratio
play a crucial role and a maximum is observed. If other phenomenon (adsorption
or desorption) had been the r.d.s., this parameter would have been irrelevant
and a straight line would have been observed.

7
Cycle 1
Formation rate (mol min g )
-1

6 Cycle 2
-1

Cycle 3
5
Cycle 4
4
3
2
1
0 CO MeOH
425 450 475 500 525 425 450 475 500 525
Temperature (K)

Figure 6.3. Stability test with four reaction cycles from 423 to 523 K.

16

15
rCO2 (mol min g )
-1
-1

14

13

12
0.1 0.2 0.3 0.4 0.5 0.6 0.7
CO2/H2 (p/p)

Figure 6.4. Influence of the CO2/H2 ratio on the CO2 reaction rate.

234
CHAPTER 6

The synthesis of MeOH by CO2 hydrogenation involves three main


reactions: Eq. (1), Eq. (2) and CO hydrogenation (Eq. (6.8)):

CO + 2H2 ⇆ CH3 OH ∆H25°C = –90.5 kJ∙mol-1 (6.8)

It was initially assumed that all reactions occur simultaneously on the


surface of the catalyst. The mechanism reported by Lim et al. [13] was used to
propose the reaction rate equations (Table 6.2), with reactions marked with
(r.d.s.) being the corresponding rate determining steps. Regarding the
hydrogenation of CO2, it has also been reported [22] that the hydrogenation of
formate is the most likely r.d.s. Therefore, the following reaction rate equations
were proposed:

CO2 hydrogenation:
PMeOH PH2O
(PCO2 P3
H2 -KeqMeOH-CO2 )
k'MeOH-CO2 KCO2,s1 KH2,s2
P2
H2
rMeOH-CO2 = (6.9)
DMeOH-CO2

Reverse WGS:
P P
(PCO2 PH2 KCO H2O )
eqRWGS
k'RWGS KCO2,s3 K0.5
H2,s2
P0.5
H2
rRWGS = (6.10)
DRWGS

CO hydrogenation:
PMeOH PH2O
(PCO2 P3
H2 KeqMeOH-CO )
k'MeOH-CO KCO,s1 KH2,s2
P2H2
rMeOH-CO = (6.11)
DMeOH-CO

In this work, three different Langmuir–Hinshelwood models were


proposed: in the first one, reactants were supposed to adsorb competitively on
the catalyst active sites. The second one considered two adsorption sites: Pd-
based nanoparticles that adsorbed carbonaceous species (CO, CH3OH and
CO2), and ZnO, on which H2 and H2O were adsorbed. Finally, a three-site kinetic
235
CHAPTER 6

model was proposed and this is explained below. The difference between these
models was related to the adsorption term of the reaction rate equations, which
will be explained in the corresponding section for each model. The rate and
adsorption constants were given by the orthogonalized Arrhenius and Van't Hoff
equations, respectively:

̅̅̅x ∙exp(- Eax ⁄R ∙ 1⁄θ )


k'x =k' (6.12)

̅̅̅
k'x =k'x,∞ ∙ exp(- Eax ⁄R ∙ 1⁄T̅) (6.13)

Ki =K̅i ∙exp( -∆H0i ⁄R ∙ 1⁄θ) (6.14)

K̅i =Ki,∞ ∙exp( -∆H0i ⁄R ∙ 1⁄T̅ ) (6.15)

1⁄θ = 1⁄T - 1⁄T̅ (6.16)

The equilibrium constants of the reactions involved were estimated using


Aspen Hysys:
6610
lnKeqMeOH-CO2 = -23.462 (6.17)
T(K)

-4762.4
lnKeqRWGS = +4.539 (6.18)
T(K)

KeqMeOH-CO2
KeqMeOH-CO = (6.19)
KeqRWGS

The flow of the different species through the catalyst bed was modeled
with a pseudohomogeneous, one-dimensional plug flow model, since mass transfer
resistance was not observed and a very large reactor L/D ratio was used. In this
way, the following expression for the axial flow profiles through the reactor (rx)
for each of the species i could be used:

rx = dFPi ⁄dW (6.20)

where FPi is the species molar flow and w the catalyst mass. Therefore,
the mass balance equations for all the species are given by:
236
CHAPTER 6

dFPCO2⁄dW =-rMeOH-CO2 -rRWGS (6.21)

dFPH2⁄dW =-3 rMeOH-CO2 -rRWGS -2 rMeOH-CO (6.22)

dFPCH3OH ⁄dW =rMeOH-CO2 +rMeOH-CO (6.23)

dFPCO ⁄dw =rRWGS -rMeOH-CO (6.24)

dFPH2O ⁄dw =rMeOH-CO2 +rRWGS (6.25)

The parameter estimation was performed using an iterative nonlinear


regression procedure. For given initial parameters, the Matlab subroutine ode15s
was used to solve the system of ordinary differential equations. The optimum
kinetic parameters were then determined by minimizing the objective function,
using the Matlab subroutine lsqnonlin (Levenberg–Marquardt). The Weighted
Residual Sum of Squares (WRSS) between the theoretical and the experimental
partial pressure data was set as the objective function to be minimized:

WRSS= ∑Nspecies
i=1 wi ∙ ∑Ndata
j=1 (Pi,j,exp -Pi,j,theo )
2
(6.26)

The weighting factor of each component was calculated by the inverse


of their corresponding residuals (si2):

wi = 1⁄√s2i ⁄(n-p ) (6.27)

where n and p are the number of data and parameters, respectively. The
weighting factors were also calculated using an iterative procedure. A first
estimation of the parameters was performed using wi =1. The corresponding wi
were then calculated and used for the next parameter estimation. This procedure
was repeated until significant changes in any of the wi values were not observed.
All kinetic models were subjected to a rigorous statistical analysis. Firstly, the
quality of the fit was evaluated by calculating the regression coefficient:
2 2
2 ∑Ndata ̅̅̅̅̅̅
j=1 (Pj,exp -P
Ndata
exp ) - ∑j=1 (Pj,exp -Pj,theo)
r = ̅̅̅̅̅̅ 2 (6.28)
∑Ndata
j=1 (Pj,exp -Pexp )

237
CHAPTER 6

Table 6.2. Reaction mechanism of the CO2 hydrogenation [10].

Species adsorption
CO2 + s1 ⇆ CO2 s1 / CO2 + s3 ⇆ CO2 s3
H2 + 2s2 ⇆ 2H s2
CH3OH s1 ⇆ CH3OH + s1
CO s3 ⇆ CO + s3 / CO + s1 ⇆ CO s1
H2O s2 ⇆ H2O + s2
MeOH synthesis from CO2 hydrogenation (MeOH-CO2)
CO2 s1 + H s2 ⇆ HCO2 s1 + s2
HCO2 s1 + H s2 ⇆ H2CO2 s1 + s2 (r.d.s.)
H2CO2 s1 + H s2 ⇆ H3CO2 s1 + s2
H3CO2 s1 + H s2 ⇆ H2CO s1 + H2O s2
H2CO s1 + H s2 ⇆ H3CO s1 + s2
H3CO s1 + H s2 ⇆ CH3OH s1 + s2
Reverse Water-Gas-Shift (RWGS)
CO2 s3 + H s2 ⇆ HCO2 s3 + s2 (r.d.s.)
HCO2 s3 + H s2 ⇆ CO s3 + H2O s2
MeOH synthesis from CO hydrogenation (MeOH-CO)
CO s1 + H s2 ⇆ HCO s1 + s2
HCO s1 + H s2 ⇆ H2CO s1 + s2
H2CO s1 + H s2 ⇆ H3CO s1 + s2
H3CO s1 + H s2 ⇆ CH3OH s1 + s2 (r.d.s.)

The statistical significance of the parameters was then evaluated using the
t-test, by the procedure described previously [23], with a confidence interval (α)
of 95%. Finally, due to the potential correlation issues related to Langmuir–
Hinshelwood models, the correlation matrix pij was evaluated in each kinetic

238
CHAPTER 6

model. The results obtained in all kinetic models (not shown) revealed that there
were no correlation issues.

3.2.3. Model 1: competitive adsorption

Preliminary results revealed that, on the one hand, the adsorption


constants of MeOH, CO and H2O were negligible because of the low
concentration of these species in all experiments. Although CO adsorption was
expected to affect the kinetic model (as observed experimentally), initial
attempts to calculate its adsorption constants (KCO and ΔHCO) led to unusual
values, which indeed did not meet the associated constraints (Ki > 0 and ΔHi <
0). On the other hand, the kinetic parameters for the synthesis of MeOH from CO
(Eq. (6.8)) were also negligible because of the very low concentration of CO, as
mentioned above, and this reaction was therefore not considered. It is important
to note that this behavior was also observed for the other two kinetic models, so
that neither the adsorption of the species mentioned above nor MeOH from CO
hydrogenation were considered.

In this model, the catalyst was supposed to have one type of active site
and, as a consequence, the species involved in the process were supposed to
adsorb competitively on these sites. The model proposed by Skrzypek et al. [10]
is the best example in the literature in this respect. Therefore, the adsorption terms
could be written as follows according to the preliminary observations:

DMeOH-CO2 =DRWGS =Dx (6.29)

s1 =s2 =s3 =s (6.30)

Dx =(1+K CO2,s PCO2 +√K H2,s PH2 )2 (6.31)

In this model, 8 parameters [2 kinetic constants at the reference


temperature (225 °C), 2 activation energies, 2 adsorption constants at the
reference temperature and 2 adsorption enthalpies] were estimated. The parity

239
CHAPTER 6

plot for the prediction of the species partial pressures is shown in Fig. 6.5a. The
parameter estimation results are listed in Table 6.3 along with the corresponding
t-test values.

3.2.4. Model 2: two-site mechanism

Most of the kinetic models for MeOH synthesis reported in the literature
using Cu/ZnO-based catalysts [11, 13, 24] concern the two-site adsorption
mechanism. In this respect, ZnO was believed to adsorb H2 and H2O, whereas
carbonaceous species were adsorbed on Cu. In this study, a similar two-site
adsorption mechanism was proposed and the adsorption terms are as follows:

DMeOH-CO2 =DRWGS =Dx (6.32)

s1 =s3 =s (6.33)

Dx =(1+KCO2,s PCO2 ) (1+√KH2,s2 PH2 ) (6.34)

As in the previous model, 8 parameters were estimated in this model. The


corresponding parity plot is shown in Fig. 6.5b and the parameter estimation
results are listed in Table 6.3.

3.2.5. Model 3: three-site mechanism

In previous work by our group [15] it was concluded that the Pd-Cu-Zn
catalyst, which showed the best results, had three different active sites: PdZn
alloy, on which MeOH synthesis from CO2 hydrogenation occurred, PdCu alloy,
which was prone to catalyze the RWGS reaction, and ZnO sites, which are
involved in both reactions. In line with this conclusion, a three-site adsorption
mechanism was proposed in this work and it was considered that carbonaceous
molecules (in this case CO2) adsorbed differently on PdZn or PdCu, whereas ZnO
interacted with H2. Therefore, the adsorption terms for this model can be written:

DMeOH-CO2 =(1+KCO2,s1 PCO2 ) (1+√KH2,s2 PH2 ) (6.35)

240
CHAPTER 6

a)
0.0008
0.75
0.0005

0.60 0.0002

Ppredicted (bar)
0.0002 0.0005 0.0008
0.45

0.30

0.15

0.15 0.30 0.45 0.60 0.75


Pexperimental (bar)
b)
0.0008
0.75
0.0005

0.60 0.0002
Ppredicted (bar)

0.0002 0.0005 0.0008


0.45

0.30

0.15

0.15 0.30 0.45 0.60 0.75


Pexperimental (bar)
c)
0.0008
0.75
0.0005

0.60 0.0002
Ppredicted (bar)

0.0002 0.0005 0.0008


0.45

0.30

0.15

0.15 0.30 0.45 0.60 0.75


Pexperimental (bar)

Figure 6.5. Parity plot of (a) competitive adsorption (b) two-site mechanism and (c) three site
mechanism.

241
Table 6.3. Parameter estimation results.
Parameter Competitive adsorption Two-sites Three-sites
CHAPTER 6

Value t-Test Value t-Test Value t-Test

̅̅̅̅̅̅̅̅
ln(k'MeOH +K̅̅̅̅̅̅̅̅̅ ̅̅̅̅̅̅̅
CO2,s1 +K H2,s2 )
-2.79 ± 0.07 42.79 -3.12 ± 0.06 53.00 -3.08 ± 0.15 20.90

(EaMeOH +∆HCO2,s1 +∆HH2,s2 ⁄(R∙T̅) -0.72 ± 0.05 14.03 -1.20 ± 0.29 4.19 -0.61 ± 0.08 7.71

' ̅̅̅̅̅̅̅̅̅ ̅̅̅̅ -3.26 ± 0.04 86.20 -3.57 ± 0.05 65.97 -3.69 ± 0.07 50.03
ln(k̅̅̅̅̅̅̅̅
RWGS +K CO2,s3 +0.5∙K H2 )

(EaRWGS +∆HCO2,s1 +0.5∙∆HH2,s2 ⁄(R∙T̅) 8.46 ± 0.39 21.69 9.08 ± 0.51 17.72 10.63 ± 0.84 12.69

̅̅̅̅̅̅̅̅̅
ln(K CO2,s1 )
1.98 ± 0.07 28.76 2.91 ± 0.08 35.08 2.96 ± 0.23 13.03

242
∆HCO2,s1 ⁄(R∙T̅) -12.69 ± 0.67 18.85 -14.08 ± 0.66 21.44 -18.18 ± 2.29 7.94

̅̅̅̅̅̅̅
ln(0.5∙K H2,s2 )
-0.22 ± 0.04 5.46 -0.46 ± 0.02 18.71 -0.45 ± 0.09 5.22

0.5∙ ∆HH2,s2 ⁄(R∙T̅) -7.85 ± 0.39 20.18 -8.47 ± 0.95 8.96 -5.28 ± 0.69 7.60

̅̅̅̅̅̅̅̅̅
ln(K CO2,s3 )
̅̅̅̅̅̅̅̅
=ln(K CO2,1 )
- ̅̅̅̅̅̅̅̅
=ln(K CO2,1 )
- 2.72 ± 0.15 18.28

∆HCO2,s3 ⁄(R∙T̅) =∆HCO2,1 ⁄(R∙T̅) - =∆HCO2,1 ⁄(R∙T̅) - -13.16 ± 1.70 7.76


CHAPTER 6

DRWGS =(1+KCO2,s3 PCO2 ) (1+√KH2,s2 PH2 ) (6.36)

In this model there were three adsorption species (CO2 at two sites and
H2) and therefore 10 parameters were estimated. The corresponding parity plot
is depicted in Figure 6.5c and the parameter estimation results are listed in Table
3.

4. Discussion

It can be observed from the results in Table 6.3 that all kinetic models
were developed using lumped parameters, with the aim of minimizing the
potential correlation issues associated with the proposed equations (for example,
the use of the full Arrhenius and Van't Hoff equations). The results of the t-test
revealed that all parameters were statistically meaningful, as all t values were
higher than the corresponding threshold (1.96, see Table 6.3). This fact - together
with the lack of difference between the parity plots (Fig. 6.5) - made it difficult
at first glance to carry out the model discrimination.

As a consequence of the above, the kinetic model results should be


compared in more detail. The variables that were compared to select the best
kinetic model are listed in Table 6.4. If we consider the first column, it can be seen
that the regression coefficient for all models (r2, Eq. (6.28)) was 0.9998, which
confirmed (i) that all of the models predicted the experimental results well and
(ii) that it was not possible to discriminate any model according to this variable.
The unweighted residual sum of squares showed slight differences between the
proposed kinetic models. In this regard, the three-site model led to the lowest
value, which confirmed that this model predicted the experimental results better
than the other ones. It is important to note that the WRSS values (Eq. (6.26)) could
not be used to compare the kinetic models since the weighting factors were not
the same in each case.

243
CHAPTER 6

Table 6.4. Model discrimination parameters.

Hydrogen: parameter constrain


(𝑃𝑖,𝑗,𝑒𝑥𝑝
Kinetic model r2 (eq. 28) (eq. 35)
− 𝑃𝑖,𝑗,𝑡ℎ𝑒𝑜 )2 0 0 ⁄
−∆𝑆𝑎𝑑𝑠 ⁄𝑅 𝑆𝑔𝑎𝑠 𝑅
Competitive
0.9998 0.003488 16.13 15.72
adsorption
Two-sites 0.9998 0.003486 17.86 15.72
Three-sites 0.9998 0.003475 11.46 15.72

Finally, as proposed by Graaf et al. [11], the kinetic parameters should


meet the following constraints: kx > 0, Eax > 0, Ki > 0, ΔH°i < 0, and the
following one, which is included in Table 6.4:

0< -∆S0ads ⁄R < S0gas ⁄R (6.37)

where

K=exp (-∆S0ads ⁄R )∙exp( -∆H0ads ⁄R∙T) (6.38)

It can be seen from the results in Table 6.4 that the corresponding H2
parameter only meets this constraint in the three-site kinetic model (CO2
parameters meet it in all models). On considering these findings, the three-site
kinetic model was chosen as the one that provided the best prediction of the
experimental results, since it provides the least unweighted residual sum of
squares and all of its parameters satisfied their corresponding constraints.

Once the model discrimination had been performed, the kinetic


parameters were calculated from the corresponding lumped ones (Table 6.5).
These values were also compared with those of some other models reported in
the literature. It is important to note that, since each model was developed for
different mechanisms, reaction conditions and catalysts, it was practically
impossible to make direct comparisons, especially in terms of kinetic and
adsorption constants.

244
Table 6.5. Kinetic parameters of three-site model and comparison with the literature.

Parameter This work Graaf (1988) Skrzypek (1991) Park (2014) Units

k'MeOH (498K) 2.4∙10-5 6.3∙10-5 a 5.2∙10-4 4.9∙10-3 c mol∙g-1∙min-1

EaMeOH 116.5 65.2 104.7 68.3 kJ∙mol-1

k'RWGS(498K) 8.4∙10-6 8.3∙10-5 a2 4.3∙10-4 6.1∙10-4 c2 mol∙g-1∙min-1

EaRWGS 120.4 123.4 104.7 126.6 kJ∙mol-1

KCO2,s1 (498K) 19.2 1.2 0.4 - bar-1

245
ΔHCO2,s1 -75.3 -67.4 -75.4 - kJ∙mol-1

KH2,s2(498K) 0.4 3.8b 0.1 - bar-1

ΔHH2,s2 -43.7 -104.5b -75.4 - kJ∙mol-1

KCO2,s3 (498K) 15.2 - - - bar-1

ΔHCO2,s3 -54.5 - - - kJ∙mol-1

a [k'
MeOH (498K)]: mol∙s-1∙kg-1∙bar-1, [k'RWGS (498K)]: mol∙s-1∙kg-1∙bar-1/2
b Refers to H2O/H21/2
c [k'
MeOH (498K)]: mol∙s-1∙kg-1∙bar-1.5, [k'RWGS (498K)]: mol∙s-1∙kg-1∙bar-1
CHAPTER 6
CHAPTER 6

1.5
a) Predicted
Experimental

CO2 conversion (%)


1.0

0.5

0.0
0 488 K
1 513 K
2 3
488 K 513 K
4 5

21 Exp.number 22
100
b) Predicted
Experimental
80
CH3OH selectivity (%)

60

40

20

0
0 488 1K 513 K
2 3
488 K 4
513 K 5

21 Exp.number 22
CH3OH formation rate (mol min-1 g-1)

5
c) Predicted
Experimental
4

0
0 1
488 K 2
513 K 3 4 5
488 K 513 K

21 Exp.number 22
7
d)
CO formation rate (mol min-1 g-1)

Predicted
6 Experimental

0
0 4881K 513 2K 4883K 513 4K 5

21 Exp.number 22

Figure 6.6. Comparison between experimental and predicted catalytic results in (a) CO 2 conversion,
(b) CH3OH selectivity, (c) CH3OH formation rate and (d) CO formation rate.
246
CHAPTER 6

However, comparison of the activation energies and adsorption


enthalpies was easier. On the one hand, the EaMeOH in this work was higher than
those reported in the literature, especially when compared to the values of Graaf
[11] and Park [12], whereas all EaRWGS values were similar. The ΔHCO2 values
were all similar to each other. Generally, the estimated parameters obtained in
this work were of the same order of magnitude as those reported in the literature.

Finally, the quality of the model was checked by using it to predict the
species partial pressures under different experimental conditions that were not
used in the parameter fitting procedure. These predicted values are shown in Fig.
6.6 along with the experimental values for the sake of comparison. It was
observed that the proposed kinetic model predicted the experimental results well
and the only significant deviations were observed in the CO counterparts.

5. Conclusions

The following conclusions can be drawn from the work described above:

- The sensitivity study revealed that there was a temperature at which the
maximum MeOH formation rate was attained and this is due to the exothermic
nature of the reaction. Moreover, higher MeOH selectivity was obtained when
the CO2/H2 ratio was lower. Finally, the presence of MeOH or CO in the feed
stream did not have an appreciable effect on the catalytic performance beyond
shifting the corresponding reaction equilibria.

- A proper kinetic analysis of the CO2 hydrogenation to MeOH was carried out.
Three Langmuir–Hinshelwood (LH) kinetic models, in which the nature of the active
sites was varied, were successfully proposed.

- All kinetic models predicted the experimental results well and wer statistically
meaningful. Among them, the three-site LH kinetic model provided the least
unweighted residual sum of squares and met all the established constraints. As a
consequence, this was selected as the best kinetic model.

247
CHAPTER 6

- This model provided a good prediction of the experimental results obtained in


extra experiments that were carried out under different conditions.

- The kinetic model proposed in this work is consistent with the catalytic
observations reported previously.

References

[1] Rahman, F. A.; Aziz, M. M. A.; Saidur, R.; Bakar, W. A. W. A.; Hainin, M. R.;
Putrajaya, R.; Hassan, N. A. Pollution to solution: capture and sequestration of
carbon dioxide (CO2) and its utilization as a renewable energy source for a
sustainable future. Renewable and Sustainable Energy Reviews, 71 (2017) 112-
126.

[2] Saeidi, S.; Amin, N. A. S.; Rahimpour, M.R. Hydrogenation of CO2 to value-
added products - a review and potential future developments. Journal of CO2
Utilization, 5 (2014) 66-81.

[3] Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M.; Joshi, J. B. Catalytic carbon
dioxide hydrogenation to methanol: a review of recent studies. Chemical
Engineering Research and Design, 92 (2014) 2557-2567.

[4] Karim, N. A.; Kamarudin, S. K.; Loh, K. S. Performance of a novel non-platinum


cathode catalyst for direct methanol fuel cells. Energy Conversion and
Management, 145 (2017) 293–307.

[5] Olah, G. A. Beyond oil and gas: the methanol economy. Angewandte Chemie
International Edition, 44 (2005) 2636-2639.

[6] Wisaijorn, W.; Arporn, Y. P.; Marin, P.; Ordóñez, S.; Assabumrungrat, S.;
Praserthdam, P.; Saebea, D.; Soisuwan, S. Reduction of carbon dioxide via
catalytic hydrogenation over copper-based catalysts modified by oyster shell-
derived calcium oxide. Journal of Environmental Chemical Engineering, 5 (2017)
3115-3121.

248
CHAPTER 6

[7] Zhao, F.; Gong, M.; Cao, K.; Zhang, Y.; Li, J.; Chen, R. Atomic layer deposition
of Ni on Cu nanoparticles for methanol synthesis from CO2 hydrogenation,
ChemCatChem 9 (2017) 3772-3778.

[8] An, B.; Zhang, J.; Cheng, K.; Ji, P.; Wang, C.; Lin, W. Confinement of Ultrasmall
Cu/ZnOx Nanoparticles in Metal-organic Frameworks for Selective Methanol
Synthesis From Catalytic Hydrogenation of CO2, 139 (2017) 3834-3840.

[10] Skrzypek, J.; Lachowska, M.; Moroz, H. Kinetics of methanol synthesis over
commercial copper/zinc oxide/alumina catalysts. Chemical Engineering Science,
46 (1991) 2809-2813.

[11] Graaf, G. H.; Stamhuis, E. J.; Beenackers, A. A. C. M. Kinetics of low-pressure


methanol synthesis. Chemical Engineering Science, 43 (1988) 3185-3195.

[12] Park, N.; Park, M. J.; Lee, Y. J.; Ha, K. S.; Jun, K. W. Kinetic modeling of
methanol synthesis over commercial catalysts based on three-site adsorption, Fuel
Processing Technology, 125 (2014) 139-147.

[13] Lim, H. W.; Park, M. J.; Kang, S. H.; Chae, H. J.; Bae, J. W.; Jun, K. W.
Modeling of the kinetics for methanol synthesis using Cu/ZnO/Al2O3/ZrO2
catalyst: influence of carbon dioxide during hydrogenation. Industrial Engineering
Chemistry Research, 48 (2009) 10448-10455.

[14] Chiavassa, D. L.; Collins, S. E.; Bonivardi, A. L.; Baltanás, M. A. Methanol


synthesis from CO2/H2 using Ga2O3-Pd/silica catalysts: kinetic modeling,
Chemical Engineering Journal, 150 (2009) 204-212.

[15] Díez-Ramírez, J.; Díaz, J. A.; Sánchez, P.; Dorado, F. Optimization of the
Pd/Cu ratio in Pd-Cu-Zn/SiC catalysts for the CO2 hydrogenation to methanol at
atmospheric pressure. Journal of CO2 Utilization, 22 (2017) 71-80.

[16] Díez-Ramírez, J.; Sánchez, P.; Valverde, J. L.; Dorado, F. Electrochemical


promotion and characterization of PdZn alloy catalysts with K and Na ionic

249
CHAPTER 6

conductors for pure gaseous CO2 hydrogenation. Journal of CO2 Utilization, 16


(2016) 375-383.

[17] Choi, Y.; Futagami, K.; Fujitani, T.; Nakamura, J. The difference in the active
sites for CO2 and CO hydrogenations on Cu/ZnO-based methanol synthesis
catalysts. Catalysis Letters, 73 (2001) 27-31.

[18] Fujitani, T.; Saito, M.; Kanai, Y.; Kakumoto, T.; Watanabe, T.; Nakamura, J.;
Uchijima, T. The role of metal oxides in promoting a copper catalyst for methanol
synthesis. Catalysis Letters, 25 (1994) 271-276.

[19] Joo, O. S.; Jung, K. D.; Han, S. H.; Uhm, S. J.; Lee, D. K.; Ihm, S. K. Migration
and reduction of formate to form methanol on Cu/ZnO catalysts. Applied
Catalysis A, 135 (1996) 273-286.

[20] Froment, G. F.; Bischoff, K. B.; De Wilde, J. Chemical Reactor Analysis and
Design, Wiley, New York, 1990.

[21] Chapter 9, Heterogeneous catalysis, Comprehensive Chemical Kinetics, 40


2004, pp. 273-308.

[22] Kunkes, E.L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M.
Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2O3: is there a common
intermediate or not? Journal of Catalysis, 328 (2015) 43-48.

[23] García-Vargas, J. M.; Valverde, J. L.; Díez, J.; Dorado, F.; Sánchez, P.
Catalytic and kinetic analysis of the methane tri-reforming over a Ni-Mg/β-SiC
catalyst. International Journal of Hydrogen Energy, 40 (2015) 8677-8687.

[24] Peter, M.; Fichtl, M. B.; Ruland, H.; Kaluza, S.; Muhler, M.; Hinrichsen, O.
Detailed kinetic modeling of methanol synthesis over a ternary copper catalyst.
Chemical Engineering Journal, 203 (2012) 480–491.

250
CHAPTER 7
ELECTROCHEMICAL PROMOTION AND
CHARACTERIZATION OF PDZN ALLOY
CATALYSTS WITH K AND NA IONIC
CONDUCTORS FOR PURE GASEOUS
CO2 HYDROGENATION
CHAPTER 7

Abstract
A PdZn alloy has been used for the first time as a catalytic electrochemical film
for CO2 hydrogenation. Three different electrochemical systems were prepared
to study the influence of the Pd/Zn ratio and the presence of potassium and
sodium ions. These catalysts were characterized by XRD, SEM and cyclic
voltammetry and then tested in the hydrogenation of CO2 at three different
H2/CO2 feed ratios (3, 9 and 39) and three temperatures (300, 320 and 340
ºC) with electrochemical promotion by Na+ and K+ ions. CH3OH and CO were
the only two products detected. During the electrochemical promotion, a low
quantity of ions increases the formation rate of CH3OH and a high quantity leads
to the poisoning of the PdZn active sites. The catalysts which have palladium and
the palladium-zinc alloy in their structure show an electrophilic behavior in the CO
formation and when only the alloy is formed in the catalyst, the CO has the same
behavior found for the methanol. Apparent Faradaic efficiency values above
1000 were obtained under the different conditions tested.

CO2 H2 CH3OH CO

PdZn Pd PdZn ZnO

Na+ Na+ Na+ K+ K+ K+ K+ K+ K+


Au Au

253
CHAPTER 7

1. Introduction
The valorization of carbon dioxide to desirable compounds is one of the
targets of the scientific community in general, especially considering the recent
increase in the emission of this pollutant, which is well-known as the most important
greenhouse gas. The carbon dioxide molecule is practically unreactive but it is
unquestionably a major carbon source. Therefore, the challenge for scientists is to
transform CO2 into chemical compounds such as hydrocarbons, alcohols,
aldehydes, etc. [1-4].

Hydrogenation of carbon dioxide is one of the possible ways to valorize


this compound and this reaction continues to be studied in the field of conventional
catalysis [5-7]. Indeed there are several pilot plants in Japan [8, 9] and a
commercial CO2-to-methanol recycling plant in Iceland [10]. The wide range of
products that are allowed in the general equation (Eq. (7.1)) means that high
conversion and high selectivity towards the desired product are the two most
important parameters in this reaction. In this sense, promotion by alkali has been
reported to be an effective approach to control the activity and selectivity of a
catalyst [11]. Although this promotion could be carried out by conventional
catalysis, electrocatalysis allows control of this alkali promotion by varying the
polarization of a catalyst-electrode through the effect of electrochemical
promotion of catalysis (EPOC), also known as non-Faradaic electrochemical
modification of catalytic activity (the NEMCA effect).

xCO2 + (2x – z + y/2) H2 ⇆ CxHyOz + (2 – z) H2O (7.1)

The NEMCA effect was first observed by Stoukides and Vayenas [12].
The effect is based on the controlled migration of promoting ions (i.e., O2-, Na+,
K+ or H+) from an electroactive support, such as β"-Al2O3 (K+ and Na+ conductor),
YSZ (yttrium-stabilized-zirconia,an O2- conductor), SZY (SrZr0.95Yb0.05O3± δ, a H+
conductor), CZY (CaZr0.9In0.1O3-α, a H+ conductor) or BZY (BaZr0.8Y0.2O3± δ, a H+

254
CHAPTER 7

conductor), to the catalytic metal/gas interface. The migration of the ions allows
in situ control of the catalytic behavior of the system by changing the binding
strength of chemisorbed species and reaction intermediates [13, 14].

Very few EPOC studies have been devoted to the reaction discussed here
and the vast majority was previously summarized by our group [15]. A total of
fourteen such papers have been published, including the most recent paper [16],
and this total is a very low compared to other fields of research.

The most commonly used active metals are Cu, Pd, Pt, Ru and Rh supported
on Na-β´´Al2O3 [17], K-β´´Al2O3 [18-21], YSZ [17, 20, 22-27], or SZY [28]. The
most common products are carbon monoxide and methane-except for the works
by Esperanza et al. [18, 19, 25], where alcohols and hydrocarbons were
obtained on working under realistic post-combustion CO2 capture exiting gas
compositions on a bench-scale plant. The novelty of the work described here is
the use of a PdZn alloy as the catalyst-electrode in the electrocatalytic system.
The PdZn alloy has previously been studied in conventional catalysis by our group
[29, 30] and it has also been extensively studied for the steam reforming of
methanol [31-36]. This catalyst showed interesting properties in terms of the
selectivity and activity to methanol. The study reported here is the first in which
the only two products obtained in a pure gaseous hydrogenation of CO2 in EPOC
studies are carbon monoxide and methanol, as described by the following
equations:

CO2 + H2 ⇆ CO+ H2O (2)

CO2 + 3H2 ⇆ CH3OH + H2O (3)

In this work three different electrocatalytic systems (PdZn/K-βAl2O3/Au,


PdZn/K-βAl2O3/Au with a different PdZn molar ratio and PdZn/NaβAl2O3/Au)
were studied by different characterization techniques and with catalytic activity
experiments under electrochemical promotion conditions. The influence of the

255
CHAPTER 7

Pd/Zn molar ratio on the formation of the PdZn alloy or the way in which
promotional ions (K+ or Na+) affect the electrochemical results were studied.
Moreover, an in-depth electrochemical study was carried out with PdZn/K-
βAl2O3/Au using a Pd/Zn molar ratio of 1.

2. Experimental

2.1. Electrochemical cell preparation

Each electrochemical cell consisted of a 19-mm-diameter, 1-mm-thick (K,


Na)- β“Al2O3 (Ionotec) pellet as the solid electrolyte. In each electrochemical cell
the Au counter and reference electrodes were first deposited on one side of the
electrolyte by annealing a gold organometallic paste (Fuel Cell Materials ref-
231001) in two correlative stages, firstly at 300 ºC for 1 h and secondly at 800
ºC for 2 h (heating ramps of 5 ºC/min). The PdZn catalyst films were subsequently
deposited by impregnation. Thus, two solutions of palladium (II) nitrate
[Pd(NO3)2.H2O, Aldrich] and zinc nitrate 6-hydrate [Zn(NO3)2.6H2O, Panreac]
were prepared. The first solution was 0.1 M in both metals (Pd/Zn molar ratio of
1), whereas the second one was 0.1 M in palladium with a Pd/Zn molar ratio of
0.13 (the best molar ratio found in a previous study carried out with conventional
catalysts [29]). In each impregnation, 20 ml of solution were added and then the
sample was calcined at 500 ºC with a heating rate of 5 ºC/min. After all of the
required impregnations (36 for each cell) had been completed, different
quantities of metals were deposited in ~1.5 cm2 areas.

The description and nomenclature of each electrochemical catalyst are


summarized in Table 7.1, along with the different palladium and zinc weights,
particle sizes and active surface area of palladium and PdZn alloy. The active
surface mol were calculated in each case taking into account the total amount of
deposited mol of each metal, the EDX measurements, the particle diameter and
dispersion values for each compound, which were estimated from XRD via the
Scherrer equation. For the calculation of product formation rates, it has been
256
CHAPTER 7

considered that the methanol is preferentially produced on the PdZn active sites
whereas the carbon monoxide is produced on both active sites [30].
Table 7.1. Summary of the different PdZn based electrochemical catalyst.

Nomenclature PdZn/Na PdZn/K PdZn0.13/K

PdZn/Na- PdZn/K- PdZn/K-


Description
βAl2O3/Au βAl2O3/Au βAl2O3/Au

Pd/Zn molar ratio 1 1 0.13

Total Pd weight (mg) 7.66 7.66 7.66

Total Zn weight (mg) 4.71 4.71 36.22

PdZn alloy particle size after


78.0 84.5 52.6
reaction (nm)

Pd particle size after reaction


72.6 59.3 -
(nm)

ZnO particle size after reaction


- - 60.4
(nm)

NPd (active mol of metallic


6.57 x 10-7 9.71 x 10-8 -
palladium)

NPdZn (active mol of PdZn alloy) 9.00 x 10-7 1.88 x 10-6 3.25 x 10-6

NPd+PdZn (total active mol) 1.56 x 10-6 1.98 x 10-6 3.25 x 10-6

2.2. Electrochemical catalyst characterization

A temperature-programed reduction (TPR) experiment was carried out in situ


during the reduction of the catalysts under a stream of hydrogen (10 vol%)
diluted with nitrogen at a flow rate of 25 ml min-1. The temperature was increased
at a heating rate of 1.3 ºC min-1 up to 500 ºC. During the TPR, the H2 consumption
was continuously monitored by a thermal conductivity detector, whereas the in-
plane PdZn film surface electrical resistance between two points separated by 1
cm was measured using a digital multimeter. XRD analyses were carried out on a
Philips X’Pert instrument using nickel-filtered Cu-Ka radiation. Samples were
scanned at a rate of 0.02º step-1 over the range 5º≤2θ≤90º (scan time = 2 s
257
CHAPTER 7

step-1). The morphology of the different catalyst films was evaluated using a
Phenom Pro X scanning electron microscope (SEM). This instrument was equipped
with an Energy Dispersive X-ray Spectroscopy (EDX) analyzer to determine the
average composition of the samples. Cyclic voltammetry studies were carried out
using a Voltalab PGZ 301 potentiostat-galvanostat (Radiometer Analytical). The
potential was scanned between 2 and 2.5 V (starting potential) and -2 V at a
scan rate of 50 mV s-1. The current density generated by the applied cyclic
polarization was recorded as a function of the applied potential relative to the
counter-reference electrode.

2.3. Catalytic activity measurements

The experimental setup was described in detail in a previous publication


[37]. Reaction gases were Praxair certified standards of CO2 (99.999% purity),
H2 (99.999% purity) and N2 (99.999% purity). The gas flows were controlled by
a set of calibrated mass flowmeters (Brooks 5850 E and 5850 S).

Prior to reaction, catalysts were reduced as explained above for the TPR
experiment. The catalytic runs were carried out at atmospheric pressure. The total
flow rate was of 120 ml min-1, which is sufficiently high to avoid any mass transfer
limitation phenomena. Three H2/CO2 ratios were selected: 3 (75% H2 and 25%
CO2), 9 (90% H2 and 10% CO2) and 39 (97.5% H2 and 2.5% CO2). The
experiments were carried out at three temperatures (300, 320 and 340 ºC).
Reactant and product gases were on-line analyzed using a micro gas-
chromatograph (Varian CP-4900) equipped with two columns (Molsieve and
Poraplot Q column) and two thermal conductivity detectors (TCD). The detected
reaction products were CO and CH3OH. The error in the carbon atom balance
did not exceed 1%. In order to carry out the electrochemical promotion (EPOC)
experiments, the three electrodes (working, counter and reference) were
connected to a Voltalab PGZ 301 potentiostat- galvanostat (Radiometer

258
CHAPTER 7

Analytical) using gold wires. The CO2 conversion and the CO and CH3OH
selectivities were calculated as follows (Eqs. (7.4)-(7.6)):

F0CO2 -FCO2
CO2 conversion (%)= x 100 (7.4)
F0CO2

FCO
CO selectivity (%)= F0 x 100 (7.5)
CO2 -FCO2

FCH3 OH
CH3 OH selectivity (%)= F0 x 100 (7.6)
CO2 -FCO2

2.4. EPOC parameters

In the electrochemical promotion of catalysis there is a correspondence


between the different potentials applied and the amount of ions
electrochemically transferred to the metal film, thus modifying the adsorption of
the different gases and ultimately the catalytic behavior. Several different
parameters are commonly used to quantify the electropromotional effect:

- The rate enhancement ratio, r, defined by Eq. (7.7):

ρ = r/r0 (7.7)

where r is the electropromoted catalytic rate and r0 is the unpromoted


rate.

- The apparent Faradaic efficiency, Λ, defined by Eq. 7.8:

Λi = Δrcatalytic/(I/F) (7.8)

where Δrcatalytic is the current- or potential-induced observed change in


catalytic rate and I is the applied current. For this work, it implies:

ΛCO = 2ΔrCO/(I/F) (7.9)


ΛCH3OH = 6ΔrCH3OH/(I/F) (7.10)
where ΔrCO and ΔrCH3OH are in mol/s.

259
CHAPTER 7

For the cyclic voltammetry studies, the corresponding coverage of


potassium and sodium species established on the PdZn surface for each scan can
be calculated according to Faraday’s law (Eq. (7.11)) by integration of the
current density versus time or, equivalently, versus potential curves obtained for
the cathodic and anodic scan, respectively.
I A
θK or Na = ∫ dt= ∫ j dt (7.11)
FN F

where I is the current, F is the Faraday constant, N is the surface mol (mol
of active sites) of the PdZn catalyst electrode, A is the superficial surface area of
the electrode/electrolyte interface and j is the current density.

3. Results and discussion

3.1. Electrochemical catalyst characterization

The variation in the surface electrical resistance of the PdZn catalyst film
and the H2 consumption rate during a temperature programed reduction
experiment (TPR) for the PdZn/K sample is shown in Fig. 7.1. Significant
differences were not detected between the three electrochemical catalysts in the
TPR experiment. At the beginning of the experiment, the PdZn film showed a very
high electrical resistance (~23.7 MV). The decrease in the electrical resistance
started at around 60 ºC, when the first peak appeared in the TPR analysis. This
peak was attributed to the reduction of PdO to metallic palladium [32]. The
electrical resistance then stabilized around 4 V (See enlargement in Fig. 7.1). The
second peak (450 ºC) was related to the formation of the PdZn alloy [29, 32],
which is the active phase responsible for the formation of methanol.

260
CHAPTER 7

27.5

Resistance ()
25.0 5.5
22.5 5.0
4.5
20.0
4.0
Resistance (M)
17.5 100 200 300 400 500

Intensity (a.u.)
15.0 Temperature (ºC)
12.5
10.0
PdZn
7.5
5.0
2.5 K-β”Al2O3
0.0

50 100 150 200 250 300 350 400 450 500


Temperature (ºC)

Figure 7.1. TPR profile and variation of the PdZn Surface electrical resistance with temperature.

The XRD patterns of the samples PdZn/Na, PdZn/K and PdZn0.13/K are
shown in Fig. 7.2. The main peaks observed for the samples with a Pd/Zn molar
ratio of 1 were those of metallic palladium (JCPDS 87-0645) and PdZn alloy
(JCPDS 06-0620). The other small peaks that appeared were related to the solid
electrolytes (JCPDS 02-0921). In the case of the sample PdZn0.13/K, which has
a lower Pd/Zn molar ratio and therefore a higher zinc concentration in the
impregnating solution, peaks corresponding to ZnO crystals were observed
(JCPDS 80-0075) while the metallic palladium peaks were not. A difference in
the intensities of the peaks was observed for the samples PdZn/K and PdZn/Na:
for the latter the peaks due to the metallic palladium were higher and those for
the PdZn alloy were lower. The intensities of the peaks are related to the amount
of these species present [29, 38].

261
CHAPTER 7

º* * PdZn alloy
º Metallic palladium
+ ZnO
º º º
*
PdZn/K * * * * º * **
*
Intensity (a.u.)

+
++ + + +
* +
*
PdZn0.13/K * *+ + * + + * *
*
º
*

º º º
* *
PdZn/Na * * * º **

20 30 40 50 60 70 80 90 100
2 (º)

Figure 7.2. XRD profile of PdZn/K, PdZn0.13/K and PdZn/Na samples after reduction.

This situation is consistent with EDX mapping analysis on SEM images (Fig.
7.3d). The sample PdZn/Na showed the highest Pd percentage even though
PdZn/Na and PdZn/K samples should have the same quantity of palladium in the
film. This discrepancy could be due to the position of the metallic palladium
particles. These particles could be hidden in the pores of the film structure of the
PdZn/K sample but located on the surface for the PdZn/Na sample. Such a
phenomenon would have a consequence on the catalytic behavior, with the
PdZn/Na sample being more active for carbon monoxide production due to the
higher level of metallic palladium particles exposed. SEM images (Fig. 7.3a–c),
all at a resolution of 30 mm, showed the different external morphologies of the
catalyst films. The sample PdZn/Na (Fig. 7.3a) contained more holes and cavities
in a similar way to the sample PdZn/K (Fig. 7.3b). However, the surface of the
sample PdZn0.13/K is very different as it seems less compact, rougher and it
contains small holes. This morphology allows particles to be hosted within a high

262
CHAPTER 7

surface area. Therefore, one would expect the dispersion for this catalyst to be
higher.

a) b) c)

d)

Figure 7.3. SEM images of catalysts (a) PdZn/K (b) PdZn/Na and (c) PdZn0.13K. The EDX mapping
composition is also included in (d) for each simple.

The results of the cyclic voltammetry study are represented in Fig. 7.4. The
cathodic peaks, which appear with the application of decreasing potentials, are
related to species formed on the surfaceof the catalyst when the ions of the
electroactive support migrate to the catalyst interface and react with the
reactants. The anodic peaks, which appear with the application of increasing
potentials, are related to the decomposition of these species.

The voltammetry analysis for each sample at 300 ºC with a feed


composition of 90% H2 and 10% CO2 and a total flow rate of 120 ml min-1 are
shown in Fig. 7.4a. The low intensity shown by the sample PdZn/K is related to
the formation of only low levels of species on the surface. In contrast, the sample
PdZn/Na showed the highest intensity, probably due to the high capacity of this
sample to capture reactants on the surface and form Na compounds such as

263
CHAPTER 7

sodium carbonate [39, 40]. Moreover, this high capacity to capture reactants on
the surface could lead to deactivation of the active sites where these species are
formed and this could therefore reduce the activity of the catalyst [39]. This
possibility is supported by measurement of the potassium and sodium coverage
(θK and θNa) (see Table 7.2). The marked difference between the formation and
decomposition scan rates for the sample PdZn/Na suggests that decomposition
of the formed species does not occur. These species remain on the surface of the
catalyst film and block the active sites. The formation and decomposition scans
for the samples PdZn/K and PdZn0.13/K are very similar (the small differences
can be ascribed to the error inherent in the area integration). Therefore, a
reversible promotional phenomenon was observed during cyclic voltammetry at
all temperatures.
Table 7.2. Coverage of potassium and sodium species in Figure 4a.

PdZn/Na PdZn/K PdZn0.13/K

Decomposition scan 1.30 x10-3 2.46 x10-4 5.89 x10-4

Formation scan 2.42 x10-3 2.70 x10-4 5.76 x10-4

Relative change (%) 46.3 8.8 2.2

Another point that should be highlighted is the potential at which the


highest cathodic peak was obtained. These values were -0.25 V, -0.75 V and -1
V for PdZn/K, PdZn0.13/K and PdZn/Na, respectively. This potential increased
with temperature in each case (not shown here). At higher temperatures the same
effect was observed with lower electricity consumption, as higher temperatures
improve the migration of ions, thus increasing the electrolyte ionic conductivity
with a concomitant increase in the number of available potassium storage sites.
High temperature also increases the number of species formed.

264
CHAPTER 7

) )
2 2
300

(A/cm
300 a) a)

(A/cm
200
200
100
100

density
00

density
-100 300 ºC
300 ºC
-100
90% H2 10%10% COCO2

Current
-200 90% H
-200 2 2

) ) Current
PdZn/Na
PdZn/Na
-300
-300 PdZn/K
PdZn/K
-400
-400 PdZn0.13/K
PdZn0.13/K
500
2

500
(A/cm
400
400 b)
b)
2
(A/cm

D1
300
300
200
200 D2 D3
density

100
100
density

00
-100
-100
Current

F3 340 ºC
340 ºC
-200
) ) Current

-200 75% H2 25%25% COCO2


75% H
-300
-300 F2 2 2

-400
-400
F1
-500
-500
2 2

800 c)
A/cm

800 c)
density((A/cm

600
600
400
400
200
200
Currentdensity

00
-200
-200 340 ºC
340 ºC
-400
-400 75% H 25% CO22
CO
Current

-600
-600 75% H2 25% N N22
-800
-800 75% N2 25% CO CO22
-1000
-1000
-3
-3 -2
-2 -1 0 1 2 3
UPotential
WR
(V) vs(V)
Au

Figure 7.4. Cyclic voltammograms recorded over a) PdZn/K (blue), PdZn/Na (red) and PdZn0.13/K
(green) at 300 ºC under conditions of 10% CO2 and 90% H2; b) PdZn0.13/K at 340 ºC under 25%
CO2 and 75% H2; and c) PdZn0.13/K under different conditions: 25% CO2 and 75% H2 (blue), 25%
N2 and 75% H2 (red) and 25% CO2 and 75% N2 (green). (For interpretation of the references to colour
in this figure legend, the reader is referred to the web version of this article.

265
CHAPTER 7

A cyclic voltammogram for the sample PdZn0.13/K at 340 ºC is shown in


Fig. 7.4b, where the cathodic and anodic peaks reach the highest values with
75% H2/25% CO2 because, under these conditions, there is a greater quantity
of CO2 available, which is responsible for the formation of the cathodic peaks.
These cathodic peaks. In an effort to understand which species are formed, these
peaks were identified and denoted with the letters F (Formation of species) and
D (Decomposition of species). Three peaks appear in both sides of the cyclic
voltammogram. Two additional experiments, in which both reactants were
combined with N2, were carried out at 340 ºC (Fig. 7.4c) in order to establish a
correspondence between the position of the peaks and the identity of the phases
formed as a result of the potassium pumping. The peaks observed for 75% H2
and 25% N2 appeared at lower potentials (-1 V and -1.5 V), while the peak for
75% N2 and 25% CO2 appeared when the potassium ions were transferred to
the PdZn film (~0.5 V). Therefore, the formation of peak F1 is related to
carbonate species (potassium carbonate and bicarbonate) and the peaks F2 and
F3 to hydrogen storage as potassium hydrides. D1 and D3 were attributed to the
decomposition of these hydrides and D2 to the decomposition of carbonates. This
assignment is consistent with data reported in literature. Thus, a similar study was
carried out by Esperanza et al. [41] with CO2 and O2 in order to identify the
compounds formed when a cathodic peak appears. The formation of these
compounds has been also reported in other works: potassium carbonates were
detected by XPS under similar reaction conditions [42], as well as potassium
bicarbonates due to the formation of water during the reaction [43]. In addition,
in another work, the formation of potassium carbonate and potassium
bicarbonate was observed and confirmed by FTIR measurements [44].

3.2. Effect of the gas flow rate

The effects of the gas flow rate on the CO2 reaction rate, CO2 conversion
and selectivity to CO and CH3OH, at the reference state UWR = 2 V, for the
sample PdZn/K are represented in Fig. 7.5. The aim was to find the flow rate
266
CHAPTER 7

required to avoid mass transfer limitation phenomena. The reference


(unpromoted) CO2 reaction rate initially increased with the flow rate, reaching a
plateau at 120 ml min-1. The CO2 conversion decreased in a linear mannerupon
increasing theflow rate. The methanol selectivity increases atlower CO2
conversion, this behavior was also found in conventional catalysis [30].

3
(mol CO2·mol Pd+PdZn · s ) x 10
4.2
0.16
-1
4.0
0.14
3.8

CO2 conversion (%)


0.12
3.6
0.10
2
rCO

3.4
0.08
-1

3.2
0.06
3.0
0.04
2.8
0.02
90 90
CH3OH Selectivity (%)

80 80

CO Selectivity (%)
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
60 80 100 120 140 160
Total flow rate (ml·min-1)
Figure 7.5. Effect of the overall gas flow rate on the CO2 consumption rate, CO2 conversion and
CH3OH and CO selectivity, under +2 V polarization for the PdZn/K catalyst. Reaction conditions: 10%
CO2 and 90% H2 at 340 ºC.

3.3. PdZn/Na electrochemical catalyst

In order to study the catalytic behavior in depth, all samples (PdZn/K,


PdZn/Na and PdZn0.13/K) were tested under different reaction conditions and

267
CHAPTER 7

at different potentials (UWR) ranging from +2 V to -2 V. The study of each


catalyst separately was necessary due to the differences found in the dynamic
response of the reaction rates for each catalyst. It should be noted that for this
sample two types of particles are present, namely metallic Pd and PdZn alloy
particles, as shown by the results of an XRD analysis. These particles will have a
different electrocatalytic behaviors when Na+ ions migrate to the catalytic film.

The dynamic response of the reaction rates for CO and CH3OH formation
for the catalyst PdZn/Na with a feed composition of 90% H2 and 10% CO2 at
300 ºC is shown in Fig. 7.6. This sample showed that a low quantity of sodium
ions improves the PdZn alloy activity towards methanol (potentials of 0.5 V and
-0.5 V) compared to the reference state (2 V). Moreover, it can be seen that the
methanol production decreased at lower potentials due to the migration of large
amounts of Na+ onto the catalyst-working electrode, which would weaken the
chemical bond between PdZn and Pd and the electron donor adsorbate (H2) and
strengthen the bond with electron acceptors (CO2). The increase in the binding
strength of CO2 on the PdZn surface would disfavor methanol production because
the kinetic is disfavored by lower levels of the adsorbate H2. At the same time,
the increase in the binding strength of CO2 on the Pd surface would favor the
dissociative adsorption of CO2 through the reverse water gas-shift reaction (Eq.
(7.2)). Therefore electrophilic behavior was obtained for CO because a cathodic
(negative) polarization enhanced the carbon monoxide production rate. The
highest CO production was obtained at -1 V, where the cathodic peak was found
in the cyclic voltammetry study (Fig. 7.4). From this potential, the CO formation
rate showed sharp positive peaks whereas the CH3OH formation rate showed
negative peaks. This behavior is consistent with the results of the cyclic
voltammetry analysis. As explained above, the high intensity found in the
voltammetry study at -1 V is related to the formation of sodium carbonates and
sodium hydrides. These compounds can block the active sites of the catalyst and
decrease its activity. The differences found in the values for the formation and
268
CHAPTER 7

decomposition peak areas in the voltammetry study suggest that these species
are not totally decomposed after each cycle.
3
25 3

(mol CO·mol Pd+PdZn·s )


25

-1
2
2
20
1

UWR (V) vs Au
20
1
15

3
0

x 10
rCO 15 0
-1
10 -1
10 -1
-2
5 -2
5
8.5
8.5 2
(mol CH3OH·mol PdZn·s )
-1

8.0 2
8.0
1
7.5

UWR (V) vs Au
1
7.5
-1

7.0 0
3
rCH3OH

x 10

7.0 0
6.5 -1
6.5 -1
6.0
6.0 -2
-2
5.5
5.5 -3
0 100 200 300 400 500 -3
600
0 100 200 300 400 500 600
Time (min)

Figure 7.6. Influence of the applied potential vs. time on the reaction rate values of CO production and
CH3OH formation for the catalyst PdZn/Na under conditions of 10% CO2 and 90% H2 at 300 ºC. Total
flow rate = 120 ml min-1.

This situation is consistent with the catalytic results: the highest CO


production rate found at -1 V (where more species are formed) is not obtained
at -1.5 V or -2 V due to the blocking effect of the species that are not
decomposed. Hence, the trend in the methanol formation rate is opposite to that
of the CO formation rate because the kinetic of the latter is favored by the
strengthening with the CO2 (electron acceptor) while the methanol kinetic is
disadvantaged. In addition, although CO is formed preferentially, when
269
CHAPTER 7

poisoning in the CO formation rate occurs due to the formation of certain


compounds, the blocking produced at the active palladium sites leads to a
recovery in the formation rate of methanol because the free H2 (not used in
blocked Pd active sites) migrates to PdZn active sites to produce methanol.

3.4. PdZn/K electrochemical catalyst

The dynamic responses of the reaction rates for CO2, CO and CH3OH for
the catalyst PdZn/K under all the conditions used in this work are shown in Fig.
7.7. The reaction rates of CO2 and CO are fairly similar since the contribution of
the CH3OH formation rate to the total is low. Although methanol is obtained in a
lower order of magnitude, the production of this compound is remarkable and
this is the first time that a dynamic and stable response has been reported for
this product in the hydrogenation of CO2.

Although the PdZn/Na and PdZn/K samples have similar particle sizes for
the Pd and the PdZn alloy, the PdZn/Na sample showed high activity towards
methanol and CO. This behavior could be due to the higher amount palladium
exposed in the form of PdZn or metallic palladium at the external surface, as
shown by EDX analysis in SEM images (Fig. 7.3). As for the PdZn/Na sample, the
methanol formation rate increases at 0.5 V and decreases at lower potentials,
whereas electrophilic behavior was found for the carbon monoxide formation
rate. When the K+ ions migrate to the catalytic film, an increase in the CO2
(electron acceptor) production is observed and, consequently, this change favors
CO production and inhibits methanol production. The difference between the
promotional effects is that in the case of the PdZn/K sample the effect is stable.

The reaction rates of CO2 and CO increased with temperature due to the
enhancement of the CO reaction (Eq. (7.2)). When the H2/ CO2 ratio in the feed
atmosphere was decreased, higher reaction rates were obtained for CO and
CO2 but more time was required to return to the value of the unpromoted state
(+2 V). In some cases, however, this value was not reached at high temperatures,
270
CHAPTER 7

thus leading to an increase in the global activity of the catalyst. The reaction rate
of CH3OH also increased with temperature at lower H2/CO2 ratios (3 and 9),
but this was not the case for the 2.5% CO2/ 97.5% H2 mixture. Although the
methanol reaction rate increased at +0.5 V, the selectivity towards this compound
decreased at this potential (Table 7.3) due to the increase found for the carbon
monoxide reaction rate at +0.5 V. The highest methanol selectivity was found for
the highest H2/CO2 ratio (39, corresponding to the 2.5% CO2/97.5% H2
experiment) at 300 ºC and 2 V. The selectivity was around 76% but this
decreased to 60% and 41% at the ratios 9 and 3, respectively. Therefore, higher
H2/CO2 ratios led to higher selectivity towards methanol. In general, the methanol
selectivity decreased with the temperature due to the thermodynamical limitations
of this reaction and the preferential formation of CO. The optimum of selectivity
was found at 300 ºC and this is consistent with the equilibrium for methanol
production (Eq. (7.3)), which is favored at lower temperatures.

The variations with applied potential of the CO reaction rate


enhancement ratio (ρCO) and the apparent Faradaic efficiencies for both
products, under all the feed compositions and for two temperatures (300 ºC and
340 ºC), are represented in Fig. 7.8. The figure also shows the CH3OH reaction
rate enhancement ratio (ρCH3OH) at the potential +0.5 V, where the ratio is higher
than 1.
Table 7.3. Methanol selectivity and CO2 conversion for the catalyst PdZn/K.

2.5%CO2 97.5%H2 10%CO2 90%H2 25%CO2 75%H2


SCH3OH (%) XCO2 (%) SCH3OH (%) XCO2 (%) SCH3OH (%) XCO2 (%)
2V 76 0.1 60 < 0.1 41 < 0.1
300ºC
0.5V 54 0.2 45 < 0.1 31 < 0.1
2V 65 0.2 46 < 0.1 20 < 0.1
320ºC
0.5V 33 0.3 25 0.1 17 0.1
2V 61 0.2 35 0.1 18 0.1
340ºC
0.5V 31 0.3 16 0.2 12 0.1

271
30 4

-1
2.5% CO2 97.5% H2 10% CO2 90% H2 25% CO2 75% H2
25 3
300 ºC 320 ºC 340 ºC 300 ºC 320 ºC 340 ºC 300 ºC 320 ºC 340 ºC
CHAPTER 7

20 2

3
1

-1
15

rCO2
x 10
0
10
UWR (V) vs Au

-1
5
-2

(mol CO2·mol Pd+PdZn·s )


0

-1
2.5% CO2 97.5% H2 10% CO2 90% H2 25% CO2 75% H2
25 3
300 ºC 320 ºC 340 ºC 300 ºC 320 ºC 340 ºC 300 ºC 320 ºC 340 ºC
20 2

3
1
15

-1

rCO
x 10
0

272
10
UWR (V) vs Au

-1
5
-2

(mol CO·mol Pd+PdZn·s )


0

-1
2.5% CO2 97.5% H2 10% CO2 90% H2 25% CO2 75% H2
2 3
300 ºC 320 ºC 340 ºC 300 ºC 320 ºC 340 ºC 300 ºC 320 ºC 340 ºC
2

-1
3
1
1

320 (green) and 340 ºC (red). Total flow rate = 120 ml min-1.
x 10

rCH3OH
0

3
UWR (V) vs Au

-1
0 -2

(mol CH OH·mol PdZn·s )


-3
0 75 150 225 300 375 450 525 600 75 150 225 300 375 450 525 600 75 150 225 300 375 450 525 600

CO production and CH3OH formation for the catalyst PdZn/K under three different feed ratios: 25%
Time (min) Time (min) Time (min)

CO2 and 75% H2, 10% CO2 and 90% H2, 2.5% CO2 and 97.5% H2 at the temperatures of 300 (blue),
Figure 7.7. Influence of the applied potential vs. time on the reaction rate values of CO2 consumption,
CHAPTER 7

10
9 2.5 % CO 97.5% H 10 % CO 90% H 25 % CO 75% H
2 2 2 2 2 2
8
7 300 ºC 340 ºC
6

CO
5
4
3
2
1
3500 2.5 % CO 97.5% H 10 % CO 90% H 25 % CO 75% H
2 2 2 2 2 2
3000
2500
2000
CO

1500
1000
500
0
300 2.5 % CO 97.5 % H 10 % CO 90% H 25 % CO 75 % H
2 2 2 2 2 2
200  = 1.06  = 1.04
CH3OH
100 CH3OH
 = 1.03
0 CH3OH
CH OH

-100  = 1.02

3

-200  = 1.02 CH3OH


= 1.06 CH3OH
CH3OH
-300
-400
-500
-600
-2 -1 0 1 2 -2 -1 0 1 2 -2 -1 0 1 2
UWR (V) vs Au UWR (V) vs Au UWR (V) vs Au

Figure 7.8. Effect of the applied potential (UWR) on the rate enhancement ratio (ρi) and on the apparent
Faradaic efficiency (Λi) for the catalyst PdZn/K under three different feed ratios: 25% CO 2 and 75%
H2, 10% CO2 and 90% H2, 2.5% CO2 and 97.5% H2 at the temperatures of 300 and 340 ºC. Total
flow rate = 120 ml min-1.

The CO reaction rate enhancement ratio increased at higher H2/CO2


ratios, at higher temperature, and at negative potentials, which proves that the
latter favors CO2 adsorption and consequently the kinetic for the production of
CO. The highest CH3OH reaction rate enhancement ratio at 300 ºC was found
for the conditions 10% CO2/ 90% H2. The efficiency achieved for CO (ΛCO =
3250) is the highest value found in the literature for a hydrogenation reaction.
For instance, in a recent study [16] Λ values as high as 500 were claimed for the
first time but these are significantly lower than the values found in this study.
Interestingly, both apparent Faradaic efficiencies exhibit a pronounced maximum
with varying anodic potential, as observed in other work [16]. This behavior may
be due to the fact that the first migration of potassium ions to the PdZn film is
273
CHAPTER 7

produced easily because it is favored by the concentration gradient of ions. As a


consequence, only a very small current is necessary for this process. However,
when the potential continues to decrease, the movement of potassium ions
becomes more difficult due to the existence of ions on the PdZn film. Thus, higher
currents are necessary and this leads to a decrease in the Λ values.

3.5. PdZn0.13/K electrochemical catalyst

The PdZn0.13/K catalyst was prepared in order to study the influence of


the Pd/Zn molar ratio in the catalytic film formed. Two main differences were
found during the characterization: (i) a different external morphology of the
catalyst film (Fig. 7.3), i.e., less compact, rougher and with small holes and (ii)
only PdZn alloy particles were evidenced in the XRD analysis (Fig. 7.2). The
consequence of the first characteristic is that the particles formed were smaller
and the dispersion was higher. These differences are essential to understand the
catalytic results because, in this case, the electrocatalytic behavior depends only
on the PdZn particles and how they act when the migration of K+ ions is produced.

The dynamic response of the reaction rates for CO and CH3OH formation
for the catalyst PdZn0.13/K with a feed composition of 90% H2 and 10% CO2
at 300 ºC is shown in Fig. 7.9. Although this experiment was selected as an
example, the experiments carried out under different conditions showed the same
trend. This sample shows the same catalytic behavior for both products (CO and
CH3OH). The migration of a small quantity of K+ ions produced at 0.5 V improves
the kinetics of both reactions but at lower potentials (<0.5 V) the migration of K+
ions to the PdZn film inhibits the global activity of the electrochemical catalyst.

The Pd/Zn molar ratio did not show any influence on methanol production
because similar formation rates were obtained for both catalysts (PdZn/K and
PdZn0.13/K). The same behavior was also found in other work on conventional
catalysts [30], where it was found that a certain number of PdZn alloy particles

274
CHAPTER 7

provide the maximum methanol production regardless of further changes to the


Pd/Zn molar ratio.

33
8.0
8.0

(mol CO·mol -1PdZn·s8 )


22

-1
7.5
7.5

x 10 ) x 10

UWR (V) vs Au
7.0
7.0 11
6.5
6.5

3
-1

rCO (mol CO·s


rCO
00
6.0
6.0
5.5
5.5 -1-1
5.0
5.0
-2-2
4.5
4.5
)9
10
-1

22
(mol CH3OH·mol PdZn·s

0.8
0.8
) x
-1

UWR (V) vs Au
11
rCH3OH (mol CH3OH·s

0.6
0.6
-1

3
rCH3OH

x 10

00
0.4
0.4
-1-1
0.2
0.2
-2-2
0.0
0.0

00 100
100 200
200 300
300 400
400 500
500 600
600
Time (min)

Figure 7.9. Influence of the applied potential vs. time on the reaction rate values of CO production and
CH3OH formation for the catalyst PdZn0.13/K under conditions of 10% CO 2 and 90% H2 at 300 ºC.
Total flow rate = 120 ml min-1.

The high CO production rate when compared with the PdZn/K sample
occurs because the PdZn particles are smaller and the dispersion is higher,
therefore the activity of the catalyst is higher. It w.as noted that while the
methanol reaction rate seems similar in the different cycles (+2 to a negative
potential), the carbon monoxide formation rate decreased. Thus, deactivation of
the electrochemical catalyst is produced in this process. The authors of this work
suggest that the deactivation is produced in PdZn alloy particles. This effect is
275
CHAPTER 7

only evident in CO formation rates because the order of magnitude in which this
compound is produced is higher, so the effect of deactivation is easier to see.
Evidence for this phenomenon is provided by the existence of carbonaceous
deposits (Fig. 7.10) on the catalyst surface, with a carbon content greater than
60%. The authors suggest that the deactivation is produced in all of the PdZn
alloy particles (also in the hidden particles), but carbonaceous deposits were only
found on the surface at the points where the Au wires were connected with the
PdZn alloy. This finding is due to the high activity and migration of ions in the
vicinity of the connections. The deactivation observed between the first and the
last reference state in the CO formation rate is around 25% for all temperatures
tested.

Figure 7.10. SEM image of a carbonaceous deposit found on the surface of the PdZn0.13/ K sample.

276
CHAPTER 7

4. Conclusions
The following conclusions can be drawn from this study:

- The PdZn alloy can be used as a catalytic film in an electrochemical system.


Methanol and carbon monoxide were the two only products obtained.

- The formation of potassium carbonates and potassium hydrides was


demonstrated by a cyclic voltammetry study on the sample PdZn0.13/K. The
sample PdZn/Na showed high intensities in the cyclic voltammetry study. These
values are related to a higher capture of reactants through the formation of
sodium compounds.

- PdZn/Na and PdZn/K samples shows how a low quantity of ions increases the
formation rate of CH3OH and a high quantity leads to the poisoning of the PdZn
active sites. An electrophilic behavior for the CO formation rate was found. These
behaviors are attributed to the coexistence of large particles of metallic
palladium and PdZn alloy.

- The PdZn/Na sample showed an unstable promotional effect in the dynamic


response. As shown by cyclic voltammetry, sodium carbonates and sodium
hydrides are formed and these could block the active sites, thus decreasing the
reaction rate.

- The PdZn/K sample showed a stable promotional effect for both products. A
complete dynamic study was carried out under different conditions: feed H2/CO2
ratios of 3, 9 and 39 and temperatures of 300, 320 and 340 C. Apparent
Faradaic efficiency values up to 1000 were obtained.

- The PdZn0.13/K sample showed the same catalytic behavior for both products.
CO and CH3OH formation rates increase at +0.5 V compared to the unpromoted
state (+2 V) and these decreased at negative potentials. This behavior was
attributed to the fact that the EPOC effect in this catalyst is produced in PdZn
alloy particles.

277
CHAPTER 7

References
[1] Olajire, A.A. Valorization of greenhouse carbon dioxide emissions into value-
added products by catalytic processes. Journal of CO2 Utilization, 3-4 (2013)
74-92.

[2] Ali, K. A.; Abdullah, A. Z.; Mohamed, A. R. Recent development in catalytic


technologies for methanol synthesis from renewable sources: a critical review.
Renewable and Sustainable Energy Reviews, 44 (2015) 505-518.

[3] Saeidi, S.; Amin, N. A. S.; Rahimpour, M. R. Hydrogenation of CO2 to value-


added products - A review and potential future developments. Journal of CO2
Utilization, 5 (2014) 66-81.

[4] Gutiérrez-Guerra, N.; Moreno-López, L.; Serrano-Ruiz, J. C.; Valverde, J. L.;


de Lucas-Consuegra, A. Gas phase electrocatalytic conversion of CO2 to syn-fuels
on Cu based catalysts-electrodes. Applied Catalysis B, 188 (2016) 272-282.

[5] Sun, K.; Fan, Z.; Ye, J.; Yan, J.; Ge, Q.; Li, Y.; He, W.; Yang, W.; Liu, C.-j.
Hydrogenation of CO2 to methanol over In2O3 catalyst. Journal of CO2
Utilization, 12 (2015) 1-6.

[6] Huang, C.; Chen, S.; Fei, X.; Liu, D.; Zhang, Y. Catalytic hydrogenation of CO2
to methanol: study of synergistic effect on adsorption properties of CO2 and H2
in CuO/ZnO/ZrO2 system. Catalysts, 5 (2015) 1846-1861.

[7] Jiang, X.; Koizumi, N.; Guo, X.; Song, C. Bimetallic Pd-Cu catalysts for selective
CO2 hydrogenation to methanol. Applied Catalysis B, 170-171 (2015) 173-185.

[8] Saito, M. R&D activities in Japan on methanol synthesis from CO2 and H2.
Catalysis Surveys from Japan, 2 (1998) 175-184.

[9] Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M.; Joshi, J. B. Catalytic carbon
dioxide hydrogenation to methanol: a review of recent studies. Chemical
Engineering Research and Design, 92 (2014) 2557-2567.

278
CHAPTER 7

[10] Shulenberger, A. M.; Jonsson, F. R.; Ingolfsson, O.; Tran, K. C. Process for
producing liquid fuel from carbon dioxide and water. Google Patents, 2012.

[11] de Lucas-Consuegra, A. New trends of alkali promotion in heterogeneous


catalysis: electrochemical promotion with alkaline ionic conductors. Catalysis
Surveys from Asia, 19 (2015) 25-37.

[12] Stoukides, M.; Vayenas, C. G. The effect of electrochemical oxygen pumping


on the rate and selectivity of ethylene oxidation on polycrystalline silver, J. Catal.
70 (1981) 137–146.

[13] Vayenas, C. G.; Bebelis, S.; Ladas, S. Dependence of catalytic rates on


catalyst work function, Nature 343 (1990) 625-627.

[14] Vayenas, C. G.; Bebelis, S.; Pliangos, C.; Brosda, S.; Tsiplakides, D.
Electrochemical Activation of Catalysis: Promotion, Electrochemical Promotion, and
Metal- Support Interactions, Springer US2007.

[15] Gutiérrez-Guerra, N.; González-Cobos, J.; Serrano-Ruiz, J. C.; Valverde, J.


L.; de Lucas-Consuegra, A. Electrochemical activation of Ni catalysts with
potassium ionic conductors for CO2 hydrogenation. Topics in Catalysis, 58 (2015)
1256-1269.

[16] Kalaitzidou, I.; Katsaounis, A.; Norby, T.; Vayenas, C. G. Electrochemical


promotion of the hydrogenation of CO2 on Ru deposited on a BZY proton
conductor, Journal of Catalysis, 331 (2015) 98-109.

[17] Bebelis, S.; Karasali, H.; Vayenas, C. G. Electrochemical promotion of the


CO2 hydrogenation on Pd/YSZ and Pd/β´´-Al2O3 catalyst-electrodes, Solid
State Ionics 179 (2008) 1391–1395.

[18] Ruiz, E.; Cillero, D.; Martínez, P. J.; Morales, Á.; Vicente, G. S.; de Diego, G.;
Sánchez, J. M. Bench scale study of electrochemically promoted catalytic CO2
hydrogenation to renewable fuels. Catalysis Today, 210 (2013) 55-66.

279
CHAPTER 7

[19] Ruiz, E.; Cillero, D.; Martínez, P. J.; Morales, Á.; Vicente, G. S.; de Diego, G.;
Sánchez, J. M.; Electrochemical synthesis of fuels by CO2 hydrogenation on Cu in
a potassium ion conducting membrane reactor at bench scale. Catalysis Today,
236 (Part A) (2014) 108-120.

[20] Theleritis, D.; Makri, M.; Souentie, S.; Caravaca, A.; Katsaounis, A.; Vayenas,
C. G. Comparative study of the electrochemical promotion of CO2 hydrogenation
over Ru-Supported catalysts using electronegative and electropositive promoters.
ChemElectroChem, 1 (2014) 254-262.

[21] Makri, M.; Katsaounis, A.; Vayenas, C. G. Electrochemical promotion of CO2


hydrogenation on Ru catalyst-electrodes supported on a K-β´´Al2O3 solid
electrolyte, Electrochimica Acta, 179 (2015) 556-564.

[22] Papaioannou, E. I.; Souentie, S.; Hammad, A.; Vayenas, C. G.


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.

[23] Bebelis, S.; Karasali, H.; Vayenas, C. G. Electrochemical promotion of CO2


hydrogenation on Rh/YSZ electrodes, Journal of Applied Electrochemical, 38
(2008) 1127-1133.

[24] Jiménez, V.; Jiménez-Borja, C.; Sánchez, P.; Romero, A.; Papaioannou, E. I.;
Theleritis, D.; Souentie, S.; Brosda, S.; Valverde, J. L. Electrochemical promotion
of the CO2 hydrogenation reaction on composite Ni or Ru impregnated carbon
nanofiber catalyst-electrodes deposited on YSZ. Applied Catalysis B, 107 (2011)
210-220.

[25] Ruiz, E.; Cillero, D.; Martínez, P. J.; Morales, Á.; Vicente, G. S.; de Diego, G.;
Sánchez, J. M. 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.

280
CHAPTER 7

[26] Theleritis, D.; Souentie, S.; Siokou, A.; Katsaounis, A.; Vayenas, C. G.
Hydrogenation of CO2 over Ru/YSZ electropromoted catalysts. ACS Catalysis, 2
(2012) 770-780.

[27] Pekridis, G.; Kalimeri, K.; Kaklidis, N.; Vakouftsi, E.; Iliopoulou, E. F.;
Athanasiou, C.; Marnellos, G. E. Study of the reverse water gas shift (RWGS)
reaction over Pt in a solid oxide fuel cell (SOFC) operating under open and
closed-circuit conditions. Catalysis Today, 127 (2007) 337-346.

[28] Karagiannakis, G.; Zisekas, S.; Stoukides, M. Hydrogenation of carbon


dioxide on copper in a H+ conducting membrane-reactor. Solid State Ionics, 162-
163 (2003) 313-318.

[29] Díez-Ramírez, J.; Valverde, J. L.; Sánchez, P.; Dorado, F. CO2 Hydrogenation
to Methanol at atmospheric pressure: influence of the preparation method of Pd/
ZnO catalysts. Catalysis Letters, 146 (2016) 373-382.

[30] Díez-Ramírez, J.; Sanchez, P.; Rodríguez-Gomez, A.; Valverde, J. L.; Dorado,
F. Carbon nanofiber-based palladium/zinc catalysts for the hydrogenation of
carbon dioxide to methanol at atmospheric pressure. Industrial & Engineering
Chemistry Research, 55 (2016) 3556-3567.

[31] Chin, Y. H.; Wang, Y.; Dagle, R. A.; Li, X. S. Methanol steam reforming over
Pd/ZnO: Catalyst preparation and pretreatment studies. Fuel Processing
Technology, 83 (2003) 193-201.

[32] Chin, Y. H.; Dagle, R.; Hu, J.; Dohnalkova, A. C.; Wang, Y. Steam reforming
of methanol over highly active Pd/ZnO catalyst. Catalysis Today 77 (2002) 79-
88.

[33] Wang, Y.; Zhang, J.; Xu, H. Interaction between Pd and ZnO during reduction
of Pd/ZnO catalyst for steam reforming of methanol to hydrogen. Chinese
Journal of Catalysis, 27 (2006) 217-222.

281
CHAPTER 7

[34] Iwasa, N.; Mayanagi, T.; Ogawa, N.; Sakata, K.; Takezawa, N. New
catalytic functions of Pd-Zn, Pd-Ga Pd-In, Pt-Zn, Pt-Ga and Pt-In alloys in the
conversions of methanol. Catalysis Letters, 54 (1998) 119-123.

[35] Zhang, H.; Sun, J.; Dagle, V. L.; Halevi, B.; A.K. Datye, Y. Wang, Influence of
ZnO facets on Pd/ZnO catalysts for methanol steam reforming. ACS Catalysis, 4
(2014) 2379-2386.

[36] Iwasa, N.; Masuda, S.; Ogawa, N.; Takezawa, N. Steam reforming of
methanol over Pd/ZnO: effect of the formation of PdZn alloys upon the reaction,
Applied Catalysis A, 125 (1995) 145-157.

[37] de Lucas-Consuegra, A.; Gutiérrez-Guerra, N.; Caravaca, A.; Serrano-Ruiz,


J. C.; Valverde, J. L. Coupling catalysis and electrocatalysis for hydrogen
production in a solid electrolyte membrane reactor. Applied Catalysis A, 483
(2014) 25-30.

[38] Iwasa Suzuki, N. H.; Terashita, M.; Arai, N. Methanol synthesis from CO2
under atmospheric pressure over supported Pd catalysts, Catal. Lett. 96 (2004)
75-78.

[39] Filkin, N. C.; Tikhov, M. S.; Palermo, A.; Lambert, R. M. A kinetic and
spectroscopic study of the in situ electrochemical promotion by sodium of the
platinum-catalyzed combustion of propene. The Journal of Physical Chemistry A,
103 (1999) 2680-2687.

[40] Vernoux, P.; Gaillard, F.; Lopez, C.; Siebert, E. In-situ electrochemical control
of the catalytic activity of platinum for the propene oxidation. Solid State Ionics,
175 (2004) 609–613.

[41] Ruiz, E.; Cillero, D.; Morales, Á.; Vicente, G. S.; de Diego, G.; Martínez, P.
J.; Sánchez, J. M. Bench-scale study of electrochemically promoted CO2 capture
on Pt/ K-βAl2O3, Electrochimica Acta, 112 (2013) 967-975.

282
CHAPTER 7

[42] Urquhart, A. J.; Keel, J. M.; Williams, F. J.; Lambert, R. M. Electrochemical


promotion by potassium of rhodium-catalyzed Fischer-Tropsch synthesis: XP
spectroscopy and reaction studies. The Journal of Physical Chemistry B, 107
(2003) 10591.

[43] Vernoux, P.; Gaillard, F.; Lopez, C.; Siebert, E. In-situ electrochemical control
of the catalytic activity of platinum for the propene oxidation, Solid State Ionics
175 (2004).

[44] de Lucas-Consuegra, A.; Dorado, F.; Valverde, J.L.; Karoum, R.; Vernoux, P.
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.

283
CONCLUSIONES
CONCLUSIONES

Las conclusiones generales obtenidas de este trabajo son:

- La síntesis de metanol compite con las reacciones de formación de monóxido de


carbono y metano, por lo cual es crucial el uso de un catalizador activo y selectivo
hacia metanol.

- Entre los metales probados en reacción: Pd, Cu y Co, tan solo los dos primeros
muestran velocidades de formación de metanol a partir de dióxido de carbono
y selectividades hacia metanol a presión atmosférica, similares a las de otros
catalizadores encontrados en bibliografía. El cobalto, sin embargo, cataliza en
un mayor grado (>90%) la formación de metano.

- Las etapas de calcinación y reducción son de vital importancia, tanto en los


catalizadores de Pd/ZnO, como en los de Cu/ZnO. En el primero, condicionan el
tamaño y la cantidad de centros activos de aleación PdZn. En el segundo, el
tamaño, el estado de oxidación (Cu2+, Cu1+, Cu0) y la especie activa (Cu metálico
o aleación CuZn) del catalizador.

- El soporte utilizado puede modificar completamente el comportamiento de las


partículas de aleación PdZn. En el caso de las nanofibras de carbono tipo
platelet, la interacción con el soporte da lugar a nuevas partículas con una
relación molar 48.5:51.5 (Pd48.5Zn51.5), que muestran una mayor selectividad y
actividad hacia metanol en comparación con la aleación de relación molar
50:50.

- El uso de nanofibras de carbono desplaza hacia temperaturas mayores la curva


de formación de metanol, obteniéndose metanol a temperaturas superiores a
300ºC. Este hallazgo permitiría que fuesen utilizadas en reactores
electrocatalíticos, cuyas membranas protónicas necesitan altas temperaturas
para el desplazamiento de los iones.

- En los catalizadores de Pd/ZnO las partículas grandes de aleación PdZn


(50:50) son más selectivas hacia metanol puesto que disminuyen la conversión

287
CONCLUSIONES

total de CO2. Sin embargo, al combinar Pd, Cu y Zn en un mismo catalizador la


selectividad hacia metanol no disminuye sólo porque el tamaño de las partículas
de aleación PdZn (50:50) aumenta, sino porque las partículas de paladio
metálico se incorparan a la estructura de la aleación disminuyendo el porcentaje
de las mismas. Esta disminución del porcentaje de partículas de paladio metálico
(invisbles al XRD) hace que disminuya la producción de CO y aumente por tanto,
la selectividad hacia metanol.

- El efecto sinérgico del Pd, Cu y Zn queda demostrado en los catalizadores tri-


metálicos, que muestran una mayor selectividad hacia metanol en comparación
con los bimetálicos PdZn/SiC y CuZn/SiC. El cobre actúa como inhibidor del
paladio metálico formando la aleación PdCu, menos activa hacia la síntesis de
monóxido de carbono.

- El mejor catalizador en términos de selectividad y actividad hacia metanol


obtenido en este trabajo tiene una composición de 37,5 % Pd, 12,5% Cu y 50%
Zn. Con este catalizador se ha desarrollado un modelo cinético que predice con
garantías los resultados experimentales esperados a diferentes composiciones y
temperaturas.

- La aleación PdZn es suceptible de ser utilizada como electrodo en un reactor


electrocatalítico de cámara simple dando lugar únicamente a metanol y
monóxido de carbono. Al estudiar el efecto NEMCA en el sistema, la migración
de pocos iones K+ y Na+ (0.5V) favorece la síntesis de metanol, sin embargo, un
mayor porcentaje de ambos iones (< 0.5 V) produce una inhibición de la
formación de metanol.

288
FUTURE WORK/
RECOMENDACIONES
FUTURE WORK

Further studies could be conducted following the work developed in this


thesis in the methanol synthesis using heterogeneous catalysis and electrocatalysis.
The following proposals can be stated in order to complete and extent this
research work:

- To study the catalyst stability in long term experiments, analyzing the


deactivation process.

- To perform a simulation of a methanol synthesis industrial plant using the kinetic


model obtained in the present work in order to analyze its behavior under
different conditions or analyse the energetic and exergetic losses.

- To evaluate new metals, supports and different catalyst preparation methods.

- Once the development of protonic conductors has been carried out and the ionic
migration will be able to be performed at temperatures around 300ºC,
experiments in a double chamber reactor could be done to obtain methanol from
CO2 and H2O.

291
RECOMENDACIONES

Tomando de base el estudio de investigación abordado en esta tesis


doctoral sobre la síntesis de methanol a partir de CO2 e H2 mediante catálisis
heterogénea y electrocatálisis, se plantean las siguientes propuestas para
completar y ampliar el mismo:

- Estudiar la estabilidad del catalizador en experimentos de larga duración y


analizar el proceso de desactivación.

- Realizar una simulación de la planta industrial de síntesis de metanol utilizando


el modelo cinético obtenido en el presente trabajo para analizar su
comportamiento bajo diferentes condiciones y analizar las pérdidas energéticas
y exergéticas de la planta.

- Evaluar nuevos metales, soportes y distintos métodos de preparación de


catalizadores.

-Una vez se haya llevado a cabo el desarrollo de conductores protónicos que


puedan trabajar a temperaturas alrededor de los 300ºC consiguiendo un
movimiento de protones aceptable, sería interesante realizar experimentos en un
reactor de doble cámara para obtener metanol a partir de CO2 y H2O.

292
ANEXO:
INSTALACIONES
EXPERIMENTALES
ANEXO: INSTALACIONES EXPERIMENTALES

En este trabajo se han utilizado dos instalaciones experimentales


principales: la instalación de catálisis convencional y la de electrocatálisis. En
ambas instalaciones se pueden diferenciar tres zonas comunes que se detallan a
continuación: sistema de alimentación, sistema de reacción y sistema de análisis.
Además de estas, en la instalación de electrocatálisis se encuentra el sistema de
polarización para la aplicación controlada de potenciales.

1. Catálisis convencional

En la Figura A.1 se muestra el esquema de la instalación experimental


donde se llevaron a cabo los experimentos de catálisis heterogénea.

Controlador de
Reactor temperatura

CO2
H2
Horno
N2

Campana de
gases
Controladores
de caudal
Medidor de caudal

Micro
cromatógrafo

Figura A.1. Instalación experimental: catálisis convencional.

295
ANEXO: INSTALACIONES EXPERIMENTALES

1.1. Sistema de alimentación


El sistema de alimentación está constituido por tres líneas de flujo continuo,
análogas e independientes, para la alimentación de los gases de reacción: CO2
(99.999% pureza) e H2 (99.999% pureza) y el N2 (99.999% pureza) utilizado
para diluir la corriente de entrada en la etapa de reducción. El dióxido de
carbono se alimenta desde una bala que contiene el gas a alta presión, mientras
que el hidrógeno y el nitrógeno proceden de una canalización general de gases.
Cada línea de flujo está compuesta por una tubería de acero inoxidable de
1/8’’ con un manorreductor, un filtro de acero inoxidable, un controlador
indicador de caudal másico (FIC) y una válvula antirretorno (VR).

Los controladores másicos, de la marca BROOKS INSTRUMENTS (modelo


5850E para el nitrógeno y modelo 5850S para el resto de gases), están
constituidos por un sensor de conductividad térmica, un indicador controlador y
una electroválvula. Estos son controlados mediante un software informático.

1.2. Sistema de reacción

El sistema de reacción está constituido por un reactor tubular de lecho fijo


y flujo descendente construido en cuarzo, de 1 cm de diámetro interno y 45 cm
de longitud. En la parte superior se inserta la conducción de entrada de los gases
reaccionantes.

El termopar para medir la temperatura del lecho catalítico se sitúa de


forma que el extremo del mismo queda situado al mismo nivel que el lecho de
catalizador, situado sobre el soporte, de este modo la temperatura medida por
el termopar coincide con la de la reacción. Este termopar está conectado a un
controlador de temperaturas y horno eléctrico LENTON THERMAL DESIGN, que
permite monitorizar y controlar la temperatura en el reactor, así como realizar
rampas de calentamiento y enfriamiento. El reactor se sitúa dentro del horno
eléctrico.

296
ANEXO: INSTALACIONES EXPERIMENTALES

1.3. Sistema de análisis

La línea de salida del reactor hasta el sistema de análisis se mantiene

S-1
calefactada para evitar la condensación de productos como el metanol e impedir

S-1
que entren líquidos al sistema de análisis. Así, la corriente gaseosa resultante,

S-1
libre de productos líquidos, se analiza en un micro cromatógrafo de gases

S-1

S-1
VARIAN CP-4900. Este equipo permite el análisis del H2, N2, CH4, CO, CO2,

S-1
C2H4, C2H6, C3H8, C3H6 Y CH3OH. Los cromatogramas obtenidos son
almacenados y cuantificados por el propio software informático suministrado con
el equipo.

S-1
2. Electrocatálisis

En la Figura A.2 se muestra el esquema de la instalación experimental


donde se llevaron a cabo los experimentos de electrocatálisis en un reactor de
cámara sencilla (Figura A.3).

Línea
FIC
Líneacalefactada
calorifugada
N22
FIC

CO
CO22
FIC

H2/N
N 2/H22 FIC

N
H22/H
/N22 Controlador
Controlador
BROOKS INSTRUMENT
de
temperatura
temperatura
Reactor
Reactor 300
300 32

Controladores
300
300

Controlador de caudal
de caudal Horno
Venteo
Venteo

Medidor de
Caudalímetro
caudal Micro
Microcromatógrafo de
cromatógrafo
gases Potenciostato/
Galvanostato
PGZ 301

VoltaLab
cell

PC Potenciostato/Galvanostato

Figura A.2. Instalación experimental: electrocatálisis.

297
S-1
S-1

S-1
ANEXO: INSTALACIONES EXPERIMENTALES

Electrolito
1 mm sólido Electrodo de trabajo (WE)
K-βAl2O3
20 mm o
Na-βAl2O3 Contra-electrodo (CE) Termopar

Salida Entrada
Contra-electrodo Electrodo de trabajo
(Au) (Ni)
(PdZn) Refrigerante

Electrodo de
referencia (Au)

Tubo de
Tubo de cuarzo
alúmina con 4
perforaciones

Hilos de Au

Figura A.3. Esquema del reactor de cámara sencilla.

2.1. Sistema de alimentación

El sistema de alimentación es idéntico al de la instalación de catálisis


convencional, ya descrito en el apartado 1.1. de este anexo.

2.2. Sistema de reacción

El componente principal de la instalación es el reactor, que se


esquematiza en la Figura 3.

Este reactor es de tipo mezcla perfecta y consiste en un tubo de cuarzo


con un cabezal de acero en el que se encuentran las conducciones de entrada y
salida de los gases. Dentro del mismo, un tubo de cuarzo más fino conduce los
reactivos hacia el fondo del reactor donde se encuentra el catalizador (PdZn) o
electrodo de trabajo (W) soportado sobre un electrolito sólido K-βAl2O3 o Na-
βAl2O3 (Capítulo 7). Sobre el electrolito sólido además se encuentran el
contraelectrodo (C) y un electrodo de referencia (R) de oro inerte. Los tres
electrodos se exponen a la misma atmósfera de reacción al ser el reactor de
298
ANEXO: INSTALACIONES EXPERIMENTALES

cámara sencilla. Por ello, el contraelectrodo y electrodo de referencia


seleccionados, así como los hilos de oro, deben estar compuestos de un metal
catalíticamente inactivo en la reacción estudiada, con el objetivo de que no
interfiera en la actividad catalítica del sistema catalizador/electrodo de trabajo.
Los electrodos están conectados con un potenciostato mediante hilos de oro que
son conducidos hacia el exterior del reactor a través de un tubo de alúmina
perforado que impide el contacto entre los mismos. El catalizador electroquímico
se encuentra sujeto por unas pinzas de macor.

El calentamiento del reactor se lleva a cabo en un horno eléctrico cilíndrico


JH (modelo HCV 0611), con un termopar de control situado en la pared del horno
que está conectado a un controlador e indicador de temperatura CONATEC
CN300-P. La temperatura en el interior del reactor se mide con un termopar
THERMOCOAX tipo k, que se encuentra en el interior de una varilla de cuarzo a
la altura del catalizador. Las temperaturas registradas por este último se
registran con un dispositivo Pico technology TC-08 asistido por el software
PicoLog.

2.3. Sistema de polarización

El sistema de polarización consiste en un potenciostato/galvanostato,


modelo VoltaLab PGZ 301 de Radiometer Analytical, que es controlado por un
software avanzado que permite el tratamiento de los resultados (VoltaMaster
4).

En este trabajo el equipo se ha utilizado principalmente para la


realización de estudios potenciostáticos mediante la aplicación controlada de
potenciales. Además, permite llevar a cabo técnicas de caracterización
electroquímicas, como voltametrías lineales o cíclicas.

299
ANEXO: INSTALACIONES EXPERIMENTALES

2.4. Sistema de análisis

El efluente del sistema del reactor se conduce mediante una línea


calorifugada, para evitar la condensación del mismo, hacia un
microcromatógrafo de gases VARIAN CP-4900, que permite el análisis tanto de
los reactivos (N2, H2 y CO2) como de los productos detectados, los cuales fueron
CO y CH3OH. Los cromatogramas obtenidos son almacenados y cuantificados
por el software informático suministrado con el equipo.

300
ABOUT THE AUTHOR
ABOUT THE AUTHOR

Javier Díez Ramírez was born in Ciudad


Real in 1990. He grew up there, where he
studied a BSC in Chemical Engineering in
University of Castilla-La Mancha (UCLM,
Spain) from 2008 to 2013. He obtained
the “Extraordinary Award End of BSC in
Chemical Engineering” and the “REPSOL
Award to the best final research project”
in 2013.

He continued studying the MSC in Chemical Engineering in the same university


from 2013 to 2014. He obtained the “Extraordinary Award End of MSC in
Chemical Engineering” and the “CO2 Spanish Technology Platform Award to the
best Master final research project” in 2014.

He started his PhD in September of 2014 in the University of Castilla-La Mancha.


His PhD is related to methanol synthesis from CO2 and H2 in conventional catalysis
and electrocatalysis under the supervision of Prof. Paula Sánchez Paredes and
Prof. Fernando Dorado Fernández. He developed a stay of research for three
months in 2016 in the Centre for Research & Technology Hellas (Thessaloniki,
Greece) under the supervision of Prof. Michael Stoukides and Prof. George
Marnellos, where he could continue researching about the electrocatalytic
reduction of CO2 in a double chamber reactor.

He obtained the Second National Career Award in Chemical Engineering from


the Ministry of Education, Culture and Sport of Spain in 2017.

Thesis publications

1. J.Díez-Ramírez, J. A. Díaz, F.Dorado, P. Sánchez. Kinetics of the hydrogenation


of CO2 to methanol at atmospheric pressure using a Pd-Cu-Zn/SiC catalyst. Fuel
Processing Technology 173 (2018) 173-181

303
ABOUT THE AUTHOR

2. J.Díez-Ramírez, J. A. Díaz, P. Sánchez, F.Dorado. Optimization of the Pd/Cu


ratio in Pd-Cu-Zn/SiC catalysts for the CO2 hydrogenation to methanol at
atmospheric pressure. Journal of CO2 utilization 22 (2017) 71-80.

3. J. Díez-Ramírez, P. Sánchez, V. Kyriakou, S. Zafeiratos, G.E. Marnellos, M.


Konsolakis, F. Dorado. Effect of support nature on the cobalt-catalyzed CO2
hydrogenation. Journal of CO2 utilization 21 (2017) 562-571.

4. J. Díez-Ramírez, F. Dorado, A. R. de La Osa, J. L. Valverde, P. Sánchez.


Hydrogenation of CO2 to Methanol at Atmospheric Pressure over Cu/ZnO
Catalysts: Influence of the Calcination, Reduction, and Metal Loading.
Industrial & Engineering Chemistry Research 56 (2017) 1979-1987.

5. J. Díez-Ramírez, P. Sánchez, J. L. Valverde, F. Dorado. Electrochemical


promotion and characterization of PdZn alloy catalysts with K and Na ionic
conductors for pure gaseous CO2 hydrogenation. Journal of CO2 Utilization
16 (2016) 375-383.

6. J. Díez-Ramírez, P. Sánchez, A. Rodríguez-Gómez, J. L. Valverde, F.


Dorado. Carbon Nanofiber-Based Palladium/Zinc Catalysts for the
Hydrogenation of Carbon Dioxide to Methanol at Atmospheric Pressure.
Industrial & Engineering Chemistry Research 55 (2016) 3556-3567.

7. J. Díez-Ramírez, J. L. Valverde, P. Sánchez, F. Dorado. CO2 Hydrogenation


to Methanol at Atmospheric Pressure: Influence of the Preparation Method of
Pd/ZnO Catalyst. Catalysis Letters 146 (2015) 373-382.

Related publications

1. J. Díez-Ramírez, V.Kyriakou, I.Garagounis, A. Vourros, E. Vasileiou, P.


Sánchez, F. Dorado, M. Stoukides. Enhancement of Ammonia Synthesis on a
Co3Mo3N-Ag electrocatalyst in a K-βAl2O3 solid electrolyte cell. ACS
Sustainable Chemistry & Engineering. 5 (2017) 8844-8851.

304
ABOUT THE AUTHOR

2. A. R. de La Osa, A. Romero, J. Díez-Ramírez, J. L. Valverde, P. Sánchez.


Influence of a Zeolite-based cascade layer on Fischer-Tropsch Fuels Production
over Silicon Carbide Supported Cobalt Catalyst. Topics in Catalysis. 60
(2017) 1082-1093.

3. J. Díez-Ramírez, F. Dorado, Á. Martínez-Valiente, J. M. García-Vargas, P.


Sánchez. Kinetic, energetic and exergetic approach to the methane tri-
reforming process. International Journal of Hydrogen Energy 41 (2016)
19339-19348.

4. D. Lopez-Gonzalez, J. Couble, M. Aouine, L. Massin, P. Mascunan, J. Díez-


Ramírez, M. Klotz, C. Tardivat, P. Vernoux. Effect of the Reduction Step on the
Catalytic Performance of Pd–CeMO2 Based Catalysts (M = Gd, Zr) for
Propane Combustion. Topics in Catalysis 59 (2016) 1638-1650.

5. J. M. García-Vargas, J. L. Valverde, J. Díez, P. Sánchez, F. Dorado.


Catalytic and kinetic analysis of the methane tri-reforming over a Ni-Mg/β-
SiC catalyst. International Journal of Hydrogen Energy 40 (2015) 8677-
8687.

6. J. M. García-Vargas, J. L. Valverde, J. Díez, P. Sánchez, F. Dorado.


Preparation of Ni–Mg/β-SiC catalysts for the methane tri-reforming: Effect of
the order of metal impregnation. Applied Catalysis B: Environmental 164
(2015) 316-323.

7. J. M. García-Vargas, J. L. Valverde, J. Díez, P. Sánchez, F. Dorado.


Influence of alkaline and alkaline-earth cocations on the performance of
Ni/βSiC catalysts in the methane tri-reforming reaction. Applied Catalysis B:
Environmental 148-149 (2014) 322-329.

305
ABOUT THE AUTHOR

Conference contributions

-Oral presentations

1. Hydrogenation of CO2 to methanol at atmospheric pressure: optimization of


the molar percentage of Pd and Cu in PdCuZn/SiC catalysts. J. Díez-Ramírez,
P.Sánchez, E. Monge-Ruiz, J. A. Díaz, F. Dorado. EUROPACAT 2017 “Catalysis-
A bridge to the future”. Florence (Italy), August, 2017.

2. Electrochemical promotion of PdZn alloy catalysts with K and Na-βAl2O3 ionic


conductors for CO2 hydrogenation to CO and methanol. J. Díez-Ramírez, P.
Sánchez, J.A. Díaz, J. l. Valverde, F. Dorado. Secat 2017. Oviedo (Spain) June,
2017.

3. Activation of Pd-CeMO2 based catalysts (M=Gd, Zr) for propane combustion.


D. Lopez-Gonzalez, J. Couble, M. Aouine, L. Massin, P. Mascunan, J. Díez-Ramírez,
M. Klotz, C. Tardivat, P. Vernoux. 9th International Conference on Environmental
Catalysis (ICEC). Newcastle, Australia. 10-13 July, 2016.

4. Hydrogenation of CO2 to CH3OH: Influence of the type of nanofiber in the


formation of the PdZn alloy. J. Díez-Ramírez, F. Dorado, J. L. Valverde, P.
Sánchez. II Encuentro de Jóvenes Investigadores de la SECAT. Ciudad Real
(Spain). June, 2016.

5. Kinetic, energetic and exergetic approach to the methane tri-reforming


process. J. M. Garcia-Vargas, J. L. Valverde, J. Díez-Ramírez, A. Martínez-
Valiente, F. Dorado, P. Sanchez. Hyphothesis XI. Toledo (Spain). September,
2015.

6. Methanol synthesis from CO2 and H2O using conventional and electrochemical
catalysis. J. Díez-Ramírez, F. Dorado. P. Sánchez. IX Simposio de Ciencia Joven.
Ciudad Real (Spain). May, 2015.

306
ABOUT THE AUTHOR

7. Tri-reforming of methane: Converting CO2 and CH4 into valuable chemical


compounds. J. M. García-Vargas, J. L. Valverde, J. Díez, F. Dorado. P. Sánchez.
Energy and Environment Knowledge Week. Toledo (Spain). November, 2013.

- Poster presentations

1. Methanol: The “green fuel” of the future. J. Díez-Ramírez, F. Dorado, P.


Sánchez. VI Jornadas doctorales. Toledo (Spain). October, 2016.

3. CO2 Hydrogenation to methanol at atmospheric pressure: influence of the


preparation and the reaction conditions of Cu/ZnO catalysts. J. Díez-Ramírez, F.
Dorado, A.R. de la Osa, J. L. Valverde, P. Sánchez. CCESC 2016. Madrid (Spain)
September, 2016.

3. CO2 valorization to obtain methanol by electrocatalysis. J. Díez-Ramírez, F.


Dorado, J. L. Valverde, P. Sánchez. V Jornadas doctorales. Ciudad Real (Spain).
October, 2015.

4. Influence of the Preparation Method of Pd/ZnO Catalyst in the CO2


hydrogenation to methanol at atmospheric pressure. J. Díez-Ramírez, F. Dorado,
J. L. Valverde, P. Sánchez. SECAT 2015. Barcelona (Spain). July, 2015.

5. Methanol synthesis from H2 and CO2 at atmospheric pressure using the Pd/ZnO
catalyst. J. Díez-Ramírez, F. Dorado, J. L. Valverde, P. Sánchez. IV Jornadas
doctorales. Toledo (Spain) October, 2014.

6. Influence of the impregnation method on the performance of Ni/Mg/SiC


catalysts in the methane tri-reforming. J. Díez, J. M. García-Vargas, J. L.
Valverde, P. Sánchez, F. Dorado. I Encuentro de Jóvenes Investigadores de la
SECAT. Málaga (Spain). June, 2014.

7. Ni/Mg/SiC: Influence of the preparation method. I Workshop of Chemical


Engineering. J.Díez, J.M. García-Vargas, J.L.Valverde, P.Sánchez, F.Dorado.
Ciudad Real (Spain). November, 2013.

307
ABOUT THE AUTHOR

Participation in projects

European project: CO2 and H2O toward methanol at atmospheric pressure in


co-ionic electrochemical membrane reactors “GREEN MEOH”. CAPITA ERA-
NET seventh framework programme of the European Union ACE.07.016 (1
January 2014-31 December 2016).

Teaching experience

Assistant professor in laboratory practices (2014-2017):

- Mass and Energy Balances lab. Second year of the degree in Chemical
Engineering in the University of Castilla-La Mancha. 120 hours.
- Emerging renewable energy and environment technologies. Master in
Chemical Engineering in the University of Castilla-La Mancha. 12 hours.

Co-supervision of:

- Degree final project called “Pd/Zn catalyst synthesis over nanofibers for
CO2 hydrogenation”, which was developed by Alberto Rodríguez Gómez
(2015).
- Master final project called “Electrochemical promotion of PdZn catalysts
for CO2 hydrogenation to methanol”, which was developed by Paola
Anguita Fernandez (2016).
- Master final project called “Optimization of Pd-Cu-Zn catalyst for the
CO2 hydrogenation”, which was developed by Alberto Rodríguez Gomez
(2017).

308

You might also like