Tesis Doctorado Samantha Yadira Pinedo Hernández
Tesis Doctorado Samantha Yadira Pinedo Hernández
Tesis Doctorado Samantha Yadira Pinedo Hernández
FACULTAD DE QUÍMICA
TESIS
QUE PARA OBTENER EL GRADO DE:
P R E S E N T A:
DIRIGIDA POR:
1
Agradecimientos
2
A la Facultad de Química de la UAEM por haber facilitado los medios y herramientas
necesarias para llevar a cabo todas las actividades propuestas durante el desarrollo de
esta tesis.
Así como al Dr. Uvaldo Hernández B. del Laboratorio de Difracción de Rayos X de Polvos del
Centro Conjunto de Investigación en Química Sustentable UAEM-UNAM por el apoyo
brindado.
Al Dr. Óscar Fernando Olea Mejía del laboratorio de Microscopio Electrónico de Transmisión
del Centro Conjunto de Investigación en Química Sustentable UAEM-UNAM por el apoyo
brindado para las caracterizaciones de TEM.
3
Al Dr. Raúl Alberto Morales Luckie del Centro Conjunto de Investigación en Química
Sustentable UAEM-UNAM por el apoyo brindado en el material para las caracterizaciones de
TEM.
A los técnicos de los laboratorios de la Facultad de Química de la UAEM por todo el apoyo
brindado durante la estancia del doctorado.
4
ÍNDICE
pág.
RESUMEN………………………………………………………………………….......... x
ABSTRACT……………………………………………………………………………..... xi
Lista de Tablas…………………………………………………………………………… vii
Lista de Figuras…………………………………………………………………………... vii
Abreviaturas………………………………………………………………………………. viii
INTRODUCCIÓN……………………………………………………………………….. 16
1 ANTECEDENTES……………………………………………………………………….. 13
1.1 Contaminación del agua………………………………………………………….. 19
1.2 Colorantes………………………………….………………………………………. 20
1.3 Zeolitas………………………………………………………..…………………..... 22
1.3.1 Clinoptilolita………………………………………………………………. 24
1.4 Nanomateriales…………………………………………………………………….. 26
1.5 Sorción……………………………………………………………………………… 30
1.5.1 Cinética de sorción………………………………………………………. 30
1.5.1.1 Modelo cinético de pseudo primer orden (Lagergren)……… 31
1.5.1.2 Modelo de Elovich…………………………………………....... 32
1.5.1.3 Modelo cinético de pseudo segundo orden (Ho)…………… 33
1.5.2 Estudio de sorción en lote………………………………………………. 34
1.5.2.1 Modelo de Freundlich ……………………………….……….... 35
1.5.2.2 Modelo de Langmuir ………………………………..…………. 36
1.5.2.3 Modelo de Langmuir – Freundlich ……..…………………….. 37
1.5.2.4 Modelo Sips……………………………………………………. 37
1.6 Procesos de oxidación avanzada ……………..………………………………… 38
1.6.1 Proceso Fenton…………………………………………………………... 39
2 JUSTIFICACIÓN, HIPÓTESIS Y OBJETIVO …………………….………………….. 36
2.1 Justificación……………………………………………………………………….. 43
2.2 Hipótesis…………………………………………………………………….......... 43
2.3 Objetivo General………………………………………………………………..... 43
5
2.3.1 Objetivos Específicos……………………………………………………. 43
3 DESARROLLO EXPERIMENTAL…………….……………………………………….. 45
3.1 Obtención de los sorbentes…………………………………………………….. 46
3.2 Molienda y tamizado de los materiales ………………………………………. 46
3.3 Acondicionamiento y modificación del material zeolítico…………………… 46
3.4 Acondicionamiento de la zeolita sódica con hierro………………………….. 47
3.5 Sistemas nanoestructurados de Fe-Cu……………………………………….. 47
3.6 Obtención del composito de material zeolítico/Fe-Cu ……………………… 47
3.7 Caracterización de los materiales………………………….………………….. 48
3.7.1 Microscopía electrónica de barrido con microanálisis elemental…. 48
3.7.2 Análisis por activación neutrónica……………………………………….. 48
3.7.3 Espectroscopia de infrarrojo (IR)………………………………………… 48
3.7.4 Análisis termogravimétrico (TGA)……………………………………….. 48
3.7.5 Microscopía electrónica de transmisión (TEM)………………………… 48
3.7.6 Difracción de rayos X (XRD)…………………………………………… 49
3.7.7 Área superficial específica……………………………………………… 49
3.7.8 Determinación del punto de carga cero (pHZ) ……………………….. 49
3.8 Métodos Analíticos…….……………………..…………………………………. 50
3.8.1 Espectroscopia de UV / Vis.……………………………………………… 50
3.8.2 Determinación de la cantidad de H2O2 ……………………………….... 50
3.9 Cinética del proceso de adsorción de azul 1…………………………………. 50
3.10 Isoterma del proceso de adsorción de azul 1………………………………… 50
3.11 Efecto de la temperatura sobre la adsorción de azul 1……………………… 51
3.12 Efecto del pH sobre la adsorción azul 1………………………………………. 51
3.13 Efecto Fenton……………………………………………………………………. 51
3.14 Efecto Fenton-UV………………………………………………………………. 52
4 RESULTADOS Y DISCUSIÓN…………………………………...…………………… 53
4.1 Artículo publicado……………………………………………...……………...... 54
4.2 Articulo enviado……………………………………………………………….. 81
4.3 Discusión General……………………………………………………………….. 112
CONCLUSIONES...………………………………………………………..…………… 116
6
ANEXOS……...………………………………………………………………………….. 119
REFERENCIAS..……………………...………………………………………………… 125
LISTA DE FIGURAS
LISTA DE TABLAS
7
ABREVIATURAS
IR Espectroscopia de infrarroja
dispersa
nm Nanómetro
R2 Coeficiente de correlación
Ze Zeolita natural
8
Ze-Fe(Fe-Cu)PF Composito de hierro-cobre tratadas después del proceso Fenton
9
Resumen
10
RESUMEN
Una roca zeolítica modificada con hierro (Ze-Fe) y un composito de esa zeolita modificada y
nanopartículas de Fe-Cu (Ze-Fe (Fe-Cu)) se investigaron para la eliminación de azul 1. El
compuesto se sintetizó por reducción in situ de sales de Fe y Cu usando borohidruro de sodio.
Ambos materiales se caracterizaron por espectroscopía IR, BET, XRD, SEM y TEM. El análisis de
TEM demostró que las nanoestructuras de Fe-Cu, con un tamaño promedio entre 11 y 15 nm, se
dispersaron exitosamente en el material zeolítico.
En los experimentos de cinética de adsorción para los materiales se observó que al inicio, la
velocidad de adsorción es alta, esto es de 0 a 7 h de contacto; el equilibrio de adsorción se
alcanzó a las 72 h de contacto. Los experimentos por lotes mostraron que la adsorción del
colorante es más favorable para Ze-Fe (Fe-Cu) que para Ze-Fe, los datos de adsorción cinética
siguieron el modelo cinético de segundo orden. Los resultados también mostraron que la
eliminación del colorante fue mayor para ambos materiales a un pH entre 3 y 5. La eliminación del
azul 1(azul brillante) fue de 87.02% para Ze-Fe (Fe-Cu) y de 75.29% para Ze-Fe. Los valores de
ΔH ° de 52.60 kJ / mol para Ze-Fe y 126.29 kJ / mol para Ze-Fe (Fe-Cu) indicaron que los
procesos de adsorción son endotérmicos para ambos materiales.
La segunda etapa fue evaluar la remoción de azul 1 de una solución acuosa mediante un el
composito de la clinoptilolita modificado con nanopartículas de Fe-Cu, mediante los procesos de
oxidación avanzada (proceso Fenton con y sin radiación UV). Se preparó un composito en dos
etapas, la primera fue la obtención del material zeolitíco acondicionado con FeCl3, en condiciones
de reflujo, y la segunda fue la obtención del composito de la clinoptilolita con nanopartículas de
Fe-Cu, mediante la técnica de reducción in situ de sales metálicas en presencia del material
zeolítico.
Algunos factores como el pH de la solución, la cantidad de adsorbente, el tiempo de contacto, la
concentración inicial de la solución y la temperatura tuvieron un impacto en la eficiencia de la
11
adsorción. Los resultados muestran que la degradación del colorante fue similar de 97 a 99% y
casi 100% a un pH de 3, es importante mencionar que la acidez es un factor crucial para los
procesos de oxidación de Fenton. Los datos cinéticos se ajustaron mejor al modelo de segundo
orden para ambos materiales bajo procesos Fenton y foto-Fenton. Hubo ligeras diferencias en las
eficiencias de eliminación del colorante entre los dos materiales: 98.8% de degradación para Ze-
Fe.PF y 95.94% para el composito Ze-Fe(Fe-Cu). Sin embargo, el composito Ze-Fe(Fe-Cu)
muestra una mayor tasa de degradación, ya que el equilibrio se alcanzó en aproximadamente 20
horas y para la zeolita (Ze-Fe) en aproximadamente 50 horas. Ambos materiales muestran una
adsorción similar en el equilibrio de 0,99 y 0,96 mg / g para Ze-Fe y Ze-Fe (Fe-Cu),
respectivamente. Por lo tanto, la activación heterogénea de H2O2 fue el principal responsable del
proceso de degradación del colorante. De acuerdo con los parámetros obtenidos de los diferentes
modelos de isotermas, el mejor modelo que se ajusta a la adsorción del azul 1 por los materiales
fue el modelo de Freundlich, que indica que la adsorción se realiza en superficies heterogéneas.
12
Abstract
13
ABSTRACT
An iron-modified zeolitic rock (Ze-Fe) and a composite of that modified zeolite and Fe-Cu
nanoparticles (Ze-Fe (Fe-Cu)) were investigated for the elimination of blue 1. The compound was
synthesized by reduction of Fe and Cu salts using sodium borohydride. Both materials were
characterized by IR, BET, XRD, SEM and TEM spectroscopy. TEM showed that the Fe-Cu
nanostructures, with an average size between 11 and 15 nm, were successfully dispersed in the
zeolitic material.
The physicochemical characterization was carried out, the morphology of the samples was
analyzed with scanning electron microscopy (SEM), and the presence of typical crystals of
clinoptilolite was observed. The elemental chemical composition analysis was performed by X-ray
energy dispersion spectroscopy (EDS); the increase in iron was recorded after the zeolite was
conditioned. It was further observed that the specific area of the zeolite increased after the
treatments of 37.61 (m2 / g) in the natural zeolite to 220.34 (m2 / g) in the ferrous zeolite.
In the adsorption kinetics experiments for the materials it was observed that at the beginning, the
adsorption speed is high, this is from 0 to 7 h of contact; the adsorption equilibrium was reached
after 72 h of contact. The batch experiments showed that the adsorption of the dye is more
favorable for Ze-Fe (Fe-Cu) than for Ze-Fe, the kinetic adsorption data followed the second-order
kinetic model. The results also showed that the elimination of the dye was greater for both
materials at a pH between 3 and 5. The elimination of blue 1 (bright blue) was 87.02% for Ze-Fe
(Fe-Cu) and of 75.29% for Ze -Faith. The values of ΔH ° of 52.60 kJ / mol for Ze-Fe and 126.29 kJ
/ mol for Ze-Fe (Fe-Cu) indicated that the adsorption processes are endothermic for both
materials.
The second stage was to evaluate the removal of blue 1 from an aqueous solution by means of
the modified clinoptilolite composition with Fe / Cu nanoparticles, through the advanced oxidation
processes (Fenton process with and without UV radiation). A composite was prepared in two
stages, the first one was the obtaining of the zeolitic material conditioned with FeCl3, under reflux
conditions, and the second was the obtaining of the clinoptilolite composition with Fe-Cu
nanoparticles, by means of the in situ reduction technique of metal salts in the presence of the
zeolitic material.
14
Some factors such as the pH of the solution, the amount of adsorbent, the contact time, the initial
concentration of the solution and the temperature had an impact on the efficiency of the
adsorption. The results show that the degradation of the dye was similar from 97 to 99% and
almost 100% at a pH of 3; it is important to mention that the acidity is a crucial factor for the
oxidation processes of Fenton. The kinetic data were better adjusted to the second order model for
both materials under the Fenton and photo-Fenton processes. There were slight differences in the
elimination efficiencies of the dye between the two materials: 98.8% degradation for ZeFe.PF and
95.94% for the composition Ze-Fe (Fe-Cu). However, the composition Ze-Fe (Fe-Cu) shows a
higher rate of degradation, since the equilibrium was reached in about 20 hours and for the zeolite
(Ze-Fe) in about 50 hours. Both materials show a similar adsorption at the equilibrium of 0.99 and
0.96 mg / g for Ze-Fe and Ze-Fe (Fe-Cu), respectively. Therefore, the heterogeneous activation of
H2O2 was the main responsible for the degradation process of the dye. According to the
parameters obtained from the different isotherm models, the best model that adjusts to the
adsorption of blue 1 by the materials was the Freundlich model, which indicates that the
adsorption is carried out on heterogeneous surfaces.
15
INTRODUCCIÓN
La contaminación del agua es un problema a nivel mundial generado principalmente por las
actividades humanas
Las aguas residuales industriales a menudo han demostrado ser una seria amenaza para el
medio ambiente si no se tratan adecuadamente antes de las descargas. De manera similar, las
aguas residuales con color pueden degenerar el medio ambiente debido al alto contenido de
sustancias químicas, sólidos suspendidos y un color intenso muy visible, entre otras cosas.
Los colorantes son fuentes importantes de contaminación del agua y sus productos de
degradación pueden ser carcinógenos y tóxicos para los mamíferos. Se estima que
aproximadamente el 15% de la producción total de colorantes se pierde y se descarga en el
efluente durante la producción de colorante. Los colorantes azoicos constituyen la mayoría (60–
70%) de los colorantes aplicados en el procesamiento textil y se consideran recalcitrantes, no
biodegradables y persistentes .Se han investigado diversos métodos de tratamiento, como la
coagulación y la floculación, la adsorción y la ultrafiltración para eliminar los colorantes azoicos de
las aguas residuales. Estos procesos de alto costo no destruyen las moléculas de colorante, sino
que sólo las transfieren de una fase a otra .Los procesos de oxidación avanzada (AOP) son las
tecnologías más atractivas para el tratamiento de aguas residuales de colorante, capaces de
oxidarse rápidamente.
Las zeolitas se han caracterizado por tener una importante capacidad de eliminación de
contaminantes orgánicos. Por lo tanto, en este estudio se ha intentado optimizar clásicamente la
eliminación del color, utilizando nanoparticulas de Fe-Cu soportadas en zeolita.
16
En el primer capítulo se abordan los aspectos generales del trabajo tales como contaminación del
agua, materiales sorbentes, tipos tratamiento, modelos de sorción.
17
1
Antecedentes
18
1. ANTECEDENTES
El agua es uno de los recursos más importantes para los seres vivos, el agua tiene varias
propiedades importantes como medio de vida, como disolvente, en el comportamiento
ambiental y en usos industriales (Manahan, 2007).
Las aguas residuales de industrias textiles contienen colorantes, sólidos suspendidos, y otras
sustancias solubles tales como materia orgánica y metales pesados. Los colorantes y los
iones de metales pesados son importantes agentes contaminadores, causando problemas
ambientales y de salud para el ser humano (Wang y Ariyanto 2007). Los colorantes presentes
en el agua afectan la naturaleza del medio, ya que inhiben la penetración de la luz solar en la
corriente y favorecen la reducción de la fotosíntesis. Algunos colorantes son tóxicos e incluso
cancerígenos (Wang et al., 2005). La eliminación de colorantes y metales pesados de las
aguas residuales se pueden lograr por varias técnicas, por ejemplo la precipitación, la
floculación, la adsorción, el intercambio de iones, y la separación por membranas, sin
19
embargo es bien conocida la baja eficiencia de estos procesos. La adsorción se considera
como una de las técnicas más simples y rentables (Wang y Ariyanto 2007; Zanin et al., 2016).
1.2 Colorantes
Los colorantes suelen ser compuestos orgánicos que sirven para dar color a diversas
sustancias: fibras animales, vegetales o sintéticas y productos similares (tales como lana,
seda, algodón, lino, rayón, nylon, papel, etc.); u otros materiales como aceites, ceras o
plásticos. Los colorantes se pueden clasificar según su procedencia en naturales y sintéticos.
(Cubero et al., 2002; AVQTT, 2009; Lara, 2015).
Aunque actualmente resulta más factible obtenerlos por síntesis orgánica o inorgánica; en
este último, se emplean metales de transición interna, los cuales pueden existir en dos o más
estados de oxidación, de esta forma, el color depende del estado de oxidación del ión
metálico y del tipo y disposición de las demás moléculas que se unen a él. En los colorantes
sintéticos orgánicos, el color se debe a los cromóforos, que son secuencias de átomos unidos
por dobles enlaces ya sea en cadenas o anillos acoplados con una cadena lateral para que
haya resonancia y así se pueda impartir color (Pinedo,2010; Lara, 2015).
Los colorantes están constituidos por tres grupos funcionales: cromógeno, cromóforo y un
auxocromo. El cromógeno es la estructura aromática que contiene anillos de benceno,
naftaleno o antraceno. Un grupo cromóforo es aquél que proporciona el color. La estructura
cromogeno-cromóforo a menudo es capaz de propiciar la solubilidad y causar la adherencia
del colorante a la fibra. Por lo cual se requiere de grupos con afinidad de enlace llamados
auxócromos que aumentan la intensidad del color (Kirk-Othmer, 2001; Gutiérrez, 2009;
Trujillo, 2012; Lara, 2015).
.
La industria utiliza cerca de 10 000 diferentes colorantes, entre éstos, destaca el azul 1,
objeto de este estudio (Figura 1). Es uno de los colorantes con gran campo de aplicación en
la industria alimenticia, donde es utilizado como aditivo en la producción de gelatinas,
helados, confitería, productos lácteos, bebidas gaseosas, dulces, productos de pastelería,
productos de panadería, medicamentos (tabletas, cápsulas) y lociones para cabello. Además
de ser usado en la industria textil y del cuero (Kirk-Othmer, 2001; Mittal, 2006; Directiva,
20
2008; Panreac, 2008; Ni et al., 2009; Sigma Aldrich, 2009; EPA, 2009: C.I., 2009; Trujillo et
al., 2012).
21
En general, un sorbente puede ser considerado como de bajo costo si éste es abundante en
la naturaleza, requiere pocos procesos para su elaboración o si es un subproducto o material
residual industrial. Mediante esta investigación se determinarán las cualidades de un sorbente
de bajo costo, tales como son los materiales zeolíticos.
1.3. Zeolitas
Estos materiales poseen características tales como una adecuada área de contacto, gran
capacidad de intercambio iónico, efecto de tamizado molecular y naturaleza hidrófila que
favorecen su uso como adsorbentes. La fórmula general de las zeolitas es (Tsitsishvili et al.,
1992):
22
Mx/n [ AlxSiyO2(x+y)]pH2O……………1.1
En donde:
M = catión (Na+, K+, Li+) y/o (Ca2+, Mg+, Ba2+, Sr2+)
n = la carga del catión
x = número de átomos de aluminio
y = número de átomos de silicio
p = número de moléculas de agua
x/y = 1 a 6
p/x = 1 a 4
x+y= indica el número total de tetraedros de silicio y aluminio por celda unitaria.
Hay alrededor de 40 tipos conocidos de zeolitas naturales, sin embargo, sólo unos pocos son
ampliamente utilizados, incluida la clinoptilolita. Los estudios con clinoptilolita como
adsorbente demuestran su eficacia como material de eliminación de colorantes (Zanin et al.,
2016.
23
1.3.1 Clinoptilolita
Los iones Ca2+ y Na+, ocupan dos sitios (M1 y M2) en los canales A y B de la clinoptilolita. En
la estructura de la clinoptilolita, además hay dos sitios: M3 ocupados por K +, y M4 ocupados
por Mg2+.
El proceso de acondicionamiento se lleva a cabo cuando una zeolita se trata con un ion
específico en solución, por ejemplo Na+, con ello lo que se obtiene es el incremento del
contenido de este catión en la zeolita. El material resultante recibe el nombre de “homoiónico”
porque la mayoría de los cationes han sido intercambiados por un ion específico (Díaz, 2006).
24
Curkovic y colaboradores (1997) observaron que se favorece la homoionización cuando el
pretratamiento de la zeolita se lleva a cabo a altas temperaturas.
Las zeolitas naturales pueden ser modificadas por surfactantes catiónicos o mediante un
acondicionamiento (contacto entre una zeolita natural y una solución de un catión) para
convertir la zeolita a una forma homoiónica, de tal forma que no se pierde su capacidad de
sorber cationes. Durante el proceso de modificación, los cationes ya sea del surfactante o del
tipo de acondicionamiento seleccionado (una sal en solución), se intercambian
irreversiblemente con los cationes nativos como Na+, K+ o Ca+2 (Villalva, 2009).
El inconveniente para usar óxidos de hierro puros en polvo para remover colorantes están
asociadas a la dificultad de separar estos óxidos de hierro saturados con colorantes de la
fase acuosa. Otra alternativa es su uso en forma granular empacado en columnas, aunque
tiene la limitante de ser un material con un alto peso específico que dificulta su manejo si se
requiere fluidificar el lecho para su operación o limpieza (Rivera y Piña, 2008).
Debido a que los óxidos de hierro pueden representar una dificultad en el proceso de
acondicionamiento, algunas investigaciones han reportado que el uso del hierro en solución
es eficaz ya sea usando FeCl3+H2O2 (Reactivo de Fenton) o FeCl3. Torres y colaboradores
(2008) estudiaron la remoción del colorante amarillo remazol usando HDTMA
(hexadeciltrimetilamonio) para modificar la zeolita (clinoptilolita). Se logró observar que al
regenerar el material zeolítico saturado, usando FeCl3+H2O2 (Reactivo de Fenton) o FeCl3
que el surfactante (HDTMA) fue removido por el reactivo de Fenton, además de observarse
un aumento en la capacidad de sorción mayor al que originalmente tenía la zeolita.
25
Se observó que las zeolitas (clinoptilolitas) modificadas con FeCl3, incrementan el área
específica, propiciando condiciones adecuadas para la sorción del colorante anaranjado
remazol (Villalva, 2009). Otra aplicación es la de Gutiérrez y colaboradores (2009), quienes
observaron una remoción conveniente del colorante Índigo carmín, usando una zeolita
modificada con FeCl3.
1.4 Nanomateriales
26
Tabla 1.1 Métodos sintéticos para obtener materiales nanoestructurados
Métodos físicos Métodos químicos
Formación de nanopartículas a partir de Método de microemulsión
vapores Método de reducción química de sales
Método de Joule, calentamiento – metálicas
condensación en gas inerte Método electroquímico
Método de dispersión de átomos Método de sol-gel
solvatados Método de síntesis de nanopartículas
Método de pulverización con plasma y soportadas
condensación en gas inerte Método de síntesis de nanopartículas
Método de ablación láser empleando nanoreactores
El método de reducción química de sales metálicas, se considera uno de los métodos más
sencillos para obtener nanopartículas metálicas, se basa en la reducción de un precursor
metálico con agentes químicos bajo la protección de estabilizadores. Cuando se adiciona un
agente reductor a la solución de una sal metálica, se forman pequeñas partículas metálicas
(Trujillo, 2012).
Como es el caso particular también del sistema Fe-Cu que ha sido utilizado en sorción de
contaminantes orgánicos por Trujillo y colaboradores (2012b); donde obtuvieron
nanoestructuras bimetálicas de Fe–Cu y Fe–Ni y composito: material
carbonoso/nanopartículas de Fe–Cu y Fe–Ni, a partir del método de reducción de sales
metálicas y evaluaron las propiedades de adsorción de los materiales para remover amarillo
remazol de soluciones acuosas. Las capacidades de adsorción de las nanoparticulas Fe–Cu
y Fe–Ni fueron 157,8 mg / g y 117,6 mg / g, lo que resulta en casi el 83% y el 70% de la
eliminación del colorante, respectivamente, utilizando 100 mg / L de concentración inicial del
colorante y 10 mg de cada material. Las capacidades de adsorción de nanoestructuras
bimetálicas de Fe–Cu y Fe–Ni y composito: material carbonoso/nanopartículas de Fe–Cu y
Fe–Ni proporcionan mejores resultados a pH entre 3 y 5. Diversos autores han utilizado los
nananocatalizadores en la degradación de compuestos orgánicos recalcitrantes, en la Tabla
1.2 se muestran algunos estudios.
28
Tabla 1.2.Uso de nanocatálisis para degradación de varios compuestos orgánicos.
CATALIZADOR CONTAMINANTE REFERENCIA
La funcionalización de la superficie de las zeolitas con nanopartículas está siendo usada para
la degradación de diversos contaminantes (Garrido et al., 2010). En este contexto, el uso de
zeolitas modificada con Cu han sido utilizadas para la síntesis de amoníaco, en la oxidación
de p-xileno, de CH4, C3H6, CO, y metanol, en la descomposición de NO (Pérez R., González
M. and Bentrup U. 2012). El CuO se ha incorporado como nanopartículas sobre la zeolita
para la degradación de colorantes (Nezamzadeh-Ejhieh A. and Zabihi-Mobarakeh H., 2014).
La modificación de la superficie de las zeolitas con Fe también ha sido utilizada para la
adsorción de colorantes (Shahwana et al., 2011; Akgül 2014; Nairat, Shahwan et al., 2015)
para contaminantes orgánicos (Georgi et al., 2010.) e inorgánicos (Kim Ah. et al., 2013). La
zeolita modificada con FeO ha sido utilizada para la remoción de farmacos (Narges A. and
Nezamzadeh-Ejhieh A., 2015).
29
1.5 Sorción
En el proceso de sorción, el sorbato penetra en las cavidades y espacios libres entre cada
partícula de material sorbente, por lo tanto la cantidad que es retenida puede variar de un
material a otro hasta alcanzar el equilibrio cinético, esto va a depender principalmente de
variables tales como temperatura, concentración, pH, tipo de interacciones existentes entre el
material sorbente y las moléculas del contaminante adsorbidas (Gutiérrez, 2011). El análisis
cinético se puede realizar utilizando los siguientes modelos: dentro de los más empleados se
30
encuentran el modelo cinético de primer orden (Lagergren), el modelo cinético de Elovich y el
modelo cinético de pseudo segundo orden (Ho).
Si el proceso de adsorción sigue una cinética de pseudo primer orden, las constantes de
sorción de este modelo se obtienen por medio de la ecuación 1.1 (Torres et al., 2007; Kong X.
et al., 2016) conocida como ecuación de Lagergren, la cual se basa principalmente en la
capacidad de adsorción del sólido (adsorbente) (Lagergren, 1898 citado en Díaz, 2009).
Donde:
qe = Cantidad de soluto removido en el equilibrio por cantidad de material adsorbente
(mg/g).
qt = Cantidad de soluto removido en el tiempo t por cantidad de material adsorbente
(mg/g).
KL = Constante de velocidad en el equilibrio de una adsorción de primer orden (h-1).
t = Tiempo (h).
31
1.5.1.2 Modelo de Elovich
dqt
=αexp(-βqt )………1.3
dt
Donde:
qt = Es la cantidad de soluto removido en el tiempo t (mg/g)
α = Es la velocidad de sorción inicial (mg/(g*min))
β = Es la constante de desorción (g/mg)
Para simplificar la ecuación de Elovich, Chien y Clayton asumen que αβ>>1 y las condiciones
de frontera qt =0 hasta t=0 y qt = qt hasta t=t por lo que la ecuación queda como:
qt =βIn(αβ)+βIn(t)………1.4
32
1.5.1.3 Modelo cinético de pseudo segundo orden (Ho)
De 1996 a la fecha (Ho, 2006), la ecuación de pseudo segundo orden (ecuación 1.5) ha sido
considerada como uno de los modelos más apropiados para explicar diferentes sistemas de
adsorción.
t 1 1
= 2 + t………1.5
qt kqe qe
Donde:
qe = Cantidad de soluto removido en el equilibrio (mg/g)
qt = Cantidad de soluto removido en un tiempo t (mg/g)
k = Constante de velocidad en el equilibrio de una adsorción de pseudo-
segundo orden (g/mg*h)
En este modelo se puede observar que la velocidad de sorción está condicionada por la
capacidad de sorción en la fase sólida del adsorbente y no por la concentración del
adsorbato. Además de que la capacidad de sorción depende del tiempo, por lo que si se
conoce la capacidad de sorción del adsorbato en el equilibrio y la constante de sorción k,
entonces se puede calcular la capacidad de sorción en cualquier tiempo (t) graficando t
versus t/qt, obteniendo la pendiente de la recta igual a 1/qe y con la intercepción a la ordenada
1/(kqe2) podemos obtener la constante de velocidad (Kong X. et al., 2016).
33
1.5.2 Estudio de sorción en lote
(𝐶𝑜 − 𝐶𝑒 )𝑉
𝑞𝑒 = … … … 1.6
𝑀
(𝐶0 −𝐶𝑒 )
𝑅𝑒𝑚𝑜𝑣𝑎𝑙 (%) = 100 … … … 1.7
𝐶𝑜
Representa la forma más simple de una isoterma de sorción. Cuando la sección inicial de una
isoterma se comporta en forma lineal (ecuación 1.8), la pendiente representa el coeficiente de
distribución Kd (Fall, et. al., 2001). El coeficiente de distribución es la relación entre el
contenido de la sustancia en la fase sólida y la concentración de la sustancia en la fase
acuosa. Un valor bajo de Kd implica que la mayor parte del sorbato se encuentra en la
solución; mientras que un valor elevado de Kd indica que el sorbato tiene gran afinidad por el
sorbente (Alemayehu and Lennartz, 2009).
qe = Kd C𝑒 … … … … 1.8
34
Donde:
qe = Capacidad de sorción (mg/g)
Kd = Coeficiente de distribución (L/g)
Ce = Cantidad de soluto adsorbido y concentración de soluto en el equilibrio
(mg/L)
La ecuación (1.9) que describe este modelo relaciona la distribución de un soluto con los
coeficientes de actividad y refleja las interacciones intermoleculares del soluto en ambas
fases (Torres, 2007).
1⁄n
qe = Kf Ce … … … 1.9
1
log(qe )= log(Ce )+log(Kf ) … … … 1.10
n
Donde:
qe = Cantidad de adsorbato por unidad de peso del adsorbente (mg/g)
Kf = Constante de equilibrio que indica la capacidad de adsorción
n = Constante de adsorción, el reciproco es una medida de la
35
heterogeneidad de la superficie de adsorción ,indica la intensidad de
adsorción
Ce = Concentración del adsorbato en solución al equilibrio (mg/L)
El modelo de Langmuir está basado en las siguientes suposiciones (Venkata et al., 2002;
Volesky, 2003; Wang et al., 2005; Torres et al., 2007; Hor, K.Y., et al., 2016; Ghasemi M., et
al., 2016)
El modelo de Langmuir (la captación de iones en una superficie mediante la adsorción de una
sola capa, todos los sitios son idénticos y equivalentes desde el punto de vista energético, el
adsorbente es estructuralmente homogéneo sin interacción entre las moléculas adsorbidas en
los sitios vecinos) se expresó en la siguiente ecuación:
qo bCe
qe = … … … 1.11
1+bCe
Donde:
Ce = Concentración del soluto en la solución al equilibrio (mg/L)
qe = Cantidad de soluto adsorbido (mg/g)
b = Constante relacionada con la energía o la entalpía de adsorción
qo = Cantidad de soluto adsorbido por unidad de peso del adsorbente necesaria
para formar una monocapa en la superficie (mg/g)
Al linearizar la ecuación (1.10) se obtiene la ecuación 1.12 que permite conocer los
parámetros del modelo al graficar Ce versus Ce / qe, en la cual se obtiene una línea recta
con pendiente 1/ qo y la ordenada al origen corresponde a 1/ (qo* b):
36
Ce 1 1
= + ( ) Ce … … … 1.12
qe qo 𝑏 qo
1⁄n
KLF Ce
qe = 1⁄n
………1.13
1+aLF Ce
Donde:
𝑞𝑚 (𝐾𝑎 𝐶𝑒 )1/𝑛
𝑞𝑒 = ………1.14
1+(𝐾𝑎 𝐶𝑒 )1/𝑛
37
indicativo de la heterogeneidad del adsorbente de superficie, Ka es la constante de afinidad para la
adsorción (L / mg) 1 / n.
Los POAs son particularmente apropiados para aguas residuales que contienen materiales
no-biodegradables, recalcitrantes o tóxicos. Estos ofrecen algunas ventajas sobre los
procesos físicos y biológicos entre los que se encuentran:
De las tecnologías disponibles de oxidación, los procesos de Oxidación Avanzada han atraído
gran interés, ya que es una alternativa atractiva para el tratamiento de los contaminantes
orgánicos tóxicos presentes en aguas residuales; debido a su alta eficiencia para la
degradación de los contaminantes orgánicos persistentes en aguas residuales (Kuˇsi´c et al.,
2006; Zelmanov and Semiat 2008; Iurascu et al., 2009; Garrido et al., 2010; Huan et al.,
2012).
Los procesos de oxidación avanzada utilizan una combinación de fuertes oxidantes como el
ozono, el oxígeno, peróxido de hidrógeno, permanganato de potasio, cloro y ácido
hipocloroso, peróxido de hidrógeno con catalizadores. Se han utilizado catalizadores que van
desde sistemas homogéneos (uso de iones y complejos de Fe, Cu, Mo, Co, V y Cr,) hasta los
heterogéneos (Fe y Cu soportados en carbón activado, aluminatos, silicatos y titania) estos
tiene un alto estado de óxido-reducción (Garrido et al., 2010; Muñoz, 2011; Trujillo et al.,
2012; Bokare y Choi∗W, 2014.; Rahim et al., 2014) . El peróxido de hidrógeno es un oxidante
muy interesante dentro de los oxidantes convencionales más utilizados debido a que no
produce compuestos tóxicos, su uso no genera precipitados y se descompone en agua y
oxígeno que aumentan la biodegradabilidad del efluente. No es necesario un aporte de
energía para activar el peróxido de hidrógeno, el tiempo de reacción es corto entre todos los
procesos de oxidación avanzada; no hay limitación de transferencia de masa debido a su
naturaleza catalítica (Trujillo et al., 2012; Nidheesh and Gandhimathi, 2012).
39
Es un proceso que tiene la ventaja de producir simultáneamente la oxidación y coagulación
química; consiste entonces en la oxidación por combinación del peróxido de hidrógeno con
hierro (II) realizándose en medio ácido, que corresponde a la máxima relación de formación
de radicales libres en el sistema H2O2 / Fe2+ donde el hierro se utiliza como catalizador,
controlándose la concentración de peróxido de hidrógeno, la de hierro y el pH. El radical
hidroxilo ataca a los compuestos orgánicos, esto causa la descomposición de estos
compuestos consiguiéndose la eliminación del contaminante (Trujillo et al., 2011; Li B. and
Zhu, 2014 Wang et al., 2014).
Se han desarrollado catalizadores con partículas de tamaño nanométrico con una elevada
área superficial que puede acelerar la reacción Fenton sin necesidad de la radiación UV.
Estos nanocatalizadores son muy reactivos debido a que los sitios activos se encuentran en
la superficie. Como tales, tienen una resistencia a la baja difusión, y son fácilmente
accesibles, a las moléculas de sustrato (Garrido et al., 2010; Rusevova et al., 2012; Choi y
Lee 2012; Alireza Nezamzadeh-Ejhieh A. and Banan Z. (2012).).
En vista de los numerosos beneficios potenciales que se pueden obtener a través de su uso,
los catalizadores basados en Nanotecnología, han sido objeto de una atención considerable
en los últimos tiempos. En busca de nuevas alternativas para la remoción de contaminantes
orgánicos, diversas investigaciones están recurriendo al uso de oxidaciones tipo Fenton y el
uso de nanopartículas para las reacciones tipo Fenton.
40
Por su parte Gu y colaboradores (2013) realizaron un estudio sobre la preparación de carbón
activado basado en los lodos de aguas residuales para la adsorción de 2-naftol mediante el
uso de reactivo de Fenton, se observó que la capacidad de adsorción fue de 111,9 mg / g en
el carbono con 5% con pretratamiento de H2O2, mientras que sin ningún tratamiento el carbón
mostró una capacidad de sorción de sólo 51,5 mg / g. Wang y colaboradores (2013) lograron
la degradación del colorante verde malaquita utilizando extracto de té verde para sintetizar
nanopartículas de hierro. Kim y colaboradores (2013) lograron la degradación del 96% de Pb
en agua mediante un composito (Zeolita/Fe) teniendo una capacidad de 806 mg/g.
De esta manera, se espera que la nanotecnología sea una de las más importantes
tecnologías en este siglo ya que ofrece varias soluciones a problemas como salud, medio
ambiente, informática, entre muchas otras. Después de revisar la bibliografía se observa que
nuestra propuesta de utilizar un material zeolítico modificado para la obtención de un
composito de material zeolítico con nanopartícuals de Fe-Cu (zeolita/Fe-Cu), para la sorción-
Fenton, está sustentada en los antecedentes descritos, en los cuales no se reporta el uso de
este tipo de composito. Por todo lo anterior, es posible que se lleven a cabo fenómenos de
oxidación avanzada y de superficie activa en el composito, por lo cual, al combinar los dos
métodos (oxidación-sorción) se espera que la remoción del colorante sea mayor al aplicar el
material zeolítico modificado.
41
2
Justificación,
Hipótesis y
Objetivos.
42
2.1 Justificación
2.2 Hipótesis
43
3
Desarrollo Experimental
44
3. DESARROLLO EXPERIMENTAL
Etapa I
Efecto del pH
Análisis de resultados
Conclusiones
45
La metodología seguida para el desarrollo de este estudio se dividió en cuatro etapas. En la
primera se obtuvo el material zeolítico, se acondicionó y modificó para la obtención del
sistema nanoestructurado de Fe-Cu y el composito (Material zeolitico/Fe-Cu), y después se
procedió a su caracterización fisicoquímica.
En la tercera etapa se determinaron las isotermas de adsorción para obtener las capacidades
máximas de adsorción en el equilibrio.
Después se realizó el proceso Fenton donde se llevaron a cabo los procesos de remoción de
azul 1 con la adición de H2O2 para la adsorción-oxidación. Finalmente se analizaron los
resultados.
El material zeolítico empleado en este estudio se obtuvo de una mina de la zona de Villa de
Reyes en el Estado de San Luis Potosí, México.
Una vez que se realizó el tamizado de la zeolita natural, ésta se puso en contacto con 500 mL
de una solución 0.5 N de cloruro de sodio (NaCl) para favorecer el intercambio de iones Na+.
Dicho proceso se efectuó en condiciones de reflujo durante 3 h. Al término de este periodo se
46
separaron las fases y se agregaron otros 500 mL de NaCl a la zeolita, repitiendo el
procedimiento durante 3 h más. Al finalizar el tiempo de reflujo, se procedió a enfriar,
decantar y almacenar la solución de reflujo. La zeolita obtenida se lavó con agua destilada,
para lograr la eliminación de cloruro, lo cual se verificó con la prueba de nitrato de plata
(AgNO3 al 1%). Después de este proceso, se dejó secar la zeolita a temperatura ambiente.
Después de acondicionar la zeolita con NaCl; se procedió a acondicionarla con cloruro férrico.
Para esto se puso en contacto con 500 mL de una solución 0.1M de FeCl3, en condiciones de
reflujo por un total de 10 h, se dividió en periodos de 5 h cada uno; al finalizar el tiempo de
contacto, se realizaron lavados con 500 ml de agua destilada, hasta verificar la ausencia de
cloruros en la solución residual, posteriormente se secó la zeolita a temperatura ambiente.
47
minutos, luego se filtró y se lavó 3 veces con etanol absoluto. Finalmente, el material se secó
a 90 ° C durante 6 h.
El material zeolítico, el composito Ze-Fe(Fe-Cu), así como del sistema nanoestructurado Fe-
Cu fueron caracterizados por microscopía electrónica de barrido de alto vacío (MEB) y
microanálisis elemental por espectroscopia de rayos-X de energía dispersa (EDS) para
conocer la morfología y composición elemental de estas muestras, antes y después del
proceso de sorción. Todos estos análisis se realizaron en un microscopio electrónico (JEOL
8810LVelectron).
Las irradiaciones se realizaron en un reactor nuclear MARCA TRIGA III. Se usaron las
referencias de la IAEA para calcular las concentraciones elementales, y los datos nucleares
de los isótopos identificados en los espectros.
48
3.7.5 Microscopía electrónica de transmisión (TEM)
Los análisis se realizaron con un difractómetro SIEMENS D500 acoplado a un tubo de rayos-
X con un ánodo de Cu. Los difractogramas obtenidos se compararon con los patrones de
difracción de rayos-X. La difracción de rayos-X se emplea para obtener los patrones de
difracción de los minerales antes y después del proceso de sorción, así como obtener
información sobre su cristalinidad y su composición mineralógica.
El área superficial específica de los materiales sorbentes se determinó por el método de BET.
Se utilizó un equipo BELSORP; las muestras fueron calentadas a 100 ºC durante 2 horas
antes de la determinación del área específica de las mismas.
El punto de carga cero es el valor de pH para el cual la densidad de carga superficial neta es
cero. Este experimento se realizó pesando 75 mg de muestra en frascos ámbar, se agregaran
25 mL de solución de NaCl (0.01 M) a diferente pH (1-12), ajustados con solución de HCl y/o
NaOH (0.1 y 2.5 M). Se agitó durante 24 horas y transcurrido el tiempo se midió el pH final de
cada muestra. Para encontrar el pHz se graficaron los valores de pH iniciales contra los
finales, sobre una línea de pendiente igual a uno y el cruce de ambas líneas es el punto de
carga cero.
49
3.8 Métodos Analíticos
3.8.1 Espectroscopia de UV / Vis
Para determinar la cantidad de H2O2 que se utilizó en los procesos, y para evaluar la
influencia de esta dosis en el proceso de Fenton, se preparó una solución al 10% (v / v) de
H2O2 (Merck) a partir de un 30% (v / v) solución reactiva de H2O2 (Merck). Luego, 1, 2, 3, 4 y
5 mL de esta solución se pusieron en contacto con 9, 8, 7, 6 y 5 mL de una solución azul 1 de
10 mg / L, y esta mezcla se puso en contacto con 100 mg. de zeolita férrica (Ze-Fe) durante
72 horas a 25 ° C. Estos experimentos se realizaron por duplicado.
Este experimento relaciona la cantidad de contaminante sorbido por cada material vs.
concentración en el equilibrio, para ello se prepararon diluciones de azul 1 cuyas
concentraciones fueron de (10, 20, 40, 60, 80, 100, 120, 140, 160 and 180 mg/L, y 10 mL de
50
cada dilución se pusieron en contacto con 100 mg de material adsorbente (zeolita, composito
(Ze-Fe (Fe-Cu)). Fueron agitados a 120 rpm, los experimentos se realizaron por duplicado, el
tiempo de contacto fue de acuerdo a la cinética de sorción. Posteriormente transcurrido el
tiempo de contacto fueron separadas las fases, y se determinaron la concentración del azul 1
en solución por medio de espectrometría UV/Vis.
51
3.14 Efecto Fenton-UV
52
4
Resultados y Discusión
53
4. RESULTADOS Y DISCUSIÓN
54
55
Removal of brilliant blue by modified zeolitic materials
1
Universidad Autónoma del Estado de México, Facultad de Química. Paseo Colón esq. Paseo
Tollocan, S/N. C.P. 50180, Toluca, México;
2
Instituto Nacional de Investigaciones Nucleares, Departamento de Química. Carretera México-
Toluca S/N, km. 36.5. La Marquesa Ocoyoacac, Apartado Postal 18-1027, México DF. Tel. +52
5553297200x2271
Abstract
Ferric modified zeolitic tuff (Ze-Fe) and a composite of that modified zeolite bearing Fe-Cu
nanoparticles (Ze-Fe(Fe-Cu)) were investigated for the removal of brilliant blue. The composite
was synthesized by in-situ reduction of Fe and Cu salts using borohydride. Both materials were
characterized by BET, XRD, SEM, TEM, and IR spectroscopy. TEM demonstrated that
nanoestructures of Fe-Cu (11.70-15.85 nm) were successfully dispersed on the zeolitic tuff. Batch
experiments showed that the adsorption of dye is more favorable for Ze-Fe(Fe-Cu) than Ze-Fe, the
kinetic adsorption data followed the second-order kinetics model (R2 =0.94). The results also
showed that the removal of the dye was highest between pH 3 and 5 for both materials. The
removal of brilliant blue was 87.02 % for Ze-Fe(Fe-Cu) and 75.29 % for Ze-Fe. The ΔH° values of
52.60 kJ/mol) for Ze-Fe and 126.29 kJ/mol) for Ze-Fe(Fe-Cu) indicated that the adsorption
processes are endothermic for both materials.
Introduction
Nowadays, organic dyes are one of the major groups of pollutants found in wastewaters produced
from different industries. It is estimated that around 700,000 tons of dyes are produced annually
around the world; about 20% is discharged as industrial wastes without a previous treatment. The
chemical types of dyes utilized more frequently in the water industry are azo, anthraquionce, sulfur,
indigoid, triphenylmethane and phthalocyanine [1] (Nezamzadeh-Ejhieh A. and Banan Z., 2012).
The effluents arising from these industries cause great concern owing to their negative impact on
the environment. Treatment of residual effluents containing dyes have been a very active area of
research in recent times [2] (Rathinam A. et al., 2006). The brilliant blue Brilliant Blue(Azul
brillante FCF, Azul 1 FD&C, Food blue No. 1, Erioglaucina,E133 (dye), C.I. Acid blue 9,CI Food blue 2
,C.I.: 42090 ), molecular formula: C37H34O9N2S3Na2 , wt: 792.85 g/mol, lmax: 630 nm, pka: 5.83, 6.58). [3-
7] (Flury y Flühler, 1995; Mittal, 2006; Directiva, 2008; Ni et al., 2009; EPA, 2017)
(C37H34N2Na2O9–S3; M.W. =792.85) is a weak acid organic molecule with polar and non-polar
components, which may lead to a complex sorption behavior, its negative charge arises from
sulphonic acid groups, thus, depending on pH, the dye is either neutral or dissociated to a mono or
divalent anion. The anionic molecule can bind to anion exchange sites such as the edges of clay
minerals or hydrous oxide surfaces [8] (Morris et al., 2008).
56
The presence of dyes in water reduces light penetration which can in turn modify the
photosynthetic activity. Also, many dyes, or their metabolites, have toxic effects; carcinogenic,
mutagenic and teratogenic effects on aquatic life and humans [9] (Akgül, 2014). A wide range of
methods has been developed to remove dyestuffs from wastewaters, such as adsorption on organic
or inorganic matrices, chemical precipitation, flocculation and coagulation, oxidation by chlorine,
ozone electrolysis, electrochemical, microbiological treatments, etc. Some dyes are not completely
removed from wastewater because, most dyes are not amenable to common chemical, physical or
biological treatments due to their chemical stability, thus causing dangerous accumulation in the
environment [10,1,11] (Chen et al., 2010; Nezamzadeh-Ejhieh et al. ,2012; Arabpour et al.,
2015). Adsorption is one of the processes, which is widely used in wastewater treatments. The
removal of dyes and other organic pollutants from industrial wastewater is considered an important
application of the adsorption process using suitable adsorbents. Some of the main adsorbents used
in different processes are: silica gel, alumina, zeolites, activated carbon, sawdust, peat, lignite, red
mud [12] (Gupta and Suhas, 2009). Though activated carbon demonstrates high potential to
remove dyes, its high initial cost, coupled with problems associated with regeneration/reuse have
generate researches to look for other alternatives [2] (Rathinam et al., 2006). Studies with
clinoptilolite as adsorbent demonstrate its efficiency as a dye removal material [13] (Zanin et al.,
2016).
Among different supports, zeolites are considered to be important owing to their special properties
such as: unique nanoscale pores, thermal stability, hydrophobic and hydrophilic properties, eco-
friendly nature and ion exchange properties which could be used in efficient removal systems [1]
(Nezamzadeh-Ejhieh A. and Banan Z.,2012. Zeolites are crystalline aluminosilicates with cavities
in the range from one to several tens of nanometers depending on the type of aluminosilicates
frameworks, Si/ Al ratio and ion exchange cations; clinoptilolite is the most abundant zeolite in
nature. It has a monoclinic framework consisting of a ten-membered ring (7.5×3.1 Å) and two eight
member rings (4.6×3.6 Å, and 4.7×2.8 Å). [14] (Nezamzadeh-Ejhieh and Moeinirad 2011). The
three-dimensional crystal structure of zeolite contains two-dimensional channels which embody
some exchangeable cations such as Na+, K+, Ca2+ and Mg2+. These cations can be exchanged with
organic or inorganic cations from a solution [15] (Armagan et al., 2003)
It is well-known, through surface functionalization, the properties of the zeolite’s external surfaces
can be tailored for specific applications. These modifications can change the physicochemical
properties of the zeolites and the resulting modified materials can be used for a variety of
applications such as ion-exchange, adsorption, catalysis, etc. [9] (Akgül, 2014). Efforts have been
made to develop heterogeneous catalysts, such as Fe-treated laponite, iron exchanged zeolite, and
iron-loaded resin. Many researchers have reported that pillared clays intercalated with iron cations
have been used as active heterogeneous catalysts in decoloration and mineralization of dyes [10]
(Chen, et al., 2010)
Modified zeolites exchanged with metals such as Fe, Cu, Mn, Ca and Ba have been reported to be
more efficient under mild conditions (low temperatures and pressures) with less by-products.
Zeolites have open structures, which makes them capable of exchanging cations including iron.
Additionally, cations like iron and copper are not likely to leach out since they are strongly attached
to the zeolite network with chemical bonds and within the pores [16] (Haji S. et al., 2015)
Nanoparticles of metals and metal oxides have been extensively used as catalysts in many organic
reactions because of their high surface area and facile separation. Nanoparticles are intrinsically
unstable and often tend to form agglomerates to reduce the energy associated with the high surface
area/volume ratio because of their small sizes. Consequently, it is beneficial to develop the nano
sized particles on a matrix, in order to obtain a better dispersion of the particles. Recently, several
methods for the preparation of nanoparticles have been developed, such as thermal decomposition,
57
coprecipitation from solution and laser pyrolysis, microwave plasma, electrochemical synthesis,
sol–gel method, and mechanical activation, by using microemulsion and other methods [17] (Dutta
et al. ,2010)
Nanoscale iron particles are recently gaining great interest in environmental remediation, one of the
applications in this regard is the removal of organic pollutants from aqueous solutions. Iron
nanoparticles can be used as catalyst for the degradation of aqueous organic solutes because the
nano-scale size offers high surface area and reactivity. Iron nanoparticles can be readily synthesized
using sodium borohydride, as a reducing agent. Various chemical and physical methods are being
applied to obtain iron nanoparticles of specific sizes and morphologies; iron nanoparticles are being
synthesized in the presence of supporting inorganic material [18] (Shahwan et al. , 2011). In recent
years, the use of heterogeneous catalysts based on metal nanoparticles immobilized on materials
like natural zeolites has attracted great attention. The use of zeolites with metal nanoparticles have
produce, more active catalysts due to their pore sizes, which are comparable to the sizes of the
reacting molecules, and their significant surface areas. Therefore, they behave as selective
adsorbents as well as in-situ oxidation catalysts [16,19,20] (Haji et al., 2015; Legese Hailu et al.,
2015; Radomskii et al., 2015).
Previous studies have shown that a Fe/Cu bimetallic material was a promising adsorbent for the
removal of toxic organic matter from wastewater and it was proposed as a cost-effective
pretreatment process. Studies suggested that Cu planted on the iron surface (Fe/Cu bimetallic
particles) could accelerate the corrosion of Fe and the generation of [H+] in the absence of dissolved
oxygen (DO). However, the influence of DO was not clear and little information on the
performance under oxic conditions is known. DO can enhance the generation of hydroxyl ions and
improve the formation of iron hydroxides [21] (Sun et al., 2016). Recently, it has been found that
Fe/Cu bimetallic particles prepared by deposition of Cu° on the surface of Fe° and chemical
reduction method, can enhance rates of pollutants reduction due to the high potential difference
(0.78 V) between Cu and Fe [22] (Ren et al., 2016). Clinoptilolite is, in addition, a potential
heterogeneous catalyst, only a few studies concerning the application of Cu-containing
clinoptilolite in heterogeneous catalyzed reactions have been reported [23] (Pérez et al., 2012).
However, no studies have been reported on the application of Fe-Cu in heterogeneous catalyzed
reactions.
In the present work the application of a composite and a modified zeolite for the adsorption of a
dye is reported. The introduction of Fe-Cu nanoparticles to zeolites changes their properties of
adsorption. The brilliant blue adsorption behavior was evaluated with both a ferric zeolite materials
(Ze-Fe) and a composite (Ze-Fe (Fe-Cu)).
2. Material and methods
2.1 Materials
The clinoptilolite-rich tuff was obtained from Villa de Reyes, State of San Luis Potosí, Mexico, it
was milled and sieved. The grain size used was between 0.8 and 1.0 mm. The zeolitic (Ze) material
was treated with a hydrochloric solution and shacked for 4 h (50 g of material with 500 ml of 0.5 M
HCl solution). Afterwards, the zeolitic material was washed with distilled water until no presence
of chloride ions was observed in the washing solution tested with silver nitrate. The acid-treated
zeolitic tuff (Ze-HCl) was then dried at room temperature. Fe-zeolitic tuff (Ze-Fe) was prepared by
following the method reported elsewhere [24, 25] (Gutiérrez et al. 2009; Solache-Ríos et al.
2010) by refluxing 47.5 g of zeolitic tuff (Ze-HCl) with 500 ml of 0.1 M FeCl3·solution and left for
5 hours. This procedure was carried out twice with fresh solution. The weight of the zeolitic
58
material decreased after each conditioning; this behavior could be attributed to the removal of fine
particles during the processes.
2.2 Characterization
2.2.1 Scanning electron microscopy
The materials were mounted directly on the holders and then observed at 10 and 20 kV in a JEOL
JSM-5900-LD electron microscope. The microanalysis was done with an energy X-ray dispersive
spectroscopy (EDS) system.
2.2.3 IR spectroscopy
The IR spectra in the 4,000–500 cm−1 range were recorded at room temperature using an IR
Prestige-21Shimadzu. Samples were prepared following the standard KBr pellets method.
2.2.6. Zero charge point Aliquots of a 0.01 M NaCl solution were adjusted between pH 1 and 12 by
adding 0.1 M HCl or NaOH solutions. Aliquots were left in contact with each adsorbent for 24 h,
then the samples were centrifuged, decanted, and pH was analyzed in the remaining liquid phases
with a Hanna Instruments HI2550 Ph/ORP pH meter.
59
2.3. Adsorption kinetics
Kinetic removal of brilliant blue dye by the zeolitic tuff (Ze), acid-treated zeolitic tuff (Ze-HCl),
iron modified zeolitic tuff (Ze-Fe) and the composite Ze(Fe-Cu), was performed as follows:
samples of 100 mg of each adsorbent and 10 mL aliquots of a 10 mg/L solution (pH 5.8) were
placed in centrifuge tubes and shaken for different time intervals (5, 15 and 30 min, and 1, 3, 5, 7,
24, 48 and 72 h) at 120 rpm at 25°C. Later, the samples were centrifuged and decanted; the
experiments were carried out in duplicate. The brilliant blue dye concentrations in the solutions
were determined by using a UV/Vis Genesys 10S spectrophotometer analyzer, with λ=630 nm. The
pH of each solution was measured before and after the treatments.
(𝐶𝑜 −𝐶𝑒 )𝑉
𝑞𝑒 = (1)
𝑀
where qe (mg/g) is the equilibrium sorption capacity, Co and Ce are the initial and equilibrium
concentration (mg/L) of brilliant blue, respectively, V (L) is the volume and M (g) is the weight of
adsorbent.
The percentage removal (%) of dye was calculated using the following equation:
(𝐶0 −𝐶𝑒 )
𝑅𝑒𝑚𝑜𝑣𝑎𝑙 (%) = 100 (2)
𝐶𝑜
2.5 Effect of pH
In order to check the effect of pH on brilliant blue uptake by the different sorbent materials,
experiments were carried out by putting in contact 10 mL aliquots of a 10 mg/L solution of brilliant
blue at pH between 1 and 12 and 100 mg of each adsorbent. The pH values of the solutions were
adjusted by adding 0.1 M HCl or NaOH solution. The solutions were shaken for 72 h; the samples
were centrifuged and decanted. The pH was measured in the remaining liquid phases with a Hanna
Instruments HI2550 pH/ORP pH meter and dye concentrations were determined as above by UV-
Vis. The experiments were performed in duplicate.
3.1.1 Scanning electron microscopy (SEM) and transmission electron microscopy (TEM)
60
The morphology, particle size and elemental analysis of Ze-Fe and Ze-Fe (Fe-Cu) were studied by
scanning electron microscopy and the corresponding SEM photographs are presented in Fig. 1(a-
c).The crystallites of the zeolite have well defined cubic shapes, which are characteristic
morphologies of the clinoptilolite. [27] (Mumpton and Clayton, 1976). Fig. 1 shows the
composite with Fe-Cu nanoparticles embedded on the surface of the matrix. The Fe-Cu particles are
distributed as agglomerates; the surface is relatively rough in comparison with the original sample.
The EDS analysis was carried out in different areas for each adsorbent material (Ze, Ze-HCl, Ze-Fe
and Ze-Fe(Fe-Cu)). The main elements found of the adsorbent materials were O, Na, Mg, K, Ca,
Fe, Al, Si (Table 1), Na, Mg, K and Ca are the extra framework cations that compensate for the
deficiency of negative charge in the zeolite network[28,29] (D.W. Breck. 1974; Tsitsishvilli G.
,1992). Ca, Na and K concentrations decreases after treating the zeolitic tuff with FeCl 3. Treatment
with HCl removes impurities that block the pores, progressively the cations are exchanged by H+ to
get the H-form, and finally dealuminates the core zeolite structure creating secondary pores and the
pore size and surface area increase [30] (Motsa M.M.,et. al, 2011). Na and Ca atoms were replaced
by H+ or Fe since the content of this last element was highest in the Fe-zeolitic tuffs, and as a result,
it might have induced oxidation processes. [31] (Doula 2007). The presence of iron in the zeolite
may be located differently: as high spin Fe3+ in framework tetrahedral sites, in extraframework
octahedral sites as free Fe(H2O)6 3+, and as high spin Fe2+ in octahedral coordination in
extraframework sites or in another aluminosilicate sites associated with the zeolite. Iron is also
found as magnetite component in the zeolite rocks [32] (Xingu-Contreras et al.,2016) .Trgo and
Perić (2003) [33] have shown that the amphoteric nature of hydroxyl surface groups (=(Al/Si)–OH)
can lead to the formation of sites with different energies, this effect increases the number of
possible adsorption locations. It can be observed that the results of the EDS before and after the
adsorption experiments are similar; however, the content of Cu increases slightly. The
mineralogical composition persisted in the treated forms of clinoptilolite, confirming that the
conditioning mainly affects the concentrations of exchangeable ions without causing significant
alterations in the network of the zeolite; these findings are corroborated by the work done by [34]
Pinedo-Hernández (2007).
The morphology of the nanoparticles on the zeolite network was characterized by TEM. The TEM
image in Fig.1(g-i) confirms that the dimensions of the Fe-Cu particles remained in the nano
regime during the different treatments, the size distribution shows particles that vary between 11.70
nm and 15.85 nm and the TEM image reveals spherical particles (core-shell) as the main
morphology, connected in strings of pearls or agglomerates of different sizes. The formation of iron
oxide particles and copper oxide is confirmed from the diffraction measurements,indicate that the
iron oxide particles were deposited on the surface of the zeolite matrix. The TEM images Fig. 1(g-
i) shows the different sizes of the nano particles found.
3.1.2. IR Spectroscopy
The infrared spectra obtained from the samples of the natural zeolitic rock and the acidic zeolite
rock were similar, only slight changes in the percentage of transmittance were observed. The
zeolite samples chemically treated with iron shows no differences in the location of the vibrations
with respect to the unmodified material. The absorption peaks observed are assigned mainly to
asymmetric and symmetric stretching, which are the characteristics of this kind of material [23]
61
(Breck 1974). The IR spectrum of the different zeolitic materials shows bands at 1640 cm -1
corresponding to O-H bands of the adsorbed water molecules. The absorption band at 3400 cm-1
arises from stretching vibrations of O-H bands and increased considerably when the zeolitic
material was treated with Fe3+, this change could indicate an increment of the hydroxyl groups,
which means the formation of additional Brønsted acid sites. In addition, a broadening of the
absorption band at 1000–1100 cm-1 was observed due to the incorporation of Fe3+ into the zeolite
structure. Furthermore, peaks corresponding to typical frequencies of Fe-O (490, 599, 1370 and
1580 cm−1) were not observed, probably due to the overlap of these peaks with the characteristic
peaks of the zeolite. [9] (Akgül M. 2014). Other bands corresponding to the structure of the zeolite
were observed at 1040 cm-1, and 1089–1087 cm-1 related to Si–O–Si asymmetric stretching
vibration, Al–O–Si and A–O stretching vibrations. Also, band located in the ranges from 606 to
798 cm-1 and from 479 to 600 cm-1 are related to the stretching vibrations of zeolite framework.
Peaks located at about 1500 and 2000 cm-1 are from non-zeolitic impurities which disappeared after
successive acid-washing (with dilute HCl) of natural clinoptilolite [35,36] (Bahrami and
Nezamzadeh-Ejhieh 2014;Zabihi-Mobarakeh and Nezamzadeh-Ejhieh A. 2015). When a
transition metal cation enters a zeolitic network by ion exchange process, only slight changes occur
in the peaks located on the right side of the Si (Al)-O bands because the introduction of these
cations into the zeolite structure does not cause a significant change in the zeolite network [28]
(Breck 1974). The spectra are very similar to those reported in the literature [37,38] (Gutiérrez-
Segura et al., 2012, Montes-Luna et al., 2015), where slight changes in the bands due to the
modifications and high crystallinity of the zeolite was observed.
The BET specific surface areas for the untreated, acid, Fe modified zeolitic materials, and
composite (Ze-Fe (Fe-Cu)) samples were 37.6, 190.1, 220.3 and 120.1 m2/g, respectively. The
specific surface area of the material treated with HCl increases because HCl removes impurities
that block the pores [30] (Motsa M. et. al, 2011). The specific surface area increased when the
sample was treated with iron chloride and decreased with the addition of nanoparticles of Fe-Cu.
This suggests the presence of non-crystalline iron formations located at cationic positions in the
zeolite channels, as well as the formation of iron complexes outside the cell, which makes the
material more amorphous than the source material [31] (Doula, 2006). The specific surface areas of
this materials are higher than the values reported in the literature for this kind of adsorbents [24,39]
(Gutiérrez-Segura,2009; Jimenez- Cedillo et al, 2009).
62
3.1.5. X-ray diffraction (XRD)
Figure 3 shows the diffractograms corresponding to the natural zeolite clinoptilolite tuff (Ze), the
iron modified zeolite (Ze-Fe) and the iron modified zeolite treated with the Fe-Cu nanoparticles
(Ze-Fe (Fe-Fe)), the diffraction patterns obtained were compared with that of clinoptilolite (JCPDS
01-089-7538) and quartz (JCPDS 03-065-0466). The diffractograms obtained revealed the presence
of clinoptilolite and mordenite, similar results have been reported in the literature for clinoptilolite
type zeolites from other natural sources [29,24,38] (Tsitsishvilli et al., 1992; Gutierrez-Segura et
al.,2009; Montes-Luna et al., 2015). These results suggest that there were not any notable change
in the structure of clinoptilolite after it was treated with hydrochloric acid and iron chloride
solutions, since new peaks or displacements were not observed. The powder diffraction analysis of
natural clinoptilolite showed that clinoptilolite and quartz are the main mineral phases. The
mineralogical composition persisted after treatments, confirming that the conditioning mainly
affects the concentrations of interchangeable ions without causing significant structural alterations
within the tetrahedral nucleus; similar behaviors have been observed by Motsa et al. (2011) [30].
Changes in peak intensities are attributed to the different processes to which the zeolite rock was
subjected.
The XRD diffractograms of the zeolitic materials show characteristic peaks at 2Ɵ of 9.876, 11.184,
13.066, 17.357, 19.099, 22.357, 22.493, 23.213, 23.817, 25.061, 25.366, 26.057, 31.361 and 36.152
which are similar to the XRD data of crystalline structure of clinoptilolite corresponding to the data
[JCPDSNo.70-1859] and the literature [30,9,40]. (Motsa Machawe et al., 2011; Akgül M. (2014;
Sharafzadeh S. and Nezamzadeh-Ejhieh, 2015). The Ze-Fe (Fe-Cu) diffractogram indicates the
presence of Fe1.966O2.963 as shown in 2Ɵ in 11, 31, 35 which also correspond to those reported by
Nairat et al. ( 2015) [26] Also the diffraction peaks at 44.5 and 64.5 correspond to the presence of
Cu3Fe17 and indicate that it was deposited on the surface of the material. The average size of the
crystals was determined based on the width of XRD peak, the Bragg angle using the Scherrer
equation [41,40,42] (Bahrami M. and Nezamzadeh-Ejhieh A. 2014;Sharafzadeh S. and
Nezamzadeh-Ejhieh, 2015;Jafari S. and Nezamzadeh-Ejhieh A.2017):
where d is the mean diameter of the crystal, λ the wavelength of the X-rays (Cu Ka radiation l =
1.5406 A˚ used, β the line of the excess width of the diffraction peak in radians and Ɵ is the Bragg
angle in grade, and k is a constant, 0.89. The mean Ze-Fe (Fe-Cu) size was estimated to be about
19.12 nm for CuO and 46.40-87.04 for FeO.
Batch kinetic tests were performed with the aim of determining the time necessary to achieve
equilibrium between the adsorbate and the adsorbent. The kinetic results are presented as removal
percent of the brilliant blue from the liquid phase as a function of time. The dye removal kinetics of
Ze-Fe was compared with Ze-Fe(Fe-Cu) the results are shown in Fig 4. There was a significant
high adsorption in the first 5h for Ze-Fe(Fe-Cu) and 7 h for Ze-Fe and the adsorption rate decreased
63
up to equilibrium was researched. Equilibrium uptake of brilliant blue was attained in 72 h; this can
be attributed to the hydrophobic nature of the matrix, resulting in slow sorption; the percentage
removal at equilibrium was 75.29±0.90 and 87.02±2.40 % for Ze-Fe and Ze-Fe(Fe-Cu)
respectively. The adsorption capacities for Ze and Ze-HCl were very lower than 0.07 and 0.1 mg/g
respectively, then these materials were not considered for the other studies.
The adsorption in the first hours was faster until 3h this is due to a massive diffusion of the solution
to the surface of the adsorbent. Then a slower adsorption was observed where no significant
changes were observed after 24 h; where the adsorption was constant indicating that the
equilibrium was reached. The speed of adsorption was faster for the Ze-Fe (Fe-Cu) composite than
for Ze-Fe, in addition to achieving a greater adsorption capacity for the composite (Fig. 4a).
Y. Zeng, et al. 2016 [45] showed that the bimetallic nanocomposite system Cu-Fe exhibits higher
activity for nitrate reduction achieving 100% in 6 h compared to nanoparticles. Iron plays an
important role in the reduction process. In the bimetallic nanocomposite system Cu-Fe, the role of
Fe is as an electron donor; the second metal Cu acts as a promoter and thus increases the reactivity.
The presence of Cu in the bimetallic nano composite, at least at certain levels, can prevent iron
particles from being oxidized by O2, which can also enhance the reduction of contaminants. The
results indicate that the Ze-Fe(Fe-Cu) composite has faster dye removal kinetics and larger removal
capacity in comparison with Ze-Fe. This is attributed to the distribution of nano iron chains in the
clinoptilolite network, which make them kinetically more accessible for dye molecules, and
increase their surface area available for discoloration reaction. [26] (Nairat M. et. al, 2015). The
adsorption equilibrium time, coupled with a higher removal of Ze-Fe(Fe-Cu) in comparison with
Ze-Fe indicates higher affinity of nanoparticles for the dye, this has been observed in other studies.
The adsorption of dyes usually takes place in the mesopores [9] (Akgül M. 2014). Experimental
data were fitted to the pseudo-first-order [44,45] [(Mirzaei et al., 2015. Zanin, et. al, 2016.)],
second-order [43,44,45) (Ho Y.S.,2014; Mirzaei et al., 2015. Zanin, et. al, 2016), and pseudo-
second order [44,47] (Ho, 2006, N. Mirzaei et al., 2015) to evaluate the interaction mechanism
between the adsorbent-adsorbate complex to determine the kinetics parameters of the adsorption
process. The experimental results are shown in Fig. 4, the adjusted of data were performed by non-
linear regression using the program Origin 8.0. and the kinetic parameters obtained are shown in
Table 3:
The equations of the models are the following:
Pseudo-first order model (Langergren):
q t = q e (1 − ekl t ) (4)
Second order model (Elovich):
q t=1 In(1+abt) (5)
a
Pseudo second-order (Ho):
q2 kt
q t = 1+ eq (6)
e kt
where kL (h-1) is the pseudo first order rate constant, qe (mg/m) and qt(mg/g) are the adsorption
capacities at the equilibrium and time t (h) respectively, a and b (g/mg) are the adsorption and the
desorption constant, k (g/mg*h) relates to the constant of pseudo-second order adsorption. The
kinetic parameters and the coefficients of determination for each model are shown in Table 2. It is
evident that data were best fitted to second order model because it shows the highest correlation.
64
Table 2 shows the adsorption a (mg/g*h) and desorption b (g/mg) constants, and r obtained by
applying the Elovich model to the experimental data. This model has proven to be suitable for
highly heterogeneous systems. The adsorption of the dye with modified clinoptilolite is an example
of this case, since they are composed of different minerals and, therefore, have sites with different
energies for chemisorption. While the pseudo-first order kinetic model assumes that the rate of
solute uptake is directly proportional to the difference between the saturation concentration and the
amount of solids absorption over time. This model did not have a good agreement with
experimental data. The pseudo-first order kinetic model has the lowest R. Table 2 shows that the
pseudo-second order model provided a higher correlation coefficient than the pseudo-first order
model for both Ze-Fe and Ze-Fe (Fe-Cu).
The equilibrium between the dye adsorbed and retained in the aqueous solution is shown in Fig. 5.
It can be observed that at higher concentrations of brilliant blue, large amounts of dye are absorbed
and it can be seen that there is a high affinity of the adsorbents for the dye. The adsorption isotherm
models of Freundlich [46-50] (Panic VV and Velickovic SJ2014; Lin, et al, 2016; Huong Pham-Thi
et al., 2016), Langmuir[48,51] (K.Kalantari et al., 2014; Huong Pham-Thi, 2016) and the
Langmuir-Freundlich [52] (Moamen AOA et al., 2016), were used to treat the adsorption data of
brilliant blue by a nonlinear regression analysis using ORIGIN Pro 8.0 (for Windows). The
parameters obtained by the different models are shown in Table 3.
Freundlich
qe = Kf C1e ⁄n (7)
Langmuir
oq bCe
qe = 1+bC (8)
e
Langmuir–Freundlich
⁄
KLF Ce1 n
qe = ⁄ (9)
1+bLF C1e n
Where Kc is the adsorption equilibrium constant ((ml/g)), R is the gas constant (8.314 KJ/mol K ),
T is the absolute temperature and ΔG is the Gibbs free energy (KJ/mol). The enthalpy change ΔH
(KJ/mol) and entropy change ΔS (KJ/mol K) were obtained by calculating the slope and intercept
from plot of ln Kc versus 1/T and the data obtained are showed in Fig. 6 and Table 4. It can be seen
that the thermodynamic equilibrium constant, Kc, increased with temperature, which indicates that
sorption is an endothermic process and the values of ΔS and ΔH were both positive. The positive
value of ΔS ° suggests greater randomness at the solid/liquid interface in the sorption system and
increases throughout the sorption process (Fosso-Kankeu et al., 2017) [59].
The positive value ΔH ° for both materials indicated that the adsorption process is endothermic and
the anions of the dyes would preferably be attached to the adsorbent materials (Abdel Moamen
O.A. et al., .2016.) [60]. Also, the value of ΔH can give information about the adsorption
mechanism as chemical ion exchange or physical sorption. For the case of physical adsorption, ΔH
should be lower than 80 kJ / mol and while for chemical adsorption it ranges between 80 and 400
kJ / mol. [58] (Ghasemi M., et al., 2016.) As observed in Table 4, physical sorption process takes
place for the case of Ze-Fe and chemical sorption for the case of Ze-Fe (Fe-Cu)). This is because
the rate of intraparticle ion diffusion increases with increasing solution temperature. [54] (S.M. Al-
Jubouri et al., 2016). Positive ΔG values are observed for both materials.
66
3.2.4 Effect of pH
pH is one of the most important parameters in adsorption processes, especially due to its effects on
the loading of the adsorbents surfaces. Experiments were conducted in the pH range from 3 to 9.
Fig 7 shows the sorption efficiency of the brilliant blue at different pH conditions. At pH 7 the
sorption capacities were qe = 0.63 mg/g and 0.77 mg/g for Ze-Fe and Ze-Fe (Fe-Cu) respectively.
The efficiencies at pH 3 were highest qe = 0.78 mg/g (Ze-Fe) and 0.85 mg/g (Ze-Fe (Fe-Cu)) and
for pH 5 = 0.62 mg/g (Ze-Fe) and 0.78 mg/g (Ze-Fe (Fe-Cu)). A similar phenomenon was observed
in sorption of acid dye from aqueous solution by surfactant modified natural zeolites where the
sorption capacities decreased as the pH increased from 3 to 9. [44] (N. Mirzaei et al.,2015). Mao,
Y., et al. (2015) [61] investigated the efficiency of the degradation of soluble dyes over the pH
range 5.0–9.0 by using nanoscale zero-valent iron, the dye removal efficiency increased
significantly with decreasing pH. Also, Huong et al. (2016) [48] found that the pH of the solution
strongly influences the adsorption process between Fe-nano zeolite and nitrophenols; in acidic
conditions (pH 2-5), they achieved removal efficiencies of 95.6% for o-nitrophenol and 98.6% for
p-nitrophenol. The pH affects the charge of the adsorbents surfaces and ionic forms of contaminant
molecules in solution. The dye molecules appear to be protonated in acid solutions, especially
between pH 3 and 5. Therefore, the force between these protonated molecules and the surface of
the negatively charged adsorbents attracts the solutes to the surface of the adsorbent, resulting in a
greater sorption capacity. At pH 7, one might expect that the repulsive force between the surface of
the negatively charged adsorbent and the free electrons of the dye will decrease the sorption
capacity, however the hydroxyl ions can interact with the dye to carry out the sorption the dye. It is
observed that from pH 7 to 9 a plateau is formed, and there is not any significant change in the
sorption process.
Brilliant blue is a synthetic dye, azo type is an acid dye and contains negative sulfonic groups(-SO3)
the color part of the acid dye molecule is anionic. In acid conditions leads to an increase in the
concentration of H+ ions. On the other hand, a layer formed by the naoparticles on the zeolite
surface increases the positive charges on the external surface of the zeolite. Therefore, the strong
electrostatic attraction between the positively charged sorption site and the dye molecules leads to a
high adsorption capacity of the dye. Legese H. S., et. al. 2015 [62] reported that clinoptilolite as a
support for nanoparticles shows effectiveness as catalyst, showing that the maximum color removal
efficiency of 99 % was achieved using iron encapsulated in zeolite in acid pH.
According to the BET results, the nanoparticles provide a greater surface area in the
nanocomposite, compared to zeolite and increase the capacity of zeolites in the adsorption of
contaminants rapidly. This shows that most of the active adsorption sites of the nanoparticles can
be found outside the adsorbent and are easily accessible by the contaminating ion species, resulting
in rapid adsorption.
Conclusions
The adsorption of brilliant blue by Ze-Fe (Fe-Cu) and ferric modified zeolite was investigated in
batch system. The composition was synthesized by in situ reduction and characterized by SEM,
TEM, XRD and IR. The characterization of both materials by XRD reveals the presence of
clinotlilolite and the crystallinity after the different treatments. The specific area increases from
37.67 to 220.24 m2/g after treatments.Transmission electron microscope (TEM) confirmed that Fe-
Cu nanoparticles (11.70-15.85 nm) have been successfully loaded and efficiently dispersed on the
zeolite. The adsorption was faster using Ze-Fe(Fe-Cu) composite than Ze-Fe. The removal of
brilliant blue was 75.29±0.90 % and 87.02±2.40 % for Ze-Fe and Ze-Fe(Fe-Cu) respectively.
Adsorption isotherms of brilliant blue by Ze-Fe(Fe-Cu) composite and Ze-Fe showed that the Ze-
Fe(Fe-Cu) composite had a higher ability than Ze-Fe for the adsorption of the dye in aqueous
solution . It was observed that these materials are particularly effective in removing the dye and that
the removal of dye increases when the pH decreases, according to the results the pH affects the
surface charge of the adsorbent. The positive values of ΔH ° for both materials indicated that the
adsoprtion process are endothermic. The results provide knowledge for the development of novel
technologies using bimetallic composites to treat wastewater.
Acknowledgments
68
We acknowledge financial support from CONACYT, project 554061
References
2. Rathinam A., Nishtar F., Jonnalagadda Rao and Balachandran N. Wet oxidation of acid brown
dye by hydrogen peroxide using heterogeneous catalyst Mn-salen-Y zeolite: A potential catalyst
Journal of Hazardous Materials.B138 (2006) 152–159.
3. Flury M. and Flühler. 1995. “Tracer Characteristics of Brilliant Blue FCF.The Soil”. Science
Society of America Journal.Vol. 59. No 1. pp. 22-27.
4.Mittal A. 2006. “Use of hen feathers as potential adsorbent for the removal of a hazardous dye,
Brilliant Blue FCF, from wastewater.” Journal of Hazardous Materials. Vol B128 .pp. 233–239
6.Ni Y. ,Wang Y., Kokot S. 2009. “Simultaneous kinetic spectrophotometric analysis of five
synthetic food colorants with the aid of chemometrics”. Talanta Vol. 78. pp. 432–441.
8. Morris C., Mooney S.J. and Young S.D. Sorption and desorption characteristics of the dye tracer,
Brilliant Blue FCF, in sandy and clay soils. Geoderma.146 (2008) 434–438
10.Chen Q., Wu P., Dang Z., Zhu N., Li P., Wu J. and Wang X.. Iron pillared vermiculite as a
heterogeneous photo-Fenton catalyst for photocatalytic degradation of azo dye reactive brilliant
orange X-GN. Separation and Purification Technology .71 (2010) 315–323
12. V.K. Gupta and Suhas S, Application of low-cost adsorbents for dye removal—a review, J.
Environ. Manage. 90 (2009) 2313–2342.
13. Zanin E., Scapinello J., De Oliveira M., Rambo C., Franscescon F., Freitas L. ,Muneron de
Mello J.M., Fiori M. A., Vladimir de Oliveira J. and Dal Magro J. Adsorption of heavy metals from
wastewater graphic industry using clinoptilolite zeolite as adsorbent. Process Safety and
Environment Protectionhttp://dx.doi.org/10.1016/j.psep.2016.11.008
69
14.Nezamzadeh-Ejhieh A. and Moeinirad S. Heterogeneous photocatalytic degradation of furfural
using NiS-clinoptilolite zeolite Desalination. 273 (2011) 248–257
15.Armagan B., Ozdemir O., Turan M. and Celik MS. The removal of reactive azo dyes by natural
and modified zeolites. J Chem Technol Biotechnol 78 (2003) 725–732
16. Haji S., Khalaf M., Shukrallah M., Abdullah J. and Ahmed S. A kinetic comparative study of
azo dye decolorization by catalytic wet peroxide oxidation using Fe–Y zeolite/ H2O2 and
photooxidation using UV/H2O2. Reac Kinet Mech Cat. 114 (2015) 795–815.
17. Dutta B., Jana S., Bhattacharjee A., Gütlich P., Iijima Sei-Ichiro and Koner S. c-Fe2O3
nanoparticle in NaY-zeolite matrix: Preparation, characterization, and heterogeneous catalytic
epoxidation of olefins. Inorganica Chimica Acta. 363 (2010) 696–704
18. Shahwan T., Sirriah S. A., Nairat M., Boyacı E., Ero˘glu A.E., Scott T.B. and Hallam K.R.
Green synthesis of iron nanoparticles and their application as a Fenton-likecatalyst for the
degradation of aqueous cationic and anionic dyes. Chemical Engineering Journal. 172 (2011) 258–
266
19. Legese Hailu S., Unni Nair B., Redi-Abshiro M., Aravindhan R., Diaz I. and Tessema M.
Synthesis, characterization and catalytic application of zeolite based heterogeneous catalyst of
iron(III), nickel(II) and copper(II) salen complexes for oxidation of organic pollutants. J Porous
Mater. 22 (2015) 1363–1373.
20. Radomskii V. S., Astapova E. S. and Radomskii S. M. Structure and Thermal Properties of
Zeolites Modified with Fe and Cu Nanopowders. Inorganic Materials, 51 (2015) 999–1007.
21. L. Sun, H. Song, Q. Li, A. Li. Fe/Cu bimetallic catalysis for reductive degradation of
nitrobenzene under oxic conditions, Chem. Eng. J. 283 (2016) 366–374.
22.Y. Ren, Y. Yuan, B. Lai, Y. Zhou, J. Wang, Treatment of reverse osmosis (RO) concentrate by
the combined Fe/Cu/air and Fenton process (1stFe/Cu/air- Fenton-2nd Fe/Cu/air), J. Hazard. Mater.
302 (2016) 36–44.
23. Pérez R., Elizalde M.P. and Bentrup U.Preparation and in situ spectroscopic characterization of
Cu-clinoptilolite catalysts for the oxidative carbonylation of methanol. Microporous and
Mesoporous Materials. 164 (2012) 93–98.
24. Gutiérrez, S. E., Solache-Ríos, M., & Colin, A. Sorption of indigo carmine by a Fe-zeolitic tuff
and carbonaceous material from pirolyzed sewage sludge. Journal of Hazardous Materials, 170
(2009). 1227-1235.
27. Mumpton, F.A., Clayton, O.W., Morphology of zeolites in sedimentary rocks by scanning
electron microscopy. Clays Clay Min. 24, (1976). 1e23.
70
28.D.W. Breck. 1974. Zeolite Molecular Sieves, John Wiley & Sons, New York.
30. Motsa M.M.,Mambaa B.B., Thwala J.M. and Msagati T.A.M. 2011. Preparation,
characterization, and application of polypropylene–clinoptilolite composites for the selective
adsorption of lead from aqueous media. Journal of Colloid and Interface Science 359, 210–219
32. E. Xingu-Contreras G. , Garcıa-Rosales A., Cabral-Prieto I., Garcıa-Sosa .2016. Degradation of methyl
orange using iron boride nanoparticles supported in a natural zeolite. Environmental Nanotechnology,
Monitoring & Management. 7,121-129
33. Trgo, M., & Perić, J. Interaction of the zeolitic tuff with Zn-containing simulated pollutant
solutions. Journal of Colloid and Interface Science, 260, (2003). 166–175.
34. Pinedo-Hernández S., Díaz-Nava C. and Solache-Ríos M. Sorption Behavior of Brilliant Blue
FCF by a Fe-Zeolitic Tuff. Water Air Soil Pollut.7, (2007). 4-5.
35.Bahrami M. and Nezamzadeh-Ejhieh A. 2014. Effect of supporting and hybridizing of FeO and
ZnO semiconductors onto an Iranian clinoptilolite nano-particles and the effect of ZnO/FeO ratio in
the solar photodegradation of fish ponds waste water. Materials ScienceinSemiconductorProcessing
27,833–840.
36. Zabihi-Mobarakeh H. and Nezamzadeh-Ejhieh A.. Application of supported TiO2 onto Iranian
clinoptilolite nanoparticles in the photodegradation of mixture of aniline and 2, 4-dinitroaniline
aqueous solution. Journal of Industrial and Engineering Chemistry 26, (2015), 315–321.
39. Jimenez-Cedillo M.J., Olguin M.T. and Fall Ch.. Adsorption kinetic of arsenates as water
pollutant on iron, manganese and iron–manganese-modified clinoptilolite-rich tuffs. Journal of
Hazardous Materials 163 (2009) 939–945
71
ingredient for determination of anionic ascorbic acid species in presence of cationic dopamine
species. Electrochimica Acta 184 (2015) 371–380
42. Jafari S. and Nezamzadeh-Ejhieh A.20171. Supporting of coupled silver halides onto
clinoptilolite nanoparticles as simple method for increasing their photocatalytic activity in
heterogeneous photodegradation of mixture of 4-methoxy aniline and 4-chloro-3-nitro aniline.
Journal of Colloid and Interface Science 490 (2017) 478–487
43. Zeng Y., Walker H. and Zhu Q. (2016), Reduction of nitrate by NaY zeolite supported Fe,
Cu/Fe and Mn/Fe nanoparticles, J. Hazard. Mater. 324- B, 605-616
44. Mirzaeia N., Hadib M.,Gholamic M., Fouladi Fardd R., and Solaimany Aminabad M. (2015),
Sorption of acid dye by surfactant modificated natural zeolites, Journal of the Taiwan Institute of
Chemical Engineers.000,1-9.
45. Zanin, Evandro, Scapinello, Jaqueline, de Oliveira, Maickson, Rambo, Cassiano Lazarotto,
Franscescon, Francini, Freitas, Lucimaira, de Mello, Josiane Maria Muneron, Fiori, Marcio
Antonio, de Oliveira, Jos´e Vladimir, Dal Magro, Jacir, 2016. Adsorption of heavy metals from
wastewater graphic industry using clinoptilolite zeolite as adsorbent.Process Safety and
Environment Protection. 105,194-200.
46. Y.S. Ho, Using of pseudo-second-order model in adsorption, Environ. Sci. Pollut. Res. 21
(2014) 7234–7235.
47. Ho, Y. S. (2006). Review of second-order models for adsorption systems. Journal of Hazardous
Materials, 136 B, 681–689.
48. Huong Pham-Thi, Lee Byeong-Kyu, Kim J. and Lee Chi-Hyeon .(2016) Nitrophenols removal
from aqueous medium using Fe-nano mesoporous zeolite. Materials and Design 101, 210–217
49. Panic V. V. and Velickovic S. J.2014. Removal of model cationic dye by adsorption onto
poly(methacrylic acid)/zeolite hydrogel composites: Kinetics, equilibrium study and image
analysis. Separation and Purification Technology 122 ,384–394.
50. Lin, Lidan, Lin, Yan, Li, Chunjie, Wu, Deyi, Kong, Hainan, Synthesis of zeolite/hydrous metal
oxide composites from coal fly ash as efficient adsorbents for removal of methylene blue from
water, International Journal of Mineral Processing (2016), doi: 10.1016/j.minpro.2016.01.010
51. K.Kalantari et al., Rapid and high capacity adsorption of heavy metals by
Fe3O4/montmorillonite nanocomposite using response surface methodology:
Preparation,characterization,optimization,equilibrium isotherms, and adsorption kinetics study,
72
Journal of the Taiwan Institute of Chemical Engineers(2014),
http://dx.doi.org/10.1016/j.jtice.2014.10.025
52. Moamen A. O.A.,Ibrahim H.A., Abdelmonem N., and Ismail I.M.. 2016. Thermodynamic
analysis for the sorptive removal of cesium and strontium ions onto synthesized magnetic nano
zeolite, Microporous and Mesoporous Materials 223 ,187e195
53. Trujillo-Reyes J., Solache-Ríos M., Vilchis-Nestor A. R., Sánchez-Mendieta V. and Colín-
Cruz A. Fe–Ni Nanostructures and C/Fe–Ni Composites as Adsorbents for the Removal of a
Textile Dye from Aqueous Solution. Water Air Soil Pollut (2012) 223:1331–1341
54. S.M. Al-Jubouri, S.M. Holmes, Hierarchically Porous Zeolite X Composites for Manganese
Ion-exchange and Solidification: Equilibrium Isotherms, Kinetic and Thermodynamic Studies,
Chemical Engineering Journal (2016), doi: http://dx.doi.org/10.1016/j.cej.2016.09.081
55. Huang Y., Wang W., Feng Q. and Dong F. (2017). Preparation of magnetic
clinoptilolite/CoFe2O4 composites for removal of Sr2+ from aqueous solutions: Kinetic,
equilibrium, and thermodynamic studies. Journal of Saudi Chemical Society .21-1, 58-66
56. Akhtar M., Hasany S. M.,Bhanger M.I., and Iqbal S. (2007). Low cost sorbents for the removal
of methyl parathion pesticide from aqueous solutions. Chemosphere 66 1829–1838.
57. Z. Cao et al., The cationic dye removal by novel Si-Zn composites prepared from zinc ash,
Journal of the Taiwan Institute of Chemical Engineers (2016),
http://dx.doi.org/10.1016/j.jtice.2016.12.005
58.Ghasemi M., Javadian H., Ghasemi N., Agarwal S. and Gupta V. K. 2016. Microporous
nanocrystalline NaA zeolite prepared by microwave assisted hydrothermal method and
determination of kinetic, isotherm and thermodynamic parameters of the batch sorption of Ni (II).
Journal of Molecular Liquids .215, 161–169.
59. Fosso-Kankeu E., Mittal H., Waanders F. and Ray S. S. 20171. Thermodynamic properties
and adsorption behaviour of hydrogel nanocomposites for cadmium removal from mine effluents.
Journal of Industrial and Engineering Chemistry 48 (2017) 151–161
60. Abdel Moamen O.A. Ibrahim H.A. a, Abdelmonem b N., and Ismail I.M. 2016.
Thermodynamic analysis for the sorptive removal of cesium and strontium ions onto synthesized
magnetic nano zeolite. Microporous and Mesoporous Materials 223 (2016) 187e195
61. Mao, Y ,Xi Z. , Wang W., Ma C, Yue Q. (2015), Kinetics of Solvent Blue and Reactive Yellow
removal using microwave radiation in combination with nanoscale zero-valent iron, J. Environ. Sci.
XXX – XXX.
62.Legese H. S., Unni N. B., Redi-Abshiro M., Aravindhan R., Diaz I. and Tessema M..2015.
Synthesis, characterization and catalytic application of zeolite based heterogeneous catalyst of
iron(III), nickel(II) and copper(II) salen complexes for oxidation of organic pollutants. J Porous
Mater ,22:1363–1373
73
63.Nezamzadeh-Ejhieh A. and Zabihi-Mobarakeh H. 2014.Heterogeneous photodecolorization of
mixture of methylene blue and bromophenol blue using CuO-nano-clinoptilolite. Journal of
Industrial and Engineering Chemistry 20,1421–1431
Fig 1. Scanning electron microscopy image of a) Zeolite natural (Ze), b) zeolite treated with HCl,
c) zeolite (Ze-Fe), d) composite (Ze-Fe (Fe-Cu)), e) (Ze-Fe) after contact brilliant blue dye and f)
composite (Ze-Fe (Fe-Cu)) after contact brilliant blue dye ;(g, h, i) TEM of Fe-Cu nanoparticles.
Fig. 2 . IR spectrum of (a) Ze, (b) Ze-HCl, (c) Ze-Fe and (d) Ze-Fe (Fe-Cu)
Fig. 3 XRD pattern of zeolite natural (Ze), Zeolite ferric (Ze-Fe) and Ze-Fe(Fe-Cu) composite.
(C=Clinoptilolite, FeO= Maghemite, CuFe=Copper Iron)
Fig. 4. (a) Sorption kinetics of brilliant blue by Ze-Fe and composite (Ze-Fe(Fe-Cu)), (b). Ze-Fe
and (c) composite (Ze-Fe(Fe-Cu)) data adjusted to kinetic models.
Fig. 5. Isotherms of brilliant blue adsorption by Ze-Fe and composite (Ze-Fe(Fe-Cu)).
Fig. 6. lnKc vs. 1/T for the adsorption of brilliant blue by Ze-Fe and Ze-Fe(Fe-Cu).
Fig 7. pH effect on the adsorption of brilliant blue by Ze-Fe and Ze-Fe (Fe-Cu)
74
a) b) c)
d) e) f)
g) h) i)
20
18
16
14
12
Frecuency(%)
10
0
0 5 10 15 20 25 30 35 40
Particle size (nm)
Fig 1.
75
(d) 180
160
(c)
Transmittance (%)
140
(b)
120
(a)
100
80
60
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumber (cm )
Fig. 2 .
Fig. 3
76
a) 1.0
b) 1.0
c)
1.0 [ 0.9 [ 0.9
[
0.9
E 0.8 E 0.8 E
0.8
sc 0.7
sc 0.7
sc
0.7 0.6 0.6
ri ri ri
qt (mg/g)
qt (mg/g)
0.6 0.5 0.5
qt (mg/g)
0.5
b 0.4
b 0.4
b
Ze-Fe
0.4 a 0.3 Pseudo first model (Lagergren) a 0.3 Ze-Fe(Fe-Cu) a
Second order model (Elovich) Pseudo first order model (Lagergren)
Ze-Fe(Fe-Cu)
0.3 Ze-Fe u 0.2 Pseudo second order model (Ho) u 0.2 Second order model (Elovich) u
Pseudo second order model (Ho)
0.2 n 0.1
n 0.1
n
0.0 0.0
0.1
0 10 20 30 40 50 60 70
a 80
0 10 20 30 40 50 60 70 80 a 0 10 20 30 40 50 60 70 80 a
Time (h) Time (h)
Time (h) ci ci ci
ta ta ta
d d d
el el el
Fig. 4.
d d d
o o o
c 3.0 c c
u u u
m 2.5 m m
e e e
nt 2.0 nt nt
o o o
qe(mg/g)
o 1.5 o o
el el el
re 1.0 re re
su su su
m 0.5 m
Ze-Fe(Fe-Cu) m
e Ze-Fee e
n 0.0 n n
d 0 20 40 60 80 100 120
d140 160 d
Ce (mg/L) e
e e
u u u
n Fig. 5. n n
p p p
u u u
nt nt nt
o o o
in in in
te te te
re re re
sa sa sa
nt nt nt
e. e. e.
P P P
u u u
e e e
d d d
e e e
si si 77si
tu tu tu
ar ar ar
el el el
1/T (K-1)
0
0.00305 0.0031 0.00315 0.0032 0.00325 0.0033 0.00335
-1
-2
y = -15987x + 45.502
-3 R² = 0.9795
In Kc Ze-Fe
-4
Ze-Fe(Fe-Cu)
-5
-6
-7
y = -6326.9x + 14.438
R² = 0.8316
-8
Fig. 6.
1.00
0.95 Ze-Fe(Fe-Cu)
Ze-Fe
0.90
0.85
0.80
qe(mg/g)
0.75
0.70 Fig 7.
0.65
0.60
0.55
3 4 5 6 7 8 9
pH
78
Table 1. Elemental composition of the zeolitic materials obtained by SEM
Table 2. Kinetic parameters of the blue brilliant adsorption by Ze-Fe and composite (Ze-Fe(Fe-
Cu))
Material Kinetic models
Pseudo-first order Second order Pseudo-second order
qe (mg/g) KL (h-1) R2 a (mg/g*h) b (g/mg) R2 qe (mg/g) K R2
(g/mg*h)
Ze-Fe 0.72 0.36 0.67 14.57 12.34 0.94 0.74 1.17 0.77
Ze-Fe(Fe- 0.68 31.38 0.20 2648.78 18.28 0.94 0.69 68.14 0.42
Cu)
79
Table 3. Langmuir, Freundlich and Langmuir-Freundlich parameters for by Ze-Fe and composite
(Ze-Fe(Fe-Cu))
Material Sorption isoterms
Langmuir Freundlich Langmuir-Freundlich
qo (mg/g) b R2 KF n R2 KLF (L/g) aLF(mg/L) n R2
(L/mg) (mg/L)
Ze-Fe 3.44 0.01 0.77 0.25 2.23 0.89 Was not adjusted
Ze-Fe(Fe- 1.65 0.70 0.77 0.90 73.34 0.89 1.23 0.44 3.73 0.87
Cu)
80
4.2 Artículo enviado
de México, Paseo Colón y Tollocan s/n., Toluca, Estado de México, México. C.P. 50180.
b
Facultad de Química, Universidad Autónoma del Estado de México, Paseo Colón y Tollocan s/n.,
Abstract
Iron modified zeolitic tuff (Ze-Fe) and a composite (Ze-Fe(Fe-Cu)) in the presence of H2O2 and
with and without UV light were investigated for the removal of brilliant blue. The composite was
synthesized using the ferric modified zeolitic tuff and in-situ reduction of Fe and Cu salts via
81
sodium borohydride. The parameters considered in this study were pH, dose of H2O2, contact time
and initial concentration of brilliant blue. The results show that the degradation of the dye was
similar from 97 to 99% at a pH of 3 for the materials, it is important to mention that acidity is a
crucial factor for Fenton oxidation processes. The kinetic data were best adjusted to the second
order model for both materials under Fenton and photo-Fenton processes. There were slight
differences in the dye elimination efficiencies between the two materials: 98.8% of degradation for
(Ze-Fe) and 95.94% for Ze-Fe(Fe-Cu) composite. However, the Ze-Fe(Fe-Cu) composite shows a
higher degradation rate, since, the equilibrium was reached in about 20 hours and for zeolite (Ze-
Fe) in around 50 hours. Both materials exhibit similar adsorption at equilibrium 0.99 and 0.96 mg/g
for Ze-Fe and Ze-Fe(Fe-Cu), respectively. Therefore, the heterogeneous activation of H2O2 was the
main responsible for the dye degradation process. According to the parameters obtained from the
different isotherms models, the best model that fits the adsorption of brilliant blue by the materials
was the Freundlich model, which indicates that the adsorption is carried out on heterogeneous
surfaces.
1. Introduction
Water pollution is a major challenge due to the unregulated water discharges in aquatic
environments containing synthetic chemicals such as dyes, causing problems to humans and
since there is not any adequate color removal and mineralization of organic compounds (Thiam et
al., 2015; Karthikeyan et al., 2016). Azo dyes, which are approximately 70% of synthetic chemical
dyes, are characterized by the presence of one or more azo bonds (-N = N-) associated with
aromatic rings and auxochrome groups (eg, -OH and - SO3). Due to the relatively stable azo
82
structure, wastewater containing azo dyes has been an environmental concern in recent years (Cai
et al., 2017).
The adsorption process has been widely used in the removal of dyes; it is important to use materials
that are readily available, therefore, the adsorbents are usually modified to improve their adsorption
properties (Almazán-Sánchez et al., 2016). There is a wide range of adsorbent materials that have
been studied for the elimination of organic and inorganic contaminants from wastewater, such as
zeolites (Gutiérrez-Segura et al., 2009; Akgül, 2014; Humelnicu et al., 2017), nanomaterials (Cao et
al., 2017) and composites (Lara-Vásquez et al., 2016; Ahmed et al., 2017; Dinu et al., 2017). Some
studies have reported the adsorption and degradation of organic pollutants in aqueous solutions by
using iron modified aluminosilicates and hydrogen peroxide to improve their adsorptive properties
(Torres-Pérez et al., 2007). The modifying processes include acid and basic treatments, surfactant
impregnations (Díaz-Nava et al., 2009), or ion exchange using salts such as FeCl3 (Jiménez-Cedillo
et al., 2009; Gutiérrez-Segura et al., 2012) Several authors have described the oxidation of organic
matter by using sunlight and advanced oxidation processes (AOP) such as catalytic wet oxidation,
heterogeneous Fenton and photocatalysis by using TiO2 (Flores et al., 2016). Fenton reagents are
Advanced oxidation processes (AOPs) are used as alternatives for the elimination of toxic and non-
biodegradable compounds from water. These AOPs operate at room temperature and pressure and
can produce strongly oxidizing radical species such as •OH for the complete decomposition of
organic contaminants in non-toxic products such as CO2, H2O and inorganic salts. The
homogeneous Fenton process, which generates •OH radicals from Fe2+ and H2O2, is one of the most
common systems proposed for the treatment of organic pollutants (Hassan et al., 2016). However,
the main disadvantage of the homogeneous reaction of Fenton is the production of large amounts of
sludge that occur in the effluents and the narrow pH range (pH 2-3) is another limitation for its
83
application. In order to solve these problems, heterogeneous Fenton catalysts have been developed
and used as an alternative to these treatments (Wang et al., 2015, 2017). The Fenton heterogeneous
reaction can oxidize aqueous contaminants at a wider pH range and reduces the release of iron in
the water after treatment. When iron oxide and H2O2 are used, the efficiency to oxidize the organic
matter is low due to the low reactivity of the iron oxide to produce •OH radicals (Hassan et al.,
2016). Nano-sized iron particles have advantages over the micro size due to their large specific
surface area and high reactivity. Iron nanoparticles are good electron donors and dye molecules are
excellent electron acceptors. The iron nanoparticles in the aqueous medium and the hydroxyl,
hydrogen ions generated during the reduction process, react with the dye molecules to induce the
breakdown of the chromophore bond. The iron nanoparticles also have to break the auxochrome
bond in order to discolor the dye molecules and the resulting intermediate organic compounds
mineralize in CO2, H2O and inorganic ions to achieve a complete degradation (Raman and
Kanmani, 2016). Most of these catalysts show a lower catalytic activity than Fe2+ and they need the
Recently, nanometric materials have been considered as adsorbents given their high potential in
environmental remediation. One of the outstanding applications in this regard is the elimination of
organic and inorganic contaminants from aqueous solutions. It has been observed that the bimetal
particles of Fe-Cu are being used since they can reduce the pollutants rate degradation due to the
high potential difference that they present (0.78V) between Cu and Fe. The bimetallic iron-copper
catalyst system has attracted increasing attention as a catalyst in Fenton processes to treat
Fe-Cu and Fe-Ni nanoparticles were used to remove remazol yellow dye from water (Trujillo-
Reyes et al., 2012). In addition it has also been reported that Fe-Cu nanoparticles were supported on
carbon for the removal of phenol from wastewater (Pinedo-Hernández et al., 2017). H2O2 and
84
Fenton reagents with UV have been used and this system is one of the most common homogeneous
systems proposed for the treatment of textile wastewater. The use of heterogeneous solid Fenton
catalysts, such as iron modified zeolite can be an alternative, the iron salts are adsorbed on the
surface of a zeolite and it can react with H2O2 or with UV radiation, allowing iron ions to
participate in the Fenton catalytic cycle (Tekbaş et al., 2008). Nanoscale iron particles represent a
new generation of environmental remediation that could provide low cost and effective solutions
for one of the most difficult environmental cleanup problems. Adsorption is a promising technique
that is applied in the removal of organic and inorganic contaminants from water, and if it is coupled
with the use of nanoparticles, the removal capacity increases to eliminate contaminants and solve
environmental problems.
The aim of this research was to study and compare the adsorption and oxidation behavior of
brilliant blue 1 dye by Fenton process in presence and absence of UV light by using ferric zeolite
2.1 Materials
The zeolitic material was obtained from Villa de Reyes, San Luis Potosí, Mexico, it was milled and
sieved. The grain size used was between 0.8 and 1.0 mm. The zeolitic (Ze) material was treated
with a hydrochloric acid solution and shacked for 4 h (50 g of material with 500 ml of 0.5 M HCl
solution). Afterwards, the zeolitic material was washed with distilled water until no presence of
chloride ions, tested with silver nitrate, was observed in the washing solution. The acid-treated
zeolitic tuff (Ze-HCl) was then dried at room temperature. Fe-zeolitic tuff (Ze-Fe) was prepared by
following the method reported elsewhere (Gutiérrez-Segura et al., 2009); 47.5 g of zeolitic tuff (Ze-
HCl) were refluxed with 500 ml of 0.1 M FeCl3·solution and left for 5 hours, this procedure was
85
carried out twice with fresh solution. The weight of the zeolitic material decreased after each
conditioning; this behavior could be attributed to the removal of fine particles during the processes.
Iron modified zeolitic tuff (Ze-Fe) was prepared as follows: 47.5 g of zeolitic tuff (Ze-HCl) was
added to 500 mL of 0.1 M FeCl3·solution, and this mixture was refluxed for 5 h, this procedure was
performed twice, then the zeolitic tuff was separated and dried. Composites of iron and copper
nanoparticles supported on clinoptilolite, Ze-Fe(Fe-Cu) (70/30% wt% ratio) were prepared using
the reduction method (Nairat et al., 2015). 5.34 g of FeCl2·4H2O were dissolved in 25.0 mL of
ethanol/water (4:1) and mixed with 10 mL of a 0.2 M Cu(NO3)2 solution. Subsequently, 1.5 g of
Fe-zeolitic tuff (Ze-Fe) were added to the solution and mixed with a magnetic stirrer for 15 min.
NaBH4 solution was prepared separately by dissolving 2.5 g in 70.0 mL deionized water, then, it
was added to the mixture under continuous stirring at a constant addition rate of 0.5 mL/s. After
borohydride addition, the solution was kept under continuous stirring for another 15 min then
filtered and washed three times with absolute ethanol. Finally, the material was dried in the oven at
90 °C for 6 h.
2.3 Characterization
The materials were mounted directly on the holders and then observed at 10 and 20 kV in a JEOL
JSM-5900-LD electron microscope. The microanalysis was performed with an energy X-ray
86
Raman spectra were obtained using a confocal microprobe Raman system (HORIBA JBIN YVON,
Powder diffractograms of the materials were obtained with a Bruker D8 Advance diffractometer
coupled to a copper anode X-ray tube, operated with an accelerating voltage of 30 kV, with Bragg–
Bretano chamber, and current emission of 25 mA. The conventional diffractograms were compared
In order to determine the amount of H2O2 to be used in the processes, and to evaluate de influence
of the dose on the Fenton process, a 10% (v/v) solution of H2O2 (Merck) was prepared from a 30%
(v/v) reagent solution of H2O2 (Merck). Then, 1, 2, 3, 4 and 5 mL of this solution were put in
contact with 9, 8, 7, 6 and 5 mL of a 10 mg/L blue brilliant solution, and then with 100 mg of Ze-Fe
2.5 Effect of pH
The pH of the solution of brilliant blue was modified to 3, 4 and 5 by adding HCl or NaOH,
accordingly, considering the results of the H2O2 dose, 2 mL of this reagent together with 8 mL of
blue brilliant solutions at different pH were added to 100 mg of Ze-Fe and left for 72 h with
constant stirring.
Samples of 100 mg of each adsorbent (Ze-Fe and Ze(Fe-Cu) composite), 2 mL of 10% H2O2 and 8
mL aliquots of a 10 mg/L solution (pH 5) were placed in centrifuge tubes and shaken for different
time intervals (5, 15, and 30 min, and 1, 3, 5, 7, 24, 48 and 72 h) at 120 rpm and 25°C. Then, the
samples were centrifuged and decanted; the experiments were carried out in duplicate. The brilliant
blue concentrations in the solutions were determined by using a UV-Vis Genesis 10S
87
spectrophotometer analyzer, with λ = 630 nm. In addition, these series of experiments were
(𝐶𝑜 −𝐶𝑒 )𝑉
𝑞𝑒 = (1)
𝑀
where qe (mg/g) is the equilibrium adsorption capacity, C0 and Ce are the initial and equilibrium
concentrations (mg/L) of brilliant blue, respectively, V (L) is the volume, and M (g) is the mass of
the adsorbent.
The percent removal (%) of dye was calculated using the following equation:
(𝐶0 −𝐶𝑒 )
𝑅𝑒𝑚𝑜𝑣𝑎𝑙 (%) = x100 (2)
𝐶𝑜
Samples of 100 mg of each adsorbent (Ze-Fe and Ze-Fe(Fe-Cu) composite) were put in contact
with 2 mL of 10% H2O2 and 8 mL aliquots solutions (pH 5) of different concentrations of brilliant
blue (10, 20, 40, 60, 80, 100, 120, 140, 160, and 180 mg/L) for 48 h at room temperature, and the
samples were centrifuged and decanted. Dye concentrations were determined in the liquid phases as
described above, and the pH was measured in each solution. The experiments were performed in
duplicate. Also, these series of experiments were replicated with UV light using a 6 W lamp. The
brilliant blue dye concentrations in the solutions were determined by using a UV/Vis Genesis 10S
spectrophotometer analyzer, with λ=630 nm. The pH of each solution was measured before and
88
The materials Ze, Ze-HCl, Ze-Fe and Ze(Fe-Cu) composite were already characterized by scanning
electron microscopy, X-ray diffraction, specific surface areas and the points of cero charge were
determined and the results were reported elsewhere (Pinedo-Hernández et al., 2019).
SEM analyses were carried out to determine the morphology of the adsorbents before and after dye
removal (Figure 1). Micrograph of the natural zeolite (Ze) shows the typical morphology of a
sedimentary zeolite (Figure 1a), no changes were observed in the morphology of the zeolite tretated
with HCl and FeCl3 (Figure 1b and 1c), the results are similar to a previews work (Pinedo-
Hernández et al., 2019). Zeolitic tuff and the material treated with Fe show smooth surfaces (Figure
1d-1h). Micrographs of the zeolite after contact with brilliant blue dye (Ze-Fe)PF, (Ze–Fe)PF-UV,
The TEM image of Fe-Cu nanoparticles (Figure 1i) indicated that the nanoparticles were mostly
spherical and formed aggregates, also the Fe-Cu nanoparticles show a normal distribution and the
average particle diameter was 13.72.93 nm. These measurements were in agreement with those
found in the literature. Some studies reported the synthesis of Fe-Cu particles with a diameter
between 5 and 50 nm (Lara-Vásquez et al., 2016). Previous works (Pinedo-Hernández et al., 2019)
Table 1 shows the elemental composition determined by EDS of the materials after Fenton or UV
light processes treatments. The elemental composition of the materials is similar considering the
standard deviations.
Raman spectroscopy is a reliable technique and has been applied to identify the nanomaterials
synthesized onto zeolites. The Raman spectra of Ze-Fe(Fe-Cu) composite (Figure 2) shows bands
89
around 336, 502, and 677 cm-1 corresponding to -Fe2O3 (Soler and Qu, 2012) and a broad band at
The samples after the Fenton process and treated with UV light (FP, FP-UV) were analyzed by
XRD. The diffractograms obtained were compared with the standards of the Committee of Powder
Diffraction Policies (JCPDS), the samples showed diffraction peaks corresponding to clinoptilolite
(JCPDS 71-1425), mordenite (JCPDS 11-0155), quartz (JCPDS 01-0649 ) and iron oxide hydrate
(JCPDS 01-0662). The differences found among the samples were the intensities of the diffraction
peaks, the results show that there were not any changes in their structures after Fenton and UV light
processes.
Different volumes of H2O2 solution were employed in the degradation of blue brilliant by Ze-Fe to
optimize the amount of H2O2 to be used in the Fenton process. The results show that the
degradation of the dye by the materials is similar from 97 to 99 % in this concentration range.
Besides this, several studies have reported that the degradation of dyes increases as the
concentration of H2O2 increases; however, at high concentrations the degradation may decrease, the
chemical species formed are responsible for this behavior (Cai et al., 2017). Taking into account
Chemical oxidation of organic compounds is a powerful method used for the degradation of
pollutants in water. The oxidation process is based on the hydroxyl radical (OH•) action, which is
generated in aqueous solution by the well-known Fenton reagent, a combination of Fe2+ and H2O2.
In a classical Fenton system, the hydroxyl-radical formation cycle can be depicted as:
90
Equations (3) and (4) show that the reaction is started by the ferrous ions, which leads to the
production of hydroxyl radicals. These radicals react with the organic pollutant molecules and these
last compounds are degraded. The iron oxides, the iron oxo-hydroxide and the zero-valence iron
can be employed as sources of ferrous ions in a similar process to the Fenton one. These materials
have been used to catalyze the degradation of pollutants such as organic dyes and other organic
contaminants.
Recently, the synergistic effect of the Fe2+/Fe3+ and Cu2+ ions has been observed in the Fenton
processes degradation of some aromatic contaminants. The presence of small amounts of Cu 2+ ion
as a cocatalyst promotes a faster mineralization attributed to the contribution of the Cu2+ / Cu+ pair
to the production of • OH. For this, first Cu2+ is reduced to Cu+ on the reaction surface (5) or with
H2O2 giving rise to the weakest oxidative hydroperoxyl radical (HO2•) reaction (6). From then on,
the Cu+ behaves like a catalyst type "Fenton" and the "•OH" can be generated through the reaction
(7). In addition, Cu+ can contribute to the regeneration of Fe2+ reaction (8) (Garcia-Segura et al.,
2016).
3.6 pH effect
degradation of the dye by Ze-Fe was almost 100% at a pH of 3 (Figure not shown). It is important
to mention that acidity is a crucial factor for Fenton oxidation processes. In an acid solution, the
iron oxide/oxo-hydroxide surface is corroded yielding ferrous ions, which generate •OH radicals.
91
These radicals can react with dye’s molecules that are in solution or attached to the catalyst surface;
for the azo type dyes, the scission of the azo moiety in the chromophore leads to the solution
discoloration (Shahwan et al., 2011). Thus, it is expected that the dye could be degraded in high
acidity medium. This phenomenon was also observed previously where the degradation of Orange
G reached almost 100% in about 30 min (Cai et al., 2017). Some studies suggest that at higher
acidity the hydroxyl moieties in this dye can be protonated, and then, the repulsion force among
these substituents and the positive charges in the catalyst surface, preclude the dye’s molecules
from reaching the catalyst surface places, where the hydroxyl radicals are formed; therefore,
degradation of the dye diminishes (Bahrami and Nezamzadeh-Ejhieh, 2014). According with these
Experimental data were fitted to the pseudo-first-order (Zanin et al., 2017), second-order (Jiménez-
Cedillo et al., 2009), and pseudo-second order (Ho, 2006) models to evaluate the interaction
mechanism between the adsorbent-adsorbate and to determine the kinetics parameters of the
adsorption process. The experimental results are shown in Figure 4, the fitting of data was
performed by non-linear regression using the program Origin 8.0 and the kinetic parameters
q t = 𝑞𝑒 (1 − ekl t ) (9)
92
𝑞 2 kt
q t = 1+𝑒𝑞 (11)
𝑒 kt
where kl (h-1) is the pseudo first order rate constant, qe (mg/g) and qt (mg/g) are the adsorption
capacities at the equilibrium and time t (h), respectively, a and b are the adsorption and the
desorption constant, k (g/mgˑh) relates to the constant of pseudo-second order adsorption. The
kinetic parameters and the coefficients of determination for each model are shown in Table 2. It is
evident that data were best fitted to the second order model for both materials under Fenton process
because it shows the highest correlation. Table 2 shows the adsorption a (mg/gˑh) and desorption b
(g/mg) constants, and R2 obtained by applying the Elovich model to the experimental data.
Similarly, data were best fitted to second order model for the photo-Fenton process in both
materials.
There were slight differences in the dye elimination efficiencies between the two materials; 98.88%
of degradation was observed for (Ze-Fe)FP and 95.94% for Ze-Fe(Fe-Cu)FP composite. However,
the Ze-Fe(Fe-Cu)FP composite shows a higher degradation rate, the equilibrium was reached in
about 20 hours and for zeolite (Ze-Fe)FP in about 50 hours. Both materials exhibit similar
adsorption at equilibrium 0.99 mg/g for (Ze-Fe)FP and 0.96 mg/g for (Ze-Fe(Fe-Cu))FP composite.
Therefore, it can be said that the heterogeneous activation of H2O2, throughout superficial
reactions, was the main responsible for the dye degradation process.
The use of iron salts precursors in presence of H2O2 generates hydroxyl radicals under acidic
conditions, and this will degraded the original dye molecules. Nonetheless, as the hydroxyl radical
has a very short lifetime, the rate of reaction control capacity is limited using those salts. Thus, the
nanoparticles will bring a constant rate of reaction throughout the time until the use of H 2O2 or the
reaction conditions change (i.e. pH changes) (Truskewycz et al., 2016). The removal behavior of
brilliant blue in the presence and absence of UV light is similar as shown in Figure 4, 5 and Table
2.
93
3.8 Adsorption isotherms
The adsorption isotherm models of Langmuir, Freundlich, and Sips were used to treat the
adsorption data of brilliant blue by a nonlinear regression analysis using ORIGIN Pro 8.0 (for
Windows). The parameters obtained by the different models are shown in Table 3.
Langmuir isotherm model (Gao et al., 2016; Dinu et al., 2017) is represented by Eq. (12):
q 𝐾 C
m 𝐿 e
qe = 1+𝐾 (12)
C𝐿 e
The Freundlich model (Gutiérrez-Segura et al., 2009; Gao et al., 2016) is expressed by Eq. (13):
qe = KF C1e ⁄n (13)
The Sips isotherm (Langmuir-Freundlich model) (Corral-Capulin et al., 2019) is represented by Eq.
(14):
𝑞𝑚 (𝐾𝑎 𝐶𝑒 )1/𝑛
𝑞𝑒 = (14)
1+(𝐾𝑎 𝐶𝑒 )1/𝑛
where qe is the adsorption capacity at equilibrium (mg/g), Ce is the dye concentration at equilibrium
heterogeneity of surface adsorbent, Ka is the affinity constant for the adsorption (L/mg)1/n.
According to the parameters obtained from the different isotherms models, and R2, the model that
best fits the process of adsorption of brilliant blue by the composite in both processes (FP, FP-UV)
is the Freundlich model (Figures 6-7 and Table 3) which indicates that the adsorption is carried out
on heterogeneous surfaces.
The Freundlich constant (KF), which is proportional to the adsorption capacity is higher for
composite in the presence of UV light. The value of the constant 1/n, is lower than 1 for both
94
According to Hernández-Hernández et al., (2013) the removal of brilliant blue FCF from aqueous
solutions by using iron modified bentonite, they found that the adsorption was higher for the iron
modified material than the unmodified bentonite. The adsorption capacity was 14.22 mg/g for the
iron modified clay, this value is similar to the capacity found in this work for composite in the
presence of H2O2 and UV light and the times to reach the equilibrium were longer for the clay
materials than the materials used in this work. It is important to note that it is easier to separate the
In a previous work (Pinedo-Hernández et al., 2019) the adsorption capacities of these materials (Ze-
Fe and Ze(Fe-Cu) composite) were determined and they were lower than 2.5 mg/g, then both H2O2
4. Conclusions
Iron modified zeolitic tuff (Ze-Fe) and a composite (Ze(Fe-Cu)) were prepared and characterized.
The materials were used for the removal of a dye by the Fenton process with and without UV light.
In the experimental conditions, the concentration of H2O2 is not an important parameter in the
process, since the degradation was similar using different concentrations of hydrogen peroxide and
the adsorption was the highest at pH 3. The kinetic data showed a chemisorption mechanism, the
data were best adjusted to the second order model for both materials under Fenton and photo-
Fenton processes, the degradation rate was faster to the composite than to the iron modified zeolitic
material. The isotherms showed that the adsorbents are heterogeneous since the experimental data
were best fitted to the Freundlich model. The results show that these low-cost, ease to prepare,
adsorbents are efficient for the removal of dyes from aqueous solutions.
Declaration of interest
95
Acknowledgements
The authors are grateful to CONACYT for the financial support of this research, project 254665
References
Ahmed, M.B., Zhou, J.L., Ngo, H.H., Guo, W., Johir, M.A.H., Sornalingam, K., 2017. Single and
https://doi.org/10.1016/j.cej.2016.11.106
Akgül, M., 2014. Enhancement of the anionic dye adsorption capacity of clinoptilolite by Fe3+-
Textile wastewater treatment using iron-modified clay and copper-modified carbon in batch
and column systems. Water. Air. Soil Pollut. 227, 1–14. https://doi.org/10.1007/s11270-016-
2801-7
Bahrami, M., Nezamzadeh-Ejhieh, A., 2014. Effect of supporting and hybridizing of FeO and ZnO
semiconductors onto an Iranian clinoptilolite nano-particles and the effect of ZnO/FeO ratio in
the solar photodegradation of fish ponds waste water. Mater. Sci. Semicond. Process. 27, 833–
840. https://doi.org/10.1016/j.mssp.2014.08.030
Cai, M.Q., Zhu, Y.Z., Wei, Z.S., Hu, J.Q., Pan, S.D., Xiao, R.Y., Dong, C.Y., Jin, M.C., 2017.
https://doi.org/10.1016/j.scitotenv.2016.12.047
Cao, Z., Yue, Y., Zhong, H., Qiu, P., Chen, P., Wen, X., Wang, S., Liu, G., 2017. The cationic dye
96
removal by novel Si—Zn composites prepared from zinc ash. J. Taiwan Inst. Chem. Eng. 71,
464–473. https://doi.org/10.1016/j.jtice.2016.12.005
Colomban, P., Schreiber, H.D., 2005. Raman signature modification induced by copper
https://doi.org/10.1002/jrs.1379
Díaz-Nava, C., Olguín, M.T., Solache-Ríos, M., Alarcón-Herrera, M.T., Aguilar-Elguezabal, A.,
https://doi.org/10.1016/j.jhazmat.2009.01.138
Dinu, M.V., Lazar, M.M., Dragan, E.S., 2017. Dual ionic cross-linked
enhanced sorption properties for methylene blue. React. Funct. Polym. 116, 31–
40. https://doi.org/10.1016/j.reactfunctpolym.2017.05.001
Flores, N., Sirés, I., Garrido, J.A., Centellas, F., Rodríguez, R.M., Cabot, P.L.,
3–12. https://doi.org/10.1016/j.jhazmat.2015.11.040
Gao, Y., Guo, Y., Zhang, H., 2016. Iron modified bentonite: Enhanced adsorption
97
light photo-Fenton process at circumneutral pH. J. Hazard. Mater. 302, 105–113.
https://doi.org/10.1016/j.jhazmat.2015.09.036
Garcia-Segura, S., Brillas, E., Cornejo-Ponce, L., Salazar, R., 2016. Effect of the
Fe3+/Cu2+ ratio on the removal of the recalcitrant oxalic and oxamic acids by
https://doi.org/10.1016/j.solener.2015.11.033
https://doi.org/10.1016/j.jhazmat.2009.05.102
Gutiérrez-Segura, E., Solache-Ríos, M., Fall, C., Colín-Cruz, A., 2012. Influence of
the pH on distribution of denim blue in water Fe-Zeolitic tuff system. Sep. Sci.
Hassan, M.E., Chen, Y., Liu, G., Zhu, D., Cai, J., 2016. Journal of Water Process
O 3 / TiO 2 nanoparticles under visible light. J. Water Process Eng. 12, 52–57.
https://doi.org/10.1016/j.jwpe.2016.05.014
brilliant blue FCF from aqueous solutions using an unmodified and iron-
fern: A comparison of linear and non-linear methods. Water Res. 40, 119–125.
https://doi.org/10.1016/j.watres.2005.10.040
Humelnicu, I., Băiceanu, A., Ignat, M.E., Dulman, V., 2017. The removal of Basic
Blue 41 textile dye from aqueous solution by adsorption onto natural zeolitic
tuff: Kinetics and thermodynamics. Process Saf. Environ. Prot. 105, 274–287.
https://doi.org/10.1016/j.psep.2016.11.016
Jiménez-Cedillo, M.J., Olguín, M.T., Fall, C., 2009. Adsorption kinetic of arsenates
https://doi.org/10.1016/j.jhazmat.2008.07.049
Karthikeyan, S., Pachamuthu, M.P., Isaacs, M.A., Kumar, S., Lee, A.F., Sekaran, G.,
323–330. https://doi.org/10.1016/j.apcatb.2016.06.040
dye behaviors in the presence of biosorbents from maize (Zea mays L.), their Fe-
1594–1603. https://doi.org/10.1016/j.jece.2016.02.008
Luo, M., Yuan, S., Tong, M., Liao, P., Xie, W., Xu, X., 2014. An integrated catalyst
https://doi.org/10.1016/j.watres.2013.09.029
Nairat, M., Shahwan, T., Eroğlu, A.E., Fuchs, H., 2015. Incorporation of iron
nanoparticles into clinoptilolite and its application for the removal of cationic
https://doi.org/10.1016/j.jiec.2014.05.027
https://doi.org/10.5004/dwt.2017.20633
M., 2019. Efficient removal of brilliant blue by clinoptilolite tuff modified with
https://doi.org/10.5004/dwt.2019.23623
Raman, C.D., Kanmani, S., 2016. Textile dye degradation using nano zero valent
https://doi.org/10.1016/j.jenvman.2016.04.034
Ren, Y., Yuan, Y., Lai, B., Zhou, Y., Wang, J., 2016. Treatment of reverse osmosis
Shahwan, T., Abu Sirriah, S., Nairat, M., Boyaci, E., Eroĝlu, A.E., Scott, T.B.,
Hallam, K.R., 2011. Green synthesis of iron nanoparticles and their application
Soler, M.A.G., Qu, F., 2012. Raman Spectroscopy of Iron Oxide Nanoparticles. In:
https://doi.org/10.1007/978-3-642-20620-7
Tekbaş, M., Yatmaz, H.C., Bektaş, N., 2008. Heterogeneous photo-Fenton oxidation
https://doi.org/10.1016/j.micromeso.2008.03.001
Thiam, A., Sirés, I., Brillas, E., 2015. Treatment of a mixture of food color additives
(E122, E124 and E129) in different water matrices by UVA and solar
https://doi.org/10.1016/j.watres.2015.05.057
Torres-Pérez, J., Solache-Ríos, M., Olguín, M.T., 2007. Sorption of azo dyes onto a
318. https://doi.org/10.1080/01496390601069879
nanoscale oxides and their carbonaceous composites. Environ. Technol. 33, 545–
554. https://doi.org/10.1080/09593330.2011.584571
Truskewycz, A., Shukla, R., Ball, A.S., 2016. Iron nanoparticles synthesized using
green tea extracts for the fenton-like degradation of concentrated dye mixtures at
https://doi.org/10.1016/j.jece.2016.10.008
Wang, Y., Lin, X., Shao, Z., Shan, D., Li, G., Irini, A., 2017. Comparison of Fenton,
https://doi.org/10.1016/j.cej.2016.10.133
Wang, Y., Zhao, H., Zhao, G., 2015. Iron-copper bimetallic nanoparticles embedded
Zanin, E., Scapinello, J., de Oliveira, M., Rambo, C.L., Franscescon, F., Freitas, L.,
de Mello, J.M.M., Fiori, M.A., Oliveira, J.V., Dal Magro, J., 2017. Adsorption of
https://doi.org/10.1016/j.psep.2016.11.008
102
a) b) c)
d) e) f)
Frequency Counts of B
g) h) i)2018
16
14
Frecuency(%)
12
10
8
6
4
2
0
0 5 10 15 20 25 30 35
Particle size (nm)
Figure 1.
103
Ze-Fe
Ze-Fe(Fe-Cu) composite
Figure 2.
1.0
ZeFe FP
ZeFe FP-UV
ZeFe(Fe-Cu)FP
0.8 ZeFe(Fe-Cu) FP-UV
Counts (c.p.s.)
0.6
0.4
0.2
0.0
0 10 20 30 40 50 60 70 80
2 theta
Figure 3.
104
a) b) c)
1.0 1.1 1.1
[E
0.9 sc 1.0 1.0
0.7
ba 0.8 0.8
u 0.7 0.7
0.6
na
qt (mg/g)
qt (mg/g)
0.6 0.6
qt (mg/g)
0.5
ci 0.5 0.5
0.4 ta 0.4 0.4
0.3 de Ze(Fe-Cu)FP
(Ze-Fe)FP 0.3 (Ze-Fe)FP 0.3 Pseudo first order model
0.2 Ze(Fe-Cu)FP
l Pseudo first order model Second order model
0.2 0.2
d Second order model Pseudo second order model
0.1 0.1 Pseudo second order model 0.1
oc
0.0 0.0 0.0
0 10 20 30 40 50 60 70 80
u 0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
Time (h) m Time (h) Time (h)
en
to
o
el
Figure 4. re
su
m
en
de
a) u b) c)
1.1 [E 1.3 1.3
n ( (
1.2 1.2
1.0 sc p K K
1.1 1.1
0.9 ri u a a
1.0 1.0
0.8 ba nt rt rt
0.9 0.9
0.7 u o 0.8 hi 0.8 hi
na
qt (mg/g)
in k qt (mg/g)
0.6 k
qt (mg/g)
0.7 0.7
0.5
ci te 0.6 e 0.6
e
ta re 0.5
y 0.5
0.4 y
de sa 0.4 (Ze-Fe)PF-UV 0.4
Ze-Fe (Fe-Cu)PF-UV
0.3
Pseudo first order model a a
(Ze-Fe)FP-UV
l nt 0.3 0.3 Pseudo first order model
0.2
0.2
Second order model n 0.2 Second order model n
Ze-Fe(Fe-Cu)FP-UV d e. Pseudo first order model
0.1
oc 0.1 S. 0.1 Pseudo second order model S.
P et
0.0
u 80 0.0 0.0 et
ue 0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70
0 10 20 30 40 50 60 70
m al al80
Time (h) de Time (h) Time (h)
en ., .,
sit 201 201
to ua 6) 6)
o r
Figure 5. el el
re cu
su ad
m ro
en de
de te
u xt
n o
p en
u cu
nt al
o q
in ui
te er
re lu
sa ga
nt 105
r
e. de
P l
ue d
de
14 14 12
a) b) c)
12
[E ( (
12
sc K 10 K
ri a a
10 10
ba rt 8 rt
u hi hi
8 8
na
qe (mg/g)
qe (mg/g)
k
qe (mg/g)
6
k
ci e e
6 ta 6
y y
de 4
4 4
a a
l Ze-Fe(Fe-Cu)FP
(Ze-Fe)FP n n
d Langmuir
Langmuir S.
2 (Ze-Fe)FP oc 2 2 FreundlichS.
Ze-Fe(Fe-Cu)FP Freundlich
u et Sips et
Sips
0 m 0
al 0
al
0 10 20 30 en40 0 1 2 3 4 5 ., 6 0 10 20 30 .,40
201 201
Ce (mg/L) to Ce (mg/L) 6) Ce (mg/L) 6)
o
el
re
su
m
Figure 6. en
de
u
n
p
u
nt
o
in
16 te 16
re a) b) c)
12
14 sa [E ( 14 (
nt sc K K
12 ri 10 12
e. a a
P ba rt rt
10 10
ue u 8
hi hi
na
qe (mg/g)
qe (mg/g)
8 de k k
qe (mg/g)
8
sit ci 6 e e
ua ta y
6 y 6
r de 4 a a
Ze-Fe(Fe-Cu)FP-UV
4 el l (Ze-Fe)FP-UV 4
(Ze-Fe)FP-UV Langmuir n Langmuir n
d
2
Ze-Fe(Fe-Cu)FP-UVcu 2 Freundlich S. Freundlich S.
ad oc Sips 2 Sips
u et et
0
ro al al
m10 0 0
0 2 4 6 8 de 0 2 4 6 8
en ., 10 0.0 0.5 1.0 1.5 2.0 2.5 3.0 .,3.5
Ce (mg/L) te Ce (mg/L) 201 201
to 6) Ce (mg/L) 6)
xt
o o
en el
cu re
Figure 7. al su
q m
ui en
er de
u 106
lu
ga n
r p
de u
nt
Table 1.
Materials
Element
(Ze-Fe)FP (Ze-Fe)FP-UV Ze-Fe(Fe-Cu)FP Ze-Fe(Fe-Cu)FP-UV
107
Table 2.
qe KL R2 a b R2 qe K R2
(Ze-Fe)FP 0.91 0.14 0.61 6.09 9.77 0.78 1.00 0.20 0.67
qexp=0.99
Ze-Fe(Fe-Cu)FP 0.75 20.91 0.60 254.99 12.98 0.96 0.79 22.41 0.70
qexp=0.96
(Ze-Fe)FP-UV 0.94 0.47 0.77 18.77 9.25 0.93 0.98 0.93 0.83
qexp=0.99
Ze-Fe(Fe-Cu)FP-UV 0.85 9.90 0.74 330.51 12.23 0.94 0.88 13.57 0.82
qexp=0.98
108
Table 3.
qo KL R2 KF 1/n R2 q Ka 1/n R2
(Ze-Fe)FP 10.13 11.63 0.7343 7.93 0.16 0.8088 154.49 3.87E-8 5.88 0.7700
Ze-Fe(Fe-Cu) FP 10.74 0.21 0.9451 2.89 0.35 0.9477 15.85 0.06 1.66 0.9483
(Ze-Fe)FP-UV 8.42 3.19 0.5822 5.97 0.17 0.4963 8.49 3.22 1.04 0.4989
Ze-Fe(Fe-Cu)FP-UV 14.49 2.43 0.9291 8.98 0.36 0.9373 25.57 0.37 1.79 0.9300
Figure Captions
Figure 1. Scanning electron microscopy image of (a) natural zeolite (Ze), (b) zeolite treated with
HCl (Ze-HCl), (c) zeolite (Ze–Fe), (d) composite (Ze(Fe–Cu)), (e) (Ze–Fe)FP after Fenton process,
(f) (Ze–Fe)FP-UV after Fenton process and UV light, (g) composite (Ze(Fe–Cu))FP after Fenton
process, (h) composite (Ze(Fe–Cu))FP-UV after Fenton process and UV light, (i) TEM image of
Fe–Cu nanoparticles.
Figure 3. XRD pattern of zeolite (Ze-Fe)FP, zeolite (Ze-Fe)FP-UV, Ze(Fe-Cu)FP composite and
Ze(Fe-Cu)FP-UV composite.
109
Figure 4. (a) Kinetic adsorption of brilliant blue by Ze-Fe and Ze-Fe(Fe-Cu) composite in the
presence of H2O2, (b) Experimental data of zeolite (Ze-Fe)FP adjusted to models and c)
Figure 5. (a) Kinetic adsorption of brilliant blue by (Ze-Fe) and Ze-Fe(Fe-Cu) composite in the
presence of H2O2 and UV light, (b) Experimental data of zeolite (Ze-Fe)FP-UV adjusted to models
Figure 6. (a) Adsorption of brilliant blue by Ze-Fe and Ze(Fe-Cu) composite in the presence of
H2O2, (b) Experimental data of zeolite (Ze-Fe)FP adjusted to models and c) Experimental data of
Figure 7. (a) Adsorption of brilliant blue by (Ze-Fe) and Ze-Fe(Fe-Cu) composite in the presence of
H2O2 and UV light, (b) Experimental data of zeolite (Ze-Fe)FP-UV adjusted to models and c)
Table Captions
Table 1. Elemental composition (wt%) of (Ze-Fe)FP, (Ze-Fe)FP UV, Ze(Fe-Cu)FP and Ze(Fe-
Table 2. Kinetic parameters of the blue brilliant adsorption by Ze-Fe and composite (Ze(Fe-Cu)) by
Table 3. Langmuir, Freundlich and Sips parameters of the blue brilliant adsorption by Ze-Fe and
110
4.3 Discusión General
Los resultados de TEM confirman que las dimensiones de las partículas de Fe-Cu se encuentran
en el régimen nano, la distribución del tamaño muestra partículas que varían entre 11.70 y 15.85
nm y la imagen TEM revela partículas esféricas como la morfología principal. La presencia de
óxido de hierro y óxido de cobre se confirma a partir de las mediciones de difracción indican que
las partículas de óxido de hierro se depositaron sobre la superficie de la matriz de zeolita
Los materiales preparados fueron evaluados para la remoción del colorante azul 1 de soluciones
acuosas, se llevaron a cabo la cinética con cada uno de los materiales. En base a los resultados
111
obtenidos, se seleccionó el tipo de acondicionamiento con HCl y posteriormente con FeCl3. Se
preparó un composito de material zeolítico, mediante la técnica de reducción de sales metálicas.
Se determinó el área específica de los materiales (Ze, Ze-HCl y Ze-Fe), estas fueron más altas
que las reportadas en la literatura para este tipo de materiales, para la zeolita natural fue de 37.61
m2/g, la zeolita ácida de 190.15 m2/g y para la zeolita férrica de 220.34 m2/g. El tamaño de
partícula para este material es de 0.84 mm.
Además se evaluó la capacidad de remoción del azul 1 mediante las cinéticas e isotermas de
adsorción, realizando experimentos por lotes con cada material adsorbente zeolita natural (Ze), la
zeolita ácida (Ze-HCl) y la zeolita férrica (Ze-Fe) y se estudió el efecto del pH para evaluar la
capacidad de remoción.
La cinética fue ajustada con los modelos de cinética de pseudo primer orden (Lagergren),
segundo orden (Elovich) y pseudo segundo orden (Ho). Los datos se ajustaron mejor al modelo
de Elovich, lo cual indica que la quimisorción podría ser el principal mecanismo de adsorción que
prevalece en el composito.
El composito mostró una mayor adsorción del colorante azul 1 a valores de pH ácido. El efecto del
pH sobre la remoción de azul 1, mostró una mayor adsorción del azul 1a valores de pH ácido (3)
tanto para la zeolita natural (Ze), la zeolita ácida (Ze-HCl) y la zeolita férrica (Ze-Fe). La estructura
de los diferentes materiales adsorbentes (Ze, Ze-HCl y Ze-Fe) no fue afectada por la inserción de
las nanopartículas, ni durante la adsorción del azul 1.
Se comprobó también que la estructura del composito no fue afectada por la inserción de las
nanopartículas, ni durante la adsorción del azul 1.
Se evaluó la capacidad de remoción del azul brillante 1 mediante las cinética e isotermas de
adsorción con el composito (Ze-Fe (Fe-Cu)) y la zeolita férrica (Ze-Fe), además se estudió el
efecto de la temperatura y del pH.
112
El punto de equilibrio de azul brillante 1 se alcanzó en 72 h, el porcentaje de remoción en
equilibrio fue de 75.29 y 87.02% para Ze-Fe y Ze-Fe (Fe-Cu) respectivamente.
La cinética fue ajustada con los modelos de pseudo primer orden (Lagergren), segundo orden
(Elovich) y pseudo segundo orden (Ho). Los datos se ajustaron para ambos materiales mejor al
modelo de Elovich, lo cual indica que la quimisorción podría ser el principal mecanismo de
adsorción que prevalece en el composito.
De acuerdo con los parámetros obtenidos de los diferentes modelos de isotermas, y R2, el modelo
que mejor se adapta al proceso de adsorción de azul brillante 1 por ambos adsorbentes es el
modelo de Freundlich que indica que la adsorción se lleva a cabo en superficies heterogéneas.
Ze-Fe y Ze-Fe(Fe-Cu) mostraron una mayor adsorción del colorante azul brillante 1 a valores de
pH ácido. Se comprobó también que la estructura del composito no fue afectada por la inserción
de las nanopartículas, ni durante la adsorción del azul brillante 1.
Se evaluó la capacidad de remoción del azul brillante 1 mediante las cinéticas e isotermas de
adsorción con el composito (Ze-Fe (Fe-Cu)) y la zeolita férrica (Ze-Fe) en presencia de H2O2 y luz
UV. Se determinó la cantidad de H2O2 y el pH óptimos para llevar a cabo las reacciones Fenton.
Se observó que las mejores condiciones fueron 2 mL de H2O2 a pH 5, teniendo un 98. 57% de
remoción bajo estas condiciones.
Mediante los estudios cinéticos se encontró que ambos materiales Ze-Fe.PF y el composito Ze-
Fe (Fe-Cu).PF tienen buena capacidad de adsorción para el azul brillante 1 y los resultados
fueron similares Ze-Fe(Fe-Cu).PF (qe= 0.99 mg/g), seguido del Ze-Fe.PF (qe= 0.96 mg/g). Se
113
observó que con el composito se tienen velocidades de sorción más rápida en los primeros
tiempos en comparación con la zeolita férrica.
Los resultados mostraron que las capacidades de adsorción en presencia de H2O2 y luz UV fueron
más altas que en las otras condiciones experimentales y las capacidades para Ze-Fe.PF-UV (qe=
0.99 mg/g) y para Ze-Fe(Fe-Cu).PF-UV (qe= 0.98 mg/g) fueron similares.
La cinéticas fueron ajustadas a los modelos de pseudo primer orden (Lagergren), segundo orden
(Elovich) y pseudo segundo orden (Ho). Los datos se ajustaron para ambos materiales, al modelo
de Elovich, lo cual indica que la quimisorción podría ser el principal mecanismo de adsorción que
prevalece en ambos materiales
114
5
Conclusiones
115
El material compuesto de clinoptilolita-férrica y nanopartículas de Fe-Cu ((Ze-Fe(Fe-Cu)) se
sintetizó a través de un método de reducción in situ, en presencia del material zeolítico, con
borohidruro de sodio.
Los materiales fueron caracterizados por SEM, TEM, XRD e IR. La caracterización de ambos
materiales por DRX revela la presencia de clinoptilolita y la cristalinidad del material después de
los diferentes tratamientos. El área específica del material zeolitico aumenta de 37.67 a 220.24
m2/g después de los tratamientos. La microscopía electrónica de transmisión (TEM) confirmó que
las nanopartículas de Fe-Cu, con tamaños entre 11.70-15.85 nm, se depositaron con éxito y se
dispersaron de manera eficiente en la zeolita, generando así el composito Ze-Fe(Fe-Cu).
Las isotermas de adsorción de azul brillante por el composito Ze-Fe(Fe-Cu) y Ze-Fe mostraron
que el composito Ze-Fe(Fe-Cu) tiene una capacidad mayor que la Ze-Fe para la adsorción del
colorante en solución acuosa. Se observó que estos materiales son particularmente efectivos para
eliminar el colorante, y que la eliminación del este aumenta cuando el pH disminuye. Los valores
positivos de ΔH° para ambos materiales adsorbentes indicaron que los procesos de adsorción son
endotérmicos.
Los materiales se utilizaron también para la eliminación del colorante mediante el proceso de
Fenton con y sin luz UV. En las condiciones experimentales, la concentración de H2O2 no es un
parámetro importante en el proceso, ya que la degradación fue similar al usar diferentes
concentraciones de peróxido de hidrógeno y la adsorción fue la más alta a pH 3.
Los datos cinéticos mostraron un mecanismo de quimisorción, los datos se ajustaron mejor al
modelo de segundo orden para ambos materiales bajo procesos Fenton y foto-Fenton, la tasa de
degradación fue más rápida en el composito que en el material zeolítico modificado con hierro.
Las isotermas mostraron que los adsorbentes son heterogéneos, ya que los datos experimentales
se ajustaron mejor al modelo de Freundlich.
116
Los resultados muestran que estos adsorbentes, de bajo costo y fáciles de preparar, son
eficientes para la eliminación de colorantes de soluciones acuosas. Los resultados proporcionan
conocimiento para el desarrollo de nuevas tecnologías que utilizan materiales compuestos
bimetálicos nanoestructurados para tratar aguas residuales.
117
Anexos
118
Anexo 1. Gráfica de barrido del colorante azul 1
119
Anexo 2.Curva de calibración
2.5
2
Absorbancia (nm)
1.5
Ecuación de la recta
1 y=0.01329x
r2= 0.9997
0.5
0
0 5 10 15 20 25
Concentración (ppm)
120
Anexo 3.Tabla de propiedades de azul 1
121
Anexo 4. Análisis termogravimétrico (TGA)
En las figuras 1.5 y 1.6 se muestran los patrones de degradación termica de la zeolita natural
y al zeolita ferrica. En el caso de la zeolita natural se observa perdidas de masa entre 89.68°
C y 469.81° C (estos datos se obtuvieron de la primera derivada de la curva de la perdida de
masa), la cual se debe al agua que se encuentra debilmente ligada a la estructura zeolitica. El
porcentaje obtenido en todos los casos es de 9.14% y 0.77%.
Bajo condiciones normales, las moleculas de agua se encuentran en los espacios de los
canales y cavidades de la estructura de la roca zeolitica (Tsitsishvilli et al., 1992). Se ha
discutido en algunas referencias bibliográficas sobre las pérdidas en peso de esta zeolita,
mencionan que las pérdidas entre 25 a 600 °C es debido a las pérdidas de agua retenidas en
la superficie higroscópica de la zeolita, las cuales se encuentran enlazadas mediante
diferentes interacciones de puentes de hidrógeno, con los grupos funcionales y los iones
metálicos, entre los canales de la misma, por lo que el agua se evapora a temperaturas
superiores a los 100 °C.
De tal forma que las pérdidas entre 100° C y 400° C corresponden a agua enlazada al Na+ y
Ca++, y por encima de los 400°C puede ser de agua que interacciona mediante puentes de
hidrógeno con los grupos OH de la zeolita (Montes-Luna A. de J. et al., 2008).
122
La zeolita férrica fue analizada también por TGA. En la Figura 1.6 se puede observar el
termograma en el cual muestra una pérdida de peso de 9.99% a 60.62 °C y otra pérdida de
1.73 % a una temperatura de 470.62° C , esto es debido a la humedad presente en el
material; por tanto ninguno de los acondicionemientos realizados al material modifican el
patrón de degradación térmica de los mismos.
123
Anexo 5. Análisis por activación neutrónica
Las irradiaciones se realizaron en un reactor nuclear MARCA TRIGA III. Se usaron las
referencias de la IAEA para calcular las concentraciones elementales, y los datos nucleares
de los isótopos identificados en los espectros.
Las irradiaciones se realizaron en un reactor nuclear TRIGA MARK III a un flujo de neutrones
de 1 × 1013 cm−2 s−1. Se irradiaron muestras de 200 mg de CM durante 2 h. Se dejaron
decaer durante 2, 13,40 y 60 días y los espectros de rayos γ se registraron durante 2 h. En la
Tabla 1.4 se muestran los datos nucleares de los isótopos identificados en los espectros de
rayos; algunos elementos no se determinaron debido a la vida media de los isotopos que se
forman y a la interferencia con otros elementos.
Tabla 1.4 Análisis por activación neutrónica de la zeolita modificada con hierro (Ze-Fe) y del
composito (Ze-Fe (Fe-Cu)).
124
Referencias
125
Referencias
Abdala, M., Fábregas, O., Lamas, G., Fantini, A., Craievich, F., y Walsöe de Reca, E. 2011
Materiales nanoestructurados. Síntesis, caracterización y aplicaciones, Asociación Argentina de
Materiales, 8: 5-17.
Akgül M. 2014. Enhancement of the anionic dye adsorption capacity of clinoptilolite by Fe3+-
grafting. Journal of Hazardous Materials. 267:1– 8.
Ayodele O.B, Lim J.K., Hameed B.H. 2012. Degradation of phenol in photo-Fenton process by
phosphoric acid modified kaolin supported ferric-oxalate catalyst: Optimization and kinetic
modeling. Chemical Engineering Journal. 197: 181–192.
Bokare, AD. and Choi-W. 2014 Review of iron-free Fenton-like systems for activating H2O2 in
advanced oxidation processes Journal of Hazardous Materials.275:121–135.
Bosch Pedro y Schifter, I. 1995. La zeolita: Una piedra que hierve. Fondo de Cultura Económica,
4ª Reimpresión, México
Breck D.W., 1974. Zeolite Molecular Sieves. Editorial Wiley Interscience. New York. U.S.A.
Bullent A., Turan M., Ozdemir, and Celik M.S. 2004. Color Removal of Reactive Dyes from Water
by Clinoptilolite. Journal of Environmental Science and Health. 5:1251–1261.
Camilo García J., Castellanos M. P., Uscátegui Á., Fernández J., Pedroza A. M. y Daza C.
E.2012. Remoción de colorantes sintéticos mediante el proceso Fenton heterogéneo usando
126
Fe2O3 soportado en carbón activado obtenido a partir de residuos de rosas. Universitas
Scientiarum. 17: 303-314.
Conesa V. F. 2003. Guía Metodológica para la Evaluación del Impacto Ambiental. 3ra. ed. Ed.
Mundi- Prensa.
Corral-Capulin, N.G., Vilchis-Nestor, A.R., Gutiérrez-Segura, E., Solache-Ríos, M., 2019. 401
Comparison of the removal behavior of fluoride by Fe 3+ modified geomaterials from 402 water.
Appl. Clay Sci. 173, 19–28. https://doi.org/10.1016/j.clay.2019.03.003
Cortés, M.R, Martínez, M. V, Solache, R.M. 2004.Evaluation of natural and surfactant modified
zeolites in the removal of cadmium from aqueous solutions. Separation Science and Technology.
39: 2711-2730.
Cubero N., Monferrer A. y Villalta J. 2002. Aditivos alimentarios. Editorial Grupo Mundi Prensa,
España. Consultada el 20 de Octubre de 2009, en: http://books.google.com.mx/books
Curkovic L., and Cerjan, S.S. 1997. Metal ion exchange by natural and modified zeolites. Water
Research. 31:1379-1382.
127
Díaz Nava María del Carmen. 2006. Organo-minerales mexicanos (zeolítico y arcillosos) como
adsorbentes de fenol, contaminante del agua. Tesis de Doctorado en Ciencia y Tecnología
Ambiental. Centro de Investigaciones en materiales avanzados, S.C.
ElShafei G.M.S., Yehia F.Z., Dimitry O.I.H., Badawi A.M.and Eshaq Gh.2014. Ultrasonic assisted-
Fenton-like degradation of nitrobenzene at neutral pH using nanosized oxides of Fe and Cu.
Ultrasonics Sonochemistry. 21:1358–1365.
Fan C., Horng C.-Y. and Li S.-J. 2013. Structural characterization of natural organic matter and its
impact on methomyl removal efficiency in Fenton process. Chemosphere.30:20-30.
F. Oluwakemi Kehinde, H. Abdul Aziz. 2015. Classical optimization of process variables in the
treatment of real textile wastewater using clinoptilolite, J. Environ. Chem. Eng. 30
Garrido E.G., Theng B.K.G y Mora M.L. .2010. Clays and oxide minerals as catalysts and
nanocatalysts in Fenton-like reactions-A review. Applied Clay Science 47:182–192
Georgi A., Gonzalez R., Ko¨hler R., and Kopinke F. 2010.Fe-Zeolites as Catalysts for Wet
Peroxide Oxidation of Organic Groundwater Contaminants: Mechanistic Studies and Applicability
Tests. Separation Science and Technology. 45:1579–1586.
Ghasemi M., Javadian H. , Ghasemi N., Agarwal S. and Gupta V.K . 2016 .Microporous
nanocrystalline NaA zeolite prepared by microwave assisted hydrothermal method and
determination of kinetic, isotherm and thermodynamic parameters of the batch sorption of Ni (II).
Journal of Molecular Liquids 215; 161–169
128
Giannetto P. Giuseppe. 1990. Zeolitas: características, propiedades y aplicaciones industriales.
2da ed. Ed. Innovación Tecnológica. Caracas, Venezuela.
Gupta V. K. and Suhas. 2009. Application of low-cost adsorbents for dye removal – A review.
Journal of Environmental Management. 90:2313–2342.
Gutierrez S .E. Solache R. M. and Colin C. A. 2009. Sorption of indigo carmine by a Fe-zeolitic tuff
and carbonaceous material from pirolyzed sewage sludge. Journal of Hazardous
Materials.170:1277-1235.
Gutiérrez, E., Solache-Ríos, M., Colín-Cruz, A ,Fall C. 2012. Adsorption of cadmium by Na and Fe
modified zeolitic tuffs and carbonaceous material from pyrolyzed sewage sludge. Journal of
Environmental Management. 97: 6-13.
Kar Yan Hor, Jasmine Mun Cheng Chee, Meng Nan Chong, Bo Jin, Christopher Saint, Phaik Eong
Poh, Rupak Aryal.2016.Evaluation of physicochemical methods in enhancing the adsorption
performance of natural zeolite as low-cost adsorbent of methylene blue dye from wastewater,
Journal of Cleaner Production 30:1-13
Huang Y.-H., Huang Y.-J., Tsai H.-C. and Chen H.-T.2010. Degradation of phenol using low
concentration of ferric ions by the photo-Fenton process. Journal of the Taiwan Institute of
Chemical Engineers. 41: 699–704.
Iurascu B., Siminiceanu I.,Vione D., Vicente M.A. and Gil A.2009.Phenol degradation in water
through a heterogeneous photo-Fenton process catalyzed by Fe-treated laponite. Water
Research.43:1313-1322.
129
heterogeneous photodegradation of mixture of 4-methoxy aniline and 4-chloro-3-nitro
aniline.Journal of Colloid and Interface Science 490; 478–487
Kerkez D. V, Tomasˇevic´ D. D., Kozma G., Becˇelic´-Tomin M. R., Prica M. Dj., Roncˇevic´ S. D.,
Kukovecz A., Dalmacija B. D. and Ko´ nya Z. 2014.Three different clay-supported nanoscale zero-
valent iron materials for industrial azo dye degradation: A comparative study Journal of the Taiwan
Institute of Chemical Engineers 30.
Kalantari K, Ahmad MB, Masoumi HR, Shameli K, Basri M, Khandanlou R..2014. Rapid and high
capacity adsorption of heavy metals by Fe3O4/montmorillonite nanocomposite using response
surfacemethodology: Preparation, characterization, optimization, equilibriumisotherms,
andadsorptionkineticsstudy.Journal of the Taiwan Institute of Chemical Engineers 2014),
http://dx.doi.org/10.1016/j.jtice.10.025
Kim S., Seralathan K-K, Lee K-J., Park Y-J, Shea P., Lee W-H, Kim H-M and Oh B-T .2013.
Removal of Pb(II) from aqueous solution by a zeolite–nanoscale zero-valent iron composite.
Chemical Engineering Journal. 217: 54–60.
Klamerth N., Malato S., Agüera A. and Fernández A. 2 0 1 3 . Photo-Fenton and modified photo-
Fenton at neutral pH for the treatment of emerging contaminants in wastewater treatment plant
effluents: A comparison. Water Research.47: 833-840.
Kilinc A. S., Özbayrak Ö. and Alpat S. 2008. The adsorption kinetics and removal of cationic dye,
Toluidine Blue O, from aqueous solution with Turkish zeolite. Journal of Hazardous
Materials.151:213-220.
Kirk-Othmer. 2001. Enciclopedia Tecnológica Quimica.3ª edición. Ed Limusa. México. pp. 117-
184, 131-133, 293, 337-355, 468-473.
Kong X.,Han Z., Zhang W., Song L. and Li H. (2016) Synthesis of zeolite-supported microscale
zero-valent iron for the removal of Cr6+ and Cd2+ from aqueous solution. Journal of
Environmental Management 169; 84 – 90.
Kuˇsi´c H., Koprivanac N., Boˇzi´c A. L and Selanec I.2006. Photo-assisted Fenton type processes
for the degradation of phenol: A kinetic study. Journal of Hazardous Materials.136:632–644.
130
Legese H. S.,Unni N. B.,Redi-Abshiro2 M., Aravindhan R.,Diaz I. and Tessema M.2015.
Synthesis, characterization and catalytic application of zeolite based heterogeneo us catalyst
of iron(III), nickel(II) and copper(II) salen complexes for oxidation of organic pollutants. J
Porous Mater 22:1363–1373
Ma, M; Ding, G; Gao, Y; Zhou, F; Xu, Y. y Liu, J. 2004.Discoloration of methylene blue and
wastewater from a plant by a Fe/Cu bimetallic system, Chemosphere. 55:1207 – 1212.
Margeta K., Stefanović Š. C., Kaučič V. and Logar N. Z. 2015.The potential of clinoptilolite-rich
tuffs from Croatia and Serbia for the reduction of toxic concentrations of cations and anions in
aqueous solutions. Applied Clay Science 116–117; 111–119.
Mittal A. 2006. Use of hen feathers as potential adsorbent for the removal of a hazardous dye,
Brilliant Blue FCF, from wastewater. Journal of Hazardous Materials.128:233–239.
Moon B.-H., Park Y.-B. and Park K.-H.2011.Fenton oxidation of Orange II by pre-reduction using
nanoscale zero-valent iron .Desalination .268: 249–252.
Nairat M., Shahwan T., Erog˘lu A. and Fuchs H..2015. Incorporation of iron nanoparticles into
clinoptilolite and its application for the removal of cationic and anionic dyes. Journal of Industrial
and Engineering Chemistry. 21:1143–1151.
131
Narges A., Nezamzadeh-Ejhieh A. 2015. Modificationofclinoptilolitenano-particleswithiron oxide:I
ncreasedcompositecatalyticactivity for photodegradationofcotrimaxazoleinaqueoussuspension.
Materials Science in Semiconductor Processing.31; 684–692.
Nidheesh P.V. and Gandhimathi R. 2012. Trends in electro-Fenton process for water and
wastewater treatment: An overview Desalination. 299:1–15.
Nogueira P., Oliveira M., y Paterlini W. 2005. Simple and fast spectrophotometric determination of
H2O2 in photo-Fenton reactions using metavanadate.Talanta. 66:86-91.
Ni Y. ,Wang Y., Kokot S. 2009. Simultaneous kinetic spectrophotometric analysis of five synthetic
food colorants with the aid of chemometrics. Talanta. 78:432–441.
Peralta J. R., Zhaoa L., Lopez M. L., De la Rosa G., Hongb J. and Gardea J. L. 2011.
Nanomaterials and the environment: A review for the biennium 2008–2010. Journal of Hazardous
Materials.186:1–15.
132
Pérez R., González M. and Bentrup U. 2012. Preparation and in situ spectroscopic
characterization of Cu-clinoptilolite catalysts for the oxidative carbonylation of methanol.
Microporous and Mesoporous Materials. 164: 93–98.
Pham-Thi H. , Byeong-Kyu L., Jitae K.and Chi-Hyeon L.2016. Nitrophenols removal from aqueous
medium using Fe-nano mesoporous zeolite. Materials and Design 101; 210–217.
Pinedo S. 2010. Remoción del colorante azul 1 de soluciones acuosas utilizando zeolita férrica,
Tesis de Ingeniería Química, Instituto Tecnológico de Toluca, México.
Rahim P.S., Abdul Raman A. A. and Wan Daud W. M. A. 2014. Review on the application of
modified iron oxides as heterogeneous catalysts in Fenton reactions .Journal of Cleaner
Production. 64: 24-35.
Rajeshwar K. y Ibañez J. 1997. Enviromental electrochemistry. Ed. Academic Press. pp. 776.
Ríos M. 2005. Contaminación la tierra agredida. Ed. Equipo Sirius. pp. 12-15.
Rusevova K., Kopinke F. y Georgi A. 2012. Nano-sized magnetic iron oxides as catalysts for
heterogeneous Fenton-like reactions—Influence of Fe(II)/Fe(III) ratio on catalytic performance.
Journal of Hazardous Materials.241– 242: 433– 440.
Shahwana T., Abu S., Nairata M., Boyacıb E., Ero˘glub, A.E., Scottc, T.B., Hallamc K.R. 2011.
Green synthesis of iron nanoparticles and their application as a Fenton-like catalyst for the
degradation of aqueous cationic and anionic dyes. Chemical Engineering Journal .172: 258– 266
133
Shimizu, A., Tokumura M., Nakajima K. and Kawase Y. 2012. Phenol removal using zero-valent
iron powder in the presence of dissolved oxygen: Roles of decomposition by the Fenton reaction
and adsorption/precipitation Journal of Hazardous Materials. 201– 202: 60– 67.
Srivastava M., Singh J., Mishra R. K. and Ojha A. K. 2013. Electro-optical and magnetic
properties of monodispersed colloidal Cu2O nanoparticles. Journal of Alloys and Compounds.
555:123–130.
Sun S-P, Zeng X., Lemley A. T. 2013. Nano-magnetite catalyzed heterogeneous Fenton-like
degradation of emerging contaminants carbamazepine and ibuprofen in aqueous suspensions and
montmorillonite clay slurries at neutral pH. Journal of Molecular Catalysis A: Chemical. 371: 94–
103.
Torres P.J., Solache R. M. and Olguín M. T. 2007. Sorption of Azo Dyes onto a Mexican
Surfactant-Modified Clinoptilolite-Rich Tuff. Separation Science and Technology. 42: 299-318.
Trujillo, J., Solache-Ríos, M., Vilchis, A., Sánchez, V., and Colín-Cruz, A. 2012. Removal of
remazol yellow from aqueous solution using Fe–Cu and Fe–Ni nanoscale oxides and their
carbonaceous composites . Environmental Technology .33:545–554.
Trujillo, J., Solache-Ríos, M., Vilchis, A., Sánchez, V., and Colín-Cruz, A. 2012. Fe–Ni
nanostructures and C/Fe–Ni composites as adsorbents for the removal of a textile dye from
aqueous solution. Water Air Soil Pollut, DOI 10.1007/s11270-011-0948-9.
Tsitsishvili G. V., Andronikashvili T. G,, Kirov G N, Filizova L. D.1992. Natural Zeolites. Ed. Ellis
Horwood.
134
Villalva Coyote Rafael. 2009. Sorción y desorción del colorante anaranjado remazol de una zeolita
modificada con hierro. Tesis de Licenciatura. Instituto Tecnológico de Toluca.
Vinita M., Dorathi R., Palanivelu K. 2010. Degradation of 2,4,6-trichlorophenol by photo Fenton’s
like method using nano heterogeneous catalytic ferric ion. Solar Energy. 84: 1613–1618.
Wang S., Boyjoo Y., Choueib A. and Zhu Z.H. 2005. Removal of dyes from aqueous solution using
fly ash and red mud. Water Research.39: 129-138.
Wang S. and Ariyanto E. 2007. Competitive adsorption of malachite green and Pb ions on natural
zeolite. Journal of Colloid and Interface Science. 314: 25-31.
Wang, X., Huanga L., Chena Z., Megharaj M. and Naidub, R.2013.Synthesis of iron-based
nanoparticles by green tea extract and their degradation of malachite. Industrial Crops and
Products. 51: 342– 347.
Wang W., Liu Y., Li T. and Zhou M. 2014. Heterogeneous Fenton catalytic degradation of phenol
based on controlled release of magnetic nanoparticles. Chemical Engineering Journal. 242:1–9.
Weber, W.J. Jr. y Smith E.H. 1987. Simulation and Design models for adsorption processes.
Environmental Science and Technology .21:1040-1050.
White, R., Luque, R., Budarin, L., Clark, J., and Macquarrie, D. 2009 Supported metal
nanoparticles on porous materials. Methods and applications. Chem. Soc. Rev, 38: 481–494.
Zelmanov G. and Semiat R. 2008. Phenol oxidation kinetics in water solution using iron(3)-oxide-
based nano-catalysts .Water Research. 4 2: 3848 – 3856.
135
Zeng X., Hanna K. and Lemley A. T. 2011. Cathodic Fenton degradation of 4, 6-dinitro-o-cresol
with nano-magnetite .Journal of Molecular Catalysis A: Chemical. 339:1–7.
136