Вы находитесь на странице: 1из 37

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/320972332

A Review of Preparation Methods for Supported Metal Catalysts

Chapter  in  Advances in Catalysis · January 2017


DOI: 10.1016/bs.acat.2017.10.001

CITATIONS READS
4 770

5 authors, including:

Bahareh Tavakoli Sonia Eskandari


University of South Carolina University of South Carolina
15 PUBLICATIONS   41 CITATIONS    7 PUBLICATIONS   4 CITATIONS   

SEE PROFILE SEE PROFILE

Umema Khan Rembert White


University of South Carolina University of South Carolina
1 PUBLICATION   4 CITATIONS    3 PUBLICATIONS   4 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Catalyst View project

Catalysis View project

All content following this page was uploaded by Sonia Eskandari on 01 August 2018.

The user has requested enhancement of the downloaded file.


Provided for non-commercial research and educational use only.
Not for reproduction, distribution or commercial use.

This chapter was originally published in the book Advances in Catalysis, Vol. 61 published by Elsevier,
and the attached copy is provided by Elsevier for the author's benefit and for the benefit of the author's
institution, for non-commercial research and educational use including without limitation use in instruction
at your institution, sending it to specific colleagues who know you, and providing a copy to your
institution’s administrator.

All other uses, reproduction and distribution, including without limitation commercial reprints, selling or
licensing copies or access, or posting on open internet sites, your personal or institution’s website or
repository, are prohibited. For exceptions, permission may be sought for such use through Elsevier's
permissions site at:
http://www.elsevier.com/locate/permissionusematerial

From Bahareh A.T. Mehrabadi, Sonia Eskandari, Umema Khan, Rembert D. White and John R.
Regalbuto, A Review of Preparation Methods for Supported Metal Catalysts. In: Chunshan Song, editor,
Advances in Catalysis, Vol. 61, Burlington: Academic Press, 2017, pp. 1-35.
ISBN: 978-0-12-812078-1
© Copyright 2017 Elsevier Inc.
Academic Press
Author's personal copy

CHAPTER ONE

A Review of Preparation Methods


for Supported Metal Catalysts
Bahareh A.T. Mehrabadi, Sonia Eskandari, Umema Khan,
Rembert D. White, John R. Regalbuto1
University of South Carolina, Columbia, SC, United States
1
Corresponding author: e-mail address: Regalbuj@cec.sc.edu

Contents
1. Introduction 2
1.1 Impregnation Methods 3
1.2 Deposition Precipitation 4
1.3 Reductive Deposition 5
1.4 Colloidal 6
2. Literature Study of Synthesis Efficacy 7
2.1 Method Surveys 8
2.2 Particle Size Ranges 10
3. Newer Methods: SEA and CEDI 12
3.1 Strong Electrostatic Adsorption 12
3.2 Charge-Enhanced Dry Impregnation 26
4. Summary 29
References 32
About the Authors 34

Abstract
To gauge the efficacy of the various methods to synthesize supported metal catalysts,
over 1500 literature articles from the past 3 years have been surveyed. Platinum catalysts
over silica (SiO2), alumina (Al2O3), titania (TiO2), and carbon (C) supports have been
selected as examples. We include our work on the methods of strong electrostatic
adsorption (SEA) and a newer version of it, charge-enhanced dry impregnation
(CEDI). These methods are compared and contrasted to those most prevalent in the
literature. SEA and CEDI yield ultra-small nanoparticles, usually less than 1.5 nm average
particle size, and are simple and scalable, especially CEDI, which is a simple adaptation of
common incipient wetness impregnation. These methods can be extended to the syn-
thesis of supported bimetallic nanoparticles with equally efficacious results.

Advances in Catalysis, Volume 61 # 2017 Elsevier Inc. 1


ISSN 0360-0564 All rights reserved.
https://doi.org/10.1016/bs.acat.2017.10.001
Author's personal copy
2 Bahareh A.T. Mehrabadi et al.

1. INTRODUCTION
Catalysts play a vital role in chemical reactions in the chemical
industry. The active component of many catalysts is a supported metal such
as nickel, platinum, or chromium, and since only the surface of the metal is
available to catalyze the reaction, it is normally advantageous to maximize the
metal surface area for a given weight of metal. Thus, it is usually desired to
synthesize small metal crystallites, typically less than about 1–10 nm, anchored
to a thermally stable, high-surface-area support such as alumina (Al2O3),
silica (SiO2), titania (TiO2), or carbon (C). The synthesis of nanoparticles
with controllable nanoscale sizes and shapes and the prevention of the int-
rinsic propensity of nanoscale aggregation are the most important challenges
for catalyst preparation. In addition, the preparation method of supported
catalysts significantly affects their activity, selectivity, and lifetime (1).
A number of books on supported catalyst synthesis (2–5) have appeared
in the last decade or so, as well as a book chapter (6) and several review arti-
cles (7,8). Schwarz (8) described two main routes to synthesize supported
metal nanoparticles: in solution with “three-dimensional chemistry,” after
which they can be deposited onto catalyst supports, or by “two-dimensional
chemistry” involving deposition of metal precursors at the liquid–support
interface, after which the precursors can be thermochemically converted
to metal particles. This review mostly covers the latter set of methodologies,
as these serve as the most direct comparison of our own SEA and CEDI
methods. For a thorough review of the former, the reader is referred to
the classic Schwarz review (8) or a recent book chapter (6).
The common criteria for a high-performance catalyst are narrow size dis-
tribution and high dispersion on support. According to these criteria, several
innovative and cost-effective preparation methods have been developed
beyond the oldest, most common method of impregnation and show prom-
ise for reaching performance optimization by controlling synthetic proce-
dures and conditions. Among them there are deposition precipitation
(DP), reductive deposition, and colloidal syntheses. In addition to those
methods, strong electrostatic adsorption (SEA) and charge-enhanced dry
impregnation (CEDI) methods are discussed in depth in later sections. These
methods have been recently introduced as simple techniques to maximize
the interaction between a support and a catalytic metal precursor and can give
more control over particle size and size distributions. The advantages and dis-
advantages of each method will be summarized at the end of the review.
Author's personal copy
Preparation of Supported Metal Catalysts 3

1.1 Impregnation Methods


Among the many methods to prepare supported metal catalysts, impregna-
tion is the simplest and most widespread method for Pt catalyst preparation
(9). Using an amount of the precursor solution in excess of the pore volume
of the support, yielding a thin slurry, is termed wet impregnation (WI). Lim-
iting the solution amount to just fill the pore volume is called dry or incipient
wetness impregnation (DI). In WI, the impregnated support is filtered out,
leaving excess liquid containing any precursor that was not retained by the
support. This entails a need to recycle the excess liquid to minimize wastage
of the precursor. The use of DI eliminates excess liquid and the need for a
filtration step. However, the lack of filtration step during DI synthesis of cat-
alyst means that any counterions from the metal precursor salt, such as the
chloride from platinum tetrammine chloride, (NH3)4PtCl2, will be retained
in the dried catalyst. If removal of these other substances is desired, further
processing is necessary. To obtain the final catalysts with the zero-valent
metal particles anchored onto the support, the impregnated and dried pow-
der is subjected to thermal treatment in a calcining and/or reducing envi-
ronment. Precursor ligands are removed, either by decomposition or by
reacting with air/hydrogen-producing gaseous molecules. Other compo-
nents that do not form gaseous products, such as alkali metals or their salts,
remain in the final catalyst.
Catalysts made by these impregnation methods do not usually produce
particles with high dispersion due to the lack of induced interaction between
the precursor and the support, which allows mobility of the precursor during
drying. In DI, the pH of the resulting paste of support and impregnating
solution is buffered by the support surface to the point of zero charge
(PZC) of the support, as the amount of hydroxyl groups on the support
surface is orders of magnitude more than the acid or the base in solution
that can charge (protonate or deprotonate) the surface (10). Even in WI,
with excess liquid, the pH of the solution is still buffered to around the
PZC of the support. Without a strong interaction between the precursor
and the support, as drying ensues, loss of the solvent causes the moisture
to migrate to the external surface of the support, causing the precursor to
agglomerate together (11). The major weakness of the impregnation
method is the lack of size control of metal particles except when the porous
substrate has a narrow pore size distribution, e.g., in highly ordered meso-
porous carbon. A distribution of particle sizes from nanometer to micron
scale is commonly observed. A schematic illustration of impregnation
Author's personal copy
4 Bahareh A.T. Mehrabadi et al.

method is shown in Fig. 1. Typical DI-prepared Ni and Pt catalysts with


large particles and polydisperse size distributions are shown later.
Besides “wet” and “dry,” other variations of impregnation exist. If two
or more metal precursors are impregnated simultaneously, the method is
usually termed “coimpregnation.” Other times, dry impregnation is used
for successive metals; this can be termed “sequential impregnation.” Both
of these methods are similar to the single metal analog in that solution
pH is not controlled and the particle size is not small or monodisperse; addi-
tionally, the multiple metals are usually poorly alloyed.

1.2 Deposition Precipitation


The DP method involves the conversion of a highly soluble metal salt precur-
sor into a less-soluble substance which precipitates only on the support and
not in solution (13). Typically, this process is achieved by a change in solution
pH, addition of a precipitation agent, addition of a reducing agent, or change
in the concentration of a complexation agent. There are two main conditions
which must be fulfilled to make sure that the precipitation occurs only on the
support instead of in solution: a strong interaction between the soluble metal
precursor and the surface of the support and controlled concentrations of
the precursor in solution to avoid spontaneous precipitation. Usually in the
presence of the support, the solubility limit shifts to lower concentration com-
pared to the solubility limit in solution to favor deposition on the support.
The concentration of the metal salt should be kept between the solubility
point and the supersolubility point in solution to prevent precipitation in
the liquid. The supersolubility curve is the boundary of the metastable zone
and labile zone (liquid and precipitate). In practice, the support is slurried in
the solution containing the soluble precursor. The most important drawback
of this method is the poor control of metal distribution and surface composi-
tion, which also makes it difficult to prepare true bimetallic catalysts with con-
trolled compositions (14). Fig. 2 shows the schematic of the procedure for the
preparation of the Pt–Au/CeO2 (RDP) catalyst by deposition precipitation.

Precursor Impregnation Filtration or Reduction


PtCl62– H2, >300°C, inert
Mix with evaporation
carbon

Supported Pt

Fig. 1 A schematic illustration of impregnation method to synthesize supported plat-


inum nanoparticles (12). Modified from Chan, K. Y.; et al. J. Mater. Chem. 2004, 14(4),
505–516. Copyright 2004 Royal Society of Chemistry.
Author's personal copy
Preparation of Supported Metal Catalysts 5

Fig. 2 (A) A schematic of the procedure for the preparation of the Pt–Au/CeO2 catalyst
by deposition precipitation. (B) TEM images and EDX spectra of the Pt–Au/CeO2 catalyst
(15). Modified from Hong, X.; et al. Catal. Sci. Technol. 2016, 6, 3606–3615. Copyright 2015
Royal Society of Chemistry.

1.3 Reductive Deposition


Among various methods to prepare nanometer-sized particles, the liquid-
phase reduction method is one of the easiest procedures, since nanoparticles
can be directly obtained from various precursor compounds soluble in a spe-
cific solvent (16). For the synthesis of metal nanoparticles in this method,
first precursor metal salts were dissolved in aqueous or organic media. By
adding a solution of reducing agent, the reduction occurs and then metal
particles are selectively deposited on supports. Reductive deposition is
related to DP, where precipitation is induced in the solution phase but close
to the surface, and not on the surface as in DP. Metal precursors are generally
reduced through thermal decomposition at high-temperature, electrochem-
ical routes and chemical pathways using various reducing agents such as
sodium borohydride, hydrazine, ethylene glycol, and ascorbic acid. One
of the weaknesses of this method is that most of the strong reducing agents
are toxic, and in addition, electrochemical routes require additional electro-
chemical devices. Thus, the simple and ecofriendly reduction routes of metal
Author's personal copy
6 Bahareh A.T. Mehrabadi et al.

precursors are highly desired for the improvement in productivity and sus-
tainability of synthetic processes of metal-based nanomaterials (17). Another
problem with this method is that the type of reducing agent employed has
been found to greatly affect the resulting particles. Fig. 3 depicts the sche-
matic representation of size-controlled silver nanoparticles synthesized
employing the coreduction approach.

1.4 Colloidal
Colloidal syntheses are “three-dimensional” methods involving several steps:
(1) preparation of catalyst precursors in a solvent using a protective agent, such
as a surfactant (e.g., cetyltrimethylammonium bromide), (2) deposition of the

Fig. 3 (A) A schematic representation of size-controlled silver nanoparticles synthesized


employing the coreduction approach. (B) FEG-TEM images of silver nanoparticles of size
ranges 5  0.7 nm (18). Based on Agnihotri, S.; et al. RSC Adv. 2014, 4(8), 3974–3983. Open
access publication.
Author's personal copy
Preparation of Supported Metal Catalysts 7

Fig. 4 (A) A schematic illustration of the synthesis of the polyol method (12). (B) TEM
micrograph for a sample of ternary Pt32V13Fe55 nanoparticles prepared by colloidal
method (20). Reprinted with permission from Chan, K. Y.; et al. J. Mater. Chem. 2004,
14(4), 505–516. Copyright 2004 Royal Society of Chemistry and from Luo, J.; et al. Chem.
Mater. 2005, 17(21), 5282–5290. Copyright 2005 American Chemical Society.

colloids into the support, and (3) reduction of the mixture using chemicals
(19). Colloidal metals can be prepared in an aqueous or organic medium,
depending on the stabilizing agent used. The colloidal route can provide very
small particle sizes, but the use of surfactant and protective agents required
that the catalyst will be washed in an appropriate solvent several times;
or use high temperatures in an inert environment to decompose or remove
these foreign compounds. Before performing this step, the catalyst needs to
be adsorbed into the support to prevent the agglomeration of metal nano-
particles. Therefore, it is desirable to use an alternative route for create small
nanoparticles without the use of protecting agents, thus decreasing the
level of difficulty and avoiding contamination. Fig. 4 shows the schematic
illustration of the synthesis of the polyol method.

2. LITERATURE STUDY OF SYNTHESIS EFFICACY


Despite the large number of publications and patents about catalyst
preparation, the field of catalyst preparation method can be still considered
too much an empirical art and not sufficiently a science, as different methods
show different particle sizes and wide ranges of size distributions. This is
borne out by a comprehensive review of the literature to examine the
Author's personal copy
8 Bahareh A.T. Mehrabadi et al.

efficacy of the most important techniques for catalyst preparation which


have been used or developed over the recent past. To this end, the common
catalyst metal, platinum, on the most common supports, alumina (Al2O3),
silica (SiO2), titania (TiO2), and carbon (C) have been selected for the
review to limit the number of papers surveyed to a tractable number (about
1500). The preparation methods and particle sizes, where available, were
culled from these papers. For the purpose of this review, metal particle size
will be considered the chief metric of synthesis efficacy.
Since the number of papers surveyed is far too many to list in the refer-
ences, we have placed all papers that were analyzed only for method and
particle size in Supplementary Information, grouped according to the
support.

2.1 Method Surveys


Methods of catalyst preparation are very diverse, and each catalyst may be
produced via different routes. The different preparation methods for the
synthesis of supported Pt catalysts are summarized in Fig. 5 for Pt/SiO2
[74 references], Pt/Al2O3 [212 references], Pt/TiO2 [376 references], and
Pt/C [853 references] catalysts as reported in the literature from 2014 to
2017. Impregnation is separated into simple impregnation with a single
metal, and co- and sequential impregnation for multiple metals. While
not often employed, SEA and CEDI are included in the figure as they will
be contrasted with the other methods in the subsequent sections of this
chapter. Other methods include sol gel, microemulsion, reverse micro-
emulsion, wet chemistry, ion exchange, chemical vapor deposition, electroless
deposition (ED), electrodeposition, physical deposition, photodeposition,
atomic layer deposition, dendrimer-encapsulated metal, aerosol spray pyroly-
sis, low-temperature electrostatic self-assembly method, cool sputtering,
electron beam evaporation, simple direct adsorption, extractive-pyrolytic
method, fluoride-induced self-transformation approach, evaporative-
crystallization deposition, arc plasma deposition, liquid-phase synthesis, super-
critical fluid reactive deposition, sol immobilization, and extractive-pyrolytic
method. These results show that the various modes of impregnation are the
most prevalent methods for preparation of Pt on SiO2, Al2O3, and TiO2
and made up 49%, 72%, and 22% of the preparations reported in 2014 and
27%, 76%, and 9% in 2017. For Pt on carbon, in addition to the impregnation
method, colloidal and reductive deposition methods are used more frequently;
these stem from the electrochemistry literature and arise from the need for
Author's personal copy
Preparation of Supported Metal Catalysts 9

relatively high metal loadings. Also of note is that alumina employs the
highest fraction of impregnation preparations. This is likely because of
the effectiveness of DI with alumina. Alumina is often impregnated with
chloroplatinic acid (CPA, H2PtCl6). When dissolved in solution, CPA
gives anionic Pt hexachloride, [PtCl6]2, and two protons, which can

Fig. 5 The methods of preparation for Pt catalysts reported in the literature from 2014
to 2017 for (A) silica, (B) alumina, (C) titania, and (D) carbon.
Author's personal copy
10 Bahareh A.T. Mehrabadi et al.

Fig. 5—Cont’d

charge the alumina surface and, in doing so, create electrostatic attraction
between the anionic metal precursor and the protonated, positively
charged alumina surface (21).

2.2 Particle Size Ranges


One of the most widely used metrics of the efficiency of catalyst preparation
is metal “dispersion,” defined as the ratio of exposed metal surface atoms
Author's personal copy
Preparation of Supported Metal Catalysts 11

to the total number of metal atoms. Dispersion has a reciprocal relation with
particle size: in 1 nm particles, virtually 100% of the metal atoms are
exposed, at 2 nm, about 50% are exposed, at 3 nm, dispersion is about
33%, and at 5 nm, dispersion is about 20% (22). It is of interest to
compare the particle size and size distributions which were obtained by
different catalyst preparation methods to see which methods can obtain
the smallest particle size and narrowest size distribution. To this purpose,
we have summarized the average Pt particle sizes, averaged over all papers
for a given year and support, which obtained by each method over the
last 4 years. Fig. 6A–D shows the average Pt particle sizes and size ranges
for Pt/SiO2, Pt/Al2O3, Pt/TiO2, and Pt/C catalysts. The Pt particle size
standard deviations obtained by different methods, again averaged for all
papers for a given year and support, are also shown in Fig. 6A–D as error
bars. The standard deviations for most years and most supports are larger
than the particle sizes for most methods, with the exception of SEA
and CEDI.
The histograms show that the preparation of Pt on different supports
by impregnation, reductive deposition, and colloidal methods gives larger
particle size with larger size distributions. As an example, the Pt particle
size obtained by the impregnation method in different references has been
reported widely different (i.e., as small as 1 nm and as big as 20 nm). In con-
trast, the limited number of SEA papers used for Fig. 6 suggests that tight
size distributions are achieved by the SEA method. The average particle size
of the catalyst that was prepared by SEA is 1.5 nm for carbon, 1.8 nm for
silica, and 2.9 nm for alumina. These are much smaller particles than
those prepared by DI, 10.3, 10, and 10 nm, respectively. This reveals that
the development of supported catalysts by the SEA method is a promising
new approach that, in principle, allows for much better control of the
active metal phase during catalyst synthesis. Among the most exciting
prospects are the ability to tightly control the particle size distribution of
supported metal catalysts, with averages in the 1–3 nm or even sub-
nanometer range.
The histograms also show that the standard deviations from the average
particle size for the SEA method are much smaller in comparison to other
methods. The large standard deviations in the other methods suggest that
these methods do not have great control over particle size and size distribu-
tions and have resulted in widely different and sometimes contradicting
results.
Author's personal copy
12 Bahareh A.T. Mehrabadi et al.

Fig. 6 The average particle size and size distribution calculated for Pt catalysts for dif-
ferent preparation methods reported in the literature from 2014 to 2017, for (A) silica,
(B) alumina, (C) titania, and (D) carbon.

3. NEWER METHODS: SEA AND CEDI


3.1 Strong Electrostatic Adsorption
The method known as SEA has been discussed in detail in a number of book
chapters (5,6,9,21–25). In this method, the electrostatic interactions can be
Author's personal copy
Preparation of Supported Metal Catalysts 13

Fig. 6—Cont’d

induced by controlling the pH of impregnation. This idea was promulgated


to the catalysis community in 1978 by Brunelle (26) and further developed
by the pioneering work of Schwarz (27,28). They postulated that the
adsorption of noble metal complexes onto common oxide supports was
essentially coulombic in nature. The hydroxyl (OH) groups that populate
oxide surfaces become protonated and so positively charged below a char-
acteristic pH value, while the same hydroxyl groups become deprotonated
Author's personal copy
14 Bahareh A.T. Mehrabadi et al.

and negatively charged above this characteristic pH value. This pH, at which
the surface is neutral, is termed the PZC. Brunelle explained that oxides
placed in solutions at pH values below their PZC would adsorb anions such
as hexachloroplatinate[PtCl6]2; at pH values above their PZC, the same
support would adsorb cations such as platinum tetrammine [Pt(NH3)4]2+.
Rational synthesis techniques were then developed by Regalbuto et al.
(29), initially to deposit Pt and Pd nanoparticles onto inorganic supports.
The technique is, however, quite versatile: it was adapted to various sup-
ports, like silica (30,31), alumina (32), and carbon (33,34), and can be
extended to other metals and to bimetallic nanoparticles (35,36).
Regalbuto et al. started with the chloroplatinic acid (H2PtCl6 or CPA)/
alumina system (37–39). In this system, anionic chlorides (PtCl6 6 ) and
oxychlorides (PtCl5(OH)2, PtCl4(OH)(H2O)) complexes adsorb over a
positively charged alumina surface in the low pH range. This method has
also extended to cationic platinum tetrammine ([(NH3)4Pt]2+ or PTA, from
platinum tetrammine chloride salt, Pt(NH3)4Cl2), over deprotonated and
negatively charged silica at high pH (30,31). The latter two works especially
demonstrate the practical consequence that when strongly adsorbed at the
optimal pH, the monolayer of adsorbed coordination complexes retains
its high dispersion through the catalyst pretreatment process such as drying
and reduction steps. Finally, the SEA method was applied to activated
carbon surfaces. Carbon has great potential for SEA because controlled
oxidation of the carbon surface at mild or rigorous conditions leads to a
lesser or greater amount of oxygen functional groups on the surface (40–42),
irreversibly altering the PZC and influencing the adsorption of Pt complexes
in a way that is systematic and controllable. In a previous work on activated
carbons (33), the highest PZC active carbons adsorb the largest amount of
anions [PtCl6]2 and the lowest amount of cations [Pt(NH3)4]2+, while the
lowest-PZC active carbons adsorb the lowest amount of anions and the
largest amount of cations. Pt uptakes reach a maximum with respect to pH
for a given metal precursor and active carbon PZC.
SEA is normally performed with an excess of solution. This keeps
pH shifts to a minimum and facilitates sampling the solution. Thus, SEA
at high liquid/support ratios is a special case of WI in which the pH of
the solution is controlled to achieve strong electrostatic attraction between
precursor and surface. For benchtop experiments, surface loadings (SLs) are
generally held low (500–2000 m2/L) in order to minimize pH shifts for
the sake of convenience (i.e., surface loading is the amount of support
surface per liter of preparation solution). The SEA preparations give much
smaller particle size with much smaller standard deviations than the DI
Author's personal copy
Preparation of Supported Metal Catalysts 15

preparations in which agglomeration is likely caused by the accumulation of


metal in the solution phase during drying (6).
The procedure of SEA method is shown in Fig. 7. The first step is to
determine the PZC of the support, which is easily determined by measuring
final pH (43). Then second step of SEA approach is to perform an uptake–
pH survey to determine the pH of strongest interaction. Once strongly
adsorbed, the idea is to perform the pretreatment steps of calcinations or
reduction, often referred to in industry as catalyst finishing, in such a way
that the monolayer morphology of the precursor is maintained as the metal
is reduced, such that high metal dispersion is achieved (6). This procedure is
schematized in Fig. 8, depicting an electrostatically adsorbed layer of a

Fig. 7 The strong electrostatic adsorption procedure: (1) PZCs are determined and the
appropriate metal complex is selected (a cation for silica), (2) uptake–pH surveys are
conducted to determine the pH of strongest interaction, and (3) catalysts are synthe-
sized at this pH (10.5), and after mild reduction, ultra-small nanoparticles with a narrow
size distribution result (6). Based on Zhong, C. J.; et al. In Comprehensive Inorganic Chem-
istry II; Reedijk, J.; Poeppelmeier, R. K., Eds. Elsevier Inc., 2013; pp 75–102. Copyright 2013
Elsevier.

Fig. 8 The hydration sheaths and ligands attached to the metal core of the precursors
are removed via catalyst finishing, i.e., reduction or calcination to preserve the highly
dispersed metal nanoparticles.
Author's personal copy
16 Bahareh A.T. Mehrabadi et al.

mixture of hydrated metal precursors, followed by a thermal reduction in H2


needed to remove the metal ligands, reduce the metals to their zero-valent
state, and nucleate the atoms into nanoparticles.
The overarching hypothesis of SEA is that to produce highly dispersed
metal nanoparticles on the support, the metal precursors must themselves be
highly dispersed onto the support during the adsorption stage of preparation.
A parameter-free model dubbed the revised physical adsorption model
(RPA) has been proposed by Santhanam et al. (44) as a way to model the
adsorption of different complexes with only physical forces, omitting any
chemical terms and mechanisms from the adsorption model. This model
allows for the prediction of the pH shift of metal oxides in solution as a func-
tion of surface loading, and Park and Regalbuto (43) were the first to sim-
ulate the pH shift effect of the bulk solution with the addition of higher
loadings of oxide. The RPA model was proposed as a novel technique to
accurately measure the PZC of oxides. From the RPA model, the PZC
of the support material plays a large role in complex adsorption. Fig. 9 shows
that supports with low PZC values have a very broad adsorption peak where
the uptake is at a maximum for PTA over low PZC silica, the electrostatic
potential develops quickly after the solution passes the PZC, and we see narr-
ower maximum uptake bands for CPA over alumina as the support PZC

1.8

1.6 PZC = 11

PZC = 9
1.4

1.2
G in mol/m2

PZC = 7
1

0.8 PZC = 1
PZC = 3
0.6
PZC = 5
0.4
PZC = 7
PZC = 5
0.2
PZC = 3
PZC = 9
PZC = 1
PZC = 11
0
0 2 4 6 8 10 12 14
pH
Fig. 9 Simulation of Pt uptake over carbons of various PZCs vs final pH, CPA (lower pH
range), or PTA (higher pH range) (34). Based on Hao, X.; Barnes, S.; Regalbuto, J. R. J. Catal.
2011, 279(1), 48–65. Copyright 2011 Elsevier.
Author's personal copy
Preparation of Supported Metal Catalysts 17

increases. The increased adsorption of CPA, compared to PTA, is attributed to


the sizes of the complexes as CPA is assumed to retain one hydration sheath
when adsorbed on the support surface and PTA retains two; thus, the steric
limitations may be a contribution to the difference in total uptake.
The experimental results for Pt uptake over different alumina (38), silica
(30), carbon (34), and titania (21) supports have been shown in Fig. 10A–D,
respectively. The qualitative trends of electrostatic adsorption are indeed
seen. In all pH regimes, the uptake–pH curves are volcano shaped, which
is a further indication of electrostatic adsorption.
There are some other recent studies that successfully used the SEA
method. Hao et al. (34) studied a series of carbons of different types (acti-
vated, black, and graphitic) with different surface areas and PZC. Cationic
Pt tetrammine, [(NH3)4Pt]2+, was adsorbed over low- and mid-PZC
carbons in the high pH range, while anionic hexachloroplatinate (IV),
[PtCl6]2, was adsorbed over high-PZC carbons in the low pH range. Their
results showed that particles synthesized by SEA are normally in the 1–2 nm
range and are as small as or smaller with narrower size distributions than by
other methods, especially at high metal loadings. Lambert et al. (46) applied

Fig. 10 The experimental results for Pt uptake over different alumina (38), silica (45),
carbon (34), and titania (21) supports. Data from Regalbuto, J. R.; et al. J. Catal. 1999,
184(2), 335–348. Copyright 1999 Elsevier; Schreier, M.; Regalbuto, J. R. J. Catal. 2004,
225(1), 190–202. Copyright 2004 Elsevier; Hao, X.; Barnes, S.; Regalbuto, J. R. J. Catal.
2011, 279(1), 48–65. Copyright 2011 Elsevier; and Regalbuto, J. R. In De Jong, K. P., Ed.:
Wiley-VCH Verlag GmbH & Co.: The Netherlands, 2009; pp 33–58. Copyright 2009 John
Wiley and Sons.
Author's personal copy
18 Bahareh A.T. Mehrabadi et al.

SEA method to get highly dispersed Pt/carbon xerogels. They observed


very small platinum particles (1.1–1.3 nm) by TEM. They also showed that
these Pt particles are accessible and the Pt/carbon xerogel catalysts are very
active for the hydrogenation of benzene into cyclohexane. Samad et al. (47)
studied the SEA of anionic H2PtCl6 (CPA) and cationic (NH3)4PtCl2 (PTA)
platinum precursor complexes on low (silica, titania) and high (alumina)
PZC supports. Their XRD profiles of the washed and then reduced samples
confirm the presence of very small Pt particles (<1.5 nm).
Fig. 11 shows STEM characterization of highly dispersed Pt on different
supports, silica, alumina, carbon, and titania, from the lab of the authors. Pt
nanoparticles show up in Z-contrast imaging as bright white spots and are
seen for all catalysts. From all of the particle size analysis results, a common
trend found is that very small and well-dispersed Pt particles with narrow size
distribution and at similar particle spacing were achieved via SEA method for
all cases.

Fig. 11 The STEM results for Pt uptake over (A) silica, (B) alumina, (C), carbon, and
(D) titania supports.
Author's personal copy
Preparation of Supported Metal Catalysts 19

The SEA method also has been used for many other noble and base
metal precursors. Jiao et al. (48) extended the SEA method to the synthesis
of other silica-supported noble and base metal catalysts and investigated the
correlation between strong electrostatic interaction and highly dispersed
metals on reduced catalysts. Pd, Cu, Co, Ru, and Ni over SiO2 were
prepared via SEA and compared with those prepared via the traditional
dry impregnation method. Fig. 12 shows representative Z-contrast STEM
images of reduced catalyst samples prepared via SEA at monolayer loading
and the corresponding loading for DI preparations for Ru and Ni metals
(48). Metal particle sizes of the reduced catalysts prepared by SEA remain
relatively small, and the size distributions are limited to a narrow range.
These results also showed that reduced catalysts prepared by DI give much

Fig. 12 Data from representative STEM images and corresponding particle distributions
of reduced Ru and Ni catalyst samples: (A) 3.0% Ru, SEA preparation; (B) 3.0% Ru, DI
preparation; (C) 1.6% Ni, SEA preparation; and (D) 1.6% Ni, DI preparation (48). Jiao,
L.; Regalbuto, J. R. J. Catal. 2008, 260(2), 329–341. Copyright 2008 Elsevier.
Author's personal copy
20 Bahareh A.T. Mehrabadi et al.

larger particle sizes, and in a particular sample, the particle sizes occur in a
wide distribution, and the average particle size increases significantly with
increasing metal weight loading.
Ling et al. (49) also made cobalt (Co) catalysts supported on silica by
using the SEA method. The TEM images of their samples showed that
the Co particle size is narrowly distributed and have an average value
of 5.2 nm with a standard deviation of 1.2 nm. In another work, Jiao
et al. (50) carried out the adsorption surveys of a series of metal ammines
(Pt, Pd, Cu, Co, Ru ammine complexes) at various pH values and further
characterized Pt/SBA-15, Pd/SBA-15, and Co/SBA-15 materials syn-
thesized at the optimal pH. Their results show that the SEA method
yields small metal particles, which appear to form homogeneously
throughout the pore volume of the SBA-15. Liu et al. (51) synthesized
supported Ru and Pt nanoparticles over amorphous SiO2 and SBA-15
supports by the method of SEA and subsequently treated them under dif-
ferent steaming-reduction conditions to achieve a series of catalysts with
controlled particle sizes, ranging from 1 to 8 nm. They demonstrated that
the SEA methodology allows the control of particle which can be applied
to high surface area supports with common metal precursors. Metal
particle sizes of the reduced catalysts prepared by SEA show very narrow
size distribution throughout the SBA-15 pore channels. The average
particle size for 9.7 wt% Ru/SBA-15 is 1.1 0.2 nm, and that of the
5.4 wt% Pt/SBA-15 is 1.3 0.3 nm.
The comparison of STEM images of Pt/SBA-15 made by SEA and DI,
Pd/SBA-15, and Ru/SBA-15 is shown in Fig. 13A–D respectively.
Miller et al. (31) synthesized Pt over silica by adsorption of PtðNH3 Þ4 2 +
from strongly basic impregnation solutions. They established that by
selecting the method of preparation and calcination temperature, very small
to large metallic nanoparticles can be prepared. Their results also showed
that the highest dispersions were obtained at lower Pt loading and drying
in air at 100°C followed by reduction in H2 at 250°C. With increasing cal-
cination temperature, there was a nearly linear decrease in the Pt dispersion.
It was suggested that the dispersion is dependent on the distribution of Pt
species on silica at the time of reduction. Klaigaew et al. (52) prepared
the Co supported on silica fibers (Co/SF) catalysts with different methods
of the SEA, DI, DP, and hydrothermal synthesis. The activity of Co/SF cat-
alysts were compared for Fischer–Tropsch synthesis (FTS) reaction, and the
results show that the catalysts prepared by the SEA method have the highest
Author's personal copy
Preparation of Supported Metal Catalysts 21

Fig. 13 Representative STEM images of reduced Pt particles: (A) 8.0 wt% Pt/SBA-15
(SEA, 1-h contact, reduced at 160°C) (50), (B) 8.0 wt% Pt/SBA-15 DI sample (1-h contact,
reduced at 350°C) (50), (C) 8.7 wt% Pd/SBA-15 (50), and (D) 9.7 wt% Ru/SBA-15 (51).
Based on Jiao, L.; et al. J. Catal. 2008, 260(2), 342–350. Copyright 2008 Elsevier;
Liu, Q. L.; et al. Phys. Chem. Chem. Phys. 2014, 16(48), 26431–26435. Copyright 2014 Royal
Society of Chemistry.

catalytic activity for FTS reaction. They also showed that the excellent cat-
alytic activity of Co/SF which has been prepared by SEA may be attributed
to the smaller size of cobalt particles and higher cobalt dispersion on this
catalyst.
Akbarzadeh et al. (53) used the SEA method for the deposition of Co
nanoparticles on carbon nanotubes (CNTs) support. Their TEM and
FESEM images indicated well-dispersed Co particles on the CNTs
support, and their TPR results was proven reduction peak at high temper-
ature (530°C), indicating strong interactions between Co and CNT support.
They proved that the SEA method is the desired method in preparing
supported cobalt catalysts. Zhang et al. (54) synthesized Pd nanoparticles
Author's personal copy
22 Bahareh A.T. Mehrabadi et al.

supported on CNTs by the SEA method. The TEM images of their catalysts
showed an average Pd size of around 1 nm. They also showed that a higher
Pd nanoparticle dispersion was obtained for the catalysts prepared by the
SEA method than the catalysts prepared by the IWI method, which contrib-
uted to the higher specific activity value for the Suzuki coupling reaction.
Kyriakidou et al. (55) used the SEA method for deposition of Ag on different
oxide supports. They showed that the SEA with Ag diammines at high pH is
demonstrated to be a simple, reliable method to synthesize small Ag
nanoparticles (3.8 nm) over all oxide supports, but reacts to form large par-
ticles on carbon.
As supported metal nanoparticles prepared by SEA have high dispersion
and narrow size distribution, therefore, they are desirable as seeds for addi-
tion of secondary metals using methods like ED, since the prepared bime-
tallic catalysts should be of similar dispersion as the base, or seed, catalyst.
Ryapan et al. (56) prepared the uniform deposition of Ag on the surfaces
of small Pd particles using the combined methods of SEA and ED resulted
in significant improvement of the bimetallic Ag–Pd surface for the selective
hydrogenation of acetylene. Mager et al. (57) provided Ru on nanocarbon
support catalysts from water-soluble molecular clusters by pH-controlled
impregnations in order to probe the interactions occurring between the sup-
ports and the clusters and to maximize them. They showed that in addition
to the interactions in the form of π-bond coordination between the cluster
and the carbon aromatics, electrostatic interactions however play a deter-
mining role, allowing more or less adsorption in the case of attractive or
repulsive interactions, respectively.
In addition to the adsorption of single metals on single supports, the SEA
method can also be used for the selective adsorption of single metals over
composite supports. Feltes et al. (35) suggested that pH control for selective
adsorption in the preparation of promoted catalysts is a viable strategy for the
synthesis of any other promoted and/or bimetallic catalyst systems where
intimate contact between the different catalyst components is highly desired.
They investigated the preparation of a Mn-promoted Co/TiO2 catalyst for
FTS using pH control for deposition of Mn selectively onto Co to enhance
promoter–metal interaction.
Samad et al. (58,59) have explored the potential to control the location
and mechanism of Pt adsorption on mixed oxides silica–alumina composites
through the choice of metal precursor and simple pH adjustment. They
also showed that mixed oxides intimate mixing between the two phases
Author's personal copy
Preparation of Supported Metal Catalysts 23

(particularly, small alumina domains) so led to a more complicated behavior.


Using cationic Pt precursors, Pt(NH3)42+, at moderate to high pH, two
adsorption mechanisms (SEA and ion exchange at the acid sites present
on the interface between the two oxides) were identified by comparing
adsorption uptakes on mixed oxides with those on mixtures of pure com-
ponents. Their results indicated that electrostatic adsorption can be exploited
to achieve selective deposition, and all the 0.7 wt% Pt catalysts synthesized
using cationic Pt precursor yielded small (< 3.0 nm) Pt particle size and the
SEA method can be applied to many other composite oxides for selective
deposition (58).
Zecevic et al. (60) have also exploited the ability to control metal pre-
cursor deposition with pH control. They showed in a bifunctional catalyst—
comprised of an intimate mixture of zeolite Y and alumina binder, and
with platinum metal controllably deposited onto either the zeolite or the
binder—that atomic-scale proximity between metal and zeolite acid sites
can be detrimental. Specifically, the selectivity when cracking large hydro-
carbon feedstock molecules for high-quality diesel production is optimized
with the catalyst that contains platinum on the binder, that is, with a nano-
scale rather than closest intimacy of the metal and acid sites. Fig. 14 shows
HAADF-STEM imaging and EDX elemental mapping of 70-nm-thick
ultra microtomed sections of Pt-Y/A (i.e., Fig. 14A and B) and confirms
that Pt particles with a narrow size distribution, around 2.5 nm, are exclu-
sively present in the zeolite crystals. Observations of Pt-Y/A (Fig. 14A)
reveal Pt nanoparticles located within the zeolite crystals even though they
exceed the size of the micropores (roughly 1 nm). In the Pt-A/Y sample, Pt
particles with an average size of 3.5 nm and narrow size distribution reside
exclusively on the alumina phase of the extrudates (Fig. 14C and D). The
absence of Pt particles in the zeolite crystals was confirmed by electron
tomography (Fig. 14B).
The SEA method can also be used for higher metal loadings in a cyclic
fashion. The metal loading obtainable by SEA is limited to the amount of
metal precursor which can be adsorbed in a monolayer, and the precursors
contain not only a ligand sphere but also hydration sheaths (see fig. 9 of Ref.
(22)). Typical loadings obtainable are a few weight percent over supports of
surface area of hundreds of meters per gram. If higher loadings are desired,
the metal precursor can be applied via SEA, reduced to form nanoparticles,
which regenerates the support surface, and the two steps can be repeated. Job
et al. (61) used three cycles of SEA/reduction for Pt over carbon xerogel to
Author's personal copy
24 Bahareh A.T. Mehrabadi et al.

Fig. 14 Controlled deposition of platinum (Pt) on either the zeolite Y or the alumina
component of Y/A extrudates. (A, B) Pt-Y/A. The HAADF-STEM image (A) is of a
70-nm-thick section of a Pt-Y/A sample, with Pt particles of 2.5 nm residing exclusively
within the zeolite crystals, as evident from the EDX map (B) showing Pt (yellow), Si
(green), and Al (red) signals. (C, D) Pt-A/Y. Shown are an HAADF-STEM image (C) and
EDX map (D) of a 70-nm-thick section of Pt-A/Y, with Pt particles (yellow) of 3.5 nm
residing exclusively on the alumina platelets (red), while the zeolite crystals (green)
are empty. Scale bars, 50 nm (60). Data from Zecevic, J.; et al. Nature 2015, 528(7581),
245–248. Copyright 2015 Nature.

increase the Pt loading from 7 to 15 and then 22.5 wt%. They showed that in
each case, the metal particles were found to be highly dispersed (particle size
2 nm); in addition, their results showed that the average particle size did
not change upon repeated impregnation. Fig. 15 presents TEM micrographs
of samples from first and third cycles. These pictures show that both catalysts
are well dispersed and present an extremely low degree of agglomeration;
the particles are homogeneous in size and shape. Both samples contain
Pt particles with 2 nm in diameter, the number of which increases
with the Pt weight percentage while keeping the same metal dispersion.
Author's personal copy
Preparation of Supported Metal Catalysts 25

Fig. 15 TEM micrographs of the catalysts: (A) after one impregnation (X-Pt-1) and
(B) after three successive impregnations (X-Pt-3) (61). Data from Job, N.; et al. Catal.
Today 2010, 150(1–2), 119–127. Copyright 2010 Elsevier.

30 30 30
A B C
Number of particles (%)

Number of particles (%)

Number of particles (%)

20 20 20

10 10 10

0 0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6 0 1 2 3 4 5 6
Size (nm) Size (nm) Size (nm)

Fig. 16 Pt particle size distributions of the Pt/carbon xerogel catalyst issued from the
image analysis of TEM micrographs: (A) X-Pt-1, (B) X-Pt-2, and (C) X-Pt-3 (61). Based
on Job, N.; et al. Catal. Today 2010, 150(1–2), 119–127. Copyright 2010 Elsevier.

Fig. 16 shows their particle size distributions of the three catalysts obtained
by image analysis of TEM micrographs. In each case, the particle size ranges
from 1 to 5 nm, and no significant statistical difference is detected in the size
distributions (61).
Author's personal copy
26 Bahareh A.T. Mehrabadi et al.

Finally, the SEA method can be used for bimetallic catalyst preparations
(62,63) by simply mixing metal precursors in solution and adsorbing them
into a mixed monolayer. Recently, simultaneous electrostatic adsorption
(co-SEA) of two metals has been applied to Pt and Pd precursors on silica,
alumina, and carbon supports by Cho et al. (62) and Wong et al. (64). The
former work showed that co-SEA yields well-dispersed, homogeneously
alloyed Pt–Pd particles on silica, carbon, and alumina supports (two of
which, silica and carbon, are shown in Fig. 17), while co-DI gives large,
agglomerated, inhomogeneous particles.

3.2 Charge-Enhanced Dry Impregnation


There is no reason other than experimental convenience that SEA is applied
at high ratios of liquid to solid—to minimize pH shifts and for ease in
obtaining solution samples for measurement of metal concentration.
SEA-type impregnations performed in a large excess of Pt precursor solution
will result in metal loss if the amount of metal in solution exceeds the pre-
cursor monolayer capacity. Normally, however, the electrostatic attraction
of the precursor to the support is so large that the partitioning of the metal
from the solution to the surface is very effective, and little metal loss occurs in
this case.
There is no reason in principle why SEA cannot be performed at incip-
ient wetness conditions, which will add the practical advantages in dry
impregnations of preventing any possible metal loss and eliminating the fil-
tering step. Zhu et al. (65) recently coined this combined SEA-DI approach
as “CEDI” which, in practice, means conducting SEA at incipient wetness.

Fig. 17 Pt/Pd-alloyed nanoparticles over (A) silica, with an X-ray line scan and
(B) carbon, with an X-ray map of an individual nanoparticle (62). From Cho, H.R; et al.
Catal. Today 2015, 246,143–153. Copyright 2015 Elsevier.
Author's personal copy
Preparation of Supported Metal Catalysts 27

That is, a minimal amount of solution is employed, that to just fill the pore
volume of the support, but the solution is sufficiently acidified or basified, as
the case may be, to charge the surface and so induce electrostatic attraction.
They synthesized 2 wt% Pt catalysts supported on oxidized active carbon or
γ-alumina.
Fig. 18 shows the difference between the SEA and CEDI methods sche-
matically. In the CEDI method, the optimal impregnation pH was first
determined, and then dry impregnation was performed.
Zhu et al. (65) also showed that the average size of nanoparticles prepared
by CEDI are much smaller than those prepared by DI. Fig. 19 shows rep-
resentative Z-contrast STEM images for Pt on different support syntheses by
DI and CEDI.
Cao et al. (66) showed CEDI synthesis of highly dispersed Pt/carbon
xerogel catalysts with relatively high Pt loading without any noble metal loss.
In another work, Cao et al. (64) used CEDI to prepare Pd/Mo carbon-
supported bimetallics. Their results for the oxygen reduction reaction indi-
cated that increased activity of the CEDI-prepared samples compared
DI-prepared catalysts, which they claimed to be due to the CEDI method
giving stronger interactions between Pd and Mo. The CEDI sizes are not
quite as small as those obtained from SEA but are close and at any rate
are far better than those obtained with DI. Samad et al. (46) demonstrated
that high metal loadings are achievable even at high SLs. Hence, a well-
dispersed 5.3% Pt (PTA) on silica catalyst at incipient wetness (for which

SEA Dry impregnation CEDI


Excess liquid Pore filling Pore filling
pH pHopt pH PZC pH pHopt
Fig. 18 The difference between the SEA, DI, and CEDI methods schematically (65).
From Zhu, X. R.; et al. ACS Catal. 2013, 3(4), 625–630. Copyright 2013 American Chemical
Society.
Author's personal copy
28 Bahareh A.T. Mehrabadi et al.

Fig. 19 STEM image of Pt catalysts after reduced at 200°C. (A) Pt/C DI, (B) Pt/C CEDI,
(C) Pt/SiO2 DI, (D) Pt/SiO2 CEDI, (E) Pt/Al2O3 DI, and (F) Pt/Al2O3 CEDI (65). From
Zhu, X. R.; et al. ACS Catal. 2013, 3(4), 625–630. Copyright 2013 American Chemical Society.

the support is barely wet by the impregnating solution) could be prepared


without having to lower the metal loading. Their XRD profiles of the
washed and then reduced samples confirm the presence of very small Pt par-
ticles (<1.5 nm). Finally, D’Souza et al. (67) also synthesized Pd/C and
Author's personal copy
Preparation of Supported Metal Catalysts 29

Fig. 20 STEM images of (A) 5 wt% Co + 5 wt% Pd, (B) 5 wt% Co + 2.5 wt% Pd, (C) 5 wt%
Co + 1.25 wt% Pd, and (D) 5 wt% Co supported on Vulcan XC-72 carbon black (VXC) pre-
pared by CEDI (67). Data from D’Souza, L.; et al. Catalysts 2016, 6(5), 72, https://doi.org/10.
3390/catal6050072. Open access publication.

bimetallic CoPd/C heterogeneous catalysts by adopting CEDI. The particle


size distribution, their high metal surface-to-bulk ratios, and synthesis feasi-
bility were held to be superior to any other noble metal bimetallic hetero-
geneous catalyst preparation technique. As shown in Fig. 20, their results
demonstrated that the particle sizes obtained were 3.3 nm in the case of
20 wt%, and 2.9 nm in the case of 5.3 wt% Pd/C.

4. SUMMARY
The nanoparticle synthesis methods reviewed in this chapter have
been compared one against the other in Table 1 which has been excerpted
from Ref. (6). For simplicity’s sake, reductive deposition and deposition
precipitation have been combined into one column. Each method can be
Author's personal copy

Table 1 A Comparison of Methods for Nanoparticle Synthesis on High Surface Area Supports

Single Metals
Charge-Enhanced Deposition
Dry Impregnation Wet Impregnation Strong Electrostatic Dry Impregnation Precipitation
Colloidal (DI) (WI) Adsorption (SEA) (CEDI) (DP)
Simplicity  ++ + + ++ 
Scalability  ++ + + ++ +
Reproducibility
Macroscopic ++ + + ++ ++ ++
Nanoparticle ++   ++ + ++
Efficacy
Size +   ++ + +
Loading ++ ++  + ++ ++
Author's personal copy
Preparation of Supported Metal Catalysts 31

rated for the simplicity of the procedure and its scalability to large quan-
tities (both related to expense), reproducibility from both a macroscopic
level (composition) and a microscopic level (tightness of nanoparticle
size distribution), and efficacy of metal utilization or dispersion (higher
for smaller particles) as well as the ability to achieve high metal loadings.
The range of marks varies from “++” for the best to “ ” for the worst.
Comparing the methods for each criterion (row by row), the very simplest
method to synthesize metal nanoparticles is the fundamentally one-step
method of DI, while the most complexes are the solution methods which
require multiple steps in solution followed by painstaking removal of
the scaffolds (although deposition of bare metal nanoparticles from aqueous
solutions is an exception to this). WI and SEA are a bit more complicated
in that they involve a filtering step after the adsorption step, and the pH
control necessitated by DP makes that method a bit more complicated
than SEA.
The methods of DI, WI, SEA, and CEDI are as applicable to large quan-
tities of catalyst support as they are to small quantities. Scalability is a likely to
be more complex in the solution synthesis of encapsulated metal particles in
large batches, and also in controlling pH for DP at large scale. All methods in
theory are reproducible, though lower marks are given for DI and WI as
drying conditions can exhibit deleterious effects on metal particle size and
distribution in systems in which there is no precursor–support interaction.
Microscopic reproducibility is a measure of the tightness of particle size dis-
tributions, and there is wide variation of rankings for this. Narrow particle
size distributions of supported nanoparticles can also be achieved by SEA and
DP. Dry and wet impregnation methods frequently give poorly dispersed
particles with wide particle size distributions including large agglomerates
of metal. Efficacy in terms of size refers to metal utilization, or metal disper-
sion, with the smallest sizes giving highest dispersion. For metal utilization
over high surface area supports, there is perhaps no method more effective
than SEA. Finally, the best methods simply in terms of depositing high metal
loadings are DI, CEDI, and DP. DP and CEDI are especially effective for
synthesizing relatively small nanoparticles.
Summarizing the table on a column-by-column (method) basis, DI is a
cheap, simple method that very often gives poor and uneven metal disper-
sion as no provision is made to achieve precursor–support interaction. WI
possesses no advantages over DI but has the additional disadvantages of
requiring a filtration step and being limited to low metal loadings. SEA is
more complex than DI, but the payoff of an additional filtration step is
Author's personal copy
32 Bahareh A.T. Mehrabadi et al.

the production of ultra-small metal particles with possibly the tightest size
distributions that can be attained for preparations with high surface area sup-
ports. The results for CEDI obtained to date suggest that particle size is
somewhat larger than can be achieved with SEA conducted in excess liquid.
DP is also a higher complexity method but can give relatively small particles
at relatively high metal loadings.

REFERENCES
1. Nalwa, H. S. Handbook of Surfaces and Interfaces of Materials. Vol. 1. Academic Press:
United States, 2001.
2. Ertl, G.; Kn€ ozinger, H.; Sch€uth, F.; Weitkamp, J. Handbook of Heterogeneous Catalysis.
Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2008; p 4270.
3. Ozkan, U. S. Design of Heterogeneous Catalysts. Wiley-VCH: Weinheim, 2009.
4. Regalbuto, J. R. Catalyst Preparation Science and Engineering. CRC Press: Boca Raton,
FL, 2007.
5. De Jong, K. P. Synthesis of Solid Catalysts. Wiley-VCH Verlag GmbH & Co.:
Netherland, 2009.
6. Zhong, C. J.; et al. In: Comprehensive Inorganic Chemistry II; Reedijk, J.,
Poeppelmeier, R. K., Eds.; Elsevier Inc., 2013; pp 75–102.
7. Munnik, P.; de Jongh, P. E.; de Jong, K. P. Chem. Rev. 2015, 115(14), 6687–6718.
8. Schwarz, J. A.; Contescu, C.; Contescu, A. Chem. Rev. 1995, 95(3), 477–510.
9. Regalbuto, J. R. Catalyst Preparation, Science and Engineering. CRC Press, Taylor &
Francis Group: Boca Raton, FL, 2007.
10. Galhenage, R. P.; et al. Phys. Chem. Chem. Phys. 2015, 17(42), 28354–28363.
11. Tengco, J. M. M. Synthesis of Well Dispersed Supported Metal Catalysts by Strong Electrostatic
Adsorption and Electroless Deposition, Ph.D. Dissertation. University of South Carolina,
SC, United States, 2016.
12. Chan, K. Y.; et al. J. Mater. Chem. 2004, 14(4), 505–516.
13. Louis, C. Catalyst Preparation: Science and Engineering. CRC Press: Boca Raton, 2006.
14. Diao, W. Preparation and Characterization of Pt-Ru Bimetallic Catalysts Using Electroless
Deposition Methods and Mechanistic Study of Re and Cs Promoters for Ag-Based, High Selec-
tivity Ethylene Oxide Catalysts. Ph.D. Dissertation. University of South Carolina, SC,
United States, 2015.
15. Hong, X.; et al. Catal. Sci. Technol. 2016, 6, 3606–3615.
16. Sunagawa, Y.; et al. Catal. Today 2008, 132(1–4), 81–87.
17. Dong-Wook Lee, M.-H. J.; Lee, Y.-J.; Park, J.-H.; Lee, C.-B.; Park, J.-S. Sci. Rep.
2016, 6(26474), 1–9.
18. Agnihotri, S.; et al. RSC Adv. 2014, 4(8), 3974–3983.
19. Liu, H. S.; et al. J. Power Sources 2006, 155(2), 95–110.
20. Luo, J.; et al. Chem. Mater. 2005, 17(21), 5282–5290.
21. Regalbuto, J. R. In: Synthesis of Solid Catalysts; De Jong, K. P. Ed.; Wiley-VCH Verlag
GmbH & Co.: The Netherlands, 2009; pp 33–58.
22. Regalbuto, J. R. In: Silica and Silicates in Modern Catalysis; Halasz, I. Ed.; Transworld
Research Network: Kerala, India, 2010.
23. Regalbuto, J. R. In: Catalyst Preparation Science and Engineering; Regalbuto, J. R. Ed.;
CRC Press: Boca Raton, FL, 2007; pp 297–318.
24. Regalbuto, J. R. In: Surface and Nanomolecular Catalysis; Richards, R. Ed.; Taylor and
Francis Group/CRC Press: Boca Raton, FL, 2006; pp 161–194.
Author's personal copy
Preparation of Supported Metal Catalysts 33

25. Regalbuto, J.R., in Surface and Nanomolecular Catalysis, R. Richards, 2006, CRC
Press: Boca Raton, FL.
26. Brunelle, J. P. Pure Appl. Chem. 1978, 50(9–10), 1211–1229.
27. Heise, M. S.; Schwarz, J. A. J. Colloid Interface Sci. 1985, 107(1), 237–243.
28. Heise, M. S.; Schwarz, J. A. J. Colloid Interface Sci. 1986, 113(1), 55–61.
29. Regalbuto, J. R. In: Synthesis of Solid Catalysts; de Jong, K. P. Ed.; Wiley-VCH:
Weinheim, Germany, 2009.
30. Schreier, M.; Regalbuto, J. R. J. Catal. 2004, 225(1), 190–202.
31. Miller, J. T.; et al. J. Catal. 2004, 225(1), 203–212.
32. Spieker, W. A.; et al. Appl. Catal. A Gen. 2003, 243(1), 53–66.
33. Hao, X.; et al. J. Mol. Catal. A Chem. 2004, 219(1), 97–107.
34. Hao, X.; Barnes, S.; Regalbuto, J. R. J. Catal. 2011, 279(1), 48–65.
35. Feltes, T. E.; et al. J. Catal. 2010, 270(1), 95–102.
36. D’Souza, L. and J.R. Regalbuto, Scientific Bases for the Preparation of Heterogeneous
Catalysts: Proceedings of the 10th International Symposium, 2010. 175: p. 715–718.
37. Spieker, W. A.; Regalbuto, J. R. Chem. Eng. Sci. 2001, 56(11), 3491–3504.
38. Regalbuto, J. R.; et al. J. Catal. 1999, 184(2), 335–348.
39. Hao, X.; Spieker, W. A.; Regalbuto, J. R. J. Colloid Interface Sci. 2003, 267(2), 259–264.
40. Sepulveda-Escribano, A.; Coloma, F.; Rodriguez-Reinoso, F. Appl. Catal. A Gen.
1998, 173(2), 247–257.
41. Figueiredo, J. L.; et al. Carbon 1999, 37(9), 1379–1389.
42. Fraga, M. A.; et al. J. Catal. 2002, 209(2), 355–364.
43. Park, J.; Regalbuto, J. R. J. Colloid Interface Sci. 1995, 175(1), 239–252.
44. Santhanam, N.; et al. Catal. Today 1994, 21(1), 141–156.
45. Schreier, M.; Regalbuto, J. R. J. Catal. 2004, 225(1), 190–202.
46. Lambert, S.; et al. J. Catal. 2009, 261(1), 23–33.
47. Samad, J. E.; Hoenig, S.; Regalbuto, J. R. ChemCatChem 2015, 7(21), 3460–3463.
48. Jiao, L.; Regalbuto, J. R. J. Catal. 2008, 260(2), 329–341.
49. Ling, C. K.; Zabidi, N. A. M.; Mohan, C. J. Appl. Sci. 2011, 11(8), 1436–1440.
50. Jiao, L.; et al. J. Catal. 2008, 260(2), 342–350.
51. Liu, Q. L.; et al. Phys. Chem. Chem. Phys. 2014, 16(48), 26431–26435.
52. Klaigaew, K.; et al. Chem. Eng. J. 2015, 278, 166–173.
53. Akbarzadeh, O.; et al. Process Adv. Mater. Eng. 2014, 625, 328–332. https://doi.org/
10.4028/www.scientific.net/AMM.625.328.
54. Zhang, L. Y.; et al. ChemCatChem 2014, 6(9), 2600–2606.
55. Kyriakidou, E. A.; et al. J. Catal. 2016, 344, 749–756.
56. Riyapan, S.; et al. Catal. Sci. Technol. 2016, 6(14), 5608–5617.
57. Mager, N.; et al. Phys. Chem. Chem. Phys. 2016, 18(47), 32210–32221.
58. Samad, J. E.; et al. J. Catal. 2016, 342, 213–225.
59. Samad, J. E.; et al. J. Catal. 2016, 342, 203–212.
60. Zecevic, J.; et al. Nature 2015, 528(7581), 245–248.
61. Job, N.; et al. Catal. Today 2010, 150(1–2), 119–127.
62. Cho, H. R.; et al. Catal. Today 2015, 246, 143–153.
63. Wong, A.N.; Kyriakidou, E.A.; Toops, T.J.; Regalbuto, J.R.; Catal. Today 2017, 267,
145–156.
64. Cao, C. J.; et al. Appl. Catal. B Environ. 2014, 150, 101–106.
65. Zhu, X. R.; et al. ACS Catal. 2013, 3(4), 625–630.
66. Cao, C. J.; et al. J. Power Sources 2014, 272, 1030–1036.
67. D’Souza, L.; et al. Catalysts 2016, 6(5). https://doi.org/10.3390/catal6050072.
Author's personal copy
34 Bahareh A.T. Mehrabadi et al.

ABOUT THE AUTHORS


Bahareh Alsadat Tavakoli Mehrabadi is a
postdoctoral fellow at the Center of Catalysis
for Renewable Fuels at the University of
South Carolina (USC). She received her
PhD in chemical engineering in 2016 from
USC, performing research in the electro-
chemical field in the area of monometallic
and bimetallic electrocatalysts, fuel cell, and
corrosion. Prior to her PhD, she received her Bachelor’s degree in Chemical
Engineering (2007) and her Master’s degree in Energy Systems Engineering
(2010) both from Sharif University of Technology (SUT) in Tehran, Iran.
She received Industrial Electrochemistry and Electrochemical Engineering
(IEEE) Division Student Achievement Award and University of South
Carolina Student Research Award at 2017.

Sonia Eskandari came to South Carolina


after receiving a Master’s degree from Amir
Kabir University of Technology. Now she
is a fourth-year PhD student in chemical
engineering department at the University
of South Carolina. Her research interests
focus on design and preparation of novel
monometallic and bimetallic catalysts for
the application ranging from biomass con-
version to large-scale catalytic processes.

Umema Khan is a third-year chemical engi-


neering undergraduate student at the Uni-
versity of South Carolina. She is currently
the president of the International Students
Association and has also been a part of the
Society of Women Engineers. The area of
research that she is most interested in is
renewable and alternate energy. She plans
on focusing her career in research in these
areas to further extend her knowledge and
impact.
Author's personal copy
Preparation of Supported Metal Catalysts 35

Rembert White is a second-year under-


graduate student from Irmo, South Carolina.
He studies chemical engineering with an
emphasis in energy at the University of South
Carolina Honors College. Rembert is an
active member of the American Institute of
Chemical Engineers and Engineers Without
Borders USC Chapters and competes in the
Fencing Club. He has a passion for sustain-
able alternatives for energy production and
plans to pursue a career in research of this
field.

John R. (JR) Regalbuto joined the Depart-


ment of Chemical Engineering at the Uni-
versity of South Carolina in 2011 as the
Smartstate (endowed) Chair of the center
of Catalysis for Renewable Fuels, after
25 years at the University of Illinois at Chi-
cago, which included 3 years as a rotator
directing the Catalysis and Biocatalysis Pro-
gram in the Engineering Directorate of the
National Science Foundation. He chaired
the 2014 Gordon Research Conference on
Catalysis and was elected AIChE Fellow in 2016. At USC he is directing
his passion for rational catalyst synthesis toward the optimization of
supported metal catalysts for biomass conversion and solar fuels.

View publication stats

Вам также может понравиться