You are on page 1of 77

Master’s Thesis

SYNTHESIS AND SCREENING OF


BIMETALLIC THREE-WAY CATALYSTS

NGUYEN VU LE

Supervisor: Prof. Dr. Toshiaki Taniike

Graduate School of Advanced Science and Technology


Japan Advanced Institute of Science and Technology

(Material Science)

September 2021
Abstract
Three-way catalysts have played a pivotal role in minimizing the air
pollution from automobiles. However, a non-negligible amount of harmful emission
is still accounted for the cold-start engine. Furthermore, three-way catalysts require
nanoparticles of expensive and unabundant noble metals, especially rhodium, to
realize required performance. Hence, this work applied a series of high-throughput
experiments to construct a library of bimetallic three-way catalysts that allow
exploring information useful in finding more powerful and economic alternatives
for conventional three-way catalysts. In this library, 19 bimetallic combinations
were successfully synthesized, and 60 catalysts were prepared from them by high-
throughput synthetic setups. Later, the catalysts were subjected to in-house high-
throughput fixed bed reactors to evaluate their catalytic activities. With the aid of
statistical analysis, Rh’s alloys on any support materials, and Pd’s alloys when
loaded on CeO2-ZrO2 were found to be superior. Furthermore, Cu’s alloys on CeO2-
ZrO2 demonstrated high selectivity towards CO oxidation. This thesis elucidates the
relationship among constituent elements and supports in designing three-way
catalysts as well as suggests several less costly and/or non-rhodium alternatives.
Acknowledgements
I would like to extend my deepest gratefulness to my supervisor – Prof. Dr.
Toshiaki Taniike – for his enthusiastic instruction throughout this research.

I also wish to express my most genuine gratitude to the lab members of


Taniike’s laboratory for their warm welcome and sincere assist during my research
progress.

Sincere thanks to MITANI SANGYO CO for the financial support program


and to TOYOTA MOTOR Corporation for the supply of support material, which
helped this work to be successful.

1
Contents
List of Tables .............................................................................................................7

Chapter 1: Introduction ...........................................................................................8

1.1 Bimetallic nanoparticles (NPs) .........................................................................8

1.2 Bimetallic alloy synthesis – the polyol process ..............................................10

1.3 Bimetallic alloy and their catalytic activity ....................................................13

1.4 Bimetallic Alloy and Three-way catalyst .......................................................15

1.5 Objectives........................................................................................................18

Chapter 2: Materials andMethods ........................................................................19

2.1 Materials..........................................................................................................19

2.2 HTP Nanoparticle Synthesis ...........................................................................19

2.3 TEM analysis ..................................................................................................21

2.4 XRD measurement ..........................................................................................21

2.5 SEM-EDS measurement .................................................................................22

2.6 Non-destructive quantitative analysis with XRF ............................................23

2.7 Catalyst preparation ........................................................................................25

2.8 HTP catalytic evaluation .................................................................................26

Chapter 3: Results and Discussion ........................................................................29

3.1 Bimetallic alloy library ...................................................................................29

3.1.1 TEM images ..........................................................................................30

3.1.2 XRD results ...........................................................................................37

3.1.3 SEM-EDS results ..................................................................................42

3.2 Library of bimetallic three-way catalysts .......................................................45

3.2.1 Al2O3-based catalysts ............................................................................45

2
3.2.2 TiO2 based catalysts..............................................................................48

3.2.3 CeO2-ZrO2-based catalysts ...................................................................49

3.2.4 CO oxidation .........................................................................................54

3.2.5 Cluster analysis.....................................................................................57

Chapter 4: Conclusion ............................................................................................61

Appendix ..................................................................................................................63

I. TEM of other combinations ............................................................................63

II. Pictures of the synthesizes ..............................................................................64

References ................................................................................................................65

3
List of Figures
Figure 1: Schematic representations of basic mixing patterns: (a) core-shell, (b)
sub-cluster segregated and (c) alloy. Reproduced from Ref. 13. .......................................9
Figure 2: Advantages of the polyol process. Reproduced from Ref. 21. .................... 11
Figure 3: World gross demand by applications of rhodium between 2009 and 2013.
Reproduced from Ref. 49. ............................................................................................................ 14
Figure 4: World gross demand by applications of palladium between 2009 and
2013. Reproduced from Ref. 49. ............................................................................................... 14
Figure 5: World gross demand by application of platinum between 2009 and 2013.
Reproduced from Ref. 49. ........................................................................................................... 15
Figure 6: Parallel reactor setup for the synthesis of nanoparticles. .............................. 20
Figure 7: Synthetic scheme of bimetallic nanoparticles. .................................................. 20
Figure 8: (A) Design and (B) appearance of a XRD sample holder. ........................... 22
Figure 9: Concentration vs XRF intensity graphs for palladium, platinum, rhodium,
ruthenium, and iridium. ................................................................................................................. 24
Figure 10: Catalyst preparation scheme. ................................................................................ 25
Figure 11: Schematic representation of a high-throughput screening system.
Reproduced from Ref. 74. ............................................................................................................ 28
Figure 12: Copper product: (A) after synthesis, (B) after synthesis couple days. .. 30
Figure 13: TEM image library of bimetallic NPs. .............................................................. 33
Figure 14: Aqueous precursors solution of AuPt, AuPd, PdPt, PdRh, PtRh and
AuRh containing PVP. ................................................................................................................... 34
Figure 15: Library of particle size distribution. ................................................................... 35
Figure 16: TEM image of PdAg. .............................................................................................. 36
Figure 17: XRD library of bimetallic NPs. ........................................................................... 39
Figure 18: XRD pattern of PdAg with the powder diffraction file, Ag PDF 00-004-
0783. ..................................................................................................................................................... 40
Figure 19: XRD pattern of RuCu with the powder diffraction file, Cu PDF 00-004-
0836. ..................................................................................................................................................... 40

4
Figure 20: XRD pattern of RhCo with the powder diffraction file, Co PDF 00-015-
0806. ..................................................................................................................................................... 41
Figure 21: XRD pattern of RuCo with the powder diffraction file, Co PDF 01-071-
4239. ..................................................................................................................................................... 41
Figure 22: (A) Appearance of aqueous precursors solution containing PVP (B)
Appearance after the synthesis for RhRu, RhCu, RhCo, RuCu, RuCo and CuCo. .. 42
Figure 23: Appearance after the synthesis, appearance after adding NaOH and
appearance the synthesis with NaOH (pH = 9) for Fe, Cu and Co (left to right). .... 44
Figure 24: Temperature of 50 % C3H6 conversion for Al2O3-based (filled)
bimetallic catalysts and (stripe marked) monometallic catalysts.................................... 46
Figure 25: Temperature of 50 % NO conversion for Al2O3-based (filled) bimetallic
catalysts and (stripe marked) monometallic catalysts. ....................................................... 46
Figure 26: Temperature of 50 % CO conversion for Al2O3-based (filled) bimetallic
catalysts and (stripe marked) monometallic catalysts. ....................................................... 47
Figure 27: Temperature of 50 % C3H6 conversion for TiO2-based (filled)
bimetallic catalysts and (stripe marked) monometallic catalysts.................................... 48
Figure 28: Temperature of 50 % NO conversion for TiO2-based (filled) bimetallic
catalysts and (stripe marked) monometallic catalysts. ....................................................... 49
Figure 29: Temperature of 50 % CO conversion for TiO 2-based (filled) bimetallic
catalysts and (stripe marked) monometallic catalysts. ....................................................... 49
Figure 30: Temperature of 50 % C3H6 conversion for CeO2-ZrO2-based (filled)
bimetallic catalysts and (stripe marked) monometallic catalysts.................................... 50
Figure 31: Temperature of 50 % NO conversion for CeO2-ZrO2-based (filled)
bimetallic catalysts and (stripe marked) monometallic catalysts.................................... 50
Figure 32: Temperature of 50 % CO conversion for CeO2-ZrO2-based (filled)
bimetallic catalysts and (stripe marked) monometallic catalysts.................................... 51
Figure 33: Catalytic activity comparison of bimetallic and monometallic
nanoparticles loaded on different supports in C3H6 oxidation by T50: CeO2-ZrO2
(orange), Al2O3 (gray) and TiO2 (purple). .............................................................................. 52

5
Figure 34: Catalytic activity comparison of bimetallic and monometallic
nanoparticles loaded on different supports in NO reduction by T 50: CeO2-ZrO2
(orange), Al2O3 (gray) and TiO2 (purple). .............................................................................. 53
Figure 35: Catalytic activity comparison of bimetallic and monometallic
nanoparticles loaded on different supports in CO oxidation by T50: CeO2-ZrO2
(orange), Al2O3 (gray) and TiO2 (purple). .............................................................................. 53
Figure 36: T50 comparison among C3H6 (red), NO (blue) and CO (green)
conversion for CeO2-ZrO2 based catalysts. Catalysts with selectivity toward CO
oxidation are caro marked. ........................................................................................................... 54
Figure 37: T50 comparison among C3H6 (red), NO (blue) and CO (green)
conversion for Al2O3-based catalysts. ...................................................................................... 55
Figure 38: T50 comparison among C3H6 (red), NO (blue) and CO (green)
conversion of TiO2-based catalysts. .......................................................................................... 55
Figure 39: Dendogram of agglomerative hierarchical clustering of 58 catalysts.
AuPt/Al2O3 and AuPt/CeO2-ZrO2 were omitted due to the lack of (a part of) T50
values. .................................................................................................................................................. 57
Figure 40: Symmetric plot in multiple correspondence analysis for 60 catalysts. .. 58
Figure A1: TEM images of CuRh, RhCo, RuCo and RuCu. .......................................... 63
Figure A2: Apperance after the synthesis for PdPt, PdRh, PtRh, AuRh, AuPt and
AuPd. ................................................................................................................................................... 64
Figure A3: Apperance after the synthesis for Au, Rh, Ru, Cu, Co and Fe. ............... 64

6
List of Tables
Table 1: Statistical parameters of XRF analysis of palladium, platinum, rhodium,
ruthenium, and iridium. ................................................................................................................. 24
Table 2: Feed gas composition for three-way catalyst evaluation. ............................... 26
Table 3: Target elements and their precursors ..................................................................... 29
Table 4 Summary of successfully synthesized nanoparticles and their
characteristics .................................................................................................................................... 32
Table 5: T50 values for all the catalysts ................................................................................... 47

7
Chapter 1
Introduction
Ancient human has already known how to enhance the physical properties of
metallic material by mixing two or more elements together. One of the very first
metallic combinations humans ever made is bronze, which is an alloy of copper and
tin. The mixed material shows better stiffness than that of copper alone and it was
widely used during the Bronze Age. Comprising metallic elements together not only
strengthens the physical properties of the constituents but also improves other
properties such as chemical properties. As an example, to prevent rusting of steel
due to iron oxidation, a small amount of chromium is blended, and the hybrid
material is nowadays well-known as stainless steel. Likewise, till these modern
days, this strategy is still being applied and extended to nano-sized metals in order
to enhance certain desired properties.

1.1 Bimetallic nanoparticles (NPs)

Unlike in bulk size, metallic nanomaterials consisting of two metallic


elements are generally termed bimetallic NPs. Depending on their distinctive
properties such as atomic ordering, crystal structure, and configuration, bimetallic
NPs can be divided into three main categories: core-shell,1 subcluster segregated,2
and mixed.3
A core-shell bimetallic NP comprising a core of element A contained in a
shell of element B is commonly denoted as A@B (Figure 1-a). The synthesis of
core-shell NPs often employs a seed-mediated growth method: In presence of nano-
seeds of element A, metal B is generated from chemical reduction or thermal
decomposition of a metal B precursor and grows on the seeds.

8
Subcluster segregated bimetallic NPs contain two sub-clusters of the two
constituents sharing an interface or only having a small number of bonding between
elements A and B (Figure 1-b).
A mixed bimetallic NP, commonly known as a bimetallic nanoalloy, is a
solid mixture of two metallic elements (Figure 1-c). The mixture can be randomly
organized or well ordered. A bimetallic nanoalloy can be synthesized by many
methods, for example, thermal decomposition,4–7 electrochemical synthesis,8,9 ion-
implantation,10 chemical co-reduction,11,12 and so on. In this research, this category
of bimetallic NP is mainly focused.

Figure 1: Schematic representations of basic mixing patterns: (a) core-shell, (b)


sub-cluster segregated and (c) alloy. Reproduced from Ref. 13.

9
1.2 Bimetallic alloy synthesis – the polyol process

As mentioned in the above section, mixed bimetallic NPs can be generated in


various ways, such as molecular beam,14,15 electrochemical synthesis,8,9 ion
implantation,10 and thermal decomposition.4–7 Nonetheless, the most common and
direct way to obtain bimetallic nanoalloys is a co-reduction method. 11,12
The co-reduction method is a wet-synthetic pathway involving the
simultaneous reduction of two precursors of the two constituent elements by a
reducing agent. The most important point that needs attention in this method is the
difference in reduction potentials of the two component elements, in which the
element having the higher reduction potential tends to be reduced before the other
with the lower reduction potential. In particular synthesis, precursors such as metal
salts and metal complexes are dissolved in an appropriate solvent in the presence of
a surfactant or called a capping agent and then reduced by a reductant such as H2,
NaBH4, and N2H4.16–18 Two common techniques in this synthetic method are heat-
up and hot-injection methods. In the heat-up method, all ingredients such as
precursors, a capping agent, a reductant, and a solvent are placed together in one
mixture then heated to the temperature where the reaction occurs. While the hot-
injection method requires the mixture of a capping agent, a reductant, and a solvent
heated separately to a desire reaction temperature and then injected with a cold
precursor solution. The hot-injection method is generally more advantageous due to
its versatility in growth control, allowing access to various types of epitaxial
heterostructures.19
Depending on the choices of precursors, reductant, and solvent, co-reduction
synthesis has a wide range of variants. An example is the organometallic pathway
generating NPs from organometallic compounds.20 The method offers mild reaction
conditions that provide better control of particle formation. Furthermore, the
organometallic species, due to their nature, allow no contamination such as halides
or other ions, introduced in the final product. However, organometallic complexes
are expensive, sensitive to be handled, and not widely available. Some other

10
synthetic ways employ a low-boiling point solvent in order to dissolve or dilute a
reducing agent like citrate ion, NaBH4, and N2H4, leading to poorly crystallized
products. Hence, those methods require further thermal treatments, which are
energy consuming to collect well-crystallized particle. Unlike the examples
mentioned above, the polyol process has been proven as a simple and versatile
solution for the problems of the other methods, thus prevailing mixed bimetallic
nanoparticles synthesis on industrial scale.

Figure 2: Advantages of the polyol process. Reproduced from Ref. 21.

The term “polyol process” was firstly introduced in the late 80’s by Fietvet,
Lagier, and Figlazr. They discovered an ability of polyol compounds, which
distinctively possess multiple hydroxyl groups, in metal ion reduction to the
metallic state of noble metals, copper, and some other more electropositive

11
elements such as cobalt and nickel.22–27 Similar to any variants of the co-reduction
method, a particular polyol process consists of metal precursors, a capping agent
(commonly polyvinylpyrrolidone, hereafter abbreviated as PVP), and a polyol,
which works as both the solvent and reductant. These water-like solvents with their
unique characteristics offer numerous benefits as well summarized in Figure 2.
These are productions of well-crystallized nanoparticles due to high temperature
conditions facilitated by their boiling points, protection from oxidation by the
reducing medium, coalescence minimization due to the coordination of polyol
species with metal precursors and metal particles, and good morphological control
because of the diffusion-controlled regime created by their high viscosity.28 This
synthetic strategy was firstly deployed in large scale to generate Co and Ni NPs as
an ingredient of cemented carbides, in the early 90’s by Eurotungstene
Company.29,30 Thereafter, the use of the polyol process is widely expanded in the
industry. For example, due to the absence of contamination when using the polyol
process, Cu and Ag NPs are synthesized for polyurethane production by
Construction Research and Technology GmbH.31 In 2007, the process was also
exploited for the development of polymer-stabilized Ag NPs, which is introduced in
water-based silver ink contained in inkjet-printer products of SEMCO.32 Nowadays,
the polyol process is expanding to fabricate nano-metals in a vast spectrum of
compositions, structures, and application functions. The products from the polyol
process are surely promising candidates for tremendous applications such as in
sensing, fuel-cell, and catalysis.

12
1.3 Bimetallic alloy and their catalytic activity

As well as their monometallic counterparts, bimetallic NPs exhibit unique


optical,33 thermal,34,35 physical,36,37 and chemical properties38,39 those are distinctive
from those of themselves in bulk scale and likely superior to those of their
monometallic constituents. For instance, Pd and Pt NPs are commonly known for
their catalytic activity in the hydrogenation of aromatic hydrocarbons in fuel.
However, catalyst deactivation is the major challenge in this process. The issue is
caused by the agglomeration of Pt NPs in the catalyst when H2S coming from sulfur
impurities of fuel adsorbs and loosens the bonding between the Pt NPs and the
support material.40,41 It has been reported that when making a bimetallic alloy from
Pt and Pd, the product has better catalytic efficiency and stronger resistance to
sulfur poison compared to the monometallic counterparts.38
Catalytic reactions happen on surfaces of catalysts. Due to their high surface
area, metallic nanoparticles play a very important role in catalysis. The most
common catalytic elements are Rh, Pd, Pt and others platinum group metals
(PGMs). These elements catalyze a wide range of reactions including
hydrogenation,42 dehydrogenation,43 coupling,44 carbonylation,45 hydrolysation,46
and oxidation.39,47 Among all, the biggest consumption for Rh, Pd and Pt in the
market comes from the automobile three-way catalyst (TWC), where the global
demand for the three elements accounts for 79 %, 72 %, and 37 % respectively
(Figures 3–5).48,49

13
12

Hundreds oz
10

8 Other
Glass
6
Electrical
Chemical
4
Autocatalyst

0
2009 2010 2011 2012 2013

Figure 3: World gross demand by applications of rhodium between 2009 and 2013.
Reproduced from Ref. 49.

11
Thousands oz

9
Other

7 Jewellery
Investment
5 Electrical
Dental
3 Chemical
Autocatalyst
1

-1 2009 2010 2011 2012 2013

Figure 4: World gross demand by applications of palladium between 2009 and


2013. Reproduced from Ref. 49.

14
9

Thousands oz
8
Other
7
Petroleum
6 Medical & Biomedical
5 Jewellery

4 Investment
Glass
3
Electrical
2
Chemical
1 Autocatalyst
0
2009 2010 2011 2012 2013

Figure 5: World gross demand by application of platinum between 2009 and 2013.
Reproduced from Ref. 49.

1.4 Bimetallic Alloy and Three-way catalyst

Air pollution from automobiles have raised a major concern to modern


society. Ever since the US Environmental Protection Agency (EPA) released the
new emission standard in 1975, tremendous efforts in the development of catalytic
converters have been made. A vehicle catalytic converter is typically composed of a
honeycomb ceramic monolith covered by a wash coat containing TWC. The TWC
wash coat majorly consists of catalytic nanoparticles and support materials. The
most common support material is γ-Al2O3. In addition, CeO2 or CeO2-ZrO2 acting
as an oxygen storage promoter is also added to the wash coat. Other oxides, which
have high surface area and good thermal stability such as SiO2 and TiO2 can also be
used as support materials.50 TWC controls vehicle emission by three-way reactions,
in which Pd or Pt majorly catalyze the oxidations of light hydrocarbons and CO to
CO2 while Rh is mainly responsible for the reduction of NOx to N2.47,51 The three-
way reactions are described below:

15
Reduction of NOx to N2:

𝐶 + 2𝑁𝑂 → 𝐶𝑂 + 2𝑁𝑂

2𝐶𝑂 + 2𝑁𝑂 → 2𝐶𝑂 + 𝑁

4𝐶𝑂 + 2𝑁𝑂 → 4𝐶𝑂 + 𝑁

2𝐻 + 2𝑁𝑂 → 2𝐻 𝑂 + 𝑁

Oxidation of carbon, CO and hydrocarbons (e.g C3H6) to carbon dioxide:

𝐶 + 𝑂 → 𝐶𝑂

2𝐶𝑂 + 𝑂 → 2𝐶𝑂

2𝐶 𝐻 + 9𝑂 → 6𝐶𝑂 + 6𝐻 𝑂

Despite the best performance of these elements, the production cost of TWC
still remains high. Especially Rh which is the most essential component for NO x
reduction is one of the least abundant elements in the Earth’s crust. 52,53
Furthermore, low-temperature exhaust, or light-off exhaust contributes a notable
portion to the total emission of automobiles.54 To overcome these challenges, many
strategies have been proposed including modifying support materials,55,56 doping
rare earth57,58 and alkaline metals59,60 promoters, changing ratios of PGM
loading,61,62 and applying bimetallic alloy as catalytic particles.63 In this research,
substituting conventional PGM with bimetallic alloys in TWC is the main focus.

In past literature, numerous bimetallic combinations, such as PdAg,64


PdCu,65,66 PdIr,67 PdNi,68 PdRh,69–71 PdRu,3,72 PtRh,39 and RhIr,73 have been

16
designed to serve the purpose. Depending on the nature of the parent elements and
the atomic ratios, the bimetallic alloys seem to exhibit better catalytic performances.
For instance, Pd60Rh40 has been reported to show better activity and higher stability
as compared to Pd and Rh standalone particles.70 Another example is the pseudo-Rh
nanoparticles, synthesized from two naturally immiscible elements Pd and Ru by
Kusada et al. and later by Sato et al.3,72 Their Pd50Ru50 alloy was demonstrated to
have a similar electronic state to that of Rh nanoparticles, which are commonly used
for NOx reduction.

Despite many successful publications on this topic, it is believed that there


are still a lot more corners of these bimetallic NPs, which have not yet been
explored. Most of the past works were done in one-pot synthesis direct only on
support materials. Hence, potential weaknesses in the majority of previous studies
are at their improper characterization methods. In addition, the employed synthetic
routes in the previous research are different from each other. These limitations of
traditional experimental design provide ambiguous and sometimes inadequate
knowledge about bimetallic combinations. For example, the improvement effect in
TWC by alloying two elements is still a controversial subject. In some articles,
bimetallic catalysts such as PtRh39 and PdRh63,69 were reported to have dominant
catalytic activity and better thermal stability in three-ways reactions. Meanwhile,
some others proclaimed bimetallic alloys to be inferior to conventional ones. 47,54
Furthermore, the performance comparison of the reported bimetallic alloys is
difficult to be made since the experimental procedures applied on them are unalike.

17
1.5 Objectives

To overcome the weakness mentioned in the previous section and also to


bring a different viewpoint in the relevant fields, this thesis adopts a series of
specially designed high-throughput experiments. The bimetallic alloy synthesis was
conducted using an in-house designed high-throughput parallel reactor system
following a single synthetic procedure. A series of the synthesized NPs were
characterized with a proposed set of analytical methods to construct a
comprehensively informative library of bimetallic alloys. Moreover, the bimetallic
alloys in the library were loaded on various types of support materials in a high-
throughput manner. The prepared catalysts were subjected to the performance
evaluation using an in-house developed state-of-the-art high-throughput fixed-bed
reactor system. The high-throughput experimental setups only require a small
sample amount and can generate massive data in a short time. With the aid of
statistical analysis, this library unravels major parameters, such as the choices of
metal elements and support materials, which enormously influence to the three-way
reactions. Thereby, the findings proved that bimetallic alloys have predominant
catalytic abilities in three-way processes due to their alloying synergism.
Furthermore, some of the more economical or non-rhodium candidates which can
potentially substitute rhodium NPs were suggested. In short, the data library from
this research is to bring a broader picture in NPs synthesis as well as catalyst design.

18
Chapter 2
Materials and Methods

2.1 Materials

Hydrogen tetrachloroaurate (III) trihydrate (HAuCl3-H2O) was purchased


from MP Biomedicals. Hydrogen hexachloroplatinate (IV) hexahydrate (H2PtCl6-
6H2O, ≥ 98.5 %), potassium tetrachloropalladate (II) (K2PdCl4, 99 %), rhodium (III)
chloride trihydrate (RhCl3-3H2O, 99.5 %), copper (II) chloride dihydrate (CuCl2-
2H2O, 99 %), triethylene glycol (99 %), and polyvinylpyrrolodine (PVP,
(C6H9NO)n, K30, average MW = 40000) were purchased from Wako. Colbalt (II)
chloride (CoCl2, 97 %) palladium (II) chloride (PdCl2, ≥ 99.9 %), ruthenium (III)
chloride hydrate (RuCl3-xH2O, Ru 46.8 %) and iridium chloride hydrate (IrCl3-
xH2O, Ir 54.6 %) were purchased from Sigma Aldrich. Hydrochloric acid (35.0 % ~
37.0 %), acetone, hexane, and methanol of research grade are purchased from Kanto
Chemical. γ-Al2O3 (PURALOX SCFa100, SASOL, BET surface area of 106 m 2/g),
TiO2 (Aeroxide P25, EVONIK, BET surface area of 55 m2/g), and CeO2-ZrO2
(nanosized) were used as support materials.

2.2 HTP Nanoparticle Synthesis

All the samples were synthesized using an in-house parallel reactor setup,
which has the maximum capacity of carrying 6 parallel reactions (Figure 6). In a
particular synthesis (Figure 7), 88 mg of polyvinylpyrrolidone (PVP) as a capping
agent was dissolved in 20 ml of triethylene glycol (TEG) as a solvent in a 40 ml
glass vessel. In the cases of Cu-based bimetallic alloys, pH was controlled by
adding an adequate amount of a 30 % NaOH aqueous solution to adjust the pH of
the TEG-PVP solution to 9. The solution was vigorously stirred under N 2, followed
by heating to 190 oC. Subsequently, 2.0 ml of an aqueous solution dissolving metal

19
precursors (0.050 mmol per metal element) and 22 mg of PVP were swiftly injected
into the TEG-PVP mixture. After a dark color indicating the formation of NPs
appeared, the reaction mixture was kept for 30 min, and then naturally cooled to
room temperature. Synthesized nanoparticles were washed with an acetone/hexane
mixture (3/1 v/v), and recovered by centrifugation, which was repeated three times.
Finally, the product was re-dispersed and stored in methanol before use.

Figure 6: Parallel reactor setup for the synthesis of nanoparticles.

Figure 7: Synthetic scheme of bimetallic nanoparticles.

20
2.3 TEM analysis

Transmission electron microscopic (TEM) images were collected on a


Hitachi H-7100 electron microscope operated at 100kV. To prepare samples, one
drop of nanoparticles in methanol was dropped onto a Cu mesh and allowed to dry
naturally for 24 hours. A sample pool of 200 particles was taken to determine the
average particle size and the particle size distribution.

2.4 XRD measurement

Powder X-ray diffraction data (XRD) was acquired on a Rigaku SmartLab


X-ray diffractometer using a Cu-Kα radiation source. Sample preparation was
performed by dropping a few drops of a condensed solution of nanoparticles in
methanol onto a hole (0.5 mm in dia.) of a specially designed silica sample holder.
The holder design is detailed in Figure 8.

Vegard’s law states that a lattice parameter of a solid solution of two


constituents is approximated by a weighted mean of the two constituents’ lattice
parameters in the same temperature. Hence, when alloy formation is successful, the
peak position of the bimetallic nano-alloy must be different from those of the two
relevant monometallic NPs. By applying this outcome, alloy formation was
confirmed by checking the difference in the peak position of the bimetallic NPs
from those of the monometallic counterparts.

21
A B

1.2 cm

1.7 cm

4.4 cm
1.6 cm
1.3 cm

3.0 cm

2 mm
1.75
mm

Figure 8: (A) Design and (B) appearance of a XRD sample holder.

2.5 SEM-EDS measurement

Elemental analysis was performed using a scanning electron microscope


equipped with an energy-dispersive X-ray spectrometer (SEM-EDS, TM3030 Plus
Microscope). Sample preparation was made by adding one drop of a NPs solution
onto a carbon tape fixed on a sample holder. The sample holder with a sample
droplet was left to dry in the atmosphere for 24 hours before the analysis. In a
particular measurement, EDS signals were captured at nine different locations on
the dried sample droplet and averaged for deriving the chemical composition. In
addition to the XRD peak position shifting, the homogeneity of EDS signals in
different areas was also used as indirect support for alloy formation.

22
2.6 Non-destructive quantitative analysis with XRF

The concentration of nanoparticles in a solution was quantified on a X-ray


fluorescence spectrometer (XRF, PanAlytical Epsilon 3) based on the external
calibration method for Pd, Pt, Rh, Ru, and Ir, and with the aid of SEM-EDS results
for the other elements. The instrument is equipped with an Ag X-ray tube, operated
with a Cu filter having 500 μm thickness. The X-ray tube and generator can operate
under a potential range from 4 kV to 50 kV and a current range from 1 μA to 3000
μA, providing the maximum power of 15W. Measurements were conducted under
air with the acquisition time of 60 seconds. 1000 ppm (mg/kg) stock standard
solutions of five elements, Pd, Pt, Rh, Ru, and Ir, were prepared by dissolving
respectively PdCl2, H2PtCl6-6H2O, RhCl3-3H2O, RuCl3-xH2O and IrCl3-xH2O in a
5% hydrochloric aqueous solution. Later, multi-element standard solutions in
different concentrations were prepared from the stock to make calibration curves.
Nanoparticles dispersed in 2/3 v/v methanol/DI water were diluted from 10 to 100
times with DI water (16 MΩ). 3.0 ml of a standard solution or a sample solution
was added into a sample cup (35 mm in dia.) supported by a 6 μm-thick
polypropylene film. This analytical method was validated by comparing the
elemental composition of Pd, Pt, Rh, Ru and Ir-containing bimetallic nanoparticles
between XRF and SEM-EDS based on t-test at 95% confident level. Furthermore,
the linearity, the lowest detection limit (LoD), and the lowest quantitation limit
(LoQ) for each element were evaluated and listed in Table 1.

23
Table 1: Statistical parameters of XRF analysis of palladium, platinum, rhodium,
ruthenium, and iridium.

LoD* LoQ**
Element Analytical Equation R2
(mg/kg) (mg/kg)

Pd I = (1.3274 ± 0.0535)[Pd] – (0.6137 ± 3.5670) 0.9995 5.0 15.0

Pt I = (3.9318 ± 0.1496)[Pt] – (4.4908 ± 9.6716) 0.9996 1.5 4.5

Rh I = (1.3484 ± 0.0234)[Rh] – (2.7276 ± 1.4398) 0.9997 9.6 28.8

Ru I = (0.9378 ± 0.0330)[Ru] – (0.6494 ± 1.9761) 0.9994 1.5 4.5

Ir I = (2.6228 ± 0.1118)[Ir] – (1.1883 ± 6.8874) 0.9991 5.0 15.0

*Lowest detection limit (LoD)


**Lowest quantitation limit (LoQ)

450
400
350
300
Intensity (cps)

Pd
250
Pt
200
Rh
150
Ru
100
Ir
50
0
0 20 40 60 80 100 120
Concentration (mg/kg)

Figure 9: Concentration vs XRF intensity graphs for palladium, platinum, rhodium,


ruthenium, and iridium.

24
2.7 Catalyst preparation

In order to prepare catalysts from the synthesized nanoparticles, a wet


impregnation method was employed (Figure 10). 200 mg of support material and
roughly 2 ml (depending on the concentration) of a nanoparticle solution in 2/3 v/v
methanol/DI water were added into a 5 ml vial in order to have the final product
containing 0.3 wt% of nanoparticles. The suspension was then stirred thoroughly
for 24 hours at 60 oC. After that, the suspension was dried in an oven at 80 oC
overnight. The dried product was milled with mortar and pestle, and subjected to
the calcination for 2 hours at 550 oC under a normal atmosphere. The calcined
catalyst was milled to fine powder and used for the catalytic test. For one batch of
preparation, 20 catalysts were prepared simultaneously with a parallel heating
magnetic stirrer VELP AREX 6 equipped with MultilAluBlock.

Figure 10: Catalyst preparation scheme.

25
2.8 HTP catalytic evaluation

Table 2: Feed gas composition for three-way catalyst evaluation.

Feed gas component Concentration (ppm)

CO 26000

C3H6 4000

NO 6000

O2 28000

CO2 (10%) 100000

He (83.6%) 836000

The HTP instrument, developed in our laboratory, consists of seven major


components, a gas mixer, a flow distributor, a hollow furnace, quartz reaction tubes,
an auto-sampler, a quadrupole mass-spectrometer (QMS, Transpector CPM 3,
INFICON), and an exhausted system. In the beginning, the constituent gasses are
mixed thoroughly in the gas mixer (MU-3504, HORIBA STEC) to make a model
automobile exhaust gas with desired concentrations for each component. Next, the
mixed gas travels up to the flow distributor, where the flow is equally split into 20
channels. Later, the divided flows pass through 20 reaction quartz tubes located in
the middle of the hollow furnace. Each quartz tube has inner diameters of 4 mm for
the inlet end and 2 mm for the outlet end. Catalyst powder is fixed by quartz wool at
the middle section of each tube, corresponding to the neck position between two
different inner diameters. The furnace possesses three heating sectors and in each of
which, a ceramic heater with a thermocouple is set. The heating program was set to
rise from 100 oC to 400 oC with the ramping rate of 4 oC/min and for every 20 oC

26
increment, the temperature is kept for half an hour in order to evaluate catalytic
activity at that particular temperature before it continues to rise.

After the reaction, the effluent gas continues to enter the autosampler that is
connected to the QMS. Auto sampling is performed by a programmed action
sequence of pneumatically actuated diaphragm valves (MEGA ONE, Fujikin). In a
typical sampling, a sampling time is 1.2 seconds followed by vacuum evacuation
for 10 seconds. For one lap of measurement, the time is 224 seconds in total. Any
leftover effluent gas is evacuated to the exhaust system. During the measurement,
the sampled gas is continuously transferred from the autosampler to the QMS, and
signals of pre-designated mass fractions are recorded.
In a typical test, before performing the three-ways reaction, loaded catalysts
were pre-activated by calcination at 550 oC for one hour in the hollow furnace. After
that, the model exhaust gas is prepared by blending all components with the gas
mixer, following the gas composition listed in Table 2, and then the TWC reaction
was started. The light-off performance of catalysts was evaluated as the temperature
of 50 % conversion (T50) for C3H6, CO, and NO, respectively.

27
Figure 11: Schematic representation of a high-throughput screening system.
Reproduced from Ref. 74.

28
Chapter 3
Results and Discussion

3.1 Bimetallic alloy library

Table 3: Target elements and their precursors.

25 26 27 28
Fe Co Ni Cu
(FeCl3) (CoCl2) (NiCl2) (CuCl2)
44 45 46 47
Ru Rh Pd Ag
(RuCl3) (RhCl3) (K2PdCl4) (AgNO3)
77 78 79
Ir Pt Au
(IrCl3) (H2PtCl6) (HAuCl3)

The scope of this library originally focused on bimetallic combinations of


transition metals and noble metals, which are listed in Table 3. These elements were
chosen because in history they were found to have catalytic activity in three-way
reactions.75 To keep synthesis conditions in constant, all metal precursors are salts
or metal complexes of chloride. Furthermore, to understand the reducibility of each
element, a molar ratio of 50/50 mol/mol was aimed in all synthesis.

The characterization was conducted based on three comprehensive analytical


methods, TEM, XRD, and SEM-EDS. These methods were chosen due to their
comprehension in characterization, availability, simplicity, and compatibility with
the throughput of the other experiments. To recall, TEM was employed to analyze
the particle size distribution, as well as the dispersion of NPs XRD, was used to
confirm the formation of alloy based on the peak position shifting, and SEM-EDS
to measure the average atomic ratios and check their homogeneity. The data was
then used to construct a bimetallic alloy library and is discussed below.

29
3.1.1 TEM images

Figure 12: Copper product: (A) after synthesis, (B) after synthesis couple days.

Based on the polyol process, six monometallic nanoparticles and 19


bimetallic combinations, which are listed in Table 3, were attempted to synthesize.
Monometallic nanoparticles were first synthesized in order to get XRD references
and also to check the reducibility of each element. The difficult elements to be
reduced in this library were elements such as Cu, Co, and Fe, where pH control was
implemented to generate more anionic species from TEG as the main reduction
agent in the polyol process.21 In the case of Cu, in spite of the successful reduction,
the product was unstable and oxidized after a few days when kept dispersed in
methanol (Figure 12). On the other hand, Au NPs has the biggest sizes, in which the
smallest particle has a diameter larger than 100 nm. (Figure 13). It is believed that
this overgrowth phenomenon is because of the low coordination of Au precursor
with PVP, which will be discussed later with AuPt bimetallic nanoparticles. Nano-
sized products were successfully synthesized for other noble metals, Pt, Pd, Rh, Ru,
and Ir, with good average particle sizes of 3.6 ± 0.6 nm, 3.2 ± 0.9 nm, 2.5 ± 0.4, 0.8
± 0.1 nm, and below 0.7 nm respectively. Furthermore, all of these NPs showed
good dispersion in methanol.

30
For bimetallic synthesizes, AuRh, AuPd, and AuPt had average particle sizes
of 8.5 ± 2.5 nm, 9.6 ± 1.7, and 11.6 ± 2.1 respectively, which were 3 to 10 times
higher than those of the other bimetallic nanoparticles in the library (Table 4). The
results indicate that weak coordination between Au precursor and PVP pays the
major cause for the large size distribution. Furthermore, good dispersions were
found in nearly all cases except for AuPt. Unlike the synthesis of the other
bimetallic alloys, the addition of a small amount (22 mg) of PVP into the precursor
solution was omitted for AuPt and Au to prevent clogging during hot-injection (the
presence of PVP in Au precursors caused precipitation which can be seen in Figure
14). Severe precipitation of PVP was also observed in the synthesis of Au
nanoparticles. The precipitation of PVP suggests weak coordination between the
capping agent and Au precursor causing the lack of growth control as well as the
poor dispersion. Later in the catalyst evaluation section, this poor dispersion
appeared to influence greatly to the catalytic activity of AuPt regardless of the
support on which it was loaded.

Large deviations in particle size distribution were also found in Cu’s alloys
(Figure 15) in which Cu is the lowest reduction potential element among the used
elements. Non-uniform growth behavior of these alloys can be explained by a large
difference in the reduction rate caused by the reduction potential gaps between two
elements. Indeed, the Pt, Pd, Rh, Ru, and Ir precursors possess similar reduction
potentials76 and in this library, the bimetallic combinations among them led to
nicely uniform particle size distributions (Figure 15). Unlike any other bimetallic
NPs, PdAg grew into rod-like shape particles (Figure 16), plausibly due to the
formation of a stable AgCl precipitate during the synthesis. AgCl traps and slowly
releases Ag+ into the TEG medium, causing the slow reduction of silver. Another
explanation could be that AgCl suppresses the reduction of Ag+ to Ag by
modulating its reduction potential.76 In any case, the inhibition of Ag+ reduction can
be clearly seen later in the SEM-EDS results.

31
Table 4 Summary of successfully synthesized nanoparticles and their
characteristics.

Average Average
Crystal
Name (AB) Element A Element B element ratio particle size
structure
(A : B, at%) (nm)
Au Au - - Fcc ≥100
Pt Pt - - Fcc 3.6 ± 0.6
Pd Pd - - Fcc 3.2 ± 0.9
Rh Rh - - Fcc 2.5 ± 0.4
Ru Ru - - Hcp 0.8 ± 0.1
Ir Ir - - Fcc ≤ 0.7
AuRh Au Rh 46 : 54 Fcc 8.5 ± 2.5
AuPd Au Pd 46 : 54 Fcc 9.6 ± 1.7
AuPt Au Pt 50 : 50 Fcc 11.6 ± 2.1
PtRh Pt Rh 47 : 53 Fcc 2.8 ± 0.4
PdPt Pd Pt 55 : 45 Fcc 3.2 ± 0.4
PdRh Pd Rh 51 : 49 Fcc 2.7 ± 0.6
RhRu Rh Ru 51 : 49 Fcc 1.7 ± 0.3
PdRu Pd Ru 57 : 43 Fcc 2.6 ± 1.5
PtRu Pt Ru 47 : 53 Fcc 2.5 ± 0.6
RhIr Rh Ir 51 : 49 Fcc 1.7 ± 0.2
PdIr Pd Ir 57 : 43 Fcc 2.5 ± 0.5
PtIr Pt Ir 47 : 53 Fcc 2.1 ± 0.4
RhCu* Rh Cu 73 : 27 Fcc 1.7 ± 0.7
PdCu* Pd Cu 46 : 54 Fcc 2.6 ± 1.3
PtCu* Pt Cu 57 : 43 Fcc 2.0 ± 0.5
CuRh Cu Rh 6 : 94 Fcc ~ 5.0
RhCo Rh Co 78 : 22 Fcc 1.4 ± 0.2
RuCo Ru Co 80 : 20 Hcp 0.7 ± 0.2
RuCu Ru Cu 85 : 15 Fcc 0.9 ± 0.2
PdAg Pd Ag 70 : 30 Fcc n.d.
*pH-controlled reduction with NaOH, in which pH = 9

32
Figure 13: TEM image library of bimetallic NPs.

33
Figure 14: Aqueous precursors solution of AuPt, AuPd, PdPt, PdRh, PtRh and
AuRh containing PVP.

34
Figure 15: Library of particle size distribution.

35
Figure 16: TEM image of PdAg.

36
3.1.2 XRD results

As mentioned in the method section, XRD was employed in order to confirm


the formation of bimetallic alloy by Vegard’s law. In 1921, Lars Vegard, a
Norwegian physicist, discovered that the lattice parameter of a solid solution of two
constituents is approximately a weighted mean of the two constituents’ lattice
parameters at the same temperature:

𝑎 ( )
= 𝑥𝑎 + (1 − 𝑥)𝑎

where 𝐴 and 𝐵 are the two constituents, 𝑥 is the molar fraction of 𝐴 . 𝑎 , 𝑎 ,


𝑎 ( )
are the lattice parameters of 𝐴, 𝐵, and the solid mixture constituted of 𝐴

and 𝐵 with the molar fraction 𝑥 of 𝐴, respectively. His empirical finding dictates
that the XRD peak position of a solid solution lies between the positions of its two
components. In the case of bimetallic alloy, the XRD peak of the nano-alloy is
located in between the XRD peaks of the two monometallic counterparts.

Successful alloy formation was confirmed for all of the combinations among
Au, Pt, Pd, Rh, Ir, and Cu, (Figure 17) as well as for PdAg, RuCu, RhCo, and RuCo
(Figure 19-21 respectively). From the XRD patterns, the face-centered cubic (fcc) is
the dominant crystal structure in most of the cases except Ru monometallic and
RuCo NPs that exhibited the hexagonal close-packed (hcp) crystal lattice. In the
cases of bimetallic alloys made of either Pt, Pd, and Ir, more careful consideration
was required to judge their alloying since the references of those elements have
very similar XRD peak positions and so do their alloys. The successful alloying was
judged from the fact that the products were composed of a single crystalline phase
and that the chemical composition was uniform in SEM-EDS. These would not be
observed when a core-shell structure or a segregated structure was formed.

As mentioned earlier, because Cu has a reduction potential largely different


from those of noble metals in the list, the non-uniform reduction was expected.

37
Indeed, in SEM-EDS that will be discussed in the later section, RhCu did not
achieve the desired average element ratio of 50:50 even with the use of pH control.
Nonetheless, RhCu possessed a board XRD peak, which was suspected to arise
from a mixture of nanoparticles with different chemical compositions, probably
including 50:50 ones.

The successful alloy formations were also confirmed for RhCo, RuCo and
RuCu. These bimetallic combinations possessed very small average particle sizes as
can be seen in TEM images (Figure A1). Hence, the XRD signals of these
combinations were less intense. However, their alloying could be justified with
SEM-EDS results (discussed later).

38
Rh Pd Pt
Rh Au
Pd
AuRh AuPt
AuPd Au
Au
Au

Intensity (a.u)

Intensity (a.u)

Intensity (a.u)
Au

20 40 60 80 20 40 60 80
20 40 60 80
2θ (degree) 2θ (degree)
2θ (degree)

Rh
PtRh Pt
Pt PdPt
Pd

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

Pt

20 40 60 80 20 40 60 80
2θ (degree) 2θ (degree)

Rh
PdRh
Pd
Intensity (a.u)

Pd

20 40 60 80
2θ (degree)

Rh Pd Pt
RhRu PdRu PtRu
Ru Ru Ru

Ru
Intensity (a.u)

Intensity (a.u)

Intensity (a.u)

20 40 60 80 20 40 60 80 20 40 60 80
2θ (degree) 2θ (degree) 2θ (degree)

Rh Pd Pt
RhIr PdIr PtIr
Ir Ir Ir

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

Intensity (a.u)

20 40 60 80 20 40 60 80 20 40 60 80

2θ (degree) 2θ (degree) 2θ (degree)

Rh Pd Pt
RhCu PdCu PtCu
Cu (PDF 00-004-0836) Cu (PDF 00-004-0836) Cu (PDF 00-004-0836)
Intensity (a.u)
Intensity (a.u)

Intensity (a.u)

Cu*

20 40 60 80 20 40 60 80
20 40 60 80
2θ (degree) 2θ (degree)
2θ (degree)

* pH-controlled reduction with NaOH, in which pH = 9

Figure 17: XRD library of bimetallic NPs.

39
Pd
PdAg
Ag (PDF 00-004-0783)
Intensity (a.u)

20 40 60 80
2θ (degree)
Figure 18: XRD pattern of PdAg with the powder diffraction file, Ag PDF 00-004-
0783.

Ru
RuCu
Cu (PDF 00-004-0836)
Intensity (a.u)

20 40 60 80
2θ (degree)
Figure 19: XRD pattern of RuCu with the powder diffraction file, Cu PDF 00-004-
0836.

40
Rh
RhCo
Co (PDF 00-015-0806)
Intensity (a.u)

20 40 60 80
2θ (degree)
Figure 20: XRD pattern of RhCo with the powder diffraction file, Co PDF 00-015-
0806.

Ru
RuCo
Co (PDF 01-071-4239)
Intensity (a.u)

20 40 60 80
2θ (degree)
Figure 21: XRD pattern of RuCo with the powder diffraction file, Co PDF 01-071-
4239.

41
3.1.3 SEM-EDS results

Figure 22: (A) Appearance of aqueous precursors solution containing PVP (B)
Appearance after the synthesis for RhRu, RhCu, RhCo, RuCu, RuCo and CuCo.

Besides checking the desired average atomic ratio of 50:50 for each
constituent, the homogeneity of EDS results over the measured nine positions of the
nanoparticles droplet was also exploited as supplemental information to confirm the
formation of alloys. The homogeneity of EDS signals was found for all the
bimetallic combinations, which supports the XRD results indicating the successful
formation of bimetallic alloys. Desired average element ratios were achieved in
most of the cases except PdAg, RhCu, RhCo, RuCo, and RuCu (Table 4). Recall
from Section 3.1.1, TEM images, due to the formation of AgCl, the reduction of
Ag+ reduction was inhibited causing the rod-like shape particles as well as the
unwanted molar element ratio of 70 % for Pd and 30 % for Ag. In the cases of Cu,
obtaining the 50:50 atomic ratio tended to be difficult. As also mentioned in Section

42
3.1.1, in order to successfully reduce a low reduction potential element such as Cu,
Co, and Fe, pH control was required. Nearly desired elemental ratios were
confirmed for Pd46Cu54 and Pt57Cu43, while it was not achieved for RhCu. Without
pH control, this combination stood at 6 % of Cu and 94 % of Rh, whereas with the
TEG solution at pH = 9, the content of Cu in the new product was increased to 30
%. Nevertheless, as mentioned in the previous section, from the XRD pattern, it was
considered that RhCu produced under the pH control partly contained nanoparticles
with an elemental ratio close to 50:50.

To demonstrate the significance of pH control in metal reduction with the


polyol process, the first attempt to synthesize monometallic particles of Cu, Co, and
Fe, and CuCo bimetallic alloy was without pH control. After the reaction, the
solution was clear and exhibited the absence of any solid products (Figure 22). To a
solution, 30 % NaOH was dropwise added to make the pH of the TEG solution to 9
and this caused the color change (Figure 23). The solution was then heated up to
190 oC and kept for 30 minutes. After the reduction, a dark-colored solid product
was obtained (Figure 23).

In the case of RhCo, RuCo and RuCu, due to the reduction potential gap or
in other word, the difficulty in the reduction of Cu and Co, none of these bimetallic
combinations successfully led to the 50:50 ratio. However, the EDS results
demonstrated that reduction of low reduction potential elements such as Cu and Co
was more viable in presence of noble metals than the reduction of these elements
alone. In fact, Cu and Co were partly reduced in presence of Rh and Ru while it was
failed to do so without the noble metals. In addition, homogeneity in the EDS signal
of these species somewhat indicated them as bimetallic alloys.

43
Figure 23: Appearance after the synthesis, appearance after adding NaOH and
appearance the synthesis with NaOH (pH = 9) for Fe, Cu and Co (left to right).

44
3.2 Library of bimetallic three-way catalysts

The library of bimetallic three-way catalysts was composed of 60 different


catalysts prepared from 20 nano units of Rh, Pd, Pt, Ru, and Ir, including 15
bimetallic combinations and 5 monometallic counterparts, loaded on three different
support materials of Al2O3, TiO2, and CeO2-ZrO2. Here, the trending and unique
phenomenon of the catalysis data for these 60 catalysts are discussed and compared
with results and statements of some other previous studies from other research
groups.

3.2.1 Al2O3-based catalysts

As shown in Figure 24-26, and Table 5, PtRh led to the best light-off
performance on Al2O3 with 50% conversion (T50) at 263 oC, 251 oC, and 238 oC for
C3H6, NO, and CO respectively. In addition, the majority of bimetallic alloys
exhibited lower T50 values than their monometallic counterparts and this is
absolutely in contrast with many previous statements.47,54,77 Performat combinations
mostly contained Rh, which indicates that the three-way reactions favor Rh as an
essential element. Bimetallic alloys such as AuRh, RhCu, and RhRu, PdRh and
PtRh had very competitive performance to Rh monometallic NPs, hence it is indeed
feasible to reduce the production cost of conventional TWC with more economical
element than Rh while enhancing its activity by alloying with Rh. Besides, some
non-rhodium-containing alloys like PdCu and PdRu could also be considered as less
expensive and more abundant alternatives since the abundance of Rh on Earth’s
surface is the rarest in PGM.52,53 The result of PdRu from this library is likely
consistent with that of PdRu synthesized by Kusada et al. and later by Sato et al.
The reason behind PdRu competitive catalytic activity to Rh is the formation of
pseudo – Rh particles that mimic the electronic state of Rh NPs by alloying two
immiscible elements, Pd and Ru.3,72 On the other hand, the remarkable catalytic
activity of PdCu nano-alloy is likely explained due to the stabilization of Pd

45
metallic state, where is the adsorption site for hydrocarbon and CO, made by Cu
when the alloy of Pd-Cu formed.47,66,78,79 In contrast, AuPt was the only species that
did not show any activity between the temperatures of 100 oC and 400 oC. As
mentioned in Section 3.1.1, the disability in three-way catalysis seemed to be
contributed by the poor dispersion of AuPt.

400

350

300
T50 (oC)

250

200

150

Figure 24: Temperature of 50 % C3H6 conversion for Al2O3-based (filled)


bimetallic catalysts and (stripe marked) monometallic catalysts.

400

350

300
T50 (oC)

250

200

150

Figure 25: Temperature of 50 % NO conversion for Al2O3-based (filled) bimetallic


catalysts and (stripe marked) monometallic catalysts.

46
400

350

300

T50 (oC)
250

200

150

Figure 26: Temperature of 50 % CO conversion for Al2O3-based (filled) bimetallic


catalysts and (stripe marked) monometallic catalysts.

Table 5: T50 values for all the catalysts.


T50 for Al2O3 (oC) T50 for CeO2-ZrO2 (oC) T50 for TiO2 (oC)
Name
C3H6 NO CO C3H6 NO CO C3H6 NO CO
Pt 361.5 354.5 365.4 324.4 320.0 333.2 348.2 348.0 383.7
Pd 315.6 340.0 351.9 268.0 265.2 268.8 327.4 326.8 338.2
Rh 288.3 284 262.4 266.0 260.0 253.8 287.7 282.1 254.8
Ru 361.1 356.0 342.9 330.1 362.7 278.0 330.0 328.7 334.0
Ir 363.7 374.4 351.0 366.2 379.9 356.0 378.6 386.7 347.2
AuRh 289.6 285.0 276.1 264.7 255.2 253.8 288.8 282.5 277.4
AuPd 375.6 375.3 387.6 312.6 311.8 312.2 348.9 342.9 346.1
AuPt - - - - - 351.3 372.7 366.9 391.4
PtRh 262.7 251.0 237.9 251.0 248.1 247.1 266.6 247.8 234.6
PdPt 337.2 355.1 379.0 283.0 284.2 278.3 328.0 327.6 338.8
PdRh 268.4 268.0 255.9 251.0 249.4 254.6 288.7 284.0 351.9
RhRu 288.3 282.0 274.2 282.2 267.3 254.8 304.2 286.8 263.2
PdRu 290.9 289.0 279.5 254.4 254.2 250.6 328.4 326.5 336.8
PtRu 366.4 345.7 354.7 317.5 353.3 283.4 335.8 325.5 347.6
RhIr 286.1 285.0 273.9 266.8 261.3 253.5 287.9 284.5 255.7
PdIr 312.1 335.8 344.6 280.0 279.6 275.2 331.2 329.6 337.1
PtIr 367.6 370.3 379.6 328.3 348.3 306.3 367.8 360.0 387.0
RhCu* 285.2 281.0 254.0 268.0 266.2 216.2 306.6 294.5 278.2
PdCu* 296.2 337.5 296.3 267.5 268.0 210.2 327.8 328.0 338.3
PtCu* 352.2 351.7 339.2 314.4 324.9 230.4 327.6 349.1 356.4
*pH-controlled reduction with NaOH, in which pH = 9

47
3.2.2 TiO2 based catalysts

TiO2-based catalysts series generally exhibited a similar trending as in the


Al2O3 equivalents (Figure 27-29) that emphasized the favoritism of three-way
reaction toward rhodium. As well as PtRh/Al2O3, PtRh/TiO2 was also the best
catalytic unit in the series. In addition, the catalytic activity of most combinations
was much close to those of Al2O3-based counterparts. However, enhanced activity
was observed for AuPd, PdPt, and AuPt when loaded on TiO2, as compared to
Al2O3. Especially in the case of AuPt, the activity was detected for the three
pollutants in the temperature range of 100 oC to 400 oC, which was unlike its
siblings on Al2O3 previously reported to have no activity, and in the later section,
CeO2-ZrO2 having only CO conversion. Oppositely, despite relatively strong
catalytic abilities to the other units loaded in TiO2, when comparing to their Al2O3
equivalents, PdRu, PdCu, and RhCu were weaker in the performance of the three
reactions. In addition, although PdRh still performed well in NO conversion and
C3H6 conversion, its performance in CO conversion was significantly poor. It was
suspected that, with TiO2, PdRh appeared to have more selectivity toward NO
reduction with CO and/or incomplete oxidation of C3H6 to CO.

400

350

300
T50 (oC)

250

200

150

Figure 27: Temperature of 50 % C3H6 conversion for TiO2-based (filled)


bimetallic catalysts and (stripe marked) monometallic catalysts.

48
400

350

300
T50 (oC)

250

200

150

Figure 28: Temperature of 50 % NO conversion for TiO2-based (filled) bimetallic


catalysts and (stripe marked) monometallic catalysts.
400

350

300
T50 (oC)

250

200

150

Figure 29: Temperature of 50 % CO conversion for TiO2-based (filled) bimetallic


catalysts and (stripe marked) monometallic catalysts.

3.2.3 CeO2-ZrO2-based catalysts

Once again, similar to the trending in the Al2O3- and TiO2-based catalysts,
the preference of Rh-containing particles in TWC reactions was demonstrated in
Figure 30-32. The general catalytic mechanism in supported noble metal is the
reversible redox process between the metal state, M0, and an oxidized state
MOx.47,77,80–82 The reason for Rh superior activity in three-way reactions perhaps is
that each Rh atom can exchange 3 electrons while in other elements, such as Pd, Pt,
and Cu, only 2 electrons are exchanged.77,83

In the mid-1990s, a synergy between Pt and Rh was discovered. 84,85 Pt and


Rh in their study were in form of separate monometallic particles, rather than

49
bimetallic alloy. The proposed mechanism of the synergy was that the Rh metal
state is stabilized by H2 spillover from neighboring Pt particles. However, in this
study, the employed feed gas did not contain H2 and only use CO as the main
reductant. The outstanding performance of PtRh in this research surely did not
come from the same mechanism proposed by Hu et al, but from alloy formation.
The alloy synergism in PtRh could be because of the Rh3+-Rh0 light-off redox
process facilitated by Pt2+. 77,86
400

350

300
T50 (oC)

250

200

150

Figure 30: Temperature of 50 % C3H6 conversion for CeO2-ZrO2-based (filled)


bimetallic catalysts and (stripe marked) monometallic catalysts.

400

350

300
T50 (oC)

250

200

150

Figure 31: Temperature of 50 % NO conversion for CeO2-ZrO2-based (filled)


bimetallic catalysts and (stripe marked) monometallic catalysts.

50
400

350

300
T50 (oC)

250

200

150

Figure 32: Temperature of 50 % CO conversion for CeO2-ZrO2-based (filled)


bimetallic catalysts and (stripe marked) monometallic catalysts.

As seen from Figure 33-35 the catalytical powers of nanoparticles on CeO2-


ZrO2 were overall better than their equivalents on Al2O3 and TiO2. The superiority
of CeO2-ZrO2 in TWC conversion is due to the ability of CeO2 to directly take part
in the redox reactions by the reversible process between Ce3+ and Ce4+.
Furthermore, this redox process also allows the material to store and releasing
oxygen, thus it improve the catalytic power.47,54,77,87

𝐶𝑒𝑂 + 𝛿𝐶𝑂 → 𝐶𝑒𝑂 + 𝛿𝐶𝑂

𝛿
𝐶𝑒𝑂 + 𝑂 → 𝐶𝑒𝑂
2

Besides, CeO2 also owns extraordinary metal-support interaction47,88–91 that


appears to be the major explanation for the prevailing activity of CeO 2 catalyst
family. Though many hypotheses had been proposed, the lattice oxygen migration
and the formation of strong bonding between metal and support, M-O-Ce, are
generally attributed to the mechanism of CeO2 interaction with noble metal
nanoparticles. A strong metal-support interaction indicates high dispersion and
sintering prevention of metal nanoparticles over CeO2. For instance, a model

51
involving the lability of lattice oxygen was proposed by Yentekakis et al. In their
model, labile lattice oxygen from CeO2 spillovers and covers metal particles like Ir
and Rh to create a negatively charged layer on the surface of the particles. This
negatively charged surface makes particles repel each other and prevent
agglomeration among them during high-temperature operation. 89,92,93 Similar
mechanism that also involving the migration of lattice oxygen was also reported in
the interaction of Pd and Pt with CeO2.94–96 In the cases of metal support bonding, it
was reported that strong bonding of Pt with CeO2, i.e. Pt-O-Ce, inhibits the
sintering of these nanoparticles during high-temperature performance.47,90 Similarly,
the unique effect of metal bonding with CeO2 was also discovered in Ru
nanoparticles, where the formation of stable bonding, Ru-O-Ce, led to very high
dispersion of Ru on CeO2.81,97 It was also reported that loading metal nanoparticles
facilitate the reducibility of Ce4+ by weakening Ce-O bonds.95,98

400

350
T50 of C3H6 conversion (oC)

300

250

200

150

Figure 33: Catalytic activity comparison of bimetallic and monometallic


nanoparticles loaded on different supports in C3H6 oxidation by T50: CeO2-ZrO2
(orange), Al2O3 (gray) and TiO2 (purple).

52
400

350

T50 of NO conversion (oC)


300

250

200

150

Figure 34: Catalytic activity comparison of bimetallic and monometallic


nanoparticles loaded on different supports in NO reduction by T50: CeO2-ZrO2
(orange), Al2O3 (gray) and TiO2 (purple).

400

350
T50 of CO conversion (oC)

300

250

200

150

Figure 35: Catalytic activity comparison of bimetallic and monometallic


nanoparticles loaded on different supports in CO oxidation by T50: CeO2-ZrO2
(orange), Al2O3 (gray) and TiO2 (purple).

On CeO2-ZrO2, Pd nanoparticles, as well as Pd containing bimetallic


nanoalloys, became competitive with the Rh counterparts in all three reactions. The
enhancement of the catalytic activity of Pd when on CeO 2 or relevant materials can
be generally explained by the ability of CeO2 to stabilize the Pd(I) oxidation
state.47,99 The intermediate Pd(I) state that is made by unique metal-support
interaction pays a tremendous benefit to the elimination of C3H6 and CO. As
mentioned in the previous sections, PdRu as pseudo-Rh particles was reported to
have similar catalytic activity, when it was loaded on Al2O3.3,72 Comparing with this
result, the pollutant removal power of PdRu was greatly amplified on CeO2-ZrO2
(Figure 33-35). It is believed that despite sharing similar properties with Rh
particles, PdRu still inhered the intrinsic properties of its parent elements, and to be

53
more specific it is the special metal-support interaction between Pd and CeO2. Thus,
it is certainly another type of synergistic effect of alloy formation.

Recall that AuPt exhibited an extremely poor catalytic ability toward all
three redox processes of pollutants removal on both Al2O3 and TiO2 and that this
poor performance was plausibly attributed to the poor dispersion of as-synthesized
nanoparticles themselves. With no exception, even by loading on CeO2-ZrO2 with
good oxygen storage capacity (OSC), AuPt ineffectively converted CO to CO 2.
Furthermore, there was no significant conversion of C3H6 and NO detected with
AuPt/CeO2-ZrO2.

However, in all the top-level catalysts, which showed good light-off


conversion, most of them are bimetallic nano-alloys. This trending demonstrates the
supreme catalytic ability of bimetallic alloys over their monometallic counterparts
and indicates the synergistic effect of alloying, not just the simple synergism of the
co-presence of two kinds of monometallic nanoparticles.

3.2.4 CO oxidation

400

350

300
T50 (oC)

250

200

150

Figure 36: T50 comparison among C3H6 (red), NO (blue) and CO (green)
conversion for CeO2-ZrO2 based catalysts. Catalysts with selectivity toward CO
oxidation are caro marked.

54
400

350

300

T50 (oC)
250

200

150

Figure 37: T50 comparison among C3H6 (red), NO (blue) and CO (green)
conversion for Al2O3-based catalysts.

400

350

300
T50 (oC)

250

200

150

Figure 38: T50 comparison among C3H6 (red), NO (blue) and CO (green)
conversion of TiO2-based catalysts.

Before entering this section, one distinct phenomenon is necessary to be kept


in mind that there is synchronization among CO, NO, and C3H6 conversions in most
of the catalysts in this library (Figure 36-38). Namely, the T50 values for the three
reactions tend to be close to each other. In addition, the ranking order of nano units
within each type of supports is mostly similar among the three reactions (Figure 24-
31). However, there are several exceptions, which are selective toward the CO
conversion. Hence, this sub-chapter is dedicated in order to have a deeper look at
these interesting behaviors.

The selective behavior was found when some nanoparticles were loaded on
CeO2-ZrO2. In Figure 36, catalysts with Ru or Cu containing nanoparticles have
much lower T50 values in CO conversion than those of NO and C3H6 conversions.
In particular, Ru/CeO2-ZrO2 and PtRu/CeO2-ZrO2 exhibited much higher catalytic

55
activity in CO conversion than their siblings of Al2O3 and TiO2 (Figure 35).
Previously, Ru had been reported to be favored in CO oxidation.82 The fundamental
of CO removal with Ru involves the formation of RuOx as active species.
Moreover, CeO2 was found to anchor and stabilize this oxide species.95,100 Recalling
the lattice oxygen migration mechanism mentioned above,88,89,92,93 it was believed
that the same mechanism occurred to Ru and its alloy. Once the labile oxygen from
CeO2-ZrO2 covers the Ru constituent particles, more of RuOx active site is
generated on the catalyst surface, hence the CO conversion is enhanced.

Extraordinary selectivity toward CO oxidation was also found in catalysts of


Cu-containing alloys on CeO2-ZrO2. Unlike C3H6 and NO conversions, where PtRh
is still the champion of the rankings, PdCu, RhCu, and PtCu are the top three
catalysts in the CO conversion (Figure 32). This tremendous selectivity towards CO
oxidation can be most clearly seen in PtCu/CeO2-ZrO2, which performed poorly in
NO and C3H6 conversions. Before this work, Abdelsayed et al reported relevant
results101, where PdCu, RhCu, and PtCu loaded on CeO2 exhibited better CO
conversion than PtRh and PdRh. The promotion effect of Cu for the CO oxidation
by alloying it with several noble metals could be explained that when Cu and its
alloy are supported on a CeO2 constituent material, there is the formation of Cu+
species that play a role as highly selective chemisorption sites for CO.66,78,102
Comparing to the Ru containing particles on CeO2-ZrO2, the ones of Cu have
significantly better CO conversion. This behavior will be seen once more in the
later section.

56
3.2.5 Cluster analysis

PdIr/Al2O3
Pd/Al2O3
PtCu/TiO2
Ru/TiO2
PdIr/TiO2
PdPt/TiO2
PdCu/TiO2
Pd/TiO2
PdRu/TiO2
Pt/CeO2.ZrO2
PtRu/TiO2
PdRh/TiO2
AuPd/CeO2.ZrO2
PdCu/Al2O3
Ru/CeO2.ZrO2
PtRu/CeO2.ZrO2
PtIr/CeO2.ZrO2
PtCu/CeO2.ZrO2
Ir/CeO2.ZrO2
Ir/Al2O3
Ir/TiO2
PtIr/TiO2
AuPt/TiO2
PtIr/Al2O3
AuPd/Al2O3
PtRu/Al2O3
Pt/Al2O3

Poor
Ru/Al2O3
AuPd/TiO2
PtCu/Al2O3
Pt/TiO2
PdPt/Al2O3
PtRh/CeO2.ZrO2
PdRh/CeO2.ZrO2
PdRu/CeO2.ZrO2
PtRh/TiO2
Good
PtRh/Al2O3
Pd/CeO2.ZrO2
PdRh/Al2O3
RhIr/CeO2.ZrO2
Rh/CeO2.ZrO2
AuRh/CeO2.ZrO2
RhCu/CeO2.ZrO2
PdCu/CeO2.ZrO2
RhIr/TiO2
Rh/TiO2
RhCu/Al2O3
Rh/Al2O3
RhRu/CeO2.ZrO2
AuRh/TiO2
RhRu/Al2O3
AuRh/Al2O3
RhIr/Al2O3
PdRu/Al2O3
PdPt/CeO2.ZrO2
PdIr/CeO2.ZrO2
RhRu/TiO2
RhCu/TiO2

0 1 2 3 4 5 6
Dissimilarity

Figure 39: Dendrogram of agglomerative hierarchical clustering of 58 catalysts.


AuPt/Al2O3 and AuPt/CeO2-ZrO2 were omitted due to the lack of (a part of) T50
values.

57
Figure 40: Symmetric plot in multiple correspondence analysis for 60 catalysts.

Cluster analysis is a data analytical grouping of a set of objects in the way


that objects in the same groups are more similar to each other than those in other
groups. In this section, the results of the catalyst test are interpreted statistically or
data scientifically by applying two cluster analysis methods, i.e. agglomerative
hierarchical clustering (AHC), and multiple correspondence analysis (MCA) to
investigate the common features of the bimetallic catalyst and compare the finding
to the above literature research.

The AHC iteratively classifies catalysts based on the dissimilarity among the
58 catalysts, which was calculated from the reduced Euclidean distance of the three
T50 values for C3H6 oxidation, NO reduction, and CO oxidation. The algorithm
omitted AuPt/Al2O3 and AuPt/CeO2-ZrO2 due to the missing T50 values.
Homogenous groups were created by the complete linkage method. In Figure 39, it
was found that there are two major groups presented in the dendrogram, good
catalysts (green color), whose T50 values are mostly below 300 oC, and poor
catalysts (red color), whose values are much above 300 oC. As being discussed in

58
the previous sections, Rh-containing catalysts were demonstrated to be dominant in
three-way reactions. Furthermore, CeO2-ZrO2 based catalysts, especially with Pd-
containing particles, exhibited remarkably better catalytic activity than those on the
other two supports. The same trending can also be seen in the dendrogram in which
most of the catalysts are either Rh alloys on different supports or Pd alloys loaded
on CeO2-ZrO2.

MCA generated a set of coordinate values for non-numerical variables


allowing presenting them in the graphic. A number of dimensions (factors) in the
MCA was extracted from the variables by a scree test. In Figure 40, the behavior of
three-way catalysts is presented by different categorical variables including the
presence/absence of constituent elements, support materials, and good/poor
performances in C3H6 oxidation, NO reduction, and CO oxidation. When a T50
value for a certain conversion is below 300 oC, such a catalyst is regarded as good
in terms of the said conversion, and vice versa. As mentioned in the earlier section
(3.3.4 CO oxidation), collinearity among the three reactions appears as the clearest
trending, i.e. the catalyst that performs well in one reaction usually performs well
also in the other two reactions. This behavior is described clearly in Figure 40,
where the locations for good performances in the three reactions are very close with
each other and so for those for the poor ones. Once again, the predominance of Rh-
containing catalysts and CeO2-ZrO2 based catalysts are located nearby the good
performances. On the other hand, the elemental variables for Pt, Ir, and Au as well
as the support variable for TiO2 are clustered together with poor performance dots.
Thereby, the poor catalysts tend to contain elements among Pt, Ir and Au, and/or
TiO2. Other variables, support-Al2O3, element-Pd, element-Cu, and element-Ru
scatter in the space between good performance dots and poor performance dots,
indicating that depending on the combination, catalysts that possess these
constituents can either have good or poor catalytic performances. Indeed, as
described in the previous sub-chapters, Pd-based alloys exhibited higher
performances when loaded on CeO2-ZrO2 than those loaded on Al2O3 or TiO2. Cu
and Ru alloys when anchored on CeO2-ZrO2 possess notable conversion on CO

59
oxidation. Of course, Ru constituted particles on CeO2-ZrO2 have weaker power
than Cu ones, which is once again indicated by the distance of the Ru dot from the
good CO oxidation dot is further than that of the Cu dot. In addition, Al2O3-based
catalysts equipped with Rh-containing particles are drastically better than those
without Rh.

60
Chapter 4
Conclusion

In this thesis, 19 bimetallic nanoalloys were successfully synthesized based


on a high-throughput reactor setup under a unified synthetic procedure to afford a
firm library of bimetallic nanoalloys. The particle size distribution and element
ratios were differed greatly depending on the choice of elements and precursors.
Reduction potential gaps made a major contribution to the particle size distribution.
The larger the gap between the two elements was, the less uniform the particle size
distribution became. Also, the difference in the reduction potential was identified as
the greatest challenge in acquiring the desired element ratio, especially in the cases
of low reduction potential elements, such as Cu, Co, and Fe. The majority of as-
synthesized nanoparticles were featured with the face-centered cubic crystal lattice.
This library provides an overview and a guideline serving the purpose of
engineering metallic nanoparticles and their alloys.

From the library of bimetallic nanoalloys, a library of bimetallic three-way


catalysts was established. It was proven that bimetallic alloys have superior
catalytic activities in three-way reactions compared to their monometallic
counterparts. The improvement in the catalytic power of bimetallic alloys certainly
comes from the synergistic effect of alloying metallic elements. Moreover, the
library highlighted the important relationship between the choices of elements and
the choices of support materials and thereby widely derives a diversity of possible
strategies to improve three-way catalyst performance as well as to lower its
manufacturing cost. As an example, the amount of Rh can be minimized by alloying
with less expensive elements such as Cu while the catalytic ability remains still or
even prevail to the pristine Rh. In another route, the synergism of Pd-based alloys
and CeO2-ZrO2 offers non-Rh candidates, such as PdCu/CeO2-ZrO2 and
PdRu/CeO2-ZrO2, whose performances were competitive to Rh constituent ones.
Furthermore, Cu-based alloy loaded on CeO2-ZrO2 showed exceptional CO

61
conversion. Surely, this extended version of the library brings a comprehensive
picture in catalyst design, and it is believed that with the application of statistical
analysis such as clustering, the scope of this library can be extended furthermore.

62
Appendix
I. TEM of other combinations

CuRh RhCo

RuCo RuCu

Figure 41: TEM images of CuRh, RhCo, RuCo and RuCu.

63
II. Pictures of the synthesizes

Figure 42: Apperance after the synthesis for PdPt, PdRh, PtRh, AuRh, AuPt and
AuPd.

Figure 43: Apperance after the synthesis for Au, Rh, Ru, Cu, Co and Fe.

64
References

1. Spitale A, Perez MA, Mejía-Rosales S, Yacamán MJ, Mariscal MM. Gold-


palladium core@shell nanoalloys: Experiments and simulations. Phys Chem
Chem Phys. 2015;17(42):28060-28067.

2. Jiao Z, Sivayoganathan M, Duley WW, He P, Zhou YN. Correction to


“formation and characterization of femtosecond-laser-induced subcluster
segregated nanoalloys.” J Phys Chem C. 2015;119(2):1268.

3. Kusada K, Kobayashi H, Ikeda R, et al. Solid solution alloy nanoparticles of


immiscible Pd and Ru elements neighboring on Rh: Changeover of the
thermodynamic behavior for hydrogen storage and enhanced co-oxidizing
ability. J Am Chem Soc. 2014;136(5):1864-1871.

4. Bönnemann H, Richards RM. Nanoscopic metal particles - Synthetic methods


and potential applications. Eur J Inorg Chem. 2001;(10):2455-2480.

5. Esumi K, Tano T, Torigoe K, Meguro K. Preparation and Characterization of


Bimetallic Pd-Cu Colloids by Thermal Decomposition of Their Acetate
Compounds in Organic Solvents. Chem Mater. 1990;2(5):564-567.

6. Bradley JS, Via GH, Bonneviot L, Hill EW. Infrared and EXAFS study of
compositional effects in nanoscale colloidal palladium-copper alloys. Chem
Mater. 1996;8(8):1895-1903.

7. Thomas JM, Johnson BFG, Raja R, Sankar G, Midgley PA. High-


Performance Nanocatalysts for Single-Step Hydrogenations. Acc. Chem. Res.
2003;36(1):20-30.

8. Kolb U, Quaiser SA, Winter M, Reetz MT. Investigation of


tetraalkylammonium bromide stabilized palladium/platinum bimetallic
clusters using extended X-ray absorption fine structure spectroscopy. Chem
Mater. 1996;8(8):1889-1894.

65
9. Reetz MT, Helbig W, Quaiser SA. Jl. 1995;(7):2227-2228.

10. Mattei G, Maurizio C, Mazzoldi P, et al. Dynamics of compositional


evolution of Pd-Cu alloy nanoclusters upon heating in selected atmospheres.
Phys Rev B - Condens Matter Mater Phys. 2005;71(19):1-11.

11. Sra AK, Schaak RE. Synthesis of atomically ordered AuCu and AuCu 3
nanocrystals from bimetallic nanoparticle precursors. J Am Chem Soc.
2004;126(21):6667-6672.

12. Chi M, Wang C, Lei Y, et al. Surface faceting and elemental diffusion
behaviour at atomic scale for alloy nanoparticles during in situ annealing. Nat
Commun. 2015;6:1-9.

13. Ferrando R, Jellinek J, Johnston RL. Nanoalloys: From theory to applications


of alloy clusters and nanoparticles. Chem Rev. 2008;108(3):845-910.

14. Kürten KE, Kusmartsev F V. Fractal structures in systems made of small


magnetic particles. Phys Rev B - Condens Matter Mater Phys. 2005;72(1).

15. Milani P, Iannotta S. Molecular Beams and Cluster Nucleation. Published


online 1999:5-34.

16. Burda C, Chen X, Narayanan R, El-Sayed MA. Chemistry and Properties of


Nanocrystals of Different Shapes. Vol 105.; 2005.

17. Goia D V., Matijević E. Preparation of monodispersed metal particles. New J


Chem. 1998;22(11):1203-1215.

18. Toshima N, Yonezawa T. Bimetallic nanoparticles - Novel materials for


chemical and physical applications. New J Chem. 1998;22(11):1179-1201.

19. Chen J, Ma Q, Wu XJ, Li L, Liu J, Zhang H. Wet-Chemical Synthesis and


Applications of Semiconductor Nanomaterial-Based Epitaxial
Heterostructures. Nano-Micro Lett. 2019;11(1):1-28.

20. Amiens C, Chaudret B, Ciuculescu-Pradines D, et al. Organometallic


approach for the synthesis of nanostructures. New J Chem. 2013;37(11):3374-

66
3401.

21. Fievet F, Ammar-Merah S, Brayner R, et al. The polyol process: a unique


method for easy access to metal nanoparticles with tailored sizes, shapes and
compositions. Chem Soc Rev. 2018;47(14):5187-5233.

22. F. FieÅLvet, J.-P. Lagier, B. Beaudouin and M. Figlarz, Reactivity of Solids,


Materials Science Monographs 28A, Elsevier, Amsterdam, 1985, pp. 555–
556.
23. M. Figlarz, F. FieÅLvet and J.-P. Lagier, Cobalt Metallurgy and Uses, Cobalt
Development Institute, London, 1986, pp. 185–197.
24. F. FieÅLvet, J.-P. Lagier and M. Figlarz, MRS Bull., 1989, 14, 29–34.
25. F. FieÅLvet, J.-P. Lagier, B. Blin, B. Beaudoin and M. Figlarz, Solid State
Ionics, 1989, 32-33, 198–205.
26. M. Figlarz, F. FieÅLvet and J.-P. Lagier, US4539041, University Paris VII,
1985.
27. Fievet F, Lagier JP, Blin B, Beaudoin B, Figlarz M. Homogeneous and
heterogeneous nucleations in the polyol process for the preparation of micron
and submicron size metal particles. Solid State Ionics. 1989;32-33(PART
1):198-205.

28. Nguyen TD. From formation mechanisms to synthetic methods toward shape-
controlled oxide nanoparticles. Nanoscale. 2013;5(20):9455-9482.

29. Koeleman J. Eurotungstene Poudres moves towards ISO 9002.


1994;(March):30-33.

30. Fievet F, Lagier JP, Blin B, Beaudoin B, Figlarz M. Homogeneous and


heterogeneous nucleations in the polyol process for the preparation of micron
and submicron size metal particles. Solid State Ionics. 1989;32-33(PART
1):198-205.

31. Data RUSA, Kerber RA, Sandra J, City SL. Patent Application Publication (
10 ) Pub . No .: US 2011 / 0207128A1. 2011;1(19).

32. Lee KJ, Lee Y Il, Shim IK, Jun BH, Cho HJ, Joung JW. Large-Scale

67
Synthesis of Polymer-Stabilized Silver Nanoparticles. Solid State Phenom.
2007;124-126:1189-1192.

33. Liu S, Chen G, Prasad PN, Swihart MT. Synthesis of monodisperse Au, Ag,
and Au-Ag alloy nanoparticles with tunable size and surface plasmon
resonance frequency. Chem Mater. 2011;23(18):4098-4101.

34. Sohn H, Xiao Q, Seubsai A, et al. Thermally Robust Porous Bimetallic


(NixPt1- x) Alloy Mesocrystals within Carbon Framework: High-Performance
Catalysts for Oxygen Reduction and Hydrogenation Reactions. ACS Appl
Mater Interfaces. 2019;11(24):21435-21444.

35. Liu X, Wang A, Yang X, et al. Synthesis of thermally stable and highly active
bimetallic Au-Ag nanoparticles on inert supports. Chem Mater.
2009;21(2):410-418.

36. Zhou J, Shu X, Wang Z, et al. Hydrothermal synthesis of polyhedral FeCo


alloys with enhanced electromagnetic absorption performances. J Alloys
Compd. 2019;794:68-75.

37. Ponraj R, Thirumurugan A, Jacob GA, Sivaranjani KS, Joseyphus RJ.


Morphology and magnetic properties of FeCo alloy synthesized through
polyol process. Appl Nanosci. 2020;10(2):477-483.

38. Stanislaus A, Barry HC. Aromatic hydrogenation catalysis: A review. Catal


Rev. 1994;36(1):75-123.

39. Yamashita H, Ogami T, Kanamura K. Hydrothermal synthesis and catalytic


activity of PtRh/CeO2/Al2O3 three-way catalysts for automotive exhaust gas.
J Ceram Soc Japan. 2018;126(5):394-401.

40. Lopes T, Paganin VA, Gonzalez ER. The effects of hydrogen sulfide on the
polymer electrolyte membrane fuel cell anode catalyst: H2S-Pt/C interaction
products. J Power Sources. 2011;196(15):6256-6263.

41. Chang JR, Chang SL, Lin TB. γ-Alumina-supported Pt catalysts for aromatics
reduction: A structural investigation of sulfur poisoning catalyst deactivation.

68
J Catal. 1997;169(1):338-346.

42. G. C. Bond. Platinum Metals as Hydrogenation Catalysts. Platin Met Rev.


1957;1(3):87-93. http://www.platinummetalsreview.com/article/1/3/87-93/

43. Novell-Leruth G, Valcárcel A, Pérez-Ramírez J, Ricart JM. Ammonia


dehydrogenation over platinum-group metal surfaces. Structure, stability, and
reactivity of adsorbed NHx species. J Phys Chem C. 2007;111(2):860-868.

44. Das SK, Parandhaman T, Pentela N, Maidul Islam AKM, Mandal AB,
Mukherjee M. Understanding the biosynthesis and catalytic activity of Pd, Pt,
and Ag nanoparticles in hydrogenation and Suzuki coupling reactions at the
nano-bio interface. J Phys Chem C. 2014;118(42):24623-24632.

45. Didgikar MR, Roy D, Gupte SP, Joshi SS, Chaudhari R V. Immobilized
palladium nanoparticles catalyzed oxidative carbonylation of amines. Ind Eng
Chem Res. 2010;49(3):1027-1032.

46. Rakap M. Poly(N-vinyl-2-pyrrolidone)-stabilized palladium-platinum


nanoparticles-catalyzed hydrolysis of ammonia borane for hydrogen
generation. J Power Sources. 2015;276:320-327.

47. Wang J, Chen H, Hu Z, Yao M, Li Y. A review on the Pd-based three-way


catalyst. Catal Rev - Sci Eng. 2015;57(1):79-144.

48. Supply T. PLATINUM DEMAND BY APPLICATION : REGIONS.


Published online 1999.

49. Jhonson & Matthew. Supply and demand tables. Platin 2013 Interim Rev.
Published online 2013:30-35.

50. DieselNet. Catalytic coating and materials. 2005. Retrived from


https://dieselnet.com/tech/cat_mat.php

51. Montini T, Melchionna M, Monai M, Fornasiero P. Fundamentals and


Catalytic Applications of CeO2-Based Materials. Chem Rev.
2016;116(10):5987-6041.

69
52. Ore I, Pigments IO, Rock P, Crystal Q, Earths R, Ash S. Mineral Commodity
Summaries 2021.; 2021.

53. Hans Wedepohl K. The composition of the continental crust. Geochim


Cosmochim Acta. 1995;59(7):1217-1232.

54. Rood S, Eslava S, Manigrasso A, Bannister C. Recent advances in gasoline


three-way catalyst formulation: A review. Proc Inst Mech Eng Part D J
Automob Eng. 2020;234(4):936-949.

55. Chen HY, Chang HLR. Development of low temperature three-way catalysts
for future fuel efficient vehicles. Johnson Matthey Technol Rev.
2015;59(1):64-67.

56. Liang C, Qiu J, Li Z, Li C. Synthesis of nanostructured ceria, zirconia and


ceria-zirconia solid solutions using an ultrahigh surface area carbon material
as a template. Nanotechnology. 2004;15(7):843-847.

57. Wang Q, Li G, Zhao B, Zhou R. The effect of rare earth modification on


ceria-zirconia solid solution and its application in Pd-only three-way catalyst.
J Mol Catal A Chem. 2011;339(1-2):52-60.

58. Guo J, Shi Z, Wu D, Yin H, Gong M, Chen Y. Effects of Nd on the properties


of CeO2-ZrO2 and catalytic activities of three-way catalysts with low Pt and
Rh. J Alloys Compd. 2015;621:104-115.

59. Lan L, Li H, Chen S, et al. CeO2–ZrO2–Al2O3 Modified by Selective Doping


with SrO for Improved Pd-Only Three-Way Catalyst. Russ J Phys Chem A.
2018;92(4):696-705.

60. Lan L, Chen S, Cao Y, et al. Promotion of CeO2-ZrO2-Al2O3 composite by


selective doping with barium and its supported Pd-only three-way catalyst. J
Mol Catal A Chem. 2015;410:100-109.

61. Cooper J, Beecham J. A study of platinum group metals in three-way


autocatalysts. Platin Met Rev. 2013;57(4):281-288.

62. Alikin EA, Vedyagin AA. High Temperature Interaction of Rhodium with

70
Oxygen Storage Component in Three-Way Catalysts. Top Catal. 2016;59(10-
12):1033-1038.

63. Vedyagin AA, Gavrilov MS, Volodin AM, et al. Catalytic purification of
exhaust gases over Pd-Rh alloy catalysts. Top Catal. 2013;56(11):1008-1014.

64. Fernandes VR, Bossche M Van Den, Knudsen J, et al. Reversed Hysteresis
during CO Oxidation over Pd75Ag25(100). ACS Catal. 2016;6(7):4154-4161.

65. Cai F, Yang L, Shan S, et al. Preparation of PdCu alloy nanocatalysts for
nitrate hydrogenation and carbon monoxide oxidation. Catalysts. 2016;6(7).

66. Hungría AB, Iglesias-Juez A, Martínez-Arias A, et al. Effects of copper on


the catalytic properties of bimetallic Pd-Cu/(Ce, Zr)Ox/Al2O3 and Pd-Cu/(Ce,
Zr)Ox catalysts for CO and NO elimination. J Catal. 2002;206(2):281-294.

67. Shipitcyna A, Kinnunen NM, Hilli Y, Suvanto M, Pakkanen TA.


Characterization and Activity of Pd–Ir Catalysts in CO and C3H6 Oxidation
Under Stoichiometric Conditions. Top Catal. 2016;59(13-14):1097-1103.

68. Hilli Y, Kinnunen NM, Suvanto M, Savimäki A, Kallinen K, Pakkanen TA.


Preparation and characterization of Pd-Ni bimetallic catalysts for CO and
C3H6 oxidation under stoichiometric conditions. Appl Catal A Gen.
2015;497:85-95.

69. Vedyagin AA, Plyusnin PE, Rybinskaya AA, Shubin Y V., Mishakov I V.,
Korenev S V. Synthesis and study of Pd-Rh alloy nanoparticles and alumina-
supported low-content Pd-Rh catalysts for CO oxidation. Mater Res Bull.
2018;102(March):196-202.

70. Vedyagin AA, Stoyanovskii VO, Plyusnin PE, Shubin Y V., Slavinskaya EM,
Mishakov I V. Effect of metal ratio in alumina-supported Pd-Rh nanoalloys
on its performance in three way catalysis. J Alloys Compd. 2018;749:155-
162.

71. Vedyagin AA, Kenzhin RM, Tashlanov MY, et al. Synthesis and Study of
Bimetallic Pd-Rh System Supported on Zirconia-Doped Alumina as a

71
Component of Three-way Catalysts. Emiss Control Sci Technol.
2019;5(4):363-377.

72. Sato K, Tomonaga H, Yamamoto T, et al. A Synthetic Pseudo-Rh: NOx


Reduction Activity and Electronic Structure of Pd–Ru Solid-solution Alloy
Nanoparticles. Sci Rep. 2016;6(1):28265.

73. Haneda M, Kaneko T, Kamiuchi N, Ozawa M. Improved three-way catalytic


activity of bimetallic Ir-Rh catalysts supported on CeO2-ZrO2. Catal Sci
Technol. 2015;5(3):1792-1800.

74. Nguyen TN, Nhat TTP, Takimoto K, et al. High-Throughput Experimentation


and Catalyst Informatics for Oxidative Coupling of Methane. ACS Catal.
2020;10(2):921-932.

75. Guill N, Rico V, Garc A, Lozano D, Bueno A. THREE-WAY CATALYSTS :


PAST , PRESENT AND FUTURE . 2012;(equation 1):114-121.

76. Weik MH, Weik MH. Electrochemical Series. Comput Sci Commun Dict.
Published online 2000:489-489.

77. Patil, K. C.., Rattan, Tanu., Hegde, M. S.., Aruna, S. T.. Chemistry of
Nanocrystalline Oxide Materials: Combustion Synthesis, Properties and
Applications. Singapore: World Scientific, 2008.

78. Fernández-García M, Martínez-Arias A, Belver C, Anderson JA, Conesa JC,


Soria J. Behavior of palladium-copper catalysts for CO and NO elimination. J
Catal. 2000;190(2):387-395.

79. Shinjoh H. Rare earth metals for automotive exhaust catalysts. J Alloys
Compd. 2006;408-412(July 2004):1061-1064.

80. Asakura H, Hosokawa S, Ina T, et al. Dynamic Behavior of Rh Species in


Rh/Al2O3 Model Catalyst during Three-Way Catalytic Reaction: An
Operando X-ray Absorption Spectroscopy Study. J Am Chem Soc.
2018;140(1):176-184.

81. Hosokawa S, Nogawa S, Taniguchi M, Utani K, Kanai H, Imamura S.

72
Oxidation characteristics of Ru/CeO2 catalyst. Appl Catal A Gen. 2005;288(1-
2):67-73.

82. Dey S, Dhal GC. Applications of Rhodium and Ruthenium Catalysts for CO
Oxidation: an Overview. Polytechnica. 2020;3(1-2):26-42.

83. Gayen A, Priolkar KR, Sarode PR, et al. Ce1-xRhxO2- δ solid solution
formation in combustion-synthesized Rh/CeO2 catalyst studied by XRD,
TEM, XPS, and EXAFS. Chem Mater. 2004;16(11):2317-2328.

84. Hu Z, Allen FM, Wan CZ, et al. Performance and structure of Pt-Rh three-
way catalysts: Mechanism for Pt/Rh synergism. J Catal. 1998;174(1):13-21.

85. Corporation E. A Pt-Rh synergism in Pt/Rh three-way catalysts.


1996;(c):879-880.

86. Gayen A, Baidya T, Biswas K, Roy S, Hegde MS. Synthesis, structure and
three way catalytic activity of Ce1-xPtx/2Rhx/2O2-δ (x = 0.01 and 0.02) nano-
crystallites: Synergistic effect in bimetal ionic catalysts. Appl Catal A Gen.
2006;315:135-146.

87. Li P, Chen X, Li Y, Schwank JW. A review on oxygen storage capacity of


CeO2-based materials: Influence factors, measurement techniques, and
applications in reactions related to catalytic automotive emissions control.
Catal Today. 2019;327(May 2018):90-115.

88. Yentekakis I V., Goula G, Panagiotopoulou P, et al. Stabilization of catalyst


particles against sintering on oxide supports with high oxygen ion lability
exemplified by Ir-catalyzed decomposition of N2O. Appl Catal B Environ.
2016;192:357-364.

89. Goula G, Botzolaki G, Osatiashtiani A, et al. Oxidative thermal sintering and


redispersion of Rh nanoparticles on supports with high oxygen ion lability.
Catalysts. 2019;9(6):11-15.

90. Nagai Y, Hirabayashi T, Dohmae K, et al. Sintering inhibition mechanism of


platinum supported on ceria-based oxide and Pt-oxide-support interaction. J

73
Catal. 2006;242(1):103-109.

91. Kenji W, Saburo H, Masashi I. Development of ceria-suported ruthenium


catalysts effective for various synthetic reactions. Catal. Surv. Asia. 2011,
15(1): 1-11.

92. Yentekakis I V., Goula G, Kampouri S, et al. Ir-Catalysed Nitrous oxide


(N2O) Decomposition: Effect of Ir Particle Size and Metal–Support
Interactions. Catal Letters. 2018;148(1):341-347.

93. Yentekakis I V., Goula G, Panagiotopoulou P, et al. Dry reforming of


methane: Catalytic performance and stability of Ir catalysts supported on γ-
Al2O3, Zr0.92Y0.08O2-δ (YSZ) or Ce0.9Gd0.1O2-δ (GDC) supports. Top Catal.
2015;58(18-20):1228-1241.

94. Yang M, Shen M, Wang J, et al. Model Three-Way Catalysts. Published


online 2009:12778-12789.

95. Dong C, Zong X, Jiang W, et al. Recent Advances of Ceria‐Based Materials


in the Oxidation of Carbon Monoxide. Small Struct. 2021;2(2):2000081.

96. Ke J, Zhu W, Jiang Y, et al. Strong Local Coordination Structure Effects on


Subnanometer PtOx Clusters over CeO2 Nanowires Probed by Low-
Temperature CO Oxidation. ACS Catal. 2015;5(9):5164-5173.

97. Satsuma A, Yanagihara M, Ohyama J, Shimizu K. Oxidation of CO over


Ru/Ceria prepared by self-dispersion of Ru metal powder into nano-sized
particle. Catal Today. 2013;201(1):62-67.

98. Yao HC, Yao YFY. Ceria in automotive exhaust catalysts. I. Oxygen storage.
J Catal. 1984;86(2):254-265.

99. Fernández-García M, Iglesias-Juez A, Martínez-Arias A, et al. Role of the


state of the metal component on the light-off performance of Pd-based three-
way catalysts. J Catal. 2004;221(2):594-600.

100. Li J, Liu Z, Cullen DA, et al. Distribution and Valence State of Ru Species on
CeO2 Supports: Support Shape Effect and Its Influence on CO Oxidation.

74
ACS Catal. 2019;9(12):11088-11103.

101. Abdelsayed V, Aljarash A, El-Shall MS, Al Othman ZA, Alghamdi AH.


Microwave synthesis of bimetallic nanoalloys and CO oxidation on ceria-
supported nanoalloys. Chem Mater. 2009;21(13):2825-2834.

102. Kugai J, Moriya T, Seino S, et al. CeO2-supported Pt-Cu alloy nanoparticles


synthesized by radiolytic process for highly selective CO oxidation. Int J
Hydrogen Energy. 2012;37(6):4787-4797.

75

You might also like