Colloidal Synthesis
Colloidal Synthesis
A NNA P EKKARI
Cover:
Illustration of the synthesis of Au nanoparticles in a continuous segmented flow reactor
Typeset in LATEX
Printed by Chalmers Reproservice
Gothenburg, Sweden 2020
Colloidal synthesis of metal nanoparticles
Mechanistic studies and development of flow synthesis routes
Anna Pekkari
Department of Chemistry and Chemical Engineering
Chalmers University of Technology
Abstract
Metal nanoparticles (NPs) are central in a wide range of industrial areas including cat-
alytic and biomedical applications. Due to their interesting physical properties, the
demand of these precious materials is steadily increasing, which has created a need
for the development of effective and high-performing NPs. Because the properties are
closely correlated to NP size, shape, and crystal structure, the development of con-
trolled synthesis of metal NPs has emerged. However, the challenge of understanding
how reaction parameters influence the outcome of the synthesis currently limits the pro-
duction of precisely designed metal NPs. Additionally, low reproducibility and control
during scale-up has often restricted the production to small scale which limits the use
in industrial applications.
The studies presented in this thesis focus on the controlled synthesis of uniform metal
NPs using solution-based colloidal methods. Firstly, the multiple roles of the NP stabiliz-
ers were investigated. In the synthesis of Cu NPs, alkanethiol stabilizers only provided
temporary stabilization of the Cu NPs and decomposed under heating in inert atmo-
sphere, forming Cu2 S NPs. In another study, uniform Pd NPs stabilized with a binary
surfactant combination were synthesized without using traditional reductants. The fatty
acid stabilizer contributed to the reduction of Pd-precursors, and the reduction kinet-
ics follow a pseudo-first order kinetics. The specific stabilizers investigated influence
reduction kinetics, NP sizes and shapes. Secondly, to address scale-up challenges in
NP synthesis, the development of flow synthesis routes were explored. Uniform Pd
nanocubes (NCs) and PdPt core-shell NPs were produced in a single-phase flow reactor,
and a segmented flow reactor was developed to produce uniform Au NPs. The flow
production was scaled-up, and the uniformity of Au NPs was confirmed by inline optical
spectroscopy quality control. Catalytic evaluation of the function of PdPt core-shell NPs
in a model NO2 reduction reaction showed improved catalytic activity, selectivity and
temperature stability compared to Pd NCs.
                                                                                          i
List of Publications
                                                                                   iii
Additional publications not included in this thesis:
Guided selective deposition of nanoparticles by tuning of the surface potential
J. Eklöf, A. Stolaś, M. Herzberg, A. Pekkari, B. Tebikachew, T. Gschneidtner,
S. Lara-Avila, T. Hassenkam and K. Moth-Poulsen
Published, Europhysics Letters, 2017, 1, 18004
iv
My Contributions to the Publications
Paper I
Second author. I conducted synthesis experiments, TEM and UV-Vis analysis. XRD and
electron diffraction was performed by Christian Rohner. Part of the writing and proof
reading.
Paper II
Main author. I conducted all the synthesis experiments, sample preparation, UV-Vis and
FTIR analysis. HRTEM and electron diffraction was performed by Xin Wen. NMR mea-
surements was performed by Jessica Orrego-Hernández. I wrote the first draft and was
responsible for writing the manuscript.
Paper III
Main author. I conducted all the synthesis experiments, SEM analysis and performed
data analysis and interpreted the results with my co-authors. TEM, HRTEM and EDX
analysis was performed by Victor Sebastian. The catalytic evaluation was performed by
Zafer Say. I wrote the first draft and was responsible for writing the manuscript.
Paper IV
Main author, shared equally with Orlane Nicolardot. I conducted synthesis experiments,
UV-Vis, Zeta potential measurements and TEM analysis. HRTEM was performed by Xin
Wen. I wrote the first draft and was responsible for writing the manuscript.
                                                                                    v
vi
Contents
1 Introduction                                                                           1
  1.1 Scope of thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .    2
                                                                                        vii
        3.3.9 Catalytic evaluation . . . . . . . . . . . . . . . . . . . . . . . . .    18
Acknowledgements 43
Bibliography 45
viii
Chapter    1
Introduction
The development of metal nanoparticles (NPs) (including Cu, Au, Pd and Pt) has gained
immense interest in the recent years due to their interesting and unique physical prop-
erties, and has led to their applications in a range of areas including catalysis[1, 2], fuel
cells[3], and biomedical applications[4,5]. The steadily increasing demand of these pre-
cious materials has created a need for the development of effective and high-performing
NPs. Since the properties of the metal NPs are closely related to the size, shape and com-
position, the development of synthesis methods with a high degree of control over these
properties have emerged[6, 7]. By replacing NPs with heterogeneous morphology with
precisely designed NPs with higher performance, the material loading in the defined
application can be reduced and the cost lowered, and thus a more sustainable use of
these rare metals can be achieved[8]. Among the wide range of existing NP synthesis
methods, solution-based colloidal methods have created a range of defined shaped NPs
and benefit from its simplicity, reproducibility and high precision in controlling NP prop-
erties[9–11]. However, the lack of understanding of how different reaction parameters
influence the outcome of the synthesis is currently a challenge in the development of
precisely structured metal NPs[12]. Moreover, for controlled metal NPs to target indus-
trial applications the production has to be scaled-up to the kilogram-scale[8, 13]. The
batch reactors traditionally used for the synthesis of controlled metal NPs are often lim-
ited to small scale due to inefficient mixing and heat transfer during production scale-up
which has led to poor reproducibility between batches[8, 14]. An approach to address
these challenges is to synthesize NPs in flow reactors. Reagents are continuously infused
in micro- or millimeter sized channels, which provide fast mass transfer and heating
with high control over reaction parameters[15–18]. The synthesis can be scaled up by
flow reactor parallelization[19], increase of flow channel dimensions[13, 20], and ex-
tended operation times and NP quality can be continuously monitored by inline quality
control[21].
                                                                                           1
1.1     Scope of thesis
The primary objective of the work presented in this thesis was to develop an increased
understanding of what factors control the synthesis of uniform metal NPs using solution-
based colloidal methods. The first part of this thesis focuses on understanding synthesis
mechanisms and the influence on NP properties. In Paper I, thiol-stabilized Cu NPs
were synthesized and the decomposition route of the thiol stabilizer and the forma-
tion of Cu2 S NPs was studied. In Paper II, Pd NPs stabilized with a binary surfactant
combination were synthesized without traditional reducing agents, and the reduction
mechanisms and kinetics were studied. The second part of the thesis involves the de-
velopment of continuous flow synthesis routes to address scale-up challenges in NP
synthesis. In Paper III a flow synthesis is designed for the synthesis of Pd nanocubes
(NCs) and PdPt core-shell NPs, and process scalability explored. Furthermore, the cat-
alytic performance of the Pd and PdPt NPs was evaluated in a model NO2 reduction
reaction. Paper IV focuses on exploring the development of an automated segmented
flow synthesis to produce uniform citrate-capped Au NPs in large scale. The Au NP reac-
tion phase is separated by liquid-liquid phase separation, and product quality monitored
inline by optical spectroscopy.
2
Chapter    2
Shaped metal nanoparticles
                                                                                     3
rate, solubility)[25, 27]. Nonetheless, to understand the effect of each parameter and
how it influences the outcome of the synthesis is challenging. Several models have tried
to explain NP growth and kinetics[28], but additional knowledge is needed to fully
understand reduction mechanisms involved in NP synthesis[12, 29].
Figure 2.1: The principle of NP nucleation and growth according to LaMer theory. The
curve shows the metal precursor concentration as a function of time. [29] - Published
by The Royal Society of Chemistry.
2.1.1    Shape-control
In the synthesis of metal NPs, the shape can be directed by thermodynamics and kinetics
[25, 27, 30, 31], presented in the energy diagram in Figure 2.2. The lowest free energy
of the NP system is the product formed under thermodynamic control. A kinetic product
which has a higher energy can form at a local energy minimum. The driving force for
any kinetic product is to convert to the thermodynamic state, where a certain amount
of energy is required to cross the activation energy barrier. Depending on the size of
the energy barrier and thermal energy present in the system, the kinetic product can be
stable for different periods of time [27, 30].
     For a crystalline material, surface area is not the only factor determining the sur-
face free energy. The type of crystal structure possesses different surface free energies
depending on their atom arrangements. The metals applied in this thesis work (Cu,
Au, Pd, Pt) are face-centered cubic (fcc) metals for which three of the low-index crystal
planes are (111), (100) and (110) (Figure 2.3a). The more dangling bonds a surface
contains, the more unstable it is and the higher surface free energy it has. The number
of dangling bonds increases for (111), (100) and (110) crystal planes from 3,4 to 6,
and consequently the surface free energies follow the same pattern [27]. The Wulff
polyhedron is the thermodynamically most favorable shape, having the lowest surface
energy [31]. It consists of (111) and (100) facets and is referred to as polycrystalline.
4
Figure 2.2: Schematic illustration of kinetic and thermodynamic NP products with
relation to the total free energy. The arrow marks the activation energy required to
transform a kinetic product to a thermodynamic product. Copyright (2020) Wiley. Used
with permission from Ref. [30].
     Capping agents can influence both the thermodynamic and kinetics by selective
adsorption to crystal facets, and thus anisotropically change the surface free energies.
By lowering the activation energy for a specific facet through selective adsorption, the
capping agent can direct the growth into a specific shape [23,25]. Figure 2.3b visualizes
an example of the role of the capping agent in the growth of a metal NP into different
shapes. When a capping agent selectively passivate the (100) facets, the NP grow into a
cube, whereas a different capping agent that passivates (111) facets directs the growth
into an octahedral shape. Furthermore, by selective adsorption the capping agent can
physically block a certain crystal facet and direct the NP shape through kinetic control.
     In NP synthesis, kinetics and thermodynamics are not separate events and are often
closely connected. Among the different reaction parameters, temperature has the most
power in determining if thermodynamic or kinetic control is dominating in NP synthe-
sis. Hence, a significant rise or decrease in temperature can switch the reaction from
thermodynamic to kinetic control [25, 27].
                                                                                       5
              a)
              a)
b)
Figure 2.3: a) Models of the (111), (100), and (110) planes of a fcc-metal and the
corresponding numbers of dangling bonds per surface unit cell (NB ). b) Schematic
illustration of the role of a capping agent in directing the growth of a single-crystal seed
into NPs with different shapes. Reprinted with permission from Ref. [27]. Copyright
(2015) American Chemical Society.
Due to their interesting properties, bimetallic metal NPs have been intensively studied
in recent years which has led to the development of a plethora of different structures
[32–41].To simplify, bimetallic NPs can be divided into two main structures, core-shell
or alloy NPs. In the synthesis of bimetallic NPs, elemental composition and shape-
control is more complex than for single metal NPs. Generally, the reduction can be
performed simultaneously, or by a two-step seed-growth [32, 33]. The co-reduction
approach can create either alloys or core-shell NPs, whereas the two-step seeded-growth
normally provides core-shell structures unless significant kinetic energy is provided. I
limit this section to introducing the co-reduction approach, since it is more relevant for
the experimental part of this thesis.
     In colloidal synthesis of bimetallic NPs, apart from the parameters mentioned previ-
ously, several factors are important in determining NP shape and composition including
redox potentials, interfacial energy, and metal precursor reduction rates [32, 33, 35].
The standard reduction potential is a quantitative measure on how easy a metal precur-
6
sor can be reduced. The reduction potentials of metals can be lowered by selecting a
capping agent that strongly coordinates to the metal, which slows down the reduction
rate. The interfacial energy is determined by two factors, lattice mismatch between the
two metals and the bond between the substrate metal and the surface atoms. A large
lattice mismatch will lead to large lattice strain and high interfacial energy. Hence, a
small lattice mismatch such as for Pd and Pt (0.77 %) facilitates the combination pro-
cess [32]. Finally, the reduction rate is determined by several factors such as reduction
potential, reductant and reaction temperature. By manipulating the reduction kinetics,
the shape and the elemental composition of a NP can be altered from a core-shell struc-
ture to an alloy [32, 35, 37]. In a one-pot synthesis at slow reduction rate conditions,
the metal with the highest reduction potential will be reduced prior to the other metal,
thus creating a core-shell structure. In contrast, high reduction rates normally favors
the formation of alloy NPs [32, 35].
                                                                                          7
is minimized. Additionally, the efficient turbulent mixing inside the segments created
by the slip velocity between the reaction phase and the carrier phase minimize axial
dispersion effects normally encountered in single-phase reactors that cause wide par-
ticle size distributions[46, 47]. While fouling is reduced in segmented flow reactors,
the carrier phase generates large amount of solvent waste. A recycling strategy can be
implemented by separating the immiscible phases by polarity using liquid-liquid phase
separation[46,48]. This approach is applied in Paper IV, where the organic carrier phase
can be separated and reused in the synthesis.
     Contrary to batch reactors that are limited to small production scales due to poor
reproducibility, flow reactors can successfully be scaled up[15, 49]. Scale-up can be
achieved by operation of the reactor for extended operation times, by reactor paral-
lelization [15, 18, 19], or increased flow channel dimensions [13, 20, 44]. The possibil-
ity of full automation of the process enables an effective, safe and sustainable produc-
tion[14, 15, 18]. NP production scale-up is studied in a single-phase reactor (Paper III)
and in an automated segmented flow system (Paper IV). Furthermore, the integration
of inline quality control can provide process control by the monitoring of NP growth
to provide valuable data of reaction kinetics and synthesis outcome[16, 21, 44, 46, 49].
There exist several examples of inline quality measurements applied during metal NP
flow synthesis including dynamic light scattering [50], UV-Vis spectroscopy [13, 50–53]
and small angle x-ray scattering (SAXS) [51], that can provide structural and/or el-
emental information [21, 46]. However, inline quality control is currently limited for
fully automated segmented flow production of Au NPs and is studied in Paper IV.
(a)
(b)
8
2.3     Pd and Pt nanoparticles in catalysis
Noble metals including Pd and Pt NPs are important industrial catalysts in a range
of reactions[11]. These rare noble metals are limited resources with extremely low
abundance in the earths crust, notably at parts per billion concentrations [54]. The
increasing demand of these scarce metals has sparked the development of more high-
performing catalysts to optimize the use of these metals. The catalytic performance,
i.e. reactivity and selectivity is closely related to NP size and shape, crystal facets and
elemental composition. By applying shaped NPs with enhanced catalytic properties,
the catalytic loading can be reduced which leads to lower costs and provide a more
sustainable use of these rare metals[8]. However, in order to enhance NP catalyst per-
formance for a certain reaction, understanding of the correlation between NP properties
and catalytic activity and selectivity needs to be improved.
     Shaped Pd and Pt-based NPs have been studied as catalysts in several reactions
including reduction and oxidation reactions in fuel cells, carbon-carbon bond formation
and hydrogenation reactions[11]. Especially interesting is the combination of Pd and
Pt into bimetallic NPs, which may not only combine the individual properties but could
enhance the catalytic performance and temperature stability due to synergy effects be-
tween the metals[33–35, 55]. In Paper III catalytic evaluation of Pd NCs and PdPt
NPs with a core-shell structure is performed, where the effect of shape and elemental
composition on catalytic activity and selectivity in a model NO2 reduction reaction is
evaluated.
                                                                                         9
10
Chapter    3
Synthesis and characterization methods
The synthesis of Cu NPs was performed under N2 atmosphere using a Schlenk line (Pa-
per I). A scheme for the reaction can be found in Figure 3.1. First, the copper precursor
(CuCl2 *2H2 O) was dried into brown copper chloride (CuCl2 ) powder by heating to 60
°C for 30 minutes. Subsequently, the stabilizer dodecane thiol and solvent, dodecane,
were added to the flask, and the mixture was heated to 145 °C forming a yellow dis-
persion. The reducing agent, tert-butylamineborane complex (TBAB) was added and
the reaction was left to proceed for 5-90 min. The black Cu NP dispersion was then
left to cool in room temperature. For the synthesis of Cu2 S NPs the Cu NP dispersion
was reheated to 175 °C for 2 hours or left overnight under N2 , giving an orange Cu2 S
NP dispersion. NP dispersions were transferred to N2 -filled centrifuge tubes filled with
degassed ethanol, were centrifuged and the supernatant was discarded. The procedure
was repeated twice, followed by drying in N2 followed by dispersion in dry toluene. A
detailed description of the synthesis steps and the purification of the NPs is explained
in Paper I.
                                                                                      11
Figure 3.1: (Top) Reaction Scheme of the reaction of CuCl2 and dodecane thiol yield-
ing Cu(0)-dodecane thiol and didodecyl disulfide. (Bottom) Formation of metastable
thiolate-capped Cu NPs and their decomposition paths forming Cu2 S and Cu2 O, in
the presence of excess thiol under ambient conditions or N2 atmosphere, respectively.
Reprinted with permission from Ref. [56]. Copyright (2017) American Chemical Soci-
ety.
12
3.2     Flow synthesis of nanoparticles
This section describes the flow syntheses which were developed to synthesize Pd NCs
and PdPt core-shell NPs using a single-phase flow reactor (Paper III), and Au NPs using
a segmented flow reactor (Paper IV).
Synthesis of Pd NCs was first established by adapting a batch protocol from Niu et
al[57], to a micro-sized single-phase flow synthesis (Figure 3.2). Two streams of so-
lutions were infused in micro-sized polytetrafluoroethylene (PTFE) tubing (inner di-
ameter 800 μm) at a constant flow rate by the use of syringe pumps, interfaced in a
t-junction followed by heating in a temperature controlled water bath for a certain res-
idence time. The NP dispersion was then collected and purified by centrifugation. In
the synthesis of Pd NCs, the first stream contained an aqueous solution of Pd-precursor
(H2 PdCl4 ) and stabilizer, hexadecyltrimethylammonium bromide (CTAB), which was
reduced by a second stream consisting of an aqueous solution of reducing agent (L-
Ascorbic acid). To synthesise PdPt NPs the first stream was composed of an aqueous so-
lution of CTAB and a combination of Pd-precursor(H2 PdCl4 ) and Pt-precursor(H2 Cl6 Pt)
with varying molar ratios of the metals. A detailed description of the flow synthesis
method and the purification steps can be found in Paper II.
a) b)
Figure 3.2: The single-phase flow system used to synthesize Pd NCs and PdPt NPs. a)
Photograph of the equipment, b)Schematic setup.
                                                                                     13
3.2.2   Continuous segmented flow synthesis of Au nanoparticles
A continuous segmented hydrothermal flow synthesis was developed to produce citrate-
capped Au NPs, adapted from a modified Turkevich method in batch by Kettemann et. al
[58]. The segmented flow system (Figure 3.3) featured two peristaltic pumps, an ultra-
smooth flow chemistry syringe pump, automated back-pressure control and full automa-
tion control using the connected computer with integrated software. The reagents were
pumped in high purity grade perfluoroalkoxy (PFA) tubes. Aqueous citrate solution
was interfaced with Au-precursor solution in water in an Ethylene tetrafluoroethylene
(EFTE) T-junction. The outlet was connected to a glass-capillary connected to a sec-
ond t-junction (EFTE). Microliter sized segments were produced when carrier phase,
IsoparTM L was infused to the second T-junction. The outlet of the T-junction was con-
nected to a coiled tube microreactor (PFA), heated by hot air to a constant temperature.
The outlet of the flow reactor was connected to an automated active back-pressure regu-
lator to maintain constant pressure in the reactor and avoid gas formation. The aqueous
reaction phase was separated by a liquid-liquid separator, and the organic carrier phase
IsoparTM L was reused in the synthesis. An inline UV-Vis spectrometer flow cell was used
to monitor the quality of Au NPs in the reaction phase. The carrier phase IsoparTM L was
kindly sponsored by ExxonMobil. A detailed description of the segmented flow setup
and experimental parameters can be found in Paper IV.
a) b)
                                                                                 Au NP
                                                                               suspension
Figure 3.3: Setup of the automated segmented flow reactor for the synthesise of citrate-
capped Au NPs with liquid-liquid phase separation and inline optical spectroscopy qual-
ity control, a) Photograph of the equipment, b) Schematic setup.
14
3.3     Characterization of nanoparticles
The synthesized NPs were analyzed with respect to morphology, size, crystal struc-
ture, elemental composition and catalytic activity using the characterization methods
described in this section.
                                                                                      15
3.3.2    Scanning electron microscopy
In powder x-ray diffraction (XRD), a powder sample is irradiated with an X-ray beam
scanned at different incident angles which interacts with the electrons in the sample
and result in different scattering intensities at various angles. The resulting diffraction
pattern can be identified by comparison to a known standard or a database[63]. XRD
was applied to study the changes in the crystal structure of Cu NPs that over time formed
Cu2 S NPs (Paper I).
16
3.3.5    Microwave plasma atomic emission spectroscopy
                                                                                       17
3.3.8    Zeta potential
Zeta potential can be applied to evaluate the surface charges of NPs. When a charged
particle is dispersed in a solvent, an adsorbed electrical double layer forms at its surface.
The layer closest to the particle is of opposite charge to the particle, referred to as the
Stern layer. Due to the electrostatic field of the NPs a diffuse layer consisting of opposite
and same charge as the particle form on top of the Stern layer. Together with the
Stern layer this forms the electrical double layer. Application of an electric field results
in charges in the diffuse layer moving towards the opposite electrode. The slipping
plane is a hypothetical plane acting as an interface between the moving charges and
the dispersant around them, and the zeta potential is the potential at this interface[67].
Zeta potential was used to evaluate the colloidal stability of Au NPs (Paper IV).
18
Chapter    4
Results and discussion
                                                                                       19
                  a)                          b)
Cu (O)
Cu (I) Oxide
c) d)
Cu (O)
Cu (I) Sulfide
Figure 4.1: TEM images of a) Cu NPs, c) Cu2 S NPs. b) and d) shows the corresponding
SAED of the NPs from a) and b), respectively. Reference patterns of Cu(0), JCPDS-Nr.
4-0836 (white semicircles), Cu2 O, JCPDS-Nr. 5-0667 (red semicircles), and hexagonal
Cu2 S, space group P63/mmc, JCPDS-Nr. 46-1195, (yellow semicircles). Adapted with
permission from Ref.[56]. Copyright (2017) American Chemical Society.
20
it can be concluded that during heating of the Cu NP suspension, the alkane thiolate
stabilizer decomposed through cleavage of the C-S bond at the surface of the Cu NPs,
and only provided temporary stabilization of the Cu NP surface. This agrees well with
previous findings by Vollmer et. al [70].
a) b)
a )
Figure 4.2: (a) Temporal evolution of UV-Vis absorption spectra of freshly synthesized
Cu NP suspension. Photograph of the Cu NP suspension after b) 0 min, and c) after 275
min. Adapted with permission from Ref. [56]. Copyright (2017) American Chemical
Society.
     Apart from decomposing on the surface to alter the chemical composition of NPs,
the stabilizer can have other roles in the synthesis of metal NPs. In Paper II I investigated
the influence of the stabilizers in the development of a synthesis of Pd NPs stabilized
with a binary surfactant mixture of NaOL and CTAC. The synthesis was performed in
the absence of traditional reducing agents, hence the Pd-precursors were reduced spon-
taneously in the reaction mixture at 100 °C. Based on earlier findings [71], our initial
hypothesis was that the electron dense alkyl double bond in the stabilizer NaOL was the
main contributor to the reduction of Pd-precursors. After 4 hours of reaction uniform
Pd NPs formed (Figure 4.3) with an average size of 29.7 nm ± 5.7 % (Inset in Figure
4.3a). The Pd NPs are polycrystalline, seen by multiple diffraction rings (inset (Figure
4.3b), and can be visualized in the HRTEM image of a single multiple-twinned Pd NP
(Figure 4.3c,d).
     A range of reaction parameters were investigated to study the influence on the Pd
NP properties, thoroughly described in Paper II. In this thesis I focus the discussion on
the effects of altering the stabilizers on Pd NP properties. To evaluate the hypothe-
sis that the double bond is responsible for the reduction of Pd-precursors, NaOL was
replaced with the structurally similar saturated fatty acid, sodium stearate (NaST). Sur-
prisingly, this stabilizer combination could also produce Pd NPs, which are smaller than
                                                                                          21
            a)                                        b)
                             20   25   30   35   40
                                            )
) )
Figure 4.3: Structural characterization of Pd NPs stabilized with NaOL and CTAC. a),b)
TEM images of monodisperse Pd NPs at relatively low magnification, where inset in a)
shows histogram of particle size distribution with an average size of 29.7 nm ± 5.7 %,
inset in b) shows the SAED pattern of Pd NPs. c) HRTEM image of an individual Pd
NP shows a multiple-twinned structure, where twin boundaries are marked with red
arrows, d) HRTEM image of the selected area of the Pd NP marked by a red square in
c). The inserted image in d) shows a corresponding FFT pattern.
with NaOL, with an average size of 13 nm ± 19 % (Figure 4.4a). The reaction was
slower than with NaOL, observable by a later color change of the reaction solution. The
reduction speed was further investigated with UV-Vis spectroscopy, as explored in the
following section. Reduction occurred despite the NaST lacking double bonds which in-
dicates that the reduction mechanisms are more complex than the previously proposed
hypothesis.
     When CTAC was replaced with CTAB, the equivalent ammonium salt with bromide
as counter ion in the surfactant mixture, Pd NPs with different shapes formed (Figure
4.4b). This includes bars, cubes and “arrow” shaped Pd NPs. After 4 hours reaction,
excess particle seeds are present and Pd NPs are smaller than than those stabilized with
22
         a)                                                       b)
                            5   10   15   20   25       30   35
                                                    )
Figure 4.4: a) TEM image of Pd NPs stabilized with NaST and CTAC, where Inset in a)
shows a correspoding histogram of size distributions with an average size of 13 nm ±
19 %. b) TEM image of Pd NPs stabilized with NaOL and CTAB.
NaOL and CTAC (Figure 4.4b). Despite using the same reductant (NaOL), a slower
reduction was evidenced by a slower color change of the reaction solution. The differ-
ence in shapes obtained when CTAC was replaced with CTAB could be due to bromide
ions (Br- ) present in CTAB, well-known to selectively adsorb onto (100) crystal facets
and have been applied extensively in the controlled synthesis of cubic [57, 72] and rod-
shaped Pd NPs [73]. In the synthesis of CTAB-stabilized Au nanorods, Meena et. al [74]
showed that Br- adsorption is not the only contributor to the selective surface passiva-
tion of (100) facets, but is a driving force for CTAB micelle adsorption and stabilization
of the Au nanorod surface. CTAB was shown to form dense surfactant micellar layers
on gold surfaces. In contrast, more isotropic shaped particles were obtained with CTAC,
due to less facet selectivity and the low presence of Cl- and micellar structures protect-
ing the Au surface. These findings may explain the difference in shapes observed when
CTAC was replaced with CTAB in the synthesis of Pd NPs. Furthermore, the slower re-
duction of Pd-precursors observed in the presence of CTAB could partly be explained by
the strong complexation between bromide ions in CTAB and Pd-precursors (PdBr4 2- ).
Additionally, the dense and thick surfactant layer of CTAB may lower the accessibility of
NaOL to reduce Pd-precursors which could contribute to a slower reduction rate.
     In order to produce colloidally stable Pd NPs it was necessary to use a binary sta-
bilizer combination with the anionic surfactant NaOL/NaST and the cationic surfactant
CTAC. When synthesis was performed in the absence of CTAC, reduction occurred and
Pd NPs formed but the stabilizers (NaOL/NaST) provided poor colloidal stability and
extensive visual particle aggregation in solution could be observed. On the other hand,
synthesis with only CTAC does not result in any reduction of Pd-precursors. These
                                                                                       23
findings lead to the conclusion that CTAC was necessary to provide sufficient colloidal
stability of Pd NPs, but did not contribute to the reduction of Pd-precursors. To study
the interactions between the stabilizers NaOL and CTAC with Pd NPs, qualitative anal-
ysis was performed using FTIR and NMR, which showed that the stabilizers provided
colloidal stabilization by adsorption onto the NP surface. Details of the experiments
and qualitative analysis can be found in Paper II. It is clear that NP stabilizers can have
multiple roles in the NP synthesis and their influence of the outcomes can be complex.
24
     a)                        1                                                         b)                        1
                                        a                                    5 min                                          a                                     5 min
                                                                             30 min                                                                               30 min
                              0.8                                            60 min                               0.8                                             60 min
          Absorbance (a.u.)
                                                                                              Absorbance (a.u.)
                                                                             120 min                                                                              120 min
                                                                             180 min                                                                              180 min
                              0.6                                            240 min                              0.6                                             240 min
                                                                             300 min                                                                              300 min
0.4 0.4
0.2 0.2
                               0                                                                                   0
                               250           275           300         325         350                             250               275        300         325          350
                                                   Wavelength (nm)                                                                     Wavelength (nm)
       )                 120
                                                                                           )
                                                                                                                   5
100 4.5
                                                                                                                   4
                              80
     2-
                                                                                         ln [PdCl 2- ]
          4
                                                                                                  4
                                                                                                                  3.5
      % PdCl
                                                                        NaST                                                                                      NaST
                              60
                                                                        NaOL                                                                                      NaOL
                                                                                                                   3
                              40
                                                                                                                  2.5
                              20
                                                                                                                   2
                               0                                                                                  1.5
                                    0       50     100    150    200    250     300                                     0       50     100     150    200    250      300
                                                         Time (min)                                                                          Time (min)
Figure 4.5: Analysis of the reduction kinetics in the synthesis of Pd NPs. UV-Vis spectra
of PdCl4 2- in the reaction solution after 5-300 min for Pd NPs stabilized with a) CTAC
and NaOL, b) CTAC and NaST. c) A plot showing the percentage of PdCl4 2- remaining
in the reaction solutions for Pd NPs synthesized with NaOL and CTAC (blue), and with
NaST and CTAC (red) measured from the absorbance peak at 280 nm as a function of
time. d) Plots of ln [PdCl4 2- ] over time, which shows the pseudo-first order reaction
kinetics involved in the synthesis of Pd NPs stabilized with NaOL and CTAC, and with
NaST and CTAC, respectively.
retain a constant concentration through the reaction. Since the exact mechanisms of the
reduction are still not fully clear, further studies into this aspect should be performed to
provide a better understanding. To evaluate the influence of the stabilizer on reduction
kinetics, the effect of varying the concentration of reductant, i.e. NaOL/NaST, on the
reduction kinetics could be performed. Moreover, application of the Finke-Watzky (F-
W) kinetic model [75] to our data was performed since it could provide a model for
relatively slow reduction reactions [12]. The F-W model describes the reduction through
several steps including nucleation, homogeneous aggregation and autocatalytic surface
growth. Application of the model to our data shows poor agreement, where our system
presents a slow continuous reduction as opposed to the F-W model which includes a
slow reduction followed by fast autocatalytic growth.
                                                                                                                                                                               25
     The type of metal precursor and the metal precursor and ligand coordination is
known to substantially influence reduction kinetics [12, 77]. Therefore, in the synthesis
of Pd NPs it would be interesting to evaluate the effect on reduction kinetics of changing
type of Pd-precursor, i.e. the type of metal precursor complex. It is clear from this study
that CTAC plays an important role in providing shape and colloidal stability of the Pd
NPs. The effect on shape and reduction kinetics when CTAC was replaced with CTAB,
caused by the stronger complexation between Pd-precursors and Br– ions serves as a
first example and it would be interesting to study other cationic surfactants and their
effect on reduction kinetics and NP properties.
     Another possible direction to extend investigations of reduction kinetics could be
the study of other metals such as Au. Syntheses of various shaped Au NPs have been
performed with traditional reducing agents and stabilization of NaOL and CTAB [71]
and oleic acid and CTAB [79]. During the synthesis a noticeable color change of the Au-
precursor solution have been claimed to indicate that NaOL may act as partial reducing
agent for Au(III) precursors. Yet, further studies would be needed to understand the
reduction kinetics and mechanisms for Au NPs in the absence of these reducing agents.
     From the quantitative evaluation of the reduction kinetics involved in the synthesis
of Pd NPs, it can be concluded that the alkyl double bond in NaOL is not necessary
to reduce Pd-precursors. Nonetheless, it may influence the reduction rate since reduc-
tion with NaOL is faster and more effective than Pd NP synthesis with the saturated
fatty acid NaST. However, the exact mechanisms governing the reduction needs to be
further studied to elucidate what electron donating groups that contribute to reducing
Pd-precursors. Elemental analysis of byproducts formed during NP synthesis using e.g.
NMR and Mass spectroscopy may improve understanding of the mechanisms during
reduction [80], which could elucidate the mechanisms involved in the reduction of Pd-
precursors (Paper II) but also to understand the decomposition of alkane thiolates on
the surface of Cu NPs (Paper I). The evaluation of byproduct chemistry during metal NP
synthesis may be challenging since byproducts can be present in small quantities, and
many reagents are involved in the synthesis (metal precursors, stabilizers, reductants
etc.) which creates a multitude of possible compounds and combinations to evaluate. In
this study, we applied NMR and Liquid chromatography and mass spectroscopy (LCMS)
to evaluate the byproducts formed during the Pd NP synthesis. Nonetheless, these
analyses led to inconclusive results and further analyses are needed to draw significant
conclusions on the reduction mechanisms involved in the synthesis. Despite these re-
sults, byproduct evaluation could be a valuable tool in evaluation of NP properties and
contribute in the development of metal NP synthesis procedures.
26
4.2     Flow synthesis of colloidal nanoparticles
The second focus of this thesis is the development of flow synthesis methods, due to
their potential to provide excellent control over NP properties. First, I developed a
synthesis of Pd NCs and PdPt NPs using a single-phase flow reactor (Paper III), and
an automated segmented flow synthesis was developed for the production of Au NPs
(Paper IV). A range of reaction parameters were evaluated and the effects of synthesis
temperature discussed. Furthermore, the scale-up of NP flow production was studied,
with the future aim of targeting real applications. Finally, evaluation of catalytic per-
formance in a model catalytic reaction of Pd NCs and PdPt NPs was performed (Paper
III).
         a)                                              b)
                                                                                                        2 nm
          0   5   10   15   20   25       30   35   40   0   5   10   15   20   25       30   35   40
                                      )                                              )
                                                                                                               27
and Pd NCs synthesized in flow 14.4 nm ± 11 % (inset in Figure 4.6b). Comparison
of NP uniformity shows no significant difference between the two methods. Nonethe-
less, when the particle yield was examined using MP-AES, observing the amount of Pd
precursor that has been reduced into metallic Pd NCs, a significant improvement in re-
action yield can be seen for flow synthesized Pd NCs (94 % in flow, 63 % in batch). This
improvement in synthesis yield could be explained by the more efficient heat and mass
transfer in the flow reactor[43].
     Motivated by the enhanced catalytic properties of bimetallic NPs, I modified the
synthesis to incorporate a second metal to produce bimetallic PdPt NPs. By varying the
relative molar ratios of Pd:Pt in the precursor stream, the morphology of the formed
PdPt NPs can be controlled, which gradually changes from rough cubic to a spherical
dendritic shape with increasing Pt concentration (Figure 4.7).
a) b) )
Figure 4.7: TEM-images of flow-synthesized PdPt NPs. The molar ratio of Pd:Pt in the
particles are a) 6:1, b) 3:1, c) 1:1. Scale bars are 50 nm, in inset images scale bars are
10 nm. Reprinted with permission from Ref.[81]. Copyright (2019) American Chemical
Society.
     Characterization using STEM-HAADF of the PdPt NPs with the highest Pt concen-
tration clearly visualizes the dendritic surface topography (Figure 4.8a,c,d). The ele-
mental distribution, analyzed by a line scan on the NP using STEM-EDX (Figure 4.8b)
reveals a core-shell structure with Pd dominated in the core and a Pt-rich surface. The
formation of bimetallic NPs is directed by the kinetics and reduction potential of the
two metals. Despite its lower reduction potential, Pd is reduced first to form the parti-
cle core. The slower reduction rate of Pt precursors could be explained by the stronger
complexation between Pt precursors and CTAB [82], which then lead to the formation
of PdPt core-shell NPs.
28
              a)                          b)
) )
Figure 4.8: a) HAADF-STEM image of flow synthesized PdPt NPs with the highest Pt
ratio. b) EDX line scan of a single PdPt NP which shows the relative distribution of Pd
and Pt, a core-shell structure. c) and d) shows high-resolution HAADF-STEM images
of PdPt NPs. Reprinted with permission from Ref.[81]. Copyright (2019) American
Chemical Society.
                                                                                    29
            a)                                                b)
                              7   8   9   10   11   12
                                               )
              )                                                )
                  d(111)=0.225 nm
                                                         )
                                                         a
                                                         ba
                                                         b
           d(200)=0.225 nm d(200)=0.225 nm
                                                                   a   )
Figure 4.9: a), b) TEM images of uniform citrate-capped Au NPs at relatively low
magnification. The histogram of size distribution in inset in a) shows that the Au NPs
have a size of 9.5 nm ± 0.8 nm (8.4 %). The inset in b) shows the SAED pattern of
Au NPs. c) HRTEM image of a single Au NP shows a polycrystalline structure, where
boundaries between different crystalline domains are marked with red arrows. d) UV-
Vis absorbance spectrum of Au NPs with an absorbance maximum at 518 nm.
     The narrow size distribution of the Au NPs synthesized in the segmented flow reac-
tor could be attributed to the internal back-mixing related to the slip velocity between
the segments of reaction phase and carrier phase that cause turbulent mixing. This
chaotic mixing eliminates axial dispersion effects normally experienced in single-phase
reactor that can cause wide particle size distributions [47, 52, 83, 84]. Furthermore, our
segmented flow reactor was integrated with automated back-pressure control which
enables the potential to conduct the synthesis at higher pressure and temperature, so
called hydrothermal conditions. A higher reaction temperature may shorten of reaction
time due to a faster nucleation and growth. The segmented flow reactor enables pre-
cise controlled synthesis under hydrothermal conditions, which is not easily achieved in
conventional batch reactors.
30
4.2.3   Flow synthesis optimization - effect of temperature
In the development of the flow synthesis of Pd NCs and PdPt NPs (Paper III), and Au NPs
(Paper IV), a systematic evaluation of different reaction parameters were performed to
evaluate the influence on NP properties. Details of all the tested parameters can be
found in Paper III and Paper IV. Since the temperature has a big influence on reaction
kinetics in the synthesis of metal NPs, directing the outcome of the synthesis [12, 27],
the effect on nucleation and growth of NPs by variation of the synthesis temperature
was evaluated for both flow systems.
     When the temperature was lower than the optimal conditions (60 °C for Pd NCs
and PdPt NPs, 100 °C for Au NPs), larger Pd NPs (Figure 4.10a), PdPt NPs (Figure
4.10c) and Au NPs (Figure 4.10e) form. When the synthesis temperature is lowered,
the nucleation rate is slower which results in fewer particle seeds that subsequently
grow into larger NPs. Ftouni et. al observed an increase in Au NP size when the flow
synthesis temperature was reduced from 100 °C to 60 °C [85]. At the lower synthesis
temperature, PdPt NPs (Figure 4.10c) formed highly dendritic and porous NPs. Pd NPs
synthesized at the same temperature consisted of different shapes including nanorods,
triangles, multiple twinned NPs and NCs (Figure 4.10a). The slower nucleation rate
experienced at lower synthesis temperature may have increased the prevalence of Pd
NP seeds with stacking faults or multiple twinned structures that subsequently direct
the growth into different structures such as multiple twinned particles and nanorods.
However, further structural characterization is needed to confirm the exact structures
of the Pd NPs. Both thermodynamics and reaction kinetics, largely influenced by tem-
perature, play important roles in determining the structure of NP seeds [27]. Since
the shape of the NPs strongly depends on the initial structure of the seeds (presence of
stacking faults, twinned structure or single-crystal), control of temperature can direct
these properties [12]. Wang et. al [86] correlated the reduction rate with the struc-
ture of the seeds and found that relatively slow reduction of Pd-precursors generated
seeds with stacking faults and/or twin planes, whereas higher reduction rates yielded
single-crystal Pd NPs.
     When the synthesis temperature was increased (130 °C for Pd NCs and PdPt NPs,
140 °C for Au NPs) various effects on sizes and shapes can be seen. Pd NCs (Figure
4.10b) have a wide distribution of sizes and shapes including rods, cubes, and twinned
NPs. Plenty of the Pd NPs are smaller than the Pd NCs synthesized at the optimal con-
ditions (96 °C). Similar size distribution effects are observed for the PdPt NPs (Figure
4.10d). The wide size and shape distributions observed for Pd NPs and PdPt NPs synthe-
sized at 130 °C could be explained by the difference in nucleation and growth rates. The
nucleation and growth are competing processes that occur simultaneously [12], and at
                                                                                     31
this temperature, the optimal parameters of nucleation and growth differ, which may
cause shape and size variations. In contrast, Au NPs synthesized at 140 °C experienced
extensive aggregation and formed large structures (Figure 4.10f).
a) b)
) )
) )
Figure 4.10: TEM images of a) Pd NCs and c) PdPt NPs synthesized in single-phase flow
at 60 °C. b) Pd NCs and d) PdPt NPs synthesized in flow at 130 °C. Au NPs synthesized
in segmented flow at e) 100 °C and f) 140 °C. Scale bars are 50 nm.
     When the synthesis temperature was raised, the acceleration of nucleation of growth
created many small NP seeds. Without proper colloidal stabilization, the citrate-capped
Au NPs aggregated in an effort to minimize the high surface energy of the individual
small NPs. This correlates well with previous findings [87], where nuclei aggregation
32
and polydisperse Au NPs can be seen when synthesis temperature was increased. The
aggregation may be minimized by a shorter reaction time, and the colloidal stabiliza-
tion improved by increasing the concentration of stabilizer [87,88]. The fact that citrate
also acts as reductant and pH-mediator in the synthesis makes is difficult to predict the
effects of increased citrate concentration on Au NP morphology [88]. Furthermore, the
insufficient colloidal stabilization of Au NPs at 140 °C may have originated from partial
degradation of citrate. It has been shown that degradation of citrate occurs at higher
temperature [89, 90], but its thermal stability may be affected when applied as capping
agent [90, 91].
                                                                                        33
      a)                        b)                           )
) ) )
Figure 4.11: Structural characterization by SEM of Pd NCs (a,b and c) and PdPt NPs
(d,e and f) synthesized in large-scale in the single-phase microsized flow reactor. NP
samples were deposited on Si-substrates. a) Pd NCs and d) PdPt NPs without treatment.
b) Pd NCs and e) PdPt NPs after treatment in 48-136 °C temperature interval. c) Pd
NCs and f) PdPt NPs after treatment in 48-220°C temperature interval. Prior to heat
treatments, pre-treatment was performed at 158 °C. Heat treatments were conducted
in 2200 ppm NO2 and 2.2 % H2 in Ar(g). Scale bars are 100 nm.
      After treatment in reaction conditions (48-136 °C), Pd NCs have slightly rounded
edges (Figure 4.11b), and upon further temperature increase (220 °C), particles under-
went sintering and formed large agglomerates(Figure 4.11c). The PdPt core-shell NPs
on the other hand showed good thermal stability and retained the shapes at reaction
conditions (Figure 4.11e), and the shape remained preserved at elevated temperatures
(Figure 4.11f). The improved thermal stability experienced by the PdPt core-shell NPs
may be attributed to the dendritic crystal structure, and the combination of Pd with Pt,
which has a higher temperature stability [92]. The severe aggregation of the Pd NCs ex-
perienced at elevated temperatures (>220 °C), where the NPs lost their size and shape
can be explained by sintering. Sintering can occur by two different mechanisms; Ost-
wald ripening, where small particles dissolve and redeposit onto large particles, or par-
ticle migration and coalescence[92, 93]. It is not fully understood by which mechanism
sintering occurred for the Pd NPs, but a correlation between high NP concentration and
sintering could be observed, where shorter inter-particle distance seemed to enhance
the sintering process. It should also be pointed out that for this model catalytic reaction
the NP catalysts were simply deposited on Si-substrates. Heterogeneous catalysis is nor-
mally performed with metal NP catalysts immobilized onto a porous support material.
Several strategies have been employed to improve the sintering stability of metal NP
catalysts to minimize diffusion, including enhancing metal-support interactions [94].
34
It would hence be interesting to re-evaluate the thermal stability of the NPs on meso-
porous support materials. Moreover, it has been shown that highly monodisperse NPs
experience less sintering due to Ostwald ripening effects [95]. Therefore, it would be
very interesting to evaluate and compare the temperature stability of more shape- and
size-uniform Pd NCs and PdPt NPs.
     Fouling, which was experienced during scale-up in the flow reactors in Paper III,
is a commonly encountered problem in single-phase flow reactors. Any interfaces can
act as nucleation zones during NP growth, and since the reactants in single-phase flow
reactors are in constant contact with the reactor wall there is high risk for heteroge-
neous nucleation and accelerated growth to occur on the reactor wall, which can lead
to reactor fouling [45–47]. A possible solution is to use segmented flow reactors where
the contact between the reactor wall and the reaction phase can be minimized by the
use of an organic carrier phase that effectively wets the reactor walls. In the segmented
flow synthesis of citrate-capped Au NPs no fouling was observed in the flow reactor,
even when the synthesis was scaled-up for an extended time period (120 min). This
indicates that the segmented flow effectively minimizes contact between the reaction
phase and the reactor wall. It would be interesting to develop a fouling-free segmented
flow synthesis of the Pd NCs and PdPt NPs that would enable an effective scale-up of
the synthesis and that likely would produce more uniform NPs.
     The segmented flow created large volumes of organic solvent waste (carrier phase).
In Paper IV, a recycling strategy was implemented to separate Au NPs from the organic
phase by using LLS, where the two phases efficiently can be separated based on po-
larity and the carrier phase recovered and reused in the synthesis. Au NPs after LLS
show a zeta potential of -60 mV and maintain colloidal stabilization after separation
of the reaction phase. Slight agglomeration could be observed when the Au NPs were
placed onto TEM-grids, which could be an effect from drying the NP suspension or may
originate from after the LLS.
     An important part in continuous flow production scale-up is the monitoring of
growth and NP product consistency using inline quality control [21, 46]. During the
scale-up of the hydrothermal segmented flow synthesis of Au NPs, the product con-
sistency was monitored inline using an UV-Vis spectroscopy flow cell during 54 min
of synthesis (Figure 4.12). As can be seen in the contour plot (left Figure 4.12), the
absorbance spectra of the Au NPs stay fairly consistent, as well as the absorbance max-
imum for each spectrum (marked with red diamonds) that presented an average peak
maximum of 520.3 nm with a standard deviation of 2.8 nm (0.5 %). The representative
inline UV-Vis spectrum of Au NPs (right Figure 4.12) correlates well with the ex-situ
UV-Vis spectrum (Figure 4.9d). Some points in the contour plot show slightly lower
                                                                                      35
                                                  b   ba
)
                                                                                       )
a
                                                                                       a
                                  )                                  b   ba
Figure 4.12: Left panel shows a contour plot which visualizes the inline UV-Vis ab-
sorption spectra of citrate-capped Au NPs recorded between 60-3120 seconds during
continuous production. The absorbance maxima of each spectrum are marked with red
diamonds. Right panel shows a representative inline UV-Vis spectrum of Au NPs.
absorbances and red-shift of the absorbance maxima, which may be due to slight ag-
glomeration of Au NPs. It should be noted that the spectra were recorded during short
integration time (1 ms) and the flow cell may be sensitive to external disturbances such
as vibrations and air bubbles that may contribute to deviations in absorbance. Since
the plasmon peaks are not very defined for Pd NCs and PdPt NPs, quality monitoring by
inline UV-Vis spectroscopy is more challenging for these flow synthesized NPs.
     When comparing the two systems developed to synthesize NPs in flow in larger
scale there exist several important differences. The type of pumps used is important
when it comes to scaling up the flow synthesis. In Paper III, a syringe pump was used
to infuse the reagents. This type of pump is consequently limited by the volume of the
syringes and can only produce a certain volume of NP suspension. In Paper IV, auto-
mated peristaltic pumps were used and the production volume was not as limited since
bottled reagents were used for the precursor solutions. The advantage of the fouling-
free segmented flow synthesis of citrate-capped Au NPs is that it was fully automated
and integrated with inline quality control. The full automation provides lower mainte-
nance and safer operation and this system demonstrates the capability of NP synthesis
scale-up for future real applications like the catalysis of NO2 reduction, as explored for
a test reaction in the following section.
36
4.2.5    Catalytic evaluation of Pd and PdPt nanoparticles
In Paper III, a coauthor performed a temperature programmed NO2 reduction reaction
(48-136 °C with a gas feed of 2200 ppm NO2 and 2.2 % H2 in argon) using a flow
“pocket reactor” to study the catalytic activity of Pd NCs and PdPt core-shell NPs. The
corresponding NO2 conversion efficiencies (Figure 4.13a) show that PdPt core-shell NPs
exhibited slightly higher catalytic activity at lower temperatures. The lower activity of
Pd NCs could partly be related to the cubic structure, which is composed of (100) crystal
facets, known to exhibit lower activity compared to high-index facet counterparts[96–
99] that are highly abundant in the dendritic PdPt NPs. Additionally, synergistic effects
between Pd and Pt may have contributed to the improved catalytic activity of PdPt NPs.
Product selectivity is another aspect studied. During NO2 reduction by H2 three prod-
ucts form; NO, N2 O, and N2 , where N2 is the most desired one due to its non-toxicity
compared to the others, which are greenhouse gases and/or toxic. There is a clear
difference in selectivity between the two types of NPs (Figure 4.13b), where the PdPt
core-shell NPs show higher selectivity towards N2 , as well as lower selectivity for NO
formation compared to the Pd NCs who show nearly 98 % NO selectivity. It should be
noted that due to restructuring behavior that was observed for the NPs at elevated tem-
peratures, the operation temperature was limited to 150 °C. An improved temperature
stability would enable the temperature window to be extended, likely resulting in lower
NO selectivity. Moreover, since the synthesis scale-up produced heterogeneous NPs it
would be interesting for future studies to evaluate the catalytic activity and selectiv-
ity of uniform shaped NPs, which would deepen the understanding of the relationship
between catalytic performance and shape.
                                                                                      37
a)            100
                             PdPt NPs
                                                                  b)
                             Pd NPs
                   80        Blank
                                                                                      NO     N 2O       N2
  NO2 Conversion
                   60
                                                                  PdPt NPs            68                     20   10
                   40
                                                                   Pd NPs                     98
                                                                                                              1.3 0.7
                   20
                                                                             0   20     40         60        80     100
                                                                                  Product selectivity (%)
                    0
                        0   20   40   60   80   100   120   140
                                 Temperature (° C)
Figure 4.13: Catalytic evaluation of Pd NCs and PdPt NPs. a) Relative NO2 conversion
efficiencies of Pd NCs, PdPt NPs and blank sample. b) Calculated relative percent selec-
tivity of NO, N2 O and N2 in 48 – 136 °C temperature interval for Pd NCs and PdPt NPs.
Adapted with permission from Ref. [81]. Copyright (2019) American Chemical Society.
38
Chapter    5
Conclusions and Reflections
The aim of the work presented in this thesis was to improve the understanding of the
factors that control the synthesis of shape and size-controlled metal NPs using solution-
based colloidal methods. From the studies presented in the first part of this thesis, it can
be concluded that the stabilizers strongly influence the NP properties which include sta-
bility, shape and reduction kinetics. I found that the selected stabilizers can have more
roles than providing colloidal stabilization. Thiolate stabilizers decompose at the sur-
face of Cu NPs and only provide temporary stabilization of the NP surface, even under
inert atmosphere. In the synthesis of Pd NPs in the absence of traditional reductants,
it can be concluded that the fatty acid stabilizers NaOL and NaST contribute to the re-
duction of Pd-precursors. Quantitative evaluation of the reduction kinetics show that
these stabilizers exhibit pseudo first-order reduction kinetics in the synthesis of Pd NPs,
where the unsaturated fatty acid NaOL provides faster and more effective reduction.
The different stabilizers’ influence on reduction kinetics direct the sizes and shapes of
the Pd NPs.
      The second part presented in this thesis focused on the development of flow syn-
thesis routes for the design of shaped and uniform metal NPs, and provided synthesis
scalability and catalytic evaluation. First, I developed a single-phase flow reactor that
produced shape- and size-controlled Pd NCs and PdPt NPs. Precise morphological con-
trol of core-shell PdPt NPs was achieved by varying the elemental composition. Sec-
ondly, a fouling-free segmented flow synthesis was developed to produce uniform Au
NPs. In both syntheses, among the evaluated reaction parameters temperature strongly
influenced NP properties. Synthesis scale-up in the single-phase reactor led to extensive
fouling and heterogeneous NPs. Exposure of these NPs to elevated temperatures showed
that the bimetallic PdPt NPs exhibit better stability compared to the monometallic Pd
NCs. Furthermore, from a proof-of-concept catalytic evaluation in the NO2 reduction
reaction it can be concluded that PdPt NPs also exhibit higher catalytic activity and
                                                                                        39
improved product selectivity. The scale-up of the segmented flow system provided uni-
form Au NPs and the product consistency was confirmed by inline optical spectroscopy.
The fully automated and segmented flow system thus presents a promising system for
scalable synthesis of precisely controlled NPs.
40
a combination of automated flow synthesis and machine learning to change reaction
conditions quickly, and experimentalists to choose the most likely factors to test.
     Furthermore, NP synthesis should take inspiration from flow synthesis of organic
compounds, which in recent years have seen developments in automated synthesis us-
ing AI in planning and screening of reaction conditions to predict optimal reaction con-
ditions [103]. This can also be applied to NP synthesis, where machine learning can
serve as a powerful tool for synthesis optimization by the use of automated feedback
loops. By bringing knowledge from existing NP synthesis data, reaction parameter space
optimization can be performed to estimate and optimize NP outcomes [49, 104, 105].
Fully automated segmented flow systems with in-situ monitoring coupled to machine
learning and optimization algorithms may enable effective reaction monitoring and NP
synthesis optimization. Altogether, I believe these future technological developments
could facilitate production scale-up for real industrial applications and may shorten the
time for new high-performing controlled NPs to reach industrial applications.
                                                                                      41
42
Acknowledgements
These PhD studies have been a journey for me, both professionally and personally, and
has involved deep valleys and important life lessons learned over these four years. I
want to thank all the people who have been involved in this process to support me to
finally reach this big goal.
My second supervisor Prof. Kasper Moth-Poulsen for giving me the opportunity to start
this PhD project. For all your scientific inputs and guidance, and I am glad to have
experienced how our relationship has developed over the years.
My main supervisor Prof. Hanna Härelind for your strong support and supervision and
guidance. You helped me believe in myself, all the way to the PhD. Thanks for all the
amazing times together during research projects, and outside work at the division par-
ties.
My examiner Prof. Martin Andersson. You sparked my interest in research when you
took me in as a master thesis student in 2014, which eventually led me to start this
journey. Thank you for your scientific support and guidance during this time.
Prof. Victor Sebastian for hosting me during my research stay in University of Zaragoza
in Spain. I am grateful for your supervision in developing flow chemistry methods and
the personal developments I experienced during this research stay.
The Flow team: Jessica and Robson. Thanks for great times together in the lab, with
lots of laughter, interesting scientific discussions and achievements.
My friends and colleagues at the division of applied chemistry. Thank you for amazing
times together with fikas, parties, trips, and good times in the lab. No one mentioned,
no one forgotten.
My dear friends, who mean so much to me. You have been by my side all this time. You
know who you are. Thank you all for your support.
                                                                                    43
Ett speciellt tack till min familj. Ni har stöttat mig och alltid funnits där för mig i de bra
och de svåra stunderna. Jag är väldigt tacksam för det. Jag älskar er.
Älskade Carl-Robert, min klippa. De snart två år som du har varit en del av mitt liv har
varit fantastiska. Du har funnits där och har stöttat mig varje dag i att kämpa på och
fortsätta. Jag ser nu fram emot att fortsätta vår resa tillsammans mot en spännande
framtid. Jag älskar dig.
44
Bibliography
                                                                                      45
[19] C. T. Riche, E. J. Roberts, M. Gupta, R. L. Brutchey, and N. Malmstadt, Nature
     Communications, 2016, 7, 1–7.
[20] L. Zhang, G. Niu, N. Lu, J. Wang, L. Tong, L. Wang, M. J. Kim, and Y. Xia, Nano
     Letters, 2014, 14(11), 6626–6631.
[21] L.-J. Pan, J.-W. Tu, H.-T. Ma, Y.-J. Yang, Z.-Q. Tian, D.-W. Pang, and Z.-L. Zhang,
     Lab on a Chip, 2017, 18, 41–56.
[22] T. S. Rodrigues, M. Zhao, T. H. Yang, K. D. Gilroy, A. G. da Silva, P. H. Camargo,
     and Y. Xia, Chemistry - A European Journal, 2018, 24(64), 16944–16963.
[23] A. Heuer-Jungemann, N. Feliu, I. Bakaimi, M. Hamaly, A. Alkilany, I. Chakraborty,
     A. Masood, M. F. Casula, A. Kostopoulou, E. Oh, K. Susumu, M. H. Stewart, I. L.
     Medintz, E. Stratakis, W. J. Parak, and A. G. Kanaras, Chemical Reviews, 2019,
     119(8), 4819–4880.
[24] V. K. Lamer and R. H. Dinegar, Journal of the American Chemical Society, 1950,
     72(11), 4847–4854.
[25] Z. Wu, S. Yang, and W. Wu, Nanoscale, 2016, 8(3), 1237–1259.
[26] M. Muzzio, J. Li, Z. Yin, I. M. Delahunty, J. Xie, and S. Sun, Nanoscale, 2019,
     11(41), 18946–18967.
[27] Y. Xia, X. Xia, and H. C. Peng, Journal of the American Chemical Society, 2015,
     137(25), 7947–7966.
[28] N. T. K. Thanh, N. Maclean, and S. Mahiddine, Chemical Reviews, 2014, 114(15),
     7610–7630.
[29] J. Polte, CrystEngComm, 2015, 17(5).
[30] T. H. Yang, Y. Shi, A. Janssen, and Y. Xia, Angewandte Chemie - International
     Edition, 2020, 59(36), 15378–15401.
[31] L. D. Marks and L. Peng, Journal of Physics Condensed Matter, 2016, 28(5),
     53001.
[32] J. Gu, Y. W. Zhang, and F. Tao, Chemical Society Reviews, 2012, 41(24), 8050–
     8065.
[33] A. Zaleska-Medynska, M. Marchelek, M. Diak, and E. Grabowska, Advances in
     Colloid and Interface Science, 2016, 229, 80–107.
[34] M. B. Gawande, A. Goswami, T. Asefa, H. Guo, A. V. Biradar, D.-l. Peng, R. Zboril,
     and R. S. Varma, Chemical Society Reviews, 2015, 44(21), 7540–7590.
[35] K. D. Gilroy, A. Ruditskiy, H. C. Peng, D. Qin, and Y. Xia, Chemical Reviews, 2016,
     116(18), 10414–10472.
[36] H. L. Liu, F. Nosheen, and X. Wang, Chemical Society Reviews, 2015, 44(10),
     3056–3078.
[37] M. Zhou, H. Wang, M. Vara, Z. D. Hood, M. Luo, T. H. Yang, S. Bao, M. Chi,
46
       P. Xiao, Y. Zhang, and Y. Xia, Journal of the American Chemical Society, 2016,
       138(37), 12263–12270.
[38]   J. Zhang, L. Wan, L. Liu, Y. Deng, C. Zhong, and W. Hu, Nanoscale, 2016, 8(7),
       3962–3972.
[39]   X. Huang, Y. Li, Y. Li, H. Zhou, X. Duan, and Y. Huang, Nano Letters, 2012, 12(8),
       4265–4270.
[40]   X. Xia, Y. Wang, A. Ruditskiy, and Y. Xia, Advanced Materials, 2013, 25(44),
       6313–6332.
[41]   J. S. Santana, K. M. Koczkur, and S. E. Skrabalak, Langmuir, 2017, 33(24), 6054–
       6061.
[42]   E. Shahbazali, V. Hessel, T. Noël, and Q. Wang, Nanotechnology Reviews, 2014,
       3(1), 65–86.
[43]   S. Marre and K. F. Jensen, Chemical Society reviews, 2010, 39(3), 1183–1202.
[44]   S. E. Lohse, Physical Sciences Reviews, 2018, 3(11), 1–29.
[45]   J. M. Köhler, S. Li, and A. Knauer, Chemical Engineering and Technology, 2013,
       36(6), 887–899.
[46]   H. Huang, H. du Toit, L. Panariello, L. Mazzei, and A. Gavriilidis, Physical Sciences
       Reviews, 2018, 3, 157–220.
[47]   Y. He, K. J. Kim, and C. H. Chang, Nanomaterials, 2020, 10(7), 1–21.
[48]   Y. Shen, N. Weeranoppanant, L. Xie, Y. Chen, M. R. Lusardi, J. Imbrogno, M. G.
       Bawendi, and K. F. Jensen, Nanoscale, 2017, 9(23), 7703–7707.
[49]   V. Sebastian, S. A. Khan, and A. A. Kulkarni, Journal of Flow Chemistry, 2017,
       7(3–4), 96–105.
[50]   K. Pacławski, B. Streszewski, W. Jaworski, M. Luty-Błocho, and K. Fitzner, Col-
       loids and Surfaces A: Physicochemical and Engineering Aspects, 2012, 413, 208–
       215.
[51]   L. C. McKenzie, P. M. Haben, S. D. Kevan, and J. E. Hutchison, Journal of Physical
       Chemistry C, 2010, 114(50), 22055–22063.
[52]   V. Sebastian Cabeza, S. Kuhn, A. a. Kulkarni, and K. F. Jensen, Langmuir, 2012,
       28(17), 7007–7013.
[53]   G. Liu, X. Yang, Y. Li, Z. Yang, W. Hong, and J. F. Liu, Advances in Materials
       Science and Engineering, 2015, 2015.
[54]   J. W. Park, Z. Hu, S. Gao, I. H. Campbell, and H. Gong, Geochimica et Cosmochim-
       ica Acta, 2012, 93, 63–76.
[55]   H. Zhang, M. Jin, and Y. Xia, Chemical Society Reviews, 2012, 41(24), 8035.
[56]   C. Rohner, A. Pekkari, H. Härelind, and K. Moth-Poulsen, Langmuir, 2017,
       33(46).
                                                                                         47
[57] W. Niu, Z. Y. Li, L. Shi, X. Liu, H. Li, S. Han, J. Chen, and G. Xu, Crystal Growth
     and Design, 2008, 8(12), 4440–4444.
[58] F. Kettemann, A. Birnbaum, S. Witte, M. Wuithschick, N. Pinna, R. Kraehnert,
     K. Rademann, and J. Polte, Chemistry of Materials, 2016, 28(11), 4072–4081.
[59] D. B. Williams and C. B. Carter, Transmission Electron Microscopy, Springer, 2009.
[60] J. I. Goldstein, D. E. Newbury, J. R. Michael, N. W. M. Ritchie, J. H. J. Scott, and
     D. C. Joy, Scanning Electron Microscopy and X-Ray Microanalysis, Springer, 4 ed.,
     2018.
[61] C. J. le Roux and R. J. Kriek, Hydrometallurgy, 2017, 169, 447–455.
[62] U. Kreibig and M. Vollmer, Optical properties of metal clusters, Springer, Berlin-
     New York, 1995.
[63] C. Suryanarayana and M. G. Norton, in X-Ray Diffraction: A Practical Approach,
     Springer US, Boston, MA, 1998; pp. 21–62.
[64] G. Moore in Introduction to Inductively Coupled Plasma Atomic Emission Spectrom-
     etry, ed. G. Moore, Analytical Spectroscopy Library; Elsevier, Amsterdam, 1989;
     pp. 65 – 90.
[65] P. R. Griffiths and J. A. de Haseth, Fourier Transform Infrared Spectrometry, John
     Wiley & Sons, 2 ed., 2007.
[66] H. Günther, NMR Spectroscopy : Basic Principles, Concepts and Applications in
     Chemistry., John Wiley & Sons, Incorporated, 2013.
[67] S. Bhattacharjee, Journal of Controlled Release, 2016, 235, 337–351.
[68] Y. Bu, J. W. Niemantsverdriet, and H. O. Fredriksson, ACS Catalysis, 2016, 6(5),
     2867–2876.
[69] M. D. Susman, A. Vaskevich, and I. Rubinstein, Journal of Physical Chemistry C,
     2016, 120(29), 16140–16152.
[70] S. Vollmer, G. Witte, and C. Wöll, Langmuir, 2001, 17(24), 7560–7565.
[71] X. Ye, Y. Gao, J. Chen, D. C. Reifsnyder, C. Zheng, and C. B. Murray, Nano Letters,
     2013, 13(5), 2163–2171.
[72] H. C. Peng, S. Xie, J. Park, X. Xia, and Y. Xia, Journal of the American Chemical
     Society, 2013, 135(10), 3780–3783.
[73] Y. H. Chen, H. H. Hung, and M. H. Huang, Journal of the American Chemical
     Society, 2009, 131(25), 9114–9121.
[74] S. K. Meena, S. Celiksoy, P. Schäfer, A. Henkel, C. Sönnichsen, and M. Sulpizi,
     Physical Chemistry Chemical Physics, 2016, 18(19), 13246–13254.
[75] M. A. Watzky and R. G. Finke, Journal of the American Chemical Society, 1997,
     119(43), 10382–10400.
[76] S. Özkar and R. G. Finke, Langmuir, 2016, 32(15), 3699–3716.
48
[77] S. Mozaffari, W. Li, C. Thompson, S. Ivanov, S. Seifert, B. Lee, L. Kovarik, and
     A. M. Karim, Nanoscale, 2017, 9(36), 13772–13785.
[78] M. Luty-Błocho, K. Pacławski, M. Wojnicki, and K. Fitzner, Inorganica Chimica
     Acta, 2013, 395, 189–196.
[79] K. Raghunathan, J. Antony, S. Munir, J.-P. Andreassen, and S. Bandyopadhyay,
     Nanoscale Advances, 2020, 2, 1980–1992.
[80] F. N. Stappen, K. Enemark-Rasmussen, G. P. Junor, M. H. Clausen, J. Zhang, and
     C. Engelbrekt, Journal of Physical Chemistry C, 2019, 123(41), 25402–25411.
[81] A. Pekkari, Z. Say, A. Susarrey-Arce, C. Langhammer, H. Härelind, V. Sebastian,
     and K. Moth-Poulsen, ACS Applied Materials and Interfaces, 2019, 11(39), 36196–
     36204.
[82] Y. Kim, Y. W. Lee, M. Kim, and S. W. Han, Chemistry - A European Journal, 2014,
     20(26), 7901–7905.
[83] A. Abou-Hassan, O. Sandre, and V. Cabuil, Angewandte Chemie - International
     Edition, 2010, 49(36), 6268–6286.
[84] A. A. Kulkarni and V. Sebastian Cabeza, Langmuir, 2017, 33(50), 14315–14324.
[85] J. Ftouni, M. Penhoat, A. Addad, E. Payen, C. Rolando, and J. S. Girardon,
     Nanoscale, 2012, 4(15), 4450–4454.
[86] Y. Wang, H. C. Peng, J. Liu, C. Z. Huang, and Y. Xia, Nano Letters, 2015, 15(2),
     1445–1450.
[87] W. Ding, P. Zhang, Y. Li, H. Xia, D. Wang, and X. Tao, ChemPhysChem, 2015,
     16(2), 447–454.
[88] M. Wuithschick, A. Birnbaum, S. Witte, M. Sztucki, U. Vainio, N. Pinna, K. Rade-
     mann, F. Emmerling, R. Kraehnert, and J. Polte, ACS Nano, 2015, 9(7), 7052–
     7071.
[89] A. Marcilla, A. Gómez-Siurana, M. Beltrán, I. Martínez-Castellanos, I. Blasco,
     and D. Berenguer, Journal of the Science of Food and Agriculture, 2018, 98(15),
     5916–5931.
[90] Z. Bai, L. Yang, J. Zhang, L. Li, C. Hu, J. Lv, and Y. Guo, Journal of Power Sources,
     2010, 195(9), 2653–2658.
[91] M. Bajaj, N. Wangoo, D. V. Jain, and R. K. Sharma, Scientific Reports, 2020, 10(1),
     1–7.
[92] A. Cao, R. Lu, and G. Veser, Physical Chemistry Chemical Physics, 2010, 12(41),
     13499–13510.
[93] T. W. Hansen, A. T. Delariva, S. R. Challa, and A. K. Datye, Accounts of Chemical
     Research, 2013, 46(8), 1720–1730.
[94] L. Wang, L. Wang, X. Meng, and F. S. Xiao, Advanced Materials, 2019, 31(50),
                                                                                       49
        1–18.
 [95]   V. P. Zhdanov, E. M. Larsson, and C. Langhammer, Chemical Physics Letters, 2012,
        533, 65–69.
 [96]   D. L. Feldheim, Science, 2007, 316(5825), 399–700.
 [97]   N. Tian, Z. Y. Zhou, N. F. Yu, L. Y. Wang, and S. G. Sun, Journal of the American
        Chemical Society, 2010, 132(22), 7580–7581.
 [98]   B. Y. Xia, H. B. Wu, X. Wang, and X. W. Lou, Angewandte Chemie - International
        Edition, 2013, 52(47), 12337–12340.
 [99]   E. Ozensoy, C. Hess, and D. W. Goodman, Journal of the American Chemical Soci-
        ety, 2002, 124(29), 8524–8525.
[100]   B. J. Reizman and K. F. Jensen, Accounts of Chemical Research, 2016, 49(9),
        1786–1796.
[101]   J. Nette, P. D. Howes, and A. J. DeMello, Advanced Materials Technologies, 2020,
        5(7).
[102]   R. M. Maceiczyk, I. G. Lignos, and A. J. Demello, Current Opinion in Chemical
        Engineering, 2015, 8, 29–35.
[103]   C. W. Coley, D. A. Thomas, J. A. M. Lummiss, J. N. Jaworski, C. P. Breen,
        V. Schultz, T. Hart, J. S. Fishman, L. Rogers, H. Gao, R. W. Hicklin, P. P. Ple-
        hiers, J. Byington, J. S. Piotti, W. H. Green, A. J. Hart, T. F. Jamison, and K. F.
        Jensen, Science, 2019, 365(6453), 1–9.
[104]   O. Voznyy, L. Levina, J. Z. Fan, M. Askerka, A. Jain, M. J. Choi, O. Ouellette,
        P. Todorović, L. K. Sagar, and E. H. Sargent, ACS Nano, 2019, 13(10), 11122–
        11128.
[105]   A. M. Hiszpanski, B. Gallagher, K. Chellappan, P. Li, S. Liu, H. Kim, J. Han,
        B. Kailkhura, D. J. Buttler, and T. Y. J. Han, Journal of Chemical Information and
        Modeling, 2020, 60(6), 2876–2887.
50