You are on page 1of 43

View Article Online

View Journal

Nanoscale
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: S. de Marchi
Lourenço, S. Núñez-Sánchez, G. Bodelón, J. Perez-Juste and I. Pastoriza-Santos, Nanoscale, 2020, DOI:
10.1039/D0NR06270G.
Volume 10
Number 4
28 January 2018
This is an Accepted Manuscript, which has been through the
Pages 1549-2172
Royal Society of Chemistry peer review process and has been
accepted for publication.
Nanoscale Accepted Manuscripts are published online shortly after acceptance,
rsc.li/nanoscale

before technical editing, formatting and proof reading. Using this free
service, authors can make their results available to the community, in
citable form, before we publish the edited article. We will replace this
Accepted Manuscript with the edited and formatted Advance Article as
soon as it is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes to the
text and/or graphics, which may alter content. The journal’s standard
ISSN 2040-3372 Terms & Conditions and the Ethical guidelines still apply. In no event
PAPER
Shuping Xu, Chongyang Liang et al.
Organelle-targeting surface-enhanced Raman scattering
shall the Royal Society of Chemistry be held responsible for any errors
(SERS) nanosensors for subcellular pH sensing

or omissions in this Accepted Manuscript or any consequences arising


from the use of any information it contains.

rsc.li/nanoscale
Page 1 of 42 Nanoscale

Pd nanoparticles as plasmonic material: synthesis, optical properties and


View Article Online
applications DOI: 10.1039/D0NR06270G

Sarah De Marchi,a,b,† Sara Nuñez-Sánchez,a,b,† Gustavo Bodelón,a,b Jorge Pérez-


Juste a,b and Isabel Pastoriza-Santosa,b

a CINBIO, Universidade de Vigo, Departamento de Química Física, Campus


Universitario As Lagoas, Marcosende 36310 Vigo, Spain
b Galicia Sur Health Research Institute (IIS Galicia Sur). SERGAS-UVIGO, Vigo,

Nanoscale Accepted Manuscript


Spain
† These authors have contributed equally to this work
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

e-mail corresponding autor: pastoriza@uvigo.es

Abstract
This review provides an overview of current progress into Pd nanoparticles
supporting localized surface plasmon resonance and their applications. We begin
by analyzing briefly the optical properties of Pd putting particular focus on outlining
the origin of its size- and shape-dependent LSPR, high refractive index sensitivity,
high absorption contribution. The differences in the optical behavior with Au and Ag,
the primary plasmonic materials, are highlighted. The main strategies to synthesize
Pd nanoparticles, pure or hybrid, with well-defined optical properties are then
reviewed. In this section, we include only those works that carry out the study of the
optical properties of the nanoparticles. The applications of plasmonic Pd
nanoparticles are also discussed in detail. This review is concluded with a section
devoted to the future perspectives highlighting the most relevant challenges to be
addressed to take Pd nanoparticles from the laboratory to real applications.
Nanoscale Page 2 of 42

View Article Online


DOI: 10.1039/D0NR06270G
1. Introduction

In 1804, the English chemist and physicist W. H. Wollaston published a paper


entitled “On a New Metal, Found in Crude Platina”.1 Although the new metal was
rhodium, he also reported for the first time the separation of a small amount of
palladium from crude platinum. Named by Wollaston in honor of the asteroid Pallas,
it was not until the 1990s that Pd became popular due to its use as catalytic
converter in the automotive industry. Nowadays Pd is still a technologically relevant
metal in the vehicle markets as the main component of the three-way catalysts.2, 3

Nanoscale Accepted Manuscript


Besides, it is also an outstanding catalyst in a wide variety of reactions involved in
industrial processes or devices.4-6 It is particularly important in organic synthesis, for
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

instance, Pd is the best catalyst for C–C bond formation reactions which are
essential in the pharmaceutical and agricultural industries.6,7, 8 Additionally, its
extraordinary ability to reversibly absorb hydrogen in large quantities even at
ambient conditions makes Pd play a key role in the upcoming hydrogen economy.9,
10 In fact, Pd is applied in various hydrogen technologies including hydrogen
purification, storage, and detection as well as fuel cells. The great interest for Pd as
well as its enhanced properties exhibited at the nanoscale has driven the
development of strategies to obtain Pd nanoparticles with shape and size
control.3,11, 12
Less effort has been spent on the in-depth analysis of Pd as plasmonic metal,
as well as in the design and development of plasmonic Pd nanoparticles with
relatively narrow and tunable optical responses in the UV-visible-NIR region.
Despite its lower optical efficiency in the UV region due to interband transition
processes, Pd exhibits several advantages over the primarily studied plasmonic
metals such as Au and Ag, which make it more suitable for certain applications.
However, it is still a challenge to fabricate Pd nanoparticles with well-defined and
tunable plasmonic properties. It is a rather critical aspect in terms of Pd applicability
in the near future since it would allow optimizing the amount of this precious metal
required for a certain Pd-based technology. Pd abundance in Earth´s crust is as low
as 0.015 part per million. Besides the recent interest in Pd based materials, has
contributed to an excessive price increase in the last 5 years (from $614 (Aug 15,
2015) to $2181 (August 30, 2020) per ounce, Kitco Gold Index).13
Page 3 of 42 Nanoscale

Herein we revisit the most relevant aspects of plasmonic Pd nanoparticles. We


View Article Online
begin with the analysis of Pd as plasmonic material doing a comparativeDOI:
study with
10.1039/D0NR06270G

the two most commonly used plasmonic metals, Au and Ag. We also investigate the
origin of its high refractive index susceptibility and its high absorption cross section,
among other aspects. We then examine the main approaches to synthesize Pd
nanoparticles with a well-defined and tunable optical response in the UV-visible-NIR
region. This section is divided into two subsections devoted to pure Pd nanoparticles
and Pd-based nanoparticles exhibiting Pd-like optical properties. Next, we present
an overview of the most important applications of plasmonic Pd nanoparticles

Nanoscale Accepted Manuscript


including photocatalysis, H2 sensing, photothermal therapy, or bioimaging. Finally,
we conclude with a short prospect on the potential future developments and
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

investigations in this field.

2. Plasmonic Properties of Pd.

Localized Surface Plasmon Resonances (LSPRs) are optical modes


supported by metallic nanoparticles which confine light at the nanoscale. As optical
modes, their resonance conditions and mode quality factors will be defined by the
geometry and size of the nanoparticle and the relations between the optical
properties of the metal and the surrounding dielectric medium. Therefore, for a given
nanoparticle geometry and a dielectric environment, the permittivity of the metal will
be the magnitude to study to determine the best plasmonic metal for a specific
application.14
Quality factors are magnitudes that can be used to classify material platforms
as a function of their performance for a defined application. In the case of plasmonic
materials for LSPR based devices, the plasmonic material quality factors (QLSPR)
are defined by the relation between the real and imaginary part of the permittivity of
the metal. Importantly this relation strongly depends on the geometry of the particles
as follows:
QLSPR=-ε´(ω)/ε´´(ω) for spherical nanoparticles (1);
QLSPR=ε´(ω)2/ε´´(ω) for ellipsoidal nanoparticles (2);
where ω is the frequency and ε´(ω) and ε´´(ω) are the real and imaginary part of the
permittivity, respectively.14 Therefore, metals with large negative values of the real
part of the permittivity are associated with materials confining efficiently light at the
nanoscale and large QLSPR (see Eq. (1) and Eq. (2)). While, on the other hand, large
Nanoscale Page 4 of 42

values of the imaginary part of the permittivity are mainly associated with LSPRs
View Article Online
damping of plasmonic nanoparticle modes. DOI: 10.1039/D0NR06270G

Under this scenario, if we compare the optical properties of Au, Ag, and Pd
(Fig. 1), Au and Ag are material platforms with potentially larger QLSPR than Pd. This
is related to larger negative values of the real part of the permittivity and lower values
of the imaginary part of permittivity achieved by Au and Ag in comparison with Pd
(Eq. (1) and Eq. (2)). However, Pd nanostructures have surprisingly shown some
extraordinary plasmonic properties which are better than those from Au or Ag
nanostructures (such as refractive index sensitivity or absorption capacity). This is

Nanoscale Accepted Manuscript


because not only the magnitudes of the real and imaginary parts of permittivity are
relevant for plasmonic nanoparticle applications. In fact, the dependence of
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

dielectric function with the wavelength/frequency plays also a key role. For example,
how fast the optical properties vary with the wavelength can determine the LSPR
peak shifts produced by changes in the local surrounding medium, establishing the
refractive index sensitivity of the nanoparticles. Therefore, the relevance of the
permittivity dependence with the wavelength opens a new perspective to transition
metals as Pd as plasmonic counterparts to noble metals as Au or Ag. Moreover, Pd
as third plasmonic material shows better properties in comparison with Au and Ag,
for instance higher stability at elevated temperatures for thermoplasmonic
applications.15,16,17
Page 5 of 42 Nanoscale

View Article Online


DOI: 10.1039/D0NR06270G

Nanoscale Accepted Manuscript


Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Fig. 1. Real part, ’m, (straight line) and imaginary part, ’’m, (dash line) of the
permittivity of bulk Pd18 (red), Au (green) and Ag (blue).19 Areas with coloured
shadow indicate the wavelength range where the value of ’m is between -3.5 and -
23 for each metal: Pd (red), Au (green), and Ag (blue). The slope of the curves at
𝜀 = - 23 can be estimated by the tangent of the indicated angle. This angle is only
indicated for Au and Pd curves.

2.1. Analysis of the real part of the permittivity of Pd (’m)

In 1908 Mie showed that a dielectric sphere supports electromagnetic resonances


at discrete resonant wavelengths producing field enhancement in the vicinity of the
spheres. The same theory can be applied to spherical metal nanoparticles to
determine the scattering and absorption coefficients. For homogenous spheres with
sizes much smaller than the wavelength of the incident light and under a quasi-static
approximation, Fröhlich established a simple condition for this LSPR frequency (0)
in metals as:
Nanoscale Page 6 of 42

´ ( )
= (3) View Article Online
´ ( ) ´ DOI: 10.1039/D0NR06270G

where ε´m is the real part of dielectric function of the metal, εd is the dielectric function
of the surrounding medium and l is the angular momentum which determines the
order of the resonance mode.20 Therefore, in the case of small spherical
nanoparticles dispersed in a dielectric medium as water for a dipolar resonance
mode (l=1), the LSPR condition is reached when ε´m (ω0) = -2 εd ≈ 3.5 (Fröhlich
condition, Eq. (3)). It indicates that for small spherical nanoparticles the LSPR
depends directly on the relation between the dielectric functions of the metal and

Nanoscale Accepted Manuscript


the surrounding dielectric medium (ε´m and εd). Moreover, the Fröhlich condition
shows that the LSPR frequency is independent of particle size but strongly
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

dependent on how the dielectric function of the metal varies with the wavelength.
As can be observed in Fig. 1, where it is represented the real and imaginary part of
Pd, Au, and Ag, the dipolar LSPR wavelength will match this Fröhlich condition of
Eq. (3) at 290 nm for Pd, at 383 nm for Ag and 514 nm for Au. Therefore, in the case
of small Pd nanoparticles under a quasi-static approximation, the LSPR excitation
condition is expected to be located at shorter wavelengths than for Au and Ag since
its real part of the permittivity reach more negative values at shorter wavelengths
than Au and Ag.
LSPRs in the UV range for isotropic small Pd nanoparticles is demonstrated in
the initial works on the synthesis of Pd nanoparticles. Xiong et al. show than
isotropic Pd nanoparticles as cuboctahedra21 and nanocubes22 with sizes smaller
than 50-60 nm showed their LSRR band below 300 nm.22,21 Fig. 2a shows the
simulated extinction cross section of spherical Pd nanoparticles surrounded by
water (n = 1.333) for diameters ranging from 5 to 200 nm. The simulations were
carried out using the Boundary Element Method (BEM).23, 24 Interestingly, spherical
Pd nanoparticles smaller than 30 nm exhibit the LSPR, dipolar in nature, around
200 nm. Moreover, nanoparticles with diameters between 40 and 80 nm also show
a dipolar LSPR band centered form 200 to 400 nm (UV-visible region). Above 80
nm, the optical response broadens since apart from dipolar modes, other higher
order modes (quadrupolar, octopolar, etc.) contributed to the extinction spectra.
Interestingly, Pd nanoparticles larger than 120 nm show an extinction band
extended from UV to IR region.
Page 7 of 42 Nanoscale

View Article Online


DOI: 10.1039/D0NR06270G

Nanoscale Accepted Manuscript


Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Fig. 2. Optical properties of spherical Pd, Ag and Au nanoparticles. (a)


Calculated extinction cross sections of spherical Pd nanoparticles with diameters
between 5 nm and 200 nm. The nanoparticle diameter is indicated on top of each
curve. (b-d) Calculated extinction, absorption, and scattering cross sections of
spherical nanoparticles of 30 nm (solid lines) and 120 nm (dashed line) of Pd (b),
Ag (c), and Au (d)). All calculations were carried out considering water as the
surrounding medium.

In the case of more complex geometries than spheres, for example,


anisotropic ellipsoidal nanoparticles (size << λ), the scattering (Qsca) and the
absorption (Qabs) cross sections can be estimated as a function of the real (´) and
imaginary (´´) part of the nanoparticle polarizability () as follows:

𝑄 = |𝛼| (4)

and 𝑄 = 𝑘´´ (5)

where k is the modulus of the wavector (k = 2 / λ).25 The polarizability is defined as


the ratio between the dipole induced by the incident light in the nanoparticle and the
incident displacement field.26 The LSPR peak occurs when the polarizability
Nanoscale Page 8 of 42

experiences a resonant enhancement. In the case of ellipsoidal nanoparticles, the


View Article Online
polarizability for each principal axis (i) is established by: DOI: 10.1039/D0NR06270G

( )
𝛼 =𝑉 ( ( ) )
(6)

where (i) is the polarizability in i axis direction, V the volume of the ellipsoid, and Li
the shape factor.20 The polarizability experiences a resonant enhancement when
the denominator of Eq. (6) is close to zero, that is, d + Li (m () - d) = 0. This
condition establishes a LSPR peak for each ellipsoid principal axis, i, at frequency
i when ´m (i)= -((1- Li)/ Li) d for a nanoparticle embedded in a non-lossy dielectric

Nanoscale Accepted Manuscript


medium (d = ´d). Therefore, it will be determined by the dependence of the dielectric
function of metal with the wavelength, d and Li. For example, in the case of a prolate
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

spheroid, L1 will be <1/3, satisfying LSPR condition at more negative values of the
permittivity than -2d.27 If the prolate spheroid is highly anysotropic with shape
factors L1=L2=0.07 and L3=0.83 (as cigar-like shape), the resonance condition for
the longitudinal axis will be determined by the frequency i,0 where ε´m (ωi,0) ≈ -23.
Taking into account the real part of the permittivity of Ag, Au, and Pd (Fig. 1), the
ε´m (ωi,0) ≈ -23 condition will be matched at wavelengths 697 nm, 784 nm, and 832
nm for of Ag, Au and Pd, respectively. Therefore, in the case of anisotropic Pd
nanoparticles, as this cigar-like shape, under a quasi-static approximation, the
LSPR excitation condition for the larger axis is expected to be located at longer
wavelengths than for Au and Ag nanoparticles of same size and geometry. For
instance, Langhammer et al.28 when studied highly anisotropic Pd and Ag
nanodisks, of 20 nm in height and 530 nm in diameter, found that the LSPR of Ag
nanodisks was located at around 1600 nm while for Pd nanodisks it is at higher
wavelengths (around 1800 nm). In another work, it was reported that ultrathin Pd
plates with a thickness restricted to less than 10 atomic layers and edge length
between 20 and 160 nm shown well-defined and tunable LSPRs from 826 nm to
1068 nm in the IR.16
Some LSPR based sensors are based on changes in LSPR due to the modifications
of nanoparticle size and/or shape occurring during the analytical assay.29, 30 The
presence of target analytes can either promote/inhibit the particles etching or
promote/mediate the particle growth leading to the transformation in the size, shape,
and composition of particles.31, 32 Therefore, nanoparticles with high LSPR
sensitivity (large shift range) to shape and size modifications are more interesting
for this type of colorimetric sensors. As seen in Fig. 1, the variation of ε´m with the
Page 9 of 42 Nanoscale

wavelength for Pd shows a lower slope than that for Au and Ag. This means that
View Article Online
LSPR in Pd is more sensitive than Au and Ag to changes in the particleDOI:
shape and
10.1039/D0NR06270G

size. For example, in the previous case of highly anisotropic cigar-like ellipsoids
embedded in water with L1=L2=0.07 and L3=0.83, the longitudinal LSPR condition
is fulfilled for wavelengths where ε´m (ωi,0) ≈ -23, that is at 832 nm for Pd, 697 nm
for Ag, and 784 nm for Au. Interestingly, during the transformation of these cigar-
like ellipsoids (L1=L2=0.07 and L3=0.83) into small nanospheres (L1=L2=L3=1/3), the
LSPR condition will be tuned from ε´m (ωi,0) ≈ -23 (832 nm for Pd, 697 nm for Ag and
784 nm for Au) to ε´m (ω0) = - 3.5 (290 nm for Pd, 383 nm for Ag and 514 nm for

Nanoscale Accepted Manuscript


Au). Therefore, as indicated in Fig. 1, the LSPR will be varied in a wavelength
interval of 542 nm for Pd (red area), 314 nm for Ag (blue area), and 270 nm for Au
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

(green area).
The large range tuneability of Pd LSPR from UV to IR has been theoretically
established and experimentally demonstrated by highly anisotropic Pd nanorods33
and Pd nanodisks,28, 34, 35 respectively. Additionally, Langhammer and coworkers
compared the optical response of Pd and Ag nanodisks of different sizes.36 In the
case of small nanodisks of 20 nm height and 38 nm diameter, those made of Pd
exhibited their LSPR centered at 200 nm while Ag nanosdisks showed their LSPR
at 500 nm. In the case of larger nanodisks (20 nm height and 530 nm diameter), Pd
nanodisks showed their LSPR centered 1800 nm while Ag nanosdisks showed their
LSPR at 1600 nm. Thus, the LSPR of Pd nanodisks can be tuned from 200 nm to
1800 nm (1600 nm range) while LSPR of Ag can only be tuned from 500 nm to 1600
nm (1100 nm range). Therefore, LSPRs are more sensitive to nanoparticle
morphological transformations in the case of Pd than Ag.

2.2. Analysis of the imaginary part of permittivity (ε´´m): Absorption


contribution and heat generation

As we mentioned previously, metals with large negative values of the real part
of the permittivity are associated with materials confining efficiently light at the
nanoscale. However, large values of the imaginary part of the permittivity are mainly
associated with LSPR damping of plasmonic nanoparticle modes. Fig. 1 shows the
imaginary part of the permittivity of Au, Ag, and Pd as a function of the wavelength.
The strong interband activity in Au and Ag beyond their interband threshold energies
(3.9 eV for Ag and 2.5 eV for Au)36 causes that the largest values of ´´m for Ag are
achieved for wavelengths below 317 nm (3.9 eV) while for Au the largest values of
Nanoscale Page 10 of 42

´´m are achieved below 527 nm (2.5 eV).36, 37 However, in the case of Pd, the d-
View Article Online
electrons overlap the s- and p-electrons in the whole range of interest leading to a
DOI: 10.1039/D0NR06270G

uniform imaginary part of the permittivity achieving larger values than the ones of
Au and Ag in almost the whole wavelength range, from UV to NIR.38
This large difference in the ´´m of Pd in comparison with that for Au and Ag
will play an important role in the ratios of the contribution of absorption and
scattering cross sections to the total extinction of the plasmonic nanoparticles.
Cross sections quantify the amount of incident light scattered, absorbed, or
extinguished per nanoparticle. When nanoparticles are small with sizes below λ/10,

Nanoscale Accepted Manuscript


being λ the incident light wavelength, Rayleigh scattering dominated and a quasi-
static model approximation for homogenous spheres can be used to determine Qsca
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

and Qabs (see Eq (4) and Eq (5)). As the nanoparticle gets larger, between λ/10 and
λ, the scattering is directional and called Mie scattering. Fig. 2 shows the scattering,
absorption, and extinction cross sections of spherical nanoparticles, 30 and 200 nm
in diameter, made of Pd, Au, and Ag and embedded in water. For 30 nm particles,
the extinction is dominated by the absorption contribution independently of the
metal. This larger contribution of Qabs to Qext for small nanoparticles is related to
ε´´m. As the nanoparticle size increases (data not shown) the scattering contribution
to the extinction becomes more relevant and at sizes larger than 50 nm for Ag and
100 nm for Au the scattering dominates nanoparticle extinction and the absorption
contribution is located at shorter wavelengths. Nevertheless, in the case of Pd, it
occurs at larger sizes above 130 nm and the contribution of Qabs to Qext is still
important even at large particle sizes. It is expected for a material that reaches large
values of the imaginary part in all the whole wavelength range. Fig. 2b-d shows the
differences in absorption and scattering contribution to the extinction for 120 nm
nanoparticles made of Pd, Ag, and Au. For this size, the scattering contribution
increases in the following order: Pd < Au < Ag. Moreover, the absorption contribution
for Pd extends from UV to NIR while for Ag and Au it is only relevant in the visible
region. Similar results were found by Langhammer et al. investigating the optical
response of Pd, Ag, and Au nanodisks.36,28
There are applications, which require plasmonic nanostructures made of
metals with a large imaginary part of the permittivity and large absorption cross
sections as, for example, thermoplasmonics. The heat power density q(r) generated
inside a plasmonic nanoparticle due to a LSPR excitation is given by:
Page 11 of 42 Nanoscale

𝑞(𝒓) = 𝜀 ´´ (𝜔)𝜀 |𝑬(𝒓)| (7)


View Article Online
DOI: 10.1039/D0NR06270G
where r indicates the position vector and E(r) is the electric field achieved on the
metal nanoparticle. Therefore, q generated is a local effect that depends on the
position and is directly proportional to the field intensity distribution (|E(r)|2) inside
the metal and the imaginary part of the permittivity of the metal (ε´´m) (see Eq. (7)).
Moreover, the power absorbed (and delivered) 𝑄 by a nanoparticle is directly
proportional to the absorption cross section and can be simply estimated by the
expression:
Q = Qabs I, where I is incident irradiance (power per unit surface).39, 40 Under this

Nanoscale Accepted Manuscript


scenario, nanoparticles of metals with larger absorption cross sections and with the
ability to confine light at the nanoscale as Pd, are fantastic candidates for
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

thermoplasmonic applications.16, 41 Another important issue is that in comparison


with Au and Ag, Pd exhibit higher melting point ( 1552 C vs 1063 C and 961
C for Au and Ag, respectively), which provides Pd high thermal stability. This is a
critical issue in photothermal applications. Indeed, Au and Ag have been combined
with Pd to enhance their thermal stability.42
Extinction spectra of Pd nanoparticles are asymmetric with a long tail towards
shorter wavelengths. This is related to an intrinsic Fano-type interference between
the nanoparticle plasmon resonance and an extended interband continuum. A
Fano-type interference is a coherent interaction between two excitations where one
is much broader than the other. In the case of Pd, large continuum values of the
imaginary part of the permittivity from UV to visible are related with a free electron
“Drude contribution” that overlaps with an extended continuum background.38 While
Ag and Au show interband energy thresholds, in the case of Pd, the d-electrons
overlap the s- and p-electrons in the whole wavelength range leading to this uniform
background observed in ε´´m() (Fig. 1). Therefore, the asymmetric extinction
spectra of Pd nanoparticles is due to the overlap of the interband continuum
observed in ε´´m() with the plasmon resonance which provokes a Fano-type
interference. In metals as Ag or Au, this effect could be reduced shifting the LSPR
to longer wavelengths where the contribution of intraband transitions to the
imaginary part can be reduced.43 However, in the case of Pd nanoparticles, it is not
possible to avoid this Fano-type interference as the intraband transitions
contributions to ε´´m() are a continuum maintained in the whole wavelength range
of interest.38
Nanoscale Page 12 of 42

View Article Online


3. Synthesis of plasmonic Pd nanoparticle DOI: 10.1039/D0NR06270G

3.1. The case of pure Pd


The first colloidal approaches to fabricate pure Pd nanoparticles evidencing
plasmonic response in the visible region were published by Xia´s group.44 They
reported the synthesis of Pd nanocubes, nanoplates or cages via reduction of a Pd
precursor by ethylene glycol in the presence of PVP (polyol method) and the
presence of different additives such as FeCl3 (an oxidative etchant for Pd), HCl or
H2O dictated the number, crystalline structure, and growth of the Pd seeds. For

Nanoscale Accepted Manuscript


instance, the addition of HCl and FeCl3 in the polyol reaction medium promoted the
formation of triangular or hexagonal nanoplates since it significantly slowed down
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

the reduction kinetics due to the oxidative etching capabilities of Fe3+ species and
O2/Cl- pair. Importantly, despite the polydispersity, both types of nanoparticles
exhibited, in agreement with DDA simulations, LSPR around 520-530 nm. Those
particles showed SERS activity when tested using 4-mercaptopyridine as Raman
probe. On the other hand, the presence of water induced the transformation of single
crystal Pd nanocubes (Fig. 3A) into nanoboxes (Fig. 3B) and nanocages (Fig. 3C)
via the combination of corrosive pitting and etching of Pd (Fig. 3E).45 Interestingly,
the LSPR of the Pd nanoparticles could be tuned from 410 to 520 nm by modulating
the wall thickness of the nanoboxes (see Fig. 3D). As shown in Fig. 3F, the process
starts with the water-induced generation of a pit at a specific site on the Pd cube
surface and the preferential corrosion inside the pit. The further etching of the inner
part of the nanocube gave rise to a hollow nanostructure. The polyol reduction of
the water-substituted [PdCl4]2- species, formed during the etching process, at the
edge of the surface holes produced completely enclosed Pd nanoboxes (Figura 3B).
The formation of nanocages (Fig. 3C) took place at later stages via dissolution of
Pd from the nanoboxes corners.
Page 13 of 42 Nanoscale

View Article Online


DOI: 10.1039/D0NR06270G

Nanoscale Accepted Manuscript


Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Fig. 3. Synthesis of Pd nanoboxes and nanocages via transformation of Pd


nanocubes through the combination of corrosive pitting and etching of Pd.
(A-C) Electron micrographs of Pd nanocubes (A), nanocages(B), and nanoboxes
(C). (D) Time evolution of the extinction spectra observed during the transformation
of Pd nanocubes. E) Schematic illustration summarizing all the major morphological
changes involved in the synthesis of Pd nanoboxes and nanocages by corrosive
etching: a) Selective pitting at nanocube surface where the oxidation of Pd gives
rise to the formation of hydroxide ions from O2 dissolved in the solvent; b) further
etching of the nanocube interior and formation of hollow structures, and increase of
water-substituted [PdCl4]2-; c) formation of a Pd nanobox by reducing the water-
Nanoscale Page 14 of 42

substituted [PdCl4]2-species to form Pd atoms at the edge of each hole; d) formation


View Article Online
of nanocages by dissolving Pd from the nanobox corners; and e)DOI:
nanocage
10.1039/D0NR06270G

reconstruction by relocating all the holes from the corners to side faces and
thickening of the wall via additional reduction of the water-substituted [PdCl4]2-
species. Reprinted (adapted) with permission from ref. 45. Copyright 2005 Wiley-
VCH Verlag GmbH & Co. KGaA, Weinheim.

Several authors reported the synthesis of Pd nanocubes for catalytic

Nanoscale Accepted Manuscript


applications therefore no attention was paid to the analysis of their optical
properties.46 Nevertheless, Niu and coworkers47, 48 demonstrated that regular and
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

concave Pd nanocubes can also support tunable plasmon modes from 350 to 715
nm. The Pd nanocubes were growth on cubic Pd seeds by reducing H2PdCl4 with
ascorbic acid in the presence of cetyltrimethylammonium bromide (CTAB).
Moreover, in the case of concave Pd nanocubes (Fig. 4A), a small amount of CuSO4
was required to promote the growth of tunable sub-10-nm edges and corners since
Cu2+ ions selectively activated the fast growth of Pd along the [110] and [111]
directions. In both cases, the modulation of the edge lengths was carried out by
varying the amount of seeds. Interestingly when comparing the optical response of
regular and concave nanocubes with similar sizes, that the edges and corners
dramatically shift the LSPR to longer wavelength.
Using a seed-mediated synthesis approach, Chen et al. prepared uniform
palladium nanorods with average lengths of ∼ 200 nm and 300 nm (optical
properties in the NIR region) in water (Fig. 4B).4 The key parameter inducing the
anisotropic growth was the presence of copper acetate during the growth stage.
Nevertheless, they did not demonstrate the possibility of tuning the LSPR in a
controlled manner. Interestingly, when large amounts of copper acetate were used
branched Pd nanocrystals with optical properties at 600 nm were obtained. Recently
it has been also reported the synthesis of Pd nanorods with aspect ratios from 3.1
to 6.7 via the chemical reduction of a metal precursor in the presence of a surfactant
mixture of CTAB and Pluronic copolymers.49 The aspect ratio the rods were
determined by the number-average molecular weights and the ratio between
propylene oxide and ethylene oxide of triblock copolymers. Although the TEM
analysis show uniform nanoparticles, their optical properties are rather poor.
Page 15 of 42 Nanoscale

View Article Online


DOI: 10.1039/D0NR06270G

Fig. 4. Representative SEM and TEM images of plasmonic Pd nanoparticles


with different morphologies obtaining different chemical methods. (A) Size

Nanoscale Accepted Manuscript


tunable concave Pd nanocubes prepared via a seeded-mediated method. Reprinted
(adapted) from ref. 47 Copyright 2014 American Chemical Society. (B) Pd nanorods
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

synthesized by a seeding growth approach with the addition of Cu(OAc)2 into the
growth solution. Reprinted (adapted) from ref. 4. Copyright 2009 American Chemical
Society. (C) Biogenic porous Pd nanoparticles synthesized using an aqueous chaga
mushroom extract. Adapted from ref. 50 with permission from The Royal Society of
Chemistry.

Pd nanoparticles with a remarkable tuning of their LSPR in the NIR region were
first introduced by Zheng´s group.16, 51 They synthesized ultrathin colloidal
hexagonal palladium nanosheets through the reduction of Pd(acac)2 in a solvent
(such as DMF, benzyl alcohol) containing a halide salt and PVP under CO
atmosphere. They hypothesized that CO strongly absorbed on the basal (111)
planes of Pd nanoparticles inhibiting the growth along the [111] direction and
directing the nanosheet formation. Importantly, the control of the nanosheets edge
length by varying the experimental conditions (reaction time, solvent, presence of
halides, or size of the seeds), allows to modulate the LSPR from 700 to 1068 nm.16,
51 The nanosheets highly stable upon NIR irradiation showed suitable properties for
photothermal therapy (photothermal conversion efficiency (52.0%) at 808 nm) and
electrocatalysis. Biogenic porous Pd nanoparticles (Fig. 4C) with rough
protuberances were synthesized in an aqueous extract of chaga mushroom.50 The
nanoparticles show a broad band centered at 400 nm. Interestingly, the chaga
extract adsorbed on the surface of the nanoparticles provides the particles with
anticancer effect.
Plasmonic Pd nanoparticles have been also obtained employing non-colloidal
strategies. For instance, Langhammer and coworkers introduced a hole-mask
Nanoscale Page 16 of 42

lithography method for the synthesis of supported Pd nanodisks.28 Thus, Pd


View Article Online
nanodisks (20 nm height) were obtained by thermal evaporation of Pd on
DOI: a PMMA
10.1039/D0NR06270G

resist with holes. After the Pd evaporation the PMMA resist was subsequent lift-off
in acetone. The diameter of the disks was varied from 32 nm to 511 nm which
allowed to tune the LSPR from 300 to 1900, evidencing the aspect-ratio dependence
of the LSPR. When comparing with Ag disks, Pd exhibited broad LSPR and higher
sensitivity to the aspect ratio. Besides, theoretical calculations demonstrated that
the optical response of Pd disks was dominated for the absorption contribution
which indicated different behavior respect to Ag.

Nanoscale Accepted Manuscript


Finally, Jung and coworkers52 synthesized Pd nanorods through the
potentiostatic electrochemical deposition of Pd in an anodized aluminum oxide
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

template. The diameter of the Pd nanorods was determined by the template pore
size (D=83 nm), whereas the length could be tailored by adjusting the total charge
passed through the electrochemical cell (from L=188 to 497 nm). Although the
technique allowed controlling aspect ratio of the nanorods and therefore the LSPR
from 400 nm up to the NIR region, the polydispersity of the nanorods resulted in
broad optical features. Moreover, Pd nanorods were more sensitive to variations in
rod length when comparing with Pt and Au nanorods.

3.2. Synthesis of hybrid Pd-Au/Ag nanoparticles

A wide variety of bimetallic metal@Pd core-shell nanostructures have been


prepared using Au as building blocks.53,54-57 Spite of the large lattice mismatch
between Au/Ag and Pd (about 4.7%), Pd can grow on Au/Ag surfaces releasing the
lattice strain through the formation of dislocations and stacking faults. Often the
bimetallic structures reported in the literature either are dominated by the optical
features of Au or do not exhibit well-defined optical features. Herein we are mainly
mentioning those reports showing Au@Pd nanoparticles with Pd-like optical
behavior. Xiang et al.58 showed the synthesis of rectangular Au@Pd nanorods with
sharp edges with Pd optical features. The Au@Pd nanoparticles resulting from the
epitaxial deposition of Pd atoms on the {110} facets of Au nanorods. As the Pd layer
thickness increases, a blue-shift of the LSPR band is observed, along with a
significant broadening and damping. Wang´s group54-56, 59 has dedicated a lot of
effort to study different aspects regarding bimetallic core-shell nanostructures based
on Pd. Among them, it should be pointed out the synthesis of Au@Pd nanorods with
Page 17 of 42 Nanoscale

a relatively narrow LSPR band at the visible-NIR region in the presence of


View Article Online
quaternary ammonium surfactants (CTA+).59 Importantly, the counterion of the
DOI: 10.1039/D0NR06270G

CTA+, Br- or Cl-, determines the growth of the Pd on the Au surface as a continuous
shell or forming discrete nanoparticles, respectively. Those particles were tested for
H2 sensing (see section 4.3). More recently, Rodal-Cedeira and coworkers33
synthesized plasmonic Pd-like nanorods (Fig. 5A) with relatively narrow and tunable
optical responses across the entire visible-NIR region (Fig. 5B) via the preferential
and controlled Pd deposition on the {111} tip facets of penta-twinned Au nanorod
since bromide ions from CTAB strongly interact with the lateral [100] facets of

Nanoscale Accepted Manuscript


nanorods suppressing Pd deposition on those facets. In fact, the use of CTAC with
Cl- counterions led to a rough Pd shell on the entire Au surface. Interestingly,
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

theoretical calculations demonstrated that the Au nanorods boosted the optical


response of Au@Pd nanorods while maintaining pure Pd-like surface properties.
The particles show high performance for LSPR based sensing (see section 4).

Fig. 5. Au/Pd and Au@Au nanoparticles of different morphologies. (A) TEM


image of Au@Pd core-shell nanorods prepared using pentatwinned Au nanorods
as seeds and Br¯ as directional growth agent and their corresponding UV-visible-
NIR extinction spectra (B), as a function of Pd concentration. Reprinted (adapted)
from ref. 33. Copyright 2016 American Chemical Society. (C) Au@Pd core-shell
nanocrystals with high-index {730} facets synthesized with ascorbic acid and CTAC
Nanoscale Page 18 of 42

and the corresponding size-dependent UV-visible-NIR extinction spectra (D). The


View Article Online
inset in (C) shows a schematic drawing of a tetrahexahedral nanocrystal showing
DOI: 10.1039/D0NR06270G

the axis projecting along the [730] direction. Reprinted (adapted) from ref. 34.

Copyright 2010 American Chemical Society. (E) Au/Pd octopods prepared by


coupling a seed-mediated method with co-reduction and the corresponding UV-
visible-NIR extinction spectra (F), as a function of Pd concentration. Reprinted
(adapted) from ref. 60. Copyright 2011 American Chemical Society.

Several authors reported the fabrication of Pd nanospheres with size control

Nanoscale Accepted Manuscript


through a seed-growth approach using small Au nanoparticles as seeds, but most
of them did not pay attention in the nanoparticles optical response.61,62,63
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Nevertheless, Sugawa et al.15 64demonstrated that core-shell Au@Pd nanoparticles


with a 20 nm Au core and 25 nm Pd shell exhibit optical response in the visible
region and more importantly higher refractive index susceptibility than Au or Ag
nanoparticles of similar dimensions. Recently, Fernández-Lodeiro et al.64 proposed
the synthesis of spherical Pd nanoparticles with size control through a seeded
growth approach mediated by Fe(II) on Au seeds. Alternatively, using Au nanocubes
as the structure-directing cores and varying the experimental conditions,
tetrahexahedral, concave octahedral, and octahedral Au-Pd core-shell nanocrystals
were synthesized with ascorbic acid and CTAC as reducing and capping agent,
respectively (Fig. 5C).34 The Cl- from CTAC played a crucial role in inducing the
oxidative etching of Pd in the presence of oxygen to give rise to the tetrahexahedral
morphologies with high-index {730} facets. The optical properties of these
nanoparticles could be modulated from 435 nm to 710 nm by varying the
tetrahexahedral Au-Pd nanocrystals sizes from 56 to 124 nm (Fig. 5D).
Skrabalak et al. developed a strategy to obtain different binary Au-Pd
nanocrystals by combining the seed-mediated method with the co-reduction of Pd
and Au precursor using L-ascorbic acid and CTA+, as reducing agent and stabilizer,
respectively.60, 65 The careful control of growth kinetics allowed to tune the
morphology and composition of the final binary structure, among Au/Pd octopods
(Fig. 5E) and concave and shape-controlled alloy nanostructures or concave
Au@Pd core-shell nanocrystals, among others. The systematic analysis of the
experimental conditions revealed that the main factors dictating the final binary
nanostructures were: (i) Au:Pd precursor ratio, (ii) reaction pH, and (iii) capping
agent concentration. For instance, experiments performed at different Au:Pd
Page 19 of 42 Nanoscale

precursor ratios showed that ratios from 1:0.01 to 1:0.5 induced the formation of
View Article Online
Au/Pd octopods while concave Au@Pd nanoparticles were obtained DOI:
for10.1039/D0NR06270G
smaller
ratios. It was due to the differences in pH during the growth process produced by
the H2PdCl4 concentration. In terms of optical properties we would like to stress here
the size- and shape-dependent LSPR of the Au/Pd octopods. These nanocrystals
are formed by eight symmetrical branches growing along the ⟨111⟩ directions of a
cuboctahedral Au core with the Pd localized along the tips of the branches on the
outer surface. It was demonstrated that both branch length and tip width determined
the position of the LSPR of the octopods, which can be tuned throughout the visible

Nanoscale Accepted Manuscript


and NIR region (Fig. 5F).60, 65, 66
Chen and coworkers67 showed the preparation of hollow PdAg nanoboxes with
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

LSPR tunable to the NIR region via galvanic replacement. The method consists of
simply titrating the Pd precursor into a pre-synthesized Ag nanocubes under reflux
and vigorous stirring. The modulation of the LSPR from 440 to 730 nm is achieved
by controlling the molar ratio of Na2PdCl4 to Ag. DDA calculations revealed that the
optical properties of Pd-Ag nanoboxes were dominated by the absorption
contribution, which makes them particularly useful in photothermal heating
applications.
This section is summarized in Table 1 where the most relevant methods to
obtain pure and hybrid nanoparticles with Pd-like properties are shown. Moreover,
information about particle shape, size tuning, and LSPR range has been included.

Table 1. Summary of the morphology, size, and LSPR range of the pure and
hybrid Pd nanoparticles.

Edge length
Composition Morphology LSPR range (nm) Ref.
(nm)
triangular nanoplates 28 520
44
hexagonal
 20 530
nanoplates
nanocubes 48 410 45
cubic nanocages 48 430 - 520
concave nanocubes 67 - 109 350 - 715 47

nanocubes 37 - 109 290 - 580 48

hexagonal
Pd 20 - 160 700 - 1068 16
nanosheets
hexagonal
4.4 700 51
nanosheets
rough nanospheres 77 - 92 < 400 50

supported nanodisks 32 - 511 300 - 1900 28


Nanoscale Page 20 of 42

branched
< 100 600
nanocrystals 4 View Article Online
DOI: 10.1039/D0NR06270G
nanorods 200 - 300 1800 - 2100
nanorods  100 - 49

nanorods 188 - 497 400 - 1700 52

rectangular
50 700 58
nanorods
nanorods < 100  700
nanorods 25 - 170 900 - 1215 33

Au@Pd nanospheres 73 336 15


(core-shell)
nanospheres 15 - 90 228 - 420 64

tetrahexahedral 56 - 124 435 - 710


concave octahedral - 34

Nanoscale Accepted Manuscript


octahedral  100 -
octopods
Pd - Au 60, 65, 66
concave > 100 850 - 1030
(alloy)
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

nanocrystals
Pd – Ag nanoboxes 63 440 - 730 67

4. Plasmonic driven applications of Pd nanoparticles

The intrinsic physicochemical features of Pd, such as high thermal and


chemical stability, catalytic effect, surface functionalization, as well as their tuneable
optical response, endow Pd nanostructures with great potential in different fields
including (photo)catalysis,6 biomedicine,68,69 H2 storage and sensing,9, 70, 71 etc. We
will focus this review section on plasmon based applications of Pd-like
nanoparticles.

4.1. Photothermal therapy.

Plasmonic nanoparticles supporting LSPR at the visible and NIR regions have
been exhaustively investigated for photothermal therapy72,73 due to their ability to
convert light into heat (photothermal effect) as well as the easy bioconjugation for
specific cell targeting. Compared to short-wavelength light, NIR light can penetrate
tissue deeply and is not harmful to the human body, thereby allowing a minimally
invasive treatment.74 Therefore, Au nanoparticles absorbing light at tissue
transparency windows, such as nanorods, nanoshells, or nanostars, have been
extendedly investigated.73, 75, 76 Owing to their higher thermal stability and larger
absorption contributions (see section 2), the use of Pd nanoparticles for
photothermal therapy has been explored in a number of cell and mouse model
studies.
Page 21 of 42 Nanoscale

Zheng et al. investigated different aspects of sub-10 nm hexagonal Pd


View Article Online
nanosheets with tuneable NIR LSPR for their potential application in photothermal
DOI: 10.1039/D0NR06270G

therapy (Fig. 6).16, 51, 77-80 Pd nanosheets demonstrated high biocompatibility,


efficient photothermal conversion (photothermal cell-killing efficacy), and enhanced
photostability compared to Au or Ag nanoprisms. The systematic evaluation of the
influence of the size of PEGylated Pd nanosheets (i.e., 5, 13, 30, and 80 nm)
showed that sizes up to 30 nm achieved a temperature of 53 oC, which is sufficient
for successful photothermal therapy.78 Remarkably, Pd nanosheets of 30 nm
exhibited the highest accumulation in tumors, whereas 80 nm nanoparticles only

Nanoscale Accepted Manuscript


reached 45 oC, and they were eliminated from the blood faster than the smaller
nanosheets.78 In a different study, the effects of different surface coatings (PEG-
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

NH, carboxymethyl chitosan, PEG-SH, and dihydrolipoic acid-zwitterion) on the


blood circulation, biodistribution, toxicity, and anticancer activity were investigated.79
Among the four coatings assessed, PEG-SH provided Pd nanosheets with the
longest circulation half-life in blood and the highest tumor accumulation. Moreover,
temperature measurement within tumors revealed an increment from 35 oC to 52 oC
after 5 minutes of laser exposure, leading to tumor ablation. Interestingly, another
work reported that the silica coating of Pd nanosheets promoted cell entry, thereby
enhancing the photothermal cell-killing effect.80 The high efficacy of small Pd
nanosheets (average diameter of 4.4 nm and LSPR at 700 nm) for photothermal
ablation of liver cancer cells (Fig. 6A) motivated Zheng’s group to test them for NIR
photothermal cancer therapy in vivo (Fig. 6B-C).51 Pd nanosheets were
functionalized with reduced glutathione (GSH) to avoid adsorption of serum
proteins, which significantly contribute to prolong their blood circulation. This in turn
increased the nanoparticle concentration in tumors facilitating their complete
ablation upon 808 nm laser irradiation (Fig. 6B). Additionally, it was observed fast
clearance of the nanostructures through the renal excretion route into the urine,
attributed to their size below the renal filtration limit (<10 nm) (Fig. 6C). Hollow Pd
nanospheres with tunable sizes, and therefore tailored optical responses in the
visible range, showed high photothermal efficiencies.81 The calculation of the
photothermal transduction efficiency for 90 nm hollow nanospheres revealed an
efficiency as high as 70%. Interestingly, the conjugation of the Pd nanoparticles with
the photosensitizer Chlorin e6 (Ce6) provides them the capability of single oxygen
generation. Therefore, hollow Pd@Ce6 nanoparticles might integrate photothermal
and photodinamic therapies, as was demonstrated for HeLa cells.
Nanoscale Page 22 of 42

View Article Online


DOI: 10.1039/D0NR06270G

Nanoscale Accepted Manuscript


Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Fig. 6. Application of Pd nanoparticles for photothermal therapy and


photoacoustic imaging. (A) Characterization and NIR photothermal properties of
Pd nanosheets (SPNS) in vitro. (a) TEM image and (b) visible-NIR spectrum of the
SPNS. Inset: SPNS diameter distribution (c) The temperature versus time recorded
for different SPNS concentrations (0, 7.5, 15 and 30 ppm) upon irradiation. (d)
Micrograph corresponding to photothermal therapy in vitro by PEI-coated SPNS at
an optical power density of 0.8 W cm−2. Dead cells are stained with trypan blue. (B)
Photothermal ablation of tumors in mice using intravenously injected GSH-coated
SPNS. (a) Relative Tumor Volume (RTV) of different groups after treatment. The
tumor volumes were normalized to their initial sizes (n = 10 per group). (b) Mortality
Free Survival (MFS) curves of mice bearing 4T1 tumor after various treatments
indicated. GSH-coated SPNS injected mice after photothermal therapy survived
over 40 days without any single death. (c,d) Representative photographs of the mice
before (c) and after (d) treatment. The laser-irradiated tumors on SPNS‐GSH
injected mice were completely ablated. (C) In vivo clearance and biodistribution of
SPNS modified by reduced glutathione (GSH). (a) Urinary cumulative excretion of
GSH-coated SPNS in rats (n = 3) following i.v. administration at a single dose of 10
mg kg−1. (b) Biodistribution of GSH-coated SPNS in different organs in mice (n = 3)
within 15 days after intravenous injection. The percentage of the injected dose was
calculated based on the Pd concentration measured by ICP‐MS. From A to C,
Page 23 of 42 Nanoscale

images were reprinted with permission from ref. 51. Copyright 2014, WILEY-VCH.
View Article Online
(D) Photoacoustic imaging of Pd@Au-PEG nanoparticles in tumorDOI:sites and
10.1039/D0NR06270G

quantification photoacoustic signals over time. Reprinted with permission from ref.
82 Copyright 2014, WILEY-VCH.

Functionalization of Pd nanoparticles with peptides is an attractive approach


for realizing active targeting to diseased cells and tissues as well as for enhancing
their photothermal ablation capabilities.83 For instance, Pd nanoparticles (average
size of 22 nm) bioconjugated with an RGD peptide showed specific binding to αvβ3

Nanoscale Accepted Manuscript


integrins, which are overexpressed on activated endothelial cells in tumors.
Recently, Gao et al.84 employed Pd nanosheets functionalized with TAT polypeptide
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

to inhibit cancer cell migration (i.e. metastasis). Interestingly, small Pd nanoparticles


exhibiting poor optical response in the visible-NIR region have also shown high
photothermal conversion efficiency when functionalized with TAT or RGD
peptides.85
Zheng´s group reported the fabrication of core-shell Pd@Ag and Pd@Au
nanoplates for photothermal studies. 17, 86,82,87 In such studies, Pd acts as a template
to drive the fabrication of uniform nanoparticles and to improve their thermal stability,
while gold and silver are responsible for the plasmonic response of the core-shell
nanoplates. In a different study, porous AuPd alloy nanoparticles were used as
hyperthermia agents for bone regeneration by means of photothermal therapy.88
Chemo-photothermal therapy. The synergistic effect that results from the
combination of photothermal therapy with chemotherapy could be a more effective
way to destroy cancer cells than the cytotoxic effects elicited by both approaches
individually. Several Pd based nanostructures and nanocomposites have been
investigated for NIR-triggered chemo-photothermal treatment of cancer cells.50, 89-95
Fang et al.89 demonstrated such synergistic effect in cancer cells cultured in vitro
using Pd nanosheet covered with hollow mesoporous silica nanoparticles containing
doxorubicin (DOX) molecules. It was shown that upon particle internalization, the
NIR irradiation generated cytotoxic heat, whereas the low pH in the endosomes
triggered the release of the DOX molecules, generating a chemotherapeutic effect.
Yang and coworkers fabricated a ZIF-8 metal–organic framework nanocomposite
containing Au coated Pd nanoparticles and DOX aiming at the synergistic release
of the antitumor drug in endosomes through the acid-induced dissolution of ZIF-8
and NIR irradiation.90 Porous Pd nanoparticles loaded with a therapeutic agent and
Nanoscale Page 24 of 42

conjugated with cell-penetrating peptides or transferrin showed good performance


View Article Online
in photothermal conversion and subsequent release of the therapeutic cargo.
DOI: 10.1039/D0NR06270G

Nguyen et al. also focused their study in a pH-sensitive nanocarrier consisting of


porous Pd nanoparticles loaded with paclitaxel and functionalized with poly(acrylic
acid)-poly(ethylene oxide) for providing pH-sensitivity and transferrin for active
targeting to tumor cells. The authors reported that irradiation with NIR light
contributed to the high release rate of the chemotherapeutic agent.94 A tri-modal
combination of photothermal, chemical, and biological cancer treatment was
recently investigated using biogenic anisotropic porous Pd nanostructures

Nanoscale Accepted Manuscript


fabricated with a chaga extract and functionalized with DOX.50 The nanoparticles
exhibited a surface chaga extract-derived anticancer effect, controlled delivery of
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

doxorubicin through electrostatic interaction, and a photothermal conversion effect


under NIR irradiation.

4.2 Photoacoustic and multimodal imaging.

Photoacoustic imaging (PAI) is based on photoacoustic effect, defined as the


generation of acoustic waves due to the heat produced upon visible or NIR
irradiation.96 In general, PAI can provide centimeter penetration depth in biological
tissues, which is well-suited for biomedical imaging applications, image-guided
therapy, and theranostics.97 Metal nanoparticles absorbing in the NIR region are
commonly used as contrast agents for PAI. Since the photothermal stability of the
nanoparticles is a key issue for imaging purposes, Pd is an ideal candidate as a
contrast agent for PAI. Nie et al.98 reported the first PAI system based on Pd
nanoparticles. Remarkably, it was demonstrated that hexagonal Pd nanosheets (16
nm diameter) were much more stable at maintaining both their morphology and
optical response under long-term laser irradiation than plasmonic Au nanorods. The
authors reported stable and robust photoacoustic images of nanoparticle
accumulation in tumors developed in mice, which was attributed to the EPR effect.
Assessment of the biodistribution of the nanoparticles in the tumor and main organs
(liver, spleen, heart, lung, and kidney) revealed that the liver had the highest Pd
concentration. MTT assay and ex vivo histological examination revealed no obvious
toxic-side effects as a result of nanoparticle administration, thereby suggesting good
biocompatibility of the material. Chen et al.82 employed PEGylated hexagonal
Pd@Au nanosheets (30 nm diameter) as contrast agents, which showed a gradual
accumulation in tumor sites with an increment of the photoacoustic signal after 24
Page 25 of 42 Nanoscale

hours post-injection (Fig. 6D). Analysis of the nanoparticle biodistribution revealed


View Article Online
a preferential accumulation in tumors as compared to main organs, DOI:
which was
10.1039/D0NR06270G

attributed to surface PEGylation of the nanostructures. Also, the influence of the


size of PEGylated Pd nanosheets (from 5 to 30 nm) in the nanoparticle's
performance for PAI was not significant. Bharathiraja demonstrated the efficiency of
chitosan oligosaccharide-coated Pd nanoparticles functionalized with RGD peptides
both in vitro and then in vivo.83 As an alternative approach to the use of
chemotherapeutical agents, Zhao et al. fabricated plasmonic cubic Pd hydride
(PdH0.2) nanocrystals for combined hydrogenothermal treatment of cancer in mouse

Nanoscale Accepted Manuscript


models by means of NIR-triggered release of therapeutic hydrogen gas and
photothermal effect.99 In this study, PAI was applied to monitor intratumoral
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

accumulation of PdH0.2 nanocrystals and therapy guidance, showing that the PAI
intensity within the tumor reached its maximum as early as 2 hours after vein
injection.
Pd based nanomaterials enable to perform other imaging modalities such as
single photon emission computed tomography (SPECT),100 and coherent anti-
Stokes Raman scattering (CARS).101 pH-sensitive ultrasmall Pd nanosheets (5 nm)
radiolabeled with halide ions were used as a theranostic agent for multimodal
imaging of tumors by PAI and SPECT.100 Rubio-Ruiz employed CARS to
simultaneously induce microablation of cancerous tissue in mice and imaging the
cells labeled with the plasmonic materials.101 Finaly, taking advantage of the
increase in temperature induced by the Pd based nanostructures upon NIR
irradiation, IR thermal imaging was applied as a means to evaluate the materials
and approaches.78, 79, 82, 83, 85, 90-92, 99, 100,84, 102
Nanoscale Page 26 of 42

View Article Online

A 1000
DOI: 10.1039/D0NR06270G

RIS (nm/RIU)
Ag nanoplates
800 Ag nanocubes
Au nanospheres
Au nanorods
600 Au nanobypiramids
Au nanobranches
Au nanoframes
400 Au@Pd nanorods
Au@Pd nanobars
Branched Au-Pd
200 Au@Pd tetrahexahedra
Au@Pd octopods
Au@Pd octahedra
0
400 600 800 1000 1200 1400 1600
LSPR Wavelength (nm)

Nanoscale Accepted Manuscript


Pd
B 120
Pd@Au
Au
1213 nm/RIU
LSPR shift (nm)

877 nm/RIU
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

llong= 1110 nm

80

741 nm/RIU
40

0
1.34 1.36 1.38 1.40 1.42 1.44 1.46
Refractive index

Fig. 7. Refractive index sensitivity of Pd nanoparticles. (A) Experimental


refractive index sensitivity (RIS) as a function of LSPR wavelength for nanoparticles
of various morphologies and compositions (stars, circles, and squares correspond
to Ag, Au, and Au-Pd, respectively). (B) Calculated dependences of the longitudinal
LSPR shift on refractive index changes for a single PTW nanorod of pure Au, pure
Pd, and core-shell Au@Pd with the longitudinal LSPR centered at 1100 nm.
Reprinted (adapted) from ref. 33. Copyright 2016, respectively, American Chemical
Society.

4.3. LSPR sensing

4.3.1. Refractive index sensitivity.


LSPR shifts to longer wavelengths when the refractive index surrounding a
plasmonic nanoparticle increases. This well-known phenomenon can be exploited
for detecting analytes even at low concentrations. However, to obtain a device with
an optimized sensitivity, the selection of geometry and size and nature of
nanoparticles is critical. Refractive index sensitivity (RIS) of a plasmonic material is
quantified by the refractive index susceptibility (S) which can be estimated
Page 27 of 42 Nanoscale

experimentally as S = Δλpeak / Δn, where Δλpeak is the difference in LSPR peak


View Article Online
position of a plasmonic material and Δ𝑛 the difference in the refractive index of the
DOI: 10.1039/D0NR06270G

surrounding medium. Miller et al.,103 performing electrodynamic simulations of Au


nanoparticles with small sizes relative to the wavelength of light, found that the RIS
is mainly determined by the location of the LSPR peak wavelength and the dielectric
properties of the metal. Apart from that, they also found that RIS depends linearly
with the LSPR peak wavelength regardless of nanoparticle geometry parameters.
Similar results were obtained by Zalyubovskiy et al.104 Thus, Au cylinders and
nanodisks supporting LSPR at the same wavelengths have the same RIS. Taking

Nanoscale Accepted Manuscript


this into consideration, plasmonic nanoparticles of a certain metal exhibiting LSPR
located in the NIR will be more sensitive to refractive index changes than those with
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

LSPR in the visible (Fig. 7A).33


As it was explained previously in section 2.1, anisotropic Pd nanoparticles can
show LSPR at larger wavelengths than Au and Ag for the same nanoparticle
geometry. Therefore, Pd might be potentially a better plasmonic material with higher
RIS than Au or Ag. Recently, Rodal-Cedeira et al.33 found that highly anisotropic Pd
nanorods (S= 1213 nm / RIU) were more sensitive to refractive index changes than
Au nanorods (S= 741 nm / RIU) (Fig. 7B). Also, Sugawa et al.15 showed that core-
shell Au-Pd nanospheres achieved S values 4.9 and 2.5 times larger than for Au
and Ag nanospheres, respectively. Additionally, Sugawa et al. also experimentally
demonstrated that small Pd nanodisks (16 nm in diameter and 1.5 nm in thickness)
exhibited a refractive index susceptibility comparable to those reported for highly
anisotropic Au nanoparticles (nanorodos, nanopyramids, etc.) and excellent LSPR
sensing materiales.35 Interestingly, S was strongly dependent on the thickness of
the capping layer.
To understand the relevance of the dispersion of the dielectric function of the
metals in the refractive index susceptibility, it is necessary to obtain a mathematical
relation between S and the real and imaginary part of the metal permittivity. Using
a quasistatic theoretical framework is possible to approximate an analytical
expression for the refractive index susceptibility S as follows:15,103

, ,
𝑆= = (8)

where 𝜆 is the incident light wavelength, n is the refractive index of the surrounding
medium, ´m is the real part of the permittivity of the metal and 𝜆 is the LSPR peak
Nanoscale Page 28 of 42

wavelength. According to this relation (see Eq. (8)), the RIS is inversely proportional
View Article Online
to the slope of the real part of the permittivity of the metal at the LSPR condition:
DOI: 10.1039/D0NR06270G

,
m= (9)

If we compare the dispersion of ´m of Pd with that for Au and Ag for similar
´m values (Fig. 1), the slope (see Eq. (9)) of the Pd dispersion curve is smaller than
for Au and Ag providing a larger RI susceptibility for Pd nanoparticles.
It should be pointed out that even though Pd nanoparticles have shown larger RIS,
their use in the field of LSPR sensing is almost limited to H2 detection (see section

Nanoscale Accepted Manuscript


4.3). This can be ascribed to several issues: (1) the lack of studies with Pd, (2) the
difficulty of determination of LSPR peak positions on Pd (broad LSPRs), and (3) the
low Figure of Merit (FOM) exhibited by most of the investigated Pd based
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

nanostructures.
The difficulty of determination of LSPR peak positions in a wide LSPR band
can be easily overcome by using the second derivative of the spectrum to determine
LSPR peak. On the other hand, FOM is defined as the ratio between S and the
LSPR full width at half maximum (fwhm). In the case of Pd based sensors, FOM (S
/ fwhm) can be drastically reduced by the large values of fwhm of Pd LSPR bands
due to the large Pd imaginary part of the permittivity (see Figure 2.a). However, this
effect can be easily overcome by combining Pd with Au or Ag.33, 105,66 These core-
shell bimetallic nanoparticles do not show linearly mixed dielectric values between
Au, Ag, and Pd. Theoretical calculations performed on core-shell Au@Pd nanorods
with overall dimensions of 180x30 nm, where the width of the Au nanorod was 25
nm and the length varied from 25 nm to 170 nm, showed a sharper and more intense
LSPR peak as the amount of Au present in the core increased.33 Interestingly, the
LSPR wavelength was not affected by the Au content. As Pd has also extraordinary
plasmonic properties as, for example, higher sensitivity to changes on the
surrounded refractive index, bimetallic nanoparticles maintain RIS but with a
boosted plasmonic response.106,107

4.3.2. LSPR hydrogen sensing.


The detection of H2 plays a key role in many important areas such as the
hydrogen economy, the chemical industry, as well as the health care and food
industry. Among the different emerging hydrogen-sensing technologies, plasmonic-
based sensors are considered as a future technical solution for hydrogen
detection108 Focusing on these sensors, Pd nanoparticles have been intensively
Page 29 of 42 Nanoscale

explored in recent years due to their ability to selectively and reversibly absorb H2
View Article Online
molecules into their lattice forming the corresponding hydride. Thus, Pd is one
DOI: of the
10.1039/D0NR06270G

most frequently used materials for direct or indirect LSPR hydrogen sensing.
In the case of direct sensing, the detection is based on the change in the LSPR
position, intensity, or line width upon H2 absorption due to the change in the metal
permittivity when the Pd hydride is formed.109 Langhammer and coworkers
investigated the use of Pd nanodisks on glass support as sensing platforms110,111
demonstrating their outstanding performance for hydrogen storage and real-time
monitoring. The Pd nanodisks reached sensitivities below 50 ppm. Interestingly, to

Nanoscale Accepted Manuscript


correlate the H2 concentration and the LSPR shift they used a quartz crystal
microbalance (QCM) in combination with the LSPR sensor.112 Moreover, they found
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

that size of the Pd nanodisks and the LSPR shift were also correlated. Strohfeldt et
al.113 demonstrated that the performance of Pd nanodisks could be improved by the
combination with CaF2 and Pt. They built up a layered CaF2/Pd/ Pt disk structure
that enhanced the long-term stability of the sensor. While the bottom CaF2 layer
reduced the stress of the palladium film, the top Pt cap avoided the Pd surface
poisoning. Apart from Pd disks, other Pd nanoparticles (nanorings114 and concave
nanocubes47, 115) has been also used for H2 sensing.
In addition to pure Pd, a variety of core-shell Au@Pd nanoparticles have been
successfully applied for direct LSPR sensing of hydrogen in gas or liquid phase.59,
116,117,33, 118 For instance, Rodal-Cedeira et al.33 demonstrated that Au@Pd
nanorods are extremely effective for the detection of hydrogen sensors both in
solution and in the gas phase (see Fig. 8). Thus, Au@Pd nanorods with a
longitudinal LSPR band at ca. 1010 nm underwent a red-shift of 42 nm when
exposed to 1.47 ppm hydrogen dissolved in the colloidal dispersion. For gas
sensing, Au@Pd nanorods were uniformly deposited on a glass substrate observing
ca. 38 nm shift in the longitudinal LSPR for 1% of H2 gas. In both methodologies,
the particles exhibited a reversible plasmonic response.
For indirect H2 detection,108 the role of Pd is to trap the H2, whereas plasmonic
nanostructures (typically Au) are used as probes of the hydride formation acting as
optical antennas and signal transducers. Since the Pd nanoparticles are not the
active material of the LSPR sensor it is out of the scope of this review.
Nanoscale Page 30 of 42

View Article Online


DOI: 10.1039/D0NR06270G

Nanoscale Accepted Manuscript


Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Fig. 8. LSPR sensing of Pd nanoparticles. (A) Schematic representation of the


palladium hydride formation after exposing an aqueous dispersion of core-shell
Au@Pd nanorods to hydrogen. (B) Evolution of the extinction spectra of penta-
twinned core-shell Au@Pd nanorods during the exposition to 1.47 ppm of hydrogen
gas dissolved in the aqueous colloidal dispersion. (C) Longitudinal LSPR shift of an
Au@Pd nanorods aqueous colloidal dispersion during different hydrogen
absorption–desorption cycles. (D) Evolution of the extinction spectra of Au@Pd
nanorods on glass during the exposition to 1% of hydrogen gas. (F) Longitudinal
LSPR shift of Au@Pd nanorods deposited on glass during different 1% hydrogen
cycles. Reprinted with permission from ref. 33. Copyright 2016, American Chemical
Society.

4.4 Plasmonic nanocatalysis.


Compared with Au and Ag, Pd nanoparticles have been much less explored in
photocatalysis. This can be due in part to a lack of knowledge about the plasmonic
properties of Pd. As explained in section 2, although their optical efficiency in the
UV region is poor, this could be significantly improved by shifting the LSPR to the
Visible-NIR region. On the other hand, Pd is an excellent catalyst for numerous
Page 31 of 42 Nanoscale

chemical reactions (low-temperature oxidation of CO and hydrocarbons, cross-


View Article Online
coupling, hydrogenations, etc.). Therefore, it is meaningful to combine the LSPR of
DOI: 10.1039/D0NR06270G

the Pd particles with their excellent catalytic properties to obtain outstanding


catalytic systems for plasmon photothermal or photochemical induced reactions.
Importantly, Pd shows higher thermal stability than other metals such as Au and Ag,
which is a critical issue for photothermal-induced catalysis.

Nanoscale Accepted Manuscript


Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Fig. 9. Plasmon induced thermocatalysis of Pd NCs@ZIF‐8 nanoparticles. (A)


Schematic representation of the Pd NCs@ZIF‐8 catalyst and their performance. (B)
SEM and TEM images of Pd NCs@ZIF‐8 nanoparticles. (C) Yields of the
hydrogenation of 1-hexene with 1 atm H2 over Pd NCs@ZIF-8 under full-spectrum
irradiation with different light intensities at room temperature or upon heating at
different temperatures. (D) Conversions of the hydrogenation of various alkenes
over Pd NCs@ZIF-8 and Pd NCs. Adapted with permission from ref. 119. Copyright
2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

So far, numerous reports have demonstrated significant plasmon-enhanced


photocatalysis (both photothermal and photochemical) employing Pd nanoparticle-
Nanoscale Page 32 of 42

based catalysts. Nevertheless, in many of them the Pd nanoparticles exhibit poor


View Article Online
DOI:120,
and not well-defined plasmonic response or it is centered in the UV region. 121,119,
10.1039/D0NR06270G

122, 123 Therefore, there is still a lot of room for improvement in this field. Weng and
co-workers found that, in spite of its poor optical response in the visible region, Pd-
CeO2 nanocatalyst (< 5 nm) exhibited a remarkable enhancement of its catalytic
activity for CO and toluene oxidation upon light excitation122 The plasmon excitation
produced hot electrons in Pd that could be transferred to the conduction band of
CeO2 thereby promoting the activation of oxygen. In another work, Pd nanocubes
encapsulated within a metal-organic framework (ZIF-8) demonstrated remarkable

Nanoscale Accepted Manuscript


LSPR-driven photothermal catalytic activity for hydrogenation of olefins at room
temperature under 1 atm H2 (Fig. 9).119 While Pd nanocubes provide the
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

nanocomposite with plasmonic photothermal effects, the ZIF-8 shell accelerates the
hydrogenation reaction by H2 enrichment, provides selectivity acting as a molecular
sieve for olefins with different sizes, and improves the recyclability of the catalyst by
enhancing the colloidal stability of the Pd nanoparticles.
On the other hand, the outstanding catalytic activity of Pd has led to its
combination with Au, Ag, and Al nanoparticles in different configurations (core-shell,
alloys, raspberry, etc.).57, 124-126,127, 128 Nevertheless, often Pd acts as the main
catalyst, whereas the other metal is responsible for the plasmon-enhanced effects.
Wang and coworkers investigated the harvesting of light energy for Suzuki coupling
reactions using Au nanorods and nanospheres with tightly bonded tiny Pd
nanoparticles.129 The photocatalytic activity of Pd-tipped Au nanorods and Pd-
covered Au nanorods were investigated for formic acid dehydrogenation by Zheng
et al.128 The photocatalytic experiments carried out at 5ºC showed higher
performance than the conventional thermal catalysis at 40ºC. Swearer et al.
reported the use of core-shell Pd-aluminum nanoparticles for plasmon-enhanced
photocatalytic hydrogenation of acetylene.124 These aluminum nanocrystals
decorated with palladium nanoparticles showed increased photocatalytic activity
over their individual components. It is important to note that several authors found
the combination of Pd with other plasmonic materials promoted hot-electron
generation. In this context, Mu and coworkers investigated yolk-shell nanostructures
bearing Au/Pd bimetal nanoparticles encapsulated within mesoporous SiO2 for
driving the plasmon-enhanced catalytic reduction of p-nitrophenol.125 They
proposed that upon plasmon excitation the intensity of the near electric field
generated at the surface of Au nanoparticles was significantly enhanced when a Pd
Page 33 of 42 Nanoscale

nanoparticle was coupled as it promoted charge separation of plasmonic hot


View Article Online
electrons and holes. Similar conclusions were found by Guo et al.126 investigating
DOI: 10.1039/D0NR06270G

the catalytic activity of core-shell Au-Pd superstructures (Au@Pd SS) for molecular
oxygen activation and carbon-carbon coupling reaction. As shown in Fig. 10 the
superstructures consisted of ordered Pd nanoarrays grown on the surface of Au
nanorods. Under light irradiation, the ordered arrangement of Pd on Au nanoparticle
surface promoted hot electron generation and transfer via amplified local
electromagnetic field to reactant molecules, as well as decreased electron−phonon
coupling. The plasmon-enhanced catalytic activity of Au@Pd SS reached around

Nanoscale Accepted Manuscript


200% enhancement compared to dark conditions for the Suzuki–Miyaura coupling
reaction between iodobenzene and benzeneboronic acid.
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Fig. 10. Plasmon-induced photochemical catalysis of core-shell Au-Pd


superstructures (Au@Pd SSs) for molecular oxygen activation and carbon-
carbon coupling reaction. (A) TEM image of Au@Pd SSs. (B) visible–NIR spectra
of Au NANORODs and Au@Pd SSs with increasing amounts of Na2PdCl4. (C)
Electric field enhancement (|E|/|Eo|) contour of Au@Pd SSs. (D) Correlations
between plasmon-enhanced catalytic activity of diverse Au@Pd nanostructures with
hot electron generation (ΔODmax/OD) and decay time (τe–p), respectively.E) TEM
Nanoscale Page 34 of 42

image of single Au@Pd SS, and the arrows indicate the periodic arrangement of Pd
View Article Online
nanoarray along the transverse [10] and longitudinal [01] axes of Au NANOROD,
DOI: 10.1039/D0NR06270G

respectively. F) Correlations between plasmon-enhanced catalytic activity of diverse


Au@Pd nanostructures with hot electron generation (ΔODmax/OD) and decay time
(τe–p), respectively. Adapted with permission from ref. 126. Copyright 2017 American
Chemical Society

4.5. SERS based applications.

Nanoscale Accepted Manuscript


As a plasmonic material several authors reported the ability of Pd
nanoparticles for enhancing the Raman scattering signal of molecules.130,131
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Surface-enhanced Raman scattering is a powerful analytical tool for molecule


identification and detection.132 The performance of this technique is strongly
dependent on the efficiency of the plasmonic nanostructures, among other
parameters.132 Considering that absorption is often the dominating extinction
process in Pd nanoparticles (see section 2), this metal is lesser efficient for SERS
than Au and Ag. Despite this limitation, the capability of pure Pd nanostructures to
enhance the SERS signals of Raman active molecules has been revealed.133,134
Nevertheless, it is more common to find it combined with gold or silver particles
although its main role is as a catalyst. Huang et al. reported the use of Au–Pd
bimetallic nanostructures, obtained by selectively growth of high-index Au–Pd alloy
horns at the tips of Au nanorods, as bifunctional platforms for highly sensitive
monitoring of catalytic reactions by SERS135. In this work, the aim of incorporating
Pd alloyed to Au on the surface of the nanorods is to provide the plasmonic Au
nanorods with catalytic activity. Recently, we reported the fabrication of Au@Pd
nanorods33 with palladium surface properties (plasmonic and catalytic) via the
selective deposition of Pd at the tips of Au nanorods. Interestingly the Au@Pd
nanorods demonstrated multifunctional capabilities catalyzing the hydrogenation
reaction of 4-nitrothiophenol (4-NTP) to 4-aminothiophenol (4-ATP) and allowing in
situ optical monitoring of the reaction by time-resolved SERS analysis. This strategy
allowed us to study in depth the intrinsic kinetics and mechanism of the surface
catalyzed reaction.

Conclusions and outlook


Page 35 of 42 Nanoscale

The plasmonic behaviour of Pd demonstrates that, in spite of exhibiting less


View Article Online
overall optical efficiency than Au and Ag in the visible-NIR region, displays
DOI: 10.1039/D0NR06270G

outstanding intrinsic properties as reviewed herein. Remarkably, the real part of the
permittivity makes the LSPR of Pd more sensitive to changes in the particle shape
and size than in case of Au and Ag nanoparticles. In addition, Pd presents a larger
absorption contribution to the extinction and higher refractive index susceptibility.
However, despite the knowledge gained in recent years, our understanding of the
behavior of Pd as a plasmonic material should be further improved. For instance,
new synthetic routes to fabricate Pd nanoparticles with well-defined optical features

Nanoscale Accepted Manuscript


are still necessary. This in turn will contribute to the analysis of the critical issues
affecting the application of Pd as a plasmonic material. Interestingly, throughout this
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

review, one realizes that Pd nanostructures have not been fully exploited in
plasmonics.
Herein we have summarized relevant progress in the application of plasmonic
Pd nanoparticles in various fields. Among them, the potential of plasmonic palladium
in photothermal therapy is uncovered. Regarding this application, there are still key
issues that deserve deeper investigations such as: (a) toxicity, pharmacokinetics
and biodistribution, clearance mechanism, interactions with biological systems, etc.
and (b) others related with the optical response (optimized absorption contribution)
of the nanoparticles.
In the field of catalysis, Pd is an outstanding material for performing chemical
reactions. Remarkably, its potential as plasmonic nanocatalyst remains partially
explored. Usually, the photocatalyst is a nanocomposite constituted by Pd
nanoparticles (the catalyst) combined with plasmonic Au and Ag nanostructures (the
nanoantenna that enhances the catalytic performance of Pd).
Recently, the use of Pd as the sole plasmonic material in the nanocatalyst has
been investigated. In this regard, it should be noted that often the optical properties
of Pd nanostructures have not been optimized for photocatalysis.
As shown in section 2, Pd (pure or combined with Au) shows better refractive
index sensitivity than Au and Ag, although lower FOM. Nevertheless, its application
for LSPR sensing is mostly limited to the detection of hydrogen. Pd is the most
commonly employed active material in hydrogen sensing which could be explained
owing to the capacity of this metal to absorb hydrogen molecules, and its importance
is increasing due to the high increasing expectations of H2 as a future green fuel.
Thus, there is still a lot of room for improvement in this field, and there are several
Nanoscale Page 36 of 42

aspects that have not been studied in detail. For instance, a thorough analysis of
View Article Online
the effect of particle size and shape on the sensitivity of direct H2 sensors is still
DOI: 10.1039/D0NR06270G

needed. Another critical issue to address is that of the poisoning of Pd which impairs
the performance of the sensor. In the case of indirect hydrogen sensing, it would be
interesting to investigate the use of Pd nanoparticles with well-defined plasmonic
features. Upon excitation, Pd could strongly couple with the near electric field of the
plasmonic particle of the sensor. It could contribute to fabricating sensors with better
selectivity to hydrogen, improved sensitivity and stability, or shorten response time.
In the analysis of the plasmonic properties, we showed that the absorption

Nanoscale Accepted Manuscript


contribution is typically the most important contribution to extinction in Pd
nanoparticles. This does not foresee important progress for this metal in SERS
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

sensing, nevertheless Pd based platforms could be suitable for time-resolved SERS


analysis of catalytic processes.
In summary, the expected advancements in the coming years will definitely
contribute to further improve our understanding of plasmonic Pd. Although its
application in plasmonics is still in its infancy and needs more exhaustive
investigations, we envision a bright future for this noble metal.

Conflicts of interest
There are no conflicts to declare.

Acknowledgment
I.P.-S. and J.P.-J. acknowledge financial support from the Spanish State Research
Agency Grant No. MAT2016-77809-R and PID2019-108954RB-I00), Xunta de
Galicia/FEDER (grant GRC ED431C 2016-048) and the program Interreg V-A
España-Portugal (POCTEP) 2014–2020 and the European Union (European
Regional Development Fund-ERDF) (grant: FOODSENS). We thank Carlos
Fernández-Lodeiro for his contribution in Boundary Element Method (BEM)
simulations and Daniel García-Lojo for his contribution in the TOC desing.

References

1 W. H. Wollaston, Philos. Trans. Royal Soc., 1804, 94, 419-430.


2 P. Taladriz-Blanco, P. Herves and J. Perez-Juste, Top. Catal., 2013, 56, 1154-1170.
Page 37 of 42 Nanoscale

3 M. Perez-Lorenzo, J. Phys. Chem. Lett., 2012, 3, 167-174.


4 Y. H. Chen, H. H. Hung and M. H. Huang, J. Am. Chem. Soc., 2009, 131, DOI: 9114-9121.View Article Online
10.1039/D0NR06270G
5 A. C. Chen and C. Ostrom, Chem. Rev., 2015, 115, 11999-12044.
6 K. Hong, M. Sajjadi, J. M. Suh, K. Zhang, M. Nasrollahzadeh, H. W. Jang, R. S. Varma
and M. Shokouhimehr, ACS Appl. Nano Mater., 2020, 3, 2070-2103.
7 A. R. Kapdi and D. Prajapati, RSC Adv., 2014, 4, 41245-41259.
8 I. P. Beletskaya and A. V. Cheprakov, Chem. Rev., 2000, 100, 3009-3066.
9 S. K. Konda and A. C. Chen, Mater. Today, 2016, 19, 100-108.
10 B. D. Adams and A. C. Chen, Mater. Today, 2011, 14, 282-289.
11 W. Niu, L. Zhang and G. Xu, ACS Nano, 2010, 4, 1987-1996.
12 H. Zhang, M. Jin, Y. Xiong, B. Lim and Y. Xia, Acc. Chem. Res., 2013, 46, 1783-1794.
13 http://www.platinum.matthey.com/, Accessed August 30, 2020.
14 P. R. West, S. Ishii, G. V. Naik, N. K. Emani, V. M. Shalaev and A. Boltasseva, Laser

Nanoscale Accepted Manuscript


Photonics Rev., 2010, 4, 795-808.
15 K. Sugawa, H. Tahara, A. Yamashita, J. Otsuki, T. Sagara, T. Harumoto and S.
Yanagida, ACS Nano, 2015, 9, 1895-1904.
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

16 X. Q. Huang, S. H. Tang, X. L. Mu, Y. Dai, G. X. Chen, Z. Y. Zhou, F. X. Ruan, Z. L. Yang


and N. F. Zheng, Nat. Nanotechnol., 2011, 6, 28-32.
17 X. Q. Huang, S. H. Tang, B. J. Liu, B. Ren and N. F. Zheng, Adv. Mater., 2011, 23, 3420-
3425.
18 P. B. Johnson and R. W. Christy, Phys. Rev. B, 1974, 9, 5056-5070.
19 P. B. Johnson and R. W. Christy, Phys. Rev. B, 1972, 6, 4370-4379.
20 Modern Plasmonics, 1st edn., Elsevier Science, 2014.
21 Y. J. Xiong, J. Y. Chen, B. Wiley, Y. N. Xia, S. Aloni and Y. D. Yin, J. Am. Chem. Soc.,
2005, 127, 7332-7333.
22 Y. J. Xiong, J. Y. Chen, B. Wiley, Y. N. Xia, Y. D. Yin and Z. Y. Li, Nano Lett., 2005, 5,
1237-1242.
23 F. J. G. de Abajo and A. Howie, Phys. Rev. Lett., 1998, 80, 5180-5183.
24 F. J. G. de Abajo and A. Howie, Phys. Rev. B, 2002, 65, 115418.
25 S. A. Maier, Plasmonics: Fundamentals and Applications, Springer (US), US, 2007.
26 V. Giannini, A. I. Fernandez-Dominguez, S. C. Heck and S. A. Maier, Chem. Rev., 2011,
111, 3888-3912.
27 E. C. Le Ru and P. G. Etchegoin, Principles of Surface-Enhanced Raman Spectroscopy
Elsevier, Amsterdam, 2009.
28 C. Langhammer, Z. Yuan, I. Zorić and B. Kasemo, Nano Lett., 2006, 6, 833-838.
29 Y. M. Xiong, Y. Y. Zhang, P. F. Rong, J. Yang, W. Wang and D. B. Liu, Nanoscale, 2015,
7, 15584-15588.
30 X. Liu, S. Y. Zhang, P. L. Tan, J. Zhou, Y. Huang, Z. Nie and S. Z. Yao, Chem. Commun.,
2013, 49, 1856-1858.
31 L. L. Yu, Z. R. Song, J. Peng, M. L. Yang, H. Zhi and H. He, Trends Anal. Chem., 2020,
127, 115880(115881-115818).
32 L. Saa, M. Coronado-Puchau, V. Pavlov and L. M. Liz-Marzan, Nanoscale, 2014, 6,
7405-7409.
33 S. Rodal-Cedeira, V. Montes-Garcia, L. Polavarapu, D. M. Solis, H. Heidari, A. La
Porta, M. Angiola, A. Martucci, J. M. Taboada, F. Obelleiro, S. Bals, J. Perez-Juste and I.
Pastoriza-Santos, Chem. Mater., 2016, 28, 9169-9180.
34 C.-L. Lu, K. S. Prasad, H.-L. Wu, J.-a. A. Ho and M. H. Huang, J. Am. Chem. Soc., 2010,
132, 14546-14553.
Nanoscale Page 38 of 42

35 K. Sugawa, D. Sugimoto, H. Tahara, T. Eguchi, M. Katoh, K. Uchida, S. Jin, T. Ube, T.


Ishiguro and J. Otsuki, Opt. Mater. Express, 2016, 6, 859-867. View Article Online
DOI: 10.1039/D0NR06270G
36 C. Langhammer, B. Kasemo and I. Zoric, J. Chem. Phys., 2007, 126, 194702.
37 H. Ehrenreich and H. R. Philipp, Phys. Rev., 1962, 128, 1622-1629.
38 T. Pakizeh, C. Langhammer, I. Zoric, P. Apell and M. Kall, Nano Lett., 2009, 9, 882-
886.
39 G. Baffou and R. Quidant, Laser Photonics Rev., 2013, 7, 171-187.
40 G. Baffou, P. Berto, E. B. Urena, R. Quidant, S. Monneret, J. Polleux and H. Rigneault,
ACS Nano, 2013, 7, 6478-6488.
41 G. Albrecht, M. Ubl, S. Kaiser, H. Giessen and M. Hentschel, Acs Photonics, 2018, 5,
1058-1067.
42 M. Quintanilla, C. Kuttner, J. D. Smith, A. Seifert, S. E. Skrabalak and L. M. Liz-Marzan,
Nanoscale, 2019, 11, 19561-19570.

Nanoscale Accepted Manuscript


43 J. Perez-Juste, I. Pastoriza-Santos, L. M. Liz-Marzan and P. Mulvaney, Coord. Chem.
Rev., 2005, 249, 1870-1901.
44 Y. Xiong, J. M. McLellan, J. Chen, Y. Yin, Z.-Y. Li and Y. Xia, J. Am. Chem. Soc., 2005,
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

127, 17118-17127.
45 Y. Xiong, B. Wiley, J. Chen, Z.-Y. Li, Y. Yin and Y. Xia, Angew. Chem. Int. Ed., 2005, 44,
7913-7917.
46 Z. B. Shao, W. Zhu, H. Wang, Q. B. Yang, S. L. Yang, X. D. Liu and G. Z. Wang, J. Phys.
Chem. C, 2013, 117, 14289-14294.
47 W. Niu, W. Zhang, S. Firdoz and X. Lu, Chem. Mater., 2014, 26, 2180-2186.
48 W. X. Niu, Z. Y. Li, L. H. Shi, X. Q. Liu, H. J. Li, S. Han, J. Chen and G. B. Xu, Cryst.
Growth Des., 2008, 8, 4440-4444.
49 H. M. Song and J. I. Zink, Langmuir, 2018, 34, 4271-4281.
50 Y. G. Gil, S. Kang, A. Chae, Y. K. Kim, D. H. Min and H. Jang, Nanoscale, 2018, 10,
19810-19817.
51 S. H. Tang, M. Chen and N. F. Zheng, Small, 2014, 10, 3139-3144.
52 S. Jung, K. L. Shuford and S. Park, J. Phys. Chem. C, 2011, 115, 19049-19053.
53 J. H. Song, F. Kim, D. Kim and P. Yang, Chem. Eur. J., 2005, 11, 910-916.
54 Q. Li, R. Jiang, T. Ming, C. Fang and J. Wang, Nanoscale, 2012, 4, 7070-7077.
55 H. Chen, F. Wang, K. Li, K. C. Woo, J. Wang, Q. Li, L.-D. Sun, X. Zhang, H.-Q. Lin and
C.-H. Yan, ACS Nano, 2012, 6, 7162-7171.
56 F. Wang, L.-D. Sun, W. Feng, H. Chen, M. H. Yeung, J. Wang and C.-H. Yan, Small,
2010, 6, 2566-2575.
57 G. X. Su, H. Q. Jiang, H. Y. Zhu, J. J. Lv, G. H. Yang, B. Yan and J. J. Zhu, Nanoscale,
2017, 9, 12494-12502.
58 Y. Xiang, X. Wu, D. Liu, X. Jiang, W. Chu, Z. Li, Y. Ma, W. Zhou and S. Xie, Nano Lett.,
2006, 6, 2290-2294.
59 R. Jiang, F. Qin, Q. Ruan, J. Wang and C. Jin, Adv. Funct. Mater., 2014, 24, 7328-7337.
60 C. J. DeSantis, A. A. Peverly, D. G. Peters and S. E. Skrabalak, Nano Letters, 2011, 11,
2164-2168.
61 L. H. Lu, H. S. Wang, S. Q. Xi and H. J. Zhang, J. Mater. Chem., 2002, 12, 156-158.
62 J. W. Hu, Y. Zhang, J. F. Li, Z. Liu, B. Ren, S. G. Sun, Z. Q. Tian and T. Lian, Chem. Phys.
Lett., 2005, 408, 354-359.
63 H. J. Chen, G. Wei, A. Ispas, S. G. Hickey and A. Eychmuller, J. Phys. Chem. C, 2010,
114, 21976-21981.
64 C. Fernández-Lodeiro, J. Fernández-Lodeiro, E. Carbó-Argibay, C. Lodeiro, J. Pérez-
Juste and I. Pastoriza-Santos, Nano Res., 2020, 13, 2351-2355.
Page 39 of 42 Nanoscale

65 C. J. DeSantis, A. C. Sue, M. M. Bower and S. E. Skrabalak, ACS Nano, 2012, 6, 2617-


2628. View Article Online
DOI: 10.1039/D0NR06270G
66 C. J. DeSantis and S. E. Skrabalak, Langmuir, 2012, 28, 9055-9062.
67 J. Chen, B. Wiley, J. McLellan, Y. Xiong, Z.-Y. Li and Y. Xia, Nano Lett., 2005, 5, 2058-
2062.
68 A. Dumas and P. Couvreur, Chem. Sci., 2015, 6, 2153-2157.
69 T. T. V. Phan, T. C. Huynh, P. Manivasagan, S. Mondal and J. Oh, Nanomaterials,
2019, 10, 66.
70 A. Schneemann, J. L. White, S. Kang, S. Jeong, L. W. F. Wan, E. S. Cho, T. W. Heo, D.
Prendergast, J. J. Urban, B. C. Wood, M. D. Allendorf and V. Stavila, Chem. Rev., 2018, 118,
10775-10839.
71 C. C. Ndaya, N. Javahiraly and A. Brioude, Sensors, 2019, 19, 4478.
72 S. Link and M. A. El-Sayed, Int. Rev. Phys. Chem., 2000, 19, 409-453.

Nanoscale Accepted Manuscript


73 M. Kim, J. H. Lee and J. M. Nam, Adv. Sci., 2019, 6, 1900471.
74 A. Becker, C. Hessenius, K. Licha, B. Ebert, U. Sukowski, W. Semmler, B.
Wiedenmann and C. Grotzinger, Nat. Biotechnol., 2001, 19, 327-331.
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

75 A. R. Rastinehad, H. Anastos, E. Wajswol, J. S. Winoker, J. P. Sfakianos, S. K.


Doppalapudi, M. R. Carrick, C. J. Knauer, B. Taouli, S. C. Lewis, A. K. Tewari, J. A. Schwartz,
S. E. Canfield, A. K. George, J. L. West and N. J. Halas, Proc. Natl. Acad. Sci. U.S.A, 2019, 116,
18590-18596.
76 Y. Wu, M. R. K. Ali, B. Dong, T. G. Han, K. C. Chen, J. Chen, Y. Tang, N. Fang, F. J. Wang
and M. A. El-Sayed, ACS Nano, 2018, 12, 9279-9290.
77 X. Huang, S. Tang, X. Mu, Y. Dai, G. Chen, Z. Zhou, F. Ruan, Z. Yang and N. Zheng,
Nat. Nanotechnol., 2011, 6, 28-32.
78 M. Chen, S. Z. Chen, C. Y. He, S. G. Mo, X. Y. Wang, G. Liu and N. F. Zheng, Nano Res.,
2017, 10, 1234-1248.
79 S. G. Shi, Y. Z. Huang, X. L. Chen, J. Weng and N. F. Zheng, ACS Appl. Mater.
Interfaces, 2015, 7, 14369-14375.
80 S. H. Tang, X. Q. Huang and N. F. Zheng, Chem. Commun., 2011, 47, 3948-3950.
81 Y. Liu, L. Ding, D. Wang, M. Lin, H. Sun, H. Zhang, H. Sun and B. Yang, ACS Appl. Bio
Mater., 2018, 1, 1102-1108.
82 M. Chen, S. H. Tang, Z. D. Guo, X. Y. Wang, S. G. Mo, X. Q. Huang, G. Liu and N. F.
Zheng, Adv. Mater., 2014, 26, 8210-8216.
83 S. Bharathiraja, N. Q. Bui, P. Manivasagan, M. S. Moorthy, S. Mondal, H. Seo, N. T.
Phuoc, T. T. V. Phan, H. Kim, K. D. Lee and J. Oh, Sci. Rep., 2018, 8, 500.
84 G. Gao, Y. W. Jiang, H. R. Jia, W. Sun, Y. Guo, X. W. Yu, X. Liu and F. G. Wu,
Biomaterials, 2019, 223, 119443.
85 Z. J. Zhou, Y. T. Wang, Y. Yan, Q. Zhang and Y. Y. Cheng, ACS Nano, 2016, 10, 4863-
4872.
86 S. G. Mo, X. L. Chen, M. Chen, C. Y. He, Y. H. Lu and N. F. Zheng, J. Mater. Chem. B,
2015, 3, 6255-6260.
87 M. X. Yu and J. Zheng, ACS Nano, 2015, 9, 6655-6674.
88 X. G. Zhang, G. Cheng, X. Xing, J. C. Liu, Y. Cheng, T. Y. Ye, Q. Wang, X. H. Xiao, Z. B.
Li and H. B. Deng, J. Phys. Chem. Lett., 2019, 10, 4185-4191.
89 W. J. Fang, S. H. Tang, P. X. Liu, X. L. Fang, J. W. Gong and N. F. Zheng, Small, 2012,
8, 3816-3822.
90 X. Yang, L. L. Li, D. G. He, L. Hai, J. L. Tang, H. F. Li, X. X. He and K. M. Wang, J. Mater.
Chem. B, 2017, 5, 4648-4659.
Nanoscale Page 40 of 42

91 C. C. Wang, Y. Q. Li, X. Q. Shi, J. H. Zhou, L. Zhou and S. H. Wei, Chem. Commun.,


2018, 54, 13403-13406. View Article Online
DOI: 10.1039/D0NR06270G
92 Y. J. Zhang, R. Sha, L. Zhang, W. B. Zhang, P. P. Jin, W. G. Xu, J. X. Ding, J. Lin, J. Qian,
G. Y. Yao, R. Zhang, F. C. Luo, J. Zeng, J. Cao and L. P. Wen, Nat. Commun., 2018, 9, 4236.
93 Y. P. Liu, X. W. Zhang, L. Y. Luo, L. Li, Y. C. He, J. An and D. W. Gao, ACS Biomater. Sci.
Eng., 2018, 4, 2911-2921.
94 H. T. Nguyen, Z. C. Soe, K. Y. Yang, C. D. Phung, L. T. Nguyen, J. H. Jeong, S. G. Jin, H.
G. Choi, S. K. Ku, C. S. Yong and J. O. Kim, Colloids Surf. B: Biointerfaces, 2019, 176, 265-275.
95 S. Kang, W. Shin, K. Kang, M. H. Choi, Y. J. Kim, Y. K. Kim, D. H. Min and H. Jang, ACS
Appl. Mater. Interfaces, 2018, 10, 13819-13828.
96 S. Wang, J. Lin, T. F. Wang, X. Y. Chen and P. Huang, Theranostics, 2016, 6, 2394-
2413.
97 C. Moore and J. V. Jokerst, Theranostics, 2019, 9, 1550-1571.

Nanoscale Accepted Manuscript


98 L. M. Nie, M. Chen, X. L. Sun, P. F. Rong, N. F. Zheng and X. Y. Chen, Nanoscale, 2014,
6, 1271-1276.
99 P. H. Zhao, Z. K. Jin, Q. Chen, T. Yang, D. Y. Chen, J. Meng, X. F. Lu, Z. Gu and Q. J. He,
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Nat. Commun., 2018, 9, 4241.


100 M. Chen, Z. D. Guo, Q. H. Chen, J. P. Wei, J. C. Li, C. R. Shi, D. Xu, D. W. Zhou, X. Z.
Zhang and N. F. Zheng, Chem. Sci., 2018, 9, 4268-4274.
101 B. Rubio-Ruiz, A. M. Perez-Lopez, T. L. Bray, M. Lee, A. Serrels, M. Prieto, M.
Arruebo, N. O. Carragher, V. Sebastian and A. Unciti-Broceta, ACS Appl. Mater. Interfaces,
2018, 10, 3341-3348.
102 Y. Zhang, R. Guo, D. Wang, X. Sun and Z. Xu, Colloids Surf. B: Biointerfaces, 2019,
176, 300-308.
103 M. M. Miller and A. A. Lazarides, J. Phys. Chem. B, 2005, 109, 21556-21565.
104 S. J. Zalyubovskiy, M. Bogdanova, A. Deinega, Y. Lozovik, A. D. Pris, K. H. An, W. P.
Hall and R. A. Potyrailo, J Opt Soc Am A, 2012, 29, 994-1002.
105 X. Z. Zhu, H. L. Jia, X. M. Zhu, S. Cheng, X. L. Zhuo, F. Qin, Z. Yang and J. F. Wang, Adv.
Funct. Mater., 2017, 27, 1700016.
106 A. F. Smith, S. M. Harvey, S. E. Skrabalak and R. G. Weiner, Nanoscale, 2016, 8,
16841-16845.
107 A. F. Smith, R. G. Weiner, M. M. Bower, B. Dragnea and S. E. Skrabalak, J. Phys.
Chem. C, 2015, 119, 22114-22121.
108 C. Wadell, S. Syrenova and C. Langhammer, ACS Nano, 2014, 8, 11925-11940.
109 V. M. Silkin, R. D. Muino, I. P. Chernov, E. V. Chulkov and P. M. Echenique, J. Phys.
Condens. Matter, 2012, 24, 104021.
110 C. Langhammer, I. Zoric and B. Kasemo, Nano Lett., 2007, 7, 3122-3127.
111 I. Zoric, E. M. Larsson, B. Kasemo and C. Langhammer, Adv. Mater., 2010, 22, 4628-
4633.
112 E. M. Larsson, M. E. M. Edvardsson, C. Langhammer, I. Zoric and B. Kasemo, Rev. Sci.
Instrum., 2009, 80.
113 N. Strohfeldt, A. Tittl and H. Giessen, Opt. Mater. Express, 2013, 3, 194-204.
114 C. Langhammer, E. M. Larsson, V. P. Zhdanov and I. Zoric, J. Phys. Chem. C, 2012,
116, 21201-21207.
115 W. X. Niu, W. Q. Zhang, S. Firdoz and X. M. Lu, Chem. Mater., 2014, 26, 2180-2186.
116 R. B. Jiang, F. Qin, Q. F. Ruan, J. F. Wang and C. J. Jin, Adv. Funct. Mater., 2014, 24,
7328-7337.
117 C. Y. Chiu and M. H. Huang, Angew. Chem. Int. Ed., 2013, 52, 12709-12713.
Page 41 of 42 Nanoscale

118 M. E. Nasir, W. Dickson, G. A. Wurtz, W. P. Wardley and A. V. Zayats, Adv. Mater.,


2014, 26, 3532-3537. View Article Online
DOI: 10.1039/D0NR06270G
119 Q. H. Yang, Q. Xu, S. H. Yu and H. L. Jiang, Angew. Chem. Int. Ed., 2016, 55, 3685-
3689.
120 A. M. Lacerda, I. Larrosa and S. Dunn, Nanoscale, 2015, 7, 12331-12335.
121 I. Sarhid, I. Abdellah, C. Martini, V. Hu, D. Dragoe, P. Beaunier, I. Lampre and H.
Remita, New J. Chem., 2019, 43, 4349-4355.
122 J. S. Zou, Z. C. Si, Y. D. Cao, R. Ran, X. D. Wu and D. Weng, J. Phys. Chem. C, 2016,
120, 29116-29125.
123 Y. Z. Zhu, Z. X. Xu, W. Y. Jiang, S. X. Zhong, L. H. Zhao and S. Bai, J. Mater. Chem. A,
2017, 5, 2619-2628.
124 D. F. Swearer, H. Q. Zhao, L. N. Zhou, C. Zhang, H. Robatjazi, J. M. P. Martirez, C. M.
Krauter, S. Yazdi, M. J. McClain, E. Ringe, E. A. Carter, P. Nordlander and N. J. Halas, Proc.

Nanoscale Accepted Manuscript


Natl. Acad. Sci. U.S.A, 2016, 113, 8916-8920.
125 Y. J. Mu, H. Y. Zhang, W. T. Zheng and X. Q. Cui, New J. Chem., 2017, 41, 786-792.
126 J. Guo, Y. Zhang, L. Shi, Y. F. Zhu, M. F. Mideksa, K. Hou, W. S. Zhao, D. W. Wang, M.
Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

T. Zhao, X. F. Zhang, J. W. Lv, J. Q. Zhang, X. L. Wang and Z. Y. Tang, J. Am. Chem. Soc., 2017,
139, 17964-17972.
127 M. Wen, K. Mori, Y. Kuwahara and H. Yamashita, ACS Energy Lett., 2017, 2, 1-7.
128 Z. K. Zheng, T. Tachikawa and T. Majima, J. Am. Chem. Soc., 2015, 137, 948-957.
129 F. Wang, C. H. Li, H. J. Chen, R. B. Jiang, L. D. Sun, Q. Li, J. F. Wang, J. C. Yu and C. H.
Yan, J. Am. Chem. Soc., 2013, 135, 5588-5601.
130 S. Kundu, S. I. Yi, L. Ma, Y. Y. Chen, W. Dai, A. M. Sinyukov and H. Liang, Dalton Trans.,
2017, 46, 9678-9691.
131 P. Kannan, J. Dolinska, T. Maiyalagan and M. Opallo, Nanoscale, 2014, 6, 11169-
11176.
132 J. Langer, D. J. de Aberasturi, J. Aizpurua, R. A. Alvarez-Puebla, B. Auguie, J. J.
Baumberg, G. C. Bazan, S. E. J. Bell, A. Boisen, A. G. Brolo, J. Choo, D. Cialla-May, V. Deckert,
L. Fabris, K. Faulds, F. J. G. de Abajo, R. Goodacre, D. Graham, A. J. Haes, C. L. Haynes, C.
Huck, T. Itoh, M. Ka, J. Kneipp, N. A. Kotov, H. Kuang, E. C. Le Ru, H. K. Lee, J. F. Li, X. Y. Ling,
S. A. Maier, T. Mayerhofer, M. Moskovits, K. Murakoshi, J. M. Nam, S. Nie, Y. Ozaki, I.
Pastoriza-Santos, J. Perez-Juste, J. Popp, A. Pucci, S. Reich, B. Ren, G. C. Schatz, T. Shegai, S.
Schlucker, L. L. Tay, K. G. Thomas, Z. Q. Tian, R. P. Van Duyne, T. Vo-Dinh, Y. Wang, K. A.
Willets, C. Xu, H. Xu, Y. Xu, Y. S. Yamamoto, B. Zhao and L. M. Liz-Marzan, ACS Nano, 2020,
14, 28-117.
133 T. Bhuvana and G. U. Kulkarni, Small, 2008, 4, 670-676.
134 L. M. Chen and Y. N. Liu, Crystengcomm, 2011, 13, 6481-6487.
135 J. F. Huang, Y. H. Zhu, M. Lin, Q. X. Wang, L. Zhao, Y. Yang, K. X. Yao and Y. Han, J.
Am. Chem. Soc., 2013, 135, 8552-8561.
Nanoscale Page 42 of 42

View Article Online


DOI: 10.1039/D0NR06270G

Nanoscale Accepted Manuscript


Published on 26 October 2020. Downloaded on 11/4/2020 5:33:09 AM.

Pd as alternative material for nanoplasmonics

You might also like