Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

Accepted Manuscript

Fabrication and Application of Substoichiometric Tungsten Oxide with Tunable


Localized Surface Plasmon Resonances

Jun Chen, Yumei Ren, Tianzhao Hu, Tao Xu, Qun Xu

PII: S0169-4332(18)32556-X
DOI: https://doi.org/10.1016/j.apsusc.2018.09.140
Reference: APSUSC 40444

To appear in: Applied Surface Science

Received Date: 6 June 2018


Revised Date: 7 September 2018
Accepted Date: 16 September 2018

Please cite this article as: J. Chen, Y. Ren, T. Hu, T. Xu, Q. Xu, Fabrication and Application of Substoichiometric
Tungsten Oxide with Tunable Localized Surface Plasmon Resonances, Applied Surface Science (2018), doi: https://
doi.org/10.1016/j.apsusc.2018.09.140

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Fabrication and Application of Substoichiometric Tungsten
Oxide with Tunable Localized Surface Plasmon Resonances

Jun Chen, Yumei Ren, Tianzhao Hu, Tao Xu, Qun Xu*

College of Materials Science & Engineering, Zhengzhou University, Zhengzhou,

450052, P.R.China

Abstract

Substoichiometric tungsten oxide (WO3-x) has become one of the interesting

candidates for localized surface plasmon resonance (LSPR) hosts because of its

fascinating properties. Herein, we report that substoichiometric WO3-x can be

fabricated via a facile solid phase reduction method, and at the same time it exhibits a

strong LSPR in the near infrared region. Tunable plasmon resonances can be observed

in a wide range by changing their stoichiometry. In the bovine serum albumin (BSA)

based optical bio-sensing system, the plasmonic WO3-x displays superior sensitivity.

Moreover, the hybrid nanocomposites combining the plasmonic WO3-x and plasmonic

Ag can be also obtained, which show great potentials in heterogeneous catalysis.

Specially, the WO3-x prepared at the reduction temperature of 350 °C (WO3-350)

exerts improved photocatalytic performance toward methylene blue degradation,

which is due to the synergistic effect of the strong plasmonic effect of WO3-350 and

Ag nanoparticles.

1
Keywords: Substoichiometric tungsten oxide, biosensors, Localized surface plasmon

resonance, photocatalysis

1. Introduction

The interactions between light and matter are of interest due to their fundamental

role in numerous physical processes as well as their potential applications. The field of

plasmonics has undergone tremendous growth in recent years, where coherent

oscillation of the surface free charge carriers can be excited by incident light leading to

localized surface plasmon resonance (LSPR) [1]. This phenomenon has great potential

in a variety of promising applications such as biosensing [2], surface-enhanced Raman

spectroscopy [3], catalysis [4] and high-efficiency photothermal conversion [5-7]. The

vast majority of research about LSPR has been performed on noble metal nanoparticles

(NPs), which have high free carrier concentration and show strong and broad LSPR

absorption in the visible light region [8]. However, hot electrons generated in these

NPs will be cooled by ultrafast electron-phonon scattering [9], resulting in a seriously

reduced photocatalytic efficiency. In addition, the static form of LSPR noble metal also

limits their applications in the field of optical switching.

Recently, heavily doped semiconductors have been reported due to their tunable

LSPRs, which mainly involve copper chalcogenides [10, 11], transition metal nitrides

[12] and transition metal oxides [13, 14]. Among them, transition metal oxide (TMOs),

such as TiO2-x, MoO3-x and WO3-x, can exhibit plasmon absorption in the visible and

near-infrared region by the introduction of oxygen vacancies [15, 16]. Importantly, the
2
LSPR spectra can be tuned by changing the concentration of free carriers. Compared

to noble metals, the carrier concentration in TMOs is lower so that their LSPR is more

sensitive to the variation of carrier concentration. Moreover, TMOs can serve as a

promising LSPR host owning to their characteristics of outer-d valence electron [17].

Among various plasmonic oxides, plasmonic WO3-x is of particular interest because of

its unique electronic structure and properties. To date, substoichiometric WO3-x has

been synthesized via various approaches. For example, solvothermal treatment of the

ethanol solution of W(CO)6 [18, 19], or utilization of WO2(O2)H2O [20] as precursors

in the presence of surfactants [21]. However, the tedious post-treatment process limits

its further applications. In recent years, hydrogen reduction method has become an

ideal route to prepare such substoichiometric WO3-x [15, 22], and the plasmonic WO3-x

has been exploited in various application, such as photocatalysis [23, 24], gas sensors

[25] and photothermal therapy as well [26, 27].

Herein, a facile solid phase reduction method using NaBH4 as the alternative of

hydrogen was reported to prepare substoichiometric tungsten oxide (WO3-x). Our

experimental results indicate that the obtained WO3-x shows strong plasmon resonance

absorption in the near-infrared region of the spectrum, and the plasmon resonance can

be controlled in a wide range by changing the oxygen vacancy content depending on

the different reaction temperature. It indicates the plasmonic WO3-350 alone displays

excellent sensitivity in a model biosensing system. Further in order to achieve the

maximum utilization of solar light in full solar spectrum, a semiconductor-metal hybrid

that combines excitonic absorption and LSPR absorption of WO3-x and noble metal also
3
be fabricated. It demonstrates that the as-prepared Ag/WO3-350 hybrids exhibit the

most enhanced photocatalytic degradation performance for methylene blue, which is

attributed to the wide spectral absorption and the strong pasmon-plasmon coupling

effect of WO3-350 and Ag.

2. Experimental Section

2.1. Materials

WO3 powder was purchased from Sigma-Aldrich (Fluka, product number 95410).

NaBH4 was purchased from Tokyo Chemical Industry Co., Ltd (TCI). Bovine serum

albumin was provided by Beijing Aobo Star Biotechnology Co., Ltd (China).

Trisodium citrate dihydrate, silver nitrate and ethanol in analytical grade were provided

by Sinopharm Chemical Reagent Co., Ltd (China). Ascorbic acid was purchased from

Tianjin Damao Chemical Reagent Factory (China). All the materials were used as

purchased without any purification.

2.2. Characterization

The morphologies of WO3 power and prepared samples were characterized by

field-emission SEM (JEOR JSM-6700F). Transmission electron microscopy (TEM)

and high-resolution transmission electron microscopy (HRTEM) images were

collected with a JEM -2100 (JEOL, Japan). X-ray diffraction (XRD) patterns were

examined by a Bruker D8 Focus diffractometer (Bruker AXS, Germany) with Cu Kα

radiation. UV-vis-NIR spectra (UV-vis DRS, Cary 500) were used to measure the light

adsorption of samples. Raman spectra measurements were performed on LabRAM HR


4
Evolution using laser wavelength of 532 nm. The electron paramagnetic resonance

spectroscopy (EPR) was carried out using an Endor spectrometer (JEOL ES-ED3X) at

room temperature. The photoluminescence (PL) spectroscopy excited with 365 nm

laser light was collected to disclose the efficiency of the separation and recombination

of charge carriers. X-ray photoelectron spectroscopy (XPS) analysis was conducted on

ESCLAB 280 using electrostatic lens mode with pass energy of 100 eV. All binding

energies were calibrated using contaminated carbon (C1s = 285 eV) as a reference.

2.3. Fabrication of substoichiometric WO3-x

0.3 g WO3 and 0.3 g NaBH4 were mixed and ground thoroughly for 20 min at room

temperature. Then the mixture was transferred into a pair of porcelain boat. And the

boats were placed in a tubular furnace which was heated form room temperature to

200-400 °C under an Argon atmosphere with a heating rate of 5 °C/min. The heating

temperature is maintained at the predetermined temperature for 60 minutes. Then the

temperature is naturally cooled to room temperature, and the sample was taken out and

washed for several times with ethanol and deionized water, for the purpose to remove

unreacted NaBH4. Subsequently they are dried in the oven at 60 °C, and a series of

different colors of WO3-x can be obtained by controlling the reaction temperature.

2.4. BSA Preparation and test

To evaluate the bio-sensing performance of the substoichiometric WO3-x, 2 mL

BSA solutions in the various concentration of 0, 0.5, 2, 5, 10, 20, 25, 30 and 35

mg·mL-1 were incubated with 1 mL dispersions of the substoichiometric WO 3-x for 60

min, and then their UV-vis-NIR spectra was measured.


5
2.5. Preparation of Ag/WO3-x hybrids

Here we adopt the method in a previous report [28]. WO3-x sample (36 mg) was

dispersed in deionized water (7.5 mL) and then sonicated for 1 h. Trisodium citrate

dihydrate solution (1 mL, 0.039 M) and silver nitrate solution (1 mL, 0.01 M) were

sequentially added to the dispersion of WO3-x sample with stirring constantly. Then a

solution of ascorbic acid (50 μL, 0.1 mM) was added into the above mixture. After

continuous stirring for 1 h, the obtained precipitate by centrifugation was washed

twice with deionized water and ethanol. The sample is then dried in a vacuum oven at

60 °C.

2.6. Photoelectrochemical (PEC) Measurements

PEC measurements were performed by an electrochemical workstation

(CHI660D) with a three-electrode cell. The as-prepared samples were used as the

working electrode, and a Pt wire and Ag/AgCl electrode were used as counter electrode

and reference electrode, respectively. 0.5 M Na2SO4 was used as the electrolyte. The

light ON-OFF switches were set as 80 s when measuring the I-t curves of the absolute

values under visible light. The bias for the measurement was set as 0.6 V

2.7. Photocatalytic Degradation of Methylene Blue

The photocatalytic activity of Ag/WO3-x hybrids for the degradation of methylene

blue was evaluated by using a 350 W xenon lamp with a 420 nm cutoff filter as the light

source at room temperature. Near infrared light (λ ≥ 800 nm) was obtained from a 350

W xenon lamp with a 800 nm cutoff filter. In a typical process, 12 mg of sample was

added into 50 ml solution containing 20 mg·L-1 methylene blue. Then the suspension
6
was stirred in dark for 30 min to achieve adsorption/desorption equilibrium. The

suspension was exposed to visible light irradiation under constant stirring. At fixed

intervals, 4 mL suspension was collected and centrifuged to remove solid particles, and

the supernatant was analyzed by UV-Vis spectroscopic measurements.

3. Results and discussion

3.1. Morphology and structure characterization of WO3-x

Substoichiometric WO3-x samples were prepared by the reduction of the original

tungsten oxide with NaBH4 at different temperatures, marked as WO3-T (e.g.,

WO3-350 refers to the product obtained by annealing at 350 °C). As shown in Fig. S1,

the morphology and size of the obtained sample did not change significantly compared

to the original WO3. We have also performed transmission electron microscopy

(TEM) and high-resolution transmission electron microscopy (HRTEM) analysis on

the pristine WO3 as well as the samples with the varied annealing temperature. Clear

lattice fringes can be observed in the HRTEM image of the original WO3 (Fig. S2),

revealing its high crystallinity. As shown in the corresponding fast fourier transform

(FFT) image, (200) and (020) crystal plane with angle of 90° can be clearly observed,

corresponding well to the monoclinic phase structure of WO3. Fig. 1a and b show the

images of samples calcined at 250 °C and 350 °C, respectively. As shown in the FFT

images (Fig. 1a and Fig. 1e, inset), the interior of the samples still retains high

crystallinity. While disordered regions appear on the edge of the samples, and the

disordered area increased with the increase in treatment temperature. When the
7
annealing temperature is further raised up to 400 °C, large-area disorder regions can

be obtained (Fig. S3). Based on the above observations, we suggest that the

hydrogenation of samples first happens on the surface and edge, and then it gradually

extends to the interior with the rise of the temperature.

Fig. 1. TEM images of WO3-250 and WO3-350 samples. a) HRTEM image of


WO3-250 sample. Inset: filtered images of the region enclosed by the white square
and the corresponding FFT image. b) HRTEM image of WO3-350 sample. c)
Magnified image of the region enclosed by the red square in (a). d) FFT image of the
region enclosed by the red square for WO3-350 sample. e) Filtered images enclosed by
the white square in (b). Inset: the corresponding FFT image.

Fig. 2a shows the UV-vis-NIR spectra and photographs of WO3 and

8
NaBH4-treated WO3 samples prepared at different temperature (200-400 °C). As

depicted in Fig. 2a (inset), the pristine WO3 is yellowish in color. As the annealing

temperature increases, the color of pristine WO3 gradually turns blue, signifying the

enhanced light absorption. When the annealing temperature increases to 350 °C, the

sample shows the deepest blue color, while the blue color becomes faint as the

processing temperature is up to 400 °C. The UV-vis-NIR spectra displays that the

most absorbed light of the pristine WO3 sample are UV and blue light thence it

appears yellowish in color. After annealing possess, the samples possess enhanced

absorption of red and orange light in visible region (600-760 nm). With the gradual

increase of temperature (up to 350 °C), the sample absorb more light in the vision

region, which corresponds well to the color change from yellowish to light blue then

navy blue. In the near-infrared region, samples prepared above 250 °C show

significantly enhanced light absorption.

The change of light absorption suggests possible change in crystal structure

and/or phase transition. Next we performed XRD analysis on WO 3 and NaBH4-treated

WO3 samples prepared at different temperature (Fig. S4). Diffraction peaks of pristine

WO3 can be well indexed into monoclinic WO 3 (JCPDS 72-0677) and the clear

diffraction peaks demonstrate its high crystallinity, which is consistent with the

previous analysis of HRTEM images. When the annealing temperatures are 200 °C

and 250 °C, there are no obvious shift of the diffraction peaks between pristine WO 3

and the obtained samples. The slight decrease in intensity of diffraction peak indicates

a slight decrease in crystallinity. Then these diffraction peaks partially disappear and
9
new diffraction peaks emerge at 300 °C, which can be indexed to monoclinic WO 2.9

(JCPDS 05-0386) marked by + in Fig. 2b, indicating the successful introduction of

oxygen vacancies in sample. When annealing temperature is 350 °C, the diffraction

peaks of the sample have been substantially corresponded to that of monoclinic WO 2.9

(JCPDS 05-0386). Next as processing temperature reach 400 °C, besides the

diffraction peak of WO2.9 in the XRD spectra, another new peaks highlighted by # and

* in Fig. 2b can be indexed as orthorhombic WO3 ·H2O (JCPDS 84-0886) and h-WO3

(JCPDS 85-2460). It is well-known that a large number of oxygen vacancies existed

in the sample are beneficial to the adsorption of water molecules and subsequent the

formation of WO3·H2O structure [15]. Considering sample prepared at 400 °C

containing a large number of disordered regions or amorphous regions, it can be

suggested that the amorphous structure can undergone lattice rearrangement in the

annealing process at 400 °C, then h-WO3 can be formed [29]. To more directly

identify the oxygen vacancies in samples, a room-temperature electron paramagnetic

resonance (EPR) spectroscopy was used to characterize the samples (Fig. S5). As

compared with pristine WO3, WO3-350 exhibits a significant EPR signal at g = 1.999

that can be attributed to the electron trapping at oxygen vacancies [7]. And the signal

intensity of WO3 almost disappears on account of no oxygen vacancies in the pristine

WO3. The results above demonstrate that we have successfully introduced oxygen

vacancies into the original WO3 by treatment with NaBH4, and the concentration of

oxygen vacancies can be regulated by changing the annealing temperature.

10
Fig. 2. a) UV-Vis-NIR absorbance spectra of pristine WO3 and WO3 samples
annealed at 250, 300, 350 and 400 °C. Inset: photographs of pristine WO3 and WO3
samples annealed at different temperature. b) Enlarged XRD patterns of pristine WO3
and WO3 samples prepared at varying temperature in the region of 2 θ = 15~45°. The
diffraction peaks of WO2.9, WO3·H2O and h-WO3 are highlighted by +, # and *,
respectively. c) Raman spectra of WO3 and NaBH4 treated WO3 samples at various
temperatures. d) LSPR wavelength as a function of the refractive index of the solvent.
RI for methanol, acetonitrile, ethanol and DMF are 1.329, 1.343, 1.362 and 1.430,
respectively.

Raman spectrometer measurements on WO3 and WO3-x samples prepared at

different temperature were also performed to further investigate the revolution in

crystal structure (Fig. 2c). Three major vibration bands at 272, 716, and 805 cm-1 can

11
be found in the Raman spectrum of the original WO3, which are attributed to the

bending vibration of δ(O-W-O) and the stretching vibration of ν(W-O-W) of

monoclinic phase, respectively [30]. After annealing, the characteristic Raman peaks

of samples become wider and weaker, confirming the gradual decrease in crystallinity

and the gradual increase of disordered areas. For the sample annealing above 300 °C,

all the characteristic Raman peaks for WO3 disappear, just a wide peak appears in the

range of 700-900 cm-1, which are consistent with Raman peaks of WO2.9 [31]. Given

that the introduced oxygen vacancies are electron donors [32], they can bring a large

number of free carriers in the semiconductor. When the concentration of free carriers

is above the threshold value, the collective oscillations of free carriers can induce

LSPR [33], and the samples will have a significantly enhanced light absorption in the

near-infrared region (Fig. 2a). With the gradual increase of the annealing temperature

(up to 350 °C), the content of oxygen vacancies in the samples, i.e., the concentration

of free carriers gradually increases. Thence sample annealing at 350 °C has the

strongest LSPR absorption. While further increase experimental temperature to

400 °C, there will be more amorphous regions in the sample, which can reduce the

effective free carrier density and thus quench the plasmon resonance [34]. Besides,

the adsorbed water molecules at oxygen vacancies may also reduce the effective free

carrier concentration. As a result, the light absorption of the sample annealing at

400 °C in the near-infrared region will become weaker.

To further verify the enhanced NIR absorbance arising from LSPR phenomenon

rather than the charge transfer between the defect states, we measured the absorbance
12
spectra of sample processed at 350 °C in several solvents with different refractive

indices. In Fig. 2d, we can observe that the LSPR energy decreases with increasing

refractive index of the medium, which ensures that the NIR absorbance band is

attributed to LSPR phenomenon [14]. We further calculated the free carrier density of

WO3-x prepared at 350 °C according to the Drude model [35],

p = ,

where ωp is the bulk plasma frequency, N is the free electrons density, e is the

elementary charge, ε0 is the permittivity of free space, and me is the effective mass of

an electron. For WO3-x prepared at 350 °C, LSPR peak centered at ≈ 1150 nm can be

observed. Therefore, the free electron concentration in WO3-x prepared at 350 °C is

determined to be 2.7 × 1022 cm-3, which is consistent with the values reported in the

literature [36]. Based on the above results, it is clearly seen that substoichiometric

WO3-x with surface plasmon resonances is successfully prepared, and the LSPR can be

tuned by varying the annealing temperature.

13
Fig. 3. a) High-resolution W 4f XPS of WO3 samples prepared at 300 °C and 350 °C.
The black curves correspond to the experimental data. Each black curve can be
deconvoluted into two pairs of peaks corresponding to W6+ (red curves) and W5+
(green curves). b) Normalized O 1s XPS spectra of pristine WO3 and NaBH4-treated
WO3 samples annealed at 300 °C and 350 °C. c) XPS W 4f spectrum collected from
WO3 sample obtained at 400 °C. The black curve is the experimental data that can be
deconvoluted into three pairs of peaks corresponding to W6+ (red curves), W5+ (green
curves) and W4+ (blue curves). d) XPS O 1s spectrum collected from WO3 sample
prepared at 400 °C.

To further explain the effect of NaBH4 treatment on the chemical states of WO3,

we performed the X-ray photoelectron spectroscopy (XPS) of pristine WO3 and the

samples annealing at varying temperature. As shown in the normalized high

14
resolution W 4f spectra (Fig. S6), the W 4f core level spectrum of pristine WO3

exhibits the characteristic spin-orbit doublet with binding energy peaks at 35.7 and

37.8 eV, correlating with 4f5/2 and 4f7/2 of W6+ species at high oxidation state. For

samples annealing at 300 and 350 °C, apart from W6+, low oxidation states of W

species, i.e., W5+ (centered at 34.3 and 36.4 eV) also exist (Fig. 3a) [37]. By

calculating the corresponding area of peaks assigned, the proportion of W5+ in

samples annealing at 300 °C is 15%, and it increases to 18% in the sample annealing

at 350 °C. This result shows that the degree of reduction improves as the annealing

temperature increases, which is consistent with the previous XRD and Raman results.

Hydrogenation at 300 or 350 °C does not bring about significant changes in the O 1s

XPS (Fig. 3b), indicating that plane defects formed by substoichiometric WO3-x do

not influence oxygen coordination polyhedron for each W atom [22]. When the

annealing temperature is 400 °C, in addition to W6+ and W5+ ions, new peaks centered

at 33.7 and 35.7 eV appear in the W 4f spectra of prepared sample (Fig. 3c), which

can be assigned to W4+ ions [38]. In the O 1s XPS spectra (Fig. 3d), the peak can be

deconvoluted into two peaks. A binding energy of 530.5 eV implies the existence of

oxygen bonded to hexavalent tungsten in WO3-x [39], and the other signal at 532.4 eV

is attributed to nonstoichiometric tungsten oxides [40]. These results indicate that

NaBH4-treated WO3 sample at 400 °C has higher reduction degree.

3.2 BSA test of WO3-x

In order to explore the application of these prepared WO3-x samples, we designed

15
an optical biosensor system, wherein selecting bovine serum albumin (BSA) as a

model protein to assess the sensitivity. The BSA can be immobilized on the surface of

WO3-x because of electrostatic interaction, hydrophobic interaction and ligand

exchange between BSA and oxide surface [41]. And the electrostatic effect may affect

the carrier concentration in WO3-x, thereby changing its plasmon characteristics. Here,

WO3-x samples prepared at 350 °C were tested because of its strongest LSPR

absorption in near-infrared region. As shown in Fig. 4a, when BSA concentration is

lower than 5 mg/ml, the intensity of plasmon resonances continuously increase with

the increment of BSA concentration. While as BSA concentration is up to 10 mg/ml,

the plasmon resonance begins to decline. The relationship between response factor

and BSA concentration is shown in Fig. 4b. And this phenomenon is similar with our

previous study on molybdenum oxide [42]. It is suggested there are two factors that

determine the plasmon resonance properties. On one hand, considering BSA is

negatively charged in deionized water [43], it will repel the free electrons in WO3-x,

which leads to a decrease in free electron concentration on the surface of WO3-x, and

then weakens the intensity of the LSPR peak. On the other hand, BSA may be prone to

enrich the vacancies of amorphous regions in WO3-x, resulting in increased electron

concentration and enhanced plasmon resonance. Apparently, in this work, the former

is dominant, especially in higher concentration of BSA.

16
a

Fig. 4. a) UV-Vis-NIR absorbance spectra of WO 3-x sample prepared at 350 °C after

being incubated with different concentrations of BSA solution for 1 h. b) The

corresponding response factor as a function of the concentration of BSA. The

response factor is defined as the ratio of the absorbance value of the sample after being

incubated with BSA for 1 h over that of the initial sample at the resonance peak.

3.3 Characterization and application of Ag/WO3-x hybrids

Since WO3-x sample prepared at 350 °C has a strong NIR LSPR absorption, it is
expected to observe wide spectral absorption through combining inter-band and

17
plasmon absorption of WO3-350 sample and Ag NPs. Moreover, the synergistic effect
of plasmon resonance between WO3-350 and Ag NPs is likely to strengthen the
interactions of Ag/WO3-x hybrids with light. Consequently, improved photocatalytic
performance is expected to be obtained over Ag/WO3-350 hybrids. Herein,
Ag/WO3-350 hybrids were prepared by a chemical reduction method (Experimental
section). To illustrate the role of plasmon effect of WO3-350 and the existence of
plasmon-plasmon coupling between WO3-350 and Ag, the control experiments with
the WO3-250 sample that has no LSPR under identical conditions were also
conducted for a comparison. For the XRD of Ag/WO3-350 hybrid (Fig. 5a), besides
diffraction peaks of monoclinic WO2.9 (JCPDS 05-0386), a new peak at 38.11°
corresponding to (111) face of metallic Ag (JCPDS 04-0783) can be observed. From
the HRTEM image of Ag/WO3-350 (Fig. 5b), it can be observed that the calculated
interplanar distances are found to be 0.37 nm and 0.23 nm, which correspond to the
reflection of the (302) plane of WO3-x and (111) plane of Ag, respectively. These
results confirm that Ag/WO3-350 hybrid has been successfully obtained. Similarly,
Ag/WO3-250 hybrid can also be proven to be obtained by XRD spectra and HRTEM
image (Fig. S7).

Fig. 5. a) XRD spectra of WO3-350 sample and Ag/WO3-350 hybrid. b) HRTEM

image of Ag/WO3-350 hybrid.

We also conducted XPS spectra to investigate the elemental compositions and


18
chemical states of Ag/WO3-x hybrids. As shown in Fig. 6a, only the peaks of W and O

with C as a reference can be observed for WO3-250 and WO3-350 samples. New peaks

attribute to Ag 3d in the range from 365 to 380 eV can be found for Ag/WO3-250 and

Ag/WO3-350. The Ag 3d XPS spectrum of Ag/WO3-250 and Ag/WO3-350 exhibit

two peaks at 367.6 and 373.6 eV, correlating with the binding energy of Ag 3d 5/2 and

Ag 3d3/2 of metallic Ag [44], respectively (Fig.6b). As shown in Fig. 6c, Raman

spectra were further performed to determine the evolution of bonds upon Ag

deposition. Obviously, the peaks at 805 cm-1 of all the hybrids present blue shift

compared with that of pure WO3-x samples, which demonstrates the existence of the

strong interaction between Ag NPs and WO3-x sample [45]. And the new peaks at 906

cm-1 of all hybrids do not belong to monoclinic tungsten oxide, revealing the

distortion of WO6 octahedral units forming tungsten bronze with the intercalation of

Ag atoms into WO3 tunnels at the interface [44].

The optical properties of WO3-x and Ag/WO3-x hybrids were studied by

UV-Vis-NIR absorption spectra (Fig.6d). After loading of Ag NPs on WO3-x surfaces,

the LSPR peak of Ag NPs did not appear. This may be due to the low content of Ag in

Ag/WO3-x hybrids [46]. However, the absorbance edges of WO3-250 and WO3-350

are found to be red shifted and absorption in the visible region is enhanced, which can

be due to the overlap of bandgap absorption for WO3-x and plasmonic absorption for

Ag [23]. For Ag/WO3-350 hybrid, the enhancement of light absorption also appears in

the near-infrared region, which does not happen for Ag/WO3-250 hybrid. Considering

that WO3-350 sample has LSPR effect, so the enhancement of light absorption in the
19
near-infrared region should be a result of plasmon-plasmon coupling between

WO3-350 and Ag nanoparticles [47]. Through combining inter-band absorption and

LSPR absorption of WO3-350 and Ag, we have achieved maximum light absorption.

Fig. 6. a) XPS survey spectrum of WO3-250, WO3-350, Ag/WO3-250 and


Ag/WO3-350. b) High-resolution Ag 3d XPS spectra of Ag/WO3-250 and
Ag/WO3-350. c) Raman spectra of WO3-250, WO3-350, Ag/WO3-250 and
Ag/WO3-350. d) UV/Vis-NIR absorbance spectra of WO3-250, WO3-350,
Ag/WO3-250 and Ag/WO3-350.

We explored the photocatalytic activity of Ag/WO3-x hybrids for the degradation

of methylene blue (MB) under visible light (λ ≥ 420 nm). Prior to the photocatalytic

test, 30 min pre-adsorption was performed in order to reach an adsorption–desorption

equilibrium. As shown in Fig. 7a, the black dots refer to degradation of MB over
20
catalyst in dark conditions after the pre-adsorption process for 30 min. There was no

significant change in the concentration of methylene blue within 180 minutes, which

demonstrates that the system has achieved adsorption-desorption equilibrium after 30

min of pre-adsorption. After 180 min of irradiation, only about 10% of MB was

removed in the absence of catalyst indicating its own weak photodegradation, while

the concentration of MB solution decreased by about 34% when WO3-350 sample

was applied to degradation. As to Ag/WO3-350 hybrid, the intensity of the absorption

peak at 664 nm decreased drastically over time (Fig. 7b). It is obvious that much more

MB is degraded in the presence of Ag/WO3-350 hybrid and the degradation rate of

MB increased to 78%. The improvement can be due to that silver doping can cause

enhanced light absorption (Fig. 6d) thus producing more photo-generated

electron-hole pairs. In addition, because of the relative position of energy band of

WO3-350 and Ag, effective electron transfer from the conduction band of WO3-x to

the Fermi level (EF) of Ag can occur, resulting in efficient separation of

photo-generated electron-hole pairs [48]. As shown in Fig. S8, a lower PL intensity of

Ag/WO3-350 compared to that of WO3-350 suggesting a lower recombination rate of

photo-excited charge carriers, which results in higher photocurrent density (Fig. S9)

and improved photocatalytic performance. While for WO3-250 as well as

Ag/WO3-250, they have no LSPR absorption above 600 nm (Fig. 6d). So the higher

photocatalytic efficiency of Ag/WO3-350 compared with that of Ag/WO3-250 should

be derived from the plasmon effect of WO3-350. The plasmon effect of WO3-350

endows Ag/WO3-350 enhanced light absorption in the NIR region, and then a large
21
number of hot energetic electrons are generated by LSPR excitation.

We also performed photocatalytic degradation on methyl orange. The results

illustrated that Ag/WO3-350 still exhibited better photo-degradation efficiency than

that of Ag/WO3-250 or WO3-350 (Fig. S10). In order to further reveal the impact of

plasmon effect on photocatalytic performance, we also conducted photo-degradation

experiment over MB under NIR light irradiation (λ ≥800 nm). As can be seen in

Fig.7c, Ag/WO3-250 sample shows negligible photocatalytic activity for MB

degradation for the reason of lacking of harvesting near-infrared light. While

WO3-350 and Ag/WO3-350 exhibit improved NIR-driven photocatalytic activities.

More importantly, Ag/WO3-350 presents the best photocatalytic degradation

efficiency because of plasmon-plasmon coupling effect (Fig.6d) and efficient

separation of electron-hole pairs (Fig. S8). As shown in Fig. 7d, after four continuous

cycles, the recyclability and stability of Ag/WO3-350 possessed on MB degradation

can be confirmed.

22
Fig. 7. a) Photocatalytic degradation of MB solutions in the presence of WO3-350,
Ag/WO3-250 and Ag/WO3-350 hybrids under visible light irradiation. The black dot
refers to degradation of MB over Ag/WO3-350 hybrid in the dark after 30 min
pre-absorption. b) Absorption spectra of photocatalytic degradation of MB over
Ag/WO3-350 hybrid under visible light irradiation. c) Photocatalytic degradation of
MB solutions in the presence of WO3-350, Ag/WO3-250 and Ag/WO3-350 hybrids
under NIR light irradiation. d) Recycling stability of Ag/WO3-350 hybrid for
degradation of MB solutions under visible light irradiation for 4 h.

The mechanism of enhanced photocatalytic activity of Ag/WO3-350 toward MB

degradation under NIR light irradiation can be proposed based on the above

experimental observations. Under the illumination of near-infrared light, the plasmon

excitation of WO3-350 would give rise to hot energetic electrons, which would soon

transfer to the contiguous Ag NPs, resulting in increased carrier density on Ag


23
surface. In this way, the surface of Ag NPs is negatively charged and WO3-x support is

positively charged, which is beneficial to fast interfacial charge transfer between

catalyst and reagents [49], and eventually improves the catalytic activity of MB

degradation (Fig. 8).

Fig. 8. Schematic illustration of plasmon enhanced photocatalytic activity for

degradation of MB over Ag/WO3-350 hybrid.

4. Conclusions

In summary, substoichiometric WO3-x was successfully prepared by the reduction

of the original tungsten oxide with NaBH4 at different temperature. The introduction

of a large number of oxygen vacancies in tungsten oxide can induce the surface

plasmon resonance, which can be tuned by regulating the annealing temperature to

change the stoichiometry. The application of the substoichiometric WO3-x in a


24
biosensing system was studied by using BSA as a model protein. We discover that the

LSPR intensity depends upon the concentration of BSA, which is a result of

electrostatic interaction and adsorption between BSA and WO3-x. It demonstrates the

ability of substoichiometric WO3-x can be used to quantitatively detect organism, thus

it can supply a platform for developing tunable optoelectronic devices in the future.

Moreover, we also successfully fabricated Ag/WO3-x hybrids by wet chemical

reduction method. In terms of optical performance, Ag/WO3-350 hybrid shows the

maximum light absorption in visible and near-infrared regions by combining LSPR

absorption of WO3-350 and Ag. For the photocatalytic degradation of MB,

Ag/WO3-350 hybrid proves to have the highest activity, in which plasmon effect of

WO3-350 plays an important role to generate hot electrons and facilitate the charge

carrier separation and transfer. Therefore this work supplies a facile route to fabricate

plasmonic WO3-x, and it can be anticipated that this method will be very promising for

application in photocatalysis and the related fields in the near future.

Acknowledgements

We are grateful to the National Natural Science Foundation of China

(No. 21773216, 51173170, 51473149, 21571157), the Science and Technology

Project of Henan Province (152102410010), and the Key program of science and

technology (121PZDGG213) from Zhengzhou Bureau of science and technology.

25
References
[1] K.A. Willets, R.P. Van Duyne, Localized surface plasmon resonance spectroscopy
and sensing, Annu. Rev. Phys. Chem. 58 (2007) 267-297.
[2] S.S. Aćimović, H. Šípová, G. Emilsson, A.B. Dahlin, T.J. Antosiewicz, M. Käll,
Superior LSPR substrates based on electromagnetic decoupling for on-a-chip
high-throughput label-free biosensing, Light: Sci. Appl. 6 (2017) e17042.
[3] S.-Y. Ding, J. Yi, J.-F. Li, B. Ren, D.-Y. Wu, R. Panneerselvam, Z.-Q. Tian,
Nanostructure-based plasmon-enhanced Raman spectroscopy for surface analysis
of materials, Nat. Rev. Mater. 1 (2016) 16021.
[4] X. Zhang, Y.L. Chen, R.S. Liu, D.P. Tsai, Plasmonic photocatalysis, Rep. Prog.
Phys. 76 (2013) 046401.
[5] X. Ding, C.H. Liow, M. Zhang, R. Huang, C. Li, H. Shen, M. Liu, Y. Zou, N. Gao,
Z. Zhang, Y. Li, Q. Wang, S. Li, J. Jiang, Surface plasmon resonance enhanced
light absorption and photothermal therapy in the second near-infrared window, J.
Am. Chem. Soc. 136 (2014) 15684-15693.
[6] G. Song, J. Shen, F. Jiang, R. Hu, W. Li, L. An, R. Zou, Z. Chen, Z. Qin, J. Hu,
Hydrophilic molybdenum oxide nanomaterials with controlled morphology and
strong plasmonic absorption for photothermal ablation of cancer cells, ACS Appl.
Mater. Interfaces 6 (2014) 3915-3922.
[7] X. Ming, A. Guo, G. Wang, X. Wang, Two-dimensional defective tungsten oxide
nanosheets as high performance photo-absorbers for efficient solar steam
generation, Sol. Energy Mater. Sol. Cells 185 (2018) 333-341.
[8] L. Zhang, K. Xia, Z. Lu, G. Li, J. Chen, Y. Deng, S. Li, F. Zhou, N. He, Efficient
and Facile Synthesis of Gold Nanorods with Finely Tunable Plasmonic Peaks
from Visible to Near-IR Range, Chem. Mater. 26 (2014) 1794-1798.
[9] S. Link, C. Burda, M.B. Mohamed, B. Nikoobakht, M.A. El-Sayed, Femtosecond
transient-absorption dynamics of colloidal gold nanorods: Shape independence of
the electron-phonon relaxation time, Phys Rev B, Phys. Rev. B 61 (2000)
6086-6090.
[10] X. Wang, Y. Ke, H. Pan, K. Ma, Q. Xiao, D. Yin, G. Wu, M.T. Swihart,
Cu-Deficient Plasmonic Cu2–xS Nanoplate Electrocatalysts for Oxygen
Reduction, ACS Catal. 5 (2015) 2534-2540.
[11] D. Zhou, D. Liu, W. Xu, Z. Yin, X. Chen, P. Zhou, S. Cui, Z. Chen, H. Song,
Observation of Considerable Upconversion Enhancement Induced by Cu 2-xS
Plasmon Nanoparticles, ACS Nano 10 (2016) 5169-5179.
[12] Z. Liu, R. Beaulac, Nature of the Infrared Transition of Colloidal Indium Nitride
Nanocrystals: Nonparabolicity Effects on the Plasmonic Behavior of Doped
Semiconductor Nanomaterials, Chem. Mater. 29 (2017) 7507-7514.
[13] H. Cheng, X. Qian, Y. Kuwahara, K. Mori, H. Yamashita, A Plasmonic
Molybdenum Oxide Hybrid with Reversible Tunability for
Visible-Light-Enhanced Catalytic Reactions, Adv. Mater. 27 (2015) 4616-4621.
[14] K. Manthiram, A.P. Alivisatos, Tunable localized surface plasmon resonances in
tungsten oxide nanocrystals, J. Am. Chem. Soc. 134 (2012) 3995-3998.
26
[15] J. Yan, T. Wang, G. Wu, W. Dai, N. Guan, L. Li, J. Gong, Tungsten oxide single
crystal nanosheets for enhanced multichannel solar light harvesting, Adv. Mater.
27 (2015) 1580-1586.
[16] N.M.Y. Zhang, K. Li, T. Zhang, P. Shum, Z. Wang, Z. Wang, N. Zhang, J. Zhang,
T. Wu, L. Wei, Electron-Rich Two-Dimensional Molybdenum Trioxides for
Highly Integrated Plasmonic Biosensing, ACS Photonics 5 (2017) 347-352.
[17] J.B. Goodenough, The two components of the crystallographic transition in VO2
☆, J. Solid State Chem. 3 (1971) 490-500.
[18] O.A. Balitskii, D. Moszyński, Z. Abbas, Aqueous processable WO3−x nanocrystals
with solution tunable localized surface plasmon resonance, RSC Adv. 6 (2016)
59050-59054.
[19] Z. Lou, Q. Gu, Y. Liao, S. Yu, C. Xue, Promoting Pd-catalyzed Suzuki coupling
reactions through near-infrared plasmon excitation of WO3−x nanowires, Appl.
Catal., B 184 (2016) 258-263.
[20] H. Watanabe, K. Fujikata, Y. Oaki, H. Imai, Band-gap expansion of tungsten oxide
quantum dots synthesized in sub-nano porous silica, Chem. Commun. 49 (2013)
8477-8479.
[21] M. Epifani, E. Comini, R. Diaz, T. Andreu, A. Genc, J. Arbiol, P. Siciliano, G.
Faglia, J.R. Morante, Solvothermal, chloroalkoxide-based synthesis of monoclinic
WO3 quantum dots and gas-sensing enhancement by surface oxygen vacancies,
ACS Appl. Mater. Interfaces 6 (2014) 16808-16816.
[22] G. Wang, Y. Ling, H. Wang, X. Yang, C. Wang, J.Z. Zhang, Y. Li,
Hydrogen-treated WO3 nanoflakes show enhanced photostability, Energy
Environ. Sci. 5 (2012) 6180.
[23] S. Ghosh, M. Saha, S. Paul, S.K. De, Maximizing the photo catalytic and photo
response properties of multimodal plasmonic Ag/WO3-x heterostructure nanorods
by variation of the Ag size, Nanoscale 7 (2015) 18284-18298.
[24] Z. Lou, Q. Gu, L. Xu, Y. Liao, C. Xue, Surfactant-free synthesis of plasmonic
tungsten oxide nanowires with visible-light-enhanced hydrogen generation from
ammonia borane, Chem. Asian J. 10 (2015) 1291-1294.
[25] M. Li, M. Hu, D. Jia, S. Ma, W. Yan, NO2-sensing properties based on the
nanocomposite of n-WO3−x/n-porous silicon at room temperature, Sens.
Actuators, B 186 (2013) 140-147.
[26] Z. Chen, Q. Wang, H. Wang, L. Zhang, G. Song, L. Song, J. Hu, H. Wang, J. Liu,
M. Zhu, D. Zhao, Ultrathin PEGylated W18O49 nanowires as a new 980
nm-laser-driven photothermal agent for efficient ablation of cancer cells in vivo,
Adv. Mater. 25 (2013) 2095-2100.
[27] L. Wen, L. Chen, S. Zheng, J. Zeng, G. Duan, Y. Wang, G. Wang, Z. Chai, Z. Li,
M. Gao, Ultrasmall Biocompatible WO3-x Nanodots for Multi-Modality Imaging
and Combined Therapy of Cancers, Adv. Mater. 28 (2016) 5072-5079.
[28] Y. Ren, Q. Xu, X. Zheng, Y. Fu, Z. Wang, H. Chen, Y. Weng, Y. Zhou, Building of
peculiar heterostructure of Ag/two-dimensional fullerene shell-WO3-x for
enhanced photoelectrochemical performance, Appl. Catal., B 231 (2018) 381-390.
[29] P. Basnet, Y. Zhao, Superior dye adsorption capacity of amorphous WO 3
27
sub-micrometer rods fabricated by glancing angle deposition, J. Mater. Chem. A 2
(2014) 911-914.
[30] K. Yuan, Q. Cao, H.-L. Lu, M. Zhong, X. Zheng, H.-Y. Chen, T. Wang, J.-J.
Delaunay, W. Luo, L. Zhang, Y.-Y. Wang, Y. Deng, S.-J. Ding, D.W. Zhang,
Oxygen-deficient WO3−x@TiO2−x core–shell nanosheets for efficient
photoelectrochemical oxidation of neutral water solutions, J. Mater. Chem. A 5
(2017) 14697-14706.
[31] J. Liu, Y. Zhao, Z. Zhang, Low-temperature synthesis of large-scale arrays of
aligned tungsten oxide nanorods, J. Phys. Condens. Matter 15 (2003) L453-L461.
[32] W. Wang, A. Janotti, C.G. Van de Walle, Role of oxygen vacancies in crystalline
WO3, J. Mater. Chem. C 4 (2016) 6641-6648.
[33] G.V. Naik, V.M. Shalaev, A. Boltasseva, Alternative plasmonic materials: beyond
gold and silver, Adv. Mater. 25 (2013) 3264-3294.
[34] M.J. Polking, P.K. Jain, Y. Bekenstein, U. Banin, O. Millo, R. Ramesh, A.P.
Alivisatos, Controlling localized surface plasmon resonances in GeTe
nanoparticles using an amorphous-to-crystalline phase transition, Phys. Rev. Lett.
111 (2013) 037401.
[35] H. Cheng, M. Wen, X. Ma, Y. Kuwahara, K. Mori, Y. Dai, B. Huang, H.
Yamashita, Hydrogen Doped Metal Oxide Semiconductors with Exceptional and
Tunable Localized Surface Plasmon Resonances, J. Am. Chem. Soc. 138 (2016)
9316-9324.
[36] J.A. Faucheaux, A.L. Stanton, P.K. Jain, Plasmon Resonances of Semiconductor
Nanocrystals: Physical Principles and New Opportunities, J. Phys. Chem. Lett. 5
(2014) 976-985.
[37] C. Guo, S. Yin, M. Yan, M. Kobayashi, M. Kakihana, T. Sato,
Morphology-controlled synthesis of W18O49 nanostructures and their near-infrared
absorption properties, Inorg. Chem. 51 (2012) 4763-4771.
[38] P. Uppachai, V. Harnchana, S. Pimanpang, V. Amornkitbamrung, A.P. Brown,
R.M.D. Brydson, A substoichiometric tungsten oxide catalyst provides a
sustainable and efficient counter electrode for dye-sensitized solar cells,
Electrochim. Acta 145 (2014) 27-33.
[39] P. Zhou, Q. Xu, H. Li, Y. Wang, B. Yan, Y. Zhou, J. Chen, J. Zhang, K. Wang,
Fabrication of Two-Dimensional Lateral Heterostructures of WS 2/WO3·H2O
Through Selective Oxidation of Monolayer WS2, Angew. Chem., Int. Ed. 54
(2015) 15226-15230.
[40] A. Lewera, L. Timperman, A. Roguska, N. Alonso-Vante, Metal–Support
Interactions between Nanosized Pt and Metal Oxides (WO3 and TiO2) Studied
Using X-ray Photoelectron Spectroscopy, J. Phys. Chem. C 115 (2011)
20153-20159.
[41] L. Song, K. Yang, W. Jiang, P. Du, B. Xing, Adsorption of bovine serum albumin
on nano and bulk oxide particles in deionized water, Colloids Surf., B 94 (2012)
341-346.
[42] W. Liu, Q. Xu, W. Cui, C. Zhu, Y. Qi, CO2-Assisted Fabrication of
Two-Dimensional Amorphous Molybdenum Oxide Nanosheets for Enhanced
28
Plasmon Resonances, Angew. Chem., Int. Ed. 56 (2017) 1600-1604.
[43] G. Guan, J. Xia, S. Liu, Y. Cheng, S. Bai, S.Y. Tee, Y.W. Zhang, M.Y. Han,
Electrostatic-Driven Exfoliation and Hybridization of 2D Nanomaterials, Adv.
Mater. 29 (2017).
[44] W. Zhu, J. Liu, S. Yu, Y. Zhou, X. Yan, Ag loaded WO3 nanoplates for efficient
photocatalytic degradation of sulfanilamide and their bactericidal effect under
visible light irradiation, J. Hazard. Mater. 318 (2016) 407-416.
[45] J. Ding, L. Zhang, Q. Liu, W.-L. Dai, G. Guan, Synergistic effects of electronic
structure of WO3 nanorods with the dominant {001} exposed facets combined
with silver size-dependent on the visible-light photocatalytic activity, Appl. Catal.,
B 203 (2017) 335-342.
[46] P. Dong, B. Yang, C. Liu, F. Xu, X. Xi, G. Hou, R. Shao, Highly enhanced
photocatalytic activity of WO3 thin films loaded with Pt–Ag bimetallic alloy
nanoparticles, RSC Adv. 7 (2017) 947-956.
[47] J.H. Lee, M.H. You, G.H. Kim, J.M. Nam, Plasmonic nanosnowmen with a
conductive junction as highly tunable nanoantenna structures and sensitive,
quantitative and multiplexable surface-enhanced Raman scattering probes, Nano
Lett 14 (2014) 6217-6225.
[48] N. Bate, H. Shi, L. Chen, J. Wang, S. Xu, W. Chen, J. Li, E. Wang,
Micelle-Directing Synthesis of Ag-Doped WO3 and MoO3 Composites for
Photocatalytic Water Oxidation and Organic-Dye Adsorption, Chem. Asian J. 12
(2017) 2597-2603.
[49] S. Linic, P. Christopher, D.B. Ingram, Plasmonic-metal nanostructures for
efficient conversion of solar to chemical energy, Nat. Mater. 10 (2011) 911-921.

29
Highlights
Fabrication and Application of Substoichiometric Tungsten Oxide with Tunable

Localized Surface Plasmon Resonances

Jun Chen, Yumei Ren, Tianzhao Hu, Tao Xu, Qun Xu*

College of Materials Science & Engineering, Zhengzhou University, Zhengzhou,

450052, P.R.China

 In this study substoichiometric WO3-x with tunable plasmon resonances and

remarkable NIR absorption is fabricated.

 In the bovine serum albumin (BSA) based optical bio-sensing system, WO3-x

displays superior sensitivity owing to LSPR effect.

 The synergistic effect of the strong plasmonic effect of WO3-350 and Ag

nanoparticles helps to achieve the maximum utilization of solar light in full solar

spectrum.

 The plasmon-generated hot electrons of WO3-x can transfer to Ag nanoparticles

efficiently.

 The as-prepared plasmonic Ag/WO3-x hybrid exhibits significantly enhanced

photocatalytic performance.

30
Graphical Abstract
Fabrication and Application of Substoichiometric Tungsten Oxide with Tunable

Localized Surface Plasmon Resonances

Jun Chen, Yumei Ren, Tianzhao Hu, Tao Xu, Qun Xu*

College of Materials Science & Engineering, Zhengzhou University, Zhengzhou,

450052, P.R.China

By fabricating a plasmonic Ag/WO3-x hybrid, significantly improved

photocatalytic performance can be obtained. This is considered to be a result of the

synergistic effect of the strong plasmonic effect of WO3-350 and Ag nanoparticles

31
which helps to achieve the maximum utilization of solar light in full solar

spectrum. Another key point is that the plasmon-generated hot electrons of

WO3-350 can transfer to Ag nanoparticles efficiently.

32

You might also like