Next Article in Journal
A Digital Twin of the Coaxial Lamination Mixer for the Systematic Study of Mixing Performance and the Prediction of Precipitated Nanoparticle Properties
Next Article in Special Issue
Simulation and Synthesis of Cobalt (Co) Nanoparticles by Gamma Radiation Technique
Previous Article in Journal
A Novel Memristive Neural Network Circuit and Its Application in Character Recognition
Previous Article in Special Issue
A Study of Nano-Tungsten Colloid Preparing by the Electrical Spark Discharge Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structure and Photoluminescence of WO3-x Aggregates Tuned by Surfactants

1
College of Chemistry and Chemical Engineering, Chongqing University of Technology, 69 Hongguang Rd, Lijiatuo, Banan District, Chongqing 400054, China
2
State Key Laboratory of Advanced Optical Communication and Networks, Key Laboratory for Laser Plasmas (Ministry of Education), School of Physics and Astronomy, Shanghai Jiao Tong University, Shanghai 200240, China
3
School of Materials and Chemical Engineering, Anhui Jianzhu University, Hefei 230601, China
4
School of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou 225009, China
5
Jozef Stefan Institute, Jamova Cesta 39, SI-1000 Ljubljana, Slovenia
6
School of Chemistry and Physics, Queensland University of Technology, Brisbane, QLD 4000, Australia
7
Centre for Materials Science, Queensland University of Technology, Brisbane, QLD 4000, Australia
*
Author to whom correspondence should be addressed.
Micromachines 2022, 13(12), 2075; https://doi.org/10.3390/mi13122075
Submission received: 31 October 2022 / Revised: 18 November 2022 / Accepted: 23 November 2022 / Published: 25 November 2022
(This article belongs to the Special Issue Nanoparticles: Optical Properties and Applications)

Abstract

:
The optoelectronic properties of transition metal oxide semiconductors depend on their oxygen vacancies, nanostructures and aggregation states. Here, we report the synthesis and photoluminescence (PL) properties of substoichiometric tungsten oxide (WO3-x) aggregates with the nanorods, nanoflakes, submicro-spherical-like, submicro-spherical and micro-spherical structures in the acetic acid solution without and with the special surfactants (butyric or oleic acids). Based on theory on the osmotic potential of polymers, we demonstrate the structural change of the WO3-x aggregates, which is related to the change of steric repulsion caused by the surfactant layers, adsorption and deformation of the surfactant molecules on the WO3-x nanocrystals. The WO3-x aggregates generate multi-color light, including ultraviolet, blue, green, red and near-infrared light caused by the inter-band transition and defect level-specific transition as well as the relaxation of polarons. Compared to the nanorod and nanoflake WO3-x aggregates, the PL quenching of the submicro-spherical-like, submicro-spherical and micro-spherical WO3-x aggregates is associated with the coupling between the WO3-x nanoparticles and the trapping centers arising from the surfactant molecules adsorbed on the WO3-x nanoparticles.

1. Introduction

Tungsten oxides are important semiconductor materials. Since their excellent physical and chemical properties are linked to the size, oxygen vacancies, nanostructures and aggregation states, it is still a significant challenge to tune their structures and properties [1,2]. When inorganic nanomaterials are synthesized in solutions, surfactants are universally used to tune their structures by adjusting the aggregation of nanocrystals [3,4,5,6]. For example, the In2O3 nanodots are aggregated into nanoflowers depending on myristic acid [6]. Furthermore, the studies indicate that some properties of the aggregated materials (e.g., carrier transport and photoluminescence) strongly depend on the coupling separation between the nanocrystals [5,7], which are related to the lengths of the surfactant molecules [8,9,10]. In other words, surfactants play important roles in tuning the aggregated nanomaterials’ structures and properties. In a solution containing the surfactants, there are attraction and repulsive interactions between the nanocrystals. The attraction interactions (i.e., dipole-dipole and van der Waals interactions) promote the aggregation between the nanocrystals, while the repulsive interactions (e.g., steric and electrostatic repulsion forces) hinder the aggregation [3,11]. Due to the strong attraction interactions, the nanocrystals are aggregated together [11]. Therefore, tuning of the aggregation process weakens the attractions by the introduction of the surfactants. This paper explores the tuning of the structures and photoluminescence (PL) properties of substoichiometric tungsten oxide (WO3-x, 2 < x < 3) aggregates through adjusting the aggregation process using the solvothermal synthesis technique.
Compared to stoichiometric tungsten oxides (WO3 and WO2), the WO3-x nanomaterials have fascinating luminescent and catalytic properties, which are related to the aggregation and the defect levels formed by the oxygen vacancies [12,13,14,15,16,17]. In other words, the WO3-x materials are becoming the important semiconductor materials for optoelectronic and catalytic applications. Therefore, the WO3-x materials have attracted considerable attention and great efforts contribute to the foundation and application studies in recent years [14,15,17,18,19,20,21].
Various WO3-x nanostructures (e.g., films, nanowires, nanodots and hybrid nanostructures) have been fabricated using physical and chemical methods [12,14,19,20,21]. In particular, the WO3-x nanomaterials are universally synthesized in different solvents using the solvothermal method [22]. It is well known that various WO3-x nanostructures are made up of nanoparticles, i.e., the WO3-x nanostructures can be tuned by adjusting the aggregation of the WO3-x nanoparticles. The early studies indicate that the aggregation of the WO3-x nanoparticles is indeed tuned by the surfactants [23,24]. However, the mechanism has not been well understood because the repulsion effects caused by the surfactants involve the lengths, adsorption, compress, etc. of the surfactant molecules [25]. In this paper, the special surfactants (i.e., butyric acid and oleic acid) were employed to tune the structures of the WO3-x aggregates, respectively. Based on the osmotic potential of polymers, the effects of the surfactants on the aggregation of the WO3-x nanoparticles are analyzed. Furthermore, the PL properties of the WO3-x aggregates were studied.

2. Materials and Method

2.1. Materials

In the process of synthesizing the WO3-x aggregates, the WCl6 powder, acetic, butyric and oleic acids were used as the precursors. Among these precursors, the WCl6 powder (99%) was supplied by Yanyu (Shanghai) Chemical Agent Co., Ltd.) and all the carboxylic acids including acetic acid (≥99.8%), butyric acid (99%) and oleic acid (chemical grade) were provided by Shanghai Macklin Biochemical Technology Co., Ltd. (Shanghai, China). The previous studies indicate that the metal acetate was used to fabricate the corresponding metal oxide nanocrystals [6]. Here, the WCl6 powder and acetic acid was used to synthesize the WO3-x nanocrystals, and the butyric and oleic acids were used as the surfactants to tune the structure of the WO3-x aggregates, respectively.

2.2. Synthesis

Due to differences in the viscosity of different carboxylic acids, the acetic acid and surfactant were mixed in a 100 mL Teflon-lined autoclave by magnetic stirring for 30 min to ensure the uniformity of the solution. After the WCl6 powder was poured into the solution, the solution was further stirred magnetically for 30 min, and then the autoclave was encapsulated in a stainless tank. In a heating furnace, the tank was heated to 120 °C in ~60 min and maintained for 12 h at this temperature. The loose WO3-x agglomerate was obtained at room temperature by centrifugal precipitation and drying treatment for 12 h at 50 °C. Furthermore, the agglomerate was ground into the powder to use for further characterizations. Different WO3-x samples were synthesized in the experiments and the synthesis conditions are summarized in Table 1.

2.3. Characterization and PL Measurement

The phases and composition of the as-synthesized WO3-x aggregates are determined according to the data recorded in a TTRAX3 X-ray diffractometer (XRD) using a Cu Kα radiation with a wavelength of 0.154 nm and a Horiba Scientific LabRAM HR evolution Raman spectrometer under irradiation of the 532 nm line of a semiconductor laser, respectively. The binding states of elements were analyzed depending on the data obtained from an ESCALAB 250 X-ray photoelectron spectroscopy (XPS) equipped with an Al Kα X-ray source. The studies on the surface morphology and microstructures were carried out in an S-4800 field emission scanning electron microscope (FESEM) (operated at 5 kV) and a TECNAI G2 F30 high-resolution transmission electron microscope (TEM) (operated at 300 kV), respectively.
At room temperature, the PL emission from the as-synthesized aggregates was measured in the Horiba Scientific LabRAM HR evolution Raman spectrometer under irradiation of the He-Cd laser with a wave length of 325 nm and ~1 μm diameter of laser beam. During the PL measurement, the power of laser and the acquisition time were set to 6.8 mW and 30 s, respectively.

3. Results

3.1. Phases and Composition

The phases and composition of the as-synthesized WO3-x aggregates are determined by the XRD and Raman results. The XRD patterns of samples A–D are presented in Figure 1 and the phases corresponding to the peaks are summarized in Table 2. According to the data in Table 2, the as-synthesized materials are the WO3-x materials composed of the W19O55, WO2.92 and W18O49 phases. Furthermore, Figure 1 indicates that the strongest peaks in Figure 1b become wide compared to the strongest peak in Figure 1a. In other words, the size of WO3-x nanoparticles synthesized in the solution containing the surfactant becomes small.
The Raman spectra presented in Figure 2 further evidence that the as-synthesized materials are the WO3-x aggregates. To clearly distinguish the Raman peaks, the Raman spectra in Figure 2a are deconvoluted and the results are presented in Figure 2b−f. In Table 3, the deconvoluted Raman peaks are assigned and summarized. Depending on the studies [26], the Raman peaks at 835−859 cm−1 may result from the vibration modes of different W−O bonds.
Furthermore, the Raman peaks in a range of 200−400 cm−1 are associated with the vibration of W5+−O bonds, which indicate the formation of the WO3-x aggregates [29]. In addition, the studies indicate that the Raman spectrum of pure WO3 shows the sharp characteristic peaks at about 806, 716, 326 and 272 cm−1, while the spectra of WO3-x exhibit the wide peaks at these positions and some new peaks due to the distortions in the crystalline structure [30]. According to the studies [30], the Raman peaks in Figure 2 further address that the as-synthesized materials are the WO3-x aggregates.

3.2. Binding States of Elements

The XPS spectra are usually used to analyze the binding states of elements. The XPS survey spectra presented in Figure 3 indicate that the as-synthesized aggregates contain the W, O and C elements. In order to confirm the binding states of these elements, a XPSPEAK 41 analysis software was used to deconvolute the W 4f, O 1s and C 1s XPS spectra after Shirley background subtraction. The results are presented in Figure 4 and the deconvoluted peaks are summarized in Table 4.
From Figure 4a–d, one can see that the W 4f XPS spectra feature two pair of peaks, i.e., one pair of peaks at about 34.7−34.9 eV and 36.1−36.3 eV are attributed to W5+ 4f7/2 and W5+ 4f5/2, and another pair at about 35.5−35.6 eV and 37.8 eV correspond to W6+ 4f7/2 and W6+ 4f5/2 [32,33]. The O 1s XPS spectra in Figure 4f−j are deconvoluted into three peaks located at about 530.3−530.5, 531.3−531.7 and 532.7−533.1 eV, which are related to the oxygen atoms in the lattices (OL), the oxygen in the vicinity of oxygen vacancies (OV) and the oxygen atoms of carboxylic groups (OC) [32,34], respectively. Similarly, the C 1s XPS spectra in Figure 4k–o are fitted by three peaks at about 284.7, 285.5−285.8 and 288.9−289.0 eV, and they are attributed to (CH2)n, CH2-O and O=C-O [31,34], respectively.
The appearance of the XPS peaks related to W5+ indicates the formation of the WO3-x structure and the creation of oxygen vacancies [16]. Furthermore, the O 1s XPS spectra confirm the presence of oxygen vacancies, and the carboxyl acid molecules are adsorbed on the surfaces of the WO3-x aggregates and they are bonded on the surface in the bidentate interaction modes [35].
According to Figure 4f–j, the areas of OL, OV and OC peaks are obtained and they are summarized in Table 5. Referring to Ref. [36], the coverage ratios of the oxygen vacancies on the sample surfaces (i.e., the area ratio R of OV peak to the area sum of OL and OV peaks) are calculated and the results are shown in Table 5. The studies indicate that the insulativity of metal oxides is enhanced by the reduction of oxygen vacancies [37]. In other words, the conductivity is enhanced by the oxygen vacancies. Based on the relation of the conductivity with the oxygen vacancies, the data in Table 5 indicate that the conductivity of sample B is high compared to sample A. Furthermore, the conductivity of samples C and D reduces compared to sample B. However, the conductivity of sample E becomes higher than sample B. The change in the conductivity may originate from the structural conversion of the WO3-x aggregates and the amount of carboxylic acid as well as the lengths of carboxylic acid molecules. The SEM images (see Figure 5) indicate that the structures of samples A and B convert from the nanorods and nanoflakes to the aggregates of the nanoparticles. As a result, the surface area of sample B is enhanced to increase the oxygen vacancies. Compared to sample B, Table 1 indicates that the amount of butyric acid increases for sample C, which implies that more butyric acid molecules adsorb on the surface of the WO3-x nanoparticles to increase the insulativity. Therefore, the conductivity is reduced. When sample D was synthesized, oleic acid was employed. Due to the long chain of an oleic acid molecule, the insulativity of sample D is enhanced to reduce the conductivity. For sample E, much oleic acid was used. The high viscosity of oleic acid leads to the reduction of the oleic acid molecules adsorbed on the WO3-x nanoparticles (see the analyses in Section 3.4). As a consequence, the conductivity is increased.

3.3. Morphology and Microstructure

The morphologies of the as-synthesized WO3-x aggregates are presented in Figure 5. Figure 5a exhibits that sample A contains the nanoflakes and nanorods as well as some aggregated nanoparticles, which indicate that the WO3-x nanoparticles are aggregated in an oriented mode in acetic acid solution.
Figure 5b−e exhibit the mixture of the submicro-spherical-like, submicro-spherical and micro-spherical structures mixed with the aggregated nanoparticles, respectively. These indicate that the aggregation of the WO3-x nanoparticles is changed by the butyric acid or oleic acid, i.e., they are aggregated in a disoriented mode, but the aggregation is related to the amount of the surfactants. From Figure 5b−e, one can see that the increase in the amount of butyric acid leas to the structural conversion from submicro-spherical-like to submicro-spherical shapes, while the enhancement of oleic acid results in the change from micro-spherical to submicro-spherical-like shapes.
The microstructures of samples A, B and D are revealed in Figure 6, Figure 7, Figure 8. The high-resolution TEM images presented in Figure 6 indicate that the nanoflake, nanorod and nanoparticle in sample A are well crystalized. After the thickness of five crystal planes are measured using digital micrograph software, the lattice distances of the nanoflake, nanorod and nanoparticle are obtained. They are ~0.376, 0.369 and 0.383 nm, respectively, which correspond to the d values (0.368 and 0.370 nm) of the WO2.92(-1016) and W19O55(302) [JCPDS: 30-1387 and 45-0167].
The TEM images in Figure 7 indicate that the WO3-x aggregates in sample B are the submicro-spherical-like structure mixed with some nanorods. Similar to Figure 6, the thickness of ten or five crystal planes are measured using digital micrograph software. According to the thickness values, the lattice spacing of the nanorod and nanoparticle are, respectively, ~0.378 and 0.380 nm, which correspond to the d values (0. 370 and 0.373 nm) of the W19O55(302) and W18O49(103) [JCPDS: 45-0167 and 05-0392].
The TEM images of sample D presented in Figure 8 exhibit a micro-spherical structure. According to the measured thickness of five crystal planes, the lattice spacing is about 0.389 nm, which corresponds to the d value (0.373 nm) of W18O49(103) [JCPDS: 05-0392].

3.4. Aggregation of WO3-x Nanoparticles in Different Solutions

3.4.1. Aggregation of WO3-x Nanoparticles in Acetic Acid

In an earlier paper [13], the formation of the WO3-x nanoparticles in acetic acid was studied, which involves the WCl6−CH3CO2H reaction and the condensation reaction. The WCl6−CH3CO2H reaction leads to the formation of tungsten acetate, and the further condensation reaction of tungsten acetate forms the WO3-x nanoparticles. In acetic acid solution, the WO3-x nanoparticles are covered by the acetic acid molecules. As a result, the WO3-x nanoparticles experience a quenching for a long time to form the WO3-x nanocrystals (NCs). The XPS results indicate the adsorption of the acetic acid molecules on the surfaces of the WO3-x NCs, which indicates that a protecting layer is formed to prevent the aggregation of the WO3-x NCs [38]. However, the adsorption is reversible due to the low molecular weight of acetic acid [25,39]. In other words, some acetic acid molecules can be desorbed from the WO3-x NCs. Due to the anisotropy of the WO3-x NCs, the acetic acid molecules coordinated with the tungsten atoms are changed to selectively adsorb on the special crystal planes [40]. As a result, a strong dipole−dipole interaction occurs between the WO3-x NCs. The directionality of the dipole−dipole interaction leads to the aggregation of the WO3-x NCs along a direction to form the nanorods [11]. As for the formation of the WO3-x nanoflakes, it is related to the gas-liquid interface formed by acetic acid. Due to Brownian motion, some WO3-x NCs move to the gas–liquid interface. Some acetic acid molecules adsorbed on the upper and side surfaces of the WO3-x NCs on the interface are desorbed. As a result, the strong dipole−dipole interaction between the WO3-x NPs leads to the aggregation [11]. During the aggregation, the WO3-x NCs are still constrained on the interface. Finally, the aggregation results in the formation of the WO3-x nanoflakes.

3.4.2. Aggregation under Surfactants

After the surfactant (i.e., butyric acid or oleic acid) is added, it leads to the WO3-x nanoparticles’ size, motion and distribution changes. At the same time, the addition of the surfactant results in the change of the viscosity of the solution. These changes further cause the change in the aggregation process of the WO3-x nanoparticles.
  • Changes in size, motion and distribution of WO3-x nanoparticles
In the nucleation process, the surfactant molecules are adsorbed on the surfaces of the WO3-x nuclei. Unlike the acetic acid molecules, the surfactant molecules have a high molecular wight, which lead to a relatively irreversible adsorption [25,39]. As a result, the nucleation rate is enhanced according to Gibbs equation and energy barrier for nucleation described below. The Gibbs equation is expressed by:
d γ = i Γ i d μ i
where γ , Γ i and μ i represent the surface tension, the superficial adsorption concentration and chemical potential of the adsorbed species i [41], respectively. The nucleation energy barrier Δ G crit is expressed by:
Δ G crit = 16 π 3 γ 3 V m 2 R T ln S 2
where R , T , S and V m are universal gas constant, temperature, the supersaturation and the molar volume of monomers [41], respectively. Based on Equations (1) and (2), the adsorption of the surfactant molecules reduces the surface tension, which further lowers the nucleation energy barrier and enhances the nucleation rate. At the same time, the surfactants molecules adsorbed on the surfaces of the WO3-x nuclei form the surfactant layers to prevent their aggregation and growth. After these nuclei experience quenching in the solution, they are formed into irregular crystal nanodots (CNDs) with a small size, which are evidenced by the TEM results. Here, we should note that the size refers to the size of the WO3-x CNDs without the surfactant layers. Due to the attachment of the surfactant layers on the WO3-x CNDs, the size of the WO3-x CNDs becomes relatively large due to the long carbon chain of the surfactant molecules. In addition, the surfactants result in the increase in the viscosity η of the solution, which can be addressed by the early studies [42]. In Ref. [42], the results indicate that the viscosity of a mixture composed of an organic compound with high viscosity (e.g., ethanol, butanol or octanol) and another organic compound with low viscosity (e.g., propylamine) reduces with the increase in the mole fraction of the compound with low viscosity. In other words, the viscosity of the mixtures increases with the increase in the mole fraction of the compound with high viscosity. Due to the relation of mole fraction x with volume v ( x = ρ v / M , where ρ and M are the density and molar mass), the viscosity of the mixtures increases with the increase in the volume of the compound with high viscosity. The viscosity of acetic, butyric and oleic acids at 20 °C is 1.23, 1.54 and 34.8 mPa·s [43,44], respectively. According to these data, the viscosity of the solutions increases with the increase in the volume of the surfactants. Therefore, the introduction of the surfactant leads to the change in the motion of the WO3-x CNDs. The relation of diffusion coefficient D diff with the radius of particle r and viscosity η is:
D diff = k B T 6 π r η
where k B is Boltzmann’s constant [22]. According to Equation (3), the diffusion coefficient is reversely proportional to the radius of particle and the viscosity of solution. Therefore, the motion of the WO3-x CNDs becomes slow in the solution. In other words, the aggregation of the WO3-x CNDs takes place in the interior of the solution. Because the formation of the WO3-x CNDs originates from the WCl6−CH3COO reaction, the slow speed of the WO3-x CNDs makes them located in the reaction region, i.e., their distribution in the solution is nonuniform.
2.
Aggregation of WO3-x nanoparticles
Unlike the aggregation of the WO3-x NCs in acetic acid, the WO3-x CNDs are aggregated into the submicro-spherical-like, submicro-spherical and micro-spherical structures under the surfactants (Figure 5), which are related to the repulsion interactions between the WO3-x CNDs and the property of solution. In this section, the aggregation under the surfactants is analyzed.
For the convenience of analysis, the WO3-x CNDs are considered to be a spherical structure and the surfactant molecule is represented by a bending line, which are presented in Figure 9. Due to Brownian motion, the surfactant layers are overlapped when two WO3-x CND approach. As a result, the increased concentration of the surfactant tails in the overlap region increases osmotic pressure, generating steric repulsion [25,45]. The theoretical studies indicate the osmotic potential V Osm is related to the thickness of the polymer layer and the volume fraction of polymer molecules in the layer [46], which is expressed by:
V Osm = 0 ,   d 2 l
V Osm = k B T 4 π a v 1 p 2 1 2 χ l d 2 2 ,   l d < 2 l
V Osm = k B T 4 π a v 1 p 2 1 2 χ l 2 d 2 l 1 4 ln d l ,   d < l
where a , l , d and p represent the radius of a particle, the thickness of polymer layer, the distance between the particles and the effective volume fraction of polymer molecules in the polymer layer, respectively, and v 1 and χ correspond to the volume of one solvent molecule and the Flory–Huggins solvency parameter, respectively. Here, p depends on the concentration of surfactant tails in the overlap region of the surfactant layers [25]. Form Equations (5) and (6), we obtain:
V Osm l / d = 2 k B T 4 π a v 1 p 2 d 2 1 2 χ l d 1 2 > 0 ,   l d < 2 l
V O s m d / l = k B T 4 π a v 1 p 2 1 2 χ l 2 1 2 l d < 0 ,   d < l
According to Equations (7) and (8), the steric repulsion increases with the thickness of the surfactant layers and the number of the surfactant tails in the overlap region. If the surfactant layers are thin, the steric repulsion is increased by increasing the amount of the surfactants.
In the initial process of aggregation, the distribution of the WO3-x CNDs in the reaction region is nonuniform. Due to long chains of the surfactant molecules, the diffusion is slow. As a result, a number of the WO3-x CNDs are partly covered by the surfactant molecules and others are completely capped, which are shown in Figure 10a. Under the attraction interactions, the WO3-x CNDs are aggregated in an irregular structure (Figure 10a). With the progress of aggregation, the surfactant molecules gradually come into the irregular structure through diffusion and are adsorbed on the WO3-x CNDs (Figure 10b). As shown in Figure 10b, the concentration of the surfactant tails around the WO3-x CNDs increases, which leads to the enhancement of the steric repulsion [25]. In this situation, the attraction interactions between the WO3-x CNDs are gradually weakened to tune the aggregation process. If the steric repulsion cannot balance the attraction interactions, the WO3-x CNDs are aggregated, but the arrangement of the WO3-x CNDs is tuned due to the presence of the steric repulsion (Figure 10b). Depending on the steric repulsion, different WO3-x CNDs form the submicro-spherical-like or submicro-spherical and micro-shperical structures to reduce the surface energy, which are shown in Figure 10b,c.
Due to the short chain of the butyric acid molecule (~0.6 nm) compared to an oleic acid molecule (~2.5 nm) [10], the butyric acid molecules form thin surfactant layers on the WO3-x CNDs. With the increase in the amount of butyric acid, the steric repulsion is enhanced. As a result, the aggregation is further tuned and the structure coverts from the submicro-spherical-like to submicro-spherical shapes (Figure 10b,c). These are the reasons why Figure 5b,c present the submicro-spherical-like and submicro-spherical WO3-x aggregates. For oleic acid, the long chain of molecules results in the formation of the thick surfactant layers on the WO3-x CNDs, which generates a strong steric repulsion [47]. Furthermore, the separation between the WO3-x CNDs is increased. Consequently, the WO3-x CNDs are easily rearranged in a large region to form the micro-spherical structure. Therefore, Figure 5d shows the spherical structure with a micron size.
With the increase in the amount of oleic acid in the solution, the viscosity of the solution is enhanced, which leads to a low diffusion speed of the oleic molecules. As a result, there are a small number of the oleic acid molecules to adsorb on the WO3-x CNDs. When both the WO3-x NPs approach, there are few of the oleic tails between the WO3-x CNDs. Based on Equation (5), the steric repulsion is weak, which results in the aggregation of the WO3-x CNDs. Simultaneously, some oleic acid molecules adsorbed on the WO3-x NPs are deformed by compression during the aggregation [5]. This further shortens the separation between the WO3-x NPs and improves the attraction interactions. In this case, the WO3-x CNDs are aggregated into the submicro-spherical-like structure, which is confirmed by Figure 5e.

3.5. PL Property

The PL spectra presented in Figure 11 address the PL properties of the as-synthesized WO3-x aggregates. From Figure 11a, one can see that the PL emission from samples B−E exhibits the PL quenching. To confirm the PL bands, a Labspec 5 software was used to fit the PL spectra of samples A, B and D after linear background subtraction. Figure 11a indicates that the PL spectra of samples C and E are similar to the PL spectra of samples B and D, respectively. Here, the PL spectra of samples A, B and D as the typical PL spectra are fitted. The fitted PL peaks are respectively presented in Figure 11b−d, which indicate that the as-synthesized WO3-x aggregates emit ultraviolet (UV), blue, green, red and near-infrared (NIR) light.
In the early work [22], the studies on the PL emission from the WO3-x nanomaterials indicate that the PL emission is related to the band–band transition and the defect level-level transition. The oxygen vacancies in the WO3-x materials form three defect states including the neutral ( V O 0 ), single charged ( V O + ) and doubly charged ( V O 2 + ) states [16]. In the electronic structure, the V O 0 and V O 2 + states form the resonant state levels in the valence and conduction bands, and the V O + state form a level in the bandgap [16]. Therefore, different transitions give rise to different PL emissions.
According to the early studies [22], the UV emission (384−389 nm) is attributed to the recombination between the electron occupying the resonant defect state in the conduction band and the hole in the valence band, and the blue emission (420−433 nm) can be attributed to the band−band transition. For the transitions related to defect levels, the transition from V O + to V O 0 leads to the green emission (566−576 nm) and the transitions from V O 2 + to V O + and V O 2 + to V O 0 generate the red emission (625 and 632 nm), respectively. The green emission (504−512 nm) is related to the V O 0 V O + transition (where * represents the nonrelaxed state). The NIR PL emission (749−753 nm) originates from the relaxation of polarons caused by the coupling of lattice distortion with carries [13].
Compared with Figure 11b, Figure 11c,d exhibit new PL bands at ~680 and 685 nm, which may be related to the coupling between the WO3-x CNDs [5]. The aggregation indicates the occurrence of the coupling between the neighboring WO3-x CNDs depending on their separation. In the solution containing the surfactant, the short carbon chain (~0.6 nm) of the butyric acid molecule easily causes the coupling between the WO3-x CNDs [5,10]. For the oleic acid molecules, the long carbon chain (~2.5 nm) should prevent the coupling [10]. However, their deformation reduces the separation between the WO3-x CNDs [5]. As a result, coupling can occur between the WO3-x CNDs. The coupling can form a new level in the bandgap of the WO3-x materials [5], which generates the PL emission at about 680 and 685 nm.
The PL quenching shown in Figure 11a can originate from the coupling between the WO3-x CNDs and the trapping centers arising from the surfactant molecules. Due to the coupling between the WO3-x CNDs, the charge transfer occurs between the WO3-x CNDs [5], which reduces the recombination probability of the photogenerated electrons and holes to decrease the PL emission. The XPS results indicate that the carboxylic acid molecules coordinate with the WO3-x CNDs in bidentate mode. The bidentate structure can trap the photogenerated electrons to quench PL emission [48]. When the PL emission was measured, the irradiation of laser results in the desorption of the residual acetic acid molecules due to small molecular wight, while the butyric and oleic acid molecules are not easily desorbed. As a result, the trapping centers in sample A are reduced and sample A generates strong PL emission (see Figure 11a). Compared with the PL spectra of samples B and D in Figure 11a, the PL intensity of sample D is weaker than that of sample B. This may be related to their aggregation. From Figure 5, one can see that sample D is composed of the micro-spheres, which imply that there are a number of trapping centers. Thus, Figure 11a shows that the PL emission from sample D is weak compared to sample B. Figure 5c,e indicate that there is a small difference in the morphologies of samples C and E, i.e., they may have almost the same amount of trapping centers. Therefore, the PL intensity of sample C is almost the same as that of sample E.

4. Conclusions

In summary, the WO3-x materials with different structures were synthesized under the butyric or oleic acid surfactants. The results indicate that the WO3-x CNDs can be aggregated into the submicro-spherical-like, submicro-spherical or micro-spherical aggregates depending on the steric repulsion of butyric acid or oleic acid, which are related to their amount employed in the process of syntheses. In other words, the increase in the amount of butyric acid leads to the conversion of submicro-spherical-like to submicro-spherical WO3-x aggregates, while the increased oleic acid causes the change in the micro-spherical to submicro-spherical-like conversion of the WO3-x aggregates. Based on the analyses on the osmotic potential of polymers, the structural conversion of WO3-x aggregates originates from the change in the steric repulsion caused by the surfactant layers, the adsorption and deformation of surfactant molecules on the WO3-x CNDs. Furthermore, the PL properties of the WO3-x aggregates were studied. The WO3-x aggregates generate the UV, blue, green, red and NIR PL emission, which are related to the inter-band transition and defect level-associated transition as well as the relaxation of polarons caused by the coupling of lattice distortion with carries. The PL quenching of the submicro-spherical-like, submicro-spherical and micro-spherical WO3-x aggregates are associated with the coupling between the WO3-x CNDs and the trapping centers originating from the surfactant molecules on the WO3-x CNDs. Our achievements can effectively tune the structure of WO3-based nanomaterials and contribute to the development of next-generation optoelectronic nanodevices.

Author Contributions

Conceptualization, B.W.; Investigation, B.W., X.Z., H.X., Y.Z., U.C. and K.O.; Writing—original draft, B.W.; Writing—review and editing, X.Z., H.X., Y.Z., U.C. and K.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 12175141); Natural Science Foundation of Shanghai (Grant No. 21ZR1435200); National Key Research and Development Program of China (Grant No. 2018YFA0306304); Australian Research Council and QUT Centre for Materials Science; Slovenia-China Bilateral grant BI-CN; and Slovenian Research Agency (ARRS) grant N2-0091.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pan, D.; Fang, Z.; Yang, E.; Ning, Z.; Zhou, Q.; Chen, K.; Zheng, Y.; Zhang, Y.; Shen, Y. Facile preparation of WO3-x dots with remarkably low toxicity and uncompromised activity as coreactants for clinic electrochemiluminescent diagnosis. Angew. Chem. Int. Ed. 2020, 59, 16747–16754. [Google Scholar] [CrossRef] [PubMed]
  2. Han, X.; Li, X.; Wang, R.; Liu, J.; Liu, L. Study on the gamma irradiation characteristics of a carbon nanotube sponge/polydimethylsiloxane/tungsten oxide flexible force-sensitive structure. Micromachines 2022, 13, 1024. [Google Scholar] [CrossRef] [PubMed]
  3. Hou, K.; Han, J.; Tang, Z. Formation of supraparticles and their application in catalysis. ACS Mater. Lett. 2020, 2, 95–106. [Google Scholar] [CrossRef] [Green Version]
  4. Zhao, Q.; Gouget, G.; Guo, J.; Yang, S.; Zhao, T.; Straus, D.B.; Qian, C.; Oh, N.; Wang, H.; Murray, C.B.; et al. Enhanced carrier transport in strongly coupled, epitaxially fused CdSe nanocrystal solids. Nano Lett. 2021, 21, 3318–3324. [Google Scholar] [CrossRef] [PubMed]
  5. Luo, D.; Qin, X.; Song, Q.; Qiao, X.; Zhang, Z.; Xue, Z.; Liu, C.; Mo, G.; Wang, T. Ordered superparticles with an enhanced photoelectric effect by sub-nanometer interparticle distance. Adv. Funct. Mater. 2017, 27, 1701982. [Google Scholar] [CrossRef]
  6. Narayanaswamy, A.; Xu, H.; Pradhan, N.; Peng, X. Crystalline nanoflowers with different chemical compositions and physical properties grown by limited ligand protection. Angew. Chem. 2006, 118, 5487–5490. [Google Scholar] [CrossRef]
  7. Evers, W.H.; Schins, J.M.; Aerts, M.; Kulkarni, A.; Capiod, P.; Berthe, M.; Grandidier, B.; Delerue, C.; van der Zant, H.S.J.; van Overbeek, C.; et al. High charge mobility in two-dimensional percolative networks of PbSe quantum dots connected by atomic bonds. Nat. Commun. 2015, 6, 8195. [Google Scholar] [CrossRef] [Green Version]
  8. Pereira, R.N.; Coutinho, J.; Niesar, S.; Oliveira, T.A.; Aigner, W.; Wiggers, H.; Rayson, M.J.; Briddon, P.; Brandt, M.S.; Stutzmann, M. Resonant electronic coupling enabled by small molecules in nanocrystal solids. Nano Lett. 2014, 14, 3817–3826. [Google Scholar] [CrossRef]
  9. Liu, Y.; Gibbs, M.; Puthussery, J.; Gaik, S.; Ihly, R.; Hillhouse, H.W.; Law, M. Dependence of carrier mobility on nanocrystal size and ligand length in PbSe nanocrystal solids. Nano Lett. 2010, 10, 1960–1969. [Google Scholar] [CrossRef]
  10. Konstantatos, G.; Howard, I.; Fischer, A.; Hoogland, S.; Clifford, J.; Klem, E.; Levina, L.; Sargent, E.H. Ultrasensitive solution-cast quantum dot photodetectors. Nature 2006, 442, 180–183. [Google Scholar] [CrossRef]
  11. Tang, Z.; Kotov, N.A.; Giersig, M. Spontaneous organization of single CdTe nanoparticles into luminescent nanowires. Science 2002, 297, 237–240. [Google Scholar] [CrossRef]
  12. Johansson, M.B.; Zietz, B.; Niklasson, G.A.; Österlund, L. Optical properties of nanocrystalline WO3 and WO3-x thin films prepared by DC magnetron sputtering. J. Appl. Phys. 2014, 115, 213510. [Google Scholar] [CrossRef] [Green Version]
  13. Wang, B.B.; Zhong, X.X.; He, C.L.; Zhang, B.; Cvelbar, U.; Ostrikov, K. Solvent-dependent structures and photoluminescence of WO3-x nanomaterials grown in nonaqueous solutions. J. Alloys Compd. 2021, 854, 157249. [Google Scholar] [CrossRef]
  14. Paik, T.; Cargnello, M.; Gordon, T.R.; Zhang, S.; Yun, H.; Lee, J.D.; Woo, H.Y.; Oh, S.J.; Kagan, C.R.; Fornasiero, P.; et al. Photocatalytic hydrogen evolution from substoichiometric colloidal WO3−x nanowires. ACS Energy Lett. 2018, 3, 1904–1910. [Google Scholar] [CrossRef]
  15. Chen, J.; Yu, D.; Liao, W.; Zheng, M.; Xiao, L.; Zhu, H.; Zhang, M.; Du, M.; Yao, J. WO3−x nanoplates grown on carbon nanofibers for an efficient electrocatalytic hydrogen evolution reaction. ACS Appl. Mater. Interfaces 2016, 8, 18132–18139. [Google Scholar] [CrossRef]
  16. Deb, S.K. Opportunities and challenges in science and technology of WO3 for electrochromic and related applications. Sol. Energy Mater. Sol. Cells 2008, 92, 245–258. [Google Scholar] [CrossRef]
  17. Wang, F.; Valentin, C.D.; Pacchioni, G. Semiconductor-to-metal transition in WO3−x: Nature of the oxygen vacancy. Phys. Rev. B 2011, 84, 073103. [Google Scholar] [CrossRef]
  18. Gerosa, M.; Valentin, C.D.; Onida, G.; Bottani, C.E.; Pacchioni, G. Anisotropic effects of oxygen vacancies on electrochromic properties and conductivity of γ-monoclinic WO3. J. Phys. Chem. C 2016, 120, 11716–11726. [Google Scholar] [CrossRef] [Green Version]
  19. Xiao, X.; Zhou, X.; Ma, J.; Zhu, Y.; Cheng, X.; Luo, W.; Deng, Y. Rational synthesis and gas sensing performance of ordered mesoporous semiconducting WO3/NiO composites. ACS Appl. Mater. Interfaces 2019, 11, 26268–26276. [Google Scholar] [CrossRef]
  20. Duan, G.; Chen, L.; Jing, Z.; Luna, P.D.; Wen, L.; Zhang, L.; Zhao, L.; Xu, J.; Li, Z.; Yang, Z.; et al. Robust antibacterial activity of tungsten oxide (WO3-x) nanodots. Chem. Res. Toxicol. 2019, 32, 1357–1366. [Google Scholar] [CrossRef]
  21. Wang, B.B.; Zhong, X.X.; He, C.L.; Zhang, B.; Shao, R.W.; Shvalya, V.; Cvelbar, U.; Ostrikov, K. Controlling oxygen vacancies of WOx suboxides by ZnWO4 nanophase hybridization. Mater. Sci. Eng. B 2020, 262, 114706. [Google Scholar] [CrossRef]
  22. Wang, B.B.; Zhong, X.X.; He, C.L.; Zhang, B.; Cvelbar, U.; Ostrikov, K. Nanostructure conversion and enhanced photoluminescence of vacancy engineered substoichiometric tungsten oxide nanomaterials. Mater. Chem. Phys. 2021, 262, 124311. [Google Scholar] [CrossRef]
  23. Wang, B.B.; Zhong, X.X.; Zhu, J.; Zhang, Y.; Cvelbar, U.; Ostrikov, K. Single-step synthesis of sub-stoichiometric tungsten oxide particles in mixed acetic and oleic acids: Structural conversion and photoluminescence enhancement. J. Alloys Compd. 2022, 899, 163265. [Google Scholar] [CrossRef]
  24. Wang, B.; Zhong, X.; Zhang, Y.; Xu, H.; Zhang, Y. Microspheres assembled by WO3-x nanoparticles under action of dual surfactants: Structure and photoluminescence properties. Opt. Mater. 2022, 129, 112550. [Google Scholar] [CrossRef]
  25. Piacenza, E.; Presentato, A.; Turner, R.J. Stability of biogenic metal(loid) nanomaterials related to the colloidal stabilization theory of chemical nanostructures. Crit. Rev. Biotechnol. 2018, 38, 1137–1156. [Google Scholar] [CrossRef]
  26. Fulton, G.; Lunev, A. Probing the correlation between phase evolution and growth kinetics in the oxide layers of tungsten using Raman spectroscopy and EBSD. Corros. Sci. 2020, 162, 108221. [Google Scholar] [CrossRef]
  27. Daniel, M.F.; Desbat, B.; Lassegues, J.C. Infrared and Raman study of WO3 tungsten trioxides and WO3, xH2O tungsten trioxide hydrates. J. Solid State Chem. 1987, 67, 235–247. [Google Scholar] [CrossRef]
  28. Garcia-Sanchez, R.F.; Ahmido, T.; Casimir, D.; Baliga, S.; Misra, P. Thermal effects associated with the Raman spectroscopy of WO3 gas-sensor materials. J. Phys. Chem. A 2013, 117, 13825–13831. [Google Scholar] [CrossRef]
  29. Blackman, C.S.; Parkin, I.P. Atmospheric pressure chemical vapor deposition of crystalline monoclinic WO3 and WO3-x thin films from reaction of WCl6 with O-containing solvents and their photochromic and electrochromic properties. Chem. Mater. 2005, 17, 1583–1590. [Google Scholar] [CrossRef]
  30. Yu, W.; Shen, Z.; Peng, F.; Lu, Y.; Ge, M.; Fu, X.; Sun, Y.; Chen, X.; Dai, N. Improving gas sensing performance by oxygen vacancies in sub-stoichiometric WO3-x. RSC Adv. 2019, 9, 7723–7728. [Google Scholar] [CrossRef]
  31. Wagner, C.D.; Riggs, W.M.; Davis, L.E.; Moulder, J.F.; Muilenberg, G.E. Handbook of X-ray Photoelectron Spectroscopy; Perkin-Elmer Corp., Physical Electronics Division: Eden Prairie, MI, USA, 1979; Volume 38, pp. 147–184. [Google Scholar]
  32. He, Y.; Liang, A.; Zhu, D.; Hu, M.; Xu, L.; Chao, S.; Zhou, W.; Wu, Y.; Xu, J.; Zhao, F. Organic-inorganic hybrid electrode engineering for high-performance asymmetric supercapacitor based on WO3-CeO2 nanowires with oxygen vacancies. Appl. Surf. Sci. 2022, 573, 151624. [Google Scholar] [CrossRef]
  33. Zhang, N.; Li, F.; Yin, Y.; Han, J.; Li, X.; Liu, C.; Zhou, J.; Wen, S.; Adimi, S.; Chen, Y.; et al. Gas sensor based on TiO2 nanofibers decorated with monodispersed WO3 nanocubes for fast and selective xylene detection. Mater. Sci. Eng. B 2021, 263, 114901. [Google Scholar] [CrossRef]
  34. Wilson, D.; Langell, M.A. XPS analysis of oleylamine/oleic acid capped Fe3O4 nanoparticles as a function of temperature. Appl. Surf. Sci. 2014, 303, 6–13. [Google Scholar] [CrossRef]
  35. Mahdavi, M.; Ahmad, M.B.; Haron, M.J.; Namvar, F.; Nadi, B.; Rahman, M.Z.A.; Amin, J. Synthesis, surface modification and characterisation of biocompatible magnetic iron oxide nanoparticles for biomedical applications. Molecules 2013, 18, 7533–7548. [Google Scholar] [CrossRef] [Green Version]
  36. Chen, S.-H.; Yu, C.-F.; Wang, C.-J.; Chen, S.-H.; Chen, Y.-D.; Chen, T.-C.; Lin, C.-F. Light enhancement of plasmonic nano-structure for PLEDs at RGB wavelengths. Org. Electron. 2016, 38, 337–343. [Google Scholar] [CrossRef]
  37. Chen, S.-H.; Hsu, P.-J. Luminous efficiency improvement of polymer light-emitting diodes with platinum nanolayer at the PEDOT:PSS–ITO interface. Opt. Lett. 2021, 46, 6039–6042. [Google Scholar] [CrossRef]
  38. Segets, D.; Marczak, R.; Schäfer, S.; Paula, C.; Gnichwitz, J.-F.; Hirsch, A.; Peukert, W. Experimental and theoretical studies of the colloidal stability of nanoparticles―A general interpretation based on stability maps. ACS Nano 2011, 5, 4658–4669. [Google Scholar] [CrossRef]
  39. Hotze, E.M.; Phenrat, T.; Lowry, G.V. Nanoparticle aggregation: Challenges to understanding transport and reactivity in the environment. J. Environ. Qual. 2010, 39, 1909–1924. [Google Scholar] [CrossRef] [Green Version]
  40. Patil, V.B.; Adhyapak, P.V.; Suryavanshi, S.S.; Mulla, I.S. Oxalic acid induced hydrothermal synthesis of single crystalline tungsten oxide nanorods. J. Alloys Compd. 2014, 590, 283–288. [Google Scholar] [CrossRef]
  41. Vayssieres, L. On the design of advanced metal oxide nanomaterials. Int. J. Nanotechnol. 2004, 1, 1–41. [Google Scholar] [CrossRef]
  42. Oswal, S.L.; Desai, H.S. Studies of viscosity and excess molar volume of binary mixtures. 1. Propylamineq1-alkanol mixtures at 303.15 and 313.15 K. Fluid Phase Equilibria 1998, 149, 359–376. [Google Scholar] [CrossRef]
  43. Shen, Y.; Liu, M. Viscosity of organic compounds. J. South-Cent. Coll. Natl. (Nat. Sci.) 1997, 16, 5–10. (In Chinese) [Google Scholar]
  44. Sagdeev, D.; Gabitov, I.; Isyanov, C.; Khairutdinov, V.; Farakhov, M.; Zaripov, Z.; Abdulagatov, I. Densities and viscosities of oleic acid at atmospheric pressure. J. Am. Oil Chem. Soc. 2019, 96, 647–662. [Google Scholar] [CrossRef]
  45. Johes, R.A.L. Soft Condensed Matter; Oxford University Press: Oxford, England, 2002; p. 60. [Google Scholar]
  46. Fritz, G.; Schädler, V.; Willenbacher, N.; Wagner, N.J. Electrosteric stabilization of colloidal dispersions. Langmuir 2002, 18, 6381–6390. [Google Scholar] [CrossRef]
  47. Kotov, N.K. Practica l aspects of self-organization of nanoparticles: Experimental guide and future applications. J. Mater. Chem. 2011, 21, 16673–16674. [Google Scholar] [CrossRef]
  48. Sakohara, S.; Tickanen, L.D.; Anderson, M.A. Luminescence properties of thin zinc oxide membranes prepared by the Sol-Gel technique: Change in visible luminescence during firing. J. Phys. Chem. 1992, 96, 11086–11091. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of as-synthesized WO3-x aggregates of (a) sample A and (b) samples B–E respectively, which indicate that they are composed of mixed phases.
Figure 1. XRD patterns of as-synthesized WO3-x aggregates of (a) sample A and (b) samples B–E respectively, which indicate that they are composed of mixed phases.
Micromachines 13 02075 g001
Figure 2. Raman spectra (a) and deconvoluted Raman spectra (bf) of as-synthesized WO3-x aggregates. The data in the insets are the positions corresponding to Raman peaks.
Figure 2. Raman spectra (a) and deconvoluted Raman spectra (bf) of as-synthesized WO3-x aggregates. The data in the insets are the positions corresponding to Raman peaks.
Micromachines 13 02075 g002
Figure 3. XPS survey spectra of samples A−D. The XPS peaks are assigned according to Ref. [31]. These XPS peaks indicate that the as-synthesized aggregates contain the W, O and C elements.
Figure 3. XPS survey spectra of samples A−D. The XPS peaks are assigned according to Ref. [31]. These XPS peaks indicate that the as-synthesized aggregates contain the W, O and C elements.
Micromachines 13 02075 g003
Figure 4. High-resolution W 4f, O 1s and C 1s XPS spectra of as-synthesized WO3-x aggregates. The deconvoluted peaks indicate the presence of oxygen vacancies and the absorption of carboxylic acid molecules.
Figure 4. High-resolution W 4f, O 1s and C 1s XPS spectra of as-synthesized WO3-x aggregates. The deconvoluted peaks indicate the presence of oxygen vacancies and the absorption of carboxylic acid molecules.
Micromachines 13 02075 g004
Figure 5. FESEM images of samples A−E. These images indicate that the as-synthesized WO3-x materials are the mixture of (a) nanoflakes and aggregated nanodots, (b,e) the submicro-spherical-like, (c) submicro-spherical and (d) micro-spherical structures mixed with the aggregated nanoparticles.
Figure 5. FESEM images of samples A−E. These images indicate that the as-synthesized WO3-x materials are the mixture of (a) nanoflakes and aggregated nanodots, (b,e) the submicro-spherical-like, (c) submicro-spherical and (d) micro-spherical structures mixed with the aggregated nanoparticles.
Micromachines 13 02075 g005
Figure 6. TEM images of sample A, which indicate that the nanoflake (a), nanorod (b) and nanoparticle (c) are the crystal structure with the lattice spacing of ~0.376, 0.369 and 0.383 nm, respectively.
Figure 6. TEM images of sample A, which indicate that the nanoflake (a), nanorod (b) and nanoparticle (c) are the crystal structure with the lattice spacing of ~0.376, 0.369 and 0.383 nm, respectively.
Micromachines 13 02075 g006
Figure 7. TEM images of sample B. The image (a) indicates that sample B is the submicro-spherical-like structure mixed some nanorods and the submicro-spherical-like structure is aggregated by nanoparticles. The high-resolution TEM images indicate that the nanorod (b) and nanoparticle (c) are the crystalline structure with the lattice spacing of ~0.378 and 0.380 nm, respectively.
Figure 7. TEM images of sample B. The image (a) indicates that sample B is the submicro-spherical-like structure mixed some nanorods and the submicro-spherical-like structure is aggregated by nanoparticles. The high-resolution TEM images indicate that the nanorod (b) and nanoparticle (c) are the crystalline structure with the lattice spacing of ~0.378 and 0.380 nm, respectively.
Micromachines 13 02075 g007
Figure 8. TEM images of sample D. Image (a) indicates that the edge is composed of nanoparticles. The high-resolution TEM image (b) indicates that the nanoparticle is the crystalline structure with a lattice spacing of ~0.389 nm.
Figure 8. TEM images of sample D. Image (a) indicates that the edge is composed of nanoparticles. The high-resolution TEM image (b) indicates that the nanoparticle is the crystalline structure with a lattice spacing of ~0.389 nm.
Micromachines 13 02075 g008
Figure 9. Schematic diagram of steric repulsion produced by the surfactant layers on the WO3-x CNDs. Adapted with modification from Refs. [25,45].
Figure 9. Schematic diagram of steric repulsion produced by the surfactant layers on the WO3-x CNDs. Adapted with modification from Refs. [25,45].
Micromachines 13 02075 g009
Figure 10. Aggregation schematic of spherical-like and spherical WO3-x aggregates. (a) irregular structure; (b) submicro-spherical-like structure; (c) submicro- or micro-spherical structure.
Figure 10. Aggregation schematic of spherical-like and spherical WO3-x aggregates. (a) irregular structure; (b) submicro-spherical-like structure; (c) submicro- or micro-spherical structure.
Micromachines 13 02075 g010
Figure 11. (a) PL spectra of samples A–E and (bd) deconvoluted PL peaks for the PL spectra of samples A, B and D. The data in the images (bd) correspond to the wave lengths of PL emissions.
Figure 11. (a) PL spectra of samples A–E and (bd) deconvoluted PL peaks for the PL spectra of samples A, B and D. The data in the images (bd) correspond to the wave lengths of PL emissions.
Micromachines 13 02075 g011
Table 1. Mass of WCl6, volume of carboxylic acids, growth temperature (T), and time (t).
Table 1. Mass of WCl6, volume of carboxylic acids, growth temperature (T), and time (t).
SampleWCl6 (g)Acetic Acid (mL)Butyric Acid (mL)Oleic Acid (mL)T (°C)t (h)
A2.0050--12012
B2.003020-12012
C2.002030-12012
D2.0030-2012012
E2.0020-3012012
Table 2. Phases responding to the XRD peaks in Figure 1a,b.
Table 2. Phases responding to the XRD peaks in Figure 1a,b.
PhasePeakJCPDS
A1: W19O55(-105); A2: W19O55(302); A3: W19O55(-2011) 20.5°, 24.0°, 45.9°45-0167
B1: WO2.92(-1012); B2: WO2.92(-1016); B3: WO2.92(-2128)18.5°, 24.0°, 49.8°30-1387
C1: W18O49(-301); C2: W18O49(102); C3: W18O49(103)14.5°, 16.6°, 24.0°05-0392
C4: W18O49(302); C5: W18O49(012); C6: W18O49(211)25.6°, 27.4°, 28.1°
C7: W18O49(501); C8: W18O49(502); C9: W18O49(-704)30.6°, 35.8°, 36.6°
C10: W18O49(021); C11: W18O49(123); C12: W18O49(-523)49.8°, 54.4°, 55.7°
C13: W18O49(017); C14: W18O49(-1015)56.3°, 57.3°
Table 3. Assignment of Raman peaks in Figure 2.
Table 3. Assignment of Raman peaks in Figure 2.
Raman Peak (cm−1)Assignment *References
196, 198Lattice vibration[26]
205, 254ν(W−O−W)[27]
272, 278, 281δ(O−W−O)[27,28]
660, 663, 679, 680, 682, 694ν(O−W−O)[27]
745, 756, 764, 773ν(W−O)[13,28]
801ν(O−W−O)[28]
943, 944, 948ν(W=O)[13,28]
* δ and ν represent the bending and stretching vibration modes, respectively.
Table 4. Binding energy (eV) of deconvoluted XPS peaks in Figure 4.
Table 4. Binding energy (eV) of deconvoluted XPS peaks in Figure 4.
SampleW5+ 4f7/2W6+ 4f7/2W5+ 4f5/2W6+ 4f5/2OLOVOC(CH2)nCH2OO=C-O
A34.735.536.137.8530.4531.5532.8284.7285.8289.0
B34.835.636.337.8530.4531.7533.1284.7285.6288.9
C34.735.636.337.8530.4531.7533.1284.7285.5288.9
D34.935.636.237.8530.5531.7533.0284.7285.6288.9
E34.735.636.237.8530.3531.3532.7284.7285.5288.9
Table 5. Areas of OL, OV and OC peaks and the area ratio R of OV peak to the area sum of OL and OV peaks.
Table 5. Areas of OL, OV and OC peaks and the area ratio R of OV peak to the area sum of OL and OV peaks.
SampleOL Peak AreaOV Peak AreaOC Peak AreaRatio (%)
A50,97017,02511,23025.0
B39,68516,63712,06229.5
C52,01317,02510,64924.7
D54,68416,90111,86623.6
E47,57620,92614,30130.5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, B.; Zhong, X.; Xu, H.; Zhang, Y.; Cvelbar, U.; Ostrikov, K. Structure and Photoluminescence of WO3-x Aggregates Tuned by Surfactants. Micromachines 2022, 13, 2075. https://doi.org/10.3390/mi13122075

AMA Style

Wang B, Zhong X, Xu H, Zhang Y, Cvelbar U, Ostrikov K. Structure and Photoluminescence of WO3-x Aggregates Tuned by Surfactants. Micromachines. 2022; 13(12):2075. https://doi.org/10.3390/mi13122075

Chicago/Turabian Style

Wang, Biben, Xiaoxia Zhong, Haiyan Xu, Yongcai Zhang, Uros Cvelbar, and Kostya (Ken) Ostrikov. 2022. "Structure and Photoluminescence of WO3-x Aggregates Tuned by Surfactants" Micromachines 13, no. 12: 2075. https://doi.org/10.3390/mi13122075

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop