Large-area fabrication of TiN thin films with photothermal effect via PECVD

Large-area fabrication of TiN thin films with photothermal effect via PECVD

Ceramics International xxx (xxxx) xxx–xxx Contents lists available at ScienceDirect Ceramics International journal homepage: www.elsevier.com/locate...

3MB Sizes 0 Downloads 12 Views

Ceramics International xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Ceramics International journal homepage: www.elsevier.com/locate/ceramint

Large-area fabrication of TiN thin films with photothermal effect via PECVD Wanyin Gea,∗, Zhe Changa, Awais Siddiquea, Bo Shia, Chun Liub a School of Materials Science and Engineering, Shaanxi Key Laboratory of Green Preparation and Functionalization for Inorganic Materials, Shaanxi University of Science and Technology, Xi'an, 710021, China b Xianyang Research & Design Institute of Ceramics Xianyang, 710000, China

A R T I C LE I N FO

A B S T R A C T

Keywords: TiN films Transition metal nitride PECVD Photothermal therapy

Titanium nitride alloy has triggered extensive interests for the actual application in aeronautics and astronautics, especially in the biological window serving as cancer therapy and tumor detection due to broad band absorption. However, developing a facile and reliable method for the manufacture of TiN films is still a propelled and demanding challenge. Here we present a one-step approach to obtain cubic-TiN films using plasma-enhanced chemical vapor deposition (PECVD) for the first time. Large-scale TiN films can be deposited on sophisticated substrate (e.g. flattened and curled shape) with controllable size by our PECVD approach. The chemical composition and structural analysis were conducted by X-ray diffraction, scanning electron microscope, transmission electron microscopy, and X-ray photoelectron spectroscopy. Furthermore, the TiN powders extracted from our films for photothermal effect demonstrated an excellent photothermal conversion efficiency of 47.9% under 808 nm irradiation. Our study not only offers a facile PECVD route to realize the TiN films on sophisticated substrates, but also provides a promising candidate for photothermal therapy in the future.

1. Introduction The local electromagnetic field near nanoparticles enhanced by localized surface plasmon resonance (LSPR) has aroused great interest owing to its unique application in biological imaging [1] photothermal therapy [2], and optoelectronic devices and so on [3,4]. Among them, photothermal therapy (PTT) has been considered as one of the most promising cancer treatment strategy over the last few decades due to high specificity, minimal invasiveness and fewer complications [5,6]. Up to now, the research on PTT has mainly focused on noble metals (Au [7], Ag [8,9]), metal chalcogenides [10] and carbon nanotubes [11]. However, the drawback for classic noble metals nanoparticles is their LSPR locating in the range of visible region (400–500 nm), which is naturally difficult to redshift to the biological windows (650–1350 nm) [12,13]. It is commonly known that visible light is hardly penetrating to biological tissue, which has been extremely limited the PTT application [14]. To solve this critical issue, TiN, a high biocompatible material, has received great interest due to its suitable LSPR region located in the biological window [15–19]. For example, TiN showed that near the unity absorbance is achieved at around 650 nm [20]. In addition, TiN NPs exhibit a wide absorption from 650 nm to near-infrared region. These strong absorption result from its LSPR effect of TiN NPs [21]. To satisfy the urgent need for PTT, the synthesis strategies of TiN



have been intensively investigated, including magnetron sputter [22,23], electron beam evaporation [24], laser CVD method [25], direct nitridation of titanium metal [26], and chemical vapor deposition (CVD) [27–29]. Among them, CVD approach was proved to be convenient because of its simplicity and scalability. However, this method requires two-step reaction [30], high temperature (1200 °C) [31] or complex reaction condition (employing NH3 as nitrogen source) [30,31] to obtain TiN films or particles. Currently, it is still a significant challenge to rationally design a simple and effective route for the production of high-purified TiN films. In addition, the deposition of TiN films was reported widely on flat or some special substrates. Shankernath employed gold-electroplated brass as substrate and obtained nanostructured TiN films [22]. To address these issues, we herein report a one-step approach for the first time to realize the large-area and highquality TiN films by plasma-enhanced chemical vapor deposition (PECVD). The TiN films can be deposited on sophisticated substrate (e.g. flattened and curled shape) with controllable size by our approach. Furthermore, the TiN powders extracted from our films for PTT revealed an excellent photothermal conversion efficiency, comparing with the maximum one of Au. Our study not only offers a facile PECVD route to realize the TiN films on sophisticated substrate, but also provides a promising candidate for PTT in biomedical applications.

Corresponding author. E-mail address: [email protected] (W. Ge).

https://doi.org/10.1016/j.ceramint.2019.11.231 Received 8 October 2019; Received in revised form 20 November 2019; Accepted 26 November 2019 0272-8842/ © 2019 Elsevier Ltd and Techna Group S.r.l. All rights reserved.

Please cite this article as: Wanyin Ge, et al., Ceramics International, https://doi.org/10.1016/j.ceramint.2019.11.231

Ceramics International xxx (xxxx) xxx–xxx

W. Ge, et al.

Fig. 1. (a) XRD patterns of TiN nanoparticles prepared by PECVD at various temperatures; (b) XPS spectra of the TiN nanoparticles grown at 800 °C and 1000 °C; (c, d) high-resolution XPS of Ti 2p and N 1s in the TiN nanoparticles grown at 1000 °C.

Fig. 2. (a) Schematic illustration of the PECVD setup, (b) the growth mechanism of TiN films on quartz substrates; (c, d) Photographs of TiN films grown on the flattened and curled shape quartz.

2

Ceramics International xxx (xxxx) xxx–xxx

W. Ge, et al.

then AgNO3 solution (20 mL, 0.05 mol/L) was added. Finally, the mixture solution was transferred to a polytetrafluoroethylene for 6 h at 140 °C, and the crude product was washed many times with distilled water and dried at 40 °C.

Table 1 Estimation of Ti, N, and O atomic concentration of the TiN nanoparticles grown at 800 °C and 1000 °C by XPS. Temperature (°C)

800 1000

Atomic concentration (%) Ti 2p

N 1s

O 1s

15.51 18.91

10.72 13.11

30.65 33.53

2.4. Material characterizations XRD patterns were recorded on a DX-2700BH diffractometer (Dandong Haoyuan Instrument Co. Ltd.) using Cu Kα radiation (λ = 1.5406 Å, 40 kV, 30 mA). The morphology images were collected by a scanning electron microscopy (SEM, Hitachi S-4800) equipped with an energy dispersive X-ray (EDS). TEM were carried out on the FEI Tecnai G2 F20 S, operating at 200 kV. X-ray photoelectron (XPS) spectra were recorded on an ESCALAB 250XI (Thermo Scientific) spectrometer. UV–Vis–NIR absorption spectra was conducted on a Lambda 25 spectrophotometer. The photothermal images were monitored through an infrared thermal camera (FOTRIC).

2. Experimental 2.1. Chemicals Ammonium chloride (NH4Cl), Titanium powder (Ti), sodium hydroxide (NaOH), Glucose (C6H12O6), silver nitrate (AgNO3), and Polyvinylpyrrolidone (PVP) were received as chemical pure without any purification.

2.5. Photothermal effect test of TiN nanoparticles 2.2. Deposition of TiN films The dispersions aqueous solution of as-prepared TiN powders extracted from films with different concentrations were irradiated under NIR laser (808 nm, power density of 0.5 W/cm2 and 1 W/cm2) for 15 min. Meanwhile, the temperature monitor of TiN solution was recorded by digital thermometer, and their photothermal images were acquired by an infrared thermal camera. The photostability of TiN solution and Ag nanoparticles were further investigated by using four laser ON/OFF cycles exposed to an 808 nm laser.

A two-temperature-zone furnace equipped with a 30 mm diameter quartz tube was used for the PECVD growth. Ti/NH4Cl powders mixture were placed in the upstream zone set at 700 °C, and the quartz substrate (flattened and curled shape) were placed in the downstream zone set at 600–1000 °C. Meanwhile, the PECVD system was cleaned with Ar gas (500 sccm) for 10 min to remove residual oxygen. Until the temperature was reached to 1000 °C, turned on the plasma for 20 min. Meanwhile, the Ar/H2 carrier gas with 1:1 ratio was introduced at a flow rate (40 sccm), and the pressure of PECVD system was maintained at 70 Pa. Finally, the furnace was cooled down naturally under Ar/H2 flow of 200 sccm.

3. Results and discussion The PECVD growth was carried out in a two-temperature-zones furnace. The Ti/NH4Cl mixed powders and the quartz substrates were placed in the upstream and downstream zones of the furnace, respectively. The flattened and curled shape quartz substrates were employed for deposition of TiN films. Fig. 2a shows XRD patterns of the films prepared by PECVD at different temperatures (600–1000 °C). The characteristic diffraction peaks located at 36.68°, 42.61°, 61.84°,

2.3. Preparation of Ag nanoparticles PVP (1 g) was dissolved into distilled water (20 mL), then NaOH (10 mL, 0.05 mol/L) solution and glucose solution (10 mL, 0.2 mol/L) were added, stirring well with a glass rod to make it evenly dispersed,

Fig. 3. SEM images of TiN nanoparticles grown at different temperatures, (a) 700, (b) 800, (c) 900, (d) 1000 °C. The insert images in (a–d) are the size distribution histogram. 3

Ceramics International xxx (xxxx) xxx–xxx

W. Ge, et al.

Fig. 4. (a) TEM and (b) HRTEM image of TiN nanoparticles grown at 1000 °C; (c–f) element mapping images of TiN for a selected area.

critical factor for the formation of films, since no films deposited on the substrate without the plasma. The growth mechanism of the TiN films was presented in Fig. 2b. During the PECVD process, NH4Cl was primarily decomposed into NH3 (g) and HCl (g). Subsequently, Ti reacted with HCl gas to form volatile TiClx (g). Finally, TiClx (g) reacted with NH3 (g) to form TiN. The whole PECVD reaction process as follows: NH4Cl(s)→NH3(g)+HCl(g) 2Ti(s)+2xHCl(g)→2TiClx(g)+xH2(g) 2TiClx(g)+(x-4)H2(g)+2NH3(g)→2TiN(s)+2xHCl(g) At the elevated temperature, the hydrogen provided a reducing atmosphere for the whole growth. Although the growth involves the decomposed gas such as NH3 (g) and HCl (g), the amount of them is limited compared to the reported CVD route employed persistent NH3 flow. Interestingly, the dazzling golden TiN films were obtained by PECVD for the first time. More specially, with this method, not only a film can be formed on a flattened substrate, but also a film can be formed on a curled substrate (Fig. 2c and d), which provides a possibility to obtain complex matrix films. Therefore, we herein offered a one-step PECVD route to obtain TiN films on complex substrate. The microstructure of the TiN films was characterized by SEM. Fig. 3(a–d) shows SEM images of TiN films obtained in the temperature range of 700–1000 °C. The average size of TiN films varies from 20 nm to 60 nm, showing the temperature dependence feature. Fig. 4a shows a typical transmission electron microscopy (TEM) image of the synthesized TiN alloy, revealing a catenoid morphology formed by nanoparticles. High-resolution transmission electron microscopy (HRTEM) was further used to characterize the TiN microscopic structure (Fig. 4b). The TiN sample revealed a lattice fringe of 0.21 nm, in consistent with the (220) planes of the face-centered cubic TiN. This result is in line with the XRD results. The compositional distribution of the nanocrystals was further investigated by EDS. We found the chemical molar ratio of Ti and N is very close to 1:1, revealing the stoichiometric TiN compound. Fig. 4c is a selected region of TiN films used for elemental mapping analysis. Fig. 4(d–f) shows primarily the elemental composition with Ti, N and O elements. The O element is trace, originating from oxygen absorption on the surface. The uniform element distribution revealed the high degree of alloying between Ti and N. The absorption spectra of TiN films prepared at different temperatures were studied by UV–Vis–NIR spectroscopy (Fig. 5). It reveals that TiN nanoparticles present a wide absorption peak centered at 590 nm and extend to near-infrared region, which is consistent with the

Fig. 5. UV–Vis–NIR spectra of TiN nanoparticles grown at different temperature.

74.10°, and 78.00°, corresponding to (111), (200), (220), (311) and (222) of cubic-TiN (PDF: 04-004-6867, a = b = c = 4.24 Å). No impurity diffraction peaks were observed, confirming the high purity of the products. However, the stray peaks appear when temperatures are below 1000 °C. With the increase of growth temperature, the intensity of TiN diffraction peaks increased, and the half peak gradually narrowed. The diffraction peaks of TiN at 1000 °C were sharp, indicating the high crystalline at higher growth temperature. Our results revealed that high crystalline TiN alloy obtained by PECVD route. Surface analysis of sample was further carried out by XPS (Fig. 1b). It can be seen from the XPS that the products at different growth temperatures contain Ti, N and O elements, revealing the chemical composition of the as-synthesized sample. The ratio of Ti:N gradually increases, when the temperature raised from 800 °C to 1000 °C, as shown in Table 1. This fact indicated that the purity of the formed TiN films gradually improves, which was in good consistent with the XRD analysis. The high resolution Ti 2p and N 1s spectrums of TiN nanoparticles prepared at 1000 °C are shown in Fig. 1c and d. The Ti 2p spectrum (Fig. 1c) is mainly deconvoluted into three components of TiN (454.99), TiO or TiO2-xNx (456.24) and TiO2 (458.83) [32]. The N 1s spectrum (Fig. 1d) exhibits three contributions, locating at 396.27, 397.37 eV and 398.81 eV, respectively. Notably, the N 1s peak at 398.81 eV can be attributed to interstitial N atoms or chemisorbed N-containing gas (NH3 or N2). The N 1s peaks at 396.27 and 397.37 eV are assigned to β−N in the Ti–N bond or N substituted at oxygen sites [33,34]. Fig. 2a shows the schematic of our PECVD setup. Argon was employed and served as protection for TiN growth. The plasma is the most 4

Ceramics International xxx (xxxx) xxx–xxx

W. Ge, et al.

Fig. 6. Temperature elevation of various concentrations TiN solution under laser irradiation (a) 0.5 W/cm2 and (b) 1 W/cm2, (c) Temperature increment of TiN solution vary with concentration and laser power density, (d) Temperature elevation of Ag nanoparticles and TiN nanoparticles (grown at different temperatures).

that the photothermal property of TiN nanoparticle had a strong dependence of its concentration. In addition, the TiN solution with a concentration of 1 mg/mL increased from the initial 24.41–51.1 °C in 15 min. It was reported that the tumor cells will be destroyed once the temperature over 42 °C [35]. Therefore, our TiN solution satisfies the requirements of photothermal therapy. Fig. 6d demonstrates TiN nanoparticles have stronger heat-generating capacity for higher grown temperature, revealing the crystallinity dependence. As a comparative study, Ag nanoparticles with the same concentration were tested. The photothermal property of the Ag nanoparticles is lower than that of the TiN nanoparticles, suggesting TiN nanoparticles possess the advantages of wide spectral absorption and excellent photothermal effect. Meanwhile, visible analysis for the photothermal property of TiN solution was conducted by an infrared thermal camera, providing more detailed information. (Fig. 7). According to the chromatic temperature chart, the heated color transfers gradually from black to high-light color when the temperature of the TiN solution increasing. The higher temperature of the TiN solution or the longer illumination time, the brighter color. Again, this visible analysis verify temperature variations of TiN is time-dependent and concentration-dependent, showing the similar result in Fig. 6. To further evaluate the photothermal property, we calculated the photothermal conversion efficiency (η ) of the TiN solution according to the following formula [36]:

Fig. 7. Infrared visible images of the TiN solution under laser irradiation (808 nm, 1 W/cm2).

reported plasmon resonance frequency of TiN [20,21]. Therefore, the TiN nanoparticles are expected to be a near-infrared photothermal agent for PTT. To verify the photothermal effect of TiN nanoparticles, TiN nanoparticles were irradiated for 15 min under 808 nm laser excitation. The photothermal temperature of the sample was monitored by an infrared camera simultaneously. Fig. 6 (a, b) shows the photothermal temperature rise behavior of the samples with different concentration at different irradiation power (0.5 W/cm2 and 1 W/cm2). From the floating bars (Fig. 6c), we can observe that the temperature of TiN solution increases gradually with the higher concentration, indicating

η=

hS (Tmax − Tsurr ) − Qdis I (1 − 10−A808)

(1)

here h is the heat transfer coefficient; S is the surface area for radiative heat transfer; Tmax is the highest equilibrium temperature; Tsurr is the ambient temperature; A808 is the absorbance of the sample at 808 nm; I is the laser power (1 W/cm2). Qdis presents the amount of heat absorbed by the solvent and container. The value of hS can be calculated using the following formula:

5

Ceramics International xxx (xxxx) xxx–xxx

W. Ge, et al.

Fig. 8. (a) Photothermal behavior of TiN solution response to NIR laser (808 nm, 1 W/cm2) under an on/off period; (b) Linear fitting curve between time and  − lnθ ; (c) Photothermal elevation of TiN solution (1.0 mg/mL) over four on/off cycles under NIR laser irradiation; (d) Comparative temperature variations of the TiN solution and Ag nanoparticles under laser irradiation for four cycles.

In order to assess the photothermal stability of TiN nanoparticles, cyclic heating experiments were carried out using an 808 nm laser with power density of 1 W/cm2, as shown in Fig. 8c and d. Photothermal elevation of TiN solution (1.0 mg/mL) exhibited an excellent cycling stability over four on/off cycles under NIR laser irradiation. By contrast, the cyclic heating experiments of Ag nanoparticles were also performed. Obviously, Ag nanoparticles show a slightly fluctuation four on/off cycles under NIR laser irradiation, while the cycle of temperature variations for TiN solution presents a better stability than that of Ag nanoparticles. Therefore, our results indicated outstanding photostability of TiN solution.

Table 2 Photothermal conversion efficacy of several reported photothermal agents. Photothermal agents

Photothermal conversion efficacy

Wavelength

Refs.

TiN

47.9%

808

Ag nanoparticles

40.1%

808

TiN nanoparticles Au nanorods Ag nanocages Au/Ag nanoshells WS2 nanosheets Cu9S5 nanocrystals Ta2O5 coatings

44.6% 48.5% 46.1% 41.6% 32.8% 25.7% 30.8%

808 808 808 808 808 980 808

This work This work [21] [21] [37] [38] [39] [40] [41]

hS =

4. Conclusions In summary, we have developed a successful one-step approach to fabricate cubic-TiN films on the flattened and curled shape quartz substrate. The chemical composition and structural analysis were conducted by XRD, SEM, TEM and XPS. The TiN nanoparticles solution exhibit excellent photothermal effect owing to its strong LSPR absorption. The photothermal conversion efficiency of 47.9% was obtained under 808 nm irradiation. The increasing high temperature (51.1 °C) shows the potential superiority to destroy the tumor cells. The related work highlight the TiN nanoparticles could be possibly applied to photothermal therapy.

∑i mi Cp, i τs

(2)

here, C and m are the heat capacity (4.2 J/g﹒°C) and the mass of water respectively; τs is the time constant of the thermal equilibrium, which can be deduced by the linear relationship between the time and − lnθ obtained from the cooling curve, according to following equations.

T = −τs lnθ θ=

T − Tsurr Tmax − Tsurr

(3)

(4) Declaration of competing interest

Fig. 8a plots the heating and cooling process of the TiN solution. A linear fit is performed on the cooling data to determine the time constant (τs ) (Fig. 8b), revealing an average value of 344.8 s. Hence, the photothermal conversion efficiency of TiN nanoparticle was calculated to be 47.9% by formula (1). According to numerous reported literatures (Table 2), the photothermal conversion efficiency of our TiN nanoparticles are higher than the most of common photothermal agents.

The authors declared that they have no conflicts of interest to this work. We declare that we do not have any commercial or associative interest that represents a conflict of interest in connection with the work submitted.

6

Ceramics International xxx (xxxx) xxx–xxx

W. Ge, et al.

Acknowledgments

(2018) 2869–2876. [19] U. Guler, A. Boltasseva, V.M. Shalaev, Applied physics. Refractory plasmonics, Science 344 (2014) 263–264. [20] W. Li, U. Guler, N. Kinsey, G.V. Naik, A. Boltasseva, J.G. Guan, V.M. Shalaev, A.V. Kildishev, Refractory plasmonics with titanium nitride: broadband metamaterial absorber, Adv. Mater. 26 (2014) 7921-7921. [21] W.Q. Jiang, Q.G. Fu, H.Y. Wei, A.H. Yao, TiN nanoparticles: synthesis and application as near-infrared photothermal agents for cancer therapy, J. Mater. Sci. 54 (2019) 5743–5756. [22] V. Shankernath, K.L. Naidu, M.G. Krishna, K. Padmanabhan, Optical response of ultra-thin titanium nitride films on brass and gold plated brass surfaces, Mater. Res. Bull. 85 (2017) 121–130. [23] F. Solis-Pomar, O. Nápoles, O.V. Robaina, C. Gutierrez-Lazos, A. Fundora, A. Colin, E. Pérez-Tijerina, M. Melendrez, Preparation and characterization of nanostructured titanium nitride thin films at room temperature, Ceram. Int. 42 (2016) 7571–7575. [24] F. Arezzo, P. Gimondo, M. Hashimoto, N. Ono, T. Takahashi, Characterization of TiN films deposited onto stainless steel strips by continuous dry-coating process, Thin Solid Films 290 (1996) 226–231. [25] G. Reisse, R. Ebert, Titanium nitride thin film deposition by laser CVD, Appl. Surf. Sci. 106 (1996) 268–274. [26] M.S. El-Eskandarany, Synthesis of nanocrystalline titanium carbide alloy powders by mechanical solid state reaction, Metall. Mater. Trans. A 27 (1996) 2374–2382. [27] X. Yang, C. Li, L. Yang, Y. Yan, Y. Qian, Reduction‐nitridation synthesis of titanium nitride nanocrystals, J. Am. Ceram. Soc. 86 (2003) 206–208. [28] J. Li, L. Gao, J. Sun, Q. Zhang, J. Guo, D. Yan, Synthesis of nanocrystalline titanium nitride powders by direct nitridation of titanium oxide, J. Am. Ceram. Soc. 84 (2001) 3045–3047. [29] J. PieteráDekker, P.J. ávan der Put, Vapour-phase synthesis of titanium nitride powder, J. Mater. Chem. 4 (1994) 689–694. [30] K. Kamiya, T. Nishijima, K. Tanaka, Nitridation of the sol–gel‐derived titanium oxide films by heating in ammonia gas, J. Am. Ceram. Soc. 73 (1990) 2750–2752. [31] L. Wicikowski, B. Kusz, L. Murawski, B. Susła, K. Szaniawska, AFM and XPS study of nitrided TiO2 and SiO2–TiO2 sol–gel derived films, Vacuum 54 (1999) 221–225. [32] M. Drygaś, C. Czosnek, R.T. Paine, J.F. Janik, Two-stage aerosol synthesis of titanium nitride TiN and titanium oxynitride TiOxNy nanopowders of spherical particle morphology, Chem. Mater. 18 (2006) 3122–3129. [33] S. Hoang, S. Guo, N.T. Hahn, A.J. Bard, C.B. Mullins, Visible light driven photoelectrochemical water oxidation on nitrogen-modified TiO2 nanowires, Nano Lett. 12 (2012) 26–32. [34] B. Crist, BE Lookup Table for Signals from Elements and Common Chemical Species, Wiley, Chichester, UK, 1999, pp. 77–358. [35] P.-Y. Liu, Z.-H. Miao, K. Li, H. Yang, L. Zhen, C.-Y. Xu, Biocompatible Fe3+–TA coordination complex with high photothermal conversion efficiency for ablation of cancer cells, Colloids Surf., B 167 (2018) 183–190. [36] Y. Liu, Multifunctional Nanoprobes: from Design Validation to Biomedical Applications, Springer, Germany, 2017. [37] K. Bian, X. Zhang, K. Liu, T. Yin, H. Liu, K. Niu, W. Cao, D. Gao, Peptide-directed hierarchical mineralized silver nanocages for anti-tumor photothermal therapy, ACS Sustain. Chem. Eng. 6 (2018) 7574–7588. [38] R. Zhu, Y. Li, X. Zhang, K. Bian, M. Yang, C. Cong, X. Cheng, S. Zhao, X. Li, D. Gao, Vapreotide-mediated hierarchical mineralized Ag/Au nanoshells for photothermal anti-tumor therapy, Nanotechnology 30 (2018) 055602. [39] Y. Yong, L. Zhou, Z. Gu, L. Yan, G. Tian, X. Zheng, X. Liu, X. Zhang, J. Shi, W. Cong, WS2 nanosheet as a new photosensitizer carrier for combined photodynamic and photothermal therapy of cancer cells, Nanoscale 6 (2014) 10394–10403. [40] Q. Tian, F. Jiang, R. Zou, Q. Liu, Z. Chen, M. Zhu, S. Yang, J. Wang, J. Wang, J. Hu, Hydrophilic Cu9S5 nanocrystals: a photothermal agent with a 25.7% heat conversion efficiency for photothermal ablation of cancer cells in vivo, ACS Nano 5 (2011) 9761–9771. [41] D. Ding, Y.T. Xie, K. Li, L.P. Huang, X.B. Zheng, Black plasma-sprayed Ta2O5 coatings with photothermal effect for bone tumor therapy, Ceram. Int. 44 (2018) 12002–12006.

The authors thank the supported by Research Initiation Funds for the Academic Figures from Shaanxi University of Science and Technology (No. 2016XSGG03). References [1] I.H. El-Sayed, X. Huang, M.A. El-Sayed, Surface plasmon resonance scattering and absorption of anti-EGFR antibody conjugated gold nanoparticles in cancer diagnostics: applications in oral cancer, Nano Lett. 5 (2005) 829–834. [2] X. Huang, I.H. El-Sayed, W. Qian, M.A. El-Sayed, Cancer cell imaging and photothermal therapy in the near-infrared region by using gold nanorods, J. Am. Chem. Soc. 128 (2006) 2115–2120. [3] H.A. Atwater, A. Polman, Plasmonics for improved photovoltaic devices, Nat. Mater. 9 (2010) 205–213. [4] X. Liu, L.C. He, J.Z. Zheng, J. Guo, F. Bi, X. Ma, K. Zhao, Y.L. Liu, R. Song, Z.Y. Tang, Solar-light-driven renewable butanol separation by core-shell Ag@ZIF-8 nanowires, Adv. Mater. 27 (2015) 3273–3277. [5] X.L. Liu, Y.X. Tao, H.H. Mao, Y. Kong, J. Shen, L.H. Deng, L.S. Yang, Construction of magnetic-targeted and NIR irradiation-controlled drug delivery platform with Fe3O4@Au@SiO2 nanospheres, Ceram. Int. 43 (2017) 5061–5067. [6] L. Cheng, C. Wang, L.Z. Feng, K. Yang, Z. Liu, Functional nanomaterials for phototherapies of cancer, Chem. Rev. 114 (2014) 10869–10939. [7] S. Kang, S.H. Bhang, S. Hwang, J.K. Yoon, J. Song, H.K. Jang, S. Kim, B.S. Kim, Mesenchymal stem cells aggregate and deliver gold nanoparticles to tumors for photothermal therapy, ACS Nano 9 (2015) 9678–9690. [8] J. Wu, N. Li, Y. Yao, D. Tang, D. Yang, J. Ong’achwa Machuki, J. Li, Y. Yu, F. Gao, DNA-stabilized silver nanoclusters for label-free fluorescence imaging of cell surface glycans and fluorescence guided photothermal therapy, Anal. Chem. 90 (2018) 14368–14375. [9] S. Cheng, S. Zhan, X. Wu, G. Nie, S. Wu, J. Hu, J. Li, S. Hu, Y. Zhang, Y. Liu, Insulating plasmonic photothermal heat of Ag nanoparticles by a thin carbon shell, J. Alloy. Comp. 791 (2019) 380–384. [10] X.X. Qian, S.D. Shen, T. Liu, L. Cheng, Z. Liu, Two-dimensional TiS2 nanosheets for in vivo photoacoustic imaging and photothermal cancer therapy, Nanoscale 7 (2015) 6380–6387. [11] J.T. Robinson, S.M. Tabakman, Y. Liang, H. Wang, H.S. Casalongue, D. Vinh, H. Dai, Ultrasmall reduced graphene oxide with high near-infrared absorbance for photothermal therapy, J. Am. Chem. Soc. 133 (2011) 6825–6831. [12] D. Jaque, L. Martinez Maestro, B. del Rosal, P. Haro-Gonzalez, A. Benayas, J.L. Plaza, E. Martin Rodriguez, J. Garcia Sole, Nanoparticles for photothermal therapies, Nanoscale 6 (2014) 9494–9530. [13] L.M. Maestro, P. Haro-Gonzalez, B. del Rosal, J. Ramiro, A.J. Caamano, E. Carrasco, A. Juarranz, F. Sanz-Rodriguez, J.G. Sole, D. Jaque, Heating efficiency of multiwalled carbon nanotubes in the first and second biological windows, Nanoscale 5 (2013) 7882–7889. [14] X. Huang, P.K. Jain, I.H. El-Sayed, M.A. El-Sayed, Plasmonic photothermal therapy (PPTT) using gold nanoparticles, Lasers Med. Sci. 23 (2008) 217–228. [15] M. Cortie, J. Giddings, A. Dowd, Optical properties and plasmon resonances of titanium nitride nanostructures, Nanotechnology 21 (2010) 115201. [16] A. Naldoni, U. Guler, Z. Wang, M. Marelli, F. Malara, X. Meng, L.V. Besteiro, A.O. Govorov, A.V. Kildishev, A. Boltasseva, Broadband hot‐electron collection for solar water splitting with plasmonic titanium nitride, Adv. Opt. Mater. 5 (2017) 1601031. [17] M. Chirumamilla, A. Chirumamilla, Y. Yang, A.S. Roberts, P.K. Kristensen, K. Chaudhuri, A. Boltasseva, D.S. Sutherland, S.I. Bozhevolnyi, K. Pedersen, Large‐area ultrabroadband absorber for solar thermophotovoltaics based on 3D titanium nitride nanopillars, Adv. Opt. Mater. 5 (2017) 1700552. [18] K.S. Schramke, Y. Qin, J.T. Held, K.A. Mkhoyan, U.R. Kortshagen, Nonthermal plasma synthesis of titanium nitride nanocrystals with plasmon resonances at nearinfrared wavelengths relevant to photothermal therapy, ACS Appl. Nano Mater. 1

7