Tunable charge density wave in TiS3 nanoribbons
Huang Ce1, 2, Zhang Enze1, 2, Yuan Xiang1, 2, Wang Weiyi1, 2, Liu Yanwen1, 2, Zhang Cheng1, 2, Ling Jiwei1, 2, Liu Shanshan1, 2, Xiu Faxian1, 2, 3, †
State Key Laboratory of Surface Physics and Department of Physics, Fudan University, Shanghai 200433, China
Institute for Nanoelectronic Devices and Quantum Computing, Fudan University, Shanghai 200433, China
Collaborative Innovation Center of Advanced Microstructures, Nanjing 210093 China

 

† Corresponding author. E-mail: Faxian@fudan.edu.cn

Abstract

Recently, modifications of charge density wave (CDW) in two-dimensional (2D) show intriguing properties in quasi-2D materials such as layered transition metal dichalcogenides (TMDCs). Optical, electrical transport measurements and scanning tunneling microscopy uncover the enormous difference on the many-body states when the thickness is reduced down to monolayer. However, the CDW in quasi-one-dimensional (1D) materials like transition metal trichalcogenides (TMTCs) is yet to be explored in low dimension whose mechanism is likely distinct from their quasi-2D counterparts. Here, we report a systematic study on the CDW properties of titanium trisulfide (TiS ). Two phase transition temperatures were observed to decrease from 53 K (103 K) to 46 K (85 K) for the bulk and < 15-nm thick nanoribbon, respectively, which arises from the increased fluctuation effect across the chain in the nanoribbon structure, thereby destroying the CDW coherence. It also suggests a strong anisotropy of CDW states in quasi-1D TMTCs which is different from that in TMDCs. Remarkably, by using back gate of V in 15-nm device, we can tune the second transition temperature from 110 K (at −30 V) to 93 K (at 70 V) owing to the altered electron concentration. Finally, the optical approach through the impinging of laser beams on the sample surface is exploited to manipulate the CDW transition, where the melting of the CDW states shows a strong dependence on the excitation energy. Our results demonstrate TiS as a promising quasi-1D CDW material and open up a new window for the study of collective phases in TMTCs.

PACS: 73.22.Pr
1. Introduction

Periodic modulations of electron density from electron–phonon interactions lead to a novel type of ground state called charge density wave (CDW).[1] A great deal of materials such as layered transition metal dichalcogenides (TMDCs) and transition metal trichalcogenides (TMTCs) have been found to possess the CDW states.[1,2] Recently, these interesting states show unexpected and exotic properties when approaching the 2D limit. For example, with the thickness decreasing,[3] a commensurate CDW state completely disappears in TaS ; and the transition temperature can be drastically enhanced in monolayer NbSe and TiSe .[47] More interestingly, the quantum critical behavior may exist in 2D TaS and other correlated systems[8,9] as the reduced dimensionality can strengthen Peierls instability, electron–phonon interactions, and the fluctuation effect.[1,7] Therefore, it is indispensable to investigate the effect of the dimensionality on the CDW states and uncover the intrinsic mechanism.

Like TMDCs, many bulk TMTCs exhibit layered structures with interlayer van der Waals interactions.[1012] As a typical TMTC, -type crystals (M = Ti, Zr, Hf, Ta, Nb; X = S, Se, Te) are quasi-1D materials, some of which possess charge density waves in the bulk form. And these states emerge and slide along the direction of metal-atom chains unlike in TMDC materials where the various CDW states are developed by several-atoms-clusters in 2D layers.[13,14] For instance, two phase transitions at 144 K and 59 K in NbSe are the consequence of three structurally different chains.[15] Monoclinic TaS has two types of chains which induce CDW transitions at 240 K and 160 K.[16] Importantly, the low dimensionality may exert strong effects on those chain-like quasi-1D materials due to the different crystal structures. Each CDW transition can be variously affected with the change of device scale along different crystal orientations, thus giving rise to an evident anisotropy. However, limited works have been reported on the CDW states in quasi-1D TMTC materials when the dimensionality is reduced to two dimensions.

TiS , as one of the semiconducting TMTCs, shows extremely excellent optical properties.[1721] Strong anisotropy from electric transport measurements represents 30% of increase over the black phosphorus.[19,22] The bulk crystals have phase transitions at 120, 60, and 17 K through quasi-1D transport.[2326] with the first two along the perpendicular chains. We anticipate that the nanoribbon structure can have structural fluctuations in horizontal and perpendicular directions that lead to anisotropic CDW behavior upon external electrical gating. However, to date such studies remain lacking.

Here, we report the properties of exfoliated few-layer TiS nanoribbons and find only two phase transitions at 46 K ( and 85 K ( in stark contrast to the three in thick TiS at 29, 53, and 103 K. can be tuned by the back-gate voltage because of the variable carrier concentration whereas does not show an evident change owing to its robustness to the structural fluctuations. Our transport experiments clearly exhibit a strong anisotropy of CDW states in TiS . We also reveal the light-induced melting of charge density wave under 633-nm laser illumination. When the laser intensity exceeds 10 μW, both two transitions disappear. Moreover, a low illumination intensity of 0.02 μW by a 532-nm laser can readily destroy the CDW states.

2. Methods
2.1. Growth and characterizations of TiS nanoribbons

TiS was synthetized by sulfuration of Ti using chemical vapor transport (CVT) method.[13] 0.23-g titanium powder and 0.6-g sulfur powder were sealed in an evacuated quartz ampule. The ampule was placed in a tube furnace, where it was heated up to 500 °C for 1 day, then kept at 500 °C for 1 day for sulfuration. The cooling process took place in ambient conditions. Sulfur vapor pressure at 500 °C was around 4 bars ( ).

Raman spectroscopy measurements were performed at a self-built Raman system with a laser wavelength of 532 nm. The diameter of focused laser spot is less than 2 μm when using a 100-μm pin hole.

2.2. Device fabrication and transport measurements

The electrodes of the device were fabricated by EBL using MMA/PMMA bilayer polymer. Cr/Au (5 nm/100 nm) electrodes were deposited by sputtering. The temperature-dependent transport measurements of the 4-probe device were carried out in a Physical Property Measurement System (PPMS) system (Quantum design) using an Agilent 2912.

2.3. Light-induced transport measurement

The light-induced transport measurements were carried out in an optical dry cryogenic system (Oxford Instruments, OptistatDry BL4).

3. Results

Bulk TiS crystals were synthesized via chemical vapor transport (CVT) method (sulphuration of titanium powder).[17]

Figure 1(a) shows a layered structure of TiS ; parallel chains of triangular prisms which are held by van der Waals forces constitute sheets. The parallel chains are responsible for the quasi-1D nature of the material which leads to charge density waves. Each sheet is formed by two titanium layers and four sulfide layers. The as-grown TiS naturally develops groups of black needle-like bulk crystals. Figure 1(b) shows one of the exfoliated TiS samples transferred onto SiO (285 nm)/Si substrate with a number of nanoribbons. Figure 1(c) shows a scanning electron microscopy (SEM) image, where both bulk crystals and thin nanoribbons can be identified. The purple one in the red rectangle is a thin nanoribbon with a thickness of nm. To identify the composition of sample, energy dispersive x-ray spectroscopy (EDX) has been carried out. Figure 1(d) displays the EDX results of the as-grown samples. More than five positions have been randomly picked and measured, all of which show clear peaks of Ti and S with the total error less than 0.5%. The atomic ratio of Ti and S is nearly Ti:S=1:3, indicating a correct composition of TiS . We further carried out Raman spectroscopy to explore the phonon characteristics of materials and identify the chemical bond. Figure 1(e) displays the Raman spectrum of an isolated nanoribbon thicker than 50 nm; the optical image of the sample is shown in the inset under 532-nm laser illumination. While the peak at is due to Raman mode of silicon substrate, the other four peaks located at , , , and correspond to A -type Raman modes of the TiS which are in good agreement with the bulk crystals. Besides, a previous study reports the observation of the peak which has the highest intensity along b axis in nanoribbons,[27] which is consistent with our results.

Fig. 1. (color online) Synthesis and characterizations of TiS nanoribbons. (a) The TiS crystal structure. Ti atoms are covalently bonded to six S atoms forming triangular prisms which are stacked to make chains. (b) The optical microscopy image of TiS nanoribbons on an SiO /Si substrate. (c) The SEM image of TiS nanoribbons. Inset: single TiS nanoribbon. (d) The EDX spectrum of TiS nanoribbons with the atomic ratio of Ti/S of 1:3. (e) The Raman spectrum of the individual TiS nanoribbon on SiO /Si substrate at room temperature using a 532-nm laser source.

In order to study the CDW properties, four-terminal devices with Cr/Au (5 nm/100 nm) electrodes were fabricated along the nanoribbon chain direction, i.e., the b axis, by standard e-beam lithography, metal evaporation, and lift-off process. Temperature-dependent resistance (RT) is measured from room temperature down to 10 K as shown in Fig. 2(a). The inset displays the device image by a conventional optical microscope. The length of the sample is around 20 μm and the width is 2 μm. The thickness of the nanoribbon is determined to be nm, which is relatively thick. The resistivity measured at 300 K has a value of ρ=0.4Ωcm1 in good agreement with the previously-reported bulk results.[28] The RT curves in our experiments are all measured in the cooling process to eliminate the hysteresis induced by the system. As the temperature cools down, the resistance shows several broad anomaly features. In order to clearly identify the transition temperatures, we adopted the widely-used derivative method and plotted against the temperature. It is known that near the transition temperature, the conductivity can be extracted by

based on the mean-field theory, where is the mobility and is the density of states.[16] When , the temperature derivative of the resistivity is approximately given by
Consequently, the derivative has a cusp singularity at transition temperature. Here, we use the maxima from the derivative curve.[29] to estimate the transition temperature of 29, 53, and 103 K in the nm-thick nanoribbon, which are similar to that of the bulk crystals.[23] Instead of showing a cusp, one observes a broadened maximum, which is consequence of fluctuations caused by the 2D dimensionality. At the same time, quasi-1D systems always induce nonlinear conductivity due to the sliding or creeping of the CDW states.[30] And the current–voltage (IV) curves follow the power law behavior in CDW systems (along the chains) at low temperatures. Indeed, the IV curve at 10 K exhibits a voltage-dependent conductivity whereas at 130 K it becomes linear with a constant conductivity. Figure 2(b) summarizes all IV curves in the temperature range of 10 K to 130 K where a nonlinear behavior is exclusively observed. However, owing to the possible contribution from the Schottky barrier which is observed in monolayer MoS using four-point probe measurement, the exact onset of the CDW transition remains elusive.[31] We note that the four-probe method is unsuitable if the Schottky behavior occurs between the metallic contact and the sample where the Schottky barrier induces a large contact resistance causing inaccuracy of sample resistance.[32,33] Since the nanoribbon is thick enough, the drain–source current can be defined by three-dimensional (3D) thermionic emission equation[34]
where A is the Richardson constant, n is the ideality factor. The fitting results show a Schottky barrier of 42.3 meV (see Fig. S1 in the supplementary information in Appendix A). Thus, our IV curves consist of contributions from both quasi-1D CDW and Schottky barrier. When the voltage is larger than 0.4 V, is proportional to V, matching the thermionic emission model (Fig. S1(a), and the current induced by Schottky barrier at low voltage is much smaller than that from the quasi-1D transport. Therefore, at low voltage, the quasi-1D transport dominates while the Schottky barrier plays important roles at the high-voltage regime. When the voltage is lower than the threshold, the electrons cannot be excited across the CDW gap in a two-fluid model.[1] The current can be written as , where the subscripts n and CDW refer to the current carried by the uncondensed and condensed electrons. The former has an Ohmic contribution and the latter has a nonlinear response. For CDW materials, the threshold field is around few millivolt per micron,[1] which is much lower than that in the fitting region of Fig. 2(b). Moreover, no threshold field is observed in nonlinear transport measurements in bulk TiS in the high temperature region,[23] unlike the characteristics in very thin NbSe samples with disorders.[35] At 4.2 K, the threshold field is reported to be 6 V/cm.[36] Therefore, at a specific temperature, the IV curves have two linear parts as shown in the inset of Fig. 2(c); the blue region is dominated by the quasi-1D CDW transport and the red regime corresponds to the Schottky barrier. We fit both the two regions by using the power law at different temperatures; the resultant temperature-dependent parameter is illustrated in Fig. 2(c). It can be seen that becomes larger than 1 as the temperature goes below 110 K and a sharp increase occurs at about 60 K.

Fig. 2. (color online) CDW transitions in a nm thick and 15-nm thin TiS nanoribbon. (a) The temperature dependence of the resistance and the logarithmic derivative . The source–drain current nA. Three CDW transitions are evident at 29, 53, and 103 K. Inset: optical microscopy image of the 50 nm–60 nm TiS3 FET. (b) curves of the thick TiS3 nanoribbon at different temperatures. (c) The temperature dependence of parameter using different fitting region. The red curve is fitted at V, while the blue one is at V. The inset shows the two fitting regions. (d) Temperature dependence of the resistance and the logarithmic derivative in a device whose thickness is lower than 15 nm. The source drain current nA. Unlike the thick nanoribbon, two CDW transitions appear at 46 K and 85 K. Inset: optical image of the FET device.

The phase transitions at 53 K and 103 K can be explained by the formation of two charge density waves in disordered quasi-1D system. It was found that below 110 K, a CDW state develops in the direction across the chains, like ZrTe , while the other CDW forms along the chains below 60 K.[23] Our narrow TiS nanoribbon increases the fluctuation in the direction across the chains, thus weakening the CDW, which also explains the weak non-linear transport after the first CDW state at 110 K. When the temperature is below 25 K, the noise at low voltage becomes ignorable; correspondingly the fitting parameter has large errors to be analyzed. As the voltage becomes larger, the Schottky barrier turns to be prominent (the red curve). Below 130 K, the IV curve turns to be nonlinear, and gets larger with the temperature reduced, but three peaks clearly show up with two of which corresponding to the aforementioned two CDW states. The transition at 25 K, however, is still unknown and needs further investigation. The transition temperatures from the quasi-1D transport analysis and derivative curve are quite similar, which also confirms the three CDW phase transitions.

Next, we focus on the thinner nanoribbons. The fabrication method is the same as before, but the sample thickness is reduced to < 15nm as displayed in the inset of Fig. 2(d). Note that experimentally it is challenging to reach monolayer TiS nanoribbon because of the 1D chain structure and the subsequent electric measurements are difficult owing to the extremely large resistance when approaching the atomic thickness. The RT curve shows anomaly peaks at 46 K and 85 K (the blue curve), which correspond to the two peaks in the derivative curve (in red). The third transition cannot be observed in transport measurements for which the underlying mechanism is not clear yet. We note that the carrier density can strongly affect the CDW phase.[16,37] Compared with the thick device, in the thin nanoribbons we find a rapid diminishment of the two transition temperatures. We provide two reasons to explain this behavior as follows. First, the scattering mechanism in TiS is dominated by charge impurities at low temperatures (see Fig. S2 in Appendix A). The phase transition is greatly suppressed by impurities in quasi-2D or 1D CDW materials.[16] Non-isoelectronic impurities can induce potentials which lead to changes of the phase of the condensate, destroy long-range CDW order and lead to a finite phase–phase correlation length .[38,39] Second, lowering the dimensionality of a system usually tends to reduce the phase space available for the phase parameter to adapt to imperfections, and consequently it suppresses the long-range order and lowers the transition temperature .[16] Moreover, the Coulombic interaction can modulate transport properties in two dimensions.[40] With the thickness decreased, the interaction between the SiO /Si substrate and TiS nanoribbon gradually dominates which can also jeopardize the CDW coherence. The size effect of CDW transition temperature can be described by empirical formula[41]

where is the bulk CDW transition temperature, is the critical thickness at which the CDW transition vanishes, is a temperature correction term. Alternatively, we think that the second CDW transition is more stable in the bulk or nanosheet but not in nanoribbon (chains) because is in the direction across the chains. In other words, the interlayer coupling is unfavorable for the second CDW state. The device in Fig. 3(b) has a phase transition at 40 K and 100 K, which is different from the device in Fig. 3(a). This may be due to the different thicknesses, impurities, and electron concentrations. Therefore, at large gate voltage the decreases. The transition temperature in TiS nanoribbon shows opposite thickness-dependence trend with quasi-2D TMDCs (NbSe , TiSe which will help us to better understand the relation between Peierls instabilities and fluctuation effects.[7,42]

Fig. 3. (color online) Gate-tuned CDW transitions in thin TiS nanoribbon. (a) Temperature-dependent resistance at various gate voltages from V to 70 V except 40 V in a 15-nm device. Inset: the enlarged view of the RT curve at low temperature regime. (b) Temperature-dependent logarithmic derivative of RT under different . Right panel, the curves are multiplied by 10 times to display the evolution of CDW phase transition under external gate voltages. (c) The gate voltage dependence of transition temperature and .

In bulk materials, the doping and high pressure can modulate the carrier density, for instance, to enhance the superconductivity in Cu-doped TiSe .[43] In quasi-2D materials, the electric-field effect is much more convenient than these two methods: a gate voltage can induce a large number of electrons (holes) into the system so that CDW can be easily tuned.[3,37] Here, we use a back-gate voltage to tune the phase transitions in 15-nm thick nanoribbon. Figure 3(a) illustrates a set of RT curves at gate voltage ranging from V to 70 V using a constant source–drain current of 1.5 nA. At 40 K, the resistance changes from 2.6 G to 0.07 G . From V to V, a peak can be clearly identified near the first transition temperature (Fig. 3(a)). The semiconducting phase converting to be metallic looks more evident when we use logarithmic coordinates as demonstrated in the inset. It is possible that the large CDW conduction modulation is due to changes in CDW pinning arising from gate–voltage-induced variations in the CDW gap which was also observed in semiconducting quasi-1D material TaS .[44] Likewise, figure 3(b) displays the derivative curves to ensure the extraction of the transition temperature. Apparently, two peaks are shifted as a function of the gate voltage. The gate–voltage dependence of conductance at 160 K is displayed in Fig. S2(b). At V, the conductance tends to be zero and the extracted field-effect mobility and on/off ratio are 0.8 cm and , respectively (More information is available in Figs. S2 and S3 in Appendix A). Systematic changes under different gate voltages are summarized in Fig. 3(c). For the first CDW transition, the critical transition temperature does not vary much and it positions at around 40 K. On the contrary, is identified to be 110 K when V and reduced to 93 K at V and in the whole gate region has negative dependence with the value of gate voltage. It can be naturally explained by the fact that the increase of the electron density at positive voltages suppresses , which is also supported by a similar behavior occurring in 1T-TiSe and TaS .[3,37] Specifically, a large electron density may enhance CDW phase fluctuations, strengthen disorders, and thus ultimately destroy a long-range CDW coherence. More electrons tend to be excited across the gap, the periodic structure of charge density becomes unstable and much lower temperature is needed to reduce the thermal excitation and keep the density wave. This is commonly observed in quasi-1D and quasi-2D systems.[16] In contrast, decreases a little at larger ; this is presumably caused by its robustness to the fluctuations because of the ribbon structure in which the CDW forms along the chains. However, for the thick nanoribbon sample, the gate cannot tune it well in a sharp contrast to the thin one, as shown in Fig. S4 in the supplementary information (Appendix A).

Since TiS is a light-sensitive semiconductor, a 633-nm laser (1.94 eV) is used to illuminate the TiS surface as displayed in the inset of Fig. 4(a). The RT curve shows two phase transitions at 40 K and 85 K without laser illumination (purple curve). The illuminating region of TiS is about 20 . As the laser intensity gradually increases, the resistance reduces and the phase transition becomes weaker prior to the disappearance under 1.0 μW. Figure 4(c) displays the derivative curves where the threshold intensity is about 1.0 μW and the two peaks come to be completely invisible above this threshold. Though the resistance is monotonically decreased by increasing the laser power, the position of two peaks do not change below 1.0 μW, which can exclude the heating effect. Like the process of light-induced superconductivity suppression,[45,46] laser can destroy the condensate in the distorted region in the CDW state and cause a prompt loss of charge order and equilibration of the electron–lattice system when the absorbed optical energy is higher than the condensation energy (the energy difference between the free energy of the CDW state and the normal state).[47,48] After the excitation, the periodic electron density structure is destroyed and TiS converts to the normal semiconducting state. With a low laser excitation, however, the two phase transitions still exist but are softened with the increasing of the laser intensity, which may be due to the insufficient photons absorbed per unit cell.[49] The low laser excitation can raise the fluctuation but not sufficient to destroy the condensation energy. Figure 4(b) shows the IV curves under different incident laser intensities at 25 K. The resistance reduces after the laser excitation and the IV curves become linear under 11-μW laser intensity which indicates that the CDW states and Schottky barrier both disappear. Under 1.0-μW laser intensity, the curve is nonlinear, indicating the Schottky barrier existence while CDW states are destroyed. The quasi-linear IV curve also shows the suppression of CDW states under laser illumination. Figure 4(d) displays the 532-nm laser condition: under a laser intensity of 0.02 μW, the CDW transitions are substantially weakened and two peaks in the derivative curve disappear (see the inset in Fig. 4(d)). Lower critical intensity of 532-nm laser may attribute to the larger illumination energy. In short, we conclude that the CDW state strongly depends on the excitation energy, reminiscent of superconductivity, where the CDW pseudo gap state is easily broken by external factors. Below a critical energy, the system needs much more photons to excite the redundant electrons across the gap. Otherwise, the CDW states can still survive.

Fig. 4. (color online) Melting of the CDW states by laser. (a) curves under different laser power (0 μW–11 μW). Inset shows a schematic structure of TiS device incident with a 633-nm laser. (b) I –V curves under different incident laser powers. (c) Temperature-dependent RT logarithmic derivatives under different laser powers. (d) RT curves with and without 532-nm laser. Inset displays temperature-dependent RT logarithmic derivatives.
4. Discussion

Finally, let us briefly compare TiS with previously widely studied CDW in TMDC such as NbSe . In NbSe , reducing the dimensionality strengthens the CDW phase and weakens the superconducting phase due to the Fermi surface nesting.[7] In contrast, TiS shows an opposite trend in which thin TiS exhibits a lower transition temperature. The CDW states in TiS are strong anisotropic resulting in the different gate tuning capability because of its robustness to the structural fluctuations. In conclusion, we have performed a systematic study on charge density wave properties in quasi-1D TiS nanoribbons. Two CDW transition temperatures decrease in the thin nanoribbon. The second transition temperature reduces as the increase of the electron doping, while the first transition shows a negligible change. We also observed the light-induced melting of CDW after 633-nm and 532-nm laser illuminations. CDW properties in TMTCs are different from quasi-2D materials such as TMDCs. Our study shows that TiS is a promising CDW material for the study of collective phases in TMTCs and expanding the functionalities for electronic applications like information processing.[50]

Reference
[1]GrünerG 1988 Rev. Mod. Phys. 60 1129
[2]WilsonJ ADi SalvoFMahajanS 1975 Adv. Phys. 24 117
[3]YuYYangFLuX FYanY JChoY HMaLNiuXKimSSonY WFengD 2015 Nat. Nanotechnol. 10 270
[4]ChenPChanY HFangX YZhangYChouM YMoS KHussainZFedorovA VChiangT C 2015 Nat. Commun. 6 8943
[5]UgedaM MBradleyA JZhangYOnishiSChenYRuanWOjeda-AristizabalCRyuHEdmondsM TTsaiH Z2016Nat. Phys.1292
[6]ChatterjeeUZhaoJIavaroneMDi CapuaRCastellanJ PKarapetrovGMalliakasC DKanatzidisM GClausHRuffJ PWeberFVan WezelJCampuzanoJ COsbornRRanderiaMTrivediNNormanM RRosenkranzS 2015 Nat. Commun. 6 6313
[7]XiXZhaoLWangZBergerHForróLShanJMakK F 2015 Nat. Nano 10 765
[8]StojchevskaLVaskivskyiIMerteljTKusarPSvetinDBrazovskiiSMihailovicD 2014 Science 344 177
[9]VaskivskyiIGospodaricJBrazovskiiSSvetinDSutarPGoreshnikEMihailovicI AMerteljTMihailovicD 2015 Sci. Adv. 1 e1500168
[10]HodeauJMarezioMRoucauCAyrolesRMeerschautARouxelJMonceauP1978J. Phys. C: Solid State Phys.114117
[11]MonceauPOngNPortisA MMeerschautARouxelJ 1976 Phys. Rev. Lett. 37 602
[12]MonctonD EAxeJDi SalvoF 1977 Phys. Rev. 16 801
[13]JinYLiXYangY 2015 Phys. Chem. Chem. Phys. 17 18665
[14]YoshidaMSuzukiRZhangYNakanoMIwasaY 2015 Sci. Adv. 1 e1500606
[15]WilsonJ 1979 Phys. Rev. 19 6456
[16]GrunerG2000Density waves in solidsWestview Press
[17]errerIMaciáMCarceléVAresJSáchezC 2012 Energy Procedia 22 48
[18]DaiJZengX C 2015 Angewandte Chemie 127 7682
[19]IslandJ OBuscemaMBarawiMClamagirandJ MAresJ RSánchezCFerrerI JSteeleG Avan der ZantH SCastellanos-GomezA 2014 Adv. Opt. Mater. 2 641
[20]WuKTorunESahinHChenBFanXPantAParsonsWright DAokiTPeetersF MSoignardETongayS 2016 Nat. Commun. 7 12952
[21]KangJSahinHOzaydinH DSengerR TPeetersF M 2015 Phys. Rev. 92 075413
[22]IslandJ OBieleRBarawiMClamagirandJ MAresJ RSanchezCvan der ZantH SFerrerI JD’AgostaRCastellanos-GomezA2015arXiv: 1510.06889
[23]GorlovaIZybtsevSPokrovskiiV YBolotinaNVerinITitovA 2012 Physica B: Conden. Matter 407 1707
[24]GorlovaIZybtsevSPokrovskiiV YBolotinaNGavrilkinS YTsvetkovA Y2012Physica B: Conden. Matter46011
[25]GorlovaIPokrovskiiV YZybtsevSTitovATimofeevV 2010 J. Exp. and Theor. Phys. 111 298
[26]GorlovaI GZybtsevS G EPokrovskiiV Y 2014 JETP Lett. 100 256
[27]PawbakeA SIslandJ OFloresEAresJ RSanchezCFerrerI JJadkarS Rvan der ZantH SCastellanos-GomezALateD 2015 ACS Appl. Mater. & Interfaces 7 24185
[28]HsiehP LJacksonCGrüerG 1983 Solid State Commun. 46 505
[29]MinakovaVNasretdinovaVZaitsev-ZotovS 2015 Physica B: Conden. Matter 460 185
[30]FelserCFinckhEKleinkeHRockerFTremelW 1998 J. Mater. Chem. 8 1787
[31]RadisavljevicBKisA 2013 Nat. Mater. 12 815
[32]HesseE1982IEEE Trans. Instrum. Meas.1001166
[33]BarnettCKryvchenkovaOWilsonLMaffeisTKalnaKCobleyR 2015 J. Appl. Phys. 117 174306
[34]BhuiyanAMartinezAEsteveD 1988 Thin Solid Films 161 93
[35]SlotEHolstMvan der ZantHZaitsev-ZotovS 2004 Phys. Rev. Lett. 93 176602
[36]GorlovaI GPokrovskiiV Y 2009 JETP Lett. 90 295
[37]LiL JO’FarrellE C TLohK PEdaGÖyilmazBCastro NetoA H2016Nature529185
[38]ImryYMaS K 1975 Phys. Rev. Lett. 35 1399
[39]ShamLPattonB R 1976 Phys. Rev. 13 3151
[40]RadisavljevicBRadenovicABrivioJGiacomettiVKisA2011Nat. Nanotechnol.6147
[41]YangJWangWLiuYDuHNingWZhengGJinCHanYWangNYangZ 2014 Appl. Phys. Lett. 105 063109
[42]GoliPKhanJWickramaratneDLakeR KBalandinA A 2012 Nano Lett. 12 5941
[43]MorosanEZandbergenHDennisBBosJOnoseYKlimczukTRamirezAOngNCavaR 2006 Nat. Phys. 2 544
[44]AdelmanTZaitsev-ZotovSThorneR 1995 Phys. Rev. Lett. 74 5264
[45]DemsarJAverittRTaylorAKabanovVKangWKimHChoiELeeS 2003 Phys. Rev. Lett. 91 267002
[46]KusarPKabanovV VDemsarJMerteljTSugaiSMihailovicD 2008 Phys. Rev. Lett. 101 227001
[47]TomeljakASchaeferHStäterDBeyerMBiljakovicKDemsarJ 2009 Phys. Rev. Lett. 102 066404
[48]Mör-VorobevaEJohnsonS LBeaudPStaubUDe SouzaRMilneCIngoldGDemsarJSchäerHTitovA 2011 Phys. Rev. Lett. 107 036403
[49]OgawaNShiragaAKondoRKagoshimaSMiyanoK 2001 Phys. Rev. Lett. 87 256401
[50]KhanJNolenCTeweldebrhanDWickramaratneDLakeRBalandinA 2012 Appl. Phys. Lett. 100 043109
[51]WangWLiuYTangLJinYZhaoTXiuF2014Sci. Rep.4
[52]SzeS MNgK K2006Physics of semiconductor devicesJohn wiley & Sons
[53]KaasbjergKThygesenK SJacobsenK W 2012 Phys. Rev. 85 115317