Proximity-Induced Superconductivity in New Superstructures on 2H-NbSe2 Surface

Funds: Supported by the National Natural Science Foundation of China under Grant Nos 11574372 and 11322432, and the 'Strategic Priority Research Program (B)' of the Chinese Academy of Sciences under Grant No XDB07020300.
  • Received Date: April 24, 2017
  • Published Date: June 30, 2017
  • Using scanning tunneling microscopy we observe a stripe phase smoothly interfacing with a triangular 2×2 superstructure on the surface of 2H-NbSe2 single crystal. Proximity-induced superconductivity is demonstrated in these new ordered structures by measurements of low-temperature tunneling spectra. The modulation of superconductivity by the reconstruction provides an opportunity to understand the interplay between superconductivity and charge orders.
  • Article Text

  • Recent years, two-dimensional (2D) magnetism in layered van der Waals (vdW) materials has attracted considerable attention since potential applications in 2D spintronics.[17] Directly measuring magnetization is crucial to revealing the magnetic structures underlying nontrivial behaviors,[810] such as abnormal magnetizing,[11] magnetic phase transitions,[1214] and hidden orders.[15,16] However, conventional magnetic measurement techniques encounter difficulties when probing the magnetic properties of nano-scale vdW magnets. For example, vibrating sample magnetometry has a sensitivity of 10−9 A⋅m2, which is insufficient for such studies.

    Dynamic cantilever magnetometry (DCM) has emerged as an ultra-sensitive technique,[1719] the method enables the detection of magnetic moment signals at the magnitude of 10−17 A⋅m2.[20] When a nanometer-sized sample is attached onto the free end of a cantilever, the effective spring constant of cantilever alters with external magnetic field due to magnetic restoring force exerting on the sample. This additional magnetic spring constant Δk is associated with the magnetization of the sample.[21] Based on this relationship, the magnetic properties of sample can be derived from eigen frequency shifts Δf of the cantilever sensor. So far, DCM has been employed in measurements of individual nanometer-scaled particles such as magnetic nanotube and magnetic mesocrystals,[18,2226] which can be mounted onto cantilever by using micro-manipulator. However, due to the lack of appropriate transfer technique, the thickness of vdW materials studied by DCM has been limited to micrometer scale.[27]

    In this work, we develop a two-step fabricating method for transferring exfoliated atomically thin flakes onto cantilever sensor, enabling us to extend DCM measurements to 2D magnets. Employing this technique, we investigate the magnetic properties in thin CrPS4 flake with a thickness down to 8 nm by measuring the Δk. In the case of Hc, the spin flop transition leads to a dip in Δk, followed by a gradual increase and then reaches next region with a change in slope. For the case of Hc, the low-temperature results indicate a canted antiferromagnetic state. As T increases, a bump followed by a hump in Δk is observed until Δk is completely negative above 30 K, which can be attributed to the presence of spin reorientation. The HT phase diagrams of 2D CrPS4 revealed by DCM experiment are consistent with the previous studies on bulk samples,[28] which demonstrates the reliability of DCM in investigating magnetic properties of 2D magnets.

    CrPS4, a ternary transition metal chalcogenide (TTMC), has monoclinic symmetry with space group C2/m.[29,30] Its closely packed surface is constituted by S atoms in monolayers as presented in Fig. 1(a). Magnetically, CrPS4 exhibits intralayer ferromagnetism and interlayer antiferro-coupling,[31,32] which contribute to a complex metamagnetic phase transition. The weak spin interactions in CrPS4 leads to little dependence of transition temperature on thickness.[32] Thus, CrPS4 is an appropriate candidate for demonstrating the capability of DCM in examining the magnetism in 2D magnets. The zero-field-cooled (ZFC) magnetization characterizations of bulk CrPS4 were measured with external field μ0H = 0.1 T applied both parallel to the c direction [Fig. 1(b)] and in the ab plane [Fig. 1(c)]. The Néel temperature of our sample is around 38 K, which is consistent with previous studies.[29]

    Fig. 1.  Characterization of CrPS4 and fabrication of the DCM device. (a) The crystal structure of CrPS4. [(b), (c)] The c-direction and ab-plane magnetic susceptibilities, respectively, measured under a magnetic field of μ0H = 0.1 T. (d) Schematic representation of the transferring method of the 2D CrPS4 flakes. Gray needles represent the micro manipulators for transfer. The main steps have been arranged in sequential order using numbered circles. (1) The polycarbonate (PC) film was affixed onto the polydimethylsiloxane (PDMS) and pressed against the flake, followed by heat melting to enhance adhesion. (2) Cool the substrate to solidify the PC film and peel off the flake. (3) The flake was aligned on top of the target silicon lamella and pressed against the lamella. (4) Heat the substrate and peel off the PDMS. (5) Soak the substrate in chloroform and dissolve the PC film. (6) Dry off the substrate and transfer lamella to the cantilever sensor. (e) Schematic drawing of the DCM devices S1. The inset shows the photograph of the silicon lamella with sample attached on (scale bar: 5 μm). (f) Atomic force microscopy image of the flake on S1 with a thickness of 8 nm. The thickness of the flake is shown in (g).

    We designed and fabricated sensitive cantilever beam for DCM measurements. Firstly, a thin film of SiO2 with thickness of 1 μm was grown on top Si layer of a silicon-on-insulator (SOI) substrate by plasma-enhanced chemical vapor deposition (PECVD) to balance stress beneath. Patterns of cantilevers were micro-fabricated by ultra-violet lithography on top surface of SOI. An inductively coupled plasma (ICP) deep reactive ion etch (DRIE) was adopted to etch SOI to buried oxide (BOX) layer. On bottom surface of SOI substrate, the patterned 500 μm silicon was removed by DRIE (Bosch Process) on a deep silicon etcher system. Finally, sacrificial SiO2 layer was released by dry release etch using vapor HF (hydrogen fluoride) chemistries (sacrificial vapor release etching).

    Conventionally, a focused ion beam (FIB) is used to structure micro/nano-sized samples and transfer them onto the cantilever,[18] especially for thin films.[19] For 2D flakes, FIB fabricating might induce damage in samples. In our study, a two-step transfer method is schematically illustrated in Fig. 1(d), whose process comprises two essential steps. Initially, the exfoliated 2D flake was mechanically peeled off and transferred onto a silicon lamella. Subsequently, by using the micro-manipulators, the 2D flake together with the lamella were mounted onto the free end of a cantilever. This technique provides two advantages in DCM experiment: (1) a major reduction in thickness, (2) a major boost in surface area for thin film to enhance the magnetic signal. The lamella attached to the extremity of cantilever can play a role as a mass-loaded end to enhance the sensitivity. Moreover, the use of lamella facilitates the mounting process. By utilizing this technique, we have successfully fabricated CrPS4 sample S1 with a thickness down to 8 nm and an area of 30 μm2. The silicon lamella was cut to a size of 12 μm × 8 μm × 1 μm by using FIB. Its magnetic signal can be negligible in DCM experiments. The single-crystal Si cantilever used in our study is 200 μm long, 10 μm wide and 200 nm thick. The morphological characterization depicted in Fig. 1(e) demonstrates that our method is a clean transfer technique.

    The DCM insert was sealed in a vacuum sleeve, maintaining a pressure below 10−6 mbar. By using a rotatable sample stage, a magnetic field up to 9 T can be applied parallel or perpendicular to the cantilever. The cantilever used in our DCM has a low spring constant of 0.14 mN/m. The eigen frequency f0 of 2.7 kHz can be measured by using a home-made laser interferometer. The frequency responses were precisely monitored by employing the ring-down method. When the signal is strong, due to the mechanical properties of the cantilever beam, the antisymmetric effect in magnetic fields is obvious. In order to exclude the antisymmetric effect in magnetic fields, the Δf has been symmetrized in positive and negative magnetic fields.

    Figure 2 displays the DCM experiment results in S1 for Hc. The measured Δf as a function of H at low temperatures could be classified into three regimes. In the low-field range below a critical field of around 0.8 T, Δf rapidly decreases to a minimum. In the intermediate regime, Δf gradually increases and then enters next region at around 7.6 T. The similar behavior has typically been observed in magnetic torque signals of bulk CrPS4,[29] however, its quantitative interpretation of magnetism is still missing.

    Fig. 2.  The experimental and calculated results for the magnetic field dependence of Δf in the case of Hc. (a) The field dependence of Δf in S1 with Hc from 5 K to 40 K. Inset schematically illustrates the experimental setup. The cantilever is securely fixed along the y-axis and vibrates within the yz plane. The laser is directed along the z-axis. (b) The calculated Δf for an antiferromagnetic thin film when H is applied along the easy axis. The red line illustrates the numerical results. The left inset exhibits the schematic view of the relative orientations of the cantilever axis, applied magnetic field H, and the magnetic moments in sample. Other insets depict the deduced spin alignments of CrPS4 as black and green arrows.

    In order to analyze the magnetic phase transitions at low temperatures, we treated the sample as a uniaxial particle by following the Stoner–Wohlfarth-type model presented by Weber et al.[18] For CrPS4 samples, as reported previously, there are two inequivalent magnetic structures that induce complicated anisotropies.[30] Nevertheless, at low temperature, the out-of-plane anisotropy dominates. Thereby we considered a projected anisotropy constant K (K > 0) to simplify the analysis. In this model, two sets of macrospins (ma and mb) with interactions have been presented. The spin alignments have been depicted as green and black arrows in Fig. 2(b). The ϕa and ϕb denote the deviations of ma and mb from the easy axis. The two sets of macrospins remain in equilibrium between the inherent anisotropy, magnetic torque induced by H and the antiferromagnetic interactions. An antiferromagnetic interaction constant A has been introduced. Ama can be regarded as an effective magnetic field applied to mb. Thus, the free energy of the system can be written as the sum of the cantilever energy, the Zeeman energy, an effective anisotropy energy and an antiferromagnetic interaction energy

    E=12k0l2eθ2(MaVHcos(ϕaθ)+MbVHcos(ϕbθ))+12KV(sin2ϕa+sin2ϕb)+AMaMbV2cos(ϕaϕb),

    (1)

    where k0 and le are the inherent spring constant and effective length of the cantilever, respectively. V is the volume of the CrPS4 sample. The magnetizations of two macrospins are Ma and Mb with Ma = Mb = M0. As shown in inset of Fig. 2(b), θ depicts the vibration deviation. Considering the system in an equilibrium state, the energy of the system must satisfy Eϕa=Eϕb=0, giving

    Eϕa=MaVHsin(ϕaθ)+12KVsin2ϕaAMaMbV2sin(ϕaϕb)=0,Eϕb=MbVHsin(ϕbθ)+12KVsin2ϕb+AMaMbV2sin(ϕaϕb)=0.

    (2)

    The solutions of the equation reveal the possible spin alignments in samples. One solution for θ = 0 is ϕa0 = −ϕb0 = 0, which indicates an A-type antiferromagnetic (A-AFM) spin arrangement. When H reaches a critical value Hsf, which satisfies M0Hsf+(KAVM20)(2M0HsfK+2AVM20K+2AVM20M0H2sf)=0, the macrospins prefer to stay in plane with lower total energy and the Hsf was denoted as spin-flop field. Correspondingly, the macrospins align as ϕa0=ϕb0=arccos(M0HK+2AM20V), revealing the canted spin alignments. With increasing H, the macrospins gradually rotate to the external field and orient towards the same direction at Hs=K+2AM02VM0. The connection between the spin alignment and the mechanical response of the cantilever sensor was established by the formula Δk=1l2e2Eθ2k0 (see chapter S4 in the Supporting Information for more details). According to Δf=f0Δk2k0, we can obtain the calculated Δf as a function of H,

    Δf=f0V2k0l2e(HKM0K+HM0+AM20VHKM0KHM0+AM20V),for|H|<Hsf,Δf=f0V2k0l2e2M0HK(2H2M20(K+2AVM02)21)[M0H+(KAVM20)(2M0HK+2AVM20K+2AVM20M0H2)]1,forHsf<|H|<Hs,Δf=f0V2k0l2e2M0|H|KM0|H|+KAVM20,for|H|>Hs.

    (3)

    The numerical results have been shown in Fig. 2(b), which are consistent with the experiment data. The singularities at H = ±Hsf reflects breakdown in ΔfH curve. The corresponding ϕa0 and ϕb0 indicate that the macrospins, initially aligning as an A-AFM spin alignment, flop down into the ab plane. By combining the calculated Δf and the experiment results, the dips in measured Δf shown in Fig. 2(a) can be attributed to the occurrence of spin flop. Theoretically, the flopping behavior was a sudden action, which reflects a sharp jump in calculated results. Experimentally, the Néel vector deviated from the c axis in CrPS4 for Hc[30] and the measured results show the Δf gradually decreases below Hsf. It should be noticed that the amplitude of the dip is dependent on the relative orientation of the crystal axes in CrPS4 to the direction of cantilever ˆr. It becomes more pronounced when a axis is mounted closer to ˆr (as detailed in the Supporting Information). Nonetheless, the Hsf derived from the different sample is consistent. Above Hsf, the calculated curve agrees well with the measurement results in S1 at 5 K. The best fits give parameters of M0 = 1.65 kA/m, AV = 0.0047 and K = 12.9 kJ/m3. As shown in Fig. 2(b), the calculated Δf rapidly increases with increasing H. A gradual decrease in its slope indicates that the canted AFM state undergoes a transition from initially metastable state to a steady state. Correspondingly, the spins rotate towards the external field, resulting in an increase in the total magnetic moment. Δf shows a kink at H = Hs, then eventually reaches saturation. For H > Hs, the calculated Δf shows a positive high-field asymptote behavior, where all the spins align in one direction as a field-polarized paramagnetic (FPP) type spin alignment and total magnetization satisfies Ms = 2M0.

    Figure 3 summarized the HT diagram for Hc in CrPS4 flake with different thickness. The diagram has been delineated into three regions. Below Hsf, spins align as an A-AFM ground state. The decrease in Hsf was observed as the temperature increases, indicating the reduction in anisotropy. The region of Hsf < H < Hs, corresponding to the canted spin alignments, was denoted as canted AFM. For H > Hs, the FPP region indicates a field-polarized paramagnetic-type spin alignments.

    Fig. 3.  Magnetic phase diagram of CrPS4 determined from the DCM measurements with H applied along c axis. The spin flop boundaries are represented as triangles (circles for the saturation states) and four patterns denote four different samples.

    To further investigate the temperature dependence of anisotropy in 2D CrPS4, the DCM experiments have been conducted in S1 for Hc. Figure 4(a) shows field dependence of Δf for Hc in a temperature range from 5 K to 40 K. We firstly focus on the data at temperatures below 20 K, where the Δf at low field is positive. With further increasing H, the Δf goes through a maximum, and reaches saturation with a change in slope for large H. The uniaxial model has been extended for the case of Hc and K > 0 (see the Supporting Information for more details). The numerical results, indicating two macrospins gradually rotate to the external field with ϕa0=πϕb0=arcsin(M0HK+2AM20V) from the ac plane, agree well with our experiment data.

    Fig. 4.  Magnetic field dependence of Δf in the case of Hc. (a) The magnetic field dependence of Δf in S1 with Hc from 5 K to 40 K. (b) The zoom-in ΔfH curve at selected temperatures. The data has been normalized and shifted for clarify. The positive M shape and the negative W shape have been highlighted. The frequency shift at zero field is always zero. The dashed lines indicate the value Δf = 0 at different temperatures. Black triangles mark the position of the valley.

    Now we turn to the evolution of metamagnetic phases above 20 K. In Fig. 4(b), the curves have been normalized and shifted for clarity. As temperature increases, the shape of the ΔfH curve gradually deviates from the M shape. Remarkably, a hump emerges right after the bump in the Δf at around H = 1.3 T when T reaches 25 K. This change is induced by a relatively larger in-plane magnetization. In this case, the spins aligns close to ab plane. The low-field ΔfH curve becomes negative above 30 K and maintains a W shape at low field until 35 K. The negative signal indicates the dominance of in-plane anisotropy and the reorientation of the easy axis. The evolution of the curve shape shows that the spin reorientation is a continuous transition. Above 38 K, the Δf shows a quadratic field dependence. The upward parabola indicates that CrPS4 enters a paramagnetic state where the in-plane shape anisotropy dominates.

    The HT phase diagram for Hc has been summarized from different CrPS4 samples. As shown in Fig. 5, the phase diagram is roughly delineated as four regions. The FPP region represents the saturated magnetic phase with two macrospins aligning one direction. PM denots the paramagnetic phase of CrPS4. The FPP-PM phase boundary is more distinct than the discussion before. AFM1 and AFM2 represent different cases for canted spin alignments, respectively. In AFM2, the spins align closer to the a axis. With increasing magnetic field, the spins prefer to align closer to the b axis (AFM1). The boundary between them is determined by the reversal field at the valley of Δf. Due to the different relative directions of crystal axes to the cantilever, the reversal fields vary for different samples. Nonetheless, the evolution of magnetic anisotropy induced by spin reorientation can be directly distinguished by performing DCM experiments.

    Fig. 5.  Magnetic phase diagram of CrPS4 with H applied perpendicular to c axis. The saturation fields (circles) and the reversal fields (triangles) are extracted from the ΔfH curves in Figs. 4(a) and 4(b), Figs. S2 and Fig. S4. The insets depict the corresponding spin alignments in different phases.

    The magnetic phase diagrams for Hc and Hc, constructed by DCM measurements in 2D CrPS4, are basically consistent with that reported in bulk samples. Note that the alteration of spin orientation identified by commercial magnetometry is subtle and indirect. Specifically, the canted spin alignments and the evolution of spin reorientation in CrPS4 can be only revealed by neutron scattering measurements, which is, however, unavailable for individual 2D particles. As discussed above, the effective magnetic spring constant Δk is directly related to anisotropic magnetic susceptibility. The inherent dynamic property of DCM, on the other hand, makes it sensitive to quasi-static states. Moreover, the smallest detectable magnetotropic coefficient in our study can be equivalent to a magnetic moment 4.5 × 106μB at 1 T (see Section S10 in the Supporting Information), which enables effective characterization of ultra weak magnetism even for single-layer samples. In this sense, DCM technique has the potential to investigate the exotic properties in 2D system, including skyrmions,[3335] hidden orders,[36,37] and quantum critical phenomena.[38]

    In summary, we have extended the high-precision DCM measurements to 2D antiferromagnets through a clean transfer technique. The effective magnetic spring constant Δk has been measured in atomically thin CrPS4 flake. Based on the field dependence of magnetic spring constant, the connection between the spin alignments and the Δk of cantilever sensor is established to reveal the metamagnetic phase transitions for both Hc and Hc. The temperature evolution of the ΔfH curve has been analyzed, from which we reveal the occurrence of spin reorientation. Our work pioneers magnetism measurement of a single atomically thin sample and opens a route to effectively explore novel phenomenon in 2D materials.

  • [1]
    Mazin I, Singh D, Johannes M and Du M 2008 Phys. Rev. Lett. 101 057003 doi: 10.1103/PhysRevLett.101.057003

    CrossRef Google Scholar

    [2]
    Norman M R, Pines D and Kallin C 2005 Adv. Phys. 54 715 doi: 10.1080/00018730500459906

    CrossRef Google Scholar

    [3]
    Johnston D C 2010 Adv. Phys. 59 803 doi: 10.1080/00018732.2010.513480

    CrossRef Google Scholar

    [4]
    Kivelson S A, Bindloss I P, Fradkin E, Oganesyan V, Tranquada J M, Kapitulnik A and Howald C 2003 Rev. Mod. Phys. 75 1201 doi: 10.1103/RevModPhys.75.1201

    CrossRef Google Scholar

    [5]
    Wilson J A, Di Salvo F J and Mahajan S 1974 Phys. Rev. Lett. 32 882 doi: 10.1103/PhysRevLett.32.882

    CrossRef Google Scholar

    [6]
    Wilson J A, Di Salvo F J and Mahajan S 2001 Adv. Phys. 50 1171 doi: 10.1080/00018730110102718

    CrossRef Google Scholar

    [7]
    Ali M N, Xiong J, Flynn S, Tao J, Gibson Q D, Schoop L M, Liang T, Haldolaarachchige N, Hirschberger M, Ong N P and Cava R J 2014 Nature 514 205 doi: 10.1038/nature13763

    CrossRef Google Scholar

    [8]
    Soluyanov A A, Gresch D, Wang Z, Wu Q, Troyer M, Dai X and Bernevig B A 2015 Nature 527 495 doi: 10.1038/nature15768

    CrossRef Google Scholar

    [9]
    Sun Y, Wu S C, Ali M N, Felser C and Yan B H 2015 Phys. Rev. B 92 161107 doi: 10.1103/PhysRevB.92.161107

    CrossRef Google Scholar

    [10]
    Morosan E, Zandbergen H W, Dennis B S, Bos J W G, Onose Y, Klimczuk T, Ramirez A P, Ong N P and Cava R J 2006 Nat. Phys. 2 544 doi: 10.1038/nphys360

    CrossRef Google Scholar

    [11]
    Baugher B, Churchill H, Yang Y and Jarillo-Herrero P 2013 Nano Lett. 13 4212 doi: 10.1021/nl401916s

    CrossRef Google Scholar

    [12]
    Chhowalla M, Shin H, Eda G, Li L, Loh K and Zhang H 2013 Nat. Chem. 5 263 doi: 10.1038/nchem.1589

    CrossRef Google Scholar

    [13]
    Li H, Wu J, Yin Z and Zhang H 2014 Acc. Chem. Res. 47 1067 doi: 10.1021/ar4002312

    CrossRef Google Scholar

    [14]
    Jariwala D, Sangwan V, Lauhon L, Marks T and Hersam M 2014 ACS Nano 8 1102 doi: 10.1021/nn500064s

    CrossRef Google Scholar

    [15]
    Lv R, Robinson J, Schaak R, Sun D, Sun Y, Mallouk T and Terrones M 2015Acc. Chem. Res. 48 56 doi: 10.1021/ar5002846

    CrossRef Google Scholar

    [16]
    Zhang J, Liu J, Huang J L, Kim P and Lieber C M 1996 Science 274 757 doi: 10.1126/science.274.5288.757

    CrossRef Google Scholar

    [17]
    Kim J, Park C, Yamaguchi W, Shiino O, Kitazawa K and Hasegawa T 1997 Phys. Rev. B 56 R15573R doi: 10.1103/PhysRevB.56.R15573

    CrossRef Google Scholar

    [18]
    Wang H, Lee J, Dreyer M and Barker B 2009 J. Phys.: Condens. Matter 21 265005 doi: 10.1088/0953-8984/21/26/265005

    CrossRef Google Scholar

    [19]
    Duerloo K N, Li Y and Reed E J 2014Nat. Commun. 5 4214 doi: 10.1038/ncomms5214

    CrossRef Google Scholar

    [20]
    Cho D, Cheon S, Kim K, Lee S, Cho Y, Cheong S and Yeom H 2016 Nat. Commun. 7 10956 doi: 10.1038/ncomms10956

    CrossRef Google Scholar

    [21]
    Ma L, Ye C, Yu Y, Lu X, Niu X, Kim S, Feng D, Tománek D, Som Y, Chen X and Zhang Y 2016Nat. Commun. 7 10453 doi: 10.1038/ncomms10453

    CrossRef Google Scholar

    [22]
    Huang H, Fan X, Singh D, Chen H, Jiang Q and Zheng W 2016 Phys. Chem. Chem. Phys. 18 4086 doi: 10.1039/C5CP06706E

    CrossRef Google Scholar

    [23]
    Moncton D E, Axe J D and DiSalvo F J 1975 Phys. Rev. Lett. 34 734 doi: 10.1103/PhysRevLett.34.734

    CrossRef Google Scholar

    [24]
    Ikebe M, Katagiri K, Noto K and Muto Y 1980 Phys. B+C 99 209 doi: 10.1016/0378-43638090234-X

    CrossRef Google Scholar

    [25]
    Ishioka J, Liu Y H, Shimatake K, Kurosawa T, Ichimura K, Toda Y, Oda Y and Tanda S 2010 Phys. Rev. Lett. 105 176401 doi: 10.1103/PhysRevLett.105.176401

    CrossRef Google Scholar

    [26]
    Ishioka J, Fujii T, Katono K, Ichimura K, Kurosawa T, Oda M and Tanda S 2011 Phys. Rev. B 84 245125 doi: 10.1103/PhysRevB.84.245125

    CrossRef Google Scholar

    [27]
    Iavarone M, Di Capua R, Zhang X, Golalikhani M, Moore S and Karapetrov G 2012 Phys. Rev. B 85 155103 doi: 10.1103/PhysRevB.85.155103

    CrossRef Google Scholar

    [28]
    Novello A, Hildebrand B, Scarfato A, Didiot C, Monney G, Ubaldini A, Berger H, Bowler D, Aebi P and Renner C 2015 Phys. Rev. B 92 081101R doi: 10.1103/PhysRevB.92.081101

    CrossRef Google Scholar

    [29]
    Novello A M, Spera M, Scarfato A, Ubaldini A, Giannini E, Bowler D R and Renner C 2017 Phys. Rev. Lett. 118 017002 doi: 10.1103/PhysRevLett.118.017002

    CrossRef Google Scholar

    [30]
    Nakata Y, Sugawara K, Shimizu R, Okada Y, Han P, Hitosugi T, Ueno K, Sato T and Takahashi T 2016 NPG Asia Mater. 8 e321 doi: 10.1038/am.2016.157

    CrossRef Google Scholar

    [31]
    Dai Z, Xue Q, Gong Y, Slough C G and Coleman R V 1993 Phys. Rev. B 48 14543 doi: 10.1103/PhysRevB.48.14543

    CrossRef Google Scholar

    [32]
    Soumyanarayanan A, Yee M, He Y, van Wezel J, Rahn D, Rossnagel K, Hudson E, Norman M and Hoffman J 2013 Proc. Natl. Acad. Sci. USA 110 1623 doi: 10.1073/pnas.1211387110

    CrossRef Google Scholar

    [33]
    Rahnejat K C, Howard C A, Shuttleworth N E, Schofield S R, Iwaya K, Hirjibehedin C F, Renner Aeppli G and Ellerby M 2011 Nat. Commun. 2 558 doi: 10.1038/ncomms1574

    CrossRef Google Scholar

    [34]
    Flicker F and van Wezel J 2015 Phys. Rev. B 92 201103R doi: 10.1103/PhysRevB.92.201103

    CrossRef Google Scholar

    [35]
    Hess H F, Robinson R B, Dynes R C, Valles Jr. J M and Waszczak J V 2001 J. Vac. Sci. Technol. B 19 1150 doi: 10.1116/1.1379796

    CrossRef Google Scholar

    [36]
    Xu J, Wang M, Liu Z, Ge J, Yang X, Liu C, Xu Z, Guan D, Gao C, Qian D, Liu Y, Wang Q, Zhang F, Xue Q and Jia J 2015 Phys. Rev. Lett. 114 017001 doi: 10.1103/PhysRevLett.114.017001

    CrossRef Google Scholar

    [37]
    Sun H, Zhang K, Hu L, Li C, Wang G, Ma H, Xu Z, Gao C, Guan D, Li Y, Liu C, Qian D, Zhou Y, Fu L, Li S, Zhang F and Jia J 2016 Phys. Rev. Lett. 116 257003 doi: 10.1103/PhysRevLett.116.257003

    CrossRef Google Scholar

    [38]
    He W, Zhou B, He J, Zhang T and Law K 2016 arXiv:1604.02867

    Google Scholar

    [39]
    Horcas I, Fernández R, Gómez-Rodríguez J M, Colchero J, Gómez-Herrero J and Baro A M 2007 Rev. Sci. Instrum. 78 013705 doi: 10.1063/1.2432410

    CrossRef Google Scholar

  • Related Articles

    [1]CAO Can, CHEN Ling-Na, JIA Shu-Ting, ZHANG Dan, XU Hui. First-Principles Study on Electronic Structures and Optical Properties of Doped Ag Crystal [J]. Chin. Phys. Lett., 2012, 29(3): 037101. doi: 10.1088/0256-307X/29/3/037101
    [2]LU Yong-Fang, SHI Li-Qun, DING Wei, LONG Xing-Gui. First-Principles Study of Hydrogen Impact on the Formation and Migration of Helium Interstitial Defects in hcp Titanium [J]. Chin. Phys. Lett., 2012, 29(1): 013102. doi: 10.1088/0256-307X/29/1/013102
    [3]LI Deng-Feng, XIAO Hai-Yan, XUE Shu-Wen, YANG Li, ZU Xiao-Tao. Surface Structure and Electronic Property of InP(001)-(2×1)S Surface: A First-Principles Study [J]. Chin. Phys. Lett., 2010, 27(4): 046802. doi: 10.1088/0256-307X/27/4/046802
    [4]JIANG Xiao-Shu, MI Shu, SUN Peng-Jun, LU Yuan, LIANG Jiu-Qing. Bulk Modulus and Electronic Band Structure of ZnGa2X4(X=S,Se): a First-Principles Study [J]. Chin. Phys. Lett., 2009, 26(7): 077102. doi: 10.1088/0256-307X/26/7/077102
    [5]FANG Zhi-Jie, FANG Cheng, SHI Li-Jie, LIU Yong-Hui, HE Man-Chao. First-Principles Study of Defects in CuGaO2 [J]. Chin. Phys. Lett., 2008, 25(8): 2997-3000.
    [6]DING Zong-Ling, XING Huai-Zhong, XU Sheng-Lan, HUANG Yan, CHENXiao-Shuang. First-Principles Study of Electronic Properties in PbS(100) with Vacancy Defect [J]. Chin. Phys. Lett., 2007, 24(11): 3218-3221.
    [7]LIU Ting-Yu, ZHANG Qi-Ren, ZHUANG Song-Lin. First-Principles Studies on Electronic Structures and Absorption Spectra of PbWO4 Crystals with Defect [VPb2---VO2+ --VPb2-]2- [J]. Chin. Phys. Lett., 2007, 24(8): 2361-2364.
    [8]WU Qing-Hua, HU Qian-Ku, LUO Xiao-Guang, YU Dong-Li, LI Dong-Chun, HE Ju-Long. First-Principles Study of Structural and Electronic Properties of Layered B2CN Crystals [J]. Chin. Phys. Lett., 2007, 24(1): 180-183.
    [9]CHEN Teng, LIU Ting-Yu, ZHANG Qi-Ren, LI Fang-Fei, YI Zhi-Jun, TIAN Dong-Sheng, ZHANG Xiu-Yan. First-Principles Study on the Electronic Structures for Y3+:PbWO₄ Crystals [J]. Chin. Phys. Lett., 2006, 23(12): 3323-3326.
    [10]WANG Yi, SUN Yun-Feng. Thermodynamic Properties of Sodium: A First-Principles Study [J]. Chin. Phys. Lett., 2001, 18(7): 864-867.

Catalog

    Article views (313) PDF downloads (890) Cited by()

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return