This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Weak localization competes with the quantum oscillations in a natural electronic superlattice: the case of Na1.5(PO2)4(WO3)20

Kamil K. Kolincio [email protected] Laboratoire CRISMAT, UMR 6508 du CNRS et de l’Ensicaen, 6 Bd Marechal Juin, 14050 Caen, France. Faculty of Applied Physics and Mathematics, Gdansk University of Technology, Narutowicza 11/12, 80-233 Gdansk, Poland    Olivier Pérez Laboratoire CRISMAT, UMR 6508 du CNRS et de l’Ensicaen, 6 Bd Marechal Juin, 14050 Caen, France.    Enric Canadell Institut de Ciència de Materials de Barcelona (ICMAB-CSIC), Campus Bellaterra, 08193 Bellaterra, Barcelona, Spain    Pere Alemany Departament de Ciència de Materials i Química Fisica and Institut de Química Teòrica i Computacional (IQTCUB), Universitat de Barcelona, Martí i Franquès 1, 08028 Barcelona, Spain    Elen Duverger-Nédellec Laboratoire CRISMAT, UMR 6508 du CNRS et de l’Ensicaen, 6 Bd Marechal Juin, 14050 Caen, France. Charles University, Faculty of Mathematics and Physics, Condensed Matter Department, Ke Karlovu 3, 121 16 Praha 2, Czech Republic    Arianna Minelli ESRF - The European Synchrotron, 71, Avenue des Martyrs, F-38000 Grenoble, France    Alexei Bosak ESRF - The European Synchrotron, 71, Avenue des Martyrs, F-38000 Grenoble, France    Alain Pautrat Laboratoire CRISMAT, UMR 6508 du CNRS et de l’Ensicaen, 6 Bd Marechal Juin, 14050 Caen, France.
Abstract

We report an investigation of the combined structural and electronic properties of the bronze Na1.5(PO2)4(WO3)20. Its low dimensional structure and possible large reconstruction of the Fermi surface due to charge density wave instability makes this bulk material a natural superlattice with a reduced number of carriers and Fermi energy. Signatures of multilayered 2D electron weak localization are consequently reported, with an enhanced influence of quantum oscillations. A crossover between these two antagonistic entities previously observed only in genuine low dimensional materials and devices, is shown to occur in a bulk crystal due to its hidden 2D nature.

The unfading interest in low dimensional conducting materials originates from their remarkable electronic properties. High critical temperature in superconductors Kharkov and Sushkov (2016); Cyr-Choinière et al. (2017), improved thermoelectric properties Xie et al. (2009); Mao et al. (2016); Kolincio et al. (2016a); Molina-Valdovinos et al. (2018), instability of the electron gas into a charge density wave (CDW) Monceau (2012); Grüner (1988), are consequences of a pronounced structural and electronic anisotropy. In particular, quasi 1D or 2D conductors exhibit a partial Fermi surface (FS) reconstruction at a CDW transition, accompanied by strong reduction of free electron density, increase of the remaining carriers mobilities, and enhanced magnetoresistance Yasuzuka et al. (1999); Schlenker et al. (1985).

In addition, quantum corrections may strongly impact the low temperature transport properties. Weak localization (WL) phenomenon is a well known example. Another characteristic of low temperature magnetoresistance effects in metals, especially of those with high carrier mobility, is the existence of Shubnikov de Haas (SdH) quantum oscillations. They originate from the quantization of electron density of states into Landau levels when exposed to a magnetic field, and are observed in high purity samples in the limit ωcτ\omega_{c}\tau\gg 1, where ωc=eBm\omega_{c}=\frac{eB}{m^{*}} is the cyclotron frequency, mm^{*} the effective mass of carriers, and τ\tau the elastic scattering time). WL and SdH effects are a priori antagonistic because they are respectively favored and limited by disorder. However, due to the decrease of the WL contribution with magnetic field and with the restricted condition ωcτ(kF)1\omega_{c}\tau\leq(k_{F}\ell)^{-1} (kFk_{F} is the Fermi momentum and ll the electronic mean free path), a crossover can be observed between these two quantum processes, whose magnetic field range has been related to interaction effects Sedrakyan and Raikh (2008). This crossover has been previously observed in rare systems showing strong 2D characteristics: 2D electron gas heterostructures with high mobilities Wang et al. (2018), quantum wells Sedrakyan and Raikh (2008); Tayari et al. (2015); Sarcan et al. (2018), and graphene Heer et al. (2010); et al. (2014). We show here that both WL and SdH are for the first time observed in a single crystal whose 2D character arises from the carriers confinement in conducting layers with a thickness of few atomic planes. The sample under study belongs to the monophosphate tungsten bronze (MPTB) family Ax(PO4)2(WO3)2m (A = K, Na or Pb). In this family, carriers originate from the PO4 groups, and are delocalized in the middle of WO3 layers, giving a pronounced anisotropic electronic structure Roussel et al. (2001); Foury-Leylekian and Pouget (2002). The carrier density can be tuned by changing mm due to the increase of unit cell for constant carrier numbers and/or by the insertion of A cations with metallic character Roussel et al. (2000). These A cations are inserted in hexagonal tunnels to give Ax(PO4)2(WO3)2mDomengés et al. (1988). Supplemental interest of this bronze family is that most of the investigated members have shown CDW transitions that have been explained using the hidden nesting concept Canadell and Whangbo (1991); Kolincio et al. (2016b).

Refer to caption
Figure 1: a) Projection along b of the structure of Na1.5P4W20O68Na_{1.5}P_{4}W_{20}O_{68}. Blue octahedra are (WO6WO_{6}), green tetrahedra are (PO4PO_{4}) and purple a circles are Na atoms. The [15 0 7] and [15 0 -4] directions are showing the short and long direction (WO6WO_{6}) chains; they are 90 oriented. b) (0kl)* c) (hk0)* planes assembled from the frames collected at 130 K on the ID28 beamline at ESRFGirard et al. (2019). Two directions of diffuse scattering are shown by green and blue arrows. Intensifications of the diffuse signal are observed at the intersection of the two types of diffusion scattering for 12\frac{1}{2}a* (see blue dashed rectangle).

X-ray diffraction investigations performed at room temperature reveal that the whole pattern can be described using a monoclinic cell (a=6.5483(6)a=6.5483(6)Å, b=5.2893(5)b=5.2893(5)Å, c=18.0789(15)c=18.0789(15)Å, β=94.7000(4)\beta=94.7000(4)^{\circ}) and the space group P21/mP2_{1}/m; these cell parameters are in agreement with those reported for PbxP4W20O68  on the basis of a pattern matching analysis performed on powder diffraction data Roussel et al. (1998). The final agreement factor for our data is 4.7% and the refined chemical composition is Na1.49(8)P4W20O68; Na sites are partly occupied. For details of the data collection, refinement and the atomic structure parameters see the Supplementary Material sup .

A projection along the b axis is shown in Figure 1a. The main structural characteristics expected for the MPTB family in the fundamental state are observed. Slabs of the (WO6) edge sharing octahedra alternate with slices of (PO4) tetraehdra; at the junction, hexagonal tunnels host Na atoms. Two orthogonal directions [15 0 7] and [15 0 -4] can be highlighted; they correspond to chains of 10 and 5 edge sharing (WO6) octahedra respectively (see Figure  1a).

To check the possibility of a CDW transition, full data collections were performed at RT and 130 K on the ID28 beamline at ESRF synchrotron. There is no evidence for additional scattering signal at RT but at 130K diffuse scattering can be observed on different experimental frames. An accurate study of the (hk0)* plane (see Figures 1c) reveals an enhancement of the diffuse signal at the intersection of two diffuse planes for the 12\frac{1}{2}a* wave vector. An analysis of the shift between the positions of the diffuse scattering in subsequent (0kl)*, (1kl)*, (2kl)* reciprocal layers, reveals that the diffuse sheets intersecting them run along the direction [15 0 7] (see Supplementary Material sup for details).

The planes running along [15 0 7] are perpendicular to [-7 0 15]*; but [-7 0 15]* and [15 0 -4] fortuitously correspond to the same direction (see Supplementary Materials sup ) the direction of the chain of 10 octahedra. Then the observation of the diffuse scattering can be correlated with an ordering within the longest (WO6) chain but with a loss of order both along b and along [15 0 -4] i.e. with a loss of order with the adjacent chains. The existence of diffuse scattering planes evidences the 1D ordering, possibly associated to CDW. Nevertheless, the absence of their condensation to sharp reflections shows that this ordering shows at T = 130 K a moderate coherence length, instead of establishing a long range order.

Refer to caption
Figure 2: a) Electronic band structure calculated for Na1.5(PO2)4(WO3)20. Zero energy corresponds to the Fermi level and the different points are defined as Γ\Gamma = (0, 0, 0), X = (1/2, 0, 0), Y = (0, 1/2, 0), S = (1/2, 1/2, 0) and Z = (0, 0, 1/2), respectively in the units of the reciprocal vectors. b) section of the Fermi surface for c=c^{*}= 0. c) 3D representation of the Fermi surface. d) calculated Lindhard function.

Having established the crystalographic parameters, we studied the electronic structure of Na1.5(PO2)4(WO3)20. The results of band calculations performed with Generalised Gradient Approximation (GGA) (see supplementary material for details sup ) are shown in Fig. 2. There are large similarities between the studied compound and other MPTB members showing CDW Wang et al. (1989); Canadell and Whangbo (1991). As seen in Fig. 2a, three bands are crossing the Fermi level resulting in FS composed of three quasi 1D elements: three pairs of parallel planes running along (a+b)(a^{*}+b^{*}), (ab)(a^{*}-b^{*}), and bb^{*} directions with almost no dispersion along cc^{*} as shown in Figures 2a) and c). While the presence of similar flat regions in MPTB hosting CDW gives a rise to hidden 1D nesting Wang et al. (1989); Canadell and Whangbo (1991), their almost perfect alignment in Na1.5(PO2)4(WO3)20 additionally enhances the susceptibility for such a scenario. The relevance of this mechanism is additionally supported by the calculation of a Linhard function (Fig. 2d). In particular three very intense features are near the a*/2 wave vector suggest that large parts of the Fermi surface are susceptible to destruction by either an inherent instability of the Fermi surface or coupling with another type of instability of the lattice with this wave vector.

Fig. 3a shows the thermal variation of resistivity ρ\rho of a Na1.5(PO2)4(WO3)20 single crystal. Metallic behavior (dρdT<0\frac{d\rho}{dT}<0) is observed from 400 K down to 10 K, where a smooth increase of resistivity appears. The electronic resistivity follows a -lnTT variation at low temperature, and this experimental observation suggests a WL contribution. The expression of the WL correction to conductance for non interacting electrons is Bergmann (1984):

ΔG=pe22π2lnTT0\Delta G=\frac{pe^{2}}{2\pi^{2}\hbar}ln\frac{T}{T_{0}} (1)

where pp describes the temperature dependence of the inelastic scattering rate τiTp\tau_{i}\propto T^{-p}. p is equal to 1 for phase breaking dominated by Coulomb interactions and to 3 for electron-phonon scattering Lee and Ramakrishnan (1985). Fitting the low temperature part gives dGdlnT\frac{dG}{dlnT} = 0.0517 Ω1\Omega^{-1}, to compare with e2p2π\frac{e^{2}p}{2\pi\hbar} = 1.23 10-5 Ω1\Omega^{-1} expected for a 2D WL contribution. The large difference between these values can be explained if our sample is considered as a superlattice of N independent and non-interacting 2D layers, giving rise to a multilayered WL effect Moyle et al. (1987); Stormer et al. (1986); Szott et al. (1989a, b, 1992). Dividing the measured conductivity by the theoretical WL conductivity of a single layer gives N \approx 4300, considering a value p = 1, as expected for electron scattering in d-electron systems. Multiplying N by a thickness of the W-O conducting layer - approximately equal to the c lattice constant, we obtain \approx 10 μ\mum, comparable to the thickness of the sample. The weak disorder causing WL can originate from a weakly nonuniform distribution of Na atoms in the tunnels Kolincio et al. (2016b), stacking faults reported for MPTB with high mm parameter Domengés et al. (1983) or can be contributed by the short order of charge density waves state.

Refer to caption
Figure 3: a) In - plane resistivity versus temperature. Inset: expanded view of the low temperature region shown with logarithmic horizontal scale. Dashed line is a guide for the eye. b) Magnetoresistance as a function of magnetic field applied parallel to c axis. Inset: MR(B) plots at TT = 1.8 K for various θ\theta angles between the magnetic field direction and the c axis. c) Magnetoconductance versus magnetic field for selected temperatures. The red curves show the data fits with equation (2).

Since other mechanisms can lead to a logarithmic increase of resistivity at low temperature (Kondo scattering, strong Coulomb interaction), we analyzed the magnetoresistance (MR=ρ(B)ρ(0)ρ(0)100%MR=\frac{\rho(B)-\rho(0)}{\rho(0)}\cdot 100\%) in order to confirm the underlying mechanism. The low temperature magnetoresistance depicted in Figure 3b is non-monotonic. It is negative for moderate BB, and upon increasing the magnetic field, it shows a minimum followed by a notable increase. For T \leq 18 K, large oscillations of the resistance can be observed in the increasing part of MR(B)MR(B).

As a strong argument for WL, we observe a negative MR for B perpendicular to the conducting plane (θ\theta = 0o) but it disappears in the parallel configuration (θ\theta = 90o) - see inset of Fig. 3b. This is precisely what is expected in the 2D WL regime, where the orbital effect responsible for the MR is suppressed for parallel fields Meyer et al. (2002). The low field magnetoconductance ΔG=G(B)G(0)\Delta G=G(B)-G(0) in the WL regime, can be described by the Hikami-Larkin-Nagaoka (HKN) expression Hikami et al. (1980); Lee and Ramakrishnan (1985):

ΔG=e22π2(ψ(12+1A)+ln(A)),\begin{split}\Delta G=\frac{e^{2}}{2\pi^{2}\hbar}\left(\psi\left(\frac{1}{2}+\frac{1}{A}\right)+ln(A)\right),\end{split} (2)

where ψ\psi is the digamma function and AA = 4eDτiB\frac{4eD\tau_{i}B}{\hbar}. To show the increase of magnetoconductance in the magnetic field, we have quantified and removed the classical \sim B2 magnetoresistance component. Then, we fit the low field curves with the WL expression considering the number of 4300 independent layers as found from the thermal dependence of conductivity. As shown in Fig. 3c, the fits show good agreement with the experimental data. The diffusion length of localization (Thouless length) Lϕ = (Dτi\tau_{i})1/2 was extracted from the fitting parameter and reported in the Fig. 4a. It is worth noting that Lϕ is larger than the thickness of the conducting layer and approximately equal to the periodicity along the cc axis, confirming the regime of 2D coherence. When increasing the magnetic field, the dephasing regime emerges for LB = (2eBϕ)1/2Lϕ\left(\frac{2\hbar}{eB_{\phi}}\right)^{1/2}\geq L_{\phi}. The effect of the magnetic field is then no longer relevant for WL and a negative contribution disappears. We have compared the value of BϕB_{\phi} extracted from the fitting with the condition L = LB and the magnetic field value B where the G(B) ceases to grow with B and where quantum oscillations immediately emerge. We find that these values are very close to each other (for example for 4, 8 and 10 K the values obtained from the fits yield 5.27 T, 5.36 T, and 5.49 T respectively to compare with observed B = 5.45 T, 5.31 T and 5.53 T, respectively). For lowest temperatures, LϕL_{\phi} ceases to grow as the temperature is lowered, and even drops for 1.8 K, where the negative MR term also starts to decrease. A plausible explanation for this effect is the further crossover from weak localization to weak antilocalization (WAL)Lang et al. (2013); Liu et al. (2012) due to strong spin-orbit effects observed in tungsten rich compounds Sheng et al. (2017); McCormick et al. (2017). This scenario is also supported by the emergence of sharp cusp-like features in the low field limit of magnetoresistance (see Supplementary Materials), which is characteristic for WAL regimeBergmann (1984); Thomas et al. (2016). A further examination of this term, to obtain reliable HKN fit with the introduction of the antilocalization component will require the transport measurements at temperatures below 1.8 K.

Refer to caption
Figure 4: a) Temperature variation of the localization length, obtained from the magnetoconductance fits with equation 2 b) Power spectrum of the oscillatory component of magnetoresistance. Solid line shows the Lorentzian fit to the experimental data. c) the geometrical dependence of the oscillation frequency. The dashed line is a guide for the eye. d) Thermal variation of the oscillation amplitude.

The oscillations in magnetoresistance emerge immediately after the negative WL term is suppressed, as evidenced in Fig. 3c. They are periodic as a function of 1/B, as expected for Shubnikov-de Haas quantum oscillations. We have used the Lifshitz-Kosevich (LK) theory to extract the electronic parameters Lifshitz and Kosevich (1956). The Fourier transform of the oscillating part of magnetoresistance gives the power spectrum shown in Fig. 4b. A peak can be observed at the frequency F1F_{1}= 69 T for θ\theta = 0o. This value yields the area A1A_{1} of an extremal Fermi surface cross section that is perpendicular to B, through the Onsager relation Shoenberg (1984) Ai=2πeFi/A_{i}=2\pi eF_{i}/\hbar. It corresponds to a FS cross section A1A_{1}= 0.68 nm-2, i.e. to a very small part (\approx 0.5 %) of the first Brillouin zone, estimated to be 110 nm-2. A plausible explanation for the presence of such a small pocket is the Fermi surface decomposition associated with charge density wave. This scenario is supported by the X-ray diffuse scattering and electronic structure calculations showing a potential for strong nesting around a*/2, which most likely occurs at high temperature as for other high m values of the MPTB family Ottolenghi, Alberto and Pouget, Jean-Paul (1996). The temperature dependence of the quantum oscillations amplitude is depicted in Fig. 4c. It can be related to the effective mass of quasi-particles m through the temperature damping factor. We find m= 0.83 me with me being the free electron mass, indicating that the electron-electron interactions are moderate as theoretically predicted for the WL-SdH crossoverSedrakyan and Raikh (2008). The field dependence of the quantum oscillations allows to estimate the Dingle temperature TD \approx 6 K, corresponding to an electronic mean free path ll\approx 62 nm. Since the Fermi momentum is kF(A1/π)1/2k_{F}\approx(A_{1}/\pi)^{1/2} = 0.47 nm-1, it leads to kFlk_{F}l\approx 29\gg 1 confirming the diffusive regime Minkov et al. (2002); Pusep et al. (2004). The angle θ\theta between BB and the cc axis was varied from 0 to 90o and the SdH oscillations can be observed within experimental resolution up to θ\theta\approx 50o. The frequency of the SdH oscillations clearly shifts to larger values as θ\theta is increased, with an approximate 1/cosθ\theta variation, as depicted in the inset of Figure 4b. This is expected for quasi-2D Fermi surface topology and is also consistent with the angular dependence of magnetoresistance shown in the inset of figure 3b. For BB perpendicular to the conducting plane, MR initially attenuated by the negative contribution, steeply increases with B2\sim B^{2} in the high field limit, where the WL phase coherence is broken, while for in-plane orientation MR is sublinear with BB which gives a tendency for saturation as BB\rightarrow\infty. Such a behavior is characteristic for open and closed orbits, respectively, expected for a cylinder-like FSPippard (2009).

Such a reduced, cylindrical Fermi surface is a characteristic feature of 2D systems, as high mobility ultrathin films and heterostructures Wu et al. (2017); Gillgren et al. (2014); Kozuka et al. (2009); Li et al. (2015) and unparalleled cases of pure Bi and graphene, where the FS covers only 10-5 and 10-6 of the Brillouin zone, respectively Soule et al. (1964); Bhargava (1967); Behnia et al. (2007); Édel’man (1976). The renormalization of the Na1.5(PO2)4(WO3)20 electronic structure driven by the charge density wave, causing the lowering of the Fermi wavevectors brings this bulk system close to a natural superlattice, with reduced number of carriers and/or reduced Fermi energy, which promotes the quantum phenomena.

In conclusion, we report the structural and transport properties of the monophosphate tungsten bronze Na1.5(PO2)4(WO3)20. We demonstrate that this single crystal shows multilayered 2D weak localization effects. As soon as the dephasing by magnetic field dominates and breaks the WL contribution, there is a sudden crossover to the large quantum oscillations regime. This shows that natural crystals with 2D characteristics and reduced carriers density possibly due to a CDW condensation can be used as promising systems for investigating enhanced quantum transport properties.

Acknowledgements.
Acknowledgements: A.P would like to thank Thierry Klein and Andrea Gauzzi for valuable remarks on the competition between weak localization and quantum oscillations conditions which motivated our analysis. Financial support by the ANR Projects: ANR-18-CE92-0014 and ANR-11-BS04-0004 is gratefully acknowledged. Work in Spain was supported by MICIU (PGC2018-096955-B-C44 and PGC2018-093863-B-C22), MINECO through the Severo Ochoa (SEV-2015-0496) and Maria de Maeztu (MDM-2017-0767) Programs and the Generalitat de Catalunya (2017SGR1506 and 2017SGR1289). E.Duverger-Nédellec was supported by the project NanoCent-Nanomaterials center for advanced applications, project no. CZ.02.1.01/0.0/0.0/15_003/0000485, financed by the ERDF.

References