6
PHYSICAL REVIEW B 95, 054409 (2017) Spin pseudogap in the S = 1 2 chain material Sr 2 CuO 3 with impurities G. Simutis, 1 , * S. Gvasaliya, 1 N. S. Beesetty, 2 T. Yoshida, 3 J. Robert, 4 S. Petit, 4 A. I. Kolesnikov, 5 M. B. Stone, 6 F. Bourdarot, 7 H. C. Walker, 8 D. T. Adroja, 8 O. Sobolev, 9 C. Hess, 10 T. Masuda, 3 A. Revcolevschi, 2 B. B¨ uchner, 10 and A. Zheludev 1 1 Neutron Scattering and Magnetism, Laboratory for Solid State Physics, ETH Z¨ urich, Z¨ urich, Switzerland 2 Synthese, Proprietes et Modelisation des Materiaux, Universite Paris-Sud, 91405 Orsay Cedex, France 3 Institute for Solid State Physics, University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa, Chiba 277-8581, Japan 4 Laboratoire L´ eon Brillouin, CEACNRS, CEA-Saclay, F-91191 Gif-sur-Yvette, France 5 Chemical and Engineering Materials Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA 6 Quantum Condensed Matter Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831-6393, USA 7 Mod´ elisation et dexploration de la mati´ ere, Universit´ e Grenoble Alpes et CEA, INAC, 17 rue des Martyrs, 38054 Grenoble, France 8 ISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 OQX, United Kingdom 9 Forschungsneutronenquelle Heinz Maier-Leibnitz (FRM-II), TU Munchen, D-85747 Garching, Germany 10 Leibniz Institute for Solid State and Materials Research IFW Dresden, P.O. Box 270116, D-01171 Dresden, Germany (Received 11 November 2016; revised manuscript received 23 January 2017; published 7 February 2017) The low-energy magnetic excitation spectrum of the Heisenberg antiferromagnetic S = 1/2 chain system Sr 2 CuO 3 with Ni and Ca impurities is studied by neutron spectroscopy. In all cases, a defect-induced spectral pseudogap is observed and shown to scale proportionately to the number of scattering centers in the spin chains. DOI: 10.1103/PhysRevB.95.054409 I. INTRODUCTION Defects play a special role in one dimension due to its intrinsic topology [1,2]. This is especially pronounced in quantum magnets where both the ground states and the excitations are modified upon the introduction of imperfec- tions in magnetic lattices [310]. Recent experiments have demonstrated that when impurities are added to the proto- typical Heisenberg antiferromagnetic S = 1/2 chain materials SrCuO 2 and Sr 2 CuO 3 , their magnetism is suppressed [1116]. Long-range magnetic ordering arising from residual three- dimensional interactions becomes inhomogeneous and occurs at lower temperatures [11,16]. Additionally, the spin-lattice relaxation rate 1/T 1 measured by NMR experiments drops at low temperatures, suggesting the depletion of low-energy excitations [12,15,17]. Defects also affect the heat transport and compromise its ballistic nature [18,19]. The influence of defects on the magnetic excitation spec- trum can be directly probed using neutron scattering. We have previously employed this method to demonstrate the opening of a spin pseudogap in the Heisenberg spin-chain system SrCuO 2 with chain-breaking Ni 2+ impurities [13]. We proposed a simple chain-fragmentation model based on previous theoretical work [20] and derived quantitative predictions for the excitation spectrum in the presence of defects. This model was able to account for the data over a wide temperature range, on an absolute scale, and with no adjustable parameters. Nevertheless, a few questions remained to be answered: (i) Does the size of the pseudogap indeed scale with the impurity concentration as predicted? (ii) What is the role of the double-chain structure of SrCuO 2 ? and (iii) What effect would other types of impurities have? The present paper aims to settle these outstanding issues. Here we study a related spin-chain compound, namely, Sr 2 CuO 3 . It has the benefit of having a simpler single-chain * [email protected] structure, as opposed to the paired chains in SrCuO 2 [2123]. Two types of impurities are investigated: S = 1 Ni 2+ ions, which replace the S = 1/2 Cu 2+ ions in the chain [11,16,20]; and Ca 2+ ions, which replace Sr 2+ and therefore affect the Cu 2+ chains only indirectly [17,19]. II. EXPERIMENT Single crystals of Sr 2 CuO 3 with Ni and Ca impurities were grown using floating zone furnaces as described in earlier re- ports [15,19]. They crystalize in an orthorhombic Immm space group, with lattice parameters of a = 3.9089, b = 3.4940, and c = 12.6910 ˚ A for the pure compound [24]. The spin chains formed by the Cu 2+ ions run along the crystallographic b direction. The exchange interaction between the spins has been estimated as J = 241(11) meV [23]. While the chains are very well isolated, three-dimensional ordering still takes place due to residual three-dimensional interactions. In the pure system the spins order at T N = 5.4K[22] and the ordering temperature is rapidly decreased with the introduction of impurities [11]. All of the measurements presented here were performed above the three-dimensional ordering temperatures. The measurement temperatures were much lower than the temperature corresponding to the intrachain exchange energy. In order to study the effects of disorder, diligent attention had to be given to controlling the level of impurities. In addition to carefully monitoring the ingredients in the growth procedure and performing energy-dispersive x-ray analysis measurements, we have also measured the magnetic suscep- tibility in order to estimate the extent of chain fragmentation. The small samples used for susceptibility measurements were cut from the same crystals used for neutron spectroscopy. The crystals were aligned with the b crystal axis parallel to the applied magnetic field. Data were taken with the vibrating sample magnetometer option of the Quantum Design PPMS instrument. The samples studied with neutron spectroscopy were made of a few coaligned crystals with the crystal a axis perpendicular 2469-9950/2017/95(5)/054409(6) 054409-1 ©2017 American Physical Society

Spin pseudogap in the 1 chain material Sr CuO with ...4311/datastream/… · PHYSICAL REVIEW B 95, 054409 (2017) Spin pseudogap in the S = 1 2 chain material Sr 2CuO 3 with impurities

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • PHYSICAL REVIEW B 95, 054409 (2017)

    Spin pseudogap in the S = 12 chain material Sr2CuO3 with impuritiesG. Simutis,1,* S. Gvasaliya,1 N. S. Beesetty,2 T. Yoshida,3 J. Robert,4 S. Petit,4 A. I. Kolesnikov,5 M. B. Stone,6 F. Bourdarot,7

    H. C. Walker,8 D. T. Adroja,8 O. Sobolev,9 C. Hess,10 T. Masuda,3 A. Revcolevschi,2 B. Büchner,10 and A. Zheludev11Neutron Scattering and Magnetism, Laboratory for Solid State Physics, ETH Zürich, Zürich, Switzerland

    2Synthese, Proprietes et Modelisation des Materiaux, Universite Paris-Sud, 91405 Orsay Cedex, France3Institute for Solid State Physics, University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa, Chiba 277-8581, Japan

    4Laboratoire Léon Brillouin, CEACNRS, CEA-Saclay, F-91191 Gif-sur-Yvette, France5Chemical and Engineering Materials Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA

    6Quantum Condensed Matter Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831-6393, USA7Modélisation et dexploration de la matiére, Université Grenoble Alpes et CEA, INAC, 17 rue des Martyrs, 38054 Grenoble, France

    8ISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 OQX, United Kingdom9Forschungsneutronenquelle Heinz Maier-Leibnitz (FRM-II), TU Munchen, D-85747 Garching, Germany

    10Leibniz Institute for Solid State and Materials Research IFW Dresden, P.O. Box 270116, D-01171 Dresden, Germany(Received 11 November 2016; revised manuscript received 23 January 2017; published 7 February 2017)

    The low-energy magnetic excitation spectrum of the Heisenberg antiferromagnetic S = 1/2 chain systemSr2CuO3 with Ni and Ca impurities is studied by neutron spectroscopy. In all cases, a defect-induced spectralpseudogap is observed and shown to scale proportionately to the number of scattering centers in the spin chains.

    DOI: 10.1103/PhysRevB.95.054409

    I. INTRODUCTION

    Defects play a special role in one dimension due to itsintrinsic topology [1,2]. This is especially pronounced inquantum magnets where both the ground states and theexcitations are modified upon the introduction of imperfec-tions in magnetic lattices [3–10]. Recent experiments havedemonstrated that when impurities are added to the proto-typical Heisenberg antiferromagnetic S = 1/2 chain materialsSrCuO2 and Sr2CuO3, their magnetism is suppressed [11–16].Long-range magnetic ordering arising from residual three-dimensional interactions becomes inhomogeneous and occursat lower temperatures [11,16]. Additionally, the spin-latticerelaxation rate 1/T1 measured by NMR experiments dropsat low temperatures, suggesting the depletion of low-energyexcitations [12,15,17]. Defects also affect the heat transportand compromise its ballistic nature [18,19].

    The influence of defects on the magnetic excitation spec-trum can be directly probed using neutron scattering. Wehave previously employed this method to demonstrate theopening of a spin pseudogap in the Heisenberg spin-chainsystem SrCuO2 with chain-breaking Ni2+ impurities [13].We proposed a simple chain-fragmentation model basedon previous theoretical work [20] and derived quantitativepredictions for the excitation spectrum in the presence ofdefects. This model was able to account for the data overa wide temperature range, on an absolute scale, and with noadjustable parameters. Nevertheless, a few questions remainedto be answered: (i) Does the size of the pseudogap indeed scalewith the impurity concentration as predicted? (ii) What is therole of the double-chain structure of SrCuO2? and (iii) Whateffect would other types of impurities have?

    The present paper aims to settle these outstanding issues.Here we study a related spin-chain compound, namely,Sr2CuO3. It has the benefit of having a simpler single-chain

    *[email protected]

    structure, as opposed to the paired chains in SrCuO2 [21–23].Two types of impurities are investigated: S = 1 Ni2+ ions,which replace the S = 1/2 Cu2+ ions in the chain [11,16,20];and Ca2+ ions, which replace Sr2+ and therefore affect theCu2+ chains only indirectly [17,19].

    II. EXPERIMENT

    Single crystals of Sr2CuO3 with Ni and Ca impurities weregrown using floating zone furnaces as described in earlier re-ports [15,19]. They crystalize in an orthorhombic Immm spacegroup, with lattice parameters of a = 3.9089, b = 3.4940, andc = 12.6910 Å for the pure compound [24]. The spin chainsformed by the Cu2+ ions run along the crystallographic bdirection. The exchange interaction between the spins hasbeen estimated as J = 241(11) meV [23]. While the chainsare very well isolated, three-dimensional ordering still takesplace due to residual three-dimensional interactions. In thepure system the spins order at TN = 5.4 K [22] and the orderingtemperature is rapidly decreased with the introduction ofimpurities [11]. All of the measurements presented here wereperformed above the three-dimensional ordering temperatures.The measurement temperatures were much lower than thetemperature corresponding to the intrachain exchange energy.

    In order to study the effects of disorder, diligent attentionhad to be given to controlling the level of impurities. Inaddition to carefully monitoring the ingredients in the growthprocedure and performing energy-dispersive x-ray analysismeasurements, we have also measured the magnetic suscep-tibility in order to estimate the extent of chain fragmentation.The small samples used for susceptibility measurements werecut from the same crystals used for neutron spectroscopy. Thecrystals were aligned with the b crystal axis parallel to theapplied magnetic field. Data were taken with the vibratingsample magnetometer option of the Quantum Design PPMSinstrument.

    The samples studied with neutron spectroscopy were madeof a few coaligned crystals with the crystal a axis perpendicular

    2469-9950/2017/95(5)/054409(6) 054409-1 ©2017 American Physical Society

    https://doi.org/10.1103/PhysRevB.95.054409

  • G. SIMUTIS et al. PHYSICAL REVIEW B 95, 054409 (2017)

    to the scattering plane. The sample with 1% Ni impuritiesconsisted of three coaligned single crystals with a total massof 8.6 g and an FWHM mosaic of 0.4◦ as measured at the 020and 002 peaks. The sample with 2% Ni impurities consistedof two single crystals with a mass of 4.6 g and mosaic of 0.7◦.The sample with 5% Ca impurities was made of two crystalswith a total mass of 3.5 g and a mosaic of 0.3◦.

    Neutron spectroscopy has been performed at a number ofuser facilities. The existence of the spin pseudogap in Ni-substituted samples was established using the 4F2 three-axisspectrometer at LLB, using a combination of energy andmomentum scans with the final neutron wave vector fixedat kf = 1.55 Å−1. Preliminary measurements on the Ca-substituted compound were performed on the PUMA three-

    axis spectrometer at FRM-2, using kf = 2.66 Å−1. Most ofthe measurements were then performed at the SEQUOIA [25]time-of-flight spectrometer at ORNL using incident energiesof 12, 20, and 50 meV with the high-resolution Fermi chopperfrequencies set to 240, 300, and 420 Hz, respectively. Thebackground was measured by repeating the experiment inthe identical configuration with the sample removed andonly the sample can in the beam path. This background wassubtracted from the data before analysis. Data for the 1%Ni-substituted crystal at various temperatures were collectedusing the IN22 3-axis spectrometer [Collaborative ResearchGroup (CEA) at the ILL, Grenoble, France] with the final

    neutron wave vector set to kf = 2.66 Å−1. Additional datafor samples with Ni impurities were collected using theMERLIN [26] spectrometer at ISIS with the high-resolutionFermi chopper with a Gd-slit package rotating at 250 Hzand repetition rate multiplication allowing access to incidentenergies of 10, 20, and 53 meV. In order to achieve quantitativecomparison with the theory, the neutron time-of-flight spectrawere normalized using a standard vanadium sample. Thethree-axis measurements at 8 K were then compared withthe time-of-flight data at 6 K and a correction factor thusobtained was used to normalize the three-axis data obtained atother temperatures. All the data presented here were obtainedat a temperature of T = 6 K. The only exception is themeasurement of temperature dependence shown in Fig. 8,where the temperatures are reported for each data set.

    III. EXPERIMENTAL RESULTS

    A. Low-temperature susceptibility

    To verify the actual concentration of chain-breaking defectsin our samples, we performed measurements of their magneticsusceptibility. The results are plotted versus temperaturein Fig. 1. For the Ni2+-substituted samples, impurities areexpected to take the place of Cu2+ and directly fragment thespin chains. Magnetic susceptibility curves for this scenariohave been derived theoretically [20,27,28] The most completedescription is given by Sirker et al. [28] and allows quantitativeestimation of the actual number of chain breaks. As acharacterization tool, this approach has also been validatedexperimentally [14,16]. By using the same fitting equationas in Ref. [16], we get very good fits (solid lines in Fig. 1)to our experimental data. The deviations seen at very lowtemperatures (note the log scale on the abscissa) can be

    10 1001E-9

    1E-8

    1E-7

    1E-6

    2 % Ni1 % Ni5 % Ca

    χΤ ( m

    3 K/mol

    )

    T (K)

    nominal x (%)

    Sr2CuO3

    0.0 0.5 1.0 1.5 2.0

    0.0

    0.5

    1.0

    1.5

    2.0

    fittedx(%)

    FIG. 1. Magnetic susceptibility multiplied by temperature plottedvs temperature for the studied compounds (symbols). Solid lines arefits as described in the text. Inset: Extracted impurity concentrationof Ni-substituted samples, for fit using all the data (black circles)and measurements only above 10 K (blue squares). The solid blackline corresponds to the number of impurities equal to the nominalconcentration. In the case of Ca substitution, the magnetic responseis much smaller as expected from substitution outside the chain.

    attributed to weak interchain coupling [14,16]. The fittedchain-break concentration is an excellent match with thenominal number of impurities, as shown in the inset in Fig. 1.

    In the case of Ca defects, the Ca2+ ions are expectedto replace Sr2+ and not enter the Cu2+ chains directly. Themagnetic susceptibility measurements in Fig. 1 are consistentwith that picture. Even though a large amount of Ca impuritiesis introduced into the sample, their effect on the magneticresponse is minimal. From the observed magnitude of theeffect in the 5% Ca-substituted sample, we can say withcertainty that the number of broken links in the Cu chainsis less than 0.6%.

    B. Spin excitations

    Time-of-flight neutron spectroscopy allows the measure-ment of the magnetic dynamic structure factor in largeportions of the energy-momentum space. When studyingone-dimensional systems, the data quality can be improvedby integrating over the directions perpendicular to the relevantdimension. In the case of Sr2CuO3, the ratio of inter- andintrachain exchange constants has been estimated [29] to beJ ′/J < 10−3, and therefore such integration can be made useof. A projected slice of such a data set is shown in Fig. 2 forSr2CuO3 with 1% and 2% Ni impurities. The figure displaysthe dynamic structure factor as a function of the energy transferh̄ω and momentum transfer parallel to the direction of thespin chain q‖. The vertical rods of intensity correspond tothe bottom of the two-spinon continuum. Since the exchangeconstant in this material is very large, no dispersion can beobserved and the width of the observed scattering is due to theresolution of the experimental setup. It is immediately visiblethat in both cases there is a suppression of low-energy states.

    054409-2

  • SPIN PSEUDOGAP IN THE S = 12 CHAIN . . . PHYSICAL REVIEW B 95, 054409 (2017)

    FIG. 2. Dynamical structure factor of the two Ni-substitutedsamples as measured by the SEQUOIA time-of-flight neutronspectrometer. The vertical rods at the antiferromagnetic wave vectorare the bottom of the highly dispersive two-spinon continuum. Thesuppression of low-energy states gives rise to a pseudogap whichscales in energy with an increasing number of broken links asdescribed in the text.

    The scale of this pseudogap is larger in the compound withmore impurities.

    In order to quantify this effect, the dynamical structurefactor was integrated with respect to the momentum transferalong the chain axis. Since the peak width is given by theresolution of the spectrometer, Gaussian peaks were used to fitthe momentum cuts, as shown for a selection of cuts in Fig. 3and Fig. 4 for 1% and 2% Ni impurity levels respectively.

    In low-energy transfers, the intensity of the peaks is low andextracting the correct intensity becomes difficult. Thereforethe position and width of the peak were obtained from higherenergy cuts, where the intensity is substantial, and were thenheld fixed for the low-energy cuts. The momentum-integrated(local) dynamical structure factor thus obtained, correctedfor the anisotropic form factor [30] of Cu2+, is shown inFig. 5. Most of the emphasis was drawn to the low-energy

    FIG. 3. Constant-energy cuts of the measured dynamical struc-ture factor in Sr2CuO3 with 1% Ni impurities. Data sets are offsetby 25 mb/meV/sr/f.u. for clarity. Solid lines are Gaussian fits, asdescribed in the text.

    FIG. 4. Constant-energy cuts of the measured dynamical struc-ture factor in Sr2CuO3 with 2% Ni impurities (symbols). Data setsare offset by 25 mb/meV/sr/f.u. for clarity. Solid lines are Gaussianfits, as described in the text.

    part of the spectrum where the pseudogap is observed, andmeasurements with 50-meV incident energy were performedwith shorter counting times. This is especially evident for thehigh-energy part of the spectrum for the 2% sample where,due to the smaller mass of the sample and shorter countingtime, the error bars are substantially larger. Nevertheless, thehigh-energy measurements complement the low-energy dataset and demonstrate that the dynamical structure factor staysconstant in high-energy transfers.

    As shown in Fig. 5, the system with a higher amount ofimpurities has a more pronounced suppression of low-energystates. However, the decrease in intensity is gradual andtherefore the extraction of the gap size is not straightforward.

    FIG. 5. Symbols: momentum-integrated dynamical structure fac-tor of 1% (squares) and 2% (circles) Ni-substituted samples. Thedashed line shows the prediction for an intact chain, whereas solidlines show the scattering for spin chains with 1% (black) and 2%(blue) broken links.

    054409-3

  • G. SIMUTIS et al. PHYSICAL REVIEW B 95, 054409 (2017)

    FIG. 6. Constant-energy cuts of the dynamical structure factorin Sr2CuO3 with 5% Ca impurities for selected energy transfers(symbols). Data sets are offset by 25 mb/meV/sr/f.u. for clarity.Solid lines are Gaussian fits as described in the text.

    The simplest way to assign the value is by quoting the energy atwhich the intensity is halved compared to the expected value.Such an estimate would give values of ≈3 meV for 1% Niimpurities and ≈6 meV for 2% Ni impurities. A much betterway is to consider the whole line shape by introducing a modelbased on the distribution of gap sizes as detailed in Sec. IV.

    Deviations from the dynamical structure factor of an intactchain were observed at all temperatures. It is best visible, how-ever, at low temperatures, where the momentum-integratedspectrum of an ideal chain is constant in the measured energyrange. Therefore most of the discussion is dedicated to themeasurements performed at 6 K, where the observed reductionof low-energy states is directly due to the introduction ofimpurities.

    A series of similar measurements was additionally per-formed at different temperatures. The results are plotted inFig. 8(a), together with previous results for 1% Ni-substitutedSrCuO2 [13]. Here we used scaled coordinates, with h̄ω/2kBTon the abscissa. The y axis is the momentum-integrated imagi-nary part of the susceptibility multiplied by the correspondingmaterial’s exchange constant (J = 241 meV [23] for Sr2CuO3and J = 221 meV [30] for SrCuO2), for direct comparison.

    The same type of measurements was carried out for thesample with Ca impurities. The data, while of poorer qualitydue to the smaller sample, show a clear suppression oflow-energy states (Fig. 6). Its value can be estimated fromthe momentum-integrated dynamical structure factor shownin Fig. 7. Somewhat surprisingly, it is similar in magnitude tothe pseudogaps observed in the samples with 1% and 2% Niimpurities.

    IV. DISCUSSION

    A. Pseudogap and impurity concentration

    The data shown in Fig. 2 are a vivid illustration of thepseudogap dependence on the impurity concentration. Asargued earlier [13], the pseudogap arises from the confinedmotion of spinons in chain fragments bound by defect sites.

    FIG. 7. Measured low-energy excitations in Ca-substitutedSr2CuO3 in a momentum-integrated form (symbols). The solid linerepresents a prediction for 2.1% broken links.

    A finite length L of such fragments ensures a gappingand overall discretization of the excitation spectrum withthe spacing of � = 3.65 × J/L, where J is the intrachainexchange constant [20]. The observed spectrum is defined bythe statistical length distribution of fragments in which thequasiparticles are trapped. The resulting dynamic structurefactor can be factored into that of a defect-free chain and adefect-related envelope function [13]. The envelope functiondepends on the defect concentration x and can be expressedas [13]

    F (ω) =(

    �0 × x2h̄ω

    )2sinh−2

    (�0 × x

    2h̄ω

    ), (1)

    where �0 = 3.65 × J .The envelope function effectively describes the mod-

    ification of the spin-chain spectrum in a bulk materialdue to randomly distributed impurities. Different distri-butions would lead to different shapes of the envelopefunction.

    Since the magnetic dynamical structure factor of a defect-free chain is known in absolute units, and the envelope functionis uniquely expressed through the exchange constant and thedefect concentration, the model has no adjustable parameters.Nevertheless, as shown by solid lines in Fig. 5 it providesan excellent description of our data on the absolute scale. Itis noteworthy that there seems to be a deviation of the datapoints from the prediction at energy transfers around 10 meV.While the origin of this deviation is unclear, there are twopossibilities for such behavior. First, it is conceivable thatour assumption of a random distribution of the defects is toosimplistic. A reduced probability of very short chain segmentswould lead to the observed spectrum with an earlier saturation.Second, this deviation could be due to an additional peak,which may warrant further attention from the theoretical pointof view.

    As pointed out in Ref. [13], the envelope function isindependent of the temperature. For this reason, the model

    054409-4

  • SPIN PSEUDOGAP IN THE S = 12 CHAIN . . . PHYSICAL REVIEW B 95, 054409 (2017)

    FIG. 8. The dynamical structure factor expressed in a universalscaling form (a) before and (b) after correction by the pseudogapfunction. Open (filled) symbols correspond to time-of-flight (three-axis) measurements. The solid line is the theoretical prediction forthe S = 1/2 chain. While the raw data in (a) display the breakdownof scaling behavior, it can be restored by taking into account theT-independent envelope function. The corrected data set in (b) showsscaling at all studied temperatures.

    is expected to work even at elevated temperatures. Inparticular, after a normalization by the envelope function,finite-temperature data are expected to obey the scalingrelations for spin correlations in defect-free chains. In thelatter, the imaginary part of the local susceptibility, χ ′′(ω) =S(ω)/[n(ω) + 1], is a universal function of ω/T [31–33].The scaling function is known [31,32] and is plotted as thesolid line in Fig. 8. The raw data for our Ni-substitutedsamples clearly violate scaling. However, when normalized bythe respective envelope function, they produce a convincingdata collapse across all materials, temperatures, and defectconcentrations.

    Additional confidence in this interpretation of the pseudo-gap is provided by the recent NMR experiments on 1% and 2%Ni-substituted Sr2CuO3, where the characteristic temperatureof the drop in the spin-lattice relaxation rate was found toscale with the impurity concentration [15]. Since the relaxationmechanism in such an experiment is due to the scatteringof thermally excited spinons, the observation in Ref. [15]also implies scaling of the pseudogap with the impurityconcentration. We believe that this scaling of the pseudogapwith the impurity concentration should persist up to higherconcentrations until clustering of impurities becomes relevant.Experimentally, it would become more challenging to observe

    it, since at higher energies, the phonons start to obscure theneutron scattering spectrum and NMR measurements wouldrequire a high-temperature apparatus.

    B. Defects vs scattering centers

    The spectrum measured in the 5% Ca-substituted com-pound at first glance seems to be at odds with this picture.As mentioned above, off-chain Ca2+ defects have virtually noeffect on the magnetic susceptibility. At the same time, neutronscattering clearly shows a pseudogap (Figs. 6 and 7). Usingthe previously proposed model, and treating the number ofchain breaks as a fitting parameter for the neutron data, wededuce as much as 2.1(3)% chain defects, well above the

  • G. SIMUTIS et al. PHYSICAL REVIEW B 95, 054409 (2017)

    [1] S. Eggert, in Statistical Field Theories, edited by A. Cappelliand G. Mussardo (Springer Netherlands, Dordrecht, 2002),p. 279.

    [2] C. L. Kane and M. P. A. Fisher, Phys. Rev. Lett. 68, 1220 (1992).[3] C. Dasgupta and S.-k. Ma, Phys. Rev. B 22, 1305 (1980).[4] I. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, Phys. Rev.

    Lett. 59, 799 (1987).[5] D. S. Fisher, Phys. Rev. B 50, 3799 (1994).[6] K. Damle, O. Motrunich, and D. A. Huse, Phys. Rev. Lett. 84,

    3434 (2000).[7] O. Motrunich, K. Damle, and D. A. Huse, Phys. Rev. B 63,

    134424 (2001).[8] S. Eggert, I. Affleck, and M. D. P. Horton, Phys. Rev. Lett. 89,

    047202 (2002).[9] P. Schmitteckert, T. Schulze, C. Schuster, P. Schwab, and U.

    Eckern, Phys. Rev. Lett. 80, 560 (1998).[10] T. Giamarchi and H. J. Schulz, Europhys. Lett. 3, 1287 (1987).[11] K. M. Kojima, J. Yamanobe, H. Eisaki, S. Uchida, Y. Fudamoto,

    I. M. Gat, M. I. Larkin, A. Savici, Y. J. Uemura, P. P. Kyriakouet al., Phys. Rev. B 70, 094402 (2004).

    [12] F. Hammerath, S. Nishimoto, H.-J. Grafe, A. U. B. Wolter, V.Kataev, P. Ribeiro, C. Hess, S.-L. Drechsler, and B. Büchner,Phys. Rev. Lett. 107, 017203 (2011).

    [13] G. Simutis, S. Gvasaliya, M. Månsson, A. L. Chernyshev, A.Mohan, S. Singh, C. Hess, A. T. Savici, A. I. Kolesnikov, A.Piovano et al., Phys. Rev. Lett. 111, 067204 (2013).

    [14] K. Karmakar and S. Singh, Phys. Rev. B 91, 224401 (2015).[15] Y. Utz, F. Hammerath, S. Nishimoto, C. Hess, N. S. Beesetty,

    R. Saint-Martin, A. Revcolevschi, B. Büchner, and H.-J. Grafe,Phys. Rev. B 92, 060405 (2015).

    [16] G. Simutis, M. Thede, R. Saint-Martin, A. Mohan, C. Baines, Z.Guguchia, R. Khasanov, C. Hess, A. Revcolevschi, B. Büchneret al., Phys. Rev. B 93, 214430 (2016).

    [17] F. Hammerath, E. M. Brüning, S. Sanna, Y. Utz, N. S. Beesetty,R. Saint-Martin, A. Revcolevschi, C. Hess, B. Büchner, andH.-J. Grafe, Phys. Rev. B 89, 184410 (2014).

    [18] N. Hlubek, P. Ribeiro, R. Saint-Martin, S. Nishimoto, A.Revcolevschi, S.-L. Drechsler, G. Behr, J. Trinckauf, J. E.Hamann-Borrero, J. Geck et al., Phys. Rev. B 84, 214419(2011).

    [19] A. Mohan, N. S. Beesetty, N. Hlubek, R. Saint-Martin, A.Revcolevschi, B. Büchner, and C. Hess, Phys. Rev. B 89, 104302(2014).

    [20] S. Eggert and I. Affleck, Phys. Rev. B 46, 10866 (1992).[21] N. Motoyama, H. Eisaki, and S. Uchida, Phys. Rev. Lett. 76,

    3212 (1996).[22] K. M. Kojima, Y. Fudamoto, M. Larkin, G. M. Luke, J. Merrin,

    B. Nachumi, Y. J. Uemura, N. Motoyama, H. Eisaki, S. Uchidaet al., Phys. Rev. Lett. 78, 1787 (1997).

    [23] A. C. Walters, T. G. Perring, J.-S. Caux, A. T. Savici, G. D.Gu, C.-C. Lee, W. Ku, and I. A. Zaliznyak, Nat. Phys. 5, 867(2009).

    [24] T. Ami, M. K. Crawford, R. L. Harlow, Z. R. Wang, D. C.Johnston, Q. Huang, and R. W. Erwin, Phys. Rev. B 51, 5994(1995).

    [25] G. E. Granroth, A. I. Kolesnikov, T. E. Sherline, J. P. Clancy,K. A. Ross, J. P. C. Ruff, B. D. Gaulin, and S. E. Nagler, J.Phys.: Conf. Ser. 251, 012058 (2010).

    [26] R. I. Bewley, T. Guidi, and S. Bennington, Notizi. Neutroni LuceSincrotrone 14, 22 (2009).

    [27] S. Fujimoto and S. Eggert, Phys. Rev. Lett. 92, 037206(2004).

    [28] J. Sirker, S. Fujimoto, N. Laflorencie, S. Eggert, and I. Affleck,J. Stat. Mech.: Theory Exp. (2008) P02015.

    [29] C. Yasuda, S. Todo, K. Hukushima, F. Alet, M. Keller, M. Troyer,and H. Takayama, Phys. Rev. Lett. 94, 217201 (2005).

    [30] I. A. Zaliznyak, H. Woo, T. G. Perring, C. L. Broholm, C. D.Frost, and H. Takagi, Phys. Rev. Lett. 93, 087202 (2004).

    [31] H. J. Schulz, Phys. Rev. B 34, 6372 (1986).[32] D. C. Dender, Ph.D. thesis, Johns Hopkins University (1997).[33] B. Lake, D. A. Tennant, C. D. Frost, and S. E. Nagler, Nat.

    Mater. 4, 329 (2005).

    054409-6

    https://doi.org/10.1103/PhysRevLett.68.1220https://doi.org/10.1103/PhysRevLett.68.1220https://doi.org/10.1103/PhysRevLett.68.1220https://doi.org/10.1103/PhysRevLett.68.1220https://doi.org/10.1103/PhysRevB.22.1305https://doi.org/10.1103/PhysRevB.22.1305https://doi.org/10.1103/PhysRevB.22.1305https://doi.org/10.1103/PhysRevB.22.1305https://doi.org/10.1103/PhysRevLett.59.799https://doi.org/10.1103/PhysRevLett.59.799https://doi.org/10.1103/PhysRevLett.59.799https://doi.org/10.1103/PhysRevLett.59.799https://doi.org/10.1103/PhysRevB.50.3799https://doi.org/10.1103/PhysRevB.50.3799https://doi.org/10.1103/PhysRevB.50.3799https://doi.org/10.1103/PhysRevB.50.3799https://doi.org/10.1103/PhysRevLett.84.3434https://doi.org/10.1103/PhysRevLett.84.3434https://doi.org/10.1103/PhysRevLett.84.3434https://doi.org/10.1103/PhysRevLett.84.3434https://doi.org/10.1103/PhysRevB.63.134424https://doi.org/10.1103/PhysRevB.63.134424https://doi.org/10.1103/PhysRevB.63.134424https://doi.org/10.1103/PhysRevB.63.134424https://doi.org/10.1103/PhysRevLett.89.047202https://doi.org/10.1103/PhysRevLett.89.047202https://doi.org/10.1103/PhysRevLett.89.047202https://doi.org/10.1103/PhysRevLett.89.047202https://doi.org/10.1103/PhysRevLett.80.560https://doi.org/10.1103/PhysRevLett.80.560https://doi.org/10.1103/PhysRevLett.80.560https://doi.org/10.1103/PhysRevLett.80.560https://doi.org/10.1209/0295-5075/3/12/007https://doi.org/10.1209/0295-5075/3/12/007https://doi.org/10.1209/0295-5075/3/12/007https://doi.org/10.1209/0295-5075/3/12/007https://doi.org/10.1103/PhysRevB.70.094402https://doi.org/10.1103/PhysRevB.70.094402https://doi.org/10.1103/PhysRevB.70.094402https://doi.org/10.1103/PhysRevB.70.094402https://doi.org/10.1103/PhysRevLett.107.017203https://doi.org/10.1103/PhysRevLett.107.017203https://doi.org/10.1103/PhysRevLett.107.017203https://doi.org/10.1103/PhysRevLett.107.017203https://doi.org/10.1103/PhysRevLett.111.067204https://doi.org/10.1103/PhysRevLett.111.067204https://doi.org/10.1103/PhysRevLett.111.067204https://doi.org/10.1103/PhysRevLett.111.067204https://doi.org/10.1103/PhysRevB.91.224401https://doi.org/10.1103/PhysRevB.91.224401https://doi.org/10.1103/PhysRevB.91.224401https://doi.org/10.1103/PhysRevB.91.224401https://doi.org/10.1103/PhysRevB.92.060405https://doi.org/10.1103/PhysRevB.92.060405https://doi.org/10.1103/PhysRevB.92.060405https://doi.org/10.1103/PhysRevB.92.060405https://doi.org/10.1103/PhysRevB.93.214430https://doi.org/10.1103/PhysRevB.93.214430https://doi.org/10.1103/PhysRevB.93.214430https://doi.org/10.1103/PhysRevB.93.214430https://doi.org/10.1103/PhysRevB.89.184410https://doi.org/10.1103/PhysRevB.89.184410https://doi.org/10.1103/PhysRevB.89.184410https://doi.org/10.1103/PhysRevB.89.184410https://doi.org/10.1103/PhysRevB.84.214419https://doi.org/10.1103/PhysRevB.84.214419https://doi.org/10.1103/PhysRevB.84.214419https://doi.org/10.1103/PhysRevB.84.214419https://doi.org/10.1103/PhysRevB.89.104302https://doi.org/10.1103/PhysRevB.89.104302https://doi.org/10.1103/PhysRevB.89.104302https://doi.org/10.1103/PhysRevB.89.104302https://doi.org/10.1103/PhysRevB.46.10866https://doi.org/10.1103/PhysRevB.46.10866https://doi.org/10.1103/PhysRevB.46.10866https://doi.org/10.1103/PhysRevB.46.10866https://doi.org/10.1103/PhysRevLett.76.3212https://doi.org/10.1103/PhysRevLett.76.3212https://doi.org/10.1103/PhysRevLett.76.3212https://doi.org/10.1103/PhysRevLett.76.3212https://doi.org/10.1103/PhysRevLett.78.1787https://doi.org/10.1103/PhysRevLett.78.1787https://doi.org/10.1103/PhysRevLett.78.1787https://doi.org/10.1103/PhysRevLett.78.1787https://doi.org/10.1038/nphys1405https://doi.org/10.1038/nphys1405https://doi.org/10.1038/nphys1405https://doi.org/10.1038/nphys1405https://doi.org/10.1103/PhysRevB.51.5994https://doi.org/10.1103/PhysRevB.51.5994https://doi.org/10.1103/PhysRevB.51.5994https://doi.org/10.1103/PhysRevB.51.5994https://doi.org/10.1088/1742-6596/251/1/012058https://doi.org/10.1088/1742-6596/251/1/012058https://doi.org/10.1088/1742-6596/251/1/012058https://doi.org/10.1088/1742-6596/251/1/012058https://doi.org/10.1103/PhysRevLett.92.037206https://doi.org/10.1103/PhysRevLett.92.037206https://doi.org/10.1103/PhysRevLett.92.037206https://doi.org/10.1103/PhysRevLett.92.037206https://doi.org/10.1088/1742-5468/2008/02/P02015https://doi.org/10.1088/1742-5468/2008/02/P02015https://doi.org/10.1088/1742-5468/2008/02/P02015https://doi.org/10.1103/PhysRevLett.94.217201https://doi.org/10.1103/PhysRevLett.94.217201https://doi.org/10.1103/PhysRevLett.94.217201https://doi.org/10.1103/PhysRevLett.94.217201https://doi.org/10.1103/PhysRevLett.93.087202https://doi.org/10.1103/PhysRevLett.93.087202https://doi.org/10.1103/PhysRevLett.93.087202https://doi.org/10.1103/PhysRevLett.93.087202https://doi.org/10.1103/PhysRevB.34.6372https://doi.org/10.1103/PhysRevB.34.6372https://doi.org/10.1103/PhysRevB.34.6372https://doi.org/10.1103/PhysRevB.34.6372https://doi.org/10.1038/nmat1327https://doi.org/10.1038/nmat1327https://doi.org/10.1038/nmat1327https://doi.org/10.1038/nmat1327