THESIS I 17(60 LIBRARY Michigan State University This is to certify that the dissertation entitled Local Atomic Structure and the Metal — Insulator Tran51t10n of Lal-X CaX MnO3 presented by REM) GIOVANNI DIFRANCESCO has been accepted towards fulfillment of the requirements for Ph . D. degree in PHYSICS a jor professor Date 01/18/00 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 1110:) mm.“ LOCAL ATOMIC STRUCTURE AND THE METAL-INSULATOR TRANSITION OF La1_,,.Ca,MnO3 By Remo Giovanni DiFrancesco A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 2000 ABSTRACT LOCAL ATOMIC STRUCTURE AND THE METAL-INSULATOR TRANSITION OF La1_3CaanO3 By Remo Giovanni DiFrancesco La1_zCazMnO3 is one of a class of materials which were recently found to display Colossal Magnetoresistance. This has generated a great deal of interest in attempting to explain this phenomenon. Recent work has suggested that the lattice plays an important role in the electronic behavior of these compounds by way of the Jahn— Teller effect, leading to polarons. We have probed the lattice using atomic pair distribution functions, a local structural technique, to determine if changes in the lattice are present and how they correlate with the various electronic phases. The structure is found to respond strongly to the metal-insulator transition with a distortion of the six oxygen atoms surrounding each Manganese site. We interpret this as the formation of polarons which are first seen in the ferromagnetic metallic (FM) phase and increase in density as the paramagnetic insulating phase is entered. The FM state is seen to contain two regions: a low temperature state in which all carriers are delocalized and the structure is undistorted, and a higher temperature area in which the delocalized charges coexist with polarons. The PM state in the x = 0.5 material is found to contain a similar mixture of phases for a narrow range of temperatures. A charge ordering transition eventually suppresses the delocalized state which ultimately destroys the ferromagnetism. The PDF reveals that at high temperatures the carriers remain localized but their ordering is lost through an order-disorder transition. Multas per gentes et multa per aequora uectus aduenio has miseras, frater, ad inferias, ut te postremo donarem munere mortis et mutam nequiquam alloquerer cinerem. ~Catullus iii ACKNOWLEDGMENTS Of course the first person I need to thank is Simon Billinge. I’m deeply indebted not only for everything he has taught me but also for his patience and tolerance. Thanks also Julius Kovacs, S. D. Mahanti, Wayne Repko, and Stuart Tessmer for their time and effort in serving on my committee. So many people have passed through the Billinge group during my stay that I can’t list them all, but each of them helped me in this work in some way. I hope my colleagues have enjoyed working with me as much as I have enjoyed the time I’ve spent with them. Numerous other members of the department have made my time here something I’ll remember: Kelly Page, Michael Narlock, Freddie Landry, Michael Hamlin, Charles Moreau, Declan Mulhall, and Jerry Pollack to name a few. Finally, thanks to my parents for their support and their trust. iv TABLE OF CONTENTS LIST OF TABLES vii LIST OF FIGURES viii 1 Introduction to Colossal Magnetoresistant Manganites 1 1.1 Introduction ................................ 1 1.1.1 Early work ............................. 2 1.1.2 Colossal magnetoresistance .................... 4 1.1.3 Experimental evidence for a strong electron lattice interaction 8 1.2 Measuring local structure ......................... 8 1.2.1 Average and local structure ................... 8 1.2.2 PDF technique .......................... 9 1.2.3 XAFS ............................... 9 1.3 Polarons and the Jahn-Teller effect ................... 10 1.3.1 Jahn-Teller effect ......................... 10 1.3.2 Polarons .............................. 10 1.4 La1_xCaanOs .............................. 12 1.4.1 Phase diagram .......................... 12 1.4.2 Average structure ......................... 13 1.4.3 Properties ............................. 13 1.5 Summary ................................. 14 2 The Pair Distribution Function Technique 15 2.1 Introduction ................................ 15 2.2 Data collection .............................. 18 2.2.1 Neutrons .............................. 18 2.2.2 X-Rays ............................... 21 V 2.3 Analysis .................................. 2.3.1 Neutrons .............................. 2.3.2 X—Rays ............................... 2.3.3 Obtaining the PDF ........................ 2.4 Modeling .................................. 2.5 PDF peak height and width ....................... Metal-Insulator Transition at Tc and xc 3.1 Introduction ................................ 3.2 Structural Response to Temperature .................. 3.2.1 Atomic disorder from PDF peak-heights ............ 3.3 Doping across the phase transition ................... 3.3.1 Within the metallic region .................... 3.4 Polarons and modeling .......................... 3.4.1 Fitting the Jahn-Teller distortion ................ Metal Insulator Transition at x=0.5 4.1 Introduction ................................ 4.2 Experiments and results ......................... 4.3 Summary ................................. Discussion and Conclusions 5.1 Summary ................................. 5.1.1 Results ............................... 5.2 Future work ................................ APPENDIX Bibliography vi 23 23 24 26 27 28 29 29 30 32 35 39 41 41 44 44 45 49 51 51 51 53 55 56 List of Tables 3.1 First seventeen atomic correlations and amplitudes for La1_zCa,Mn03 based on the crystal structure model. Note the strong Mn-O peak at 1.96A and the strong O-O peak at 2.74A ................. 33 vii 1.1 1.2 1.3 1.4 1.5 2.1 2.2 2.3 3.1 3.2 3.3 LIST OF FIGURES Basic structure of the perovskite cell. In Lal_,Ca,MnO3 the A and B sites are a random mixture of Lanthanum and Calcium. The oxygen sites form an octahedron about each Manganese ............. Resistivity vs. temperature curves for La1_,SrzMnOg at various com- positions. Note the large changes in slope near T6 for the intermediate- doped compounds [1] ............................ Phase diagram of temperature versus average A-site size. Open sym- bols show Tc measured by magnetization and closed symbols are Tc de- termined by resistivity. Tolerance factor is a measure of the deviation away from perfect cubic structure [2] ................... Phase diagram of average A-site radius vs. dOpant concentration. Ef- fects of Sr and Ca substitution are shown by the dashed lines. Values of T6 are shown in the FMM region[3]. ................. Electronic and magnetic phases of La1-zCa,MnOa [4] .......... X—ray parafocusing geometry. The detector and the source are kept equidistant from the sample ........................ Structure function of 1431-303.an03 up to 22A“. .......... Typical neutron PDF of La1_xCa,MnO3 . ............... PDFs of Lao,83Cao,12MnO3 measured at 150K and 188K, difference plotted below. Dashed lines indicate two-sigma level of uncertainty. Note that the difference curve remains largely within the dashed lines, showing that the structure remains largely unchanged. ........ PDFs of Lao,33Cao,12MnOg measured at 160K and 192K, difference plotted below. Note the large amplitude of the difference curve rel- ative to the uncertainty. Between these temperatures the sample has passed through the MI transition and the structure has changed no- ticeably. .................................. Peak heights vs. temperature for three compositions: x = 0.25 (top) , a: = 0.21 (middle), and a: = 0.12 (bottom). The upper two panels show a large drop in height well below Tc , marked with arrow. Bottom panel shows a composition with no MI transition. The peak heights follow the expected Debye behavior. Insets show resistivity vs. temperature. viii 7 12 21 25 26 32 34 3.4 3.5 3.6 3.7 3.8 4.1 4.2 5.1 PDFs of Lao,79Cao,21MnO3 and Lao.38Cao_12MnO3 . (A) Both measured at 220K, both in the insulating phase. The structures are remarkably similar although the compositions are different.(B) Both measured at 10K, the a: = 0.21 sample is now in the FM phase and is sharper relative to the a: = 0.12 sample. ..................... Comparison of PDF difference curves between Lao,79Cao,21Mn03 and Lao,38Cao,12MnO3 at 220K (solid line) and 10K (triangles). At low tem- perature, when the two samples are in different phases, the structures show larger differences than at higher temperatures at which they are both PI. .................................. 36 37 Difference curves: Triangles are again the difference between Lao,79Cao_21MnO3 and Lao.33Cao,12MnO3 PDFs at 10K. Solid line is the difference between Lao,79Cao,21MnO3 at 225K and 140K. The changes in the structure caused by entering the FM phase through a decrease in temperature or an increase in doping are large and correlate well ........... Height of the Mn-O peak at 10K vs. Calcium concentration. Dashed line shows the boundary composition between insulating and metallic regions at this temperature. The peak sharpens smoothly through :rM, and continues sharpening well into the FM phase. This suggests that local distortions exist close to the phase boundary even at low temperatures. ............................... Undistorted crystal model fit to Lao,75Cao,25MnO3 in the FM phase at 20K (left) and in the PI phase at 260K (right). Darker lines are data. Note that the quality of the fit is considerably poorer at the higher temperature ................................. PDFs of Lao,5oCao,5oMn03 measured at 112K. The solid line is the PDF measure while warming and the circles while cooling. Difference is plotted underneath. Dashed lines indicate two sigmas of statistical uncertainty. Note the close agreement between PDFs .......... Top: Heights of PDF peak at 16.2A as a function of temperature. Squares indicate data collected while warming, circles while cooling. Peak height rises dramatically with temperature between T N and T00- This is due to the loss of charge ordering which leads to a higher symme- try state. Bottom: Heights of PDF peak at 2.75A. Peak heights tend to follow Debye curve (solid line) except on cooling between T6 and T N . Those peak heights (filled circles) increase due to charge delocalization which is suppressed by charge-ordering .................. PDFs of Lao.75Cao,25MnOa measured at room temperature using x-rays and neutrons. ............................... ix 38 40 42 46 47 Chapter 1 Introduction to Colossal Magnetoresistant Manganites 1.1 Introduction This thesis deals with the local atomic structure of 1481-303an03 and how this affects its physical properties. First, early work on these materials from the 1950’s is described. Interest in these materials was recently reawakened with the discovery of an extremely large magnetoresistance in thin films as we discuss in Section 1.1.2. The importance of the lattice, and in particular the local structure, to this phenomenon was predicted theoretically and verified experimentally as we describe in Section 1.1.3. In Section 1.2 we introduce different methods for measuring local structure including the atomic pair distribution function (PDF) technique, which is the approach we have used. The origin of the strong electron-lattice coupling is the Jahn-Teller effect which is introduced in Section 1.3.1. This results in lattice polaron formation which is briefly described in Section 1.3.2. Finally, we describe the structure and properties of the La1_zCa3MnO3 system which is the system studied in this work. A number of recent reviews [3, 5] and collections [6, 7, 8] are available. 0’ Q X. ,0 . A&B G {9' 0 0 Mn 0’ 0 Oxygen J. 6’ a} Figure 1.1: Basic structure of the perovskite cell. In La1_zCa,Mn03 the A and B sites are a random mixture of Lanthanum and Calcium. The oxygen sites form an octahedron about each Manganese. 1.1.1 Early work In 1950 Jonker and van Santen [9] synthesized a series of manganese oxides of the general formula A1_$BanO3, where A is a trivalent and B a divalent ion. These compounds form in the perovskite structure shown in Figure 1.1 in which the A, B, and Mn ions form interpenetrating simple cubic sublattices with O at the cubic faces and edges. At finite d0ping values Mn takes on a mixed valence in order to preserve charge balance. A fraction of the ions, 2:, takes on a Mn4+ (d3) state while the remainder are left in the Mn3+ (d4) state. The doping level was found to influence the magnetic and electrical properties of these materials. For values of a: close to 0 and 1 the ground state is antiferromagnetic while for intermediate (0 < a: < 0.5) dOping levels a ferromagnetic ground state was observed with a Curie temperature reaching its maximum at a: z 0.3. Subsequently, van Santen and Jonker [10] found that above Tc the resistivity was semiconductor-like, dp/dT < 0, while below Tc there is a transition to metallic 2 103‘ 102 10‘ 10° 10'1 Resistivity (Qcm) X=0.17 10'2 X=O v / -3 1° iIIIIII'D" xstts . .x:o«i 1 04 . . l 1 l . I n l La 0 100 200 300 400 500 Temperature (K) Figure 1.2: Resistivity vs. temperature curves for La1_,,erMn03 at various com- positions. Note the large changes in slope near T6 for the intermediate-doped com- pounds [1]. behavior, dp/dT > 0, as well as a sharp reduction in resistivity as shown in Figure 1.2. A mechanism called double exchange (DE) was proposed by Zener [11] to explain the coincidental appearance of metallicity and ferromagnetism. The essential ingre- dients of DE are that Mn is in a mixed valence state so that Mn3+ and Mn4+ are both present in the structure, that there is localized spin (in this case 3/ 2 spin) in the localized tgg orbitals on each Mn ion, and that the tgg holes are Hunds-rule coupled to the electron in a delocalized eg orbital on the Mn3+ ions. Charge transport happens through the hopping of an eg electron to an empty eg orbital on a neighboring Mn4+ ion. The probability of hopping will depend on the relative alignment of the spins of neighboring Mn sites. Specifically, the hopping probability varies like cos %, where 0 is the angle between core-spins on neighboring Mn sites [12]. Therefore when the spins align ferromagnetically there is an enhancement in electron hopping. The lattice does not play an important part in DE and the crystal structure re- ceived little attention until recently. Elemans et al. [13] published structural param- eters from neutron diffraction for the series La1-zBa,Mn1_yTiyOg. The endmember LaMn03 was found to be highly distorted, with rotations and elongations of its oxygen octahedra. Although perovskites often show such rotations, the observed octahedral elongations were thought to be caused by the Jahn-Teller effect [14] which we will discuss in Section 1.3.1. With doping, these elongations seemed to vanish based on diffraction studies and the rotations decreased in magnitude. Overall, these materials remained of purely academic interest until the discovery of their magnetoresistance (MR). 1.1.2 Colossal magnetoresistance Interest was reawakened with a series of experiments [15, 16] that studied the MR of these compounds associated with the transition from paramagnetic insulator (P1) to ferromagnetic metal (FM). They found large decreases in resistance when a field sufficient to drive the system into the FM region was applied. A much larger MR was found in thin films of several manganese oxides [17, 18, 19] which led to the term colossal magnetoresistance (CMR) to distinguish it from the previously discovered gi- ant magnetoresistance (GMR). GMR is found in metallic multilayers with alternating layers of a ferromagnetic metal like Fe or C0, and normal metals like Cu with each layer being the thickness of the oscillation period of the RKKY interaction. GMR is created by reversing ferromagnetic domains defined by the layer thicknesses. The field strength needed to saturate the magnetoresistance is usually several tenths of a 4 Tesla, while a similar effect in CMR materials requires a field of over 5T. It was found that the size of the MR effect is inversely related to Tc [20]. This can be understood quite easily. The large MR occurs because the sample is driven through an insulator-metal transition by the application of a strong external field. In the insulating (high-temperature) state the resistivity increases exponentially on cooling. At Tc the sample changes to a metal and the resistivity dr0ps dramatically. If Tc is lower, the dr0p in the resistivity is larger simply because the resistivity of the insulating state is so high. The same argument can be applied to the case where the sample is driven metallic by an external magnetic field. The Tc can be changed by varying the Mn“ density, which changes the hole concentration and bandfilling. But we can also affect Tc by varying the bandwidth. This is done by changing the average size of the A ion in the ABO3 formula which changes the lattice constant. In Lao,7yPryCao_3MnO3 and Lao,7yYyCao,3MnO3 Hwang et al. [2] studied the evolution of Tc with the Ca-concentration, hence hole density, fixed. They found that as the average size of A, < ra >, decreases, Tc also decreases (Figure 1.3). Instead of creating internal pressure by changing < ra >, external hydrostatic pressure can be applied to alter the lattice parameters. A number of such exper- iments [21, 22] have found that Tc increases as external pressure is applied, with ch/dP increasing as a: is decreased, showing increased sensitivity of T6 to 1:. Sim- ilar experiments [2] on Pro_7Cao,3MnO3 and Lao,7Cao,3MnOa directly correlate with results from internal pressure studies. This sensitive dependence of Tc on pressure is one indicator that the structure has a role [23, 24] in the properties of these materials and that DE is not an adequate theory. Internal pressure studies show that a decrease in the Mn-O—Mn bond angle arises 1.15 1.20 1.25 1.30 400 r 1 I A0 ‘7" osMnOa 300 - A PHI 5; 200 - E" warming open-1'." 10° cloud-1' ' d to“ 0037C.“ on. Ca. Sr : 143.; Sr :Ba): 4 I o J l 0.89 0.90 0.91 0.92 0.93 0.94 0.95 Tolerance factor Figure 1.3: Phase diagram of temperature versus average A-site size. Open sym- bols show Tc measured by magnetization and closed symbols are Tc determined by resistivity. Tolerance factor is a measure of the deviation away from perfect cubic structure [2]. from decreasing < ra >. The increase in orbital overlap created by pressure allows electrons to delocalize, and the kinetic energy gained by the carriers would seem to be the driving force behind the relationship between < ra > and Tc . In addition, it was found that DE alone cannot give the magnitude of the change in resistivity at T6 [23]. A theory was put forth by Millis [25] which argues that the important parameter governing Tc is the e-ph coupling constant A. Millis’ theory is discussed further in Section 1.1.3. The idea that there is a universal relationship between MR and Tc can be seen if we plot To for the various phases as a function of a: and < re > as in Figure 1.4. The PM ground state is bound by the size of the La3+ ion above and by CO, AF, and spin glass states below which are strongly coupled to lattice W// é Lanthanide 1.8 6 ._ 5 Contraction ~50 (A) 1.7 ..... .. L... 1.6 . - a 7. . 0.0 0.1 0.2 0.3 0.4 0.5 X Figure 1.4: Phase diagram of average A-site radius vs. dopant concentration. Effects of Sr and Ca substitution are shown by the dashed lines. Values of T6 are shown in the FMM region[3]. distortions as we might expect since A is increasingly important in that region. While the pressure studies attempted to vary only bandwidth or bandfilling, these parameters are too connected to be varied alone. In the < ra > experiments, although a: is constant the effective filling likely changes as the e-ph coupling is changed. In the external pressure studies, varying the La concentration changes < ra >. The e-ph coupling, A, would seem to be the parameter which is actually responsible for the re sults of these experiments. In Millis’ theory [25] an effective A depends on bandwidth, which in turn depends on the spin alignment through the usual DE expression, and this would seem to fit well with the observed phenomenology. 1.1.3 Experimental evidence for a strong electron lattice in- teraction Other experiments have strongly suggested an important interaction between the elec- tron and the lattice. One of these was a study of the isotope effect in Lao,3Cao_2MnOa [26]. After a 95 per cent exchange of 180 for 160, a reduction of To from 210K to 190K was observed. No change was seen in SI‘RUOa, an itinerant ferromagnet, which had 80 per cent exchange. A number of other experiments including XAFS [27 , 28] and diffraction studies [29] have found large thermal factors on oxygen sites in the PI phase which dr0p rapidly near Tc . All of these experiments demonstrate that the transition from P1 to FM is not solely electronic in origin and that e—ph coupling must play an important part. The theory of Millis [25] attempts to explain the electron-lattice interaction in terms of polarons (see Section 1.3.2). Localized carriers in the PI state are believed to distort their immediate environment, and it is this distortion which XAFS and diffraction experiments may be sensing as exaggerated thermal factors. Other, more specialized types of experiments may be better able to detect and characterize these distortions. 1.2 Measuring local structure 1.2.1 Average and local structure A typical diffraction experiment analyzes the position and intensity of scattered parti- cles in the form of Bragg peaks. These peaks form when the change in the wavevector, Q, of the scattered particle is a vector of the reciprocal lattice, K, of the scattering structure. However, this assumes an infinite, perfectly periodic array of atoms. In the case where imperfections exist, we still recover Bragg peaks but their analysis cannot give us the true picture of the structure in question. Since Bragg scattering 8 arisies from perfect periodic arrangements of atoms, it only carries information about the average structure. Average in this sense does not carry its usual meaning, but is often used in place of long—range structure. Any arrangement of atoms which is ordered over long ranges will be discernible through the analysis of Bragg peaks. Short-range structures which are not spatially ordered do not give rise to Bragg peaks but instead produce diffuse scattering. The analysis of diffuse scattering from single crystals to determine local structure has been done for some time but is hampered by the low collection rate and difficulties of sample preparation. 1.2.2 PDF technique As we discuss in Chapter 2, the atomic pair distribution function (PDF) technique uses powder diffraction data by Fourier transforming the entire pattern: Bragg peaks and diffuse scattering. The result is a real-space radial projection of the atomic pair correlations. Because the entire diffraction pattern is accepted, this function contains information on both the local and average structure. This technique has been applied for some time to the study of liquids and glasses but with modern synchrotrons and neutron sources it has now been extended to crystalline materials [30]. 1.2.3 XAFS Another experimental approach that gives local structural information is XAFS (X- ray Absorption Fine Structure). XAF S works by examining changes in the absorption spectrum near the resonant x-ray energy of a particular ion. At the resonance energy an atom of the resonant species absorbs a photon and photoemits an electron. The outgoing electron wave is backseattered by neighboring atoms and interferes with itself. This gives rise to a measurable modulation of the absorption probability at 9 energies above the absorption edge; the so-called extended x-ray absorption fine struc- ture or XAFS. These fluctuations in the absorption cross-section depend sensitively on details of the local atomic structure and it is possible to infer the local struc- ture from them [31, 32]. This has the advantage of chemical-specificity if one desires to know the structure in the region about one particular type of atom. However, the response of the absorption spectra to structure is highly r-dependent, making information beyond the nearest neighbor somewhat problematic. The PDF has the advantage that it can be applied on all length scales. 1.3 Polarons and the J ahn—Teller effect 1.3.1 J ahn-Teller effect One type of local distortion that is present in these manganites is caused by the Jahn-Teller effect. This is an atomic distortion which certain ions create in order to break the degeneracy of their electron levels [33, 34]. In our case Mn“, which has 3 d-electrons in the tgg state and one in an eg state, is a Jahn-Teller ion. In the octahedral coordination of Mn3+ , the two eg levels are degenerate if all six oxygen atoms are equidistant from the manganese. By distorting the octahedra, the two e9 states lose their degeneracy and the electron can occupy the lower of them. In Jahn- Teller ions the energy thus gained is greater than the elastic penalty of stretching bonds. In Mn3+ ions this leads to an elongation of the octahedron with the d,2_,2 orbital occupied. 1.3.2 Polarons The term polaron is used to refer to a large class of quasiparticles in which a localized charge polarizes the background medium where it resides. In Millis’ theory of electron- 10 phonon coupling the type of polarons created specifically are small lattice polarons. These form when carriers become self-trapped [35]. The condensed charge cloud exerts a force on the previously charge-balanced environment: attracting and repelling the variously charged ions in its vicinity. This “polarizes” the lattice, hence the name. As the local ions adjust to the new situation, the carrier in turn responds to the new potential created by the shifting ions. It is forced to spatially condense further until it is finally contained on a single atomic site. We might expect to see polarons on Mn sites in the PI phase as a distortion in the position of neighboring oxygen sites. In the case of La1_,CaanO3 , the existence of polarons and the Jahn—Teller effect depends on the behavior of the carriers. The carriers may be fully localized onto discrete manganese sites, or they may be itinerant, with each site sharing them evenly. The ability of manganese to adopt a mixed valence state means that at finite doping levels, when the carriers are localized a fraction of the Mn ions, 1 — x, can still be in the 3+ state and so will still display a Jahn-Teller distortion. This means the remaining Mn ions carry a 4+ charge which would attract the surrounding oxygen ions, contracting those MnOe octahdera and creating a polaron. The two distinct charge states create two distinct local structures which we hope to detect. On the other hand, if the carriers are delocalized, all the manganese ions will carry the equivalent charge of 3 + 2:. In that case all the octahedra will be identical, with neither Jahn-Teller nor polaronic distortions. Thus the presence of these two effects are not only intimately linked to each other, but also tied to the electronic state. 11 - T. H, mm,“ ' [mic-Moo. Temperature (K) N 8 180- ~ ’s 100 , § 140 F" N Am insulator met-l § Insulator 120. § § \ 1m IL L 1 § 1 r n 0 10 20 30 40 50 60 70 80 90 100 9603 Figure 1.5: Electronic and magnetic phases of La1_zCa;Mn03 [4]. 1.4.1 Phase diagram The phase diagram of La1_,CazMn03 has been studied by a number of groups [4, 36] and is reproduced in Figure 1.5. At a: = 0, the ground state is an antiferromagnetic insulator (AFI) and a ferromagnetic insulator when 0 < :L‘ < 0.2. For 0.2 < :1: < 0.5 the ground state is FM with a Tc which can be driven upwards by the application of a magnetic field, hence the CMR effect. For a: > 0.5 the ground state is again AFI although there is a well-defined line in the T6 -x plane in this region which defines a charge-ordering (CO) transition seen from TEM [37]. For a: z 0.5 there are three distinct states: COI, FM, and PI. We will discuss this special case further in Chapter 4. 12 1.4.2 Average structure The structure of La1_xCa,MnOg was first studied by several groups [38] in the 1950’s. They found for the undoped material orthorhombic, rhombohedral, and monoclinic symmetries could all be found depending on the method of preparation. Later, El- emans et at. [13] found the compound to be orthorhombic, space group Puma. The average structure shows a large, ordered, Jahn-Teller distortion which disappears with the presence of the dopant. 1.4.3 Properties At temperatures above Tc , transport is characterized by an activated resistivity, p(T) or exp(Ap/T) [9] where Ap z 2500 - 1000K for La1-,Ca¢MnO3 , 0.1 < :1: < 0.6 [39]. Thermopower measurements also show semiconductor-like behavior: 3 (T) or AS/ T, where AS z 120K. In simple semiconductor models with a single carrier type Ap = AS, while the experiments show an order of magnitude difference strongly suggesting an additional excitation. A number of papers [40, 41] have demonstrated that these results are consistent with transport via small polarons. Hall efl'ect measurements done above and below Tc on films of La1_zCa1MnOg with a Tc around 260K show a carrier density n z 1 hole/ Mn site in the FM region, or about three times the Ca concentration [42]. They also find that the mobility is independent of field near Tc for strengths up to several Tesla. This suggests that CMR is due to a field-induced increase in the density of delocalized carriers, and not to a change in the hopping rate as DE holds. Optical conductivity measurements of a Lao,825Cao,1-,5Mn03 single crystal [43] found a broad peak centered at about 1.4eV. A similar structure was found in thin film Ndo,7Sro,3MnOa [44]. The authors argue that this peak is the result of a charge- 13 transfer transition from a Mn3+ eg level split by the Jahn— Teller effect to an unoccu- pied adjacent Mn4+ eg level. One group [45] studied the optical reflectivity of RMO3 (R = La, Y; M = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu) and found a charge gap for most of these compounds but not in LaMnOg, which has no Mn“. The presence of the 1eV feature in the doped versions of this compound supports the polaron theory of transport. 1.5 Summary The phenomenon of colossal magnetoresistance is of great technological and scientific interest. A qualitative understanding of the phenomenology was made possible by the insights of Zener and his Double exchange model which coupled magnetism and charge transport. Two things have happened recently to renew interest in these materials and this phenomenon. First, an extremely large magnetoresistance (so- called colossal magnetoresistance) has been observed in thin-film samples exciting interest in these materials in possible magnetic field sensing applications. Second, it has been shown that DE alone does not explain the magnitudes of the resistivity changes that are observed. As a result, the importance of the electron-lattice coupling through the Jahn-Teller effect has been proposed by Millis, Roder, and others. These theories suggest that the strong e-ph coupling leads to polaronic effects. It is clearly important to verify these proposals to understand better the CMR phenomenon. The local structure is a useful probe to do this because when lattice polarons form they distort the lattice in their vicinity. This has a small effect on the average crystal structure, but is evident in the local structure. A number of techniques can be used to study local structure but in this thesis we have used exclusively the atomic pair distribution function technique. This is described in detail in Chapter 2 and the results are described and summarized in Chapters 3-5. 14 Chapter 2 The Pair Distribution Function Technique 2.1 Introduction All information contained in the PDF comes from the scattering of x-rays and neu- trons. Each scattering event involves the interaction of a particle with an atomic potential, causing a change in the particle’s momentum and, in general, its energy. By measuring the final states of scattered particles one can determine the nature of the potentials they interact with if the incident state is well known. In our case [46, 47], to measure atomic structure we are interested in elastic scattering, in which the energy of the scattered particle is unchanged. As soon as we have more than one atom acting as a scattering center we create interference among our scattered particles. When we introduce an infinite periodic array of atoms, we arrive at the well known von Laue condition, yielding Bragg peaks wherever the change in wave vector Q of the scattered particle is a vector of the reciprocal lattice, K. The analysis of Bragg peaks has been a tool for determining crystal structure since 1913, but is limited only to information on the average structure. In the case of a perfect crystal this is no drawback since the average and actual structures are identical on all length scales, but in most materials there are important local deviations away 15 from the average. Because Bragg scattering arises only from long-range periodicity, it cannot tell us anything about such structures. Fortunately, disorder within a crystal gives rise to diffuse scattering. As the name implies, this scattering has a broad, oscillatory appearance with very low intensity relative to Bragg peaks. In principle, diffuse scattering can be analysed analogous to Bragg peaks to determine the exact nature of the disordered structures. In practice this is extremely difficult, especially in powder diffraction experiments due to intensity problems and unwanted contributions from Bragg peaks. We need not bother with such issues if we accept both diffuse and Bragg scattering. The total elastic scattering amplitude for a crystal is given by fk(9,¢) : le—ik-x U(x) eikf-xdax (2.1) where k is the wave vector of the incident plane wave, k' of the scattered wave, and U (x) is the scattering potential from the array of atoms. We can write this potential as the sum of individual atomic potentials located at each point in the array U(x) = 2 mm.) (2.2) where R.- = x — xi. Thus MM) = Z 300w the inelastic scattering dominates the difl'raction pattern, where How is the Debye-Waller temper- ature factor. This is the width of the Gaussian envelope which damps the intensity of Bragg peaks as a function of Q due to the uncorrelated motion of the atoms. The error in Q-binning still should not produce a large effect because features in inelastic scattering such as phonon dispersions tend to be broad and vary slowly with Q. In 23 addition, assuming we understand the sources of the inelastic scattering, we can re- bin a predicted quantity of neutrons into their correct Q-values. This is known as a Placzek correction and is discussed by a number of authors [51, 52]. Corrections for absorption and multiple scattering are calculated for cylindrical samples following the procedure of Price [53] and Blech and Averbach [54]. This takes into account the absorption and multiple scattering from the sample as well as the container, and corrects these effects before the container scattering is subtracted from the total. Finally, the source spectrum and detector efficiency are accounted for. This is done by taking the ratio of structure factors from the sample to structure factors from a reference vanadium rod. Vanadium is an almost perfectly incoherent scatterer of neutrons and so the scattering intensity measured from a vanadium rod reflects the incident spectrum modified by the detector efficiency. Before calculating the PDF, we convert S(Q) into the single-particle structure function by dividing by the total number of scatterers. This should yield a normalized function which goes to 1 as Q approaches infinity. In practice, the density of the sample during the experiment and the volume irradiated by the beam are not precisely known so the effective density is varied as a parameter until the proper normalization condition is met. 2.3.2 X-Rays Extracting a PDF from x-ray data uses the same fundamental mathematics, but some different corrections must be made. In our case, no sample holder or shielding is in the path of the beam so no characterization runs are needed as with neutron data. Because x-rays are scattered off electrons, the finite size of the atom must be taken into account. This is done by dividing the scattered intensity by the average form factor, which contains the Q—dependence of the scattering due to the atom’s physical 24 15 cu _ _ v-t )- a b _ 6" °¢ V a: c)- _ j. l A LILLAJJAALJ L...‘ 1111 0 5 10 15 20 0(3)“) Figure 2.2: Structure function of La1_zCa,MnO3 up to 22A‘1. extent. As neutrons scatter off of the atomic nucleus, we can ignore the neutron form factor. Compton scattering is another challenge posed by x—rays. While Compton-scattered photons should always have a lower energy than elastic scattering, it is difficult to discriminate between them, especially at low angles. Therefore, both Compton and elastic scattering is accepted. The contribution due to Compton is calculated from theory and subtracted from the data. Absorption and multiple scattering are removed similarly as in the case of neutrons. After these corrections are applied, the resulting structure function can be processed to obtain a PDF without regard to the method of data collection. Figure 2.2 shows a typical structure function from La1_xCazMnOg . 25 .1ffifir.1.r...r,rs.. c” -1 i m... ,A i [ [l H at: 5° ”H H H H w ~ L o_ d I iamLirnrlrkalmnnk 5 10 15 20 r(A) Figure 2.3: Typical neutron PDF of La1_xCa,MnO3 . 2.3.3 Obtaining the PDF After S(Q) is known, the PDF is calculated through direct Fourier transform, accord- ing to Equation 2.8. Figure 2.3 shows a PDF calculated from the structure function in Figure 2.2. Care must be used in determining the range of Q to keep when carrying out the integral. Going to higher Q improves the resolution of the PDF, but the count rate tends to fall quickly with rising Q. Thus, we Fourier transform a higher fraction of noise as we increase QM and the benefit of improved resolution is washed out by increasing statistical fluctuations. Fourier transforming the structure function with its finite Q—range will introduce high-frequency termination ripples into the PDF, which will be discussed in the next section. These arise from the convolution of S (Q) with a step function. Applying a 26 damping function to high-Q data so that S (Q) smoothly approaches 1 at high Q is a common practice but its benefits are mainly cosmetic. It does reduce termination ripples but at the cost of broadening PDF peaks due to the loss of high-Q information. The same can be achieved by simply choosing a lower Qm. Since the termination ripples produced by the convolution can be mimicked in the modeling stage, we have not used any damping. 2.4 Modeling Although modeling is an important element in determining local structural infor- mation from PDFs, most of this work will not refer to results obtained from any modeling. Therefore, I will only briefly discuss the features of the procedure. In general, the average structure of a material is known from a Rietveld refinement of powder diffraction data or some other means. These parameters, including atomic positions, thermal factors, and occupancies, can be taken as a starting point to cal- culate a model-PDF. Parameters may be refined using a least-squares fitting routine while preserving the long-range order of the system. The agreement of this model- PDF with data tells us how well the long-range ordered structure does at fitting the local structure. Once the best fit has been achieved using the constraints of the average structure, local distortions and / or occupancy inhomogeneities may be introduced into the local structure. Once again these variables may be refined to look for a local structure which is superior to the average. Clearly the parameter space can be enormous so unless there is some indication of the type of distortion necessary to fit the data it can be very difficult to identify through modeling. 27 2.5 PDF peak height and width A simple yet important way to learn about disorder from a PDF without resorting to modeling is to examine the height and width of PDF peaks. Each peak in the PDF arises from a pair of atoms separated by a particular distance. A perfect experiment on a fully—ordered crystal with static atoms would yield a series of delta functions, located at every r-value which separates an atomic pair. Actual peaks have a finite width because of the intrinsic thermal and static disorder within a solid, as well as convolution effects. However, the integrated area of a given peak must be conserved since that depends on the total number of contributing atoms. Therefore, any change in the height, and therefore width, of a peak means a change in the disorder associated with that particular correlation since we always convolute our structure functions identically. If we have a number of PDFs, measured at closely spaced values of temperature or composition, we can compare heights of various peaks to spot when local disorder increases or decreases and which atoms are affected. While this alone cannot tell us the shape of distortions, it does tell us when external parameters are having an impact on local structure. 28 Chapter 3 Metal-Insulator Transition at Tc and xc 3.1 Introduction In this research we use the local atomic structure as a probe of the electronic state of the CMR materials from the system La1_zCaanO3 . In particular we are inter- ested in determining whether local Jahn-Teller distorted octahedra can be detected, whether polaronic distortions can be detected and characterized, and whether these features have a temperature or composition dependence. As we will see, these features can be seen and have a particularly large response to the metal-insulator transition. In this chapter we describe the nature of the JT and polaronic distortions and how they change on crossing the metal-insulator transition as a function of temperature. We will also describe preliminary results of what happens as the MI transition at ~ .1: = 0.18 is crossed at low-temperature as a function of doping. In the next chap- ter we will look at the temperature dependence of a sample with :r = 0.5 which lies exactly on another low-T MI-transition. Finally, in Chapter 5, we sum up. 29 3.2 Structural Response to Temperature To look for the signs of lattice polarons at Tc we performed neutron diffraction on La1_$CazMnO3 samples with a: = 0.12, 0.21, and 0.25 at temperatures ranging from 10-300K looking for changes in the PDF corresponding with the MI transition. Sam- ples were made by Dr. J. J. Neumeier at Los Alamos National Lab (LAN L) by a solid state reaction of 1432003, CaC03 and M1102 with repeated grindings and firings at temperatures up to 1400°C, with a final slow cool at 1°C per minute. The nominal Mn valences were determined by idometric titration and were 3.25, 3.24, and 3.29 for the a: = 0.12, 0.21, and 0.25, respectively. The DC magnetization was measured at LANL using a commercial SQUID magnetometer and the temperature dependent resistivity was measured using the standard 4-probe dc technique. Neutron data were collected at the Special Environment Powder Diffractometer (SEPD) instrument at the Intense Pulsed Neutron Source (IPNS) and at the Manuel Lujan Jr. Neutron Scattering Center (MLN SC) at Los Alamos National Laboratory. Collection and Analysis procedures are described in Chapter 2. Figure 3.1 shows PDFs obtained from Lao_33Cao_12MnOa at 150K and 180K. This sample has no MI transition, hence we only expect the structure to change slightly due to thermal effects. The difi'erence in PDFs is plotted below the data and shows reproducibility within statistical uncertainty. Compare this with Figure 3.2, which shows PDFs of Lao,79Cao,21MnO3 at 160K and 192K. This sample has a MI transition, at approximately 182K. We see that the PDFs match up substantially less well, particularly in the region of the first two peaks at the low-r end of the PDF. Table 3.1 shows all the atomic pair correlations below 2.75A, and their respec- tive amplitudes based on the crystal structure. Note that the first PDF peak in La1_zCa¢MnOa arises from correlations between Mn and neighboring oxygen atoms. 30 0 5 10 15 20 r(A) Figure 3.1: PDFs of Lao,33Cao,12Mn03 measured at 150K and 188K, difference plotted below. Dashed lines indicate two-sigma level of uncertainty. Note that the difference curve remains largely within the dashed lines, showing that the structure remains largely unchanged. Any changes in the size or shape of the MnOs octahedra will be reflected in changes of this peak. The second peak at r = 2.75A is largely due to neighboring oxygen-oxygen correlations, although there are also contributions from La/Ca-oxygen and La/Ca- La/ Ca pairs. These two peaks, and no others, are directly dependent on the structure of the Mn06 octahedra. The appearance of changes in the PDF at these peaks is suggestive of polaron formation, which would distort the local arrangement of oxygen atoms. In order to confirm this, we fit models containing polaronic distortions to the data and compared the results to fits containing no distortions. Unfortunately, the structural differences proved too small for our fitting program to distinguish between. We therefore resorted to examining peak-heights. 31 car”) 0 -3 «:1 l 0 5 10 15 20 1‘01) Figure 3.2: PDFs of Lao,88Cao,12MnO3 measured at 160K and 192K, diflerence plotted below. Note the large amplitude of the difference curve relative to the uncertainty. Between these temperatures the sample has passed through the MI transition and the structure has changed noticeably. 3.2.1 Atomic disorder from PDF peak-heights In Chapter 2 we mentioned that the sharpness of an individual PDF peak is deter- mined by (among other things) the degree of static and/ or dynamic disorder present in the contributing correlations. In the case of a temperature-dependent study, as T increases peaks should become broader due to increased thermal vibration. Since the integrated intensity under each peak must remain constant, the height of that peak must decrease. The same holds true in the case of static disorder: an increase will cause associated peaks to broaden and lose height. In La1_xCa,Mn03 we are interested in polaronic and Jahn-Teller distortions which should impact short-range Mn-oxygen and oxygen-oxygen correlations. Since individual PDFs show changes in 32 Table 3.1: First seventeen atomic correlations and amplitudes for La1_xCazMnO3 based on the crystal structure model. Note the strong Mn-O peak at 1.96A and the strong O-O peak at 2.74A. Atom 1 Atom 2 r (Angst) amplitude Mn 01 1.9600 0.8000 Mn 02 1.9600 0.4000 Ca O2 2.4200 0.0250 La 02 2.4200 0.0750 Ca 02 2.4300 0.0250 La 02 2.4300 0.0750 Ca 01 2.4700 0.1000 La 01 2.4700 0.3000 Ca. 02 2.6000 0.0250 La 02 2.6000 0.0750 Ca 01 2.6400 0.1000 La 01 2.6400 0.3000 Ca O2 2.6900 0.0250 La 02 2.6900 0.0750 01 O2 2.7300 0.2000 01 01 2.7400 0.8000 01 02 2.7400 0.2000 the corresponding low-r peaks, we plot the height of these peaks as a function of temperature. Figure 3.3 shows the height of the r = 2.75A peak as a function of temperature for three difl'erent samples. Both this and the 1.9A peak show the same qualitative behavior. We concentrate on the r = 2.75A, however, because the statistical and systematic errors on this peak are smaller. The reason for the former is that the peak is much stronger (there are 12 O-O bonds contributing to this peak instead of just 6 M-0 bonds for the 1.9A peak) and for the latter we note that systematic errors tend to die out with increasing-r in the PDF since they tend to originate from long- wavelength artifacts in S (Q) coming from inadequate data corrections. Beginning with the bottom panel, peak-heights for the a: = 0.12 sample drop smoothly with 33 ' I V Y Y j I Y Y I V I 1 F L 140 Peak Height (1") 120 130 140 130 120 Temperature (K) Figure 3.3: Peak heights vs. temperature for three compositions: :0 = 0.25 (top) , :r = 0.21 (middle), and a: = 0.12 (bottom). The upper two panels show a large drop in height well below Tc , marked with arrow. Bottom panel shows a composition with no MI transition. The peak heights follow the expected Debye behavior. Insets show resistivity vs. temperature. increased temperature. The solid line is a prediction of expected thermal broadening, calculated using the Debye model [47, 55]. Again, this sample has no MI transition and we see that the drop in peak-height is well fit by thermal motion. The middle panel shows the data from the a: = 0.21 sample. The same Debye curve is plotted, and the data in this and the bottom panel have been re-scaled so that the low—temperature points lie near the curve. As Tc is approached there is a very 34 large drop in peak-height which begins at least forty degrees below Tc and continues right up to and perhaps a bit beyond the transition temperature. A similar large effect is seen in the a: = 0.25 sample in the top panel. This sudden drop in height is not explained by thermal effects and must be due to a sudden increase in structural disorder of the MnOs octahedra. Such a change, taking place at the temperatures where the sample is leaving the metallic phase, is just what one might expect if the mechanism becomes polaronic at high-T. In Section 3.4 we will discuss our efiorts to characterize the size and shape of these polarons. 3.3 Doping across the phase transition We can only conclude from the previous section that the local structure becomes more disordered above Tc and that it is relatively more ordered below. We have not shown that the metallic phase is free from polaronic or Jahn-Teller effects, merely that such distortions seem to increase upon approach of the insulating state. In order to determine this, and to help establish that the PI state is indeed polaronic, we examine local structural changes as a function of composition. Once again we carried out neutron diffraction over a range of temperatures on samples with doping values of a: = 0.21 and :1: = 0.12. Figure 3.4 shows PDFs for each sample measured at 10K in the lower panel. Clearly, there are significant differences between the two structures. This in itself is not surprising as the two are chemically distinct materials as well as being in different phases: the lower-doped sample has no metallic state and is a ferromagnetic insulator at low temperatures. The upper panel shows PDFs for the same two samples measured at 220K. Here the a: = 0.21 sample has passed through Tc and is fully in the insulating phase. Now the two structures are virtually indistinguishable. The compositional differences therefore 35 200 400 I ' r A D v 1 J e (um 4) -400-200 0 . , . " l l d b . ‘- 200400 . ., A U v 1 G (nm '8) -400 -200 0 , . , . l Figure 3.4: PDFs of Lao.79Cao_21Mn03 and Lao.3gcao.12MnO3 . (A) Both measured at 220K, both in the insulating phase. The structures are remarkably similar although the compositions are different.(B) Both measured at 10K, the a: = 0.21 sample is now in the FM phase and is sharper relative to the a: = 0.12 sample. do not appear to impact the PDFs significantly, while crossing the MI transition line seems to make both samples structurally identical. This can be explained if small polarons, present in the respective insulating phases, disappear when entering the metallic regime regardless of whether they do so because of a decrease in temperature or an increase in doping. We can test whether this is so by examining changes in the PDFs as a function of temperature and composition. Figure 3.5 shows a pair of PDF difierence curves. The solid line is the difference between a: = 0.12 and a: = 0.21 samples at 220K 36 3 t i 1 c. 2‘ z _ ‘- :2; 31$ 2:: f 5 g“ .‘ .2 .2 3e [ r ”*1 = : . I :: l l c e g . 5 r l ‘. : 1 ‘ i v .: h. .. y '. o t e “‘5‘: 4. 9. ..:: ‘ '2‘ 1‘5 '2 . :2 ,‘ i. .. ‘. 3:2 3 2. ‘4 2‘1 ‘. a :z: 7: t .e 1 “' o. S l- .e ‘ .5. .. Ci I . 3 e. ‘ 2‘ a f 5 : d l 1 L 1 l 1 l 0.2 0.4 0.5 0.8 :- (run) Figure 3.5: Comparison of PDF difference curves between Lao,79Cao,21Mn03 and Lao,33Cao,12MnOg at 220K (solid line) and 10K (triangles). At low temperature, when the two samples are in different phases, the structures show larger differences than at higher temperatures at which they are both PI. and the triangles are the differences between the PDFs at 10K from Figure 3.4. We see again that the changes at low temperature are much greater, and there appears to be no significant correlation between the two curves. Thus we can reasonably rule out chemical differences as a major source of structural changes between these two samples. However, if we again plot the difference curve at 10K in Figure 3.6, together with the difference between a: = 0.21 sample at 225K and 140K, we see a remarkable coincidence. Once again, the solid line in Figure 3.6 is the difference caused by a decrease in temperature through the insulator-metal transition of the higher-doped sample while the triangles show the difference when doping is increased through the insulator-metal transition at low temperature. There is no reason for these 37 I» A D V 1.. A A A v» I ’0'. M’.’ Figure 3.6: Difference curves: Triangles are again the difference between Lao,79Cao,21MnO3 and Lao,38Cao,12MnOa PDFs at 10K. Solid line is the difference between LangCaomMnOg at 225K and 140K. The changes in the structure caused by entering the FM phase through a decrease in temperature or an increase in doping are large and correlate well. two changes to have similar effects on a structure, yet we see that both curves are large and strongly correlated over a wide range of r. The only thing these difference curves have in common is they were both produced by crossing from the insulating to the metallic phase, albeit via different paths. These results point out that in this compositional region, the division of the phase diagram between metallic and insulating phases also separates two distinct local struc- tures. The insulating phase is identified by local distortions caused by polarons which are not seen in the metallic phase. 38 3.3.1 Within the metallic region In the previous discussion we considered only compositions which lie well inside their respective phases. This does not address how the system evolves from one state to another: whether it changes suddenly or continuously passes through a series of intermediate structures from disordered to fully ordered. Distinguishing between these two modes of behavior is important because this will tell a great deal about the electronic state of both the metallic and insulating regions. If the structure gradually becomes more ordered as we increase doping at low temperature across the phase boundary into the metallic region, then the two phases are qualitatively similar on a microscopic scale. They both contain a mixture of distorted and undistorted MnOa octahedra with a ratio set by the doping level. The structure becomes more ordered with increased doping because the number of Jahn- Teller distorted sites drops proportionally. In this case the metallic state is not created by any change in the behavior of carriers but merely by changing the relative number of holes and electrons. This would support a percolative model [56, 57] for the phase transition whereby transport is enhanced once a contiguous path of Mn4+ is established. In this case polarons would still exist in at least some of the metallic phase, albeit in smaller numbers. On the other hand, if the structure remains relatively unchanged below the critical doping value and then suddenly becomes ordered upon crossing the phase boundary, this would indicate a qualitative difference in the nature of the carriers in the two phases. The discontinuous increase in structural order could be explained by a de- localization of carriers, creating a uniform distribution of non-Jahn-Teller distorted octahedra. Therefore, determining the structure will tell us a great deal about the mechanism responsible for the change in electromagnetic properties. 39 4.5 I . r l [- I l I I I i *T I II 1 3? I I f :5 . I '33.” l 1 s f . a ’ i I at“ If ' j I l . I l l 3: | 1 'i1 1 l [[444 l LLLLLI 0 01 0.2 03 04 x(Ce) Figure 3.7: Height of the Mn-O peak at 10K vs. Calcium concentration. Dashed line shows the boundary composition between insulating and metallic regions at this temperature. The peak sharpens smoothly through 3M1 and continues sharpening well into the FM phase. This suggests that local distortions exist close to the phase boundary even at low temperatures. Figure 3.7 shows the height of the Mn—O peak taken from PDFs of 14 samples at 10K with compositions ranging from a: = 0 to :1: = 0.4, concentrated about the critical doping value of 0.19. With the exception of the :r = 0 sample, all the data were collected on the same device during one continuous run under identical conditions to maximize reproducibility. Clearly the structure is undergoing a transition over a wide span of composition from most disordered at a: = 0 until saturation at about a: = 0.22. This would strongly support the percolation argument for metallicity as we see no discontinuity at the phase transformation as well as some persistent disorder in the metallic regime. Unfortunately, susceptibility measurements showed high degrees of inhomogeneity in the higher-doped samples which would smear out the MI transition’s position even if it were occuring at a single value. The samples with lower doping levels were closer to an ideal homogeneous state, and the linear behavior of their peak-heights makes it tempting to say we are seeing a linear evolution toward the 40 ordered state but until adequate samples can be re-tested the result is inconclusive. This requires considerable beam-time and has not been carried out to date. 3.4 Polarons and modeling 3.4.1 Fitting the J ahn-Teller distortion Beyond simply measuring changes in peak-height to look for local distortions, we can learn something about the nature of the Mn06 octahedra through modeling. Looking at Figure 3.2 we immediately notice a difference between PDFs of manganites in the metallic and insulating phases. The first peak in the metallic phase, besides being sharper as we discussed previously, has a noticeable secondary peak appearing as a shoulder near 2.2A which is conspicuously absent when the same material is in the insulating phase. This qualitative difference is repeated over the entire doping range 0<:1:<0.5. At first glance, this secondary peak would seem to present a contradiction. If we presume that the metallic state is more ordered, with polaronic and Jahn-Teller distortions appearing as the insulating state is approached, we might hope to see secondary Mn-O peaks in that region arising from the inhomogeneous distribution of Mn—O bond-lengths. Instead, the reverse is true. One could argue that this secondary peak is evidence of a local, full magnitude Jahn-Teller distortion within the metallic region at all temperatures. While we would agree there are residual distortions in the metallic region near the phase boundary, we shall see that this secondary peak is an experimental artifact. In Chapter 2 we discussed the effects of convolution in the Fourier integral due to finite data in Q-space. As we know, convolution with a step function will cause sharp peaks in the PDF to broaden, as well as introducing ripples according to the 41 rm 1'01) Figure 3.8: Undistorted crystal model fit to Lao.75Cao.25MnOg in the FM phase at 20K (left) and in the PI phase at 260K (right). Darker lines are data. Note that the quality of the fit is considerably poorer at the higher temperature. equation [58] (3.1) 0.20") _ 71/000 C(T,)[sinq,,m(r — r’) _ sin qmax(r + T’)]dr’ _ T—W r+fl where G (r) is the theoretically calculated function and Ge(r) is that function mea- sured experimentally. Of course, we can convolute any model we wish to calculate in the same manner. A model based on an undistorted crystal structure, i.e. a single Mn—O bond-length, is plotted in the left panel of Figure 3.8 together with a PDF from La0,75Ca0.25MnO3 measured at 20K. The data had a Q—max of 22A“1 so the model was convoluted with the correspond- ing step function to reproduce the broadening and ripple pattern. Notice that both the 1.9A peak as well as the secondary peak at 2.3A are reproduced by the model. This model has no correlations at 2.3A , the existence of a peak in the model PDF is due to the sharpness of the 1.9A peak in the metallic phase. Since the structure is well-ordered in this region, and the motion of neighboring atoms is highly correlated, 42 this peak should have the smallest natural width in the entire PDF, hence it will show the greatest effects of the convolution. In contrast, the right panel in Figure 3.8 shows a PDF from Lao_75Cao,25Mn03 at 250K together with a fit from the same undistorted model. Notice that the model once again has a peak at 2.3A which fails to fit the data at all. Because the 1.9A peak is so broadened by distortions in the PI phase, clearly visible in its decreased height, the effects of convolution on its appearance are greatly diminished. The model, which assumes no distortions, still preserves the secondary feature. However, the data shows a great deal of intensity in the region between the two peaks which the model completely misses. This feature is in the wrong position and has the wrong shape to be a termination ripple, it can only be intensity which has escaped from the first peak. It is this feature, and not the 2.3A peak, which is the signature of a long Mn-O bond caused by Jahn-Teller distortions. The extremely good fit of the undistorted model to the data in the metallic phase convinces us that at very low temperatures well inside the FM state, the structure is free from local distortions. However, as we have seen the height of the low-r peaks begin to fall well below Tc and this model gradually fails to capture the correct low-r features as temperature increases within the FM region. This indicates that there are distortions present within much of the FM phase arising from a coexistence of delocalized carriers and polarons. In Chapter Four we will see this situation arise again. 43 Chapter 4 Metal Insulator Transition at x=0.5 4.1 Introduction In La1_zCazMnO3 there is a MI transition at low temperature on increasing Ca con- tent through :1: = 0.5 [4, 59]. The M1 boundary is almost vertical on the 1: vs. T phase diagram. The composition LalszCazMnO3 with :1: = 0.5 is clearly a very spe- cial material which has marginal properties and which can hardly decide between having a ferromagnetic metallic groundstate (characteristic of a: < 0.5) or an an- tiferromagnetic, charge-ordered, insulating ground-state (characteristic of :1: > 0.5). In fact, on cooling the sample goes from a paramagnetic insulator at high tempera- ture first to a ferromagnet with improved metallic properties. On further cooling it charge-orders with an incommensurate ordering vector. Finally, at low temperature, the spins order antiferromagnetically. The sample has the charge-ordered insulating ground-state, but only marginally. The interesting possibility exists that on cooling the charges delocalize, the small polarons evident at high-temperature are destroyed, but on further cooling they re- form in an ordered array at low-temperature. An alternative possibility is that the polarons exist at all temperatures and on charge-melting they undergo an order- 44 disorder transition. In this picture, it is not clear what gives rise to the ferromagnetic phase at intermediate temperature. Clearly, the PDF is a useful tool for diagnosing the local state of the carriers in the intermediate phases. The evolution of the phases in the x = 0.5 sample is as follows. The ground state is antiferromagnetic and charge ordered in striped arrays [60] until z 155K where there is a transition to a ferromagnetic metallic (FM) state. This region is considered more metallic only in the sense that the reisitivity curve bends downwards. Initially, charge ordering survives in the FM state but it gradually becomes more incommensurate with temperature [61]. At z 225 K a second transition results in a paramagnetic insulating (PI) phase with activated polaronic transport [4]. The existence of all three distinct phases, specific to this particular composition, allows one to study two separate MI transitions within the same sample. Thus we can determine if the polaronic distortions seen in the low-temperature insulating phase reappear after crossing the intermediate metallic state. 4.2 Experiments and results Measurements were made during both cooling and warming cycles, ranging between 15K and 325K using a closed cycle helium refrigerator. Examples of G (r) are shown in Figure 4.1. Two sets of data are plotted together to demonstrate the level of reproducibility. They were both measured at 112K, well into the charge ordered phase, first upon cooling and then while warming. The difference curve, plotted below the data, is seen to fall largely within the dashed lines, which show uncertainties due to statistical fluctuations at the level of two standard deviations. Thus we conclude that PDFs obtained as a function of temperature are reproducible within the limits of statistical 45 r(fi) Figure 4.1: PDFs of Lao,5oCao,5oMnOg measured at 112K. The solid line is the PDF measure while warming and the circles while cooling. Difference is plotted underneath. Dashed lines indicate two sigmas of statistical uncertainty. Note the close agreement between PDFs. noise. We begin looking for structural changes by examining PDF peak-heights as a function of temperature. As discussed, peak-height is a strong indicator of changes in the degree of structural disorder within a given correlation. If there were a sudden increase in the disorder of the arrangement of atoms which contribute to a given correlation, that peak would broaden and its height would show a correspondingly sudden decrease. This would be directly analogous to a sharp increase in refined thermal parameters. The top panel of Figure 4.2 shows the height of a peak at 16.2A as a function of temperature. The peak’s height clearly drops once the sample enters the charge ordered region. This behavior is reproduced in a number of other peaks located at 46 4 l 35 fi ' I m . 0100400? Peek height. (1") s . Peek height. (1") e Figure 4.2: Top: Heights of PDF peak at 16.2A as a function of temperature. Squares indicate data collected while warming, circles while cooling. Peak height rises dra- matically with temperature between TN and T00- This is due to the loss of charge ordering which leads to a higher symmetry state. Bottom: Heights of PDF peak at 2.75A. Peak heights tend to follow Debye curve (solid line) except on cooling between Tc and TN . Those peak heights (filled circles) increase due to charge delocalization which is suppressed by charge-ordering. higher (>10A) values of r. This indicates some change of the average structure into a higher symmetry state because peak-heights are increasing as the temperature rises. This is more remarkable considering that thermal effects are simultaneously broad- ening these peaks. However, on shorter length scales the structure evolves differently over the same temperature range. The strong positive peak in the PDF at r z 2.75A arises primarily from oxygen- oxygen correlations and so is a good indicator of changes in the shape of the oxygen octahedra which surround the manganese ions. The bottom panel in Figure 4.2 shows 47 the height of this peak: the large changes seen on intermediate length scales are gone. The first PDF peak, at 1.9A , shows the same qualitative behavior. This peak arises solely from Mn-O correlations and so is also sensitive to local distortions of the octa- hedra. As the carriers clearly are localized in the charge ordered state, this suggests that they remain so throughout this entire temperature range, because we can observe a large effect that delocalization has on this peak [47]. Therefore, while charge or- dering gradually disappears in the FM state, localized charges persist but the charge ordering melts away through an order-disorder transition. If we again examine the bottom panel of Figure 4.2 we notice interesting behavior in the FM state. The solid line shows shows the change in peak-height expected due to thermal motion predicted by the Debye law. Most of the points lie on or near this line, and the cooling and warming cycles show good agreement except in the hys- teretic region. This indicates that there is some response of the Mn-Os octahedron to the electronic and magnetic transitions, but certainly a subtler one than that seen on longer length scales. In particular, the cooling cycle points between T ,m (shown by the dashed line) and Ta, are shifted up away from the Debye curve even though the 16.2A peak shows no change at this point and there are no charge ordering su- perlattice peaks. The octahedra are responding to the increase in metallicity. We do not see a similar effect on the warming cycle because there is no temperature range in that case where the sample is ferromagnetic without being charge ordered also. We can compare the sharpening of this peak to that observed in lower doped samples. In this material the sharpening is much less than that seen in the a: = 0.25 sample. We see this in the inset to Figure 4.2(b) which shows the same data from the main panel with the z = 0.25 data superimposed as solid circles, with 0.25A'2 added to make them line up with the other points in the polaronic region. The dashed line drawn through the a: = 0.25 points shows power-law behavior suggesting a second- 48 order transition added to the transition from the Debye curve. The peak-heights from the a: = 0.25 data fit this curve well over the entire range. The same function is plotted in Figure 4.2(b) as a curved dashed line and the cooling cycle points, shown as filled circles, between Tim and Ta, begin following the curve as well. The first two points below T I", lie on this line, while the third point, which has just entered the charge ordered phase, is between the line and the Debye curve. We explain the electronic behavior of the ferromagnetic phase through a coex- istence of localized and delocalized charges. As we cool through T ,m the polaronic charges begin to delocalize and we see the local structure begin to sharpen accordingly. However, the remaining localized carriers begin ordering around 180K which stabi- lizes the the polaronic state. This suppresses further delocalization, and the number of delocalized carriers begins to fall until they are all back to a polaronic state. This explains why the charge ordering is originally incommensurate, since the number of carriers available for ordering is originally less than the % per Mn site required to be commensurate with the lattice. However, charge ordering increases the number of lo- calized carriers and polaron density gradually approaches 50%. Eventually, when the density of delocalized carriers becomes too small, superexchange coupling becomes favored over double exchange and ferromagnetism is lost. This point, where virtually all carriers are relocalized as polarons, as determined from the PDF coincides with the point where charge ordering becomes commensurate. 4.3 Summary These PDF measurements allow us to understand in some detail the nature of the different phases of the [18.1-303an03 a: = 0.5 compound. At high temperatures the carriers are all localized but there is no charge ordering. The FM state contains a mixture of polarons and delocalized charges. It attempts to behave as the lower- 49 doped compounds do in this phase, by delocalizing more and more carriers as the temperature falls until it reaches a ground state where there are no polarons and the structure is undistorted. But long before it can reach this point the appearance of charge ordering stops further delocalization by creating a lower energy state for localized carriers. When all the charges relocalize the charge ordering can be fully commensurate and we arrive in the insulating ground state. The behavior of the FM state at a: = 0.5 agrees with our view of the CMR region. In both cases charges delocalize in the FM phase but not instantaneously. Not until we are deep in this region do all traces of polarons disappear, but the 2: = 0.5 compound simply never gets that far. The low-temperature, z-dependent measurements shown in Figure 3.7 support the picture that much of the FM phase close to the MI transition boundary actually contains a large fraction of residual polarons. 50 Chapter 5 Discussion and Conclusions 5.1 Summary We set out to understand the phase diagram of La1_,,CazMnO3 using local structure as our guide. In particular, we wanted to understand the mechanism behind the interesting CMR behavior these materials display. The motivation for our approach was the suggestion [23, 24] that electron-lattice effects played an important part in the properties of this class of compounds through the Jahn-Teller efiect which leads to polarons in some regions of the phase diagram. Since a local structural probe is capable of detecting the distortions caused by polarons we attempted to test this idea using the atomic pair distribution function method. 5.1.1 Results Our findings confirm that polaronic effects are important in this CMR material. In Chapter Three we showed that large changes in PDF peak heights are closely corre- lated with crossing the Tc /Tm,- boundary. These changes indicate a large increase in disorder of the Mn-Oe octahedra as the PI phase is approached and then entered. This is just the sort of distortion which would be caused by small polarons. In addi- tion, there are large qualitative changes in the PDF upon crossing the MI transition 51 line as a function of temperature or composition. Similar changes in T and a: which do not cause a phase change create a much smaller change in the PDF. Clearly there is a connection between the structure and the electronic state, which we explain with the appearance of polarons. At low temperatures, the FM state contains carriers which have essentially all delocalized. The structure contains no disorder and no Jahn-Teller efl'ect is present. Models based on the average structure, containing no JT distortions, fit data in the FM state at low temperature very well but performed much worse in other parts of the phase diagram. This agreement between local and average structure means there is very little disorder in this region, hence few if any polarons. Also, at low tempearatures the peak heights in Figure 3.3 reached a saturation point. After they had risen steeply around Tc they displayed canonical behavior according to the Debye curve. This indicates that the delocalization process has stopped, presumably because there are no more carriers left as polarons. At intermediate temperatures the FM state is microscopically inhomogeneous or phase-separated. It consists of both localized and delocalized carriers. This is seen through the drop in peak height which begins much below Tc , and also through the growth of the secondary peak near 2.3A before Tc has been achieved. The transition from metal to insulator as a function of composition is qualitatively difl'erent from the temperature-dependent transition. We see the smooth evolution of peak heights through 3M; in Figure 3.7 in contrast to the dramatic effects around Tc . This suggests that am) is defined by the appearance of a percolative path of undistorted Mn ions which would allow metallic transport. Unfortunately, the poor quality of the samples involved in this experiment makes this a somewhat more tentative result at the moment. However, increase in peak height which continues until a: = 0.25 does support the earlier result that much of the FM phase contains 52 polarons as well as delocalized carriers. We can now understand the strange phenomenology of the a: = 0.5 sample. The FM transition is identical to those with a: < 0.5 : charges begin to delocalize but are still significantly polaronic. At Too the remaining polarons order, which lowers the energy of the localized phase. This causes the delocalized fraction to progressively re- enter the polaron state until they are used up. This is the point when superexchange becomes preferred over DE and we enter the AFI state. This result supports our previous conclusions because without the ability of both types of carriers to coexist there is no way to explain ferromagnetism in the a: = 0.5 compound without the carriers changing from completely localized to completely delocalized. 5.2 Future work One of the main questions which we were unable to resolve is the exact shape of the Jahn-Teller distortion. A primary difficulty had to do with the scattering lengths of our constituent atoms. While oxygen is a fairly strong neutron scatterer which made oxygen correlations show up well, manganese has a negative neutron scattering length. Because of this, all Mn—to—other correlations have a negative sign and there is considerable cancellation of signal in the final PDF. This was one reason why we considered turning to x-rays. Obviously, a problem would then be that oxygen corre- lations would be overwhelmed by those of the larger atoms. However, a combination of both techniques could be succesful in solving many of these difficulties. Figure 5.1 shows PDFs measured on Lao,75Cao,25MnOg using neutrons and x- rays. Some of the most prominent features in each PDF are barely present in their counterpart. This is not an experimental artifact but the result of very different scattering properties when x-rays are substituted for neutrons. Since Lanthanum 53 Neutrons wax-rays r I ‘ I 10 ~ 73 - .‘I l] t :1 I ’ 5 if??? ft If if j +6 .‘7 I] ff it . f e; 0. .0 'i I. I .'+ . ‘I :t 9* :1 I + 3 [I , fit f #0 , l’ . f: 0 o 2.4.7 ‘t + ‘. :' r e . I ‘1 l 1 s". .‘ 1 1 . ' 4 i :l K‘ I: -5 _ g *I . ’10 2:5 5:0 7:5 10.0 WWW) Figure 5.1: PDFs of Lao,75Cao,25MnOg measured at room temperature using x—rays and neutrons. correlations dominate the x-ray PDF, these data could be used to pin down the structure of the La/Ca sublattice. Once these positions are known, they could be used when modeling a neutron PDF. This should produce a better model and reduce the number of variables to be refined. We then would stand a better chance at being able to fit the shape of the Jahn-Teller distortion. Another experiment that should be performed is to re-measure PDFs at base tem- perature for compounds with sic-values closely spaced from 0 to 0.33. While difficulties persist in obtaining homogeneous samples in sufl'icent quantities for neutron diffrac- tion, this remains an important piece of work. If we can confirm our results from our first attempt it would not only support the percolative model of conduction, but boost all of our previous results. If the peak heights show a large response at a: M), this would force us to reexamine many of our ideas. 54 APPENDIX Publications: S. J. L. Billinge, R. G. DiFrancesco, M. F. Hundley, J. D. Thompson, and G. H. Kwei, Competition between charge localization and delocalization in Lao,5oCao,5oMnO3, Phys. Rev. Lett., submitted to Phys. Rev. Lett. (2000) V. Petkov, R. G. DiFrancesco, M. Acharya, and H. C. Foley, Local structure of nanoporous carbons, Philos. Mag. B 79, 1519 (1999) Th. Proffen, R. G. DiFrancesco, S. J. L. Billinge, E. L. Brosha, and G. H. Kwei, Measurement of the local J ahn-Teller distortion in LaMnO3mG, Phys. Rev. B 60, 9973 (1999) R. G. DiFrancesco, S. J. L. Billinge, G. H. Kwei, J. J. Neumeier, and J. D. Thompson, Local structure and polaron formation in La1_,Ca¢Mn03 , Physica B 241- 243, 421 (1998) M. S. Kane, J. F. Goellner, H. C. Foley, R. G. DiFrancesco, S. J. L. Billinge, and L. F. Allard, Symmetry breaking in nanostructure development of carbogenic molecular sieves, effects of morphological pattern formation on oxygen and nitrogen transport, Chem. Mater. 8, 2159 (1996) S. J. L. Billinge, R. G. DiFrancesco, G. H. Kwei, J. J. Neumeier , and J. D. Thomp- son, Direct observation of lattice polaron formation in the local structure of La1_zCa1Mn03 , Phys. Rev. Lett. 77, 715 (1996) 55 BIBLIOGRAPHY [1] A. Urushibara, Y. Moritomo, T. Arima, A. Asamitsu, G. Kido, and Y. Tokura, Insulator-metal transition and giant magnetoresistance in La1_,Sr,MnO3, Phys. Rev. B 51, 14103 (1995). [2] H. Y. Hwang, S.-W. Cheong, P. G. Radaelli, M. Marezio, and B. Batlogg, Lattice effects on the magnetoresistance in doped LaMnO3, Phys. Rev. Lett. 75, 914 (1995). [3] A. P. Ramirez, Colossal Magnetoresistance, J. Phys: Condens. Matter 9, 8171 (1997). [4] P. Schiffer, A. P. Ramirez, W. Bao, and S.-W. Cheong, Low temperature magne- toresistance and the magnetic phase diagram of La1_,Ca,Mn03, Phys. Rev. Lett. 75, 3336 (1995). [5] J. M. D. Coey, M. Viret, and S. von Molnér, Mixed-valence manganites, Adv. in Phys. 48, 167 (1999). [6] T. A. Kaplan and S. D. Mahanti, editors, Physics of Manganites, Physics of Manganites, Plenum, New York, 1999. [7] Y. Tokura, editor, Colossal magnetoresistance Oxides, Colossal magnetoresis- tance Oxides, Gordon and Breach, 1999. [8] B. Raveau and C. N. R. Rao, editors, Colossal magnetoresistance and related properties, Colossal magnetoresistance and related properties, World Scientific, Singapore, 1999. [9] G. M. Jonker and J. H. van Santen, Ferromagnetic compounds of manganese with perovskite structures, Physica (Utrecht) 16, 337 (1950). [10] J. H. van Santen and G. H. Jonker, Physica 16, 599 (1950). [11] C. Zener, Phys. Rev. 82, 403 (1951). [12] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675 (1955). [13] J. B. A. A. Elemans, B. Van Laar, K. R. Van Der Veen, and B. O. Loopstra, J. Solid State Chem. 3, 238 (1971). [14] J. B. Goodenough, Theory of the role of covalence in the perovskite-type man- ganites [La,M(II)]MnO3, Phys. Rev. 100, 564 (1955). 56 m-afl'flu H A [15] Z. Jirak, S. Krupiéka, Z. Simsa, M. Dlouha, and S. Vratislav, neutron-diffraction study of Pr1_,Ca,Mn03, J. Magn. Magn. Mater. 53, 153 (1985). [16] R. M. Kusters, J. Singleton, D. A. Keen, R. McGreevy, and W. Hayes, Mgne- toresistance measurements on the magnetic semiconductor NdeMnOa, Physica B 155, 362 (1989). [17] R. von Helmolt, J. Wecker, B. Holzapfel, L. Schultz, and K. Samwer, Giant negative magnetoresistance in perovskitelike Lag/3381,13 MnO, ferromagnetic-films, Phys. Rev. Lett. 71, 2331 (1993). [18] K. Chahara, T. Ohno, M. Kasai, and Y. Kozono, Magnetoresistance in magnetic manganese oxide with intrinsic antiferromagnetic spin structure, Appl. Phys. Lett. 63, 1990 (1993). [19] S. Jin, T. H. Tiefel, M. McCormack, R. A. Fastnach, R. Ramesh, and J. H. Chen, s 204, 413 (1994). [20] K. Khazeni, Y. X. Jia, L. Lu, V. H. Crespi, M. L. Cohen, and A. Zettl, Effect of pressure on the magnetoresistance of single crystal Ndo.58ro,3ono_14Mn03_5, Phys. Rev. Lett. 76, 295 (1996). [21] Y. Moritomo, Y. Tomiyaka, A. Asimitsu, Y. Tokura, and Y. Matsui, Magnetic and electronic-properties in hole-doped manganese oxides with layered structures - La1_,Sr1+,MnO4, Phys. Rev. B 51, 3297 (1995). [22] J. J. Neumeier, M. F. Hundley, J. D. Thompson, and R. H. Heffner, Sub- stantial pressure effects on the electrical-resistivity and ferromagnetic transition- temperature of La1-,Ca,Mn03, Phys. Rev. B 52, R7006 (1995). [23] A. J. Millis, P. B. Littlewood, and B. I. Shraiman, Double Exchange alone does not explain the resistivity of La1-,Sr,Mn03, Phys. Rev. Lett. 74, 5144 (1995). [24] H. R6der, J. Zang, and A. R. Bishop, Lattice efl'ects in the colossal magnetore- sistant manganites, Phys. Rev. Lett. 76, 1356 (1996). [25] A. J. Millis, B. I. Shraiman, and R. Miiller, Dynamic Jahn-Teller effect and colossal magnetoresistance in La1_,Sr,Mn03, Phys. Rev. Lett. 77, 175 (1996). [26] G. Zhao, K. Conder, H. Keller, and K. A. Miiller, Giant oxygen isotope shift in the magnetoresistive perovskite La1_,Ca,MnO3+y, Nature 381, 676 (1996). [27] C. H. Booth, F. Bridges, G. J. Snyder, and T. H. Geballe, Evidence of magnetization-dependent polaron distortion in La1_,A,MnO3, A = Ca, Pb, Phys. Rev. B 54, R15606 (1996). [28] C. H. Booth, F. Bridges, G. H. Kwei, J. M. Lawrence, A. L. Cornelius, and J. J. Neumeier, Lattice effects in La1_,Ca,Mn03 (x = 0 -+ 1): Relationships between distortions, charge distribution, and magnetism, Phys. ReV. B 57, 10440 (1998). [29] P. Dai, J. Zhang, H. A. Mook, S.-H. Liou, P. A. Dowben, and E. W. Plummer, Experimental evidence for the dynamic J ahn-Teller effect in Lao,35Cao,35Mn03, Phys. Rev. B 54, R3694 (1996). 57 [30] T. Egami, Atomic Correlations in Non-periodic Matter, Mater. Trans. 31, 163 (1990). [31] E. A. Stern, D. E. Sayers, and F. W. Lytle, Phys. Rev. B 11, 4836 (1975). [32] T. M. Hayes and J. B. Boyce, in Solid State Physics, edited by H. Ehrenreich, F. Seitz, and D. 'Ihrnbull, volume 37, page 173, Academic, New York, 1982. [33] J. R. Fletcher and K. W. H. Stephens, J. Phys. C: Solid State Phys. 2, 444 (1969). [34] M. D. Sturge, The Jahn-Teller effect in solids, in Solid State Physics, edited by F. Seitz, D. 'Ilurnbull, and H. Ehrenreich, volume 20, page 91, Academic Press, 1967. [35] D. Emin and T. Holstein, Phys. Rev. Lett. 36, 323 (1976). [36] S-W. Cheong and C. H. Chen, Striped charge and orbital ordering in perovskite manganites, in Colossal magnetoresistance and related properties, edited by B. Raveau and C. N. R. Rao, World Scientific, 1999. [37] A. P. Ramirez, P. Schiffer, S-W. Cheong, W. Bao, T. T. M. Palstra, P. L. Gam- mel, D. J. Bishop, and B. Zegarski, Thermodynamic and electron diffraction signatures of charge and spin ordering in La1-,Ca,Mn03, Phys. Rev. Lett. 76, 3188 (1996). [38] E. O. Wollan and W. C. Koehler, Phys. Rev. 100, 545 (1955). [39] T. T. M. Palstra, A. P. Ramirez, S.-W. Cheong, B. R. Zegarski, P. Schiffer, and J. Zaanen, Transport mechanisms in doped LaMnOsi Evidence for polaron formation, Phys. Rev. B 56, 5104 (1997). [40] A. Asamitsu, Y. Moritomo, and Y. Tokura, Thermoelectric effect in La1_,Sr,Mn03, Phys. Rev. B 53, R2952 (1996). [41] M. Jaime, M. B. Salamon, K. Pettit, M. Rubinstein, R. E. 'I‘t‘eece, J. S. Horowitz, and D. B. Chrisey, Magnetothermopower in Lao,37Cao,33MnO3 thin films, Appl. Phys. Lett. 68, 1576 (1996). [42] P. Matl, N. P. Ong, Y. F. Yan, Y. Q. Li, D. Studebaker, T. Baum, and G. Doubin- ina, Hall effect of the colossal magnetoresistance manganite Lal-xCaanO3, Phys. Rev. B 57, 10248 (1998). [43] Y. Okimoto, T. Katsufuji, T. Ishikawa, A. Urushibara, T. Arima, and Y. Tokura, Anomalous variation of optical-spectra with spin polarization in double-exchange ferromagnet La1_,Sr,Mn03, Phys. Rev. Lett. 75, 109 (1995). [44] S. G. Kaplan, M. Quijada, H. Drew, D. B. Tanner, G. C. Xiong, R. Ramesh, C. Kwon, and T. Venkatesan, Optical evidence for the dynamic J ahn-Teller reflect in Nd_7Sl‘o,3Mn03, Phys. Rev. LBtt. 77, 2081 (1996). [45] T. Arima and Y. Tokura, J. Phys. Soc. Jpn. 64, 2488 (1995). 58 [46] R. G. DiFrancesco, S. J. L. Billinge, G. H. Kwei, J. J. Neumeier, and J. D. Thompson, Local structure and polaron formation in La1_,Ca,Mn03, Physica B 241-243, 421 (1998). [47] S. J. L. Billinge, R. G. DiFrancesco, G. H. Kwei, J. J. Neumeier, and J. D. Thompson, Direct Observation of Lattice Polaron Formation in the Local Struc- ture of La1_,Ca,MnO3, Phys. Rev. Lett. 77, 715 (1996). [48] G. L. Squires, Thermal Neutron Scattering, Thermal Neutron Scattering, Cam- bridge University Press, Cambridge, 1978. [49] S. W. Lovesey, Theory of Neutron Scattering from Condensed Matter, Theory of Neutron Scattering from Condensed Matter, Clarendon Press, Oxford, 1984. [50] E. Fermi, Ric. Sci. 7, 13 (1936). [51] P. A. Egelstaff, Methods of Experimental Physics, Vol. 23—B, in Methods of Experimental Physics, Vol. 2.3-B, edited by D. L. Price and K. Skold, Academic Press, San Diego, 1987. [52] R. Misawa, D. L. Price, and K. Suzuki, Short-range structure of alkali disilicate glasses by pulsed neutron scattering, Journal of Non-Crystalline Solids 37, 85 (1980). [53] D. Price, IPNS Handout 14, unpublished . [54] I. A. Blech and B. L. Averbach, Multiple Scattering of Neutrons in Vanadium and Copper, Phys. Rev. 137, A1 113 (1965). [55] S. J. L. Billinge, P. K. Davies, T. Egami, and C. R. A. Catlow, Deviations from planarity of copper-oxygen sheet in Cao,35$ro,1sCuOg, Phys. Rev. B 43, 10340 (1991). [56] D. Louca and T. Egami, Local lattice distortions in La1_,Sr,Mn03 studied by pulsed neutron diffraction, Phys. Rev. B 59, 6193 (1999). [57] M. Uehara, S. Mori, C. H. Chen, and S.-W. Cheong, Percolative phase separation underlies colossal magnetoresistance in mixed-valence manganites, Nature 399, 560 (1999). [58] S. J. L. Billinge, Real-Space Rietveld: Full Profile Structure Refinement of the Atomic Pair Distribution Function, in Local Structure from Diffraction, edited by S. J. L. Billinge and M. F. Thorpe, page 137, New York, 1998, Plenum. [59] M. Roy, J. F. Mitchell, A. P. Ramirez, and P. Schiffer, A study of the magnetic and electrical crossover region of La0.5Cao.5Mn03, J. Phys: Condens. Matter 11, 4843 (1999). [60] S. Mori, C. H. Chen, and S.-W. Cheong, Pairing of charge-ordered stripes in (La,Ca)MnO3, Nature 392, 473 (1998). [61] P. G. Radaelli, G. Iannone, M. Marezio, H. Y. Hwang, S.-W. Cheong, J. D. Jorgensen, and D. N. Argyriou, Structural effects on the magnetic and transport properties of perovskite A1_,A;Mn03 (x=0.25,0.30), Phys. Rev. B 56, 8265 (1997). 59 SSSSSS II[[l]i]]]]]llli]]l[[][]]]l 209