Polar–Nonpolar Phase Transition Accompanied by Negative Thermal

May 29, 2019 - In particular, a large NTE of αL = −30 ppm K–1 coupled with a ... parameter of the oxygens was fixed at an unreasonably large valu...
0 downloads 0 Views 8MB Size
Subscriber access provided by BUFFALO STATE

Article

Polar-nonpolar phase transition accompanied by negative thermal expansion in perovskite-type Bi PbNiO 1-x

x

3

Yuki Sakai, Takumi Nishikubo, Takahiro Ogata, Hayato Ishizaki, Takashi Imai, Masaichiro Mizumaki, Takashi Mizokawa, Akihiko Machida, Tetsu Watanuki, Keisuke Yokoyama, Yoichi Okimoto, Shin-ya Koshihara, Hena Das, and Masaki Azuma Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.9b00929 • Publication Date (Web): 29 May 2019 Downloaded from http://pubs.acs.org on May 29, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Polar-­‐nonpolar  phase  transition  accompanied  by  negative  thermal   expansion  in  perovskite-­‐type  Bi1−xPbxNiO3   Yuki  Sakai,*,†,‡  Takumi  Nishikubo,‡  Takahiro  Ogata,‡  Hayato  Ishizaki,‡  Takashi  Imai,‡  Ma-­‐ saichiro  Mizumaki,§  Takashi  Mizokawa,¶  Akihiko  Machida,#  Tetsu  Watanuki,#  Keisuke  Yokoya-­‐ ma,|  Yoichi  Okimoto,|  Shin-­‐ya  Koshihara,|  Hena  Das‡,\  and  Masaki  Azuma*,†,‡ †

Kanagawa  Institute  of  Industrial  Science  and  Technology,  705-­‐1  Shimoimaizumi,  Ebina,  243-­‐0435,  Japan   Laboratory  for  Materials  and  Structures,  Tokyo  Institute  of  Technology,  4259  Nagatsuta,  Midori,  Yokohama,  226-­‐ 8503,  Japan   § Japan  Synchrotron  Radiation  Research  Institute,  SPring-­‐8,  Sayo-­‐gun,  Hyogo  679-­‐5198,  Japan   ¶ Department  of  Applied  Physics,  School  of  Advanced  Science  and  Engineering,  Waseda  University,  3-­‐4-­‐1  Okubo,   Shinjuku-­‐ku,  Tokyo  169-­‐8555,  Japan   # Synchrotron  Radiation  Research  Center,  National  Institutes  for  Quantum  and  Radiological  Science  and  Technolo-­‐ gy,  Sayo,  Hyogo  679-­‐5148,  Japan   | Department  of  Chemistry,  Tokyo  Institute  of  Technology,  Meguro,  Tokyo  152-­‐8551,  Japan   ‡

\

World  Research  Hub  Initiative,  Institute  of  Innovative  Research  (IIR),  Tokyo  Institute  of  Technology,  4259  Na-­‐ gatsuta,  Midori,  Yokohama,  226-­‐8503,  Japan    

ABSTRACT:   Perovskite-­‐oxide   Bi1−xPbxNiO3   for   0.60   ≤   x   ≤   0.80   was   found   to   show   a   polar-­‐orthorhombic   to   non-­‐polar-­‐ orthorhombic  phase  transition  accompanied  by  negative  thermal  expansion  (NTE).  Bi1−xPbxNiO3  showed  successive  crys-­‐ tal   structure   changes   depending   on   the   amount   of   Pb.   As   the   amount   of   Pb   increased,   the   crystal   structure   changed   from   a   triclinic   one   with   Bi3+/Bi5+   long-­‐range   ordering   to   an   orthorhombic   one   with   Bi3+/Bi5+   short-­‐range   ordering;   then   it   changed  into  a  polar  orthorhombic  structure  without  Bi3+/Bi5+  ordering  and  finally  to  a  polar  LiNbO3-­‐type  one.  The  key  to   the  inversion  symmetry  breaking  in  PbNiO3,  where  both  6s2  lone-­‐pair  and  Jahn-­‐Teller  (JT)  active  cations  are  absent,  is  the   high  valency  state  of  Pb4+.  Our  results  suggest  that  the  polar  orthorhombic  phase  can  be  realized  by  using  high  valence  A-­‐ site  cations  in  addition  to  controlling  the  tolerance  factor.  

21

1.  INTRODUCTION    

xBixVO3.

Nanoscale   manufacture   of   electronic   devices   and   optical   communications   require   precise   positioning,   and   thus,   even  small  amounts  of  thermal  expansion  can  be  a  prob-­‐ lem.   For   this   reason,   materials   showing   negative   thermal   expansion   (NTE)   have   attracted   much   attention   because   composites   made   from   them   are   expected   to   be   able   to   compensate   for   the   thermal   expansion.1-­‐7   Compounds   with   flexible   frameworks   in   their   crystal   structures   such   as   β-­‐LiAlSiO4,1,2   ZrW2O88   and   Cd(CN)29   can   be   catego-­‐ rized   into   the   first   generation   of   NTE   materials.   The   last   decade   has   seen   remarkable   development   in   materials   with  NTE  resulting  from  phase  transitions.  In  particular,  a   large   NTE   of   αL   =   −30   ppm   K−1   coupled   with   a   magnetic   transition   was   discovered   in   anti-­‐perovskite-­‐type   manga-­‐ nese  nitride.10-­‐16  PbTiO3-­‐based  perovskite-­‐type  oxides  were   found   to   show   NTE   originating   from   a   ferroelectric-­‐ paraelectric   transition.17-­‐21   Colossal   volume   shrinkage,   as   large  as  7.8  %,  induced  by  essentially  the  same  tetragonal-­‐ to-­‐cubic   phase   transition   was   recently   found   in   Pb1-­‐

BiNiO3-­‐based   perovskite-­‐type   oxides   show   large   nega-­‐ tive  coefficients  of  thermal  expansion.24-­‐29  The  parent  ma-­‐ terial,   BiNiO3,   has   an   unusual   valence   distribution   of   Bi3+0.5Bi5+0.5Ni2+O3.30-­‐32  Bi3+  and  Bi5+  occupy  distinct  crystal-­‐ lographic   sites   in   a   √2ap ×   √2ap ×   2ap   triclinic   unit   cell   with  space  group  P-­‐1,  where  ap  is  the  lattice  constant  of  a   simple   cubic   perovskite-­‐type  structure.   Since   Bi   6s   and   Ni   3d   levels   are   close   to   each   other   in   energy,   an   intersite   charge   transfer   between   Bi5+   and   Ni2+   takes   place   under   pressure,   resulting   in   a   Bi3+Ni3+O3   high-­‐pressure   (HP)   phase   that   crystallizes   into   a   √2ap ×   √2ap ×   2ap   ortho-­‐ rhombic  GdFeO3-­‐type  structure  with  a  Pbnm  space  group   and   unique   Bi   site.   The   change   in   the   Ni   valence   state   from   2+   to   3+   leads   to   shrinkage   of   the   Ni-­‐O   bond;   the   unit-­‐cell  volume  decreases  by  2.9%.  Lanthanide  (Ln   =  La,   Nd,   Eu,   Dy)   substitution   for   Bi   or   Fe   substitution   for   Ni   destabilizes   the   Bi3+/Bi5+   disproportionated   state,   and   the  

  An   intersite   charge   transfer   transition   was   shown   to   cause   volume   shrinkage   in   A-­‐site   ordered   per-­‐ ovskite-­‐type  oxides  upon  heating.22,23  

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 12

intersite   charge   transfer   occurs   upon   heating.24-­‐27   This   transition  is  of  first  order,  but  since  low  temperature  (LT)   triclinic  and  high  temperature  (HT)  orthorhombic  phases   coexist  in  varying  fraction  over  a  wide  temperature  range,   the  weighted  average  unit-­‐cell  volume  decreases  smooth-­‐ ly.  

at  a  weight  ratio  of  5:1  and  sealed  in  Pt  capsules.  The  sam-­‐ ples  were  treated  at  8  GPa  and  1473  K  for  30  min  in  a  cu-­‐ bic   anvil-­‐type   HP   apparatus.   The   remaining   KCl   was   re-­‐ moved   by   washing   with   distilled   water.   Sintered   pellets   for  the  HAXPES  measurements  were  prepared  by  treating   the  sample  powder  at  6  GPa  and  973  K  for  10  min.    

Bi1−xPbxNiO3   shows   a   triclinic-­‐to-­‐orthorhombic   phase   transition  at  x  =  0.20.33,34  Recently,  we  reported  a  curious   NTE   behavior   in   Bi1−xPbxNiO3   with   x   ≤   0.25.34   Samples   with   x   =   0.05,   0.10   and   0.15   showed   a   triclinic-­‐to-­‐ orthorhombic   phase   transition   accompanied   by   Bi5+   and   Ni2+   charge   transfer   on   heating,   similar   to   what   is   obser-­‐ ved   in   Bi1−xLnxNiO3   and   BiNi1−xFexO3.   On   the   other   hand,   samples  with  x  =  0.20  and  0.25  showed  an  orthorhombic-­‐ to-­‐orthorhombic  phase  transition  with  a  0.6–0.7%  volume   shrinkage.   A   comprehensive   study   involving   a   Rietveld   analysis  of  synchrotron  X-­‐ray  diffraction  (SXRD)  data,  Ni   L-­‐edge   X-­‐ray   absorption   spectroscopy   (XAS),   hard   X-­‐ray   photoemission  spectroscopy  (HAXPES),  and  pair  distribu-­‐ tion   function   (PDF)   analysis   of   the   synchrotron   X-­‐ray   total   scattering   data   revealed   that   the   LT   orthorhombic   phase   had   a   short   range   ordering   of   Bi3+/Bi5+.   Therefore,   the   LT   phases   of   (Bi3+0.5Bi5+0.5)1−xPb4+xNi2+O3,   triclinic   for   0.05   ≤   x   ≤   0.15   and   orthorhombic   for   0.20   ≤   x,   had   the   same  valence  distribution.    

SXRD   patterns   were   collected   using   a   large   Debye-­‐ Scherrer   camera   installed   at   the   BL02B239,40   and   BL19B2   beamlines  of  SPring-­‐8  and  were  analyzed  using  RIETAN-­‐ FP   programs.41   A   wavelength   of   ~0.42   Å   was   used.   The   total   scattering   data   for   PDF   analysis   were   collected   using   the  diffractometer  at  the  BL22XU  beamline  of  SPring-­‐8.42   Powder   samples   were   loaded   into   cylinder   polyimide   ca-­‐ pillaries   that   had   a   diameter   of   1   mm.   Monochromatized   synchrotron   X-­‐rays   of   λ   =   0.177926   Å   were   irradiated   in   the   sample   at   RT.   Scattering   data   were   obtained   up   to   the   maximum  momentum  transfer  Qmax  of  25  Å−1  (2θmax  =  45°).   The   scattering   signal   from   the   polyimide   capillary   was   removed  by  subtracting  the  empty  capillary  data.  Various   other  corrections  were  made,  as  shown  in  the  reference.43   The   PDF   information   was   obtained   by   using   the   PDFgetX2   program.44   Local   structural   analyses   were   per-­‐ formed   by   using   the   PDFgui   program.45   The   Qdamp   and   Qbroad  values  were  calibrated  by  refining  a  CeO2  standard.    

The   perovskite   phase   of   PbNiO3   with   a   Pb4+N2+O3   charge   distribution   was   reported   to   crystalize   into   a   Ge-­‐ FeO3-­‐type   structure.35,36   However,   the   isotropic   displace-­‐ ment  parameter  of  the  oxygens  was  fixed  at  an  unreason-­‐ ably  large  value,    i.e.,  B  =  2.0  Å2  (Uiso  =  0.0253  Å2),  during   the   Rietveld   refinement;   the   refined   value   for   nickel,   B   =   1.50(3)   Å2   (Uiso   =   0.0190(4)   Å2),   was   also   unreasonably   large.   This   suggests   that   the   actual   structure   has   a   lower   symmetry.  In  the  present  work,  we  investigated  the  evolu-­‐ tion   of   crystal   structure,   valence   distribution,   thermal   expansion   and   magnetic   properties   of   Bi1−xPbxNiO3   with   0.30   ≤   x   ≤   1.00.   All   the   synthesized   samples   had   √2ap ×   √2ap ×  2ap  orthorhombic  structures,  but  the  samples  with   x   ≥   0.50,   including   PbNiO3,   had   polar   Pbn21   symmetry.   The  samples  with   x  ≥  0.50  did  not  have  the  short-­‐ranged   ordering   of   disproportionated   Bi3+/Bi5+   that   has   been   ob-­‐ served  in  samples  with  0.25  ≤  x  ≤  0.40.  The  samples  with   0.60  ≤  x  ≤  0.80  showed  NTE,  but  it  originated  from  a  po-­‐ lar-­‐nonpolar   transition,   as   in   PbTiO3,   instead   of   intersite   charge  transfer.    

2.  EXPERIMENTAL  SECTION   Precursors  for  Bi1−xPbxNiO3  (x  =  0.25,  0.30,  0.35,  0.40,  0.50,   0.60,  0.70,  0.80  and  1.00)  were  prepared  by  using  the  pol-­‐ ymerized   complex   method.23,37,38   Stoichiometric   mixtures   of  Bi(NO3)3·∙5H2O,  Pb(NO3)2  and  Ni(NO3)2·∙6H2O  were  dis-­‐ solved   in   aqueous   solutions   of   citric   acid   and   ethylene   glycol.  The  molar  ratios  of  citric  acid  and  ethylene  glycol   to   the   mixtures   were   10:1   and   30:1,   respectively.   The   gels   were  obtained  by  stirring  the  solution  for  12  h  at  473  K  in   air.  The  gels  were  pyrolyzed  for  2  h  at  723  K  in  air  by  using   a  mantle  heater.  The  obtained  powders  were  annealed  at   973  K  for  12  h  in  air  using  a  furnace  after  grinding  with  a   mortar.   The   obtained   precursors   were   mixed   with   KClO4  

Pb-­‐4f   and   Bi-­‐4f   HAXPES   measurements   were   made   at   300  K  with  E  =  7940  eV  at  the  BL09XU  and  BL47XU  un-­‐ dulater   beamlines   of   SPring-­‐8.   A   hemispherical   photoe-­‐ lectron   analyzer   (VG   Scienta,   R4000)   was   used   in   these   measurements.   The   polycrystalline   samples   used   in   the   HAXPES   measurements   were   fractured   in   situ.   The   binding   energies   were   calibrated   using   the   Au   4f7/2   peak   (84.0   eV)   and   the   Fermi   edge   of   gold   reference   samples.   Total  energy  resolution  was  set  to  280  meV  at  E  =  7940  eV.   Second   harmonic   generation   (SHG)   measurements   of   the   powder   sample   were   performed   in   the   transparent   geometry.   The   light   source   was   a   mode-­‐locked   Ti:   sap-­‐ phire   regenerative   amplified   laser   (pulse   width   about   120   fs,   repetition   rate   of   1   kHz,   photon   energy   of   1.58   eV).   The   sample   was   irradiated   with   the   pulse,   and   the   frequency   doubled  light  (3.16  eV),  i.e.,  the  SHG  signal,  was  detected   by   a   photomultiplier   after   passing   through   high-­‐pass   fil-­‐ ters   and   a   grating-­‐type   monochromator   to   cut   out   the   fundamental  1.58  eV  pulses.   The  temperature  dependence  of  the  magnetic  suscepti-­‐ bility  was  measured  with  a  SQUID  magnetometer  (Quan-­‐ tum   Design,   MPMS   5)   in   an   external   magnetic   field   of   1   kOe.   We   used   the   first-­‐principles   density   functional   theory   (DFT)   +   U   method46   with   the   Perdew-­‐Burke-­‐Ernzerhof   (PBE)   form   of   the   generalized   gradient   approximation   (GGA)   of   the   exchange-­‐correlation   functional47   to   study   the   structural   and   magnetic   phase   stability   of   PbNiO3.   The   calculations   were   performed   using   the   plane-­‐wave   pseudopotential   method   and   projector   augmented   wave   (PAW)  potentials  as  implemented  in  the  Vienna  Ab  initio   Simulation   Package   (VASP).48,49   The   rotationally   invariant   scheme   of   Dudarev   et   al.50   was   used   for   the   Hubbard   U   correction   to   GGA.   We   used   a   kinetic   energy   cut-­‐off   of  

ACS Paragon Plus Environment

Page 3 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure   1.   Temperature   dependences   of   unit   cell   volume   calculated   from   the   SXRD   patterns   confirming   the   orthorhombic-­‐to-­‐ orthorhombic   phase   transition   accompanied   by   NTE.   (a)   Temperature   variation   of   the   magnified   SXRD   patterns   around   the   prominent  peaks  of  Bi1−xPbxNiO3  (x  =  0.10,  0.25,  0.50,  0.60  and  0.80)  upon  heating.  The  data  of  x  =  0.10  and  0.25  are  reprinted   34 with  from  the  previous  report.  λ  =  0.42002  Å  for  x  =  0.10,  λ  =  0.41996  Å  for  x  =  0.25,  λ  =  0.41842  Å  for  x  =  0.50,  λ  =  0.42091  Å  for   x  =  0.60  and   λ  =  0.42011  Å  for  x  =  0.80.  LN  represents  the  reflection  of  the  LiNbO3-­‐type  phases.  (b)  Composition-­‐temperature   phase   diagram   of   Bi1−xPbxNiO3   obtained   by   the   variable   temperature   SXRD   studies   at   ambient   pressure   (AP).   (c)   Temperature   dependence  of  the  weighted  average  unit  cell  volume  of  Bi1−xPbxNiO3.  

520   eV   to   expand   the   wave   functions   for   all   calculations   and   optimized   the   k-­‐point   mesh   following   the   crystal   symmetry.   The   Hellman–Feynman   forces   converged   to   0.01   eV   Å−1.   We   performed   calculations   considering   an   effective  U  from  2.0  eV  to  8.0  eV  for  the  Ni  3d  state.  We   calculated  the  electric  polarization  for  the  polar  phases  by   using  the  modern  theory  of  polarization.51-­‐53    

3.  RESULTS  AND  DISCUSSION   Structural   transformation.  Figure  1a  shows  the  temper-­‐ ature  variation  upon  heating  of  the  SXRD  patterns  of  Bi1− xPbxNiO3  with  x  =  0.10,  0.25,  0.50,  0.60  and  0.80.  Tempera-­‐ ture-­‐induced   triclinic-­‐to-­‐orthorhombic   phase   transitions   were   observed   for   x   =   0.10   on   heating   at   around   500   K.   The  x  =  0.25  sample  showed  an  unusual  shift  in  the  peak   to  a  lower  angle  at  550  K,  indicating  an  orthorhombic-­‐to-­‐ orthorhombic   transition.   These   were   NTEs   induced   by   charge   transfer,   as   previously   reported.   The   SXRD   pat-­‐ terns   of   Bi1−xPbxNiO3   with   x   >   0.25   are   indexed   with   a   √2ap ×  √2ap ×  2ap  orthorhombic  unit  cell,  the  same  as  the   x   =   0.25   sample.   However,   the   sample   with   x   ≥   0.50   showed   a   new   temperature-­‐induced   orthorhombic-­‐to-­‐

orthorhombic  phase  transition,  where  the  transition  tem-­‐ perature  increased  as  x  increased.  A  large  shift  in  the  peak   for   the   200   reflection   of   the   x   =   0.6   sample   between   400   and  500  K  and  the  patterns  at  420-­‐480  K  showing  that  LT   and  HT  phases  coexist  indicate  that  this  phase  transition   is   of   first   order.   Therefore,   the   crystal   structure   at   RT   changes  from  a  triclinic  one  to  an  orthorhombic  one,  and   finally  to  another  orthorhombic  one.  Figure  1b  shows  the   composition-­‐temperature   phase   diagram   of   Bi1−xPbxNiO3.   It  should  be  noted  that  this  is  not  a  true  thermodynamic   phase   diagram   because   the   samples   synthesized   at   HP   and  HT  condition  and  recovered  to  ambient  condition  are   metastable   at   AP.   We   name   the   LT   and   HT   phases   of   Bi0.75Pb0.25NiO3  orthorhombic  phases  I  and  II,  and  the  new   orthorhombic   phase   observed   in   Bi0.4Pb0.6NiO3   at   LT   or-­‐ thorhombic   phase   III.   In   addition   to   the   orthorhombic-­‐ to-­‐orthorhombic   phase   transitions,   the   x   =   0.8   sample   showed   a   temperature-­‐induced   irreversible   phase   transi-­‐ tion   from   perovskite-­‐type   to   LiNbO3-­‐type   structure,   as   has   been   observed   in   PbNiO3.36   On   the   other   hand,   the   orthorhombic  I-­‐II  and  II-­‐I  transitions  are  both  reversible.   Figure  1c  shows  the  temperature  dependence  of  the    

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure   2.  Temperature  dependence  of  the  lattice  parameters   of   Bi1−xPbxNiO3.   The   orthorhombic   lattice   parameters   are   normalized   to   a   pseudo-­‐cubic   subcell   with   ap   =   a/√2,   bp   =   b/√2   and   cp   =   c/2.   In   the   orthorhombic   I-­‐II   transition,   shrinkage  of  the  b-­‐  and  c-­‐axes  and  expansion  of  a-­‐axes  occur   upon   heating.   On   the   other   hand,   the   a-­‐   and   b-­‐axes   shrink   and  the  c-­‐axis  expand  upon  heating  in  the  orthorhombic  III-­‐I   transition.  

weighted   average   unit-­‐cell   volumes,   fLTVLT   +   fHTVHT   (where  fLT(HT)  represents  the  phase  fraction  of  the  LT(HT)   phase   and   VLT(HT)   represents   the   unit-­‐cell   volume   of   the   LT(HT)  phase)  for  Bi1−xPbxNiO3  with  0.20  ≤  x  ≤  0.80.  NTE   occurred  in  the  two  composition  ranges  of  0.20  ≤  x  ≤  0.25   and  0.60  ≤  x  ≤  0.80,  while  the  x  =  0.30  and  0.50  samples   showed   almost   zero   thermal   expansion.   The   difference   between  the  orthorhombic  phases  I  and  II  was  in  the  va-­‐ lence  distribution  and  local  structure,  namely,  phase  I  had   a   (Bi3+0.5Bi5+0.5)1−xPb4+xNi2+O3   valence   distribution   with   short-­‐ranged  Bi3+/Bi5+  ordering,  whereas  phase  II  had  a  (Bi,   Pb)3+Ni3+O3   valence   distribution   without   such   a   charge   ordering.  NTE  occurs  owing  to  the  intersite  charge  trans-­‐ fer   between   Bi5+   and   Ni2+.   The   changes   in   the   lattice   pa-­‐ rameters  in  the  III-­‐I  transition  show  a  different  anisotropy   from  those  in  the  I-­‐II  transition.  Figure  2  shows  the  tem-­‐ perature   dependence   of   the   lattice   parameters   of   Bi1−xPbxNiO3   with   0.20   ≤   x   ≤   0.80.   The   orthorhombic   lat-­‐ tice   parameters   are   normalized   to   pseudo-­‐cubic   subcell.   In   the   I-­‐II   transition,   shrinkage   of   the   b-­‐   and   c-­‐axes   and   expansion   of   the   a-­‐axis   occurred   upon   heating.   On   the   other   hand,   in   the   III-­‐I   transition,   the   a-­‐   and   b-­‐axes   shrunk   and   the   c-­‐axis   expanded   upon   heating.   If   this   transition  is  accompanied  by  an  intersite  charge  transfer,   orthorhombic  phase  III  should  have  a  Bi5+1−xPb4+xNi(1+x)+O3   valence  distribution.     Structure   refinement.  To  elucidate  the  valence  distri-­‐ bution   of   orthorhombic   phase   III   and   the   mechanism   of   NTE   in   the   III-­‐I   transition,   a   Rietveld   analysis   was   per-­‐ formed  on  the  SXRD  data.  The  reflection  conditions  of  0kl   (k  =  2n),  h0l  (h  +  l  =  2n),  h00  (h  =  2n),  0k0  (k  =  2n)  and  00l   (l  =  2n)  in  the  orthorhombic  phase  III  afford  two  possible   space   groups,   nonpolar   Pbnm   and   polar   Pbn21.   Figure   3a   shows  the  laser  power  dependence  of  the  SH  intensity  for   Bi0.2Pb0.8NiO3.  (In  this  measurement,  we  used  the  x  =  0.8  

Page 4 of 12

sample   instead   of   PbNiO3   to   investigate   the   polar   sym-­‐ metry  in  the  orthorhombic  phase  III,  because  perovskite-­‐ type   PbNiO3   gradually   changes   into   the   polar   LiNbO3-­‐ type  one  even  under  ambient  conditions,  whereas  the  x  =   0.8  sample  retains  a  perovskite-­‐type  structure  at  RT.)  The   SH   intensity   was   found   to   be   proportional   to   the   square   of  the  incident  laser  power,  clearly  indicating  breaking  of   the  inversion  symmetry  of  the  x  =  0.8  sample.  The  results   of   the   Rietveld   refinement   assuming   the   Pbn21   structure   model  for  Bi1−xPbxNiO3  with  x  =  0.60,  0.80  and  1.00  at  300   K  are  shown  in  Figure  3b  and  the  crystallographic  param-­‐ eters  are  summarized  in  Table  1.  Satisfactory  small  R  fac-­‐ tors   and   reasonable   isotropic   displacement   parameters   (Uiso)   were   obtained.   It   should   be   noted   that   the   refine-­‐ ment   assuming   nonpolar   Pbnm   structure   model   resulted   in  unusually  large  Uiso  values  for  Ni  and  O2  sites  (see  Ta-­‐ ble   1).   Figure   3c   shows   a   schematic   representation   of   the   refined   Pbn21  structure  and  that  of  Pbnm  for  comparison.   Viewing  from  the  c-­‐axis  direction,  the  NiO6  octahedron  in   adjacent   planes   overlap   in   the   Pbnm   structure,   whereas   they  don’t  in  the  Pbn21  structure.  In  other  words,  the  O2   site   of   the   Pbnm   structure   splits   into   O2   and   O3   sites   in   the  Pbn21  structure.  The  unusually  large  Uiso  value  for  the   O2  site  observed  in  the  refinement  with  the  Pbnm  model   should  be  attributed  to  this  crystallographic  site  splitting.   P-­‐E   measurement   was   attempted   on   Bi0.4Pb0.6NiO3,   but   ferroelectric   loop   was   not   observed   because   of   relatively   large  leakage  current.   Valence  distribution.  The  results  of  bond  valence  sum   (BVS)   calculations   at   RT   indicates   the   formal   valence   state   of   (Bi,   Pb)4+Ni2+O3   in   all   the   compositions,   despite   the   unstable   nature   of   the   Bi4+  state   (see   Figure  4a).  HAX-­‐ PES   measurements   were   performed   to   investigate   the   possible   disproportionation   into   Bi3+   and   Bi5+   without   long-­‐range  ordering.  Figure  4b  and  4c  compare  the  HAX-­‐ PES   spectra   for   Bi1−xPbxNiO3   with   data   for   Bi   and   Pb   standard  samples.  The  intense  peaks  at  163.2  and  157.9  eV   in  the  Bi  4f  data  for  BiNiO3  that  are  absent  from  the  data   for   Bi3+Fe3+O3   indicate   the   presence   of   Bi5+   in   addition   to   Bi3+.  The  Bi  4f  spectrum  for  Bi1−xPbxNiO3  is  similar  to  that   of   BiNiO3,   indicating   coexistence   of   Bi3+   and   Bi5+   valence   states.  It  should  be  noted  that  the  6s0  electronic  configu-­‐ ration   (Bi5+   or   Pb4+)   leads   to   lower   binding   energy   than   the   6s2   configuration   (Bi3+   or   Pb2+)   because   of   a   strong   screening   effect.34,54   On   the   other   hand,   the   Pb   4f   spec-­‐ trum   consists   of   a   single   peak,   and   the   peak   position   is   rather  closer  to  that  of  Pb4+Ni2+O3  than  that  of  Pb2+Ti4+O3,   indicating   that   the   valence   state   of   Pb   is   Pb4+.   Figure   4d   shows   the   composition   dependence   of   the   Bi3+   ratio   esti-­‐ mated  from  the  area  ratio  of  the  HAXPES  peaks.  The  Bi3+   ratio  remains  one  half  for  all  compositions,  indicating  the   formal   valence   state   of   (Bi3+0.5Bi5+0.5)1−xPb4+xNi2+O3,   the   same  as  in  the  x  =  0.20  and  0.25  samples.  Focusing  on  the   width   of   Bi3+   and   Bi5+   peaks,   the   Bi3+   ones   become   nar-­‐ rower   as   the   amount   of   Pb   increases,   whereas   Bi5+   ones   become   broader.   This   trend   may   be   thought   of   as   follows.   The   inhomogeneous   potential   energy   of   the   Bi3+   site   stemming  from  the  lone-­‐pair  effect  makes  the  peak  broad,   but   the   effect   becomes   weak   when   the   amount   of   Bi   is  

ACS Paragon Plus Environment

Page 5 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure   3.   SHG   measurement   and   Rietveld   refinement   indicating  polar   symmetry   in   the   orthorhombic   phase   III.   (a)   Laser   power   dependence   of   the   SH   intensity   for   Bi0.2Pb0.8NiO3.   (b)   Observed   (red   points),   calculated   (green   lines),   and   difference   (blue   lines)   patterns  from  the  Rietveld  analysis  of  the  SXRD  data  of  the  Bi1−xPbxNiO3  with  x  =  0.60  and  0.80  and  1.00  at  300  K.  The  green  tick   marks  in  the  x  =  0.60  and  0.80  samples  and  the  top  tick  marks  in  the  x  =  1.00  sample  correspond  to  the  positions  of  Bragg  reflec-­‐ tions  of  the  Pbn21  perovskite-­‐type  phase,  and  the  bottom  tick  marks  in  the  x  =  1.00  sample  correspond  to  that  of  the  R3c  LiNbO3-­‐ type   phase.     (c)   Schematic   representation   of   the   polar   Pbn21   structure   and   the   Pbnm   structure   models   viewed   from   the   c-­‐axis   direction.  

small.  On  the  other  hand,  the  high  valence  Bi5+  site,  which   is   considered   to   share   oxygen   holes   with   Ni   site,55,56   has   inhomogeneous   potential   energy   when   the   oxygen   holes   are  biased  by  the  Ni  off-­‐center  displacement  in  the  polar   symmetry.  The  broadening  of  the  Pb4+  peak  with  increas-­‐ ing  the  amount  of  Pb  can  be  explained  by  the  same  mech-­‐ anism  as  that  of  the  Bi5+  peak.   Local   structure   characterization.   We   examined   the   local  structures  in  the  orthorhombic  phase  III  by  perform-­‐ ing  a  PDF  analysis  of  the  synchrotron  X-­‐ray  total  scatter-­‐ ing  data  in  order  to  investigate  the  possible  short-­‐ranged   Bi3+/Bi5+  charge  ordering  in  0.25  ≤  x  ≤  0.30  samples.  Figure   5   shows   the   fitting   results   of   the   PDF   analysis   for   Bi0.2Pb0.8NiO3   at   low   r   range   (1-­‐20   Å).   We   employed   a   nonpolar   P-­‐1   triclinic   structural   model   with   two   Bi/Pb   sites,  which  is  the  local  structure  of  orthorhombic  phase  I,   in   addition   to   the   Pbnm   and   Pbn21   ones.   The   isotropic   displacement   parameter   of   each   element   was   fixed   at   Ui-­‐ 2 2 so(Bi)   =   Uiso(Pb)   =   Uiso(Ni)   =   0.00633   Å   (B   =   0.5   Å )   and   2 2 Uiso(O)  =  0.01267  Å  (B  =  1.0  Å )  for  all  models.  The  initial   analysis  assuming  the  Pbnm  structure  model  gave  a  poor   reliability  factor,  RWP  =  12.47%.  Employing  the  polar  Pbn21   model  significantly  decreased  RWP  to  7.52%.  On  the  other   hand,   the   nonpolar   P-­‐1   triclinic   model   provided   only   a   minor  improvement  (RWP  =  9.64%).  These  results  indicate   the  presence  of  a  polar  local  structure  and  the  absence  of   Bi3+/Bi5+   short-­‐range   ordering   in   the   orthorhombic   phase   III.   The   absence   of   the   Bi3+/Bi5+   ordering   is   reasonable  

because  the  Bi  occupancy  of  the  Bi/Pb  site  is  only  20%,  as   will  be  discussed  later.  Therefore,  considering  the  presen-­‐ ce  of  polar  symmetry  in  the  orthorhombic  phase  III  struc-­‐ ture   and   the   absence   of   intersite   charge   transfer   upon   heating,   NTE   occurring   at   the   III-­‐I   transition   should   be   due  to  the  polar-­‐nonpolar  phase  transition,  as  in  PbTiO3.   Such  a  temperature-­‐induced  polar-­‐nonpolar  phase  transi-­‐ tion  from  Pbn21  to  Pbnm  has  been  reported  in  perovskite-­‐ type   CdTiO3   at   77   K.57,58   These   results   strongly   suggest   that   PbNiO3   also   has   a   polar   Pbn21   structure.   However,   NTE   was   absent   in   PbNiO3   because   the   perovskite-­‐type   structure  changed  into  the  LiNbO3-­‐type  one  upon  heating   before  a  transition  to  the  Pbnm  structure.     DFT   Calculations.   A   striking   feature   of   our   present   study  is  the  stabilization  of  the  polar  Pbn21  phase  for  one   of   the   end   members,   high   pressure   stabilized   PbNiO3.   Interestingly,   previous   studies   have   identified   the   high-­‐ pressure  phase  as  a  non-­‐polar  Pbnm  structure.36,59    In  or-­‐ der  to  understand  the  stability  of  this  polar  phase,  we  first   determined   the   ground   state   structure   by   calculating   the   energies  of  the  experimentally  reported  LiNbO3-­‐type  R3c,   Pbnm   and   Pbn21     phases,   and   various   other   structures   identified   on   the   basis   of   unstable   phonon   distortions   at   the  zone  center  (q  =  0)    and  at  the  zone  boundary  (q  ≠  0)   symmetry  points  of  the  Brillouin  zone  for  the  undistorted   Pm-­‐3m  structure.  We  carried  out  full  structural  optimiza-­‐ tions   considering   various   collinear   ferromagnetic   (FM)   and   antiferromagnetic   (AFM)   arrangements   between   Ni  

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table   1.   Structural   Parameters   for   Perovskite-­‐type   Bi1−xPbxNiO3  at  300  K   atoms  

site  

g  

x  

y  

2

z  

100×Uiso(Å )  

−0.7627(3)  

0.86(1)  

0.5005(12)   −0.0116(8)   0   0.6275(13)   −0.0675(15)   −0.7618(52)   0.3353(20)   0.3565(20)   −0.5483(17)   0.8008(20)   0.2182(21)   −0.4210(14)   b x  =  0.8,  Space  Group  Pbn21  

0.78(4)   1.84(20)   1.35(30)   1.00(29)  

a

x  =  0.6,  Space  Group  Pbn21   Bi   Pb   Ni   O1   O2   O3   Bi   Pb   Ni   O1   O2   O3  

4a   4a   4a   4a   4a   4a   4a   4a   4a   4a  

0.4   0.6   1   1   1   1   0.2   0.8   1   1   1   1  

0.0111(1)  

0.0476(1)  

0.0085(1)  

0.0445(1)  

−0.7663(3)  

0.56(1)  

0.4999(9)   0.6260(9)   0.3470(14)   0.7968(14)  

−0.0169(5)   −0.0721(12)   0.3719(14)   0.2216(17)  

0   −0.7514(23)   −0.5447(12)   −0.4190(10)  

0.58(3)   1.52(21)   1.04(23)   0.76(21)  

c

Pb   Ni   O1   O2   O3  

4a   4a   4a   4a   4a  

1   1   1   1   1  

x  =  1.0,  Space  Group  Pbn21   0.0079(1)   0.0417(1)   −0.7691(2)   0.5014(6)   −0.0203(4)   0   0.6259(10)   −0.0793(10)   −0.7436(12)   0.3395(14)   0.3683(14)   −0.5439(11)   0.7951(14)   0.2232(16)   −0.4145(10)  

0.51(1)   0.65(3)   0.26(16)   0.84(19)   1.06(19)  

d

x  =  0.6,  Space  Group  Pbnm   Bi   Pb   Ni   O1   O2  

4c   4a   4c   8d  

0.4   0.6   1   1   1  

Bi   Pb   Ni   O1   O2  

4a   4c   8d  

0.2   0.8   1   1   1  

Pb   Ni   O1   O2  

4c   4a   4c   8d  

1   1   1   1  

4c  

0.0112(1)  

0.4524(1)  

0.25    

0.84(1)  

0   0   0   0.6280(13)   0.5704(15)   0.25   0.3164(13)   0.1789(13)   0.0654(8)   e x  =  0.8,  Space  Group  Pbnm  

1.33(3)   1.62(20)   4.90(24)  

0.0087(1)  

0.4554(1)  

0.25    

0.54(1)  

0   0.6240(11)   0.3212(12)  

0   0.5780(13)   0.1673(13)  

0   0.25   0.0649(8)  

1.46(3)   1.38(20)   6.56(27)  

f

x  =  1.0,  Space  Group  Pbnm   0.0080(1)   0.4583(1)   0.25     0   0   0   0.6257(12)   0.5881(14)   0.25   0.3213(14)   0.1640(15)   0.0616(9)  

0.47(1)   1.93(3)   1.35(18)   6.92(28)  

  a

  Z   =   4;   a   =   5.34660(1)   Å,  b   =   5.52299(2)   Å,   c   =   7.72237(2)   Å;   b Rwp  =  6.97%,  Rp  =  4.93%,  RB  =  1.57%,  RF  =  0.97%,  S  =  2.24.    Z   =  4;  a  =  5.35698(1)  Å,  b  =  5.49382(1)  Å,  c  =  7.71333(1)  Å;  Rwp  =   c 3.57%,  Rp  =  2.65%,  RB  =  1.28%,  RF  =  0.77%,  S  =  1.74.    Z  =  4;  a  =   5.35800(1)  Å,  b  =  5.46365(1)  Å,  c  =  7.70772(1)  Å;  Rwp  =  6.72%,   Rp   =   4.70%,   RB   =   1.24%   for   perovskite-­‐type   and   1.15%   for   LiNbO3-­‐type,   RF   =   0.59%   for   perovskite-­‐type   and   0.43%   for   d LiNbO3-­‐type,  S  =  2.87.    Z  =  4;  a  =  5.34639(2)  Å,  b  =  5.52280(2)   Å,   c   =   7.72208(2)   Å;   Rwp   =   7.37%,   Rp   =   5.25%,   RB   =   2.02%,   RF   =   e 1.24%,  S  =  2.37.    Z  =  4;  a  =  5.35651(1)  Å,  b  =  5.49332(1)  Å,  c  =   7.771263(2)  Å;  Rwp  =  4.45%,  Rp  =  3.32%,  RB  =  2.66%,  RF  =  1.76%,   f S   =   2.17.     Z   =   4;   a   =   5.35805(1)   Å,   b   =   5.46372(1)   Å,   c   =   7.70778(1)  Å;  Rwp  =  8.85%,  Rp  =  6.43%,  RB  =  2.24%  for  perov-­‐ skite-­‐type   and   1.89%   for   LiNbO3-­‐type,   RF   =   1.06%   for   perov-­‐ skite-­‐type  and  0.72%  for  LiNbO3-­‐type,  S  =  3.77.    

spins.   The   relaxed   energetics   for   the   PbNiO3   system   for   the  lowest  energy  spin  structure  is  depicted  in  Figure  6a.   We  found  that  the  ground  state  structure  crystallizes  with   LiNbO3-­‐type   polar   R3c   symmetry   and   the   Ni   spins   are   ordered  in  an  AFM  G-­‐type  pattern,  in  agreement  with  the   previous  report.59  The  R3c  phase  was  found  to  be  an  insu-­‐

Page 6 of 12

lator   with   an   energy   gap   between   O   2p   states   and   Pb   6s   states   hybridized   with   O   2p   orbitals   (see   Figure   S1).   The   Ni  3d  states  form  conduction  bands  in  the  energy  range  ~   −6.0   to   −2.0   eV   in   one   spin   channel,   while   in   the   other   spin  channel,  the  t2g  states  are  located  between  ~  −4.0  eV   and   the   Fermi   level   and   the   eg   states   form   conduction   bands  (see  Figure  S1).  This  indicates  a  +2  nominal  oxida-­‐ tion   state   for   Ni   (d8:   t2g6eg2),   where   the   computed   mag-­‐ netic   moment   is   ~1.8   µB.   On   the   other   hand,   the   Pb   6s   states   hybridize   strongly   with   the   O   2p   states   with   stabili-­‐ zation   of   the   hole   on   the   oxygen,   as   was   observed   in   β-­‐ PbO257   and   pentavalent   Bi   compounds.31,61   The   basic   na-­‐ ture   of   the   electronic   structure   remains   the   same   for   all   the   phases   under   the   present   conditions.   The   polar   R3c   structure   is   composed   of   an   a−a−a−   octahedral   tilt   distor-­‐ tion  that  has  the   R5−   symmetry   and   a   polar   distortion   that   has   the   Γ4−   symmetry   of   the   Pm-­‐3m   structure.   The   esti-­‐ mated   electric   polarization   is   ~35   µC/cm2   directed   along   the   cubic   [111]   axis.   The   Pbn21   and   Pbnm   structures   are   respectively  55  meV/f.u.  and  85  meV/f.u.  higher  in  energy   compared   to   the   lowest   energy   polar   R3c   phase   (see   Fig-­‐ ure  6a),  with  a  smaller  cell  volume.  The  details  of  the  op-­‐ timized   structures   are   provided   in   Table   S1.   Our   calcula-­‐ tion  results  explain  why  PbNiO3  tends  to  crystallize  in  the   R3c  structure  under  ambient  conditions.   Next,   we   studied   the   changes   in   the   structural   phase   stability  induced  by  the  volume  reduction,  as  is  expected   to  occur  under  hydrostatic  pressure.  To  this  end,  we  cal-­‐ culated  the  energies  of  R3c  and  the  low-­‐energy  Pbn21  and   Pbnm   phases   for   a   range   of   uniformly   varied   volumes   tak-­‐ ing  into  consideration  the  c/a  ratio  of  the  fully  optimized   structures.  The  atomic  positions  were  optimized  for  each   volume.   The   results   are   summarized   in   Figure   6b.   We   observed  a  phase  transition  from  the  polar  R3c  to  the  po-­‐ lar  Pbn21  structure  at  around  a  4.9%  reduction  in  volume   and   at   around   3.2   GPa   applied   pressure   (see   Figure   S3).   Here,   although   the   energy   difference   between   Pbn21   and   Pbnm   decreased   as   the   volume   decreased,   no   further   transformation  of  phase  was  observed  even  when  the  cell   volume   was   reduced   by   12%.   This   indicates   that   under   high  pressure,  the  Pbn21  structure  should  be  stabilized,  as   was  observed  in  the  high-­‐pressure  synthesis.  This  howev-­‐ er,   has   not   been   reported   in   the   previous   studies.33,56   We   cross  validated  our  results  by  taking  into  account  a  range   of  U  values  at  the  Ni  3d  states  (see  Figure  S2).  The  prima-­‐ ry   phonon   distortions   that   contribute   to   the   Pbn21   struc-­‐ ture  are  as  follows:  (i)  an   a−a−c0  octahedral  tilt  distortion   that   follows   R5−,   (ii)   an   a0a0c+   octahedral   rotation   that   follows  M2+  and  (iii)  a  polar  distortion  that  follows  Γ4−  of   the  Pm-­‐3m  structure.  Note  that  the  Pbnm  phase  is  primar-­‐ ily  composed  of  the  octahedral  a−a−c0  tilt  and  a0a0c+  rota-­‐ tional   distortions.   The   Pbn21   structure   is   expected   to   be   stabilized  by  the  Γ4−  polar  distortion,  which  is  unstable  in   the   Pm-­‐3m   structure.   The   estimated   electric   polarization   for   the   fully   optimized   Pbn21   structure   is   ~44   µC/cm2   along   the   cubic   [001]   axis,   in   good   agreement   with   the   value   calculated   from   the   experimentally   determined   structural   parameters   (37.8   µC/cm2)   in   Table   1.  

ACS Paragon Plus Environment

Page 7 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

3+

5+

4+

2+

Figure   4.   HAXPES   measurements   indicating   the   formal   valence   state   of   (Bi 0.5Bi 0.5)1−xPb xNi O3.   (a)   Composition   depend-­‐ ence  of  BVSs  for  Bi1−xPbxNiO3  at  RT.  (b)  Bi  4f  HAXPES  results  for  Bi1−xPbxNiO3  at  RT  together  with  those  for  BiFeO3  and  BiNiO3   3+ 3+/5+ as   standard   materials   for   Bi   and   Bi .   (c)   Pb   4f   HAXPES   results   for   Bi1−xPbxNiO3   at   RT   together   with   those   for   PbTiO3   and   2+ 4+ 3+ PbNiO3  as  standard  materials  for  Pb  and  Pb .  (d)  Composition  dependence  of  the  fraction  of  Bi  calculated  from  the  area  ratio   of  HAXPES  peaks.    

Magnetic   property.   We   performed   magnetic   meas-­‐ urements   to   evaluate   the   spin   and   valence   states   of   Ni.   Figure   7a   and   S4   shows   the   temperature   dependence   of   magnetic   susceptibility   and   the   magnetization   curves   at   100  K  for  Bi1−xPbxNiO3.  Weak  ferromagnetism,  most  prob-­‐ ably   due   to   the   canted   spins   induced   by   a   Dzyalonshin-­‐ sky-­‐Moriya  interaction,  was  observed  in  the  samples  with   0.25   ≤   x   ≤   0.80.   The   data   between   370   and   400   K,   well   above   the   Néel   temperatures   (TN),   can   be   fitted   to   the   Curie-­‐Weiss  law  with  a  temperature  independent  term  χ0,   χ  =  C/(T  − θ)  +   χ0.  Here,  C  is  the  Curie  constant,  and   θ  is   the  Weiss  temperature.  Figure  7b  shows  the  composition   dependence   of   TN   and   the   effective   magnetic   moments   (peff)   calculated   from   the   Curie   constant.   The   peff   values   of   2.89  -­‐  3.13   µB  are  close  to  2.83   µB  for  Ni2+  with  a  high  spin   configuration   with   S   =   1   and   g   =   2,   indicating   that   Ni   re-­‐ mains  divalent  in  all  the  compositions.  On  the  other  hand,   TN   is   almost   independent   of   x   until   x   =   0.40,   at   which   point  it  begins  to  decrease  with  increasing  x.  These  find-­‐ ings   are   consistent   with   the   results   of   structural   analysis   showing   that   Ni   remains   divalent   and   that   the   polar-­‐ nonpolar   phase   transition   occurs   around   x   =   0.40   at   RT.   TN   decreases   for   x   ≥   0.60,   suggesting   a   decrease   in   the   antiferromagnetic   interaction   owing   to   the   reduction   of   the   Ni-­‐O-­‐Ni   angles.   Indeed,   the   average   Ni-­‐O-­‐Ni   angles   (θ)  calculated  from  the  Rietveld  refinement  for  x  ≥  0.60  at   100   K   (see   Table   S2),   well   below   TNs,   systematically   de-­‐ crease   with   increasing   x:   θ   =   137.0°   (x   =   0.6),   135.9°   (x   =   0.8),  and  135.5°  (x  =  1.0).  In  the  polar  structure,  the  magni-­‐ tude   of   the   antiferromagnetic   interaction   changes   be-­‐

cause   of   the   off-­‐center   displacement   of   Ni   in   the   NiO6   octahedron,  which  is  absent  in  the  nonpolar  structure.   Phase   stability.  Finally,  we  discuss  the  experimentally   and   theoretically   determined   stability   of   the   polar   struc-­‐ ture.   Figure   8   is   a   schematic   illustration   of   the   Bi/Pb   layer   showing   the   successive   phase   transitions   of   Bi1−xPbxNiO3   depending  on  the  amount  of  Bi.  As  the  amount  of  Bi  de-­‐ creases,  the  crystal  structure  changes  from  a  triclinic  one   with  Bi3+/Bi5+  long-­‐range  ordering  to  an  orthorhombic  one   with  Bi3+/Bi5+  short-­‐range  ordering;  then  it  changes  into  a   polar   orthorhombic   one   without   Bi3+/Bi5+   ordering   and   finally  to  a  polar  LiNbO3-­‐type  one,  which  is  closely  related   to   the   perovskite-­‐type   one.36   Under   the   HP   synthesis   con-­‐ dition,   Bi   is   all   trivalent   and   partial   ordering   of   Bi3+   and   Pb4+  occurs  because  of  the  differences  in  the  valence  and   the   ionic   radius.   When   the   sample   is   cooled   and   the   pres-­‐ sure   is   released,   the   valence   state   changes   to   (Bi3+0.5Bi5+0.5)1−xPb4+xNi2+O3,  where  Bi3+/Bi5+  disproportiona-­‐ tion  occurs.  However,  when  x  exceeds  the  critical  concen-­‐ tration   of   0.2,   the   development   of   Bi3+/Bi5+   long-­‐range   ordering  is  hindered  by  the  presence  of  small  Pb4+  at  the   large   Bi3+   site.34   The   disappearance   of   Bi3+/Bi5+   short-­‐range   ordering  can  be  similarly  understood  considering  the  de-­‐ crease  in  the  number  of  Bi  ions.  Moreover,  the  increased   stability   of   the   LiNbO3-­‐type   structure   can   be   understood   in   terms   of   Goldschmidt’s   tolerance   factor   (t),   which   has   been   used   as   an   indicator   for   the   stability   of   perovskite-­‐ type  structures.  The  ionic  radii  of  Shannon  (r)  for  six-­‐fold   coordination  are  rPb4+  =  0.775  Å,  rBi  =  (rBi5+  +  rBi3+)/2  =  0.895   Å,  rNi2+  =  0.69  Å  and  rO2−  =  1.40  Å,  leading  to  t  of  0.78    

ACS Paragon Plus Environment

Chemistry of Materials

Figure   5.   Observed   (red   points),   calculated   (blue   line),   and   difference  (green  line)  PDFs  for  Bi0.2Pb0.8NiO3  with  nonpolar   Pbnm   orthorhombic,   nonpolar   P-­‐1   triclinic   and   polar   Pbn21   orthorhombic  structural  models  at  RT.       ~ 4.9%

(b)

(a) 0.8 0.6

P1

P1 R3c

0.4 Ima2

0.2 0 -0.2

Im3

Pbn21

Imma Pbnm !3c

Energy (eV/f.u.)

I4/mmm

Energy (eV/f.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 12

R3c Pbn21 Pbnm

0.2

2+

0.15 0.1 0.05 0 54

56

58

60

62

64

66

3

Volume (Å /f.u.)

Figure   6.   DFT   calculations   indicating   the   ground   states   of   LiNbO3-­‐type  R3c  and  perovskite-­‐type  Pbn21  structures  under   AP   or   HP   conditions.   (a)   Computed   energies   of   various   dis-­‐ torted   PbNiO3   structures   with   respect   to   lowest   energy   R3c   phase.   The   phases   with   energy   within   1   eV   are   shown.   (b)   Computed   total   energy   versus   volume   per   formula   unit   for   the   R3c,   Pbn21   and   Pbnm   phases.   The   results   are   for   the   lowest   energy   AFM   G-­‐type   spin   structure   and   for   the   effec-­‐ tive  U  =  6.0  eV  at  the  Ni  site.  The  results  for  the  other  U  val-­‐ ues  are  provided  in  Figure  S2.     (x  =  0),  0.76  (x  =  0.5)  and  0.74  (x  =  1.0)  for  Bi1−xPbxNiO3.  The   perovskite-­‐type  structure  is  expected  to  be  stable  within  the   62,63  whereas  the  LiNbO3-­‐type  structure  is   limits  0.75  <  t  <  1.1, stable   for   t   <   0.75.   The   calculated   t   values   are   near   the   boundary  of  the  perovskite-­‐type  and  LiNbO3-­‐type  structures.   The  inversion  symmetry  breaking  in  perovskite-­‐type  PbNiO3   2 seems   peculiar   because   there   are   neither   6s   lone   pairs   nor   second-­‐order   Jahn-­‐Teller   (JT)   active   cations.   However,   polar   materials   without   these   cations   have   been   found   in   the   64 65 LiNbO3-­‐type  InFeO3  (t  =  0.76)  and  ScFeO3  (t  =  0.74),  and   the   driving   force   for   the   inversion   symmetry   breaking   is   con-­‐ sidered   to   be   the   increased   Coulomb   interactions   between   64,66 the  high  valency  A-­‐site  cation  and  oxygen.    

Figure   7.   Magnetic   measurements   indicating   Ni   high   spin   state  and  Ni  off-­‐center  displacement  in  the  NiO6  octahedron.   (a)   Temperature   dependence   of   molar   magnetic   susceptibil-­‐ ity   for   Bi1−xPbxNiO3.   (b)   Composition   dependence   of   the   Néel   temperature   (TN)   and   effective   magnetic   moment   (peff).   The   peff  were  obtained  from  the  fittings  to  the  Curie-­‐Weiss  law.    

As  the  A-­‐O  Coulomb  interactions  increase,  the  R-­‐3c  struc-­‐ ture   with   a   nine-­‐coordinated   A-­‐site   cation   becomes   un-­‐ stable  and  changes  into  an  R3c  one  with  six-­‐coordinations   through  polar  displacement  of  A-­‐site  cation.  The  large  A-­‐ O  Coulomb  interactions  can  also  be  expected  to  stabilize   the   polar   symmetry   in   the   orthorhombic   perovskite-­‐type   materials,   from   analogy   with   LiNbO3-­‐type   case.   The   Γ4−   polar   distortion   is   indeed   found   to   be   unstable   for   the   Pm-­‐3m  structure  of  PbNiO3.  Therefore,  the  occurrence  of   the   polar   orthorhombic   perovskite-­‐type   structure   should   be   reasonable   considering   the   high   valency   state   of   Pb4+   and  the  small  t  value  close  to  those  of  LiNbO3-­‐type  mate-­‐ rials   and   the   polar   orthorhombic   perovskite-­‐type   BiInO3   (t  =  0.78).67    

4.  CONCLUSION   The  structural  analysis  using  SXRD  revealed  the  presence   of  the  three  different  orthorhombic  phases  in  Bi1−xPbxNiO3   and  temperature-­‐induced  orthorhombic-­‐to-­‐orthorhombic   phase   transitions   accompanied   by   NTE   in   two   composi-­‐ tion  ranges,  0.20  ≤  x  ≤  0.25  and  0.60  ≤  x  ≤  0.80.  The  for-­‐ mer   NTE   is   induced   by   intersite   charge   transfer   between   Bi5+   and   Ni2+,   as   has   already   been   clarified   in   a   previous  

ACS Paragon Plus Environment

Page 9 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

  Figure   8.  Illustration  of  the  Bi/Pb  layer  showing  systematic  phase  transition  of  Bi1−xPbxNiO3  according  to  the  amount  of  Bi.  As   3+ 5+ the   amount   of   Bi   decreases,   the   crystal   structure   changes   from   a   triclinic   one   with   Bi /Bi   long-­‐range   ordering   to   an   ortho-­‐ 3+ 5+ 3+ 5+ rhombic  one  with  Bi /Bi  short-­‐range  ordering;  then  it  changes  into  polar  orthorhombic  one  without  Bi /Bi  ordering  and  a   polar  LiNbO3-­‐type  one.    

study.  The  results  of  the  Rietveld  refinements,  PDF  anal-­‐ yses,  and  SHG  measurements  revealed  that  the  LT  ortho-­‐ rhombic   phase   with   0.40   ≤   x   ≤   1.00   has   a   polar   structure   with  space  group  Pbn21.  The  latter  NTE  originates  from  a   polar-­‐nonpolar   transition.   The   change   in   the   magnetic   properties   observed   at   x   =   0.50   is   consistent   with   off-­‐ center  displacement  of  the  Ni  in  NiO6  octahedron  due  to   the   polar   nature.   Perovskite-­‐related   materials   with   small   tolerance   factors   have   attracted   much   attention   because   they  tend  to  have  a  polar  structure.  However,  most  polar   materials   found   in   the   region   of   small   tolerance   factor   have   LiNbO3-­‐type   structures;   only   a   few   polar   ortho-­‐ rhombic   perovskite-­‐type   structures   have   been   found,   which  is  probably  due  to  the  high  stability  of  LiNbO3-­‐type   structure  in  this  region.  Our  results  suggest  that  the  polar   orthorhombic  phase  can  be  realized  by  using  high  valence   A-­‐site  cations  in  addition  to  controlling  the  tolerance  fac-­‐ tor.   This   approach   will   be   new   strategy   for   searching   for   functional   materials   having   a   polar   nature,   for   example,   ferroelectric  and  large  NTE  materials.  

ASSOCIATED  CONTENT     Supporting  Information   The  Supporting  Information  is  available  free  of  charge  on  the   ACS  Publications  website  at  DOI:   Computed   G-­‐type   antiferromagnetic   density   of   states  (DOS)  for  the  R3c  phase;  Computed  total  en-­‐ ergy   versus   cell   volume   per   formula   unit   for   R3c,   Pbn21   and   Pbnm   phases;   Calculated   enthalpy   as   a   function   of   applied   pressure   for   R3c,   Pbn21   and   Pbnm  phases;  The  magnetization  curves  at  100  K  for   Bi1−xPbxNiO3;   Optimized   structural   parameters   for   PbNiO3   for   the   ground   state   R3c   structure   and   low  

energy   Pbn21   and   Pbnm   structures;   Structural   Pa-­‐ rameters   for   Perovskite-­‐type   Bi1-­‐xPbxNiO3   at   100   K   (PDF)  

AUTHOR  INFORMATION   Corresponding  Author   *  [email protected]   *  [email protected]    

Notes  

The  authors  declare  no  competing  financial  interest.  

ACKNOWLEDGMENT     This   work   was   partially   supported   by   the   Grant-­‐in-­‐Aid   for   Scientific   Research   16H02393   and   18H05208   from   the   Japan   Society  for  the  Promotion  of  Science  (JSPS)  and  by  the  Pho-­‐ ton  and  Quantum  Basic  Research  Coordinated  Development   Program  of  the  Ministry  of  Education,  Culture,  Sports,  Scien-­‐ ce   and   Technology   (MEXT),   Japan.   The   synchrotron-­‐ radiation   experiments   were   performed   at   SPring-­‐8   with   the   approval  of  the  Japan  Synchrotron  Radiation  Research  Insti-­‐ tute   (2017A1242,   2017A1388,   2017B3751,   2017B1721,   2018A1642,   2018A3751,   2018A1667,   2018B3751,   2018B1797,   2018B1672   and   2018B1860).  

REFERENCES   (1)  Chu  C.  N.;  Saka,  N.;  Suh,  N.  P.  Negative  thermal  expansion   ceramics:  A  review.  Mater.  Sci.  Eng.  1987,  95,  303-­‐308.   (2)  Sleight,  A.  W.,  Compounds  That  Contract  on  Heating.  In-­‐ org.  Chem.  1998,  37,  2854-­‐2860.   (3)  Barrera,  G.  D.;  Bruno,  J.  A.  O.;  Barron,  T.  H.  K.;  Allan,  N.  L.   Negative   thermal   expansion.   J.   Phys.   Condens.   Matter   2005,   17,   R217-­‐R252.  

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(4)  Takenaka,  K.  Negative  thermal  expansion  materials:  tech-­‐ nological  key  for  control  of  thermal  expansion.  Sci.  Technol.  Adv.   Mater.  2012,  13,  013001.   (5)   Chen,   J.;   Hu,   L.;   Deng,   J.;   Xing,   X.   Negative   thermal   expan-­‐ sion   in   functional   materials:   controllable   thermal   expansion   by   chemical  modifications.  Chem.  Soc.  Rev.  2015,  44,  3522-­‐3567.   (6)  Mittal,  R.;  Gupta,  M.  K.;  Chaplot,  S.  L.  Phonons  and  anom-­‐ alous   thermal   expansion   behaviour   in   crystalline   solids.   Prog.   Mater.  Sci.  2018,  92,  360-­‐445.   (7)   Attfield,   J.   P.   Mechanisms   and   Materials   for   NTE.   Front.   Chem.  2018,  6,  371.   (8)  Mary,  T.  A.;  Evans,  J.  S.  O.;  Vogt,  T.;  Sleight,  A.  W.  Nega-­‐ tive  Thermal  Expansion  from  0.3  to  1050  Kelvin  in  ZrW2O8.  Sci-­‐ ence  1996,  272,  90-­‐92.   (9)  Phillips,  A.  E.;  Goodwin,  A.  L.;  Halder,  G.  J.;  Southon,  P.  D.;   Kepert,   C.   J.   Nanoporosity   and   Exceptional   Negative   Thermal   Expansion   in   Single-­‐Network   Cadmium   Cyanide.   Angew.   Chem.   Int.  Ed.  2008,  47,  1396-­‐1399.   (10)   Takenaka,   K.;   Takagi,   H.   Giant   negative   thermal   expan-­‐ sion  in  Ge-­‐doped  anti-­‐perovskite  manganese  nitrides.  Appl.  Phys.   Lett.  2005,  87,  261902.   (11)  Takenaka,  K.;  Asano,  K.;  Misawa,  M.;  Takagi,  H.  Negative   thermal   expansion   in   Ge-­‐free   antiperovskite   manganese   nitrides:   Tin-­‐doping  effect.  Appl.  Phys.  Lett.  2008,  92,  011927.   (12)   Takenaka,   K.;   Inagaki,   T.;   Takagi,   H.   Conversion   of   mag-­‐ netic  structure  by  slight  dopants  in  geometrically  frustrated  an-­‐ tiperovskite  Mn3GaN.  Appl.  Phys.  Lett.  2009,  95,  132508.   (13)   Matsuno,   J.;   Takenaka,   K.;   Takagi,   H.;   Matsumura,   D.;   Nishihata,  Y.;  Mizuki,  J.  Local  structure  anomaly  around  Ge  do-­‐ pants   in   Mn3Cu0.7Ge0.3N  with   negative   thermal   expansion.   Appl.   Phys.  Lett.  2009,  94,  181904.   (14)  Hamada,  T.;  Takenaka,  K.  Giant  negative  thermal  expan-­‐ sion  in  antiperovskite  manganese  nitrides.  J.  Appl.  Phys.  2011,  109,   07E309.   (15)  Tong,  P.;  Louca,  D.;  King,  G.;  Llobet,  A.;  Lin,  J.  C.;  Sun,  Y.   P.  Magnetic  transition  broadening  and  local  lattice  distortion  in   the   negative   thermal   expansion   antiperovskite   Cu1 − xSnxNMn3.   Appl.  Phys.  Lett.  2013,  102,  041908.   (16)  Takenaka,  K.;  Hamada,  T.;  Kasugai,  D.;  Sugimoto,  N.  Tai-­‐ loring  thermal  expansion  in  metal  matrix  composites  blended  by   antiperovskite   manganese   nitrides   exhibiting   giant   negative   thermal  expansion.  J.  Appl.  Phys.  2012,  112,  083517.   (17)  Chen,  J.;  Xing,  X.;  Sun,  C.;  Hu,  P.;  Yu,  R.;  Wang,  X.;  Li,  L.   Zero   Thermal   Expansion   in   PbTiO3-­‐Based   Perovskites.   J.   Am.   Chem.  Soc.  2008,  130,  1144-­‐1145.   (18)  Hu,  P.;  Chen,  J.;  Sun,  X.;  Deng,  J.;  Chen,  X.;  Yu,  R.;  Qiao,   L.;   Xing,   X.   Zero   thermal   expansion   in   (1−x)PbTiO3– xBi(Mg,Ti)1/2O3   piezoceramics.   J.   Mater.   Chem.   2009,   19,   1648-­‐ 1652.   (19)   Chen,   J.;   Nittala,   K.;   Forrester,   J.   S.;   Jones,   J.   L.;   Deng,   J.;   Yu,   R.;   Xing,   X.   The   Role   of   Spontaneous   Polarization   in   the   Negative   Thermal   Expansion   of   Tetragonal   PbTiO3-­‐Based   Com-­‐ pounds.  J.  Am.  Chem.  Soc.  2011,  133,  11114-­‐11117.   (20)  Chen,  J.;  Fan,  L.;  Ren,  Y.;  Pan,  Z.;  Deng,  J.;  Yu,  R.;  Xing,  X.   Unusual  Transformation  from  Strong  Negative  to  Positive  Ther-­‐ mal  Expansion  in  PbTiO3-­‐BiFeO3  Perovskite.  Phys.  Rev.  Lett.  2013,   110,  115901.     (21)   Yamamoto,   H.;   Imai,   T.;   Sakai,   Y.;   Azuma,   M.   Colossal   Negative   Thermal   Expansion   in   Electron-­‐Doped   PbVO3  Perovskites.  Angew.  Chem.  Int.  Ed.  2018,  57,  8170-­‐8173.     (22)  Long,  Y.  W.;  Hayashi,  N.;  Saito,  T.;  Azuma,  M.;  Muranaka,   S.;   Shimakawa,   Y.   Temperature-­‐induced   A-­‐B   intersite   charge   transfer   in   an   A-­‐site-­‐ordered   LaCu3Fe4O12   perovskite.   Nature   2009,  458,  60-­‐63.   (23)  Yamada,  I.;  Tsuchida,  K.;  Ohgushi,  K.;    Hayashi,  N.;  Kim,   J.;   Tsuji,   N.;   Takahashi,   R.;   Matsushita,   M.;   Nishiyama,   N.;   Inoue,   T.;   Irifune,   T.;   Kato,   K.;   Takata,   M.;   Takano,   M.   Giant   negative  

Page 10 of 12

thermal   expansion   in   the   iron   perovskite   SrCu3Fe4O12.   Angew.   Chem.  Int.  Ed.  2011,  50,  6579-­‐6582.   (24)   Ishiwata,   S.;   Azuma,   M.;   Hanawa,   M.;   Moritomo,   Y.;   Ohishi,  Y.;  Kato,  K.;  Takata,  M.;  Nishibori,  E.;  Sakata,  M.;  Terasa-­‐ ki,   T.;   Takano,   M.   Pressure/temperature/substitution-­‐induced   melting   of  A-­‐site   charge   disproportionation   in  Bi1 − xLaxNiO3  (0≤x≤0.5).  Phys.  Rev.  B  2005,  72,  045104.   (25)  Azuma,  M.;  Chen,  W.;  Seki,  H.;  Czapski,  M.;  Olga,  S.;  Oka,   K.;   Mizumaki,   M.;   Watanuki,   T.;   Ishimatsu,   N.;   Kawamura,   N.;   Ishiwata,  S.;  Tucker,  M.  G.;  Shimakawa,  Y.;  Attfield,  J.  P.  Colossal   negative   thermal   expansion   in   BiNiO3   induced   by   intermetallic   charge  transfer.  Nat.  Commun.  2011,  2,  347.   (26)  Oka,  K.;  Sakaguchi,  C.;  Sinclair,  A.;  Ritter,  C.;  Mizumaki,   M.;  Attfield,  J.  P.;  Azuma,  M.  Intermetallic  charge-­‐transfer  tran-­‐ sition  in  Bi1−xLaxNiO3  as  the  origin  of  the  colossal  negative  ther-­‐ mal  expansion.  Phys.  Rev.  B  2013,  88,  014112.   (27)   Nabetani,   K.;   Muramatsu,   Y.;   Oka,   K.;   Nakano,   K.;   Hojo,   H.;   Mizumaki,   M.;   Agui,   A.;   Higo,   Y.;   Hayashi,   N.;   Takano,   M.;   Azuma,   M.   Suppression   of   temperature   hysteresis   in   negative   thermal   expansion   compound   BiNi1 − xFexO3  and   zero-­‐thermal   expansion  composite.  Appl.  Phys.  Lett.  2015,  106,  061912.   (28)  Naka,  M.;  Seo,  H.;  Motome,  Y.  Theory  of  Valence  Transi-­‐ tion  in  BiNiO3.  Phys.  Rev.  Lett.  2016,  116,  056402.   (29)     Nishikubo,   T.;   Sakai,   Y.;   Oka,   K.;   Mizumaki,   M.;     Wa-­‐ tanuki,  T.;  Machida,  A.;  Maejima,  N.;  Ueda,  S.;  Mizokawa,  T.;  M.   Azuma,   M.   Optimized   negative   thermal   expansion   induced   by   gradual  intermetallic  charge  transfer  in  Bi1−xSbxNiO3.  Appl.  Phys.   Express  2018,  11,  061102.     (30)  Ishiwata,  S.;  Azuma,  M.;  Takano,  M.;  Nishibori,  E.;  Takata,   M.;   Sakata,   M.;   Kato,   K.   High   pressure   synthesis,   crystal   struc-­‐ ture  and  physical  properties  of  a  new  Ni  (II)  perovskite  BiNiO3.  J.   Mater.  Chem.  2002,  12,  3733-­‐3737.   (31)   Azuma,   M.;   Carlsson,   S.;   Rodgers,   J.;   Tucker,   M.   G.;   Tsu-­‐ jimoto,   M.;   Ishiwata,   S.;   Isoda,   S.;   Shimakawa,   Y.;   Takano,   M.;   Attfield,  J.  P.  Pressure-­‐induced  intermetallic  valence  transition  in   BiNiO3.  J.  Am.  Chem.  Soc.  2007,  129,  14433-­‐14436.   (32)  Mizumaki,  M.;  Ishimatsu,  N.;  Kawamura,  N.;  Azuma,  M.;   Shimakawa,  Y.;  Takano,  M.;  Uozumi,  T.  Direct  observation  of  the   pressure-­‐induced   charge   redistribution   in  BiNiO3  by   x-­‐ray   ab-­‐ sorption  spectroscopy.  Phys.  Rev.  B  2009,  80,  233104.     (33)  Ishiwata,  S.;  Azuma,  M.;  Takano,  M.  Structure  and  Physi-­‐ cal   Properties   of   Perovskite   Bi0.8Pb0.2NiO3  in   Unusual   Valence   State  A4+B2+O3.  Chem.  Mater.  2006,  19,  1964-­‐1967.     (34)   Nakano,   K.;   Oka,   K.;   Watanuki,   T.;   Mizumaki,   M.;   Machida,   A.;   Agui,   A.;   Kim,   H.;   Komiyama,   J.;   Mizokawa,   T.;   Nishikubo,  T.;  Hattori,  Y.;  Ueda,  S.;  Sakai,  Y.;  Azuma,  M.  Glassy   Distribution   of   Bi3+/Bi5+  in   Bi1 – xPbxNiO3  and   Negative   Thermal   Expansion   Induced   by   Intermetallic   Charge   Transfer.   Chem.   Mater.  2016,  28,  6062-­‐6067.     (35)  Inaguma,  Y.;  Yoshida,  M.;  Tsuchiya,  T.;  Aimi,  A.;  Tanaka,   K.;   Katsumata,   T.;   Mori,   D.   High-­‐pressure   synthesis   of   novel   lithium  niobate-­‐type  oxides.  J.  Phys.:  Conf.  Ser.  2010,  215,  012131.   (36)   Inaguma,   Y.;   Tanaka,   K.;   Tsuchiya,   T.;   Mori,   D.;   Katsuma-­‐ ta,   T.;   Ohba,   T.;   Hiraki,   K.;   Takahashi,   T.;   Saitoh,   H.   Synthesis,   structural   transformation,   thermal   stability,   valence   state,   and   magnetic   and   electronic   properties   of   PbNiO3   with   perovskite-­‐   and   LiNbO3-­‐type   structures.   J.   Am.   Chem.   Soc.   2011,   133,   16920-­‐ 16929.   (37)  Yasukawa,  M.;  Kono,  T.;  Ueda,  K.;  Yanagi,  H.;  Kim,  S.  W.;   Hosono,   H.   Thermoelectric   properties   and   figure   of   merit   of   perovskite-­‐type   Ba1-­‐xLaxSnO3  with   x=0.002-­‐0.008.   Solid   State   Commun.  2013,  172,  49-­‐53.   (38)  Kakihana,  M.  Invited  review  “sol-­‐gel”  preparation  of  high   temperature   superconducting   oxides.   J.   Sol-­‐Gel   Sci.   Technol.   1996,  6,  7-­‐55.  

ACS Paragon Plus Environment

Page 11 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

(39)   Kawaguchi,   S.;   Takemoto,   M.;   Osaka,   K.;   Nishibori,   E.;   Moriyoshi,   C.;   Kubota,   Y.;   Kuroiwa,   Y.;   Sugimoto,   K.   High-­‐ throughput   powder   diffraction   measurement   system   consisting   of  multiple  MYTHEN  detectors  at  beamline  BL02B2  of  SPring-­‐8.   Rev.  Sci.  Instrum.  2017,  88,  085111.   (40)   Nishibori,   E.;   Takata,   M.;   Kato,   K.;   Sakata,   M.;   Kubota,   Y.;   Aoyagi,   S.;   Kuroiwa,   Y.;   Yamakata,   M.;   Ikeda,   N.   The   large   De-­‐ bye–Scherrer   camera   installed   at   SPring-­‐8   BL02B2   for   charge   density   studies.   Nucl.   Instrum.   Methods   Phys.   Res.   A   2001,   467-­‐ 468,  1045-­‐1048.   (41)  Izumi,  F.;  Momma,  K.  Three-­‐dimensional  visualization  in   powder  diffraction.  Solid  State  Phenom.  2007,  130,  15-­‐20.   (42)  Watanuki,  T.;  Machida,  A.;  Ikeda,  T.;  Ohmura,  A.;  Kaneko,   H.;   Aoki,   K.;   Sato,   T.   J.;   Tsai,   A.   P.   Development   of   a   single-­‐ crystal  X-­‐ray  diffraction  system  for  hydrostatic-­‐pressure  and  low-­‐ temperature   structural   measurement   and   its   application   to   the   phase  study  of  quasicrystals.  Philos.  Mag.  2007,  87,  2905-­‐2911.   (43)   Egami,   T.;   Billinge,   S.   J.   L.   Underneath   the   Bragg   Peaks:   Structural   Analysis   of   Complex   Materials;   Pergamon   Press:   Ox-­‐ ford,  U.K.,  2003.   (44)   Qiu,   X.;   Thompson,   J.   W.;   Billinge,   S.   J.   L.   PDFgetX2:   a   GUI-­‐driven   program   to   obtain   the   pair   distribution   function   from  X-­‐ray  powder  diffraction  data.  J.  Appl.  Cryst.  2004,  37,  678.   (45)  Farrow,  C.  L.;  Juhas,  P.;  Liu,  J.  W.;  Bryndin,  D.;  Božin,  E.   S.;   Bloch,   J.;   Proffen,   T.;   Billinge,   S.   J.   L.   PDFfit2   and   PDFgui:   computer   programs   for   studying   nanostructure   in   crystals.   J.   Phys.:  Condens.  Matter  2007,  19,  335219.     (46)   Anisimov,   V.   I.;  Aryasetiawan,   F.;   Lichtenstein,   A.   I.   First-­‐ principles  calculations  of  the  electronic  structure  and  spectra  of   strongly   correlated   systems:   the   LDA+U   method.   J.   Phys.:   Con-­‐ dens.  Matter.  1997,  9,  767-­‐808.   (47)  Perdew,  J.  P.;  Burke,  K.;  Ernzerhof,  M.  Generalized  gradi-­‐ ent  approximation  made  simple.  Phys.   Rev.   Lett.  1996,  77,  3865-­‐ 3868.   (48)   Kresse,   G.;   Hafner,   J.   Ab   initio   molecular   dynamics   for   liquid  metals.  Phys.  Rev.  B  1993,  47,  558–561.   (49)   Kresse,   G.;   Furthmüller,   J.   Efficient   iterative   schemes   for   ab   initio   total-­‐energy   calculations   using   a   plane-­‐wave   basis   set.   Phys.  Rev.  B  1996,  54,  11169-­‐11186.     (50)  Dudarev,  S.  L.;  Botton,  G.  A.;  Savrasov,  S.  Y.;  Humphreys,   C.   J.;   Sutton,   A.   P.   Electron-­‐Energy-­‐Loss   Spectra   and   the   Struc-­‐ tural   Stability   of   Nickel   Oxide:   An   LSDA+U   Study.   Phys.   Rev.   B   1998,  57,  1505-­‐1509.     (51)   King-­‐Smith,   R.   D.;   Vanderbilt   D.   Theory   of   polarization   of   crystalline  solids.  Phys.  Rev.  B  1993,  47,  1651-­‐1654.   (52)   Resta,   R.   Macroscopic   polarization   in   crystalline   dielec-­‐ trics:   the   geometric   phase   approach.   Rev.   Mod.   Phys.   1994,   66,   899-­‐915.     (53)  Vanderbilt,  D.;  King-­‐Smith,  R.  D.  Electric  polarization  as   a   bulk   quantity   and   its   relation   to   surface   charge.   Phys.   Rev.   B   1993,  48,  4442-­‐4455.   (54)  Sakai,  Y.;  Yang,  J.;  Yu,  R.;  Hojo,  H.;  Yamada,  I.;  Miao,  P.;   Lee,   S.;   Torii,   S.;   Kamiyama,   T.;   Ležaić,   M.;   Bihlmayer,   G.;   Mi-­‐ zumaki,   M.;   Komiyama,   J.;   Mizokawa,   T.;   Yamamoto,   H.;   Nishikubo,  T.;  Hattori,  Y.;  Oka,  K.;  Yin,  Y.;  Dai,  J.;  Li,  W.;  Ueda,   S.;   Aimi,   A.;   Mori,   D.;   Inaguma,   Y.;   Hu,   Z.;   Uozumi,   T.;   Jin,   C.;   Long,   Y.;   Azuma,   M.   A-­‐Site   and   B-­‐Site   Charge   Orderings   in   an   s-­‐ d   Level   Controlled   Perovskite   Oxide   PbCoO3.   J.   Am.   Chem.   Soc.   2017,  139,  4574-­‐4581.  

(55)  Wadati,  H.;  Takizawa,  M.;  Tran,  T.  T.;  Tanaka,  K.;  Mizo-­‐ kawa,  T.;  Fujimori,  A.;  Chikamatsu,  A.;  Kumigashira,  H.;  Oshima,   M.;  Ishiwata,  S.;  Azuma,  M.  Takano,  M.  Valence  changes  associ-­‐ ated   with   the   metal-­‐insulator   transition   in   Bi1-­‐xLaxNiO3.   Phys.   Rev.  B  2005,  72,  155103.   (56)   Foyevtsova,   K.;   Khazraie,   A.;   Elfimov,   I.;   Sawatzky,   G.   A.   Hybridization   effects   and   bond   disproportionation   in   the   bis-­‐ muth  perovskites.  Phys.  Rev.  B  2015,  91,  121114.   (57)   Shan,   Y.   J.;   Mori,   H.;   Imoto,   H.;   Itoh,   M.   Ferroelectric   Phase  Transition  in  Perovskite  Oxide  CdTiO3.  Ferroelectrics  2002,   270,  381-­‐386.   (58)  Shan,  Y.  J.;  Mori,  H.;  Tezuka,  K.;  Imoto,  H.;  Itoh,  M.  Fer-­‐ roelectric   Phase   Transition   in   CdTiO3   Single   Crystal.   Ferroelec-­‐ trics  2003,  284,  107-­‐112.   (59)   Hao,   X.   F.;  Stroppa,   A.;   Barone,   P.;  Filippetti,   A.;  Franchini,  C.;   Picozzi,   S.   Structural   and   ferroelectric   transi-­‐ tions  in  magnetic  nickelate  PbNiO3.  New  J.  Phys.  2014,  16  015030.               (60)   Payne,   D.   J.;   Egdell,   R.   G.;   Paolicelli,   G.;   Offi,   F.;   Panac-­‐ cione,  G.;  Lacovig,  P.;  Monaco,  G.;  Vanko,  G.;  Walsh,  A.;  Watson,   G.  W.;  Guo,  J.;  Beamson,  G.;  Glans,  P.  -­‐A.;  Learmonth,  T.;  Smith,   K.  E.  Nature  of  electronic  states  at  the  Fermi  level  of  metallic  β-­‐ PbO2   revealed   by   hard   x-­‐ray   photoemission   spectroscopy.   Phys.   Rev.  B  2007,  75,  15310.     (61)  Saiduzzaman,  M.;  Takei,  T.;  Yanagida,  S.;  Kumada,  N.;  Das,   H.;  Kyokane,  H.;  Wakazaki,  S.;  Azuma,  M.;  Moriyoshi,  C.;  Kuro-­‐ iwa,   Y.   Hydrothermal   Synthesis   of   Pyrochlore-­‐Type   Pentavalent   Bismuthates   Ca2Bi2O7   and   Sr2Bi2O7.   Inorg.   Chem.  2019,  58,   1759-­‐ 1763.   (62)  Reaney,  I.  M.;  Colla,  E.  L.;  Setter,  N.  Dielectric  and  Struc-­‐ tural  Characteristics  of  Ba-­‐  and  Sr-­‐based  Complex  Perovskites  as   a   Function   of   Tolerance   Factor.   Jpn.   J.   Appl.   Phys.   1994,   33,   3984-­‐ 3990.   (63)  Li,  C.;  Lu,  X.;  Ding,  W.;  Feng,  L.;  Gao,  Y.;  Guo,  Z.  Forma-­‐ bility  of  ABX3  (X  =  F,  Cl,  Br,  I)  halide  perovskites.  Acta  Cryst.  B.   2008,  64,  702-­‐707.   (64)   Fujita,   K.;   Kawamoto,   T.;   Yamada,   I.;   Hernandez,   O.;   Hayashi,   N.;   Akamatsu,   H.;   Lafargue-­‐Dit-­‐Hauret,   W.;   Rocque-­‐ felte,   X.;   Fukuzumi,   M.;   Manuel,   P.;   Studer,   A.   J.;   Knee,   C.   S.;   Tanaka,  K.  LiNbO3-­‐Type  InFeO3:  Room-­‐Temperature  Polar  Mag-­‐ net  without  Second-­‐Order  Jahn–Teller  Active  Ions.  Chem.  Mater.   2016,  28  (18),  6644-­‐6655.   (65)  Kawamoto,  T.;  Fujita,  K.;  Yamada,  I.;  Matoba,  T.;  Kim,  S.   J.;  Gao,  P.;  Pan,  X.;  Findlay,  S.  D.;  Tassel,  C.;  Kageyama,  H.;  Stu-­‐ der,  A.  J.;  Hester,  J.;  Irifune,  T.;  Akamatsu,  H.;  Tanaka,  K.  Room-­‐ temperature  polar  ferromagnet  ScFeO3  transformed  from  a  high-­‐ pressure  orthorhombic  perovskite  phase.  J.  Am.  Chem.  Soc.  2014,   136,  15291-­‐15299.     (66)   Xiang,   H.   J.   Origin   of   polar   distortion   in  LiNbO3-­‐type   “ferroelectric”   metals:   Role   of  A-­‐site   instability   and   short-­‐range   interactions.  Phys.  Rev.  B  2014,  90,  094108.   (67)  Belik,  A.  A.;  Yu,  S.;  Lazoryak,  B.  I.;  Takayama-­‐Muromachi,   E.   BiInO3:   A   Polar   Oxide   with   GdFeO3-­‐Type   Perovskite   Structure.   Chem.  Mater.  2006,  18,  1964-­‐1968.          

 

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 12

 

 

   

ACS Paragon Plus Environment

12