The Chemical Structure of Carbon Nanothreads Analyzed by

Jan 26, 2015 - We distinguish between degree-‐2 polymer (a possible inter-‐ mediate along the addition pathway to degree-‐4 and de-‐ gree-‐6...
1 downloads 0 Views 3MB Size
Subscriber access provided by University of Massachusetts Amherst Libraries

Article

The Chemical Structure of Carbon Nanothreads Analyzed by Advanced Solid-State NMR Pu Duan, Xiang Li, Tao Wang, Bo Chen, Stephen J. Juhl, Daniel Koeplinger, Vincent H. Crespi, John V. Badding, and Klaus Schmidt-Rohr J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b03733 • Publication Date (Web): 29 May 2018 Downloaded from http://pubs.acs.org on May 29, 2018

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 11 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

Journal of the American Chemical Society

The Chemical Structure of Carbon Nanothreads Analyzed by Advanced Solid-State NMR Pu  Duan,1  Xiang  Li,2,3  Tao  Wang,3,4  Bo  Chen,5  Stephen  J.  Juhl,2,3  Daniel  Koeplinger,2,3  Vincent  H.   Crespi,2,3,4,6  John  V.  Badding,2,3,4,6,*  Klaus  Schmidt-­‐Rohr1,*   1

Department  of  Chemistry,  Brandeis  University,  Waltham,  MA  02453,  USA  

2 3

Department  of  Chemistry,  Pennsylvania  State  University,  University  Park,  PA  16802,  USA  

Materials  Research  Institute,  Pennsylvania  State  University,  University  Park,  PA  16802,  USA  

4

Department  of  Physics,  Pennsylvania  State  University,  University  Park,  PA  16802,  USA  

5

Department  of  Chemistry  and  Chemical  Biology,  Baker  Laboratory,  Cornell  University,  Ithaca,  New  York  14853-­‐ 1301,  USA     6

Department  of  Materials  Science  and  Engineering,  Pennsylvania  State  University,  University  Park,  PA  16802,  USA  

  ABSTRACT:  Carbon  nanothreads  are  a  new  type  of  one-­‐dimensional  sp3-­‐carbon  nanomaterial  formed  by  slow  compres-­‐ sion  and   decompression   of  benzene.  We   report  characterization   of   the  chemical   structure  of  13C-­‐enriched   nanothreads  by   advanced  quantitative,  selective,  and  two-­‐dimensional  solid-­‐state  nuclear  magnetic  resonance  (NMR)  experiments  com-­‐ plemented   by   infrared   (IR)   spectroscopy.   The   width   of   the   NMR   spectral   peaks   suggests   that   the   nanothread   reaction   products   are   much   more   organized   than   amorphous   carbon.   In   addition,   there   is   no   evidence   from   NMR   of   a   second   phase  such  as  amorphous  mixed  sp2/sp3-­‐carbon.    Spectral  editing  reveals  that  most  carbon  atoms  are  bonded  to  one  hy-­‐ drogen  atom  as  is  expected  for  enumerated  nanothread  structures.  Characterization  of  the  local  bonding  structure  con-­‐ firms   the   presence   of   pure   fully   saturated   “degree-­‐6”   carbon   nanothreads   previously   deduced   on   the   basis   of   crystal   pack-­‐ ing   considerations   from   diffraction   and   transmission   electron   microscopy.   These   fully   saturated   threads   comprise   be-­‐ tween  25%  and  50%  of  the  sample.  Furthermore,   13C-­‐13C  spin  exchange  experiments  indicate  that  the  length  of  the  fully   saturated  regions  of  the  threads  exceeds  2.5  nm.  Two-­‐dimensional   13C-­‐13C  NMR  spectra  showing  bonding  between  chemi-­‐ cally  nonequivalent  sites  rule  out  enumerated  single-­‐site  thread  structures  such  as  polytwistane  or  tube  (3,0)  but  are  con-­‐ sistent  with  multi-­‐site  degree-­‐6  nanothreads.  Approximately  a  third  of  the  carbon  is  in  “degree-­‐4”  nanothreads  with  iso-­‐ lated   double   bonds.     The   presence   of   doubly   unsaturated   degree-­‐2   benzene   polymers   can   be   ruled   out   on   the   basis   of   13C-­‐ 13 C  NMR  with  spin  exchange  rates  tuned  by  rotational  resonance  and   1H  decoupling.  A  small  fraction  of  the  sample  con-­‐ sists   of   aromatic   rings   within   the   threads   that   link   sections   with   mostly   saturated   bonding.   NMR   provides   the   detailed   bonding   information   necessary   to   refine   solid-­‐state   organic   synthesis   techniques   to   produce   pure   degree-­‐6   or   degree-­‐4   carbon  nanothreads.  

tion  because  large  changes  in  geometry  must  occur  as  the  

INTRODUCTION 3

Carbon   nanothreads   are   a   novel   sp -­‐bonded   one-­‐ dimensional   carbon   nanomaterial.1,2   They   fill   the   last   en-­‐ try   in   a   matrix   of   carbon   nanomaterial   hybridization   (sp2/sp3)   and   dimensionality   (0/1/2/3D).3   Fully   saturated   “degree-­‐6”  nanothreads4,5  may  uniquely  combine  extreme   strength,   flexibility   and   resilience6,7   while   partially   satu-­‐ rated   “degree-­‐4”   threads4   may   act   as   novel   organic   con-­‐ ductors   (Figure   1).1,2   In   a   remarkable   solid   state   chemical   reaction  at  ~20  GPa,  dense  high-­‐symmetry  hexagonal  sin-­‐ gle   crystal   packings   of   carbon   nanothreads   emerge   from   within   a   low-­‐symmetry   monoclinic   polycrystalline   ben-­‐ zene  molecular  solid,  apparently  from  benzene  “degree-­‐0”   (Figure  1)  molecular  stacks.2     Formation  of   a  single-­‐crystal  product  in  the  solid  state   usually   requires   topochemical   reaction8   from   single-­‐ crystal   reactants   in   which   there   is   near   commensuration   between  the  periodicities  before  and  after  reaction.  Large   changes  in  unit-­‐cell  dimensions  typically  break  up  crystal   order   and   often   lead   to   amorphous   products.   Benzene   would   not   be   expected   to   undergo   a   topochemical   reac-­‐

Figure   1.   Benzene   stack  and   example   nanothreads   with  de-­‐ grees   of   saturation   of   2,   4,   and   6.   Bonds   formed   between   benzene  rings  are  in  red.  Not  all  hydrogens  are  shown.  

van   der   Waals   separations   between   molecules   are   re-­‐ placed   by   shorter,   kinetically   stable   covalent   carbon-­‐ carbon   bonds.   Indeed,   the   reaction   products   formed   by  

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

mechanochemical   compression   of   benzene   were   de-­‐ scribed   as   amorphous9,10   prior   to   the   discovery   that   con-­‐ trol   of   reaction   kinetics   by   slow   compres-­‐ sion/decompression1   results   in   formation   of   nanothreads   that   appear   to   spontaneously   self-­‐assemble   into   single   crystal   packings.2   Lifting   the   constraint   of   topochemical   reaction   could   enable   many   aromatic   molecules   to   react   to   form   new   types   of   nanothreads   with   heteroatoms   in   their  carbon  backbone  or  substitution  of  the  exterior  hy-­‐ drogens   (e.g.,   by   halogens   or   other   functional   groups).2   For  example,  we  have  recently  shown  that  pyridine,  which   has  a  different  solid-­‐state  crystal  structure  than  benzene,   forms   carbon-­‐nitride   nanothread   crystals.2,11   Thus   the   “soft”   room-­‐temperature   solid-­‐state12   organic   synthesis   employed   for   nanothreads   allows   for   incorporation   of   much   more   nitrogen   into   crystalline   carbon   nanomateri-­‐ als   than   the   high   temperature   conditions   typically   used   for   synthesis   of   carbon   nanotubes,   fullerenes,   diamond,   and   nanodiamond.11,13   Reaction   pressures   decrease   upon   decreasing   reactant   aromaticity,   increasing   temperature,   and   photochemical   irradiation,14   suggesting   a   route   to   nanothread  synthesis  at  the  pressures  of  several  GPa  used   for   synthesis   of   >106   kg/yr   of   diamond   at   modest   cost.15   Control  of  chemical  structure  for  a  wide  range  of  proper-­‐ ties   may   thus   be   possible;   for   example,   incorporation   of   nitrogen  into  nanothreads  allows  for  tuning  of  their  pho-­‐ toluminescence11   and   bandgaps   and   likely   improves   their   solubility   or   dispersion   in   protic   solvents.16   Moreover,   in   contrast   to   sp2-­‐bonded   nanomaterials   such   as   nanotubes   and  graphene,  functionalizing  or  crosslinking  the  exterior   of  degree-­‐6  nanothreads  does  not  disrupt  or  weaken  their   carbon  backbone.15   X-­‐ray   and   electron   diffraction   experiments   reveal   sym-­‐ metric   two-­‐dimensional   spot   patterns   that   provide   com-­‐ pelling   evidence   for   hexagonal   single   crystal   packings   hundreds   of   microns   across   of   carbon   (and   carbon   ni-­‐ tride)11  nanothreads  that  are  spaced  ~6.5  Å  apart,  in  good   agreement   with   the   spacing   predicted   by   modeling   of   degree-­‐6   nanothreads.2   The   threads   and   their   hexagonal   packing   are   evident   in   transmission   electron   microscopy   (TEM)   imaging   as   well.1,17   However,   although   NMR   spec-­‐ troscopy   has   revealed   that   nanothreads   consist   primarily   (75–80%)  of  sp3-­‐bonded  carbon,1  the  details  of  the  chemi-­‐ cal   structure   along   their   length   are   not   yet   understood.   Fifty  different  degree-­‐6  nanothread  structures  with  6  sp3-­‐ carbons   per   benzene   formula   unit   (Figure   1)   have   been   enumerated   by   theory.15   There   are   also   other   classes   of   reaction   products,   specifically   “singly”   unsaturated   de-­‐ gree-­‐4   nanothreads,   which   have   4   sp3-­‐carbons   per   ben-­‐ zene   formula   unit   (Figure   1),   and   “doubly”   unsaturated   degree-­‐2  polymers,  which  have  2  sp3-­‐carbons  per  benzene   formula  unit  (Figure  1).4,5  Many  of  the  pure  degree-­‐6  and   pure   degree-­‐4   nanothread   structures   have   approximately   the   same   diameter   and   thus   cannot   be   easily   distin-­‐ guished  from  each  other  or  from  mixtures  of  both  degrees   merely   by   geometric   packing   considerations,   although   detectable  differences  in  lattice  constants  are  expected  in   certain  cases.2  We  observe  only  (hk0)   diffraction  spots  for   carbon  nanothreads,2  which  indicates  that  they  are  either   helical,  axially  disordered,  and/or  lacking  in  registry  from  

Page 2 of 11

thread   to   thread.   Nonetheless,   uniform   “beading”   pat-­‐ terns  parallel  to  the  nanothread  axes  and  ~15  nm  in  length   (i.e.,  ~100  carbon  bond  lengths)  are  observed  in  TEM  im-­‐ ages,   suggesting   that   local   order   can   be   present   over   at   least  this  distance.17   Knowledge   of   the   atomic   structure   of   nanothreads   is   important  to  understanding  both  their  structure-­‐property   relations   for   applications2,11,13   and   the   non-­‐topochemical   synthesis   reaction,   and   also   for   improving   synthesis   pro-­‐ tocols  to   enable   the   production   of  large  quantities  of   pure   nanothreads   of   desired   structure,   hybridization,   and   composition.   Solid-­‐state   nuclear   magnetic   resonance   (NMR)   spectroscopy   is   an   excellent   method   for   compre-­‐ hensive  structural  and  quantitative18  compositional  analy-­‐ sis   of   organic   materials19-­‐27   and   has   played   a   seminal   role   in  elucidating  structure–property  relations  as  new  carbon   nanomaterials   such   as   sp2-­‐bonded   nanotubes28   and   sp3-­‐ bonded  nanodiamond29  have  emerged.  Here  we  present  a   systematic   and   detailed   analysis   of   the   chemical   and   na-­‐ nometer-­‐scale   composition   of   carbon   nanothreads   that   uses   advanced   NMR   spectroscopic   techniques   comple-­‐ mented  by  infrared  (IR)  spectroscopy.  

13

Figure   2.   C   NMR   spectra   of   carbon   nanothreads   synthe-­‐ 13 sized   from   C-­‐enriched   benzene.   (a)   Quantitative   multiCP   spectrum   of   all   carbons.   (b)   CH-­‐only   spectrum   obtained   by   dipolar   DEPT,   with   a   spectrum   of   amorphous   polystyrene   13 (PS)   C-­‐labeled   in   the   backbone   CH   position  shown   for  ref-­‐ erence   (dashed   line);   (c)   CH2-­‐only   spectrum   by   three-­‐spin   coherence  selection;  (d)  dipolar  dephased  multiCP  spectrum   of  nonprotonated  and  mobile  carbon.  “ssb”  marks  a  spinning   sideband.  

Two-­‐dimensional   13C-­‐13C  NMR  on  organic  materials  can   provide  detailed  information  about  bonding  and  molecu-­‐ lar   structure.21,22,25-­‐27,30   Here   we   apply   it,   combined   with   spectral   editing,   to   13C   enriched   nanothreads,   which   ena-­‐ bles   us   to   identify   4-­‐atom   segments.30   Specifically,   on   a  

ACS Paragon Plus Environment

2

Page 3 of 11 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

Journal of the American Chemical Society

sample  synthesized  mechanochemically  from  13C-­‐enriched   benzene,2   we   quantify   the   composition   in   terms   of   CH,31   CH2,32   and   CH3   groups   as   well   as   C   not   bonded   to   H.33   Alkenes   and   aromatic   carbons   are   identified   through   di-­‐ polar   dephasing   studies   as   well   as   from   the   13C   and   1H   chemical   shifts   in   two-­‐dimensional   NMR   spectra.   We   distinguish   between   degree-­‐2   polymer   (a   possible   inter-­‐ mediate   along   the   addition   pathway   to   degree-­‐4   and   de-­‐ gree-­‐6   nanothreads;   see   Figure   1)4   and   degree-­‐4   nanothreads   (likely   intermediates   along   the   addition   pathway   to   degree-­‐6   nanothreads)4   by   13C-­‐13C   correlation   experiments   with   tuning   of   spin-­‐exchange   rates   using   rotational   resonance   and   1H   decoupling   that   reveal   the   characteristic   sp3-­‐sp2-­‐hybridized   carbon   ratio   in   the   al-­‐ kene-­‐containing   structures.   We   use   long-­‐range   13C   spin   diffusion  experiments  to  estimate  the  length  of  saturated   “diamondoid”   degree-­‐6   nanothread   segments.   Our   two-­‐ dimensional  13C-­‐13C  NMR  spectra  furthermore  enable  us  to   distinguish   between   degree-­‐6   nanothreads   that   have   a   single  C  site,  such  as  single-­‐site  polytwistane  (143652)5  or   tube   (3,0)   (123456),   and   multisite   structures.   The   com-­‐ bined   experimental   observations   provide   compelling   evi-­‐ dence   for   the   synthesis   of   organized   carbon   nanothreads   rather  than  amorphous  carbon.      

heteronuclear   decoupling.30   Correlation   selectively   of   signals   of   nonprotonated   or   mobile   carbons   with   those   of   neighboring   carbons  was  achieved  in  a  simplified  version  of  EXPANSE  NMR30   with  multiCP  followed  by  50-­‐ms  13C  spin  diffusion  and  recoupled   dipolar  dephasing  before  detection.  Rotational  resonance  and   1H   decoupling   during   the   spin-­‐diffusion   period   was   used   to   speed   up   sp2-­‐C   to   sp3-­‐C   and   slow   down   sp3-­‐C   to   sp3-­‐C   magnetization   transfer,   which   provides   more   distinctive   spin-­‐exchange   behav-­‐ ior  of  degree-­‐4  nanothread  vs.  cyclohexadiene-­‐containing  struc-­‐ tures.   Spinning   sidebands   in   the   ω2   dimension   were   suppressed   by   TOSS   before   detection,   while   the   contribution   of   the   small   spinning   sideband   in   the   ω1   dimension   was   subtracted   out   by   assuming   it   to   be   equal   to   the   corresponding   negative   spinning   sideband   observed   on   the   other   side   of   the   diagonal   peak.   Spin   exchange   dynamics   were   simulated   in   MATLAB   for   several   al-­‐ kene-­‐containing   structural   models,   as   described   in   detail   in   the   Supporting   Information.   The   exchange-­‐matrix   formalism   was   used,   with   successive   multiplications   by   a   single   matrix   exp(ΠΔt),   where   Π   is   the   exchange   matrix   containing   the   ex-­‐ change  rates  as  off-­‐diagonal  elements.  

EXPERIMENTAL Synthesis.   As  reported  previously,2   we  compressed  polycrystal-­‐ line  mixtures9,10,34  of  13C  enriched  benzene  in  solid  phases  I  and  II   to  23  GPa  in  a  Paris-­‐Edinburgh  press  over  8  hours  at  2–3  GPa/hr   from  14  GPa  to  19  GPa  and  0.6–1.2  GPa/hr  from  19  GPa  to  23  GPa,   held  them  at  pressure  for  1  hour,  and  released  them  to  ambient   pressure   over   6–8   hours   at   the   same   rates   in   the   same   pressure   ranges  as  for  compression.   NMR.  Solid-­‐state  NMR  experiments  were  performed  on  a  Bruker   Avance  DSX400  spectrometer  at  a  13C  resonance  frequency  of  100   MHz,  using  a  Bruker  double-­‐resonance  4-­‐mm  magic-­‐angle  spin-­‐ ning   (MAS)   probehead.   The   ~1-­‐mg   sample   was   center-­‐packed   between  cylindrical  glass  and  hollow  KelF  spacers.  1H  and  13C  90o   pulse  lengths  were  4.2  μs,  and   1H  TPPM  decoupling  was  applied   at   |γ   B1|/2π   =   60   kHz.   A   quantitative   13C   NMR   spectrum   was   measured  using  multiCP  at  14  kHz  MAS,  with  a  Hahn  spin  echo   generated   by   an   180°   pulse   with   EXORCYCLE   phase   cycling35   applied   one   rotation   period   after   the   end   of   cross   polarization   (CP).  The  relative  intensities  of  the  two  main  peaks  were  verified   by  direct  polarization  with  60  s  recycle  delays.  Minimal  (3%)  and   not   significantly   (2   by  dashed  squares.  MAS  frequency:  14  kHz.  

RESULTS AND DISCUSSION In   this   section,   we   first   compare   the   one-­‐dimensional   C   NMR   spectrum   of   carbon   nanothreads   with   that   of   amorphous  polymers  and  carbon  materials.  Then  we  dis-­‐ cuss   the   various   moieties,   in   particular   sp2-­‐hybridized   carbons,   identified   by   spectral   editing   and   two-­‐ dimensional   NMR   and   estimate   their   proximities   in   the   nanothread   structures   based   on   13C   spin   diffusion.   Finally,   we   quantify   the   component   moieties   and   combine   them   in  an  illustrative  schematic  model.         13

Nanothreads   vs.   H-­‐rich   amorphous   carbon.   The   quantitative   solid-­‐state   13C   NMR   spectrum   of   the   13C-­‐ enriched   sample   (Figure   2a)   exhibits   two   bands,   a   large   one   near   40   ppm   characteristic   of   sp3-­‐hybridized   carbon,   and  a  smaller  and  broader  one  near  130  ppm  characteristic   of  sp2-­‐hybridized  carbon,  the  latter  accounting  for  28%  of   the  total  intensity.  The  spectral  pattern  resembles  the  one   previously   reported,1   which   also   exhibited   the   shoulders   near  145  and  25  ppm,  indicating  that  nanothread  synthesis  

ACS Paragon Plus Environment

3

Journal of the American Chemical Society 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 4 of 11

Since  the  starting  material,  benzene,  contains  only  CH   groups,   the   identification   of   CH2/CH3   groups   in   nanothreads   suggests   that   hydrogen   transfer   must   occur   during  the  high-­‐pressure  polymerization  of  benzenes.  The   extra   hydrogens   in   CH2/CH3   groups   likely   come   from   ar-­‐ omatic   substitution   and   sigmatropic   H-­‐shifts   (see   the   analysis  of  aromatic  linkers  below).      

Figure   4.   Sheared   DQ/SQ   (double   quantum/single   quan-­‐ 13 13 tum)   C-­‐ C  spectrum  showing  cross  peaks  of  directly  bond-­‐ ed   carbons.   Alkene-­‐   and   aromatic-­‐carbon   cross   peaks   are   labeled.  MAS  frequency:  14  kHz.  

is  reproducible.  CH-­‐selection  by  dipolar  DEPT,  see  Figure   2b,   confirms   that   most   carbon   atoms   in   the   material   are   bonded  to  one  hydrogen  atom,  as  expected.  While  the  CH   band  in  the  NMR  spectrum  of  nanothreads  is  broad  com-­‐ pared   to   that   of   a   specific   site   in   an   amorphous   polymer   (see   Figure   2b),   the   spread   of   frequencies   is   far   less   pro-­‐ nounced   than   in   the   shell   of   nanodiamond24   or   in   amor-­‐ phous   carbons.19,40   Hydrogen-­‐rich   tetrahedral   amorphous   carbon   (ta-­‐C:H)40   or   diamond-­‐like   amorphous   carbon   films,19   specifically   those   made   from   cyclohexane   chemi-­‐ cally  vapor  deposited  with  dc  plasma  assistance,40  are  the   closest   analogues   to   carbon   nanothreads   but,   like   all   amorphous   carbon,   show   much   wider   spectral   bands   (~30   ppm   width)   than   carbon   nanothreads   (~12   ppm).   This   confirms   the   higher   degree   of   order   in   carbon   nanothreads,   consistent   with   the   diffraction   peaks1   that   prove  a  crystal-­‐like  superstructure.     Methyl   or   methylene   groups.   A   small   shoulder   is   consistently  observed1  around  27  ppm  in  the  13C  spectrum,   see  Figure  2a.  While  the  chemical  shift  might  suggest  CH3   groups,   its   disappearance   after   dipolar   dephasing,   see   Figure   2d,   shows   that   this   signal   cannot   be   assigned   to   rotating  CH3  groups,  which  would  remain  at  >57%  inten-­‐ sity.   The   minimal   remaining   signal   imposes   an   upper   lim-­‐ it  of  0.3%  on  the  concentration  of  rotating  methyl  groups.   Spectral   editing   by   three-­‐spin   coherence   selection,32   see   Figure  2c,  shows  that  the  shoulder  near  27  ppm  is  at  least   partially   from   CH2   or   immobile   CH3   groups.   This   is   con-­‐ sistent   with   IR   bands   of   CH2   or   CH3  groups   observed   be-­‐ tween   1350   and   1500   cm-­‐1   and   near   2900   cm–1   (Figure   S1).   Assignment   to   CH2   is   possible   since   chemical   shift   esti-­‐ mations   for   methylenes   in   a   carbon   nanothread   environ-­‐ ment,  see  Figure  S2,  yield  chemical  shifts  between  15  and   30   ppm.   The   CH2   groups   appear   to   be   incorporated   into   an   alkyl   environment,   given   that   their   short-­‐range   cross   peaks   are   observed   to   alkyl   CH   at   42   ppm,   see   Figure   3.   The   methylenes   or   immobilized   methyl   groups   together   account  for  ≤4%  of  all  C.  

Multiple   sites   in   degree-­‐6   nanothreads.  The  broad-­‐ ening  of  the  40-­‐ppm  nanothread  signal  is  inhomogeneous   according   to   the   elongated   diagonal   ridge   in   the   short-­‐ time  2D  exchange  spectrum  in  Figure  3a.  This  means  that   many  CH  sites  with  different  (though  unresolved)  chemi-­‐ cal   shifts   coexist   in   the   material.     The   sharp   peak   at   129   ppm   in   Figure   2   confirms   that   this   is   not   a   problem   of   poor  magnetic-­‐field  homogeneity  (‘shim’)  or  susceptibility   effects.   Furthermore,   broadening   due   to   unpaired   elec-­‐ trons  can  be  excluded  in  carbon  nanothreads  by  analyzing   the   T1H   and   T1C   relaxation.   Figure   S3   shows   that   both   re-­‐ laxation   processes   are   relatively   slow   and   have   exponen-­‐ tial   character,   unlike   the   characteristically   fast   relaxation24  when  unpaired  electrons  are  significant.   Off-­‐diagonal  intensity  connecting   different  frequencies   within   the   band   near   40   ppm   is   observed   in   a   2D   13C -­‐13C   spectrum   with   10   ms   of   spin   exchange,   see   Figure   3b   vs.   3a.   This   indicates   the   proximity   of,   and   in   fact   direct   bonding  between,  sp3-­‐CH  sites  with  significantly  different   chemical   shifts.   In   other   words,   the   nanothreads   have   multiple  (>6)  chemically  inequivalent  sites,  unlike  single-­‐ site   models   such   as   polytwistane   (143652)   or   tube   (3,0)   (123456).     A   mixture   of   multiple   single-­‐site   degree-­‐6   nanothreads,   while   consistent   with   broadening   in   1D   NMR  spectra  and  in  pair-­‐distribution  functions,1  does  not   account   for   the   2D   NMR   pattern   since   it   would   produce   only   peaks   on   the   diagonal.   The   multi-­‐site   degree-­‐6   car-­‐ bon   nanothreads   recently   enumerated5   can   account   for   the   quasi-­‐continuous   distribution   of   13C   chemical   shifts   observed,   with   different   thread   structures   appearing   in   distinct   columns   and/or   as   changes   in   thread   structure   along   a   given   column.   Quantum-­‐chemical   calculations   within   density   functional   theory   can   further   identify   nanothread  structures  consistent  with  the  observed  posi-­‐ tions  and  widths  of  the   13C  NMR  signals  and  indicate  that   a   disordered   degree-­‐6   thread   as   well   as   certain   degree-­‐4   and   -­‐6   threads   that   represent   natural   termination   points   of   polymer   formation   are   the   most   plausible   candidate   structures.41     sp2-­‐carbons.   The   smaller   peak   in   the   13C   NMR   spec-­‐ trum   of   Figure   2a,   near   130   ppm,   is   characteristic   of   sp2-­‐ hybridized   carbon,   which   is   found   in   both   degree-­‐4   nanothreads   and   degree-­‐2   polymers.     It   is   therefore   im-­‐ portant   to   differentiate   between   sp2-­‐carbon   in   degree-­‐4   nanothreads   and   in   other   structures   or   impurities.   We   will   systematically   show   that   there   are   at   least   three   dif-­‐ ferent   types   of   sp2-­‐hybridized   carbon   present,   one   of   which  is  associated  with  degree-­‐4  nanothreads.    Further-­‐ more,  we  will  demonstrate  that  most  of  this  sp2  carbon  is   linked  by  one  or  several  covalent  bonds  to  the  sp3-­‐carbon   nanothreads   and   thus   there   is   no   minority   phase   of   aro-­‐ matic  sp2-­‐carbon.  

ACS Paragon Plus Environment

4

Page 5 of 11 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

Journal of the American Chemical Society

 Figure  2d  shows  a   13C  NMR  spectrum  detected  after   1H   decoupling  has  been  gated  off  for  68  µs.  The  dipolar  fields   of   1H   spins   dephase   the   magnetization   of   immobile   CHn   groups,   leaving   only   the   peaks   of   carbons   not   bonded   to   H   and   of   mobile   segments,   which   are   not   present   in   any   of   enumerated   nanothread   structures   and   are   minority   constituents   of   the   sample,   as   evidenced   by   their   much   smaller  areas  relative  to  the  spectrum  of  all  carbons.  

remains   near   130   ppm   in   the   dipolar-­‐dephased   spectrum   (Figure  2d).  Therefore,  the  (130  ppm,  40  ppm)  cross  peak   cannot   be   assigned   to   an   alkyl-­‐linked   aromatic   ring   be-­‐ cause  an  aromatic  carbon  bonded  to  an  alkyl  site  cannot   be  bonded  to  H  (see  the  structures  shown  near  the  cross   peaks  in  Figure   4).   Note   that   alkyl-­‐linked  nonprotonated   aromatic   carbons   are   instead   observed   at   145   ppm   (see   below).  The   1H  chemical  shift  of  5.5  ppm   associated  with   the   130-­‐ppm   carbon   in   a   1H-­‐13C   spectrum   (Figure   S4)   is   distinct  from  aromatic   1H  at  >  6.2  ppm  and  thus  confirms   the  assignment  of  the  130-­‐ppm  C-­‐H  resonance  to  alkenes.   The   1H  chemical  shift  below  6  ppm  also  indicates  that  the   double   bonds   are   not   conjugated,   consistent   with   expec-­‐ tations  for  degree-­‐4  nanothreads  and  some  degree-­‐2  pol-­‐ ymers  (Figure  1).  

2

Figure   5.   Dynamics   of   magnetization   transfer   from   sp -­‐ 3 hybridized   carbon   at   132   ppm   (mostly   alkene)   to   sp -­‐ hybridized   C   at   their   rotational   resonance   with   and   without   1 3 H  decoupling,  and  among  sp -­‐hybridized  C   for  reference,  as   obtained  from  a  series  of  2D  exchange  spectra,  see  Figure  S6.   3 (a)  Intensity  ratio  of  sp -­‐C  exchange  peaks  to  alkene  diagonal   signals,   as   a   function   of   spin-­‐exchange   time.   Filled   symbols   1 (thick  solid  line):  with   H  decoupling;  open  symbols  (dashed   line):  without  decoupling  during  spin  exchange.  (b)  Decrease   3 1 of  sp -­‐hybridized  C  diagonal  peak;   H  decoupling  slows  down   3 3 the   spin   exchange   and   constrains   the   sp -­‐sp   exchange   rate   used  in  the  simulations  in  a).  The  fit  curves  in  a)  compare  the   dynamics   predicted   in   degree-­‐4   polymer  (upper   curves)   and   two   different   cyclohexadienes   in   an   alkyl   matrix   (lower   curves  in  a).  

Alkenes.   Cross   peaks   between   =CH   resonating   at   130   ppm  and  sp3–CH  at  40  ppm  in  the  DQ/SQ  (double  quan-­‐ tum/single   quantum)   spectrum   (Figure   4)   can   be   unam-­‐ biguously  attributed  to  alkenes,  which  are  present  in  both   degree-­‐2  polymers  and  degree-­‐4  nanothreads.4  The  bond-­‐ ing  of  both  carbons  to  hydrogen  is  clearly  shown  by  dipo-­‐ lar   dephasing,  as   no  significant   signal  of  nonprotonated  C  

13

13

Figure   6.   (a)   C-­‐ C   exchange   spectrum   after   dipolar   dephasing   at   14   kHz   MAS,   with   total   sideband   suppression   during   detection.   Spin-­‐exchange   time:   50   ms   (including   25   ms   DARR).   Vertical   cross   sections   at   145   ppm   and   129   ppm   1 13 are  shown  on  the  right.  (b)   H-­‐ C  HetCor  spectrum  with  0.5-­‐ ms   cross   polarization   and   after   dipolar   dephasing,   showing   proximity   of   multiple   alkyl   CH   groups   to   the   average   aro-­‐ matic  ring.  

Degree-­‐4   nanothreads   vs.   degree-­‐2   polymers.  Next,   we   seek   to   distinguish   between   the   cyclohexadiene   rings   of  degree-­‐2  polymers  and   the   less   concentrated   alkenes   in   degree-­‐4   nanothreads.   Alkene   CH   linked   to   sp3-­‐ hybridized  CH  occurs  in  both  degree-­‐2  polymers  and  de-­‐ gree-­‐4   nanothreads   (Figure   1).   However,   the   sp3:sp2   car-­‐ bon   ratio   is   four-­‐fold   different,   namely   2:4   in   degree-­‐2   polymers   and   4:2   in   degree-­‐4   nanothreads.   The   ratio   of   protons   in   sp3-­‐   and   sp2-­‐hybridized   CH   groups   near   the   alkene  carbons  can  be  determined  in   1H-­‐13C  HetCor  NMR   with  a  short   1H  spin  diffusion  time  of  0.2  ms  (Figure  S5).  

ACS Paragon Plus Environment

5

Journal of the American Chemical Society 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

The   cross   section   at   the   alkene   13C   chemical   shift   shows   the  4:2  ratio  characteristic  of  degree-­‐4  nanothreads.  Simi-­‐ larly,  the  sp3-­‐:sp2-­‐  carbon  ratio  can  be  probed  through  2D   13 C-­‐13C  exchange  experiments  (Figures  5  and  S6).  The  fast   increase  of  the  sp3-­‐:sp2-­‐  signal  ratio  past  1:2  to  beyond  1.0   in  Figure  5a  rules  out  the  presence  of  degree-­‐2  polymers.   Thus,  if  there  are  any  degree-­‐2  polymers  formed  along  the   pathway   to   nanothreads   (possibly   by   “para   polymeriza-­‐ tion”  of  a  diradical)4  they  do  not  survive  in  the  final  reac-­‐ tion  products.  Degree-­‐2  polymers  were  calculated  to  have   the  highest  energies  among  structures  with  degree  of  sat-­‐ uration   from   0   to   6,4   which   might   be   why   they   were   not   observed  in  the  product  sample.     Degree-­‐4   nanothreads   vs.   dispersed   cyclohexadi-­‐ ene.  The  analysis  so  far  ruled  out  a  sequence  of  multiple   cyclohexadiene   rings   as   found   in   degree-­‐2   polymer.   A   careful   analysis   of   the   13C-­‐13C   spin   exchange   data   shown   in   Figure   5   makes   it   possible   to   also   rule   out   individual   cy-­‐ clohexadiene  rings  linked  to  sp3-­‐carbon  neighbors,  which   is   more   challenging   since   the   eventual   sp3:sp2   intensity   ratio   may   be   close   to   the   4:2   value   of   degree-­‐4   nanothreads.   By   optimizing   the   spin   dynamics,   the   ex-­‐ change   processes   for   alkene   and   diene   structures   can   be   made   distinct,   see   Figure   5a.   Adjusting   the   spinning   fre-­‐ quency   to   rotational   resonance   between   sp2-­‐   and   sp3-­‐ hybridized   C   speeds   up   exchange   from   sp2-­‐   to   sp3-­‐ hybridized   carbons.   The   rotational   resonance   condition   is   fulfilled   when   the   spinning   frequency   is   equal   to   the   chemical-­‐shift   frequency   difference   between   the   coupled   sites   of   interest.42   For   the   alkene–sp3-­‐CH   spin   pairs,   the   spinning  frequency  was  therefore  chosen  as  (131  ppm  –  41   ppm)  0.1  kHz/ppm  =  9  kHz.       Furthermore,   1H  decoupling  greatly  slows  down  sp3-­‐sp3   exchange   out   of   the   cyclohexadiene   rings,   as   proved   by   the  four-­‐fold  slow-­‐down  of  the  decrease  of  the  sp3  diago-­‐ nal   peak   documented   in   Figure   5b   (see   also   Figure   S6b).   This   makes   the   4:2   ratio   of   sp3:sp2-­‐hybridized   carbons   apparent.   The   different   exchange   rates   are   indicated   by   arrows   of   different   lengths   in   the   structural   cartoons   in   Figure   5a.     Simulations   of   the   spin   exchange   process   un-­‐ der   these   conditions   for   the   different   models   (see   the   Supporting   Information   for   details)   indeed   provide   dis-­‐ tinctly  different  curves  (Figure  5a)  with  a  sloping  plateau   extrapolating   to   0.5   =   2:4   for   1,4-­‐cyclohexadiene,   and   even   lower   for   1,3-­‐cyclohexadiene,   where   the   =CH   units   are   clustered   (i.e.,   conjugated   double   bonds).   The   experi-­‐ mental  data  obtained  from  diagonal  and  cross  peaks  in  a   series  of  2D  spectra  (Figure  S6  and  5),  are  in  much  better   agreement   with   the   presence   of   linked   degree-­‐4   nanothreads  rather  than  dispersed  cyclohexadiene.      Aromatic   rings   linked   to   alkyls.  A  shoulder  consist-­‐ ently  observed  at  145  ppm1  is  resolved  as  a  peak  after  dipo-­‐ lar  dephasing,  see  Figure  2d.  Its  intensity  decreases  by  less   than  15%  after  68  μs  without   1H  decoupling,  which  shows   that   this   signal   is   due   to   a   carbon   not   bonded   to   H.   The   chemical  shift  is  typical  of  a  CH-­‐substituted   aromatic   car-­‐ bon.   An   alternative   assignment   would   be   a   substituted   alkene;   the   double   bonds   would   have   to   be   conjugated   since   the   145-­‐ppm   carbon   is   bonded   to   at   least   two   sp2-­‐ hybridized  CH  sites,  according  to  13C-­‐13C  NMR  (Figure  6a).  

Page 6 of 11

This   assignment   can   be   excluded   since   IR   spectroscopy   shows   prominent   signals   of   substituted   benzene   rings   between  700  and  850  cm-­‐1  (Figure  7  and  Table  S1).  The   1H   chemical   shift   of   6.8   ppm   in   the   1H-­‐13C   spectrum   after  

Figure   7.  IR  spectrum  of  carbon  nanothreads  (black  curve),   with   empirical  band  assignment   to  vibrations  of   mono-­‐,   di-­‐ substituted   and   1,3,5-­‐tri-­‐substituted   benzene   rings.   Wave-­‐ 44 numbers  of  the  ranges  shown  are  listed  in  Table  S1.  

dipolar   dephasing   in   Figure   6b   is   also   more   consistent   with  aromatic  than  alkene  C-­‐H.   The  2D  spectrum  and  its  cross  section  shown  in  Figure   6b   also   shows   that   the   aromatic   carbon   is   near   alkyl   1H,   indicating   that   the   145-­‐ppm   aromatic   carbon   is   mostly   bonded  to  alkyl  CH,  which  is  confirmed  by  a  cross  peak  to   alkyl   CH   in   the   DQ/SQ   spectrum   (Figure   4,   upper   left   corner).  There  is  no  (145  ppm,  145  ppm)  peak  indicative  of   a   linkage   between   two   such   benzene   rings.   Further   cor-­‐ roboration   of   the   assignment   of   the   145   ppm   peak   to   a   linkage   to   alkyl   CH   comes   from   the   chemical   shift   itself:   Two   aromatic   carbons   linked   by   a   single   bond   resonate   at   <  140  ppm  in  solid-­‐state  NMR  spectra.43     It  is  important  to  establish  whether  the  alkyl-­‐linked  ar-­‐ omatic   rings   are   just   pendant   groups   (monosubstituted   benzenes)   or   mostly   linkers   connecting   alkyl   segments   (multiply   substituted   benzenes);   this   can   be   probed   by   two-­‐dimensional   NMR.   In   a   pendant   ring,   the   nonproto-­‐ nated:protonated   carbon   ratio   would   be   1:5,   in   a   linker   2:4.   A   slice   through   a   13C-­‐13C   exchange   spectrum   (after   dipolar  dephasing)  at  145  ppm  shows  about  two  aromatic   C-­‐H  per  nonprotonated  aromatic  C    (Figure  6a).  This  in-­‐ dicates  that,  per  ring,  there  are  on  average  two  linkages  to   alkyl   carbon   sites.   The   deconvolution   of   the   sp2-­‐ hybridized   carbon   signal   (Figure   8d)   confirms   this   ratio   independently.  Magnetization  quickly  transfers  to  several   sp3-­‐hybridized   CH   groups,   according   to   the   strong   cross   peaks   with   alkyl-­‐CH   in   both   13C-­‐13C   and   1H-­‐13C   spectra   (Figure  6).     The   nonprotonated:protonated   aromatic   carbon   ratio   of   2:4   ratio   seen   in   NMR   data   might   suggest   exclusively   para-­‐substitution,   but   the   IR   spectrum   reveals   a   more   complex   picture.44   It   also   shows   signals   indicative   of   mono-­‐substitution,   ortho   or   meta-­‐disubstitution,   and  

ACS Paragon Plus Environment

6

Page 7 of 11 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

Journal of the American Chemical Society

1,3,5-­‐trisubstitution.   To   balance   the   small   (1:5)   nonproto-­‐ nated:protonated   aromatic   carbon   ratio   of   monosubsti-­‐ tuted   rings,   a   similar   amount   of   1,3,5-­‐trisubstituted   aro-­‐ matic  rings  needs  to  be  invoked.       The  identification  of  alkyl-­‐substituted  benzenes  (as  well   as   CH2   discussed   above)   in   the   sample   also   indicates   hy-­‐ drogen   transfer   occurring   in   the   high-­‐pressure   transfor-­‐ mation  of  benzene  to  nanothreads.  Possible  H-­‐transfer  via   aromatic   substitution   and   sigmatropic   H-­‐shift   mecha-­‐ nisms   are   shown   in   Figure   S7.   The   amounts   of   CH2   and   nonprotonated   aromatic   carbons   match   (3-­‐4%,   see   be-­‐ low).    

anisotropy   shows   that   the   mobility   of   the   ring   must   be   anisotropic.   The   strong   asymmetry   of   the   sideband   pat-­‐ tern   is   characteristic   of   fast   rotational   jumps   around   the   ring’s  six-­‐fold  axis.  This  motion  by  itself  in  solid  benzene   at  223  K  results  in  Δσ  ≈  180  ppm;46  the  additional  33%  nar-­‐ rowing   must   be   ascribed   to   large-­‐amplitude   wobbling   of   the   rotation   axis.   Unreacted   benzenes   might   be   embed-­‐ ded   sideways   between   threads   (particularly   where   on-­‐ thread  defects  may  form  small  voids  in  thread  packing)  or   between  the  ends  of  two  co-­‐axial  threads.  The  anisotropi-­‐ cally   mobile   benzene   is   quite   intimately   associated   with   the   alkyl-­‐rich   matrix,   showing   13C-­‐13C   cross   peaks   within   50  ms,  see  Figure  6a,  and  full  equilibration  of  magnetiza-­‐ tion   from   benzene   with   that   of   other   carbon   sites  within   1   s  (see  Figure  9  below).     In  the  DQ/SQ  spectrum  of  Figure  4,  a  diagonal  peak  is   observed   faintly   near   (135,   135)   ppm.   It   must   be   assigned   to   directly   linked   aromatic   rings.   Indeed,   extractable   naphthalene   and   diphenyl   have   been   observed   by   gas   chromatography  –  mass  spectrometry,47  and  out-­‐of-­‐plane   deformation   of   the   benzene   rings   at   lower   pressures   was   speculated  to  be  related  to  such  oligomerizations.48  Their   contribution  can  be  assessed  based  on  the  nonprotonated   aromatic   signal   between   135   and   140   ppm.   It   is   relatively   small,  accounting  for  <  3%  of  all  carbons.                                                                                                  

13

Figure   8.     Selective   C   NMR   spectra   of   13C-­‐enriched   nanothreads.   (a)   Spectrum   of   nonprotonated   carbons,   after   dipolar   dephasing.  The   band   at   40  ppm   is   a   residual   artifact   from  the  intense  CH  peak.  (b)  Spectrum  of  alkene  and  near-­‐ 1 13 1 by   alkyl   CH,   from   H-­‐ C   spectrum   at   a   H   chemical   shift   of   5.5   ppm.   (c)   Spectrum   of   aromatic   carbons   and   nearby   alkyl   CH.   (d)   Quantification   of   alkene   and   aromatic   components   13 in   a  quantitative   C   NMR  spectrum  by   deconvolution   of  the   2 sp -­‐hybridized  carbon  signal  based  on  the  experimental  spec-­‐ tra   in   (a)   –   (c).   Dashed   line:   Weighted   superposition   of   the   component  spectra.  

Free   benzene   and   linked   aromatics.   After   dipolar   dephasing,  a  sharp  peak  partially  remains  at  129  ppm  (see   Figure   8a),   indicating   a   nonprotonated   carbon,   or   a   mo-­‐ bile   molecule   or   segment.   Nonprotonated   carbon   signals   are  observed  at  this  chemical  shift  only  in  large  fused  ring   systems   or   with   COOH   substitution,45   both   of   which   are   absent   from   our   sample.   Furthermore,   1H-­‐13C   correlation   shows  that  this  sharp  13C  peak  is  associated  with  a  similar-­‐ ly  narrow  1H  signal,  see  Figure  6b.  Since  the  chemical  shift   agrees   to   within   1   ppm   with   the   reported   value   for   neat   benzene,   we   assign   it   to   mobile   benzene.   A   pronounced   “self-­‐correlation”   peak   is   observed   at   129   ppm   on   the   di-­‐ agonal  in  the  sheared  DQ/SQ  spectrum  of  Figure  4,  which   proves   that   at   least   two   carbons   of   the   same   chemical   shift  are  bonded  to  each  other,  as  is  true  for  benzene.  This   benzene  signal  accounts  for  5%  of  all  carbon  (see  below).   The   benzene   carbons   have   a   large   chemical-­‐shift   ani-­‐ sotropy  (CSA)  of  Δσ  ≈  120  ppm,  according  to  fast  dephas-­‐ ing   under   CSA   recoupling   (Figure   S8)   and   the   sideband   pattern   observed   under   3-­‐kHz   MAS   (Figure   S9).   This   shift  

13

13

Figure   9.   (a)   0.5-­‐s   C-­‐ C   spin-­‐exchange   spectrum   at   rota-­‐ tional   resonance   (with   total   suppression   of   spinning   side-­‐ bands   before   detection)   and   (b)   horizontal   slices   from   the   3 2 sp -­‐  (solid  green  line)  and  the  sp -­‐  (dashed  red  line)  carbon   signal  maximum.  The  spectra  have  been  scaled  to  match  the   2 1 sp -­‐carbon  bands.  30-­‐kHz   H  decoupling  was  applied  during   the  0.5-­‐s  mixing  time.  (c)  Corresponding  slices  from  a  spec-­‐ trum  with  1-­‐s  mixing  time  (including  0.2  s  of  30-­‐kHz  decou-­‐ pling).   The   difference   spectrum   in   c)   (thick   blue   lines)   can   3 2 be  attributed  to  sp -­‐C  not  in