Charge Trapping and Exciton Dynamics in Large ... - ACS Publications

Feb 23, 2016 - and L. Michael Hayden*,‡. †. U.S. Naval Research Laboratory, Washington, D.C. 20375, United States. ‡. Department of Physics, UMB...
0 downloads 0 Views 6MB Size
Subscriber access provided by GAZI UNIV

Article 2

Charge Trapping and Exciton Dynamics in Large-area CVD Grown MoS Paul D. Cunningham, Kathleen Michelle McCreary, Aubrey T Hanbicki, Marc Currie, Berend T. Jonker, and L. Michael Hayden J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.6b00647 • Publication Date (Web): 23 Feb 2016 Downloaded from http://pubs.acs.org on February 27, 2016

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 free 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 accessible to all readers and 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.

The Journal of Physical Chemistry C 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 15

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 Journal of Physical Chemistry

Charge   Trapping   and   Exciton   Dynamics   in   Large-­‐area   CVD   Grown   MoS2     Paul  D.  Cunningham,1*  Kathleen  M.  McCreary,1  Aubrey  T.  Hanbicki,1  Marc  Currie,1  Berend  T.   Jonker,1  and  L.  Michael  Hayden2,†     1U.S.  Naval  Research  Laboratory,  Washington,  DC  20375   2Department  of  Physics,  UMBC,  Baltimore,  MD  21250     *[email protected]   †[email protected]  

  Abstract     There   is   keen   interest   in   monolayer   transition   metal   dichalcogenide   films   for   a   variety   of   optoelectronic   applications   due   to   their   direct   band-­‐gap   and   fast   carrier   dynamics.   However,   the   mechanisms   dominating   their   carrier   dynamics   are   poorly   understood.   By   combining   time-­‐resolved   terahertz   (THz)   spectroscopy   and   transient   absorption,   we   are   able   to   shed   light   on   the   optoelectronic   properties   of   large   area   CVD   grown   mono-­‐   and   multi-­‐layer   MoS2   films   and   determine   the   origins   of   the   characteristic   two-­‐component  excited  state  dynamics.  The  photo-­‐induced  conductivity  shows  that  charge   carriers,   and   not   excitons,   are   responsible   for   the   sub-­‐picosecond   dynamics.   Identical   dynamics   resulting   from   sub-­‐optical   gap   excitation   suggest   that   charge   carriers   are   rapidly   trapped   by   mid-­‐gap   states   within   600   fs.   This   process   complicates   the   excited   state   spectrum   with   rapid   changes   in   line-­‐width   broadening   in   addition   to   a   red-­‐shift   due   to   band  gap  renormalization  and  simple  state-­‐filling  effects.  These  dynamics  are  insensitive  to   film   thickness,   temperature,   or   choice   of   substrate,   which   suggests   that   carrier   trapping   occurs   at   surface   defects   or   grain   boundaries.   The   slow   dynamics   are   associated   with   exciton   recombination,   and   lengthen   from   50   ps   for   monolayer   films   to   150   ps   for   multi-­‐ layer   films   indicating   that   surface   recombination   dominates   their   lifetime.   We   see   no   signatures   of   trions   in   these   MoS2   films.   Our   results   imply   that   CVD   grown   films   of   MoS2   hold  potential  for  high-­‐speed  optoelectronics  and  provide  an  explanation  for  the  absence  of   trions  in  some  CVD  grown  MoS2  films.       Introduction   Two-­‐dimensional   (2D)   transition   metal   dichalcogenides   (TMD)   are   of   interest   for   optoelectronic   applications   due   to   their   direct   band-­‐gaps   and   strong   photoluminescence   exhibited   in   single-­‐layers.1   Similar   to   graphene,   strong   in-­‐plane   bonds   and   weak   van   der   Waals  forces  between  atomic  layers  allow  single  layer  TMDs  to  be  isolated  and  potentially   reassembled  monolayer  by  monolayer  into  heterostructures.2  Of  the  TMDs,  MoS2  was  the   first  in  which  the  now  familiar  transition  from  bulk  indirect  band-­‐gap  to  monolayer  direct   band-­‐gap  was  observed.3-­‐4  Novel  electronic  and  photonic  devices  have  been  demonstrated   using   MoS2   including   high   mobility   GHz-­‐frequency   field-­‐effect   transistors   with   large   on/off   ratios,5   fast   response   phototransistors   that   can   be   sensitized   into   the   infrared,6   and   extremely  sensitive  chemical  sensors.7  Spin  control  of  the  K  and  K’  valley  populations  has  

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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

also   led   to   intense   interest   in   MoS2   for   valleytronics   applications.8   Though   much   of   the   work  has  focused  on  exfoliated  flakes,  other  deposition  methods  including  chemical  vapor   deposition   (CVD)9   have   been   pursued   to   produce   large   area   monolayer   films.10-­‐12   Understanding   the   photo-­‐physical   properties   of   large   area   CVD   grown   MoS2   films   is   crucial   to  advancing  many  of  the  aforementioned  optoelectronic  applications.   Previous   photo-­‐physical   studies   of   both   monolayer   and   multilayer   MoS2   have   observed   complex   multiexponential   dynamics.13-­‐14   The   interpretation   of   these   results   is   heavily   debated   and   the   underlying   physics   has   been   assigned   to   a   number   of   different   phenomena.     Fast,   picosecond   scale   dynamics   have   been   assigned   to   the   electron-­‐hole   (exciton)   radiative   lifetime,15   charge   carrier   trapping,16   charge   carrier   cooling,17   trion   formation,18  and  coupling  between  the  A  and  B  excitons.14  Slower  decays  >10  picoseconds   have   been   assigned   to   interband   recombination,13   exciton-­‐phonon   scattering,19   and   the   lifetime  of  charged  excitons  (e.g.  trions).20  Clearly,  the  excited  state  photo-­‐physics  in  MoS2   is   not   yet   well   understood   and   requires   greater   clarification   for   a   consensus   to   be   reached.   Of  these  prior  reports,  thorough  studies  of  the  excited  state  photo-­‐physics  using  transient   absorption  (TA)  techniques  have  focused  on  multilayer  MoS2,14,  17-­‐18  leaving  technologically   important  monolayers  relatively  unexplored.   Nearly   all   of   the   previous   photo-­‐physical   studies   have   relied   on   traditional   all-­‐ optical  methods,  such  as  photoluminescence  and  optical  pump-­‐probe  techniques,  to  probe   the   excited   state   dynamics.   These   methods   cannot   directly   probe   charge   conduction.   However,   time-­‐resolved   THz   spectroscopy   (TRTS)   provides   unique   insight   into   excited   state   photo-­‐physics,   with   its   ability   to   discern   between   charge   carriers   and   excitons   with   sub-­‐picosecond   resolution.21   Very   recently,   two   TRTS   studies   of   CVD   grown   monolayer   MoS2  reported  dramatically  different  observations.  Lui  et  al.  report  a  ~30  ps  lifetime  photo-­‐ induced  decrease  in  conductivity  due  to  charged  exciton  (e.g.  trion)  formation.20  Docherty   et  al.  report  a  sub-­‐picosecond  lifetime  photo-­‐induced  increase  in  conductivity  from  short-­‐ lived  charge  carriers.22  The  source  of  these  short-­‐lived  dynamics,  which  are  similar  to  what   has  been  previously  observed  with  more  traditional  optical  spectroscopy,14-­‐15  remains  the   subject  of  debate.     Here  we  utilize  both  ultrafast  TA  spectroscopy  and  TRTS  to  determine  the  origins  of   the   fast   charge   carrier   dynamics   in   mono-­‐   and   multi-­‐layer   films   of   MoS2   on   sapphire,   fused   silica,  and  cyclic  olefin  co-­‐polymer  (TOPAS®)  substrates.  TA  reveals  two  decay  components   associated  with  different  photo-­‐physical  processes.  The  fast  component  is  accompanied  by   changes   in   line   width,   associated   with   a   rapid   reduction   in   excited   states,   and   a   small   spectral   shift   due   to   band-­‐gap   renormalization.   Using   TRTS,   we   observe   a   photo-­‐induced   increase   in   conductivity   with   identical   dynamics,   which   confirms   that   mobile   charge   carriers,  and  not  excitons,  produce  the  fast  decay.  These  charge  carriers  are  trapped  within   approximately   600   fs.   The   independence   of   this   decay   on   substrate   or   film   thickness   suggests   that   carriers   are   trapped   at   either   surface   defects   or   grain   boundaries.   The   observation   of   photocurrent   for   photon   energies   below   the   optical   gap   points   to   the   presence   of   mid-­‐gap   states   associated   with   these   traps.   Weak   temperature,   fluence,   and   photon-­‐energy   dependence   suggest   that   electron-­‐phonon   scattering   and   Auger   mechanisms   do   not   play   a   major   role   in   the   observed   dynamics.   In   addition   to   the   sub-­‐ picosecond  trapping,  a  50  ps  decay  is  observed  in  the  monolayer  TA  dynamics.  This  slower   decay  component  is  associated  with  exciton  recombination  and  lengthens  to  150  ps  in  the   multi-­‐layer  film  due  to  the  transition  from  direct  to  indirect  band-­‐gap.  We  do  not  observe  

ACS Paragon Plus Environment

Page 2 of 15

Page 3 of 15

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 Journal of Physical Chemistry

signs  of  trion  formation,  which  we  attribute  to  atmospheric  adsorbates  that  neutralize  the   intrinsic  n-­‐type  behavior.     Methods     Sample  Preparation.  MoS2  films  were  prepared  by  chemical  vapor  deposition  (CVD)   in   a   2   inch   diameter   tube   furnace.12   Solid   sources   of   MoCl5   and   sulfur   serve   as   the   precursors   for   the   MoS2   growth   on   c-­‐plane   sapphire   or   Si/SiO2   (275   nm).10     The   growth   substrate  and  MoCl5  precursor  are  placed  on  a  quartz  platform  at  the  center  of  the  furnace.   To   achieve   monolayer   films   of   MoS2,   2   mg   of   MoCl5   is   used,   with   sulfur   positioned   at   the   upstream   end   of   the   furnace.   Multi-­‐layer   (3-­‐4   layer)   films   are   achieved   by   increasing   the   MoCl5  amount  to  4  mg  (6  mg).  The  pressure  is  maintained  at  2  Torr  under  a  50  sccm  flow  of   Argon  while  the  temperature  is  increased  at  a  rate  of  15°  C/min.  Upon  reaching  850  °C,  the   growth  temperature  is  maintained  for  10  minutes,  and  the  sample  is  then  allowed  to  cool   undisturbed  under  continuous  Ar  flow.       To  investigate  dynamics  of  MoS2  on  other  substrates,  the  films  were  removed  from   the  Si/SiO2  growth  substrate  and  transferred  to  the  desired  wafer.  In  the  first  step  of  the   transfer  process,  PMMA  resist  (950  A4)  is  spun  onto  the  MoS2  and  cured  on  a  hotplate  at   100°C  for  10  mins.  The  sample  is  then  submerged  in  buffered  HF  to  etch  the  oxide  layer,   freeing  the  MoS2  from  the  growth  substrate.  Once  released,  the  film  is  moved  to  DI  water,   where   it   floats   on   the   surface.   The   MoS2   is   then   lifted   from   the   water   with   the   desired   substrate.  Adhesion  is  improved  by  spinning  at  2000  rpm  and  baking  at  100°  C.  The  PMMA   is  then  removed  using  acetone  and  isopropanol.   Transient   Absorption.   A   transient   absorption   spectrometer   was   used   to   measure   the   excited  state  dynamics  in  the  MoS2  films.  The  system  is  based  on  a  1  W  150  fs  Ti:Sapphire   amplifier  (Clark  MXR,  CPA).  The  output  is  split,  and  ~300  mW  are  used  to  pump  a  visible   OPA  (Clark  MXR,  NOPA)  to  produce  tunable  excitation  pulses  with  typical  incident  fluences   ca.  1013  photons/cm2.  A  small  amount  of  power  is  focused  into  a  sapphire  plate  to  generate   the   white   light   continuum   probe.   The   white   light   pulses   are   then   sent   into   a   scanning   monochromator  to  record  the  excited  state  spectra.  The  recorded  spectra  is  corrected  for   dispersion   by   projecting   the   raw   data   onto   a   Sellmeier   fit   to   the   wavelength   dependence   of   the   pump-­‐probe   overlap   to   produce   a   spectrum   where   all   wavelengths   have   experienced   the  same  delay.     Time-­Resolved   THz   Spectroscopy.   An   optical-­‐pump   THz-­‐probe   spectrometer   was   used  to  measure  the  photo-­‐induced  change  in  conductivity  of  the  MoS2  films.  The  system  is   based   on   an   8   W   30   fs   Ti:Sapphire   amplifier   (Coherent,   Legend   Elite   Duo).   The   output   is   split  with  3.2  W  used  to  pump  an  OPA  (Coherent,  OperA  Solo)  to  produce  tunable  excitation   pulses;  the  remainder  is  used  to  generate  and  detect  THz  pulses.  Broadband  THz  pulses  are   generated   in   a   two-­‐color   air-­‐plasma   by   focusing   1.5   W   through   a   BBO   crystal   and   mixing   the   fundamental   and   second   harmonic   in   a   laser-­‐induced   air-­‐plasma   filament.     The   THz   electric   field   transients   are   detected   in   a   300   µm   thick   GaP   crystal   via   free-­‐space   electro-­‐ optic   sampling,   providing   continuous   bandwidth   from   0.1-­‐8   THz.   Details   of   how   the   photoconductivity   is   measured   via   TRTS   and   extracted   from   the   raw   data   are   given   elsewhere.23,24  The  pump  beam  is  focused  to  a  ~2  mm  diameter,  which  is  larger  than  the   750  µm  THz  beam  spot,  and  leads  to  typical  incident  fluences  of  ca.  1015  photons/cm2.     Results  

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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

  MoS2  mono-­‐  and  multi-­‐layer  films  were  prepared  by  CVD  growth.12  The  number  of   layers  was  confirmed  by  Raman  spectroscopy,  atomic  force  microscopy  (AFM),  and  steady-­‐ state   photoluminescence,   as   illustrated   in   Figure   1.   AFM   profiles   in   Figure   1b   show   step   heights  of  ~1  nm  and  ~3  nm  for  the  monolayer  and  multi-­‐layer  films,  which  is  consistent   with  films  containing  1  and  4  layers  of  MoS2  respectively.  Further,  it  is  also  well  known  that   the  transition  from  indirect  to  direct  band-­‐gap  occurs  as  MoS2  transitions  from  two-­‐layer  to   monolayer.4   Strong   emission   is   observed   only   for   the   monolayer,   Figure   1c,   confirming   the   direct   band-­‐gap.   Importantly,   the   monolayer   emission   does   not   show   signs   of   trion   emission,  previously  reported  near  1.85  eV.25  Finally,  it  is  known  that  the  out-­‐of-­‐plane  A1G   Raman   mode   blue-­‐shifts   and   the   in-­‐plane   E12G   mode   red-­‐shifts   as   the   numbers   of   layers   increases.26  We  observe  splitting  between  the  E12G  and  the  A1G  peaks  of  ~  20  cm-­‐1  for  the   monolayer  and  25  cm-­‐1  for  the  multi-­‐layer  MoS2  on  sapphire,  Figure  1d,  consistent  with  the   expected  increase  in  peak  separation.    

  Figure   1.   a)   Optical   micrograph   of   monolayer   MoS2   showing   the   large   uniform   area   that   results  from  CVD  growth.  A  purposeful  scratch  in  the  corner  shows  the  contrast  between   film   and   SiO2   substrate.   The   inset   shows   the   monolayer   MoS2   film   after   it   is   transferred   onto   a   1”   x   1”   fused   silica   substrate.   b)   AFM   image   of   monolayer   MoS2   on   SiO2.   The   inset   shows  the  step-­‐height  profile.  c)  Photoluminescence  spectra  of  mono-­‐  (red)  and  multi-­‐layer   (black)   MoS2.   d)   Raman   spectra   of   mono-­‐   (red)   and   multi-­‐layer   (black)   MoS2.   The   modes   associated  with  the  film  and  sapphire  substrate  are  labeled.    

ACS Paragon Plus Environment

Page 4 of 15

Page 5 of 15

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 Journal of Physical Chemistry

  The   absorption   spectrum   in   Figure   2a   shows   the   features   typically   observed   for   monolayer   MoS2   labeled   “A”,   “B”   and   “C”   at   1.89   eV   (655   nm),   2.04   eV   (608   nm),   2.87   eV   (432   nm),   respectively.   Spin-­‐orbit   splitting   of   the   valence   band   leads   to   the   “A”   and   “B”   features,   both   corresponding   to   excitonic   transitions   at   the   K-­‐point   of   the   Brillouin   zone.   The  “C”  feature  corresponds  to  transitions  involving  bands  located  between  the  K-­‐  and  Γ-­‐ points,  and  has  been  attributed  to  van  Hove  singularities  in  the  density  of  states27  as  well  as   to  the  band-­‐gap  of  monolayer  MoS2.28    

  Figure   2.   (a)   Absorbance   of   monolayer   MoS2   on   fused   silica.   (b)   Transient   absorption   spectra  measured  as  a  function  of  delay  for  monolayer  MoS2  photo-­‐excited  at  532  nm.  (c)   Excited  state  dynamics  probed  at  490  nm  (blue),  610  nm  (green),  660  nm  (red),  and  690   nm   (dark   red).   (d)   Fit   results   to   the   TA   spectra   showing   the   temporal   evolution   of   the   A   (red)  and  B  (blue)  absorption  feature  amplitude,  line-­‐width,  and  center  wavelength.  Solid   lines  are  exponential  fits.       TA  measurements  yield  the  time-­‐evolution  of  the  excited  state  spectrum,  providing   detailed   information   of   the   excited   state   dynamics.   The   excited   state   spectrum   of   monolayer  MoS2  show  that  the  change  in  transmission,  ∆T/T0,  alternates  between  positive   and  negative,  Figure  2b.  This  type  of  response  has  been  previously  observed  for  multi-­‐layer   MoS214,   17-­‐18  and  was  also  recently  observed  in  monolayer  WS2.29  It  is  clear  that  such  a  TA   spectrum   cannot   be   described   by   simple   band-­‐filling,   which   would   produce   ground   state  

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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

bleaching  that  mirrors  the  absorption  spectrum.  Instead,  the  absorption  features  must  shift   or  broaden  to  produce  the  observed  response.     It   is   tempting   to   interpret   the   alternating   positive   and   negative   changes   in   transmission   as   belonging   to   different   optical   transitions,   i.e.   ground-­‐state   bleaching   and   excited  state  absorption  (ESA).  This  explanation  is  inadequate  to  describe  our  observations   for   the   following   reasons:   (1)   while   the   A-­‐   trion   has   been   observed   near   670   nm   in   absorption  and  luminescence  studies  of  doped  MoS2  and  at  low  temperature,25,  30  no  similar   peaks  have  been  reported  for  the  B  and  C  features  to  explain  the  potential  ESA  near  630  nm   and   490   nm;   (2)   there   are   no   other   known   transitions   that   correspond   to   the   energies   associated   with   the   negative   ∆T/T0   features;   (3)   interpreting   each   TA   peak   as   an   optical   transition  requires  an  explanation  for  the  unexpected  apparent  higher  state  filling  of  the  B   state   as   compared   to   the   A   state.   We   instead   observe   agreement   between   the   decay   dynamics   measured   for   each   of   these   features,   Figure   2c,   which   implies   they   have   common   origins.   Our   interpretation   of   the   TA   spectra   invokes   a   spectral-­‐shift   and   line-­‐width   broadening,   which   is   more   consistent   with   not   only   our   data   but   with   observations   reported  in  the  literature  as  well.     We   simulate   the   TA   spectra   by   fitting   the   absorption   spectrum   and   then   calculate   the   changes   to   the   A   and   B   absorption   features,   by   way   of   amplitude   changes   (i.e.   bleaching),   line-­‐width   broadening   (ΔΓ/Γ   ),   and   spectral   shift   (Δhν),   that   will   produce   the   observed   change   in   transmission.   We   fit   the   absorption   spectra   with   two   Gaussians,   one   each  for  the  A  and  B  excitonic  features.  We  then  compute  the  absorption  of  the  excited  state   as   2 2 −( λ −( λ 0,A +Δλ A ))2 /(ΓA R A )2 + Be −( λ −( λ 0,B +Δλ B )) /(ΓB R B ) ,       α exc ( λ ) = Ae 1   where  Ri  is  the  ratio  of  the  unexcited  and  excited  MoS2  ground  state  absorption  line-­‐widths,   Γi    is  the  ground  state  absorption  line-­‐width,  and  ∆λ  is  the  change  in  the  center  wavelength   λ0,i.   We   can   then   calculate   the   measured   photo-­‐induced   change   in   the   transmission   spectrum  a€s     ΔT 10 −α exc ( λ ) = −α 0 ( λ ) −1,       2   T 10 where  α0(λ)  is  the  ground  state  absorption  spectrum.     The   temporal   evolution   of   the   fit   parameters   for   monolayer   MoS2   are   shown   in   Figure  2d.  After  photoexcitation,  we  observe  a  sub-­‐picosecond  decay  in  amplitude  and  line-­‐ € width  broadening.  The  line-­‐width  broadening  dominates  the  TA  spectra  and  can  be  shown   to   match   the   temporal   evolution   of   the   spectral   zeroth   moment   i.e.   the   area   under   each   excitonic   feature.   Line-­‐width   broadening   also   accounts   for   the   higher   amplitude   of   the   B   feature   in   the   TA   spectrum.   An   increase   in   B   excitonic   line-­‐width   leads   to   increased   absorption   that   overlaps   with   the   A   exciton   feature,   giving   the   appearance   that   the   A   exciton  is  less  populated,  see  Supporting  Information  for  more  details.  Over  this  same  time-­‐ scale,   a   small   red-­‐shift   grows   in.   This   is   counterintuitive,   as   the   TA   spectrum   appears   to   blue-­‐shift.   However,   it   is   readily   apparent   that   a   blue-­‐shifted   absorption   spectrum   would   produce  a  positive  ∆T/T0  feature  near  690  nm,  while  instead  we  observe  a  negative  ∆T/T0   feature.  Further,  a  shifting  absorption  spectrum  will  not  create  a  shifting  TA  spectrum  but   instead  creates  an  oscillatory  TA  spectrum  that  changes  magnitude  with  the  size  of  the  shift   in  absorption  wavelength,  see  Supporting  Information  for  details.  Instead,  changes  in  line-­‐ width   and   center   wavelength   combine   to   produce   an   apparent   blue   shift   in   the   TA  

ACS Paragon Plus Environment

Page 6 of 15

Page 7 of 15

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 Journal of Physical Chemistry

spectrum.   The   measured   red-­‐shift   is   too   small   to   be   accounted   for   by   trion   formation,   which   results   in   a   ~40   meV   red-­‐shift   of   luminescence.25   After   5   ps,   all   fit   parameters   decay   at  the  same  rate.   A   similar   analysis   of   transient   absorption   in   multi-­‐layer   MoS2   presented   in   the   literature   concluded   that   the   observed   spectra   were   produced   by   coherent   interactions   between   the   A   and   B   excitons   based   primarily   on   the   identical   A   and   B   exciton   dynamics.14   Although   our   spectral   bandwidth   is   insufficient   to   quantify   changes   to   the   C   feature,   the   negative   ∆T/T0  feature  near  490  nm  suggests  that  the  C  feature  undergoes  similar  spectral   changes  as  the  A  and  B  features.  The  matching  dynamics  measured  for  490  nm,  610  nm  and   690   nm   also   imply   a   link   between   changes   in   the   A,   B   and   C   features.     Very   recently,   all   three   features   were   observed   in   the   excited   state   spectrum   after   resonant   excitation   of   the   A   exciton.31   These   observations   appear   to   be   inconsistent   with   the   interpretation   that   coupling   between   the   A   and   B   excitons   gives   rise   to   the   observed   TA   spectrum   in   monolayer  MoS2.     Examining   the   dynamics   at   individual   wavelengths   associated   with   features   in   the   TA   spectrum   provides   information   about   the   time-­‐scale   of   the   different   photo-­‐physical   processes   in   monolayer   MoS2.   Measurements   reveal   that   dynamics   take   place   on   two   distinct  times  scales,  Figure  2c.  The  B  exciton  dynamics,  probed  at  610  nm,  clearly  show  the   two   decay   components,   which   are   present   in   both   mono-­‐   and   multi-­‐layer   MoS2,   see   Supporting  Information.  The  fast  component  decays  on  a  sub-­‐picosecond  time  scale  with  a   time   constant   of   670   ±   110   fs,   which   agrees   with   the   results   obtained   by   fitting   the   TA   spectra.   Because   of   the   B-­‐feature   line-­‐width   broadening,   only   this   fast   component   can   be   resolved   when   probing   the   A   exciton   dynamics   at   660   nm.   No   oscillations   in   the   A   or   B   exciton   dynamics   are   observed,   which   would   be   expected   if   coherent   interactions   were   present.  Dynamics  probed  far  from  the  excitonic  features  do  not  show  this  fast  component   because  they  are  relatively  unaffected  by  the  rapid  changes  in  line-­‐width.  Instead,  only  the   slow  decay  component  is  observed  when  the  dynamics  are  probed  at  490  nm  or  690  nm.   The   slow   decay   component   depends   on   film   thickness,   with   50   ±   20   ps   and   150   ±   40   ps   time  constants  for  the  mono-­‐  and  multi-­‐layer  films  respectively,  Figure  3a.    

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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   3.   (a)   Excited   state   dynamics   of   multi-­‐   (black)   and   monolayer   (red)   MoS2   photo-­‐ excited   at   532   nm   and   probed   at   490   nm.   (b)   Photoconductivity   dynamics   of   multi-­‐   (black)   and   monolayer   (red)   MoS2.   (c)   Comparison   between   the   photoconductivity   dynamics   measured  with  TRTS  and  the  excited  state  dynamics  probed  at  610  nm  resulting  from  480   nm   excitation   of   monolayer   MoS2.   (d)   The   frequency-­‐dependent   photoconductivity   of   monolayer  MoS2  on  a  TOPAS®  substrate  measured  at  15  K.  The  lines  are  fits  to  the  Drude-­‐ Smith  model.     The  photo-­‐induced  conductivity  dynamics  of  the  mono-­‐  and  multi-­‐layer  MoS2  films   were   measured   using   TRTS.   The   conductivity   dynamics   are   similar   for   both   the   monolayer   and   multilayer   films,   Figure   3b.   An   ultrafast   increase   in   conductivity   was   observed   that   decays   with   a   time   constant   of   660   ±   150   fs.   This   is   similar   to   the   dynamics   recently   reported  by  Docherty  et  al.22  On  the  other  hand,  this  is  in  sharp  contrast  to  trion  dominated   dynamics   reported   by   Lui   et   al.,20   where   a   photo-­‐induced   decrease   in   conductivity   was   observed.   Though   a   similar   fast   component   was   also   present   in   those   dynamics,   it   was   attributed   to   electron-­‐phonon   scattering   after   laser   heating   of   the   lattice.   Such   a   mechanism   cannot   be   operative   here   because   the   photo-­‐induced   conductivity   increases.   Instead,   we   observe   that   the   photoconductivity   dynamics   precisely   match   the   sub-­‐ picosecond  decay  observed  with  TA,  Figure  3c,  suggesting  a  common  origin.     The   frequency-­‐dependent   complex   conductivity   in   Figure   3d   shows   non-­‐Drude   behavior   characterized   by   a   significant   real   component   and   a   small   negative   imaginary   component.   This   is   characteristic   of   charge   carrier   localization   and   scattering   due   to   nanoscale   disorder.32   The   non-­‐zero   real   conductivity   is   a   clear   departure   from   the   THz   signature   of   excitons.23   The   observed   conductivity   is   also   quite   different   from   the   Drude-­‐ like   response   observed   in   bulk   MoS2   using   TRTS.33   One   explanation   for   the   observed   localization   is   carrier   backscattering   at   defects   or   grain   boundaries   in   our   polycrystalline  

ACS Paragon Plus Environment

Page 8 of 15

Page 9 of 15

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 Journal of Physical Chemistry

film.   Anderson   localization,   where   random   disorder   causes   destructive   interference   among   carrier  scattering  events,  leads  a  non-­‐Drude  conductivity  that  can  instead  be  described  by   the   phenomenological   Drude-­‐Smith   model.34   A   best   fit   line   using   the   Drude-­‐Smith   model   yields  a  carrier  scattering  time  of  23  fs,  a  plasma  frequency  of  251  THz,  and  a  persistence  of   velocity  of  -­‐0.65.  A  persistence  of  velocity  between  0  and   -­‐1  indicates  partial  localization.   The  DC  mobility  is  estimated  to  be  23  cm2/Vs,  assuming  an  effective  mass  of  0.61me,  which   is   in   agreement   with   thin-­‐film   transistor   measurements.35   Alternatively,   surface   plasmon   resonances   or   intraband   transitions,   in   conjunction   with   a   Drude   metal   response,   could   describe   the   observed   conductivity.   However,   surface   plasmon   resonances   are   expected   only   at   visible   wavelengths   for   MoS2   nanoparticles,   making   such   a   model   improbable.   Also,   there   is   no   expected   intraband   transition   in   MoS2   between   2-­‐4   THz:   the   valance   band   splitting  could  only  produce  transitions  near  36  THz  (150  meV)  and  we  see  no  resonance   associated   with   the   0.75   THz   (3meV)36   conduction   band   splitting.   The   photo-­‐induced   conductivity  appears  to  be  temperature  independent,  with  matching  dynamics  at  300K  and   15K.  We  do  not  observe  a  resonance  near  the  trion  binding  energy  of  20  meV  (5  THz),  even   at  low  temperature,  which  is  expected  if  trions  are  present.        

  Figure   4.   (a)   Incident   fluence-­‐dependence   of   the   photoconductivity   dynamics   for   480   nm   excitation.  (b)  Incident  fluence-­‐dependence  of  the  transient  absorption  spectrum  measured   at   a   delay   of   5ps   for   387   nm   excitation.   Fluences   are   given   in   photons/cm2.   (c)   The  

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 10 of 15

photoconductivity  dynamics  measured  for  400  nm,  480  nm,  and  800  nm  excitation.  (d)  The   excited   state   dynamics   measured   at   a   probe   wavelength   of   490   nm   for   excitation   wavelengths  of  387  nm,  480  nm,  532  nm,  615  nm,  660  nm,  and  775  nm.       The   dependence   of   excited   state   and   photoconductivity   dynamics   on   excitation   fluence   and   photon   energy   often   aids   in   understanding   underlying   photo-­‐physics.   The   photo-­‐induced   conductivity   decay   time   constant   was   insensitive   to   the   incident   pump   fluence  over  the  range  2  –  8  x1015  photons/cm2,  Figure  4a.  It  is  important  to  note  that  at   480   nm   monolayer   MoS2   has   ~91%   transmission   so   that   the   absorbed   fluence   remains   below   the   atomic   sheet   density.   The   TA   spectrum,   however,   appears   to   red-­‐shift   with   incident   fluence   over   the   range   0.1   –   1.0   x1014   photons/cm2,   similar   to   what   has   been   observed  in  multi-­‐layer  films,  Figure  4b.18  Increased  line-­‐width  broadening  and  red-­‐shift  of   the  absorption  spectrum  combine  to  produce  the  apparent  red-­‐shift  of  the  TA  spectra,  see   Supporting   Information   for   details.   This   complicates   any   examination   of   the   fluence   dependence   of   individual   spectral   features   probed   at   a   specific   wavelength.   It   is   worth   noting  that  the  absorbed  fluence  in  all  cases  is  within  an  order  of  magnitude  of  the  reported   Mott  density  (1  x  1013  cm-­‐2)  in  monolayer  MoS2.5  The  photo-­‐induced  conductivity  dynamics   were   also   insensitive   to   changes   in   photon   energy,   showing   a   similar   response   for   sub-­‐ optical  gap  excitation  at  800  nm  (1.55  eV),  Figure  4c.  This  requires  a  mechanism  for  charge   carrier   generation   at   energies   below   the   ~1.9   eV   optical   gap.   Similarly,   the   excited   state   dynamics   are   found   to   be   insensitive   to   excitation   photon   energy,   see   Supporting   Information  for  details.  Probing  at  490  nm,  we  see  that  the  slow  component  of  the  excited   state   dynamics   is   independent   of   excitation   energy,   even   for   775   nm   (1.6   eV)   excitation,   Figure  4d.  This  points  to  the  existence  of  mid-­‐gap  states  in  monolayer  MoS2.     Discussion     The   photoconductivity   dynamics   show   that   charge   carriers,   and   not   excitons,   are   responsible   for   the   sub-­‐picosecond   TA   dynamics.   Charge   carrier   trapping   causes   rapid   changes  in  amplitude  and  line-­‐width  broadening  of  both  the  A  and  B  absorption  features.   The  excitation  wavelength  and  fluence  independence  does  not  support  the  involvement  of   carrier-­‐phonon   scattering   or   Auger   mechanisms.     The   red-­‐shift   of   the   absorption   spectrum   that   grows   in   is   due   to   band-­‐gap   renormalization   associated   with   the   Coulombic   interactions  within  the  2D  film.  Here,  changes  to  both  the  electronic  bandgap  and  exciton   binding   energy   only   partially   compensate,   giving   rise   to   the   observed   red-­‐shift.31   One   explanation   for   charge   generation   in   a   two-­‐dimensional   monolayer   film   with   strong   Coulombic   interactions   is   that   the   excitation   fluences   necessitated   by   these   experiments   approach   the   Mott   density,   above   which   band-­‐like   transport   is   expected.5,   37   Similar   dynamics   observed   in   multi-­‐layer   films,   where   the   Mott   density   is   not   known,   may   suggest   that  an  intrinsic  charge  separation  pathway  could  also  exist.   We  propose  that  sub  band-­‐gap  charge  carrier  generation  occurs  by  photo-­‐excitation   of   mid-­‐gap   trap   states,   likely   located   at   surface   defects   and   grain   boundaries.   Grain   boundaries38   and   vacancies39   have   been   theorized   to   produce   mid-­‐gap   states   in   monolayer   MoS2.   Our   CVD   growth   from   a   MoCl5   precursor   results   in   lower   emission   intensities   and   smaller  grains  sizes  as  compared  to  exfoliated  flakes  or  growth  from  MoO3  precursor,  see   Supporting  Information  for  details,  which  also  suggests  a  correlation  between  trap  density   and  grain  boundary  density.  The  observed  trapping  dynamics  require  defect  densities  to  be  

ACS Paragon Plus Environment

Page 11 of 15

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 Journal of Physical Chemistry

near  the  injected  carrier  density  of  ca.  1013  cm-­‐2,  which  is  in  good  agreement  with  recent   theoretical   and   experimental   estimates   of   the   defect   density   associated   with   grain   boundaries40   and   vacancies.37,   41   This   is   consistent   with   the   observed   wavelength   independent   charge   carrier   generation   and   excited   state   dynamics.   Because   the   observed   dynamics   are   unchanged   for   fused   silica   and   TOPAS®   polymer   substrates,   we   do   not   suspect   that   the   trapping   occurs   at   the   substrate-­‐film   interface.   We   observe   no   appreciable   change   in   luminescence   intensity   upon   successive   transfers,   suggesting   that   the   defect   density  does  not  increase  during  the  transfer  process  and  that  the  defects  are  intrinsic  to   the  film.  Therefore  we  propose  that  charge  carriers  are  partially  localized  by  backscattering   effects  either  due  to  the  potential  barrier  at  the  grain  boundary  or  the  potential  well  of  a   lattice   defect,   leading   to   rapid   trapping   within   600   fs.   Interestingly,   these   trapping   mechanisms  dominate  the  optoelectronic  properties  of  both  mono-­‐  and  multi-­‐layer  MoS2.     We  assign  the  slow  TA  decay  to  tightly  bound  excitons.  This  is  consistent  with  their   absence   from   the   THz   conductivity.   Excitons   exhibit   zero   real   conductivity   at   THz   frequencies  because  the  THz  electric  field  cannot  drive  the  motion  of  neutral  particles  and   so   there   will   be   no   energy   dissipation.   Instead,   excitons   interact   with   THz   via   their   polarizability,   leading   to   a   capacitive   response   through   the   imaginary   conductivity.23   Therefore,  tightly  bound  excitons  that  lack  sufficient  polarizability  will  not  be  detected  at   THz   frequencies.   These   excitons   can   partially   occupy   and   be   localized   by   defects   states,   which  has  led  to  previous  reports  of  sub  band-­‐gap  emission  in  monolayer  MoS2,42  or  may   remain  after  traps  are  filled  by  mobile  charge  carriers.  The  strong  Coulombic  interactions   that   give   rise   to   bound   excitons   also   give   rise   to   band-­‐gap   renormalization,   so   that   the   observed  red-­‐shift  is  proportional  to  the  density  of  photo-­‐excited  excitons.    The  50  ps  decay   time  for  the  monolayer  is  consistent  with  previous  reports  of  the  luminescence  lifetime  in   MoS2.19   The   increase   in   exciton   lifetime   from   monolayer   to   multilayer   films   may   appear   consistent   with   the   transition   from   direct   to   indirect   band-­‐gap,   because   electron-­‐phonon   interaction   required   in   indirect   band-­‐gap   semiconductors   leads   to   longer   radiative   lifetimes.   However,   recent   measurements   of   the   thickness-­‐dependence   of   the   exciton   lifetime  in  MoS2  show  that  fast  surface  recombination  dominates  slow  bulk  recombination   in  few-­‐layer  MoS2.43  So  we  instead  assign  the  shorter  exciton  lifetime  in  monolayer  MoS2  to   the  surface  recombination  limited  exciton  lifetime.     We   see   no   signs   of   trions   in   the   MoS2   films.   We   do   not   observe   a   photo-­‐induced   decrease  in  the  conductivity  associated  with  trion  formation  nor  a  resonance  near  20  meV   (i.e.  5  THz)  associated  with  the  trion  binding  energy.  The  reason  for  the  absence  of  trions   may  be  related  to  the  grain  size  of  our  MoS2  films.  CVD-­‐growth  using  MoCl5  as  a  precursor   typically  results  in  polycrystalline  MoS2  films  with  small  grains,  ca.  10  nm,  as  compared  to   the  large  grains,  ca.  1  µm,  found  in  mechanically  exfoliated  flakes9-­‐10  or  by  using  MoO3  as  a   precursor,   see   Supporting   Information   for   details.   The   intrinsic   n-­‐type   doping   of   MoS2   is   neutralized  by  atmospheric  adsorbates,42,  44  e.g.  oxygen,  which  preferentially  attach  at  grain   boundaries,45   such   that   films   with   smaller   grains   are   less   likely   to   show   n-­‐type   behavior.   Without   the   intrinsic   n-­‐type   doping,   negative   trions   do   not   form.   Recently,   it   was   shown   that   trions   or   excitons   can   be   selectively   prepared   in   WS2   by   cleaning   the   surface   of   adsorbates,46   but   adsorbates   may   bind   more   strongly   to   MoS2.   Comparatively,   the   recent   study   of   trions   in   CVD   grown   MoS2   by   Lui   et   al.   involved   films   with   large   grains   formed   through   seeded   CVD-­‐growth,11,   20   where   the   intrinsic   n-­‐type   doping   was   maintained.   This   may  explain  the  recent  conflicting  reports  of  conductivity  dynamics  observed  via  TRTS  in  

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 15

CVD-­‐grown   monolayer   MoS2.   Direct   measurements   comparing   the   photoconductivity   of   monolayer  MoS2  films  with  different  grain  sizes  would  aid  in  confirming  this  supposition.     Conclusions     In   summary,   we   have   observed   an   ultrafast   photoconductivity   decay   that   matches   a   corresponding   absorption   line-­‐width   broadening,   red-­‐shift,   and   decay   of   ground   state   bleaching.   We   assign   these   dynamics   to   charge   carrier   trapping   due   to   mid-­‐gap   states   associated   with   grain   boundaries   or   surface   defects.   This   points   to   the   importance   of   growth   procedures   that   yield   high   quality   films   with   large   grain   size   for   optoelectronic   devices   based   on   monolayer   TMDs.   On   the   other   hand,   such   a   fast   conductivity   response   makes   TMDs   a   good   candidate   for   high-­‐speed   optoelectronics   including   photodetectors,   modulators,   and   THz   photoconductive   dipole   antennae.   We   also   observe   a   longer-­‐lived   excited   state   that   has   no   associated   conductivity   to   produce   THz   attenuation,   which   we   assign   to   the   decay   of   tightly   bound   excitons.   This   exciton   lifetime   depends   on   film   thickness   due   to   surface   recombination,   which   dominates   the   exciton   lifetime.   The   absence   of   trions   is   attributed   to   atmospheric   adsorbates   that   neutralize   the   intrinsic   n-­‐type   character.  Similar  observations  may  also  be  expected  for  the  neutral  excited  states  of  other   CVD-­‐grown  TMD  monolayer  films.         Supporting   Information   Available:   details   concerning   the   modeling   of   transient   absorption   spectra,   fluence   dependence   of   the   transient   absorption   spectra,   excitation   photon   energy   dependence   of   transient   absorption   spectra,   substrate   and   film   thickness   dependence,   temperature   dependence,   and   a   comparison   of   different   sample   preparation   techniques.  This  material  is  available  free  of  charge  via  the  Internet  at  http://pubs.acs.org.     Acknowledgement:  This  work  was  supported  by  core  programs  at  the  U.S.  Naval  Research   Laboratory  (NRL),  the  NRL  Nanoscience  Institute,  and  by  the  Air  Force  Office  of  Scientific   Research  under  contract  number  AOARD  14IOA018-­‐134141.  KMM  acknowledges  the   Nation  Research  Council  research  associates  program.       References:    

1.   Jariwala,  D.;  Sangwan,  V.  K.;  Lauhon,  L.  J.;  Marks,  T.  J.;  Hersam,  M.  C.,  Emerging  Device   Applications   for   Semiconducting   Two-­‐Dimensional   Transition   Metal   Dichalcogenides.   ACS  Nano  2014,  9,  1102-­‐1120.   2.   Geim,   A.   K.;   Grigorieva,   I.   V.,   Van   Der   Waals   Heterostructures.   Nature   2013,   499,   419-­‐ 425.   3.   Splendiani,  A.;  Sun,  L.;  Zhang,  Y.;  Li,  T.;  Kim,  J.;  Chim,  C.-­‐Y.;  Galli,  G.;  Wang,  F.,  Emerging   Photoluminescence  in  Monolayer  MoS2.  Nano  Lett.  2010,  10,  1271-­‐1275.   4.   Mak,  K.  F.;  Lee,  C.;  Hone,  J.;  Shan,  J.;  Heinz,  T.  F.,  Atomically  Thin  MoS2:  A  New  Direct-­‐Gap   Semiconductor.  Phys.  Rev.  Lett.  2010,  105,  136805.   5.   Radisavljevic,   B.;   Kis,   A.,   Mobility   Engineering   and   a   Metal-­‐Insulator   Transition   in   Monolayer  MoS2.  Nat.  Mater.  2013,  12,  815-­‐820.  

ACS Paragon Plus Environment

Page 13 of 15

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 Journal of Physical Chemistry

6.   Kufer,   D.;   Nikitskiy,   I.;   Lasanta,   T.;   Navickaite,   G.;   Koppens,   F.   H.   L.;   Konstantatos,   G.,   Hybrid  2D-­‐0D  MoS2-­‐Pbs  Quantum  Dot  Photodetectors.  Adv.  Mater.  2015,  27,  176-­‐180.   7.   Perkins,   F.   K.;   Friedman,   A.   L.;   Cobas,   E.;   Campbell,   P.   M.;   Jernigan,   G.   C.;   Jonker,   B.   T.,   Chemical  Vapor  Sensing  with  Monolayer  Mos2.  Nano  Lett.  2013,  13,  668-­‐673.   8.   Kioseoglou,   G.;   Hanbicki,   A.   T.;   Currie,   M.;   Friedman,   A.   L.;   Gunlycke,   D.;   Jonker,   B.   T.,   Valley  Polarization  and  Intervalley  Scattering  in  Monolayer  MoS2.  Appl.  Phys.  Lett.  2012,   101,  221907.   9.   Najmaei,   S.;   Liu,   Z.;   Zhou,   W.;   Zou,   X.;   Shi,   G.;   Lei,   S.;   Yakobson,   B.   I.;   Idrobo,   J.-­‐C.;   Ajayan,   P.   M.;   Lou,   J.,   Vapour   Phase   Growth   and   Grain   Boundary   Structure   of   Molybdenum   Disulphide  Atomic  Layers.  Nat.  Mater.  2013,  12,  754-­‐759.   10.  Yu,  Y.;  Li,  C.;  Liu,  Y.;  Su,  L.;  Zhang,  Y.;  Cao,  L.,  Controlled  Scalable  Synthesis  of  Uniform,   High-­‐Quality  Monolayer  and  Few-­‐Layer  MoS2  Films.  Sci.  Reports  2013,  3,  1866.   11.  Lee,  Y.-­‐H.;  Yu,  L.;  Wang,  H.;  Fang,  W.;  Ling,  X.;  Shi,  Y.;  Lin,  C.-­‐T.;  Huang,  J.-­‐K.;  Chang,  M.-­‐T.;   Chang,  C.-­‐S.,  et  al.,  Synthesis  and  Transfer  of  Single-­‐Layer  Transition  Metal  Disulfides  on   Diverse  Surfaces.  Nano  Lett.  2013,  13,  1852-­‐1857.   12.  McCreary,   K.   M.;   Hanbicki,   A.   T.;   Robinson,   J.   T.;   Cobas,   E.;   Culbertson,   J.   C.;   Friedman,   A.   L.;  Jernigan,  G.  C.;  Jonker,  B.  T.,  Large-­‐Area  Synthesis  of  Continuous  and  Uniform  MoS2   Monolayer  Films  on  Graphene.  Adv.  Func.  Mater.  2014,  24,  6449-­‐6454.   13.  Shi,   H.;   Yan,   R.;   Bertolazzi,   S.;   Brivio,   J.;   Gao,   B.;   Kis,   A.;   Jena,   D.;   Xing,   H.   G.;   Huang,   L.,   Exciton   Dynamics   in   Suspended   Monolayer   and   Few-­‐Layer   MoS2   2D   Crystals.   ACS   Nano   2013,  7,  1072-­‐1080.   14.  Sim,   S.;   Park,   J.;   Song,   J.-­‐G.;   In,   C.;   Lee,   Y.-­‐S.;   Kim,   H.;   Choi,   H.,   Exciton   Dynamics   in   Atomically  Thin  MoS2:  Interexcitonic  Interaction  and  Broadening  Kinetics.  Phys.  Rev.  B   2013,  88.   15.  Lagarde,  D.;  Bouet,  L.;  Marie,  X.;  Zhu,  C.  R.;  Liu,  B.  L.;  Amand,  T.;  Tan,  P.  H.;  Urbaszek,  B.,   Carrier   and   Polarization   Dynamics   in   Monolayer   Mos2.   Phys.   Rev.   Lett.   2014,   112,   047401.   16.  Wang,   H.;   Zhang,   C.;   Rana,   F.,   Ultrafast   Dynamics   of   Defect-­‐Assisted   Electron-­‐Hole   Recombination  in  Monolayer  MoS2.  Nano  Lett.  2015,  15,  339-­‐345.   17.  Nie,  Z.;  Long,  R.;  Teguh,  J.  S.;  Huang,  C.-­‐C.;  Yeow,  E.  K.  L.;  Hewak,  D.  W.;  Shen,  Z.;  Prezhdo,   O.  V.;  Loh,  Z.-­‐H.,  Ultrafast  Electron  and  Hole  Relaxation  Pathways  in  Few-­‐Layer  MoS2.  J.   Phys.  Chem.  C  2015,  119,  20698-­‐20708.   18.  Borzda,  T.;  Gademaier,  C.;  Vujicic,  N.;  Topolovsek,  P.;  Borovsak,  M.;  Mertelj,  T.;  Viola,  D.;   Manzoni,   C.;   Pogna,   E.   A.;   Brida,   D.,   et   al.,   Charge   Photogeneration   in   Few-­‐Layer   MoS2.   Adv.  Func.  Mater.  2015,  25,  3351-­‐3358.   19.  Korn,   T.;   Heydrich,   S.;   Hirmer,   M.;   Schmutzler,   J.;   Schüller,   C.,   Low-­‐Temperature   Photocarrier  Dynamics  in  Monolayer  MoS2.  Appl.  Phys.  Lett.  2011,  99,  102109.   20.  Lui,  C.  H.;  Frenzel,  A.  J.;  Pilon,  D.  V.;  Lee,  Y.-­‐H.;  Ling,  X.;  Akselrod,  G.  M.;  Kong,  J.;  Gedik,  N.,   Trion-­‐Induced   Negative   Photoconductivity   in   Monolayer   MoS2.   Phys.   Rev.   Lett.   2014,   113,  166801.   21.  Ulbricht,   R.;   Hendry,   E.;   Shan,   J.;   Heinz,   T.   F.;   Bonn,   M.,   Carrier   Dynamics   in   Semiconductors   Studied   with   Time-­‐Resolved   Terahertz   Spectroscopy.   Rev.   Mod.   Phys.   2011,  83,  543.   22.  Docherty,  C.  J.;  Parkinson,  P.;  Joyce,  H.  J.;  Chiu,  M.-­‐H.;  Chen,  C.-­‐H.;  Lee,  M.-­‐Y.;  Li,  L.-­‐J.;  Herz,   L.  M.;  Johnston,  M.  B.,  Ultrafast  Transient  Terahertz  Conductivity  of  Monolayer  MoS2  and   WSe2  Grown  by  Chemical  Vapor  Deposition.  ACS  Nano  2014,  8,  11147-­‐11153.  

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 14 of 15

23.  Cunningham,   P.   D.;   Hayden,   L.   M.;   Yip,   H.   L.;   Jen,   A.   K.-­‐Y.,   Charge   Carrier   Dynamics   in   Metallated   Polymers   Investigated   by   Optical-­‐Pump   Terahertz-­‐Probe   Spectroscopy.   J.   Phys.  Chem.  B  2009,  113,  15427-­‐15432.   24.  Knab,   J.   R.;   Lu,   X.;   Vallejo,   F.   A.;   Kumar,   G.;   Murphy,   T.   E.;   Hayden,   L.   M.,   Ultrafast   Carrier   Dynamics  and  Optical  Properties  of  Nanoporous  Silicon  at  Terahertz  Frequencies.  Opt.   Mater.  Ex.  2014,  4,  300-­‐307.   25.  Mak,   K.   F.;   He,   K.;   Changgu,   L.;   Lee,   G.   H.;   Hone,   J.;   Heinz,   T.   F.;   Shan,   J.,   Tightly   Bound   Trions  in  Monolayer  MoS2.  Nat.  Mater.  2013,  12,  207-­‐211.   26.  Lee,  C.;  Yan,  H.;  Brus,  L.  E.;  Heinz,  T.  F.;  Hone,  J.;  Ryu,  S.,  Anomalous  Lattice  Vibrations   of   Single-­‐  and  Few-­‐Layer  MoS2.  ACS  Nano  2010,  4,  2695-­‐2700.   27.  Britnell,   L.;   Ribeiro,   R.   M.;   Eckmann,   A.;   Jalil,   R.;   Belle,   B.   D.;   Mishchenko,   A.;   Kim,   Y.-­‐J.;   Gorbachev,  R.  V.;  Georgiou,  T.;  Morozov,  S.  V.,  et  al.,  Strong  Light-­‐Matter  Interactions  in   Heterostructures  of  Atomically  Thin  Films.  Science  2013,  340,  1311-­‐1314.   28.  Klots,   A.   R.;   Newaz,   A.   K.   M.;   Wang,   B.;   Prasai,   D.;   Krzyzanowska,   H.;   Lin,   J.;   Caudel,   D.;   Ghimire,   N.   J.;   Yan,   J.;   Ivanov,   B.   L.,   et   al.,   Probing   Excitonic   States   in   Suspended   Two-­‐ Dimensional  Semiconductors  by  Photocurrent  Spectroscopy.  Sci.  Reports  2014,  4,  6608.   29.  He,  J.;  Kumar,  N.;  Bellus,  M.  Z.;  Chiu,  H.-­‐Y.;  He,  D.;  Wang,  Y.;  Zhao,  H.,  Electron  Transfer   and   Coupling   in   Graphene-­‐Tungsten   Disulfide   Van   Der   Waals   Heterostructures.   Nat.   Commun.  2014,  5,  5622.   30.  Sercombe,  D.;  Schwarz,  S.;  Pozo-­‐Zamudio,  O.  D.;  Liu,  F.;  Robinson,  B.  J.;  Chekhovich,  E.  A.;   Tartakovskii,   I.   I.;   Kolosov,   O.;   Tartakovskii,   A.   I.,   Optical   Investigation   of   the   Natural   Electron   Doping   in   Thin   MoS2   Films   Deposited   on   Dielectric   Substrates.   Sci.   Reports   2013,  3,  3489.   31.  Pogna,  E.  A.  A.;  Marsili,  M.;  Fazio,  D.  D.;  Conte,  S.  D.;  Manzoni,  C.;  Sangalli,  D.;  Yoon,  D.;   Lombardo,  A.;  Ferrari,  A.  C.;  Marini,  A.,  et  al.,  Photo-­‐Induced  Bandgap  Renormalization   Governs  the  Ultrafast  Response  of  Single-­‐Layer  MoS2.  ACS  Nano  2016,  10,  1182-­‐1188.   32.  Jensen,   S.   A.;   Tielrooij,   K.-­‐J.;   Hendry,   E.;   Bonn,   M.;   Rychetsky,   I.;   Nemec,   H.,   Terahertz   Depolarization  Effects  in  Colloidal  TiO2   Films   Reveal   Particle   Morphology.  J.   Phys.   Chem.   C  2013,  118,  1191-­‐1197.   33.  Strait,   J.   H.;   Nene,   P.;   Rana,   F.,   High   Intrinsic   Mobility   and   UIltrafast   Carrier   Dynamics   in   Multilaye  Metal-­‐Dichalcogenide  MoS2.  Phys.  Rev.  B  2014,  90,  245402.   34.  Smith,  N.  V.,  Classical  Generalization  of  the  Drude  Formula  for  the  Optical  Conductivity.   Phys.  Rev.  B  2001,  64,  155106.   35.  Bao,   W.;   Cai,   X.;   Kim,   D.   S.;   Sridhara,   K.;   Fuhrer,   M.,   High   Mobility   Ambipolar   MoS2   Field-­‐ Effect  Transistors:  Substrate  and  Dielectric  Effects.  Appl.  Phys.  Lett.  2013,  102,  042104.   36.  Kormanyos,   A.;   Zolyomi,   V.;   Drummond,   N.   D.;   Rakyta,   P.;   Bukard,   G.;   Fal'ko,   V.   I.,   Monolayer  MoS2:  Trigonal  Warping,  the  Γ  Valley,  and  Spin-­‐Orbit  Coupling  Effects.  Phys.   Rev.  B  2013,  88,  045416.   37.  Qiu,  H.;  Xu,  T.;  Wang,  Z.;  Ren,  W.;  Nan,  H.;  Ni,  Z.;  Chen,  Q.;  S.,  Y.;  Miao,  F.;  Song,  F.,  et  al.,   Hopping   Transport   through   Defect-­‐Induced   Localized   States   in   Molybdenum   Disulphide.  Nat.  Commun.  2013,  4,  2642.   38.  van   der   Zande,   A.   M.;   Huang,   P.   Y.;   Chenet,   D.   A.;   Berkelbach,   T.   C.;   You,   Y.;   Lee,   G.-­‐H.;   Heinz,   T.   F.;   Reichman,   D.   R.;   Muller,   D.   A.;   Hone,   J.   C.,   Grains   and   Grain   Boundaries   in   Highly  Crystalline  Monolayer  Molybdenum  Disulphide.  Nat.  Mater.  2013,  12,  554-­‐561.  

ACS Paragon Plus Environment

Page 15 of 15

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 Journal of Physical Chemistry

39.  Zhou,  W.;  Zou,  X.;  Najmaei,  S.;  Liu,  Z.;  Shi,  Y.;  Kong,  J.;  Lou,  J.;  Ajayan,  P.  M.;  Yakobson,  B.   I.;   Idrobo,   J.-­‐C.,   Intrinsic   Structural   Defects   in   Monolayer   Molybdenum   Disulfide.   Nano   Lett.  2013,  13,  2615-­‐2622.   40.  Yu,   Z.   G.;   Zhang,   Y.-­‐W.;   Yakobson,   B.   I.,   An   Anomalous   Formation   Pathway   for   Dislocation-­‐Sulfur   Vacancy   Complexes   in   Polycrystalline   Monolayer   MoS2.   Nano   Lett.   2015,  15,  6855-­‐6861.   41.  Hong,  J.;  Hu,  z.;  Probert,  M.;  Li,  K.;  Lv,  D.;  Yang,  X.;  Gu,  L.;  Mao,  N.;  Fend,  Q.;  Xie,  L.,  et  al.,   Exploring  Atomic  Defects  in  Molybdenum  Disulphide  Monolayers.  Nat.  Commun.  2015,   6,  6293.   42.  Tongay,  S.;  Suh,  J.;  Ataca,  C.;  Fan,  W.;  Luce,  A.;  Kang,  J.  S.;  Liu,  J.;  Ko,  C.;  Raghunathanan,   R.;   Zhou,   J.,   et   al.,   Defects   Activated   Photoluminescence   in   Two-­‐Dimensional   Semiconductors:   Interplay   between   Bound,   Charged,   and   Free   Excitons.   Sci.   Reports   2013,  3,  2657.   43.  Wang,   H.;   Zhang,   C.;   Rana,   F.,   Surface   Recombination   Limited   Lifetimes   of   Photogenerated   Carriers   in   Few-­‐Layer   Transition   Metal   Dichalcogenide   MoS2.   Nano   Lett.  2015,  15,  8204-­‐8210.   44.  Tongay,  S.;  Zhou,  J.;  Ataca,  C.;  Liu,  J.;  Kang,  J.  S.;  Matthews,  T.  S.;  You,  L.;  Li,  J.;  Grossman,  J.   C.;   Wu,   J.,   Broad-­‐Range   Modulation   of   Light   Emission   in   Two-­‐Dimensional   Semiconductors  by  Molecular  Physisorption  Gating.  Nano  Lett.  2013,  13,  2831-­‐2836.   45.  Nan,   H.;   Wang,   Z.;   Wang,   W.;   Liang,   Z.;   Lu,   Y.;   Chen,   Q.;   He,   D.;   Tan,   P.;   Miao,   F.;   Wang,   X.,   et  al.,  Strong  Photoluminescence  Enhancement  of  MoS2  through  Defect  Engineering  and   Oxygen  Bonding.  ACS  Nano  2014,  8,  5738-­‐5745.   46.  Currie,   M.;   Hanbicki,   A.   T.;   Kioseoglou,   G.;   Jonker,   B.   T.,   Optical   Control   of   Charged   Exciton  States  in  Tungsten  Disulfide.  Appl.  Phys.  Lett.  2015,  106,  201907.            

TOC  Figure:  

 

 

ACS Paragon Plus Environment