• Aucun résultat trouvé

CHAPTER  V     ZnO  nanowire  arrays      

N/A
N/A
Protected

Academic year: 2022

Partager "CHAPTER  V     ZnO  nanowire  arrays      "

Copied!
28
0
0

Texte intégral

(1)

                           

CHAPTER  V     ZnO  nanowire  arrays      

    In   this   chapter,   we   compare   two   different   routes  of  using  the  nanosphere  lithography   for   the   manufacturing   of   well-­‐aligned,   density-­‐controlled   ZnO   nanowires   by   hydrothermal  growth.    

In   addition   to   crystallographic   and   microstructural   characterizations,   we   performed   dye   loading   measurements   in   order   to   compare   the   surface   area   of   the   nanowires  manufactured  by  both  routes  as   regard  to  an  unpatterned  array.    

Finally,  the  reversible  surface  wettability  of  

the  samples  was  evaluated.  

(2)

CHAPTER  V  -­‐  ZnO  nanowire  arrays  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  147  

1.  Introduction  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  149  

1.1  Synthesis  procedure  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  149  

1.2  Wetting  properties  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  151  

2.  Experimental  part  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  154  

2.1  Synthesis  of  the  samples  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  155  

2.2  Morphological  &  crystallographic  characterizations  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  156  

2.3  Dye  loading  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  156  

2.4  Wetting  properties  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  157  

3.  Results  and  discussion  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  157  

3.1  Morphological  and  crystallographic  properties  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  158  

3.2  Surface  area  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  161  

3.3  Wetting  properties  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  163  

4.  Conclusions  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  168  

5.  References  -­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐-­‐  170  

(3)

 

1.  Introduction  

ZnO   nanostructures,   particularly   nanorods,

[1]

  nanowires,

[2]

  and   nanotubes

[3]

  are   attracting   considerable   interest   because   of   their   large   effective   surface   area   resulting   from  a  high  surface-­‐to-­‐volume  ratio.

[4]

   

One-­‐dimensional  (1D)  ZnO  nanostructures  are  promising  candidates  for  applications  in   the   fields   of   catalysis,

[5]

  field-­‐emission   devices,

[6]

  chemical   sensors,

[7-­‐9]

 

nanogenerators,

[10]

  ultraviolet   lasers,

[11]

  field-­‐effect   transistors

[12]

  and   dye-­‐sensitized   solar  cells.

[13-­‐15]  

Most  of  these  applications  rely  on  ZnO  characteristics  such  as  the  wide   direct  band  gap  (3.37  eV)  or  high  electron  mobility  (100  cm

2

 V

-­‐1

 s

-­‐1

).    

Since  the  electrical  and  optical  properties  of  ZnO  nanostructures  depend  on  their  crystal   structure,   dimensions   and   morphology,

[16]

  shape   and   size   of   the   ZnO   nanostructures   play  a  vital  role  for  the  performance  of  the  devices.    

 

1.1  Synthesis  procedure  

Several   approaches   have   been   developed   for   the   growth   of   well-­‐aligned   1D   ZnO   nanostructures.   High-­‐quality   nanowires   can   be   prepared   by   vapor-­‐phase   techniques   such   as   vapor-­‐liquid-­‐solid   growth   (VLS),

[17]

  vapor-­‐solid   growth   (VS)

[18]

  or   chemical   vapor  deposition  (CVD)

 [19,  20]

 but  these  techniques  require  sophisticated  equipment,  fine   tuning  of  the  experimental  conditions,  single  crystal  substrates  and  high  temperatures   (up   to   900   °C)   which   are   not   compatible   with   organic   substrates   and   low   production   costs.   Electrodeposition   can   be   used   to   produce   pure   ZnO   nanowires   under   low   temperatures,  but  it  is  limited  to  conductive  substrates.

[21]

   

To   date,   the   hydrothermal   method   stands   out   as   the   most   attractive   alternative   for   obtaining   well-­‐aligned   nanostructures   of   ZnO   under   mild   conditions   with   simple   and   cheap  implementation.  In  the  first  step,  the  substrate  is  coated  with  ZnO  nanoparticles   (seeds).  In  the  second  step,  each  seed  acts  as  a  nucleation  site  for  the  formation  of  ZnO   nanowires   under   hydrothermal   growth   conditions.   Experimental   parameters,   such   as   Zn

2+

  concentration,   organic   additives   or   growth   time   and   temperature   determine   the   final  nanowire  dimension  and  quality.

[22]

   

The  nanowires  density  has  also  been  reported  to  have  a  strong  impact  on  the  efficiency   of  related  devices.

[23]

 Therefore,  in  order  to  control  the  size,  position  and  arrangement  of   ZnO  nanowires,  various  technologies  have  been  used  to  force  a  pattern-­‐arranged  growth   of  nanowires.    

They  include  conventional  photolithography,

[24,  25]

 laser  interference  lithography

[26]

 and  

e-­‐beam  lithography,

[27,  28]

 which  can  be  used  to  manufacture  nanopores  with  controlled  

shape  and  spacing.  However,  the  complexity  of  the  preparation  process  combined  with  

high  initial  equipment  costs  makes  those  lithographic  techniques  unfavorable  for  many  

researchers.  Recently,  Kang  et  al.

[29]

 reported  the  use  of  microcontact  printing  to  directly  

pattern   ZnO   nanoparticle   seeds.   However,   although   this   process   allows   the   rapid  

(4)

 

replication   of   a   similar   pattern,   it   does   not   control   the   nanowire   density   inside   the   patterned  area.    

Nanosphere   lithography   appears   therefore   as   a   very   promising   approach,   due   to   its   rapid  implementation  and  its  compatibility  with  wafer-­‐scale  processes.  It  consists  in  two   steps:   the   preparation   of   a   colloidal   crystal   mask   made   of   nanospheres   and   the   deposition  (often  by  sputtering)  of  the  material  of  interest  through  the  voids  between   the  spheres.  In  a  last  step  called  lift-­‐off,  the  mask  is  removed  and  the  (sputtered)  layer   keeps   the   ordered   patterning   of   the   mask   interstices.   In   recent   years,   nanosphere   lithography   has   attracted   growing   interest   due   its   potential   to   manufacture   a   wide   variety  of  1D,  2D  or  3D  nanostructures.

[30,  31]

 

In  the  ZnO  research  field,  nanosphere  lithography  has  so  far  mainly  been  used  to  create   ordered  arrays  of  metal  nanodots  that  act  as  catalysts  for  nanowire  growth  by  a  vapor-­‐

phase  technique.

[32,  33]

   

Some   recent   studies   investigated   with   some   success   the   possibility   to   combine   nanosphere  lithography  and  solution-­‐based  growth  of  ZnO.  

Li   et   al.

[34]

  employed   monolayers   of   polystyrene   colloids   to   guide   the   growth   of   hexagonally   patterned   ZnO   nanopillar   arrays   on   zinc   foils.   However,   the   developed   method   was   only   optimized   for   an   individual   nanopillar   growth   at   each   interstice   between   neighboring   colloidal   spheres.   Moreover,   they   faced   difficulties   to   obtain   perfect  arrays  of  ZnO  in  a  large  area  because  of  the  defects  in  the  mask.    

Pyun   et   al.

[35]

  first   reported   the   formation   of   ZnO   nanotubes   with   the   assistance   of   polystyrene   colloids   on   ZnO   seed   layer   prepared   by   metal-­‐organic   chemical   vapor   deposition   (MOCVD),   which   is   a   rather   expensive   technique   that   fits   poorly   with   the   expected  lowering  of  production  costs.  They  observed  that  single-­‐crystalline  nanotubes   were   formed   just   below   the   nanopsheres,   whereas   solid   nanorods   were   grown   in   the   absence   of   PS   colloids   or   at   the   apertures   between   three   adjacent   nanospheres.   They   didn’t   evidence   any   significant   effect   on   the   luminescent   properties   of   the   ZnO   nanotubes  compared  to  those  of  the  ZnO  nanorods.  

Hexagonally  packed  ZnO  hemisphere-­‐array  films  formed  by  growth-­‐hindered  nanowires   were  synthesized  by  Chang  et  al.

[36]

 Although,  their  nanostructures  did  not  correspond   to   a   templated   growth   of   well-­‐aligned   ZnO   nanowires   with   a   high   aspect   ratio,   they   evidenced   an   increased   surface   area   compared   to   unpatterned   ZnO   nanorod   films.    

However,  the  ZnO  hemisphere-­‐array  and  the  unpatterned  ZnO  films  were  respectively   synthesized   by   hydrothermal   growth   and   by   electrochemical   deposition.   According   to   us,  this  impedes  any  reliable  comparison  of  their  properties.  

Finally,   Fragala   et   al.

[37]

  attempted   to   pattern   the   seed   layer   by   using   polystyrene  

microspheres   for   the   selective   ZnO   nanorod   hydrothermal   growth.   They   observed   a  

bimodal  ZnO  nanorod  growth  as  a  result  of  a  local  nuclei  concentration  gradient  in  the  

different  regions  of  the  patterned  substrate.    

(5)

 

In   our   study,   we   compare   two   different   nanosphere   lithography   routes,   which   will   be   further   referred   to   as   the   “templated   growth”   (TG)   and   the   “templated   seeding”   (TS)   procedures,  to  establish  a  low  cost  synthesis  of  patterned  ZnO  nanowire  arrays  over  a   large  area  on  glass-­‐FTO  substrates.  In  addition  to  crystallographic  and  microstructural   characterizations,   we   performed   dye   loading   measurements   in   order   to   compare   the   surface   area   of   the   nanowires   manufactured   by   both   routes   as   regard   to   unpatterned   array.  

 

1.2  Wetting  properties  

Recently,   there   have   been   lot   of   interest   in   studying   the   wetting   properties   of   metal   oxides   nanomaterials   (mainly   of   ZnO

[38,   39]

,   TiO

2[40,   41]

  and   Al

2

O

3[42]

),   which   can   be   reversibly   switched   between   superhydrophobicity   and   superhydrophilicity   by   alternation  of  ultraviolet  (UV)  irradiation  and  dark  storage.

[38,  43]

 

Indeed,   as   nanoscale   devices   aimed   for   chemical   and   biological   sensing,   surface   wettability   plays   a   very   important   role   and   surfaces   with   controllable   and   reversible   wettability   are   highly   desirable,

[44]

  particularly   for   the   control   of   effective   micro-­‐   or   nano-­‐fluid  motion.

[45]

 

In  nature,  many  surfaces  (from  plants  or  animals)  are  highly  hydrophobic.  

The   best-­‐known   example   of   a   hydrophobic   surface   is   the   leaves   of   the   lotus   plant.

[46]

 

Numerous   studies   confirmed   that   the   chemical   composition   (a   wax)   of   the   leaves   combined   with   a   specific   micro/nano   hierarchical   surface   structure   (Figure   V   -­‐   1)   provide  the  lotus  plant  its  unique  superhydrophobic

*

 and  self-­‐cleaning  properties.  

 

  Figure  V  -­‐  1  

Lotus  leaf  image  and  SEM  micrographs  of  its  upper  side  showing  the   hierarchical  structure.  

[47]  

 

Compared   to   organic   materials,   inorganic   materials   exhibit   better   light,   thermal,   and   chemical  stabilities.    

By   transferring   the   structure   of   selected   plant   surfaces   to   practical   materials,   superhydrophobic   surfaces   could   be   manufactured   and   hence   have   recently   been   the   focus  of  considerable  scientific  interest.

[48-­‐51]

   

                                                                                                                         

*

  A  surface  is  called  superhydrophobic  when  water  contact  angle  exceeds  150

°

.  

(6)

 

This  is  due  to  the  fact  that  artificial  superhydrophobic  surfaces  are  promising  candidates   for  a  number  of  practical  applications  such  as  easy  cleaning  clothing  or  windows.  

 

Indeed,   surface   cleaning   of   building   materials   like   facades   and   glass   panes   generates   considerable   trouble,   high   consumption   of   energy   and   chemical   detergents   and,   consequently,  high  costs.  Non-­‐wettable  surfaces  may  also  improve  the  ability  to  prevent   frost  from  forming  or  adhering  to  the  surface.  

The  adhesion  of  water,  as  well  as  contaminants  is  considerably  reduced.  Water  droplets   coming   in   contact   with   a   superhydrophobic   surface   form   nearly   spherical   beads.   The   contaminants,   either   inorganic   or   organic,   on   such   surfaces   are   picked   up   by   water   droplets   or   adhere   to   the   water   droplet   and   are   removed   from   the   surface   when   the   water  droplets  roll  off  (Figure  V  -­‐  2).  The  combination  of  low  surface  energy

 and  micro-­‐  

and/or  nano-­‐structured  features,  which  can  certainly  reduce  the  contact  area  between   the  surface  and  water  droplets,  form  superhydrophobic  surfaces.  

 

 

Figure  V  -­‐  2

 

On  a  typical  surface  a  drop  of  water  slides  across  and  leaves  most  dirt  particles  sticking  to   the  object,  while  on  a  superhydrophobic  surface  the  drop  rolls  across,  picking  up  dirt  and   carrying  it  away.

[52]

 

 

On  most  surfaces,  the  motion  of  the  drop  is  opposed  by  energy  barriers,  which  leads  to   an   effect   referred   to   as   contact   angle   hysteresis.   Indeed,   there   is   usually   a   difference   between  the  angle  produced  as  the  volume  of  the  drop  is  increased  (advancing  angle)   and  that  when  the  volume  is  reduced  (receding  angle).  This  hysteresis  gives  a  measure   of  the  surface  stickiness.    

                                                                                                                         

  The  surface  energy  is  a  solid  surface  characteristic  associated  with  the  molecular  forces  of  its   interaction  with  another  material.    

Surface  tension  is  the  property  of  a  liquid  arising  from  unbalanced  molecular  forces  at  or  near   the  surface.  If  it  is  higher  than  the  surface  energy  of  a  material,  the  liquid  tends  to  form  droplets   rather  than  spread  out  or  “wet  out”  as  some  refer  to  it.  Surface  tension  is  normally  measured  in   energy  units  called  dynes/cm.  A  dyne  is  the  force  required  to  produce  an  acceleration  of  1  cm/s

2

  on  a  mass  of  1g.  

 

(7)

 

A   drop   with   a   very   low   hysteresis   easily   rolls   off   an   inclined   surface.   Figure   V   -­‐   3   presents  two  different  methods  for  the  measurement  of  contact  angle  hysteresis.  

 

 

Figure  V  -­‐  3

 

Methods  to  measure  the  contact  angle  hysteresis  (θ

advancing

 -­‐  θ

receiding

)

 [53]

 

(a)  The  tilting  plate  method  captures  the  contact  angles  measurements  on  both  the  left   and  right  sides  of  a  sessile  drop  while  the  solid  surface  is  being  inclined  typically  from  0°  

to  90°.  

(b)   The  add   and   remove   volume   method   requires   adding   (or   removing)   liquid   to   the   maximum   volume   permitted   without   inducing   a   motion   of   the   three   phase   (3φ)   contact   line  (liquid-­‐vapor-­‐solid).  

 

The   hysteresis   characterizes   the   topography   of   the   sample.   Water   droplets   on   rugged   hydrophobic  surfaces  typically  exhibit  one  of  the  following  two  states  (Figure  V  -­‐  4):  (a)   the  Wenzel  state

[54]

 in  which  water  droplets  are  in  full  contact  with  the  rugged  surface   or  (b)  the  Cassie  state

[55]

 in  which  water  droplets  no  longer  penetrate,  but  rest  on  top  of   the  roughness  features.  

 

 

Figure  V  -­‐  4

 

Two  different  wetting  state  for  a  drop  on  a  hydrophobic  textured  surface.  (a)  Wenzel  state   and  (b)  Cassie  state  or  “fakir”  state.

[49]

 

 

The  Wenzel  state  is  characterized  by  a  huge  hysteresis  (from  50°  to  100°)  compared   to   the   Cassie   state   (from   5°   to   20°),   and   hence   displays   a   more   sticky   behavior   as   contact  area  with  the  surface  is  bigger.  

In  summary,  the  wettability  of  solid  surfaces  is  therefore  a  very  important  property,  and   may  be  governed  by  both  the  chemical  composition  and  the  geometrical  structure  of  the   surface.

[50,  56]

   

!"#$ !"#$%&'(#)*+',+*-./' !%#$ 0//')%/'1+,.2+'3.#4,+'5+*-./!

!"$#$%&'#&()*%+((

!"#$ !%#$

(8)

 

Techniques   to   make   superhydrophobic   surfaces   can   be   simply   divided   into   two   categories:  making  a  rough  surface  from  a  low  surface  energy  material  and  modifying  a   rough  surface  with  a  material  of  low  surface  energy.  

 

In  our  study,  we  evaluated  the  impact  of  the  templating  of  the  ZnO  nanowire  arrays  on   the  wetting  properties,  focusing  on  roughening  of  the  material.  We  have  also  studied  the   reversibility  of  the  surface  wettability,  which  is  crucial  for  real  device  applications.  

 

2.  Experimental  part    

The  two  preparation  routes  are  schematically  depicted  in  Figure  V  -­‐  5.  

 

 

  Figure  V  -­‐  5  

Two   nanosphere   lithography   routes   proposed   for   the   growth   of   ZnO   nanowire   arrays.  

Templated  growth  process  (TG)  &  Templated  seeding  process  (TS)  

 

In  the  “Templated-­‐Growth”  TG  procedure  (Figure  V  -­‐  5,  left  side),  a  continuous  layer  of   ZnO  seeds  is  obtained  by  spin  coating  a  1:1  solution  of  zinc  acetate  and  ethanolamine  in   ethanol,  followed  by  annealing  at  350°C  for  30  min  to  obtain  oriented  ZnO  seeds  (step   1).    

!

(9)

 

The   colloidal   crystal   mask   is   then   deposited   (step   2)   by   spin   coating   a   suspension   of   monodisperse  polystyrene  nanospheres  (490  nm  diameter).  Next,  we  proceeded  to  the   hydrothermal  growth  of  the  nanowires  (step  3)  in  aqueous  suspensions  containing  zinc   nitrate  hydrate  (25  mM)  and  hexamethyltetramine  (25  mM)  at  90°C  for  2.5  hours.  We   finally  removed  the  polystyrene  nanospheres  (step  4)  by  calcination  (350  °C  for  30  min).  

The  samples  before  and  after  removal  of  the  nanospheres  will  be  further  referred  to  as   TG-­‐NS  and  TG-­‐calcined.  

In  the  “Templated  Seeding”  TS  procedure  (Figure  V  -­‐  5,  right  side),  the  seeding  and  the   colloidal  crystal  mask  were  realized  with  the  same  experimental  conditions  as  in  the  TG   procedure   but   in   a   different   order.   In   the   first   step,   a   patterned   layer   of   seeds   was   obtained  by  formation  of  the  colloidal  crystal  mask  on  the  substrate  (step  1*),  followed   by  the  spin  coating  of  the  ZnO  seeds  over  the  nanospheres  (step  2*).  We  then  removed   the  nanospheres  (step  3*)  by  calcination  (TS-­‐calcined)  or  by  sonication+calcination  (TS-­‐

sonicated).  The  last  step  is  the  hydrothermal  growth  of  the  ZnO  nanowires  (step  4*).  

 

2.1  Synthesis  of  the  samples  

All   samples   were   manufactured   on   FTO   conducting   glass   substrates   (15Ω/sq)   purchased  from  Dyesol.  Previous  to  use,  the  substrates  were  washed  by  sonication  15   min  in  acetone  and  15  min  in  ethanol  and  air-­‐dried.  

In  both  the  templated  seeding  (TS)  and  templated  growth  (TG)  routes,  the  seeding  and   the  colloidal  crystal  masks  were  realized  in  the  same  way.    

Monodisperse   polystyrene   nanospheres   with   a   mean   diameter   of   490   nm   were   purchased  from  Bangs  Laboratories  as  suspensions  in  water  (concentration  of  about  10  

%  wt)  in  order  to  prepare  single-­‐layer  colloidal  crystal  masks.    

Before   deposition,   we   diluted   the   aqueous   nanospheres   (150   μL)   suspension   in   a   surfactant  Triton  X-­‐100/MeOH  mixture  1:400  by  volume  (350  μL)  and  filtered  through   centrifugal   filter   units   (porosity   0.65   µm)   in   order   to   eliminate   aggregates   of   higher   dimensions,  which  would  disturb  the  formation  of  the  monolayer.  Then,  we  vortexed  the   suspensions  during  2  minutes  to  ensure  homogeneity.      

We   dispensed   a   drop   of   the   suspension   (50   μL)   on   the   substrate   with   an   Eppendorf   pipette.   All   the   samples   were   prepared   with   the   same   spin-­‐coating   parameters,   characterized   by   a   high   rotation   speed   reached   in   a   short   time.   The   samples   were   accelerated   to   1930   rpm   (acceleration   rate   643   rpm/s)   for   2s   and   then   spun   at   2500   rpm   (acceleration   rate   190   rpm/s)   for   another   2s,   followed   by   final   spin   at   6900   rpm   (acceleration  rate  1467  rpm/s)  for  2s.  

We  seeded  the  samples  with  an  oriented  ZnO  thin  film  by  a  sol-­‐gel  spin  coating  method.  

We  prepared  the  seeding  solution  by  dissolving  zinc  acetate  Zn(CH

3

COO)

2

.2H

2

O  (0.025  

mol)   in   ethanol   (100   mL)   at   room   temperature.   Ethanolamine   (C

2

H

7

NO)   was   used   as  

stabilization   agent   and   its   molar   ratio   to   zinc   acetate   was   kept   at   1:1.   The   resultant  

(10)

 

solution   was   stirred   at   60°C   for   1h   to   yield   a   clear   and   homogeneous   sol.   Then,   the   mixed  sol  was  aged  at  room  temperature  for  24h.  The  precursor  solution  was  dropped   onto  the  samples,  which  were  then  spinned  at  3000  rpm  for  20  s  at  room  temperature  in   a   constant   humidity   atmosphere.   The   relative  humidity  level  was  set  to  35  ±  2  %  and   was  measured  with  a  digital  hygrometer  (Testo).  To  ensure  complete  sample  coverage   with  ZnO  seeds,  the  spin  coating  process  was  repeated  nine  times.    

Seeded  substrates  were  annealed  at  350  °C  for  30  min  to  obtain  the  oriented  ZnO  seed   layers.  In  the  TS-­‐sonicated  samples,  we  previously  removed  the  nanospheres  (step  3*  in   Figure  V  -­‐  5)  by  sonication  before  this  calcination  step  while  in  the  TS-­‐calcined  samples,   the   calcination   played   the   dual   role   of   removing   the   nanospheres   and   orienting   the   seeds.  

The  growth  of  the  nanowires  (step  3  or  4*  in  Figure  V  -­‐  5)  was  carried  out  by  immersing   seeded  substrates  upside  down  in  aqueous  solutions  containing  zinc  nitrate  hydrate  (25   mM)   and   hexamethyltetramine   (25   mM)   at   90°C   for   2.5   hours.   The   arrays   were   then   rinsed   with   milliQ   water   and   dried   in   oven   in   air   at   60°C   for   30   min.     To   remove   the   template  of  polystyrene  nanospheres,  we  calcined  the  TG  samples  (step  4  in  Figure  V  -­‐  

5)  in  air  at  350°C  during  30  min.  

 

2.2  Morphological  &  crystallographic  characterizations  

The   morphology   of   the   samples   was   characterized   by   scanning   electron   microscopy   (SEM)   on   a   FEG-­‐ESEM   XL30   (FEI)   with   an   accelerating   voltage   of   5   kV   under   high   vacuum.  All  samples  were  gold-­‐coated  (60s)  before  observation.    

The  XRD  patterns  were  recorded  using  a  Bruker  AXS  D8  diffractometer  in  θ–2θ  locked   coupled  mode  (30-­‐70  °2θ,  step  size  0.04°).  

 

2.3  Dye  loading  

The   dye   loading   was   measured   by   UV-­‐vis   spectroscopy   on   EtOH/water   (1/1   v/v)   solutions   with   a   Perkin   Elmer   UV-­‐vis   Spectrometer   Lambda   14   P.   The   samples   were   soaked  in  N-­‐719  (Solaronic)  dye  solutions  (ethanolic  solution,  3.0  10

-­‐4

 M)  during  16  h  at   room  temperature.    

After   drying,   the   samples   were   desorbed   in   a   known   volume   of   KOH     solution  (10

-­‐3

 M).    

A  calibration  curve  was  used  to  calculate  the  experimental  extinction  coefficient  (at  500   nm)  of  N-­‐719  dye  (12  500  (mol/L)

-­‐1

cm

-­‐1

).    

The  dye  loading  of  the  samples  was  then  determined  from  the  desorption  solutions.  

(11)

 

2.4  Wetting  properties  

Surface   wettability   was   evaluated   by   ultra-­‐pure   water   (MilliQ)   contact   angle   (CA)   measurements  in  the  GRASP  laboratory,  using  a  CAM  200  Optical  Contact  Angle  meter   (KSV  Instruments  Ltd.)  and  the  CAM  200  software  provided  with  the  instrument.  

A   5   μL   MilliQ   water   droplet   was   deposited   on   the   surface   of   the   as-­‐manufactured   samples   using   an   automated   drop   dispensing   system.   Each   CA   measurement   was   repeated  three  times  at  various  places  on  two  different  samples.  

Light-­‐induced  hydrophilicity  was  evaluated  by  irradiating  the  samples  during  2h  inside   a   specially   designed   chamber   equipped   with   six   Eversun   UVA   fluorescent   lamps   (OSRAM,  L40W/79K).  

After   each   irradiation,   a   5   μL   water   drop   was   placed   on   the   sample   and   the   corresponding  CA  was  measured  again.  

In  order  to  study  the  reversibility  of  the  wettability  transition,  the  samples  were  either   stored   in   the   dark   for   a   week   at   room   temperature   or   annealed   during   20h   at   50°C   under  O

2

 atmosphere  before  another  CA  measurement.  

The   contact   angle   hysteresis,   defined   as   the   difference   between   the   advancing   and   receding  angle,  was  also  measured.  

 

3.  Results  and  discussion  

A   critical   point   in   nanosphere   lithography   is   that   the   experimental   conditions   of   the   colloidal  crystal  mask  formation  should  ensure  the  presence  of  large  covered  areas  as   well  as  the  accessibility  of  the  voids,  which  is  of  crucial  importance  both  for  the  growth   of   the   nanowires   in   the   TG   procedure   and   for   the   deposition   of   the   seeds   onto   the   substrate  in  the  TS  procedure.    

Electron   micrographs   of   the   colloidal   crystal   mask   prior   to   ZnO   hydrothermal   growth   are   shown   in   Figure   V   -­‐   6.     The   low   magnification   micrograph   proves   that   the   nanospheres  are  present  over  several  mm

2

.  The  inset  shows  the  hexagonal  close  packing   (hcp)  of  the  polystyrene  nanosphere  monolayer.    

Figure  V  -­‐  6

   

SEM   micrograph   of   a   monolayer   mask   of   polystyrene   nanospheres   (490   nm   diameter),  

prepared  by  spin  coating  with  a  high  coverage  rate.  The  inset  is  a  high  magnification  view  

showing  the  hexagonal  packing  of  the  monodisperse  nanospheres.  

(12)

 

3.1  Morphological  and  crystallographic  properties  

Scanning   electron   micrographs   of   the   ZnO   nanowires   grown   via   the   various   synthetic   routes  are  presented  in  Figure  V  -­‐  7.  In  the  TG  procedure,  the  ZnO  nanowires  grew  in  the   voids   between   the   nanospheres   with   hexagonal   arrangement.   The   "side   view"  

micrograph   (Figure   V   -­‐   7   (b))   highlights   the   role   of   the   colloidal   crystal   mask   in   patterning   the   growth   of   the   nanowires.   After   calcination   at   350°C   to   remove   the   polystyrene  mask,  the  hexagonal  pattern  was  preserved  (Figure  V  -­‐  7  (c)),  as  well  as  the   orientation  of  the  nanowires  (Figure  V  -­‐  7  (d)).    

In  the  TS  route,  the  polystyrene  masks  were  used  to  selectively  deposit  the  seeds  on  the   substrate   and   removed   either   by   calcination   or   sonication,   before   the   hydrothermal   growth.  At  first  sight  (Figure  V  -­‐  7  (e)  and  Figure  V  -­‐  7  (g)),  the  coverage  of  the  substrate   by  nanowires  seems  to  be  higher  than  in  the  TG  route.  However,  the  nanowires  are  not   well-­‐aligned  anymore:  the  nanowires  display  a  divergent,  bush-­‐like  structure  (see  side-­‐

view  micrographs  in  Figure  V  -­‐  7  (f)  and  Figure  V  -­‐  7  (h)).    

The   dimensions   of   the   nanowires   manufactured   are   homogenous   and   similar   for   both   procedures   (TG   –   TS),   with   typical   diameter   of   ≈   50   nm   and   height   ≈   1   μm.   This   demonstrates   the   high   reproducibility   of   the   growth   process   and   the   high   quality   (coverage  and  ordering)  of  the  colloidal  masks.    

In  contrast,  Fragala  et  al.

[37]

 observed  a  bimodal  morphology  of  ZnO  nanowires  that  was   attributed  to  a  lack  of  covering  by  polystyrene  spheres  (1  μm  diameter).  They  reported   that  ZnO  grown  on  a  homogenous  seed  layer  were  larger  (200  nm)  than  the  nanowires   in  the  patterned  regions  (80  nm).    

Besides,  Pyun  et  al.

[35]

 underlined  a  difference  in  growth  rates  for  ZnO  nanowires  grown   with  and  without  polystyrene  colloids.  They  showed  that  ZnO  nanowires  enclosing  the   polystyrene   colloids   are   much   longer   than   nanowires   formed   on   the   area   not   covered   with  the  nanospheres.  

As  we  synthesized  nanowires  for  short  period  (2h30)  we  didn’t  observe  this  difference   of  length  between  templated  and  unpatterned  nanowire  arrays.  For  equivalent  solution   and  growth  time  (2h30),  our  nanowires  are  longer  (1  µm)  than  that  obtained  by  Pyun  et   al.

[35]

 (750  nm).  The  higher  temperature  (90°C)  of  hydrothermal  synthesis  could  explain   the  enhanced  growth  rate  along  [001]  direction  of  our  ZnO  nanowire  arrays.

[22]

   

 

Figure  V  -­‐  8  shows  the  X-­‐ray  diffraction  (XRD)  patterns  of  films  prepared  by  the  TG  and   TS  synthetic  routes.  All  reflections  belong  either  to  the  wurtzite-­‐type  phase  (ZnO  –  ICDD   PDF4+  No.  04-­‐003-­‐02106)  or  to  the  fluorine-­‐doped  tin  oxide  layer  of  the  substrate  (SnO

2

  –  ICDD  PDF4+  No.  00-­‐046-­‐1088).    

The  prevailing  intensity  of  the  ZnO  (002)  peak  evidences  the  c-­‐axis  texturation  normal  

to  the  substrate.  This  result  confirms  the  role  of  the  seed  layer  to  promote  the  alignment  

of  ZnO  nanowire  arrays,  as  already  reported  for  unpatterned  ZnO  nanowire  arrays.

[22,  57]

   

(13)

         

 

Figure  V  -­‐  7

 

SEM  micrographs  of  ZnO  nanowires  manufactured  via  the  “templated  growth”  (TG)  and  the  

“template   seeding”   (TS)   routes.   (a)   &   (b)   TG-­‐NS   :   (a)   As-­‐grown   nanowires   highlighting   a   hexagonal   symmetry.   (b)   The   polystyrene   nanospheres   are   still   present   between   the   nanowires  in  the  side-­‐view.  (c)  &  (d)  TG  calcined  :  (c)  The  nanospheres  have  been  removed   by   the   annealing   treatment   and   left   a   well-­‐defined   hole   where   they   were   sitting.   (d)   Side-­‐

view  of  the  nanowires  with  no  trace  of  the  nanospheres.  Inset  scale  bars  are  500  nm.  Before   the  hydrothermal  growth  of  the  nanowires  in  the  TS  process,  the  nanospheres  were  removed   either  by  calcination  (TS  -­‐  calcined)  at  350  °C  ((e)  &  (f))  or  by  sonication  (TS  -­‐  sonicated)  in   toluene  ((g)  &  (h)).  Tilted  SEM  micrographs  ((f)  &  (h))  show  the  bush  like  orientation  of  the   nanowires.  

 

   

!

(14)

 

 

Figure  V  -­‐  8

 

XRD  patterns  of  the  samples  grown  via  the  “templated  growth”  (TG)  or  “templated   seeding”  (TS)  routes.  

 

In  order  to  quantify  the  c-­‐axis  texturation,  the  ratios  of  (002)  to  (103)

 peak  net  surface   areas  are  given  in  Table  V  -­‐  1.  The  same  pronounced  c-­‐axis  texturation  is  found  in  the  TG   samples  and  in  a  reference  unpatterned  (=continuous)  nanowire  film.  As  predicted  from   the   SEM   micrographs,   the   c-­‐axis   texturation   is   far   less   pronounced   in   samples   manufactured   by   the   TS   route,   especially   when   the   nanospheres   were   removed   by   sonication  (TS-­‐sonicated).  

 

Table  V  -­‐  1  

Comparison  between  XRD  (002)  textures  of  TG,  TS  and  unpattenerd  nanowire  arrays  

Sample     Ratio  of  XRD  peak  

surface  areas   (002)/(103)  

TG  calcined   54  

TG  -­‐  NS   51  

TS  calcined   12  

TS  sonicated   8  

Unpaterned  continuous  

Nanowire  Film   52  

                                                                                                                         

 In  the  hexagonal  lattice  (such  as  wurtzite  structure),  sometimes  the  four  Bravais-­‐Miller  

indices  (h  k  i  l)  are  used  instead  of  the  three  Miller  indices  (h  k  l).  In  the  4-­‐index  notation  

the  i  index  is  related  to  the  h  and  k  indices  by  the  relation  h  +  k  +  i  =  0.  

(15)

 

The   crystalline   domain   size   calculated   from   the   (002)   ZnO   peaks   gave   an   average   domain   size   of   ≈   45   nm,   which   agrees   with   our   SEM   observations   of   the   nanowire   diameter.   This   is   expected   since   hydrothermal   growth   usually   yields   single   crystalline   nanorods.

[58]

 

 

3.2  Surface  area  

The  performance  of  such  nanostructures  for  applications  often  depends  critically  on  the   accessible   surface   area.   In   order   to   assess   this   parameter,   we   performed   dye   loading   measurements,   after   verifying   that   the   N719   dye   does   not   adsorb   on   the   polystyrene   nanospheres.    

Dye  loading  results  are  listed  in  Table  V  -­‐  2  for  the  samples  prepared  by  the  TS  and  TG   routes   and   for   a   reference   unpatterned   nanowire   array   synthesized   in   the   same   hydrothermal  growth  conditions.    

 

The   two   samples   obtained   by   the   TS   route   (TS-­‐calcined   and   TS-­‐sonicated)   have   dye   loading   values   significantly   lower   than   the   samples   prepared   by   the   TG   route.   This   observation  points  to  a  lower  density  of  nanowires  (number  of  nanowires  per  area  unit   of  substrate)  due  to  a  less  efficient  seeding  in  the  TS  route.  The  lowest  value  for  the  TS-­‐

sonicated  sample  suggests  that  some  seeds  are  removed  from  the  substrate  during  the   sonication  step.  

 

Table  V  -­‐  2

 

Dye  loading  values  of  TG,  TS  and  unpattenerd  nanowire  arrays    

Sample     Dye  loading  N-­‐719*  

[mol/mm

2

]  

TG  calcined   2.43  10

-­‐10

 

TG  -­‐  NS   1.88  10

-­‐10

 

TS  calcined   1.60  10

-­‐10

 

TS  sonicated   1.35  10

-­‐10

 

Unpaterned  Nanowire  Film   1.68  10

-­‐10

 

*  Dye  loading  values  are  expressed  in  moles  of  N-­‐719  dye  per  mm

2

 of  soaked  area.  

 

Regarding  the  TG  samples,  the  dye  loading  of  the  TG-­‐calcined  sample  is  higher  than  the  

value  for  the  TG-­‐NS  sample,  because  of  the  removal  of  the  polystyrene  nanospheres  that  

were  "screening"  the  bottom  of  the  nanowires.  Moreover,  the  calcinations  may  also  have  

removed   organic   contaminants,   therefore   favoring   the   dye   loading.   The   dye   loading  

value   obtained   for   the   TG-­‐calcined   sample   also   significantly   exceeds   the   value   for   the  

unpatterned  nanowire  array.    

(16)

 

Since  it  is  obviously  unlikely  that  a  patterned  array  may  have  a  higher  nanowire  density   than  an  unpatterned  array,  the  reason  for  this  difference  in  dye  loading  is  to  be  found  in   a   difference   of   accessible   surface   area.   Unlike   Chang   et   al.

[36]

,   both   the   patterned   and   unpatterned   ZnO   samples   were   grown   in   the   same   conditions   of   seeding   and   hydrothermal   growth.   Therefore,   we   can   unambiguously   conclude   that   the   templated   growth  (TG)  route  significantly  contributed  to  an  increase  of  the  accessible  surface  area,   which  was  the  ultimate  goal  for  using  nanosphere  lithography  in  this  study.    

     

In   order   to   estimate   the   nanowire   density   of   the   arrays   obtained   by   the   TG   route,   we   considered   a   perfect   colloidal   mask   where   each   nanosphere   is   surrounded   by   six   interstices   (inset   Figure   V   -­‐   6).   In   the   case   of   only   one   nanowire   per   interstice,   the   density  would  be  of  9.6  10

12

 nanowires/m

2

 for  a  colloidal  mask  with  490  nm  diameter   nanospheres.   This   scenario   corresponds   to   samples   manufactured   by   nanosphere   lithography   and   vapor-­‐liquid-­‐solid   (VLS)   epitaxy   mechanism   catalyzed   by   Au   nanodots.

[59]

   

In  our  process,  electron  micrographs  (Figure  V  -­‐  7)  show  that  several  nanowires  grow   between   the   nanospheres,   leading   to   an   increase   in   the   total   surface   area.   Based   on   geometric  calculations,

[60]

 it  is  possible  to  evaluate  the  size  of  the  triangular  interstice  in   the  colloidal  mask  (inset  Figure  V  -­‐  6).  Given  the  observed  diameter  of  the  nanowires,  we   estimate  that  four  nanowires  can  take  place  in  the  voids  from  a  colloidal  mask  with  490   nm   diameter   nanospheres.   This   would   raise   the   density   up   to   3.8   10

13

  nanowires/m

2

.   Such   a   value   is   in   the   same   range   than   reported   values   for   track   etched   polymer   templates   designed   for   micro-­‐   and   nanofabrication.

[61]

  However,   the   pores   in   the   polymer  membranes  are  randomly  distributed,  which  is  not  the  case  in  the  nanosphere   lithography  process.  

 

By  changing  the  size  of  the  nanospheres,  it  is  possible  to  tune  the  density  of  nanowires,   which   may   influence   the   efficiency   of   the   applications,   as   previously   mentioned.   For   example,  in  the  field  of  dye-­‐sensitized  solar  cells,  the  fact  that  electron  diffusion  length  is   in   the   range   of   20   nm   suggests   that   the   best   well-­‐aligned   nanostructure   should   be   composed  of  vertical  nanostructures  of  about  20-­‐40  nm  diameter.

[15]

   

However,  the  size  of  the  interstice  will  decrease  with  the  nanosphere  diameter  and  a  too   small  space  could  hinder  the  growth  of  the  nanowires.

[35,  36]

 We  estimate  that  the  critical   interstice   size   for   the   growth   of   one   nanowire   would   be   reached   with   ≈   215   nm   diameter  nanospheres.  

Therefore   the   adjustment   of   the   interstice   size   is   of   major   importance   to   have   control  

over   the   manufacturing   of   NSL-­‐patterned   ZnO   nanowire   arrays.   The   use   of   an   oxygen  

plasma   etching   treatment   has   been   reported   in   literature   for   the   reduction   of   the  

polystyrene   nanosphere   diameters

[62]

  or   correction   of   the   deformation   of   soft  

polystyrene  nanospheres.

[63]    

(17)

 

This  etching  of  the  mask  would  be  necessary  if  nanospheres  with  a  diameter  <  215nm   are  envisaged  to  increase  the  packing  density  of  interstices  and  nanowires.  

 

3.3  Wetting  properties  

To  evaluate  the  patterning  effect  of  the  ZnO  nanowire  arrays  on  their  wetting  properties,   we   measured   water   contact   angles   on   the   as-­‐prepared,   unpatterned   (Figure   V   -­‐   9   (a))   and   TG   (templated   growth)   ZnO   nanowires   (Figure   V   -­‐   9   (b)),   which   present   higher   surface  area  compared  with  TS  samples  (Table  V  -­‐  2).    

Both  unpatterned  and  patterned  samples  revealed  a  hydrophobic  behavior  with  water   contact  angles  (CA)  higher  than  90°.    

The  patterning  induced  by  the  templated  growth  resulted  in  a  significant  increase  of  the   contact  angle,  almost  reaching  the  superhydrophobic  limit  (112°  ±  2  à  132°  ±  2).  The   CA   of   the   unppatterned   array   agrees   with   recently   reported   studies

[45,   64]

  on   ZnO   nanowires/nanorods,   keeping   in   mind   that   it   is   difficult   to   meet   exactly   the   same   synthesis   conditions.   Up   to   now,   no   CA   measurement   was   ever   reported   on   ZnO   nanowire  arrays  manufactured  with  the  templated  growth  process.  

It   is   well   known   that   the   surface   free   energy   and   the   surface   roughness   play   a   very   important   role   in   the   wetting   properties.   Recently,   it   has   also   been   reported   that   the   wettability  of  ZnO  depends  on  the  surface  crystal  structure.

[65-­‐67]

 

 

 

Figure  V  -­‐  9

 

Reversible   surface   wettability   transition   of   ZnO   nanowire   arrays   grown   manufactured   by   (a)   conventional  unpatterned  hydrothermal  growth  and  (b)  templated  growth  route.  

 

!

!"#$#%&'(#)#'##

%#*+#

,-.#

!"#$#%%'(#)#'#

%#*+#

!

,/.# /0$&''"+0"()*+,-'.)) !"#$%&'"()*+,-'.))

(18)

 

We  will  consider  all  these  concepts  to  understand  the  hydrophobic  behavior  of  the  as-­‐

prepared  samples  and  the  increased  hydrophobic  character  of  the  patterned  sample.  

The  anisotropic  nanowire  growth  evidenced  in  the  XRD  patterns  (Figure  V  -­‐  8)  is  due  to   different  surface  free  energies  of  the  various  growing  crystallographic  planes.

[68]

 A  fast   growing  plane  generally  tends  to  disappear  leaving  behind  slower  growing  planes  with   lower   surface   energy.

[38]

  For   the   anisotropic   ZnO   nanowire,   the   velocities   of   crystal   growth   in   various   directions   were   reported   to   be   100 > 101 > 001 ≈ 001 .

[69]

 

Therefore,   compared   with   a   sample   with   a   random   orientation,   the   as-­‐prepared,   unpatterned   and   TG   patterned   samples   would   have   the   lowest   surface   free   energy,   which  can  be  explained  by  a  closer  look  at  the  crystal  structure  (Figure  V  -­‐  10).  

 

Zinc  oxide  has  wurtzite  type  symmetry  and  thus  belongs  to  the  space  group  C

!"!

.    

On  the  non-­‐polar  planes  (e.g.  (110)  face),  both  oxygen  and  zinc  ions  are  terminated  in   the  same  plane.  On  the  other  hand,  there  are  two  possibilities  of  surface-­‐terminated  ions   on  the  polar  (001)  planes,  that  is,  oxygen  ions  or  zinc  ions.  Miyauchi  et  al.

[65]

 calculated   that  the  total  energy  of  the  oxygen-­‐terminated  surface  was  higher  than  that  of  the  zinc-­‐

terminated  surface.  These  results  imply  that  the  zinc  ions  terminated  structure  is  more   energetically   stable   than   the   oxygen   ions   termination.   Under   this   stable   structure,   oxygen  ions  are  not  exposed  at  the  surface,  which  is  not  the  case  in  the  non-­‐polar  planes.  

As  surface  oxygen  ions  are  considered  to  act  as  reactive  sites  for  increasing  OH  species   on   the   surface,   it   implies   that   the   hydrophilicizing   rate   of   ZnO   nanowires   with   high   proportion  of  nonpolar  planes  is  faster  than  those  with  low  proportion.    

However,  as  we  did  not  evidence  any  significant  difference  between  the  XRD  patterns  of   both  samples,  the  high  contact  angle  measured  on  the  TG  patterned  sample  cannot  be   attributed  to  a  lower  surface  energy.  

 

 

Figure  V  -­‐  10

 

Schematic   illustration   of   the   atomic   alignments   on   ZnO   (110)   and   (001)   crystal   faces.  

Inspired  from  ref  [65].  

 

We   therefore   suggest   that   the   variation   of   the   initial   contact   angle   among   samples   should  be  attributed  to  the  difference  in  their  surface  roughness.  

!!!"!!"#$%!

!!! !

!"#

!"

!!!"!!"#$%!

!!! !

(19)

 

As  a  reminder,  the  contact  angle  hysteresis  can  give  information  about  the  topography   of  the  sample  and  reflects  the  state  of  hydrophobicity  (Wenzel  or  Cassie  state  Figure  V  -­‐  

11  (a)).    

Both   samples   revealed   high   hysteresis   behavior   (Figure   V   -­‐   11   (b)),   which   is   a   characteristic  feature  of  the  Wenzel  state.    

 

According   to   the   Wenzel   model,   the   water   contact   angle   θ

W

,   is   given   by   the   following   equation

[54]

:  

cos !

!

= ! cos !     (Equation  V  -­‐  1)  

where  r  (roughness  factor)  is  the  ratio  of  the  unfolded  surface  to  the  apparent  surface   under  the  droplet  and  θ  is  the  contact  angle  on  a  flat  surface  of  the  same  nature  as  the   rough  one.  Since  r  is  always  greater  than  unity,  this  model  predicts  an  increase  of  the   contact  angle  with  the  surface  roughness.  

From   Figure   V   -­‐   9,   it   seems   evident   that   the   TG   patterning   process   increased   the   roughness   of   the   ZnO   nanowire   array   compared   to   the   unpatterned   process,   which   is   denser.   These   results   agree   with   the   study   performed   by   Das   et   al.

[45]

  on   ZnO   nanoneedles  and  nanorods  manufactured  by  metal  organic  chemical  vapor  deposition.  

Due  to  the  large  size  of  the  samples,  we  measured  the  roughness  with  a  portable  surface   roughness   tester   and   evidenced   a   subsequent   increase   (factor   2-­‐3)   of   the   surface   roughness   due   to   the   TG   patterning   process.   However,   a   precise   knowledge   of   the   substrate  roughness  is  difficult  to  measure  due  to  the  finite  size  (radius  and  aspect  ratio)   of  the  measuring  tips.    

 

 

Figure  V  -­‐  11

 

Contact  angle  hysteresis  

(a)  Characteristic  shape  of  receiding  drop  in  Wenzel  and  Cassie  states.

[70]  

(b)  Contact  angle  hysteresis  measurements  on  unpatterned  and  TG  patterned  samples   !

!

!"#$ !"#$%%&'"&()

)*'+,%-)) .&/#0$%&()

)*'+,%-))

!

!

1(2$"34"5) 1"50&)6&3&4(4"5) 1"50&)

!%#$

7&"8&0)) )9%$%&)

:$;;4&) )9%$%&)

&'(')*)+,$-.%/'$

(20)

 

By  the  way,  the  increase  of  the  nanowire  aspect  ratio  (with  higher  growth  time)  should   also  increase  the  roughness  of  the  samples.  

Besides   tuning   the   surface   morphology,   surface   treatment   with   different   chemical   (alkaloic   acids,

[70]

  octadecanethiol   solutions

[71]

  or   silanes

[72]

)   would   be   an   effective   method  to  generate  superhydrophobicity.  

 

We   then   submitted   our   samples   to   2h   of   UV   irradiation   and   as   expected,   a   transition   from  hydrophobicity  to  superhydrophilicity  (CA  of  0°)  was  observed.

[38]

 

The  origin  of  the  photoinduced  hydrophylicity  was  first  reported  by  Sun  et  al.

[73]

 Upon   UV   irradiation,   electron-­‐hole   pairs  are   generated   in   the   lattice   (Equation   V   -­‐   2).   These   electrons  and  holes  can  either  recombine  or  move  to  the  surface  to  react  with  species   adsorbed  on  the  surface  (Equation  V  -­‐  3).  

On   the   other   hand,   some   of   the   holes   can   react   with   lattice   oxygen   and   form   oxygen   vacancies  (Equation  V  -­‐  4),  while  some  of  the  electrons  react  with  lattice  metal  to  form   Zn

+

 defective  sites  (Equation  V  -­‐  5  /  surface  trapped  electrons).  

 

ZnO         +      2  hυ       ⟶      2  h

!

        +      2  e

_

 

O

!!

        +       h

!

      ⟶       O

!!

 

O

!!

        +       h

!

      ⟶       1

2 O

!

        +       V

!

 

Zn

!!

        +       e

!

      ⟶       Zn

!!

 

 

(Surface  trapped  hole)   (Oxygen  vacancy)   (Surface  trapped   electron)  

(Equation  V  -­‐  2)   (Equation  V  -­‐  3)   (Equation  V  -­‐  4)  

(Equation  V  -­‐  5)                                

Upon   generation   by   UV   irradiation,   the   surface   trapped   electrons   tend   to   react   immediately  with  oxygen  molecules  adsorbed  on  the  surface  (Equation  V  -­‐  6)  

Zn

!!

        +       O

!

      ⟶       Zn

!!!

        +       O

!!

    (Equation  V  -­‐  6)    

 

Meanwhile,  water  molecules  may  coordinate  into  the  oxygen  vacancy  sites  (V

O

)  and  this   leads   to   dissociative   adsorption   of   the   water   molecules   on   the   surface.     Indeed,   the   defective   sites   are   kinetically   more   favorable   for   hydroxyl   adsorption   than   oxygen   adsorption.     As   a   consequence,   the   hydrophobicity   of   ZnO   surface   is   significantly   improved.    

For   many   applications,   it   is   desirable   to   achieve   surfaces,   which   can   be   tuned   conveniently  between  hydrophobicity  and  hydrophilicity.  

The   reversible   generation   and   annihilation   of   photogenerated   surface   oxygen   defect  

sites   cause   the   change   of   the   surface   free   energy   and   the   alteration   of   surface  

wettability.  

Références

Documents relatifs

A set of numerical simulations of VING devices integrating ZnO NWs was per- formed considering the characteristics of typical NWs grown by each method using stand- ard

The electrical resistivity of SiNWs was determined using (i) the free hole concentration values, measured by the optical methods and (ii) two-probe electrical measurements of

In this work, we studied deposited Cobalt- doped ZnO ( ZnO:Co ) thin films via dip-coating technique onto glass substrate, Zinc acetate dehydrate, cobalt acetate, 2- methoxyethanol

The presence of such unintentional shell with a different chemical composition compared to the NW core poses problems for the design and fabrication of NW solar cells.. In

In order to extract structures from the distance matrix, we can build country networks through filtering the time delayed correlation matrix by removing the least correlated links,

Interestingly, the absorptance of the arrays comprising both the ZnO/CdTe nano-fibre and CdTe cap (depicted in Fig. Since the only distinction between the two structures

Samples were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), energy dispersive X-ray analysis (EDX) and Electrochemical impedance spectroscopy (EIS), to

In this paper, we report the synthesis of ZnO nanogranules by a solid state mixing of zinc nitrate tetrahydrate ZnO(NO3)2 with citric acid monohydrate C6H8O7, followed by