Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Generation of a long acting GCSF

2016

Generation  of  a  long  acting  GCSF         Abdulrahman  T  Alshehri   Department  of  Oncology  and  Metabolism     The  University  of  Sheffield   This  thesis  is  submitted  for  the  degree  of   Doctor  of  Philosophy   Jan  2016     TABLE  OF  CONTENTS     List  of  Figures  .................................................................................................................................  ix   List  of  Tables  ...................................................................................................................................  xi   Declaration  ....................................................................................................................................  xiii   PUBLICATIONS  AND  PRESENTATIONS  ...........................................................................  xiv   ACKNOWLEDGMENT  .................................................................................................................  xv   Abstract  ..........................................................................................................................................  xvi   1.   Introduction  .................................................................................................................................  1   1.1   History  of  Granulocyte  Colony  Stimulating  Factor  .............................................  1   1.2   GCSF  Structure  ...................................................................................................................  2   1.3   GCSF  Expression  and  Action  .........................................................................................  4   1.4   Regulation  of  GCSF  Expression  ...................................................................................  5   1.5   GCSF-­‐Receptor  ....................................................................................................................  7   1.5.1   Discovery,  Expression  and  Cloning  ..................................................................  7   1.5.2   Structure  and  Function  of  GCSF-­‐R  ....................................................................  7   1.5.2.1   Mobilization  of  Neutrophils  .........................................................................  12   1.6   The  Major  Clinical  Use  of  GCSF  .................................................................................  14   1.6.1   Febrile  Neutropenia  Prophylaxis  ...................................................................  14   1.6.2   Mobilization  of  Stem  Cells  .................................................................................  14   1.6.3   Controlling  of  SCN  and  AML  .............................................................................  14   1.7   The  Main  Side  Effects  of  GCSF  Administration  ..................................................  15   1.8   Available  Commercial  GCSF  Preparations  and  Their  Limitations  ............  18   1.8.1   Filgrastim  (NEUPOGEN®)  ..................................................................................  18   1.8.2   Lenograstim  (Granocyte®)  ................................................................................  18   1.9   Strategies  Used  to  Delay  the  Clearance  of  GCSF  ...............................................  19   ii   1.9.1   Extension  of  Half-­‐life  by  Increasing  the    Molecular  Weight  ................  19   1.9.1.1   PEGylation  ...........................................................................................................  19   1.9.2   Extension  of  Half-­‐life  Using  the  FcRn-­‐Mediated  Recycling  .................  22   1.9.2.1   Fusion  of  GCSF  to  Albumin  ...........................................................................  24   1.9.2.2   Fusion  of  GCSF  to  IgG-­‐Fc  ................................................................................  24   1.9.3   New  Approach  by  Asterion  ...............................................................................  27   1.9.3.1   Ligand/Receptor  Fusion  ................................................................................  27   1.9.3.2   Glycosylation  ......................................................................................................  28   O-­‐linked  Glycosylation  ....................................................................................................  29   N-­‐linked  Glycosylation  ....................................................................................................  29   1-­‐  Hyperglycosylation  via  Site  Direct  Mutagenesis  ....................................  33   2-­‐  Glycosylated  Linkers  ...........................................................................................  34   1.10   Aim  and  Hypothesis  ......................................................................................................  36   2.   Materials  ....................................................................................................................................  38   2.1   Cell  Culture  ........................................................................................................................  38   2.2   DNA  Manipulation  .........................................................................................................  39   2.2.1   Restriction  Endonucleases  ................................................................................  40   2.2.1.1   Enzyme  ..................................................................................................................  40   2.2.2   Bacterial  Cell  Culture  ...........................................................................................  40   2.2.2.1   Antibiotics  ............................................................................................................  40   2.2.2.2   Media  ......................................................................................................................  40   2.3   Protein  Analysis  ..............................................................................................................  41   2.3.1   3.   Proliferation  Assay  ...............................................................................................  43   General  Methods  .....................................................................................................................  44   3.1   Preparation  of  Luria-­‐Bertani  (LB)  Media  ............................................................  44   3.2   Preparation  of  Agar  Plates  .........................................................................................  44   3.3   Preparation  of  Chemically  Competent  Cells  .......................................................  44   iii   3.4   DNA  Cloning  of  GCSF  Tandems  for  Expression  .................................................  45   3.4.1   Polymerase  Chain  Reaction  (PCR)  .................................................................  45   3.4.2   DNA  Isolation  from  Agarose  Gel  Electrophoresis  ...................................  47   3.4.3   Single  and  Double  Restriction  Enzyme  Digests  .......................................  47   3.4.4   General  DNA  Ligation  ..........................................................................................  50   3.4.5   Transformation  of  Plasmid  into  Chemically  Competent  E.coli  ..........  51   3.4.6   Plasmid  Preparation  and  Glycerol  Stocks  ...................................................  51   3.4.7   Screening  of  Potential  Clones  from  Ligations  ...........................................  52   3.4.8   Plasmid  Sequence  Analysis  ...............................................................................  52   3.5   Cell  Culture  and  Protein  Expression  ......................................................................  52   3.5.1   Growth  and  General  Maintenance  of  CHO  Flp-­‐In  Cells  .........................  52   3.5.2   Trypan  Blue  Exclusion  Method  .......................................................................  53   3.5.3   Transient  Expression  of  GCSF  Tandems  in  CHO  Flp-­‐In  ........................  53   3.5.4   Generation   o f   S table   C HO   F lp-­‐In   C ell   l ines  ........................................  54   3.5.5   Analysis  of  Crude  Media  from  Transfected  Cells  Lines  .........................  56   3.5.6   Storage  of  Stable  Cell  Lines  in  Liquid  Nitrogen  ........................................  57   3.5.7   Adaptation  of  Stable  CHO  Cells  to  Hyclone  Media  ..................................  57   3.5.8   Expression  of  GCSF  Tandems  in  Roller  Bottle  Culture  .........................  57   3.6   Vivaflow  200  Concentrator  ........................................................................................  58   3.7   Protein  Purification  .......................................................................................................  58   3.7.1   Purification  of  GCSF  Tandems  Using  IMAC  ................................................  58   3.7.2   Purification  of  GCSF  Using  Cibacron  Blue  Sepharose  ............................  60   3.8   Analysis  of  Protein  .........................................................................................................  61   3.8.1   Bradford  Protein  Assay  ......................................................................................  61   3.8.2   Analysis  of  Proteins  by  SDS-­‐  PAGE  ................................................................  62   3.8.2.1   Preparation  of  SDS-­‐PAGE  Gels  ....................................................................  62   3.8.2.2   Preparation  of  Samples  for  SDS-­‐PAGE  ....................................................  63   iv   3.8.2.3   Visualized  Protein  Gels  with  Coomassie  Blue  ......................................  63   3.8.3   Western  Blotting....................................................................................................  64   3.8.3.1   Transfer  of  Proteins  to  PVDF  Membrane  ...............................................  64   3.8.3.2   Western  Blotting  Detection  of  GCSF  .........................................................  65   3.8.4   Enzyme  Linked  Immunosorbent  Assay  (ELISA)  ......................................  66   3.8.5   AML-­‐193  Proliferation  Assay  ...........................................................................  69   3.8.5.1   Growth  of  the  AML-­‐193  Cell  Line  ..............................................................  69   3.8.5.2   AML-­‐193  Bioassay  ............................................................................................  69   3.8.6   3.9   Short  Term  Stability  of  GCSF  Tandem  Molecules  ....................................  71   Experimental  Procedure  for  In  vivo  Study  ..........................................................  71   3.10   Statistical  Analysis  .........................................................................................................  72   4.   Cloning  and  Expression  of  GCSF  Tandems  ..................................................................  73   4.1   Summary  ............................................................................................................................  73   4.2   Introduction  ......................................................................................................................  74   4.2.1   Aim  ...............................................................................................................................  77   4.2.2   Objectives  .................................................................................................................  77   4.3   Construction  of  GCSF  Tandems  ................................................................................  78   4.3.1   Construction  of  pSecTag  GCSF-­‐L1_Hist  .......................................................  80   4.3.2   Construction  of  pSecTagGCSF-­‐L2_Hist  .........................................................  84   4.4   Generating  GCSF  Tandems  with  Variable  Linkers  ...........................................  90   4.5   Expression  and  Analysis  of  GCSF  Tandems  ........................................................  92   4.5.1   Transient  Transfection  of  CHO  Flp-­‐In  Cells  ...............................................  92   4.5.1.1   Analysis  of  Expression  by  Elisa  ..................................................................  92   4.5.1.2   Analysis  of  Expression  by  Western  Blotting  .........................................  93   4.5.2   Stable  Cell  Line  Development  in  CHO  Flp-­‐In  Cell  Lines  ........................  95   4.5.2.1   Analysis  of  GCSF  Tandems  by  Elisa  ...........................................................  95   4.5.2.2   Analysis  of  GCSF  Tandems  by  Western  Blotting  .................................  96   v   5.   4.6   In  vitro  Biological  Activity  of  Crude  Media  ..........................................................  97   4.7   Discussion  ..........................................................................................................................  99   Large-­‐scale  Production  and  Analysis  of  GCSF  Tandems  .....................................  102   5.1   Summary  ..........................................................................................................................  102   5.2   Introduction  ....................................................................................................................  103   5.2.1   5.3   Aim  .............................................................................................................................  103   Results  ...............................................................................................................................  104   5.3.1   Cell  Growth  and  Productivity  .........................................................................  104   5.3.2   Purification  of  GCSF  Tandems  Using  IMAC  ..............................................  109   5.3.2.1   Purification  of  GCSF2NAT  ...........................................................................  109   5.3.2.2   Purification  of  GCSF2QAT  ...........................................................................  113   5.3.2.3   Purification  of  GCSF4NAT  ...........................................................................  116   5.3.2.4   Purification  of  GCSF4QAT  ...........................................................................  119   5.3.2.5   Purification  of  GCSF8NAT  ...........................................................................  122   5.3.2.6   Purification  of  GCSF8QAT  ...........................................................................  127   5.3.2.7   Summary  of  GCSF  Protein  Tandems  Purification  .............................  130   5.4   6.   Discussion  ........................................................................................................................  131   In  vitro  Bioactivity  Evaluation  and  Temperature  Stability  of  GCSF………….135   6.1   Summary  ..........................................................................................................................  135   6.2   Introduction  ....................................................................................................................  136   6.2.1   6.3   Aim  .............................................................................................................................  137   Results  ...............................................................................................................................  138   6.3.1   In  vitro  Bioactivity  Evaluation  .......................................................................  138   6.3.1.1   In  vitro  Biological  Activity  of  GCSF2NAT  and  Its  Control  ..............  139   6.3.1.2   In  vitro  Biological  Activity  of  GCSF4NAT  and  Its  Control  ..............  141   6.3.1.3   In  vitro  Biological  Activity  of  GCSF8NAT  and  Its  Control  ..............  143   6.3.2   Short  Term  Stability  of  GCSF  Tandem  Molecules  ..................................  145   vi   6.3.2.1   Protein  Samples  from  the  Stability  Experiment  in  the  AML  ........  145   6.3.2.2   Temperature  Stability  of  GCSF  Tandem  Molecules  .........................  146   6.4   7.   Discussion  ........................................................................................................................  153   Pharmacokinetic  &  Pharmacodynamic  Analysis  of  GCSF  Tandems    ..............  159   7.1   Summary  ..........................................................................................................................  159   7.2   Introduction  ....................................................................................................................  160   7.3   Aim  .....................................................................................................................................  162   7.4   Results  ...............................................................................................................................  163   7.4.1   Preliminary  Test  for  the  Effect  of  Rat’s  Serum  on  Elisa  Assay  ........  163   7.4.2   Preliminary  Pharmacokinetic  Analysis  in  Sprague  Dawley  Rats  ...  165   7.4.2.1   Elisa  Results  ......................................................................................................  165   7.4.2.1.1   Pharmacokinetics  Analysis  of  rhGCSF  in  Normal  Rats  ...........  166   7.4.2.1.2   Pharmacokinetic  Analysis  of  GCSF2NAT  in  Normal  Rats  ......  168   7.4.2.1.3   Pharmacokinetic  Analysis  of  GCSF4NAT  in  Normal  Rats  ......  170   7.4.2.1.4   Pharmacokinetics  Analysis  of  GCSF8NAT  in  Normal  Rats  ...  171   7.4.2.1.5   Pharmacokinetic  Analysis  of  GCSF8QAT  in  Normal  Rats  ......  173   7.4.2.1.6   Pharmacokinetic  Analysis  of  GCSF  Tandems  in  Rats  ..............  174   7.4.2.2   Terminal  Half-­‐life  Analyses  of  GCSF  Tandems  ...................................  175   7.4.2.3   Pharmacodynamics  of  GCSF  Tandems  ..................................................  178   7.5   8.   Discussion  ........................................................................................................................  180   General  Discussion  ..............................................................................................................  187   8.1   Future  Work  ...................................................................................................................  190   8.1.1   9.   Future  Work  to  Improve  GCSF  Tandems  ..................................................  191   Conclusion  ...............................................................................................................................  193   Appendix  A  ...................................................................................................................................  194   Appendix  A.1.  Nucleotide  Sequences  of  Primers  .........................................................  194   Appendix  A.2.  Restriction  Endonucleases  Cut  Sites  ...................................................  194   vii   Appendix  B  ...................................................................................................................................  195   Appendix  B.1.  Nucleotide  and  Amino  Acid  Sequences  of  GH  Tandem  ...............  195   Appendix  B.2.  Nucleotide  and  Amino  Acid  Sequences  of  GCSF  Tandem  ..........  196   Appendix  B.3.  Nucleotide  and  Amino  Acid  Sequences  of  Linker  Regions  ........  197   Appendix  C.  pSecTag_Link-­‐Hist  Modulating  Vector:  ..................................................  198   Bibliography  ................................................................................................................................  199         viii   List  of  Figures   Figure    1-­‐1:  Human  GCSF  structure  .............................................................................................  3   Figure    1-­‐2:  Regulation  pathways  of  GCSF  expression  and  production  ……………….6   Figure    1-­‐3:  Structure  and  downstream  signal  pathways  of  GCSF-­‐R  .............................  9   Figure    1-­‐4:  Pathway  upon  activation  of  Jak-­‐STAT  signals  .............................................  11   Figure    1-­‐5:  Scheme  of  the  proposed  model  ..........................................................................  13   Figure    1-­‐6:  Comparison  of  carboxyl-­‐terminal  region  of  the  GCSF-­‐R  ………………...16   Figure    1-­‐7:  Model  of  the  pH-­‐dependent  recycling  mechanism  of  albumin………..23   Figure    1-­‐8:  The  schematic  diagram  shows  (A)  G-­‐CSF/IgG-­‐Fc  protein  and  (B)  .....  26   Figure    1-­‐9:  Glycan  structure  .......................................................................................................  31   Figure    1-­‐10:  An  example  of  2NAT  glycosylation  motifs…………………………………...37   Figure    3-­‐1:  Design  of  transferring  assembly  ........................................................................  65   Figure    4-­‐1:  The  diagram  summarizes  the  process  of  producing  GCSF  tandems...79   Figure    4-­‐2:    PCR  of  GCSF-­‐L1  ........................................................................................................  81   Figure    4-­‐3:  Double  digest  of  pSecTagGH2NAT_Hist  .........................................................  82   Figure    4-­‐4:  Double  digest  of  pSecTagGCSF2NAT_Hist-­‐L1  potential  clones  ............  83   Figure    4-­‐5:  Generation  of  PCR  fragment  GCSF-­‐L2  .............................................................  85   Figure    4-­‐6:  Double  digestion  of  pSecTagGCSF2NAT_Hist_L1  .......................................  86   Figure    4-­‐7:  Double  digest  of  pSecTagGCSF2NAT_Hist  potential  clones  ...................  87   Figure    4-­‐8:  Removal  of  GCSF  L2  from  pGCSFsecTagGCSF2NAT_Hist  .......................  88   Figure    4-­‐9:  Double  digest  of  pSecTag_link_GCSF-­‐L2  potential  clones  ......................  89   Figure    4-­‐10:  Double  digest  potentially  positive  clone  of  GCSF4NAT_Hist.  .............  91   Figure    4-­‐11:  Double  digest  potentially  positive  clone  of  GCSF8QAT_Hist  ..............  91   Figure    4-­‐12:  Western  blot  of  media  samples  from  transiently  transfected  CHO..94   ix   Figure    4-­‐13:  Western  blot  of  stable  CHO  Flp-­‐In  cell  media  expressing  GCSF  ……96   Figure    4-­‐14:  In  vitro  biological  activity  for  rhGCSF  and  GCSF  tandems  ...................  98   Figure    5-­‐1:  GCSF  tandems  expressing  cells  growth  and  productivity  ....................  106   Figure    5-­‐2:  Western  blot  analysis  of  roller  bottle  media  samples  ............................  108   Figure    5-­‐3:  purification  development  of  IMAC  for  GCSF2NAT  ...................................  111   Figure    5-­‐4:  Purification  development  of  IMAC  for  GCSF2QAT  ...................................  114   Figure    5-­‐5:  Purification  analysis  of  IMAC  for  GCSF4NAT  .............................................  117   Figure    5-­‐6:  Purification  analysis  of  IMAC  for  GCSF4QAT  .............................................  120   Figure    5-­‐7:  Western  blot  analysis  of  roller  bottle  samples  for  GCSF8NAT………122   Figure    5-­‐8:  Western  blot  analysis  of  roller  bottle  samples  for  GCSF8NAT  ...........  123   Figure    5-­‐9:  Purification  analysis  of  IMAC  samples  for  GCSF8NAT  ...........................  125   Figure    5-­‐10:  Purification  development  of  IMAC  for  GCSF8QAT  ................................  128   Figure    5-­‐11:  Purified  GCSF  tandems  analysed  by  SDS-­‐PAGE  and  western  blot  .  130   Figure    6-­‐1:    AML-­‐193  cells  in  the  presence  of  tandems  and  rhGCSF  .......................  140   Figure    6-­‐2:  AML-­‐193  cells  in  the  presence  of  tandems  and  rhGCSF  ........................  142   Figure    6-­‐3:  AML-­‐193  cells  in  the  presence  of  tandems  and  rhGCSF  ........................  144   Figure    6-­‐4:  Stability  of  GCSF  tandems  after  incubation  with  AML-­‐193  cells  .......  145   Figure    6-­‐5:  Temperature  stability  of  GCSF2NAT  .............................................................  147   Figure    6-­‐6:  Temperature  stability  of  GCSF2QAT  ..............................................................  148   Figure    6-­‐7:  Temperature  stability  of  GCSF4NAT  .............................................................  149   Figure    6-­‐8:  Temperature  stability  of  GCSF4QAT  ..............................................................  150   Figure    6-­‐9:  Temperature  stability  of  GCSF8NAT  .............................................................  151   Figure    6-­‐10:  Temperature  stability  of  GCSF8QAT  ...........................................................  152   Figure    7-­‐1:  Effect  of  rat  serum  on  the  sensitivity  of  Elisa  assay  ................................  164   Figure    7-­‐2:  Elisa  analysis  of  rhGCSF  pharmacokinetics  in  normal  rat  ....................  167   x   Figure    7-­‐3:  Elisa  analysis  of  GCSF2NAT  pharmacokinetics  in  normal  rat  ............  169   Figure    7-­‐4:  Elisa  analysis  of  GCSF4NAT  pharmacokinetics  in  normal  rat  ............  170   Figure    7-­‐5:  Elisa  analysis  of  GCSF8NAT  pharmacokinetics  in  normal  rat  ............  172   Figure    7-­‐6:  Elisa  analysis  of  GCSF8QAT  pharmacokinetics  in  normal  rat.  ............  173   Figure    7-­‐7:  Elisa  analysis  of  GCSF  tandems  in  normal  rat  models  ...........................  174   Figure    7-­‐8:  Tandem  GCSF  proteins  terminal  half-­‐life  analyses……………………….177   Figure    7-­‐9:  Percentage  change  in  blood  neutrophils  following  intravenous……179   List  of  Tables   Table  1-­‐1:  The  main  side  effects  reported  for  patients  treated  with  GCSF  .............  17   Table  1-­‐2:  Summary  of  pros  and  cons  of  PEGylation  and  other    .................................  21   Table  1-­‐3:  List  of  GCSF  tandems  ................................................................................................  37   Table  3-­‐1:  PCR  reaction  utilizing  two-­‐master  mixes  ........................................................  46   Table  3-­‐2:  PCR  stages  .....................................................................................................................  46   Table  3-­‐3:  Single  step  double  digestion  of  Insert  ...............................................................  48   Table  3-­‐4:  Two-­‐step  double  digestion  of  Plasmid  DNA  ...................................................  49   Table  3-­‐5:  Preparation  of  ligation  reactions  ........................................................................  50   Table  3-­‐6:    Concentrations  of  imidazole  ................................................................................  60   Table  3-­‐7:  Standard  curve  preparation  ..................................................................................  62   Table  3-­‐8:  Preparation  of  GCSF  standards  ............................................................................  67   Table  3-­‐9:  Initial  stock  concentrations  of  rhGCSF  and  GCSF  tandem  proteins  .....  67   Table  3-­‐10:  Preparation  of  GCSF  standard  curve  ...............................................................  70   Table  4-­‐1:  The  structure  of  GCSF  tandems  with  modified  flexible  linkers    ............  76   Table  4-­‐2:    Description  of  the  two  plasmids  that  were  used  to  produce    ................  80   Table  4-­‐3:  GCSF  sandwich  Elisa  analysis  of  transiently  expressed  tandem    ..........  93   xi   Table  4-­‐4:  Results  of  GCSF  sandwich  Elisa  for  stably  expressed  GCSF  tandems  ..  95   Table  4-­‐5:  Determined  and  observed  MW’s  of  expressed  GCSF  tandems  ...............  97   Table  5-­‐1:  Concentrations  of  GCSF2NAT  during  the  purification  process  ............  112   Table  5-­‐2:  Protein  concentrations  of  GCSF2QAT  during  the  purification  ……….115   Table  5-­‐3:  Protein  concentrations  of  GCSF4NAT  during  the  purification  ……….118   Table  5-­‐4:  Concentrations  of  GCSF4QAT  during  the  purification  process  ............  121   Table  5-­‐5:  Concentrations  of  GCSF8NAT  during  the  purification  process  ............  126   Table  5-­‐6:  Protein  concentrations  of  GCSF8QAT  during  the  purification  ……….129   Table  7-­‐1:  Tandem  GCSF  Proteins  terminal  Half-­‐life  analyses………………………..176       xii   Declaration   I  hereby  declare  that  this  thesis  has  been  composed  by  myself  and  has  not  been   accepted  in  any  previous  application  for  a  higher  degree.  The  work  reported  in   this   thesis   is   novel   and   has   been   carried   out   by   myself   with   all   source   of   information  being  specifically  acknowledged  by  means  of  references.     Abdulrahman  Alshehri     January  2016         xiii   PUBLICATIONS  AND  PRESENTATIONS   Published  Abstracts  and  Posters:   -­‐  School  Research  Meeting  Conference,  Sheffield  University  June  2012   ‘Generation  of  a  long  acting  GCSF  for  treatment  of  neutropenia  and  stem  cell   harvest  -­‐  Abdulrahman  Alshehri,  Richard  Ross  and  Ian  R  Wilkinson’   -­‐ Society  for  Endocrinology  BES  Meeting  Liverpool,  March  2014   ‘Generation  of  a  long  acting  GCSF  for  treatment  of  neutropenia  and  stem  cell   harvest  -­‐  Abdulrahman  Alshehri,  Richard  Ross  and  Ian  R  Wilkinson’   -­‐ Advanced  Biomanufacturing  Conference,  Sheffield  May  2015   ‘Generation  of  a  long  acting  GCSF  for  treatment  of  neutropenia  and  stem  cell   harvest  -­‐  Abdulrahman  Alshehri,  Richard  Ross  and  Ian  R  Wilkinson’   -­‐ Society  for  Endocrinology  BES  Meeting  Edinburgh,  November  2015   (Awarded  the  Best  junior  poster  prize)   ‘Generation  of  a  long  acting  GCSF  for  treatment  of  neutropenia  and  stem  cell   harvest  -­‐  Abdulrahman  Alshehri,  Richard  Ross  and  Ian  R  Wilkinson’     Medical  School  Presentations  and  Mellanby  Centre  Internal  Seminars:     First  year  PhD  presentation,  July  2012   Third  year  PhD  presentation,  July  2014 Mellanby  Centre  Internal  Seminar,  December  2012   Mellanby  Centre  Internal  Seminar,  December  2013       xiv   ACKNOWLEDGMENT   I   would   like   to   acknowledge   the   support   and   help   of   many   colleagues.   In   particular,   I   would   like   to   thank   Professor   Richard   Ross   for   his   continuous   support   and   guidance   during   this   project.   His   support   was   not   only   to   develop   me  as  a  scientist  but  also  as  an  individual.  I  am  indebted  to  Dr  Ian  Wilkinson  who   provided   excellent   supervision   and   critical   guidance   throughout   this   project.   Also,   I   would   like   to   thank   him   for   the   difficult   time   of   correcting   this   thesis.   Without  him  the  project  would  surely  not  have  progressed  so  successfully.     I   would   like   to   thank   Dr   Hamid   Zarkesh   for   his   help   doing   the   rat   injection,   counting   WBCs   using   an   automated   coulter   counter,   blood   smears   and   sample   collection.  Also,  I  would  like  to  thank  Dr  Miguel  Debonno  for  his  help  to  calculate   in  vivo  data  using  Win  Non  Lin.     I  am  extremely  grateful  to  Dr  Sarbandra  Pradhananga,  Sue  Justice,  Hadel  Ghaban,   Mahmoud   Habibullah   and   Jude   Akinwale   for   all   their   help   and   time   during   laboratory  work.     I   would   like   to   thank   my   wife,   daughters,   mother,   father,   brothers,   sisters   and   all   friends  for  their  endless  support  and  believe  in  me.   Also,  I  am  grateful  to  thank  the  Prince  Mohammed  Bin  Naif  (Ministry  of  Interior)   and  cultural  Bureau  of  Saudi  Arabia  for  their  financial  support.           xv   Abstract   Rationale:   Current   therapies   require   daily   injections   of   GCSF   to   treat   patients   with   neutropenia   and   response   to   treatment   is   often   unpredictable   as   GCSF   is   rapidly  cleared.  A  number  of  approaches  to  reducing  GCSF  clearance  have  been   tried  mainly  through  conjugation  with  another  moiety.  The  technologies  already   being   employed,   include   PEGylation,   immunoglobulins   or   albumin   to   increase   the   half‐life   of   GCSF.   However,   although   these   approaches   have   reduced   clearance  the  pharmacokinetic  profile  of  GCSF  has  remained  unpredictable.     Aim   and   Hypothesis:   a   glycosylated   linker   between   two   ligands   could   delay   clearance  with  out  blocking  bioactivity.     Methodology:   GCSF   tandem   molecules   with   linkers   containing   between   2-­‐8   N-­‐ linked  glycosylation  sites  (NAT  motif)  and  their  respective  controls  (Q  replaces  N   in   the   sequence   motif   NAT   so   there   is   no   glycosylation)   were   cloned,   and   sequenced.  Following  expression  in  CHO  cells,  expressed  protein  was  quantified   by  ELISA  and  analysed  by  western  blot  to  confirm  molecular  weights  and  protein   integrity.   In   vitro   bioactivity   was   tested   using   an   AML-­‐193   proliferation   assay.   IMAC  was  used  to  purify  the  protein.  Pharmacokinetic  and  pharmacodynamic  of   GCSF  tandems  were  measured  in  Sprague  Dawley  rats.   Results:   Purified   glycosylated   tandem   molecules   showed   increased   molecular   weight   according   to   the   number   glycosylation   of   their   sites   when   analysed   by   SDS-­‐PAGE.  All  GCSF  tandems  showed  increased  in  vitro  bioactivity  in  comparison   to   rhGCSF.   GCSF2NAT,   GCSF4NAT   and   GCSF8NAT   containing   2,   4   &   8   glycosylation  sites  respectively  and  GCSF8QAT  displayed  a  three-­‐fold  increased   terminal   half-­‐life   compared   to   that   published   for   GCSF,   however   there   was   no   difference   in   serum   half-­‐life   according   to   the   level   of   glycosylation.   Both   GCSF2NAT   and   GCSF4NAT   showed   a   higher   increase   in   the   percentage   of   neutrophils   over   controls   at   12   hrs   post   injection   only.   In   contrast,   GCSF8NAT   exhibited  a  higher  increase  in  neutrophil  levels  over  controls  at  48  hrs.   Conclusion:   Using   glycosylated   linkers   in   GCSF   tandems   results   in   molecules   xvi   with  increased  molecular  weight  according  to  the  number  of  glycosylation  sites.   Tandems   of   GCSF   have   increased   in   vitro   bioactivity   compared   to   monomeric   GCSF.  Tandems  with  and  without  glycosylation  had  three-­‐fold  greater  half-­‐lives   than  rhGCSF.  There  was  evidence  that  GCSF8NAT  was  biologically  active  in  vivo.     The  results  confirm  the  hypothesis  that  it  is  possible  to  predictably  increase  the   molecular   weight   of   GCSF   tandems   and   retain   biological   activity   but   this   was   not   associated   with   a   predictable   prolongation   of   the   serum   half-­‐life. xvii   1. Introduction     1.1 History  of  Granulocyte  Colony  Stimulating  Factor     Prior  to  the  1960s,  several  studies  on  animal  models  had  been  performed  to  find   the   answer   to   how   white   blood   cell   (WBC)   homeostasis   can   be   regulated   in   the   circulation.   The   specific   regulator   remained   unknown   until   1966,   when   two   groups   simultaneously   performed   a   method   for   growing   colonies   of   monocytes   and  granulocytes  from  spleen  cells,  and  bone  marrow  (BM)  in  semi  sold  cultures   in  vitro  (Bradley  and  Metcalf,  1966,  Ichikawa  et  al.,  1966).  However,  the  growth  of   these   colonies   was   dependent   on   the   presence   of   unknown   proteins   that   were   given   the   name   of   colony   stimulating   factors   (CSFs).     Since   the   middle   of   the   1980s,   efforts   have   been   made   by   several   laboratories   to   identify   and   purify   these   CSF  proteins.  These  efforts  revealed  that  there  are  four  CSF  proteins  with  different   activities.   They   were   named   dependent   on   the   type   of   colony   of   cells   they   stimulated:   M-­‐CSF   stimulated   macrophage   colonies;   GM-­‐CSF   stimulated   both   granulocyte   and   macrophage   colonies;   G-­‐CSF   stimulated   granulocyte   colony   formation   and   multi-­‐CSF   (known   as   interleukin   3,   IL3)   stimulated   multiple   of   hematopoietic  cell  colonies  (Metcalf,  2010).   In   1983,   Murine   GCSF   was   first   purified   from   mouse   lung-­‐conditioned   medium   by   Nicola  and  colleagues  in  Melbourne,  Australia  (Nicola  et  al.,  1983),  while  Human   GCSF   (hGCSF)   was   first   purified   from   the   human   bladder   carcinoma   cell   line   5637   in   1984   (Welte   et   al.,   1985).   The   molecular   cloning   of   complementary   deoxyribonucleic  acid  (cDNA)  for  GCSF  and  the  first  expression  from  E.  coli  were   attained  by  Souza  and  Boone  in  1986  (Souza  et  al.,  1986).           1   1.2 GCSF  Structure   It   has   been   reported   that   the   hGCSF   is   encoded   by   a   gene   located   on   chromosome   17   (known   as   CSF3   gene)   and   due   to   differential   splicing   of   GCSF,   this   gene   encodes   two   different   messenger   ribonucleic   acids   (mRNA)   products:   isoform   A   contains   177   amino   acids   (18.8kD)   and   isoform   B   contains   174   amino   acids   (19.6kD).   The   difference   between   these   isoforms   is   that   isoform   A   contains   an   additional   three   residues   (Valine-­‐Serine-­‐Glycine)   added   after   Leucine35.   The   short  isoform  B  (174  amino  acids)  contains  a  glycosylation  site  on  the  oxygen  (O-­‐   linked   glycosylation)   attached   to   one   threonine   at   residue   133   as   the   form   expressed  in  mammalian  cells.  The  B  isoform  obtains  high  stability  and  biological   activity   and   therefore   is   the   source   for   commercial   pharmaceutical   products   of   GCSF  (Aapro  et  al.,  2011).     The   central   structure   of   hGCSF   contains   four   antiparallel,   left-­‐handed   α-­‐helical   bundles   in   a   form   that   two   helices   (A   with   29   amino   acids   &   B   with   21   amino   acids)  extend  up  and  two  helices  (C  with  24  amino  acids  &  D  with  30  amino  acids)   extend   down   (Figure   1.1)   (Arvedson   and   Giffin,   2012).   Additionally,   hGCSF   contains  five  cysteine  residues;  Four  of  these  cysteines  form  two  internal  disulfide   bonds  at  positions  Cys36–  Cys42  and  Cys64–  Cys74,  thus  leaving  one  free  cysteine   residue  at  position  Cys17  with  a  free  sulfhydryl  group  (Werner  et  al.,  1994).       2       Figure  1-­‐1:  Human  GCSF  structure   The  molecular  structure  of  hGCSF  contains  four  antiparallel,  left-­‐handed  α-­‐helical   bundles  in  a  form  that  two  helices  A  (Red)  &  B  (Orange)  extend  up  and  two  helices   C   (White)   &   D   (Cyne)   extending   down.   C=   carboxyl-­‐terminus   and   N=   amine-­‐ terminus  (Using  PyMOL  molecular  graphics  system).     3     1.3 GCSF  Expression  and  Action   Human  GCSF  is  a  glycoprotein  that  regulates  the  proliferation,  differentiation  and   functional   activation   of   granulopoiesis   (Cox   et   al.,   2014)   since   proved   by   the   significant   decrease   of   neutrophils   in   both   GCSF   and   GCSF-­‐R   deficient   mice   (Lieschke  et  al.,  1994,  Liu  et  al.,  1996).  In  response  to  several  inflammatory  factors   such   as,   interleukin   β   (IL-­‐1β)   necrosis   factor   alpha   (TNF-­‐α)   and   lipopolysaccharide   (LPS),   GCSF   can   be   produced   by   a   variety   of   cells,   including   endothelial  cells,  fibroblasts,  macrophages,  monocytes,  and  bone  marrow  stromal   cells.  GCSF  has  recently  been  shown  to  be  highly  expressed  on  a  number  of  cancer   cell  types  including  human  gastric  and  colon  cancers  (Gascon,  2012,  Morris  et  al.,   2014),   as   well   as   acute   myeloid   leukemia   (AML)   and   other   carcinoma   cells   (Beekman  et  al.,  2012).     In   healthy   individuals,   GCSF   is   present   at   low   levels   and   is   rapidly   increased   in   severe   cases   such   as,   infection   (20   times   increase)   (Cheers   et   al.,   1988,   Kawakami   et  al.,  1990).  Therefore,  the  physiological  role  of  GCSF  in  the  body  is  to  maintain   the   production   of   neutrophils   during   steady   state   situations   and   increase   neutrophil   production   during   severe   inflammatory   conditions   such   as,   infection   (Hartung  et  al.,  1999).       Several  mechanisms  showed  that  GCSF  can  enhance  and  regulate  the  production   of   neutrophils   from   the   bone   marrow   to   the   blood   circulation.   It   enhances   the   proliferation   of   all   granulocytic   lineages   from   myloblast   (stem   cell)   to   mylocyte.   It   also   drives   neutrophil   differentiation   and   accelerates   the   maturation   of   metamyelocytes.  As  results  of  these  functions,  GCSF  shows  rapid  and  continuous   elevation  in  the  number  of  neutrophils  (Lord  et  al.,  1991,  Basu  et  al.,  2002).         4   1.4 Regulation  of  GCSF  Expression   During   infection,   several   inflammatory   factors   in   the   extracellular   microenvironment  are  elevated,  such  as  IL-­‐1β,  TNF-­‐α  and  LPS  and  thereafter  act   on   target   cells   to   stimulate   GCSF   expression   by   intracellular   signaling   transcriptional   factors,   such   as   NF-­‐κB   and   C/EBP   β.   The   GCSF   promoter   region   contains   binding   sites   for   these   factors,   which   in   turn   stimulate   the   GCSF   production   (Figure   1.2   right).     The   circulatory   levels   of   GCSF   enhance   the   production   and   mobilization   of   neutrophils   from   the   bone   marrow   to   the   blood   circulation  (Panopoulos  and  Watowich,  2008).     In  addition  to  this  pathway,  recent  studies  show  that  IL-­‐23  produced  by  dendritic   cells  and  macrophages,  induces  T  helper  17  (Th17)  to  synthesis  IL-­‐17.  IL-­‐17  then   drives   the   production   of   GCSF   from   cells   contained   in   the   stroma,   such   as   endothelial   cells,   epithelial   and   fibroblasts   through   the   IL-­‐17   receptor   (IL-­‐17R)   (Fossiez   et   al.,   1996,   Langrish   et   al.,   2005).     In   response   to   bacterial   pneumonia   infection,   deficiency   of   IL-­‐17R   resulted   in   decreased   levels   of   GCSF   and   delayed   neutrophils   production,   indicating   the   important   role   of   IL-­‐17-­‐induced   granulopoiesis   in   vivo   (Figure   1.2   left)   (Ye   et   al.,   2001,   Nguyen-­‐Jackson   et   al.,   2010).       5     Figure   1-­‐2:   Regulation   pathways   of   GCSF   expression   and   production   of   neutrophils         6   1.5 GCSF-­‐Receptor   1.5.1 Discovery,  Expression  and  Cloning     The   biological   action   of   GCSF   is   mediated   by   binding   to   its   receptor   (GCSF-­‐R).   Thus,  the  regulation,  proliferation  and  differentiation  of  neutrophilic  granulocytes   are  highly  dependent  on  this  binding  (Gascon,  2012).  GCSF-­‐R  was  discovered  as  a   membrane   protein   expressed   in   all   granulocytic   lineage   cells,   including   neutrophils  and  their  precursors,  and  myeloid  leukemia  cells  (Nicola  and  Metcalf,   1984).   Later,   GCSF-­‐R   was   detected   on   normal   B   &   T   lymphocytes,   monocytes   (Boneberg   et   al.,   2000,   Morikawa   et   al.,   2002)   and   non-­‐   hematopoietic   tissues,   such   as,   cardiomyocytes   (Harada   et   al.,   2005),   vascular   endothelial   cells   (Bussolino   et   al.,   1989),   neural   stem   cells   (Schneider   et   al.,   2005),   placenta   (McCracken   et   al.,   1996),   many   non-­‐haematopoietic   tumours   cell   lines   (Roberts,   2005)  and  has  recently  been  shown  to  be  highly  expressed  on  human  gastric  and   colon  cancers  (Morris  et  al.,  2014).  However,  GCSF-­‐R  is  predominantly  expressed   on  stem  cells,  common  myeloid  progenitors  (CMP)  and  mature  neutrophils,  where   by  the  expression  of  this  receptor  increases  during  maturation  (Manz  et  al.,  2002).   In  1990,  granulocyte  colony  stimulating  factor  receptor  (GCSF-­‐R)  was  first  cloned   from  mouse  myeloid  leukemia  cell  line  (NFS-­‐60)  and  shown  to  form  homo-­‐dimers   upon   binding   to   its   ligand   (GCSF),   resulting   in   a   complex   2:2   ligand   receptor   subunit  (Fukunaga  et  al.,  1990a).   1.5.2 Structure  and  Function  of  GCSF-­‐R   The  human  GCSF  receptor  is  a  120kDa  cell  surface  receptor,  which  belongs  to  the   hematopoietic   cytokine   receptor   super-­‐family,   HCR.   The   GCSF-­‐R   is   836   amino   acids   in   length   and   consists   of   an   extracellular   region   with   604   amino   acids,   transmembrane   region   with   26   amino   acids   and   a   cytoplasmic   (intracellular)   region   with   183   amino   acids.   The   extracellular   domains   consist   of   the   cytokine   receptor   homology   (CRH)   domain,   N-­‐terminal   immunoglobulin   (Ig)-­‐like   domain,   and  a  Trp-­‐Ser-­‐X-­‐Trp-­‐Ser  (WSXWS)  motif  required  for  ligand  binding,  and  the  rest   of   the   extracellular   region   is   formed   by   3   fibronectin   type   III   (FNIII)   domains   (Molineux  et  al.,  2012).       7   The   intracellular   region   contains   2   conserved   sub-­‐domains   termed   Box   1   and   Box   2,  and  a  membrane-­‐distal  domain  that  includes  Box  3  (less  conserved  sequence)   (Fukunaga   et   al.,   1990b).   The   intracellular   region   of   GCSF-­‐R   has   also   4   tyrosine   residues  at  locations  704,  729,  744  and  764  of  the  human  receptor  (corresponding   to   Y703,   Y728,   Y743,   and   Y763  in   the   murine   receptor);   these   conserved   residues   play   an   important   role   in   the   induction   of   GCSF   cell   survival,   proliferation   and   differentiation  (Figure  1.3)  (Molineux  et  al.,  2012).                       8     .     Figure  1-­‐3:  Structure  and  downstream  signal  pathways  of  GCSF-­‐R   The   GCSF-­‐R   contains   of   extracellular   region   and   intracellular   region.   The   extracellular   region   of   the   GCSF-­‐R   includes   3   fibronectin   type   III   (FNIII)-­‐like   domains,   WSXWS   motif,   a   cytokine   receptor   homologous   (CRH)   domain,   and   immunoglobulin   (Ig)-­‐like   domain.   Conserved   Box   1,   Box   2,   and   Box   3   and   4   tyrosines  (Y703,  Y728,  Y743  and  Y763)  mediate  downstream  signal  transduction   in   the   intracellular   region.   Binding   of   GCSF   to   its   receptor   induces   phosphorylation   of   JAKs,   resulting   in   phosphorylation   of   the   4   tyrosine   residues   located   on   the   cytoplasmic   region.   Once   these   tyrosine   residues   are   phosphorylated,  they  serve  as  docking  sites  for  numerous  proteins  characterized   by   Src  Homology   2   (SH2)   domains   resulting   in   activation   of   many   signaling   pathways  that  regulate  different  cell  processes.   9   The  binding  of  GCSF  to  its  receptor  forms  homo-­‐dimers,  resulting  in  a  complex  of   two   GCSF   molecules   and   two   GCSF-­‐R   molecules   (Larsen   et   al.,   1990).   Each   GCSF   interacts  with  the  immunoglobulin  (Ig)-­‐like  domain  of  one  GCSF-­‐R  subunit  and  the   CRH  domain  of  the  second  GCSF-­‐R  subunit,  resulting  in  a  crossover  configuration   of  the  receptor  subunits  (Figure  1.4.B)  (Tamada  et  al.,  2006).     The   binding   of   GCSF   to   its   receptor   activates   the   Janus   kinase   (Jak)/signal   transducer  and  activator  of  transcription  (STAT)  signalling  pathways.      Molecules   that   get   activated   as   part   of   the   pathway   include   Jak1,   Jak2,   STAT1,   STAT3,   and   STAT5.  It  has  been  reported  that  tryptophan  residues  localized  between  Box  1  and   Box   2   in   the   intracellular   region   of   GCSF-­‐R   serves   as   a   docking   site   for   Jaks.   Because   GCSF-­‐R   lacks   intrinsic   kinase   activity;   thus,   it   mostly   relies   on   different   non-­‐receptor  kinases,  for  instance,  activation  of  JAK  family,  mainly  through  JAK  1   and   2   (Meshkibaf,   2015).   Upon   activation,   the   GCSF-­‐R   dimerizes   and   brings   the   Jaks   together   into   proximity,   resulting   in   their   trans-­‐phosphorylation   of   one   another  which  in  turn  phosphorylate  tyrosine  (Y)  resides  (Y703,  Y728,  Y743,  and   Y763)   located   in   the   cytoplasmic   region.   Once   these   tyrosine   residues   are   phosphorylated,  they  serve  as  docking  sites  for  STAT’s.  STAT’s  are  transcriptional   factors  found  in  the  intracellular  region  (cytoplasmic  region),  and  can  interact  with   phosphotyrosine  residues  of  the  GCSF-­‐R  via  their  Src  Homology  2  (SH2)  domains   (the  function  of  this  domain  is  to  identify  the  phosphorylated  state  of  tyrosine  (Y)   residues).  STAT’s  get  phosphorylated  then  form  dimers  and  migrate  to  the  nucleus,   where  they  bind  DNA  and  activate  transcription.  Although  there  are  seven  family   members   of   STATs,   activation   of   low   level   of   GCSF   results   in   phosphorylation   of   Y703   and   Y743,   resulting   in   strong   stimulation   of   STAT3   with   slight   stimulation   of   STAT1  and  STAT5  (Figure  1.4).  In  vitro,  STAT3  activation  seems  to  push  neutrophil   differentiation   mediated   by   activation   of   neutrophil   marker   genes.   Activation   of   GCSF-­‐R   is   also   appeared   to   activate   STAT1   and   STAT5,   resulting   in   cell   proliferation.     In   addition,   it   has   been   reported   that   Y728   is   a   docking   site   for   Suppressor  of  cytokine  signalling  3  (SOCS3),  which  is  a  critical  feedback  inhibitor   of  GCSF-­‐R  signalling  pathway  (Molineux  et  al.,  2012).     10     Figure  1-­‐4:  Pathway  upon  activation  of  Jak-­‐STAT  signals   [A]  In  the  absence  of  the  ligand,  G-­‐CSFR  is  associated  with  Janus  kinases  (Jaks).  [B]   The   binding   of   the   ligand   to   the   receptor   occurs   at   a   2:2   ligand:receptor   subunit   stoichiometry,  forming  a  cross-­‐over  configuration  between  the  receptor  subunits   bring   the   Jaks   into   proximity   and   enables   their   trans-­‐phosphorylation   and   stimulation.  [C]  The  intracellular  4-­‐tyrosine  residues  of  the  GCSF-­‐R  (represented   by  stars)  are  phosphorylated  by  Jaks.  [D]  STAT  interacts  with  the  phosphotyrosine   residues   through   their   Src  Homology   2   (SH2)   domains   and   become   phosphorylated   by   the   Jak.   Phospho-­‐dimers   of   STATs   accumulate   in   the   nucleus   and   activate   transcription   factors   that   drive   the   neutrophils   from   the   bone   marrow  to  the  blood  circulation.         11   Although  that  GCSF-­‐R  is  widely  accepted  to  activate  Jak/STAT  pathways,  it  has  also   been  reported  that  GCSF-­‐R  is  linked  to  numerous  components  of  Mitogen-­‐activated   protein   (MAP)   kinase   and   phosphoinostide-­‐3-­‐kinase-­‐protein   kinase   B   (PI-­‐3K-­‐PKB)   and   pathways   resulting   in   activation   of   transcription   factors   and   regulation.   For   instance,   Y764   serves   as   docking   site   for   Growth   factor   receptor-­‐bound   protein   2(Grb2)   and   has   been   linked   to   activation   of   p21   Ras   pathway.     A   significant   reduction   of   p21   Ras   activation   and   neutrophil   proliferation   was   noticed  in  vitro   when  Y764  was  absent  (Hermans  et  al.,  2003).  Extracellular  signal-­‐regulated  1/2   (Erk   1/2)   MAP   kinases   are   considered   the   main   downstream   effectors   from   the   p21  Ras  pathway  resulting  in  signaling  proliferation  of  myeloid  progenitor  cells.  In   neural  cells,  it  is  also  reported  that  Erk1/2  is  strongly  activated  upon  exposure  to   GCSF   (Hamilton,   2008,   Panopoulos   and   Watowich,   2008,   Touw   and   van   de   Geijn,   2007).  In  Swan  71  cells,  binding  of  GCSF  to  its  receptor  leads  to  the  activation  of   both   PI3K/Akt   and   Erk1/2   pathways   leads   to   the   migration   of   NF-­‐kB   to   the   nucleus,   stimulating   an   increase   of   matrix   metalloproteinase-­‐2   (MMP-­‐2)   activity   and  Vascular  endothelial  growth  factor  (VEGF)  secretion  (Furmento  et  al.,  2014).  It   is   also   reported   that   the   activation   of   GCSF-­‐R   triggers   PI-­‐3K-­‐PKB   pathway   that   is   important  for  the  stimulation  of  cell  survival  by  inhibiting  the  apoptotic  cascades   (Figure   1.3)   (Hunter   &   Avalos,   2000;   Touw   &   van   de   Geijn,   2007).   Another   pathway   that   are   induced   by   GCSF   is   the   Tyrosine-­‐protein   kinase  (Lyn),   play   a   crucial   role   in   GCSF   mediated   cell   proliferation,   in   addition   to   the   activation   of   GCSF   primed   pro-­‐inflammatory   responses   in   neutrophils   (Sampson   et   al.,   2007,   Sivakumar  et  al.,  2015).   1.5.2.1 Mobilization   o f   N eutrophils   Previously,  it  was  reported  that  STAT3  is  the  major  transcription  factor  activated   upon   binding   of   GCSF   to   its   receptor   but   the   role   of   STAT3   in   the   mechanism   of   neutrophils  mobilization  was  not  clear  (Panopoulos  et  al.,  2002).  The  chemokines,   macrophage   inflammatory   protein-­‐2   (MIP-­‐2,   known   as   Cxcl2)   and   keratinocyte   derived   chemokine   (KC,   Cxcl1)   and   their   shared   receptor   CXCR2   induce   the   mobilization   of   neutrophils   from   the   BM   to   the   circulating   blood.   In   contrast   to   this   the  stromal  cell–derived  factor  1  (SDF-­‐1,  CXCL12)  which  is  expressed  in  the   BM   and   its   chemokine   receptor   4   (CXCR4),   expressed   on   the   surface   of   neutrophils,   contribute   to   the   retention   of   neutrophils   in   the   BM   and   requiring   12   dawn-­‐regulation   to   induce   the   releasing   of   neutrophils.     Nguyen-­‐Jackson   et   al.   (2010)   demonstrated   that   STAT3   controls   the   neutrophils   migration   from   the   BM   to   the   circulating   blood   in   response   to   GCSF   treatment   by   binding   to   the   chemokines  MIP-­‐2  and  KC  and  increasing  the  production  of  these  chemokines  and   reducing   bone   marrow   SDF-­‐1   expression   in   WT   mice   (Figure   1.5)   (Nguyen-­‐ Jackson   et   al.,   2010,   Nguyen-­‐Jackson   et   al.,   2012).   In   summary,   because   GCSF   is   itself  not  chemotactic,  this  concept  is  supported  by  the  observation  that  GCSF  fails   to   induce   circulating   neutrophil   amounts   in   CXCR2-­‐knockout   mice   (Pelus   et   al.,   2002).   Inhibiting   the   SDF-­‐1/CXCR4   interaction   is   sufficient   to   enable   neutrophil   release,   as   shown   by   use   of   the   CXCR4   antagonist   AMD3100   (Plerixafor)   (Broxmeyer  et  al.,  2005).       Figure  1-­‐5:  Scheme  of  the  proposed  model   How  STAT3  induces  the  mobilization  of  neutrophils.  Neutrophils  are  reserved  in   the  BM  in  part  throughout  their  CXCR4  expression,  which  binds  to  SDF-­‐1  (stromal   cells   express   SDF-­‐1).   Administration   of   GCSF   leads   to   down-­‐regulation   and   decreases   of   SDF-­‐1   with   its   receptor   CXCR4;   suppression   of   SDF-­‐1   needs   STAT3.   Intake   of   GCSF   also   stimulate   the   neutrophil   chemo-­‐attractants   MIP-­‐2   and   KC   in   the   BM   together   with   up-­‐regulation   of   their   shared   CXCR2   on   the   neutrophils   surface,   STAT3   is   required   in   the   stimulation   of   MIP-­‐2,   KC   and   CXCR2.   Modified   from  (Nguyen-­‐Jackson  et  al.,  2010,  Nguyen-­‐Jackson  et  al.,  2012).     13   1.6 The  Major  Clinical  Use  of  GCSF   1.6.1 Febrile  Neutropenia  Prophylaxis   Febrile   neutropenia   (FN)   complications   (defined   as   development   of   fever   >38.5OC   with  absolute  neutrophil  counts  <  1.0x106/L)  are  the  main  symptoms  observed  in   patients   treated   with   systemic   cancer   chemotherapy.   Administration   of   GCSF   shows  benefits  in  reducing  the  risk  of  FN  and  accelerates  the  neutrophils  number   (Shah  and  Welsh,  2014).     1.6.2 Mobilization  of  Stem  Cells     Administration   of   rhGCSF   promotes   and   accelerates   hematopoietic   stem   cell   (HSC)   secretion   into   peripheral   blood   in   order   to   facilitate   collection   of   large   numbers  of  stem  cells  from  the  peripheral  blood.  Thus,  GCSF  is  given  to  patients   following   chemotherapies   in   two   processes   either   autologous   or   allogeneic   stem   cell   transplantation.   Autologous   stem   cell   transplantation   is   a   process   that   depends  on  the  collection  of  a  patient's  own  stem  cells,  administration  of  GCSF  is   mainly   alone   or   in   combination   with   chemotherapeutic   drugs   (Bensinger   et   al.,   1995,   Martino   et   al.,   2014).   The   combination   of   GCSF   with   chemotherapeutic   drugs   showed   higher   numbers   of   CD34+   cells   and   lower   levels   of   apheresis   sessions  in  comparison  to  administration  of  GCSF  alone  (Pusic  and  DiPersio,  2008,   Tanhehco   et   al.,   2010).   In   contrast,   allogeneic   stem   cell   transplantation   is   a   process   in   which   the   patient   receives   stem   cells   from   a   healthy   individual   who   have   been   injected   with   GCSF   for   the   purpose   of   donation.   This   process   is   more   effective  and  safe  than  autologous  stem  cell  transplantation  (Hölig,  2013).   1.6.3 Controlling  of  SCN  and  AML   Administration   of   rhGCSF   therapy   has   been   shown   to   improve   the   conditions   of   patients  with  acute  myeloid  leukemia  (AML)  prior  to  chemotherapy  in  two  ways;   increasing  both  the  neutrophil  counts  in  the  first  24  hours  and  the  susceptibility  of   myeloid  leukemia  blast  cells  to  chemotherapy  (Löwenberg  et  al.,  2003,  Beekman   and   Touw,   2010).   Also,   it   has   shown   to   directly   reduce   the   risk   of   several   leukemia’s.   For   instance,   lymphoblastic   AML   patients   achieved   complete   remission  when  treated  with  GCSF  alone  (Nimubona  et  al.,  2002).   14   Although   GCSF   can   increase   the   number   of   neutrophils   in   patients   with   severe   congenital   anaemia   (SCN)   and   thereafter   minimize   this   risk   of   recurrent   infections,  GCSF-­‐R  mutations  have  been  reported  in  a  patient  with  SCN  who  may   developed   secondary   AML   and   myelodysplastic   syndrome (MDS)   due   to   administration   of   GCSF   (Ancliff   et   al.,   2003,   Beekman   et   al.,   2012).   More   details   about   the   adverse   effects   of   GCSF   administrations   will   be   discussed   in   the   next   section.     1.7 The  Main  Side  Effects  of  GCSF  Administration   To  improve  the  safety  of  GCSF  treatment  in  the  clinical  situation,  it  is  important  to   consider  the  main  side  effects  of  GCSF  administration.  Over  two  decades,  several   distinct   GCSF-­‐R   mutations,   which   cause   intracellular   receptor   truncations,   have   been   reported   in   a   patient   with   SCN   who   developed   secondary   AML/MDS   (Beekman   et   al.,   2012).   These   truncations   are   considered   a   crucial   step   in   the   expansion  of  the  pre-­‐leukemic  clones  and  the  possibility  that  GCSF  administration   will  give  rise  to  these  mutant  clones  and  thereafter  cause  AML.     All  GCSF-­‐R  mutants  have  no  differences  in  their  juxta-­‐membrane  and  extracellular   domains,   but   do   differ   in   their   cytoplasmic   (intracellular)   regions.     Among   all   these  mutations,  the  class  IV  isoform  mutant  (differentiation-­‐defective)  has  been   detectable   in   hematopoietic   cells   and   is   associated   with   administration   of   GCSF.   This   isoform   maintains   the   membrane   proximal   sequence   that   is   required   for   proliferative   signaling   but   loses   at   position   725   the   carboxy-­‐terminal   87   amino   acids,   which   are   replaced   with   a   unique   34   amino   acid   sequence   (White   et   al.,   1998,  White  et  al.,  2000).  As  a  result,  out  of  the  4-­‐tyrosine  residues  (Y704,  Y729,   Y744   and   Y764)   in   the   full-­‐length   form,   only   Y704   is   conserved   (Class   IV   isoform)   (Figure   1.6).   Overexpression   of   isoform   IV   has   been   observed   in   patients   with   AML/MDS,   potentially   due   to   the   ability   of   this   isoform   to   block   maturation   (Liongue   and   Ward,   2014).   In   addition   to   this   mutation,   common   side   effects   reported  for  patients  who  were  treated  with  GCSF  are  summarized  in  Table  1.1.     15     Figure  1-­‐6:  Comparison  of  carboxyl-­‐terminal  region  of  the  GCSF-­‐R  in  patients   with  AML   Class   IV   isoform   has   no   differences   in   their   juxta-­‐membrane   and   extracellular   domains,   but   differing   in   their   cytoplasmic   domains.   The   isoform   maintains   the   membrane   proximal   sequence   that   required   for   proliferative   signaling   and   loses   at   position   725   the   carboxy-­‐terminal   87   amino   acids,   which   are   replaced   with   a   unique  34  amino  acid  sequence.  As  a  result,  the  4-­‐tyrosine  residues  (Y704,  Y729,   Y744  and  Y764)  in  the  full-­‐length  form  class1  (wild  type)  truncated  and  only  Y704   is   conserved   among   this   isoform   (Class   IV).   Modified   from   (Liongue   and   Ward,   2014).       16   Table  1-­‐1:  The  main  side  effects  reported  for  patients  treated  with  GCSF   Side  effect   Organ   Possible  mechanism   References   Osteopenia  &   osteoporosis   Bone   Administration  of  G-­‐CSF  increases  bone   resorption  via  increasing  the  activity  of   osteoclasts  leading  to  significant  bone  loss.   (D'Souza  et  al.,  2008)   Joint  pain  &   generalized   weakness   Bone   1)  Expansion  of  bone  marrow,     (Lambertini  et  al.,  2014)   2)  Enhancing  of  GCSF-­‐R  on  afferent  nerve   fibers  lead  to  generate  peripheral  nociceptor   sensitization,       3)  Stimulation  of  inflammatory  cells,  such  as,   macrophages  and  monocytes  that  contributes   to  nerve  remodeling,     4)  Osteoblast  and  osteoclast  activation.   Erythematous  rash,   Skin   urticarial  and  Sweet   syndrome   Infiltration  of  neutrophils  into  dermis  and   epidermis.   (Nomiyama  et  al.,  1994,   Prendiville  et  al.,  2001,   Llamas-­‐Velasco  et  al.,   2013)   Splenomegaly  &   extramedullary   hematopoiesis   Spleen   Stimulation  of  myelopoisis.   (Litam  et  al.,  1993,   O'Malley  et  al.,  2003,   Dagdas  et  al.,  2006)   Lung   Accumulation  of  neutrophils  due  to  releasing   of  chemoattractant  molecules.  As  a  result,   these  cells  release  a  number  of  substances   injurious,  for  instance,  platelet,  leukotrienes,   proteases  and,  oxidants  that  cause  damage  in   the  alveolar  endothelium  and  epithelium.   (Asano  et  al.,  1977,  Wada   et  al.,  2011,  Yamaguchi  et   al.,  2012,  Inokuchi  et  al.,   2015)   Kidney   Stimulation  of  leukocytosis  in  the  kidneys.   (Hirokawa  et  al.,  1996)     Lung  cancer  and   acute  respiratory   distress  syndrome     Reversible  renal   impairments       17   1.8 Available  Commercial  GCSF  Preparations  and  Their  Limitations   Human   GCSF   has   been   cloned   and   is   currently   available   as   two   recombinant   human  GCSF  (rhGCSF)  preparations  for  HPC  mobilization:     1.8.1 Filgrastim  (NEUPOGEN ® )     Filgrastim   is   a   non-­‐glycosylated   form   (18.8kD),   obtained   from   E.   coli   and   has   a   methionine   group   at   its   N-­‐terminal   end.   In   1991,   Filgrastim   was   licensed   and   marketed   as   a   treatment   for   patients   with   neutropenia   following   chemotherapeutic  drugs.  Since  its  launch,  clinicians  have  recommended  the  use  of   Filgrastim   in   bone   marrow   transplantation   procedures,   aplastic   anaemia,   sever   congenital   neutropenia,   and   to   support   patients   with   AIDS   and   myelodysplastic   syndromes  (Molineux,  2004).  However,  it  was  reported  that  Filgrastim  has  a  short   half-­‐life   in   the   serum   of   3.5   hours   when   injected   in   to   healthy   volunteers   and   patients   with   malignancies   because   E.   coli   derived   Filgrastim   lacks   the   O-­‐linked   glycosylation  (Cooper  et  al.,  2011,  Hoggatt  and  Pelus,  2014).   1.8.2 Lenograstim  (Granocyte ® )   Lenograstim   is   an   O-­‐glycosylated   form   at   Thr-­‐133   position   obtained   from   Chinese   hamster   ovarian   (CHO)   cells   (Nagata   et   al.,   1986).   It   has   a   short   half-­‐life   in   the   serum  of  between3–4  hours  and  O-­‐linked  was  shown  to  have  an  important  role  in   providing   greater   stability   to   the   GCSF   by   protecting   the   cysteine-­‐17   sulfhydryl   group   from   oxidation   by   free   radicals   (Hasegawa,   1993,   Cooper   et   al.,   2011).   It   was  believed  that  the  O-­‐linked  glycosylation  might  show  clinical  advantages  over   the   non-­‐glycosylated   form   (Filgrastim),   however   the   in  vivo   comparative   studies   showed   that   no   differences   between   them   in   all   aspects   (Ataergin   et   al.,   2008).   Besides,   leukemic   patients   need   daily   injections   to   maintain   its   activity   in   the   circulation,  which  are  inconvenient,  expensive  and  painful  especially  for  children.   As  Lenograstim  is  similar  to  the  natural  GCSF,  studies  shifted  to  focus  on  this  form.       18   1.9  Strategies  Used  to  Delay  the  Clearance  of  GCSF   Since   the   main   limitations   of   the   previous   forms   of   rhGCSF   (1st   generation)   are   short  circulating  half-­‐life  and  daily  injection,  two  strategies  have  been  developed   to  overcome  these  issues.  The  first  strategy  is  to  increase  the  molecular  weight  of   the   therapeutic   proteins   (hydrodynamic   radius)   above   the   renal   filtration   threshold.   Predominantly,   this   increase   in   molecular   weight   can   be   attained   by   conjugation   with   another   moiety   such   as,   PEGylation.   The   second   strategy   to   extend  the  half-­‐life  of  the  therapeutic  proteins  offers  the  benefits  of  the  neonatal   Fc  receptor  (FcRn)  recycling  mechanism  by  forming  fusion  proteins  with  both  Fc   portion   of   immunoglobulin   and   albumin   (Natalello   et   al.,   2012,   Cox   et   al.,   2014,   Chung  et  al.,  2011).  Additional  new  approaches  using  Asterion  profuse  technology   will  also  be  discussed.       1.9.1 Extension  of  Half-­‐life  by  Increasing  the  Molecular  Weight   1.9.1.1 PEGylation       The  short  circulating  half-­‐life  of  many  recombinant  proteins  can  be  increased  via   conjugation   with   poly   ethylene   glycol   (PEG),   in   a   process   termed   PEGylation   (Natalello   et   al.,   2012).   In   the   1970’s,   It   was   first   described   by   Abuchowski   and   Davis  who  found  that  PEG  may  improve  the  immunological  properties  and  serum   half-­‐life  of  proteins  such  as  bovine  liver  catalase  and  albumin  (Abuchowski  et  al.,   1977).  Since  then,  widespread  research  has  been  carried  out  into  PEG  technology   resulting   in   highly   variable   PEGs   with   several   molecular   weights   (Jain   and   Jain,   2008).   A   range   of   Pegylated   proteins   are   now   clinically   available,   such   as,   Pegvisomant  (Pegylated  growth  hormone  antagonist,  licensed  for  the  treatment  of   acromegaly  in  2003  (Trainer  et  al.,  2000,  Hamidi  et  al.,  2006).   To   generate   a   new   2nd   generation   product,   a   20   kDa   PEG   molecule   was   attached   covalently   to   recombinant   methionyl   (r-­‐met)   human   GCSF   Neupogen   (Filgrastim).   This   new   molecule   was   marketed   as   Neulasta   (Pegfilgrastim).   In   Pegfilgrastim,   each  ethylene  oxide  unit  of  PEG  binds  to  three  water  molecules  which  increases  its   water  solubility  and  also  hydrodynamic  radius  (molecule’s  diameter)  resulting  in   increased  size  of  the  molecule  to  ~38.8kDa,  thus  reducing     19   renal   clearance.   PEGylation   also   creates   a   hydrophilic   shield   that   protects   the   protein   from   proteolysis   and   immunologic   recognition   (Bailon   and   Won,   2009,   Milla   et   al.,   2012).   The   key   behind   the   successful   progress   of   Pegfilgrastim   over   Filgrastim  was  an  understanding  of  the  GCSF  clearance  processes  in  the  body.  In   humans,   GCSF   has   two   clearance   mechanisms:   renal   clearance   and   neutrophil-­‐ mediated  clearance  (Yowell  and  Blackwell,  2002).  The  presence  of  the  PEG  moiety   decreases  the  renal  clearance  of  Pegfilgrastim  and  as  a  result  it  is  mainly  cleared   via  a  self–regulating  neutrophil-­‐mediated  mechanism,  which  is  dependent  on  the   number  of  neutrophils.  Following  administration  of  Pegfilgrastim,  concentrations   remain   high   in   patient   serums   during   neutropenia,   but   are   reduced   when   the   numbers   of   neutrophils   increase.   Therefore,   a   single   injection   of   Pegfilgrastim   per   chemotherapy   cycle   is   as   efficient   as   the   daily   administration   of   Filgrastim   (Curran  and  Goa,  2002).   The   PEGylation   of   a   protein   used   to   be   one   of   the   major   limitations   because   the   whole  PEG  is  often  processed  for  excretion  in  the  human  body  without  undergoing   an  initial  biodegradation  which  could  be  toxic  to  the  body   (Patel  et  al.,  2014).  This   possible   toxicity   was   supported   by   the   detection   of   PEG   in   bile   (Caliceti   and   Veronese,   2003).   Vacuole   formation   has   also   been   observed   in   renal   tubules   upon   administration   of   PEG   thereby   affecting   the   tissue   distribution   and   in   turn   clearance  (Zhang  et  al.,  2014).   Modification  of  GCSF  with  PEGylation  reduced  in  vitro  biological  activity  of  GCSF   (2  to  3  fold),  as  the  conjugation  of  PEG  to  the  GCSF  could  induce  structural  change   in   the   molecule   that   attenuates   the   potency   (Kinstler   et   al.,   1996,   Gaertner   and   Offord,  1996).   The   increased   cost   of   PEGylation   is   considered   to   be   one   of   the   main   limitations   because   Pegfilgrastim   requires   a   post-­‐expression   chemical   modification   and   purification   processes   (Pisal   et   al.,   2010).   The   table   below   summarizes   the   advantages   and   limitations   of   PEGylation   and   other   different   strategies   used   to   generate  long  acting  GCSF  (Table  1.2).     20   Table   1-­‐2:   Summary   of   pros   and   cons   of   PEGylation   and   other   different   strategies  used  to  generate  a  long  acting  GCSF   Strategy( Pros( Cons( References( Detection+of+vacuole+formation+++ within+the+kidney+and+ ++(Kinstler+et+al.,+1996,+ ++Extended+circulating+half9life+by+ macrophages.++ Gaertner+and+Offord,+ reduced+renal+clearance.+ 1996,+Caliceti+and+ ++Non9biodegradable.+ nd+++ Creates+a+hydrophilic+shield+ PEGylation+2 generation+ Veronese,+2003),+(Pisal+et+ ++Reduced+in#vitro+biological+++++++ Peg1ilgrastim+ that+protects+the+protein+from+ activity+of+GCSF+(2+to+3+fold).++ al.,+2010),+(Patel+et+al.,+ proteolysis+and+immunologic+ 2014),+(Zhang+et+al.,+ ++High+cost+due+to+the+chemical++ recognition+and+proteolysis.++ 2014)+ modi1ications+and+puri1ication+ process+ ++Several+proteins+showing+no+ Prolonged+circulating+half9lives+ bioactivity+when+attached+to+the+ (5+to+89fold+longer+than+G9CSF),+ Fc9IgG1+domain.+Reduced+in#vitro+ ++(Cox+et+al.,+2004,+Cox+et+ Fusion+to+Antibodies+2nd+ accelerate+number+of+ biological+activity+of+GCSF+(3+to+4+ al.,+2014,+Czajkowsky+et+ generation+G9CSF/IgG9Fc+&+G9+ al.,+2012,+Mitragotri+et+al.,+ neutrophils+in+vivo+and+ fold)+when+attached+to+the+IgG9 CSF/IgG9CH+ 2014)+ decrease+risk+of+ CH+domain+due+to+signi1icant+ amount+of+disul1ide9linked+ immunogenicity.+ aggregates/oligomers.+++ Increased+half9life+of+GCSF,+ increased+WBC+counts+of+ ++(Zhao+et+al.,+2013,+ Not+offering+any+secondary+ neutropenia+mice,+decrease+risk+ nd+ Schmidt,+2009,+Mitragotri+ functions,+for+instance,+ Fusion+to+Albumin+2 generation+ of+immunogenicity,+ cytotoxicity.+High+cost+product.++ et+al.,+2014)+ biodegradable+and+showed+high+ stability.+ ++Increased+half9life+of+GH,+ ++LR9fusion+(Asterion)+3rd++++++++ reduced+immunogenicity+and+ ++(Wilkinson+et+al.,+2007,+ ++ generation+ Ferrandis+et+al.,+2010)+ toxicity.+Naturally+occurring+ sequences+ 21   1.9.2 Extension  of  Half-­‐life  Using  the  FcRn-­‐Mediated  Recycling   Mechanism   In  the  last  years,  the  FcRn  recycling  mechanism  has  been  used  expensively   as  a  strategy  to  prolong  the  half-­‐life  of  different  proteins.    Plasma  proteins   such   as,   albumin   and   Immunoglobulins   (IgG’s)   are   found   to   recycle   by   the   neonatal   Fc   receptor   (FcRn)   pathway   resulting   in   a   longer   circulating   half-­‐ life.     It   has   been   shown   that   the   FcRn   is   responsible   for   the   extraordinary   long   circulating  half-­‐life  of  albumin  and  IgG’s  (19  days  for  albumin  and   23   days   for  IgG’s  in  human  (Dall'Acqua  et  al.,  2002,  Chaudhury  et  al.,  2003,  Anderson   et   al.,   2006,   Baker   et   al.,   2009)).   Studies   have   shown   that   IgG’s   are   catabolized  more  quickly  (Ghetie  et  al.,  1996,  Israel  et  al.,  1996)  and  albumin   is  degraded  approximately  twice  as  fast  in  FcRn  deficient  mice  than  in  wild   type   mice   (Chaudhury   et   al.,   2003,   Baker   et   al.,   2009).   Fusion   or   non-­‐ covalent   binding   of   small   proteins   (e.g.   GCSF)   to   the   Fc-­‐part   of   IgG’s   or   albumin  significantly  improved  their  pharmacokinetic  properties.     Generally,   FcRn   can   bind   tightly   to   albumin   as   well   as   IgG’s   in   a   pH-­‐ dependent  manner.  Due  to  the  presence  of  histidines  in  albumin  and  IgG’s,   the   imidazole   group   of   histidine   becomes   protonated   at   acidic   pH   6.0   and   interacts  with  the  FcRn  receptor.  The  interaction  complex  of  FcRn-­‐albumin   or   IgG’s   is   then   internalized   and   taken   by   cells   to   protect   albumin   or   IgG’s   from  lysosomal  degradation  (Chaudhury  et  al.,  2006,  Andersen  and  Sandlie,   2009,  Dumont  et  al.,  2006).  FcRn-­‐albumin  or  IgG’s  complex  is  released  back   to  the  cell  surface  membrane,  where  exposure  to  physiological  pH  7.2  within   the   blood   circulation   causes   the   release   of   albumin   or   IgG’s   from   the   receptor  (Ober  et  al.,  2004,  Andersen  et  al.,  2006).     Extensive   studies   on   the   Human   serum   albumin   (HAS)   structure   indicated   that  carboxy  terminal  domain  III  HAS  (3DHAS)  alone  is  sufficient  for  binding   FcRn   receptor   and   the   histidine   present   in   3DHSA   may   dominates   the   binding   between   HSA   and   the   FcRn   receptor   (Figure   1.7)   (Andersen   et   al.,   2010).   22       Figure   1-­‐7:   Model   of   the   pH-­‐dependent   recycling   mechanism   of   albumin  via  the  FcRn  receptor  in  serum    The   three   domains   of   albumin   are   marked   in   green   (domain   I),   blue   (domain   II)   and   red   (domain   III).   Domain   III   albumin   is   binding   FcRn   receptor   at   acidic   pH   6.0   to   protect   from   lysosomal   degradation   and   recycling   again   to   the   circulation   where   exposure   to   physiological   pH   7.2   causes  the  release  of  albumin.       23   1.9.2.1   F usion   o f   G CSF   t o   A lbumin   Human   serum   albumin   (HSA)   is   produced   in   liver   and   has   numerous   physiological   roles   including   transportation   of   fatty   acid   and   metal   ions   as   well   as   maintenance   of   plasma   pH   and   colloid   blood   pressure.   As   albumin   is   a   large   protein   (67kDa),   it   can   be   fused   to   recombinant   proteins   to   extend   their   half-­‐life.   As   a   result,   fused   proteins   will   automatically   attain   a   molecular   weight   too   large   to   be   filtered   through   the   kidney   and   increase   plasma   protein   residency   time   (Dennis   et   al.,   2002,   Andersen   and   Sandlie,   2009)     The   advantage   of   carboxy   terminal   domain   III   HAS   (3DHAS)   has   been   considered  widely  and  later  3DHAS  was  genetically  fused  to  the  N-­‐terminal   of  GCSF.  The  pharmacokinetic/  pharmacodynamics  (PK/PD)  studies  showed   increased   half-­‐life   and   increased   WBC   counts   of   neutropenia   model   mice   compared  to  native  GCSF  (Zhao  et  al.,  2013).     1.9.2.2   F usion   o f   G CSF   t o   I gG-­‐Fc     In  humans,  the  circulating  half-­‐life  of  IgG1  and  IgG4  immunoglobulins  is  23   days,   and   that   IgG1   and   IgG4   immunolglobulins   have   been   used   to   form   several   long-­‐acting   fusion   proteins   (Gaberc-­‐Porekar   et   al.,   2008).   Thus,   immunoglobulins  selected  as  the  choice  antibody  for  Fc  fusion  proteins.   Structurally,   immunoglobulins   are   composed   of   two   identical   light   and   heavy   chains   connected   by   disulphide   bonds.     Both   chains   contain   two   regions:     the   fragment   of   antigen   binding   (Fab)   (the   head   region   of   an   antibody)   responsible   for   immunogenic   detection   and   the   crystallisable   fragment  (Fc)  (the  tail  region  of  an  antibody  that  interacts  with  cell  surface   receptor)   responsible   for   maintenance   of   IgG   in   the   blood   circulation   (Gaberc-­‐Porekar  et  al.,  2008).     The  IgG  immunoglobulin  has  two  fragments:  CH  (CH1-­‐Hinge-­‐CH2-­‐CH3)  and   Fc  (Hinge-­‐CH2-­‐  CH3)  domains.  The  Hinge  domain  is  responsible  for  linking   Fab   and   Fc   regions   and   provides   more   flexibility.   Many   therapeutic   proteins   have  been  reported  to  joined  via  the  amino-­‐termini  of  CH  (CH1-­‐Hinge-­‐CH2-­‐ 24   CH3)   and   Fc   (Hinge-­‐CH2-­‐   CH3)   domains   of   human   IgGs   through   their   carboxy-­‐termini  (Cox  et  al.,  2004).    In  mammalian  cells,  IgG  fusion  proteins   are   frequently   expressed   and   secreted   as   disulfide-­‐linked   homodimers   because   of   inter-­‐chain   disulfide   bonds   that   are   created   between   cysteine   residues   sited   in   the   hinge   region   of   the   IgGs.   The   effective   size   and   circulating   half-­‐life’s   of   IgG   fusion   proteins   are   further   increased   by   the   dimeric  structure  of  IgG  fusions.     Chimeric   genes   have   been   produced   encoding   human   GCSF   that   are   fused   through   a   7-­‐amino   acid   flexible   linker   (Ser-­‐Gly-­‐Gly-­‐Ser-­‐Gly-­‐Gly-­‐Ser)   to   the   N-­‐termini   of   the   CH   (CH1-­‐Hinge-­‐CH2-­‐CH3)   and   Fc   (Hinge-­‐CH2-­‐   CH3)   domains  of  human  IgG4  and  IgG1  immunoglobulins  (Figure  1.8).  Fusions  of   GCSF   to   human   IgG   domains   were   shown   to   form   homodimers   with   high   molecular   weight,   prolonged   circulating   half-­‐lives   (5   to   8-­‐fold   longer   than   GCSF)  and  accelerate  number  of  neutrophils  in   vivo,  without  any  significant   effect  on  GCSF  biological  activity  in  vitro  (Cox  et  al.,  2004,  Cox  et  al.,  2014).       25     Figure  1-­‐8:  The  schematic  diagram  shows  (A)  G-­‐CSF/IgG-­‐Fc  protein  and   (B)  G-­‐CSF/IgG-­‐CH  fusion   The  GCSF  carboxy-­‐terminus  is  linked  through  a  7  amino  acid  fixable  linker   (L)  to  the  amino  termini  of  the  IgG-­‐Fc  and  IgG-­‐CH  domains.  The  CH1,  CH2,   and   CH3   regions   and   hinge   (H)   of   the   IgG   domains   are   also   showed.   The   dimeric   of   fusion   proteins   located   in   the   IgG   hinge   region   is   due   to   the   presence   of   disulfide   bonds   (SS)   that   form   between   cysteine   residues.   Modified  from  (Cox  et  al.,  2004,  Cox  et  al.,  2014).       26   1.9.3 New  Approach  by  Asterion   Asterion   is   a   Sheffield   University   spin   out   company   formed   in   2001.   Prof   Richard   Ross   is   one   of   the   founding   directors.   Their   strategy   is   to   develop   long   acting   biological   using   novel   platform   technologies,   which   utilises   the   fusion   of   ligand   with   soluble   extracellular   receptor,   termed   Profuse™   technology.   Over  the  last  fifteen-­‐years,  plans  have  been  developed  by  Asterion  to  focus   on   the   utility   of   Profuse™   technology   to   produce   long   acting   biopharmaceutical   products.   The   current   therapeutics   regime   for   protein   replacement   requires   daily   injections,   which   are   expensive   and   inconvenient.  Thus,  there  is  a  need  for  a  3rd  generation  of  protein  therapies   that   are   easy   to   administrator,   acceptable   and   convenient   to   patients   and   also   minimizes   manufacturing   costs.   Thus,   two   approaches   have   been   created  by  Asterion  technology: 1.9.3.1 Ligand/Receptor   F usion   Using   flexible   linker   (Gly4Ser)n   technology,   Wilkinson   et   al.   (2007)   demonstrated   that   a   fusion   of   Growth   Hormone   (GH)   to   its   extracellular   receptor  (GH  binding  protein:  GHBP)  creates  an  effective  long  acting  agonist   with   exceptional   delayed   clearance   properties.   PK   analysis   in   rats   showed   that   ligand-­‐receptor   growth   hormone   fusion   molecule   (LR-­‐fusion)   had   a   resulting   300-­‐times   reduced   clearance   rate   when   compared   to   native   GH   (Wilkinson  et  al.,  2007).     In   addition,   preclinical   work   on   a   various   number   of   long   acting   GCSF   molecules   has   been   performed   and   the   preliminary   data   for   one   construct   (4A1)   showed   that   it   is   possible   to   design,   clone   and   purify   a   GCSF   linked   to   its  extracellular  receptor.  It  was  also  shown  to  reduce  the  rate  of  clearance   following   subcutaneous   injections   (up   to   60   hours)   in   rats   consistent   with   an  increase  of  15-­‐fold  over  that  recorded  for  the  native  GCSF  and  2-­‐fold  over   that  reported  for  Pegfilgrastim.     27   1.9.3.2   G lycosylation   In   medicine,   any   protein   used   for   therapeutic   purposes   is   not   only   a   sequence  of  amino  acids  determined  by  a  particular  gene,  but  it  still  requires   editing,   altering   of   the   amino   acids   or   addition   of   carbohydrates.   These   modifications   following   the   preliminary   translation   of   the   protein   are   called   post-­‐translational  processes  (Li  and  d'Anjou,  2009).    Glycosylation  refers  to   the   post-­‐translational   process   that   attaches   oligosaccharide   to   polypeptides.   It  is  one  of  the  most  common  protein  modifications  and  more  than  50%  of   proteins  are  glycosylated  in  the  body,  which  are  mainly  secreted  or  part  of   cell  membrane  components  (Sola  et  al.,  2007).     Glycosylation   play   a   fundamental   role   in   forming   or   maintaining   glycoprotein   integrity.   Generally,   it   can   increase   the   molecular   weight   of   proteins,   provide   a   high   degree   of   protection   against   proteolytic   degradation   and   enhance   thermal   stability   by   decreasing   immunogenicity   due   to   the   presence   of   terminal   sialic   acid   that   creates   negative   charge   around  the  glycoprotein  resulting  in  delay  clearance.  Given  these  important   functions,   it   is   now   believed   that   glycosylations   contribute   to   regulating   protein-­‐protein   interactions,   which   is   highly   important   in   optimizing   and   developing   glycoprotein   drugs.   In   addition,   an   understanding   of   the   association  of  the  carbohydrate  moieties  to  receptor  binding  can  be  used  to   enhance  treatment  efficacy  (Li  and  d'Anjou,  2009).   Therefore,  efforts  have  been  made  to  improve  a  strategy  to  delay  clearance   of   GCSF   by   the   addition   of   natural   carbohydrates   (e.g.   glycosylation)   and   avoiding   limitations   that   were   observed   in   section   1.8.   For   example,   PEG   molecule   is   often   processed   for   excretion   in   the   human   body   without   undergoing  an  initial  biodegradation  which  could  be  toxic  to  the  body  (Patel   et   al.,   2014).   Whereas,   modifying   proteins   with   glycosylation   can   undergo   degradation   within   the   human   body   with   no   potential   toxicity   of   PEGylation   was  supported  by  the  detection  of  PEG  in  bile  (Caliceti  and  Veronese,  2003).   For   the   safe   conjugation   of   glycosylation   to   GCSF,   it   is   important   first   to   understand   the   forms   and   functions   of   glycosylation   and   also   the   best   conjugation  method  to  the  protein.       28   Normally,   the   majority   of   proteins   synthesized   begin   in   the   endoplasmic   reticulum   (ER)   undergo   glycosylation   and   are   completed   in   the   Golgi   apparatus.   Five   classes   of   glycosylation   are   produced;   N-­‐linked   glycosylation,   O-­‐linked   glycosylation,   Phospho-­‐serine   glycosylation,   C-­‐ mannosylation  and  formation  of  GPI  anchors.  However,  the  two  major  forms   are  O-­‐linked  and  N-­‐linked  glycosylation  (Saint-­‐Jore-­‐Dupas  et  al.,  2007).     O-­‐linked  Glycosylation   O-­‐linked  glycosylations  are  attached  to  the  hydroxyl  groups  (-­‐OH)  of  serine   or   threonine   residues   within   a   protein   (Wongtrakul-­‐Kish   et   al.,   2012).   However,  no  particular  consensus  sequences  have  been  recognized  for  this   reaction   and   it   is   still   ambiguous   why   certain   Ser/Thr   residues   are   glycosylated   as   opposed   to   others.   One   theory   is   that   it   could   be   down   to   alternative   structural   properties   of   the   protein   that   may   contribute   to   the   availability  of  the  glycosylation  site  (Sinclair  and  Elliott,  2005).     In   human   GCSF,   O-­‐linked   glycosylation   is   located   at   Thr-­‐133   and   is   necessary   for   the   increased   stability   of   the   GCSF   molecule.   The   actual   molecular  mechanism  of  the  glycosylated  forms  increased  stability  remains   to   be   established,   but   a   study   carried   out   by   Hasegawa   (1993)   suggested   that   O-­‐linked   glycosylation   might   be   involved   in   the   protection   of   the   cysteine-­‐17   sulfhydryl   group   by   preventing   oxidation   by   free   radicals   (Hasegawa,  1993).     N-­‐linked  Glycosylation   N-­‐linked  glycosylation  is  attached  to  the  amide  nitrogen  of  asparagine  (Asn)   residues   within   the   common   consensus   sequence   Asn-­‐X-­‐Ser/Thr   where   X   can  be  any  amino  acid  except  proline  (Pro)  (Kornfeld  and  Kornfeld,  1985).   In   nature,   N-­‐linked   is   the   most   common   form   of   glycosylation   and   is   thus   preferentially   used   in   many   technologies   of   protein   modification   (Spiro,   2002).  Therefore,  it  is  the  focus  of  further  discussion.       Glycosylation   biosynthetic   pathways   start   in   ER   (Figure   1.9).   At   this   early   stage,   a   9-­‐Mannose   glycan   is   added   to   the   peptide   of   an   N-­‐linked   glycan   29   (known   as   a   high   mannose   type).   The   addition   of   these   glycans   is   fundamental  to  the  control  of  the  folding  of  the  newly  synthesized  proteins.   Upon   successful   folding   of   the   protein,   the   glycoprotein   migrates   into   the   Golgi  apparatus  where  mannosidases  facilitate  the  removal  of  the  mannose   groups.   This   is   then   followed   by   the   addition   of   different   monosaccharides   to   the   growing   glycan   chain   by   particular   glycosyltransferases   (This   is   known   as   a   hybrid   type   which   contains   both   high   mannose   and   complex   type).     At   this   late   stage   the   biosynthetic   process   is   completed   in   the   Golgi   apparatus  with  a  fully  sialylated  glycan  complex,  which  contains  six  sugars:   N-­‐acetylglucosamine  (GlcNAc),  mannose  (Man),  fucose  (Fuc),  galactose  (Gal)   and   sialic   acid     (NeuAc),   which   are   linked   by   different   α-­‐   or   β   glycosidic   linkages  (Kim  et  al.,  2009,  Butler  and  Spearman,  2014).     Sialic   acid   (NeuAc)   occupies   the   terminal   site   of   the   N-­‐linked   glycan   and   has   been   found   to   be   critical   in   maintaining   the   circulating   half-­‐life   of   glycoproteins   (Sola   and   Griebenow,   2009,   Kim   et   al.,   2009).     It   contains   a   large   group   of   nine   carbon-­‐containing   carbohydrates   the   most   predominant   of  which  is  N-­‐  acetylneuraminic  acid  (Neu5Ac).  Byrne  et  al.,  (2007)  showed   that   many   properties   conferred   onto   proteins   are   due   to   the   presence   of   the   negative  charge  on  C1  of  sialic  acid.         30     Figure  1-­‐9:  Glycan  structure   Initial   input   glycan   (9-­‐Mannose   glycan)   starts   biosynthetic   pathway   in   ER   (Top),   the   glycoprotein   migrates   into   Golgi   where   the   removal   of   mannose   group   and   the   addition   of   different   monosaccharides   in   a   process   called   hybrid  type  (mid).  The  biosynthetic  processed  then  is  completed  in  the  Golgi   as   fully   sialylated   glycan   complex   (bottom).   Man:   mannose;   Gal:   galactose;   GlcNAc:  N-­‐acetylglucosamine;  Fuc:  fucose;  NeuAc:  sialic  acid.       31   Glycan  and  sialic  acid  function   In   naturally-­‐glycosylated   proteins,   oligosaccharides   (glycans)   and   their   terminal  sialic  acids  fulfil  a  number  of  functions:     1.   Overexpression   charge   repulsion   between   glomerular   filtration   barrier   and  glycoprotein  reducing  renal  clearance   The   presence   of   sialic   acid   on   cell   membrane   surface   proteins   within   the   glomerular   basement   membrane   creates   over-­‐expression   of   a   negative   charge   barrier   alongside   the   glomerular   filtration   barrier   preventing   the   passage   of   glycoproteins   through   charge   repulsion.   This   is   applicable   to   therapeutic  glycoproteins  as  a  possible  method  to  increase  circulatory  half-­‐ life  (Varki,  2008).   2.  Prevention  of  proteoylatic  degradation     Sialic   acid   occupies   a   comparatively   large   volume   in   glycosylation,   which   protects   the   underlying   peptides   from   protease   detection   and   cleavage   within   the   circulation   (e.g.   papain)   (Sola   and   Griebenow,   2009).     A   study   carried   out   by   Raju   and   Scallon   (2007)   demonstrated   that   removal   of   terminal   sugars   from   Fc   antibody   fragments   (Antibodies   are   naturally   glycosylated   in   the   CH2   of   the   Fc   fragment)   resulted   in   increased   sensitivity   to  papain.   3.  Reduced  receptor  binding  affinity     Elliott  et  al.  (2004)  showed  that  non-­‐glycosylated  erythropoietin  (EPO)  has   seven   times   greater   receptor   binding   affinity   than   native   EPO   (highly   glycosylated).  This  is  due  to  the  presence  of  the  negative  charge  created  by   sialic  acid,  which  reduces  binding  affinity  via  charge  repulsion.  Also,  in   vivo   activity   of   non-­‐glycosylated   analogue   was   significantly   decreased   due   to   increased  receptor  binding  affinity  resulting  in  increased  internalization  and   protein  degradation  (Elliott  et  al.,  2004).       32   4.  Stabilization  of  the  protein     Glycosylations   offer   stability   for   therapeutic   proteins   from   two   points   of   view:   Glycosylation  may  help  decrease  the  immunogenicity  of  polypeptides  in  two   ways:  Walesh  and  Jefferis  (2006)  reported  that  one  potential  mechanism  is   via   increased   solubility   of   polypeptides   due   to   interaction   with   H2O   and   shielding   of   hydrophobic   residues   reducing   the   possibility   of   aggregate   formation  that  could  form  a  stationary  precipitate  for  antibody  recognition.     Another   possible   mechanism   is   via   shielding   the   polypeptide,   reducing   the   accessible   surface   area   exposed   for   antibody   recognition,   resulting   in   a   masking  effect  provided  by  sialic  acid.  This  is  shown  with  Darbepoetin  alpha   (a   rhEPO   analogue   composed   of   two   additional   N-­‐linked   glycosylations   sequences)   whereby   ELISA   tests   were   unable   to   detect   the   protein   in   the   circulation   from   different   patients   (Sinclair   and   Elliott,   2005,   Byrne   et   al.,   2007).         From   a   manufacturing   point   of   view,   glycosylations   offer   stability   for   therapeutic   proteins   against   chemical   denaturation   and   pH   by   increasing   the  potency  of  internal  forces  and  this  is  essential  in  maintaining  therapeutic   protein  conformation.  It  is  also  important  to  increase  the  protein  structural   compactness,   resulting   in   a   decreased   available   surface   area   to   denaturing   pH  or  chemicals  (Sola  and  Griebenow,  2009).       From   the   previous   discussion,   it   can   be   shown   that   N-­‐linked   is   the   most   common   form   of   glycosylation   and   is   thus   preferentially   used   in   many   technologies   of   protein   modification   (Spiro,   2002).   N-­‐linked   strategies   can   be  divided  in  two  classes:   1-­‐   H yperglycosylation   v ia   S ite   D irect   M utagenesis   Hyperglycosylation  is  defined  as  the  process  of  increasing  glycosylation  on  a   protein   to   alter   its   pharmacokinetic   and   biological   activity   (Sola   and   Griebenow,   2010).   DNA   mutagenesis   is   one   of   the   main   strategies   to   increase   glycosylation.   In   vivo,   mutagenesis   of   the   DNA   can   incorporate   33   additional   glycosylation   sites.   This   process   can   be   achieved   by   identifying   Thr/Ser  residues  occupying  the  third  position  in  the  sequence  of  a  protein   and   mutating   the   first   amino   acid   in   the   sequence   to   Asn   or   by   identifying   Asn   residues   within   the   sequence   of   a   protein   and   mutating   the   third   amino   acid   to   Thr/Ser.   For   example,   Elliott   et   al   (2003)   used   site   directed   DNA   mutagenesis  in  the  development  of  Darbepoetin  alpha.  The  process  started   by  mutating  Ala-­‐30,  His-­‐32  to  Asn-­‐30,  Thr-­‐32,  and  Pro-­‐87,  Trp-­‐88,  Pro-­‐90  to   Val-­‐87,  Asn-­‐88,  Thr-­‐90.  All  mutations  were  shown  to  be  glycosylated  with  an   increase   in   molecular   weight   from   35kDa   to   approximately   43kDa,   whilst   maintaining  biological  activity.     Site  direct  mutagenesis  has  also  been  performed  on  GCSF.  Hee  et  al.  (2011)   produced   an   N-­‐linked   glycosylation   site   on   rhGCSF   by   mutating   Phe140   to   Asn140producing   a   novel   form   of   human   GCSF   mutant.   The   new   mutant   rhGCSF  was  shown  to  be  glycosylated  and  more  effective  than  native  GCSF   for  stimulating  differentiation  and  proliferation  of  hematopoietic  cells.     2-­‐   G lycosylated   L inkers   The   uses   of   glycosylated   linkers   are   another   glycan   strategy   that   can   be   utilised   to   extend   the   half-­‐life   of   therapeutic   proteins.   The   glycosylated   linkers  can  be  inserted  between  subunits  of  the  same  ligand  as  observed  in   recombinant   human   Follicle-­‐Stimulating   Hormone   (rhFSH)   between   the   α-­‐   and   β-­‐   subunits   in   which   either   N-­‐   linked   or   O-­‐linked   glycosylated   linkers   were  used.  The  effect  of  the  glycosylated  linker  was  to  increase  the  half-­‐life   of   the   molecules   by   as   much   as   2-­‐fold   compared   to   rhFSH   (Weenen   et   al.,   2004).   This   method   can   also   be   used   to   insert   glycosylated   linkers   between   two   ligands  (tandem  molecules).  The  insertions  of  glycosylated  linkers  between   tandem   molecules   are   preferable   as   they   alleviate   potential   problems   with   direct   glycosylation   of   the   ligand   itself,   which   may   inhibit   bioactivity   and   potentially  introduce  immunogenic  sites.     34   The  Ross  group  (UoS)  have  used  the  advantages  of  both  glycosylated  linker   design   and   tandem   molecules   to   create   a   long   acting   GH   molecules.   They   created   a   tandem   of   two   GH   molecules   joined   by   a   flexible   (Gly4Ser)n     linker   containing   variable   glycosylation   motifs.   The   results   successfully   showed   that  the  use  of  glycosylated  linkers  between  two  GH  ligands  results  in  their   glycosylation  and  increased  molecular  weight  whilst  maintaining  biological   activity   and   delaying   clearance   in   a   rat   model.   This   methodology   could   be   applied  to  other  cytokine  hormones  such  as  GCSF.       35   1.10 Aim  and  Hypothesis   The  role  of  recombinant  human  rhGCSF  has  been  successful  in  the  treatment   of  patients  with  neutropenia  and  stem  cell  mobilization  in  the  circumstance   of   bone   marrow   transplantation,   making   it   a   multi-­‐million   pound   market.   The   fast   growing   market   size   will   increase   appetite   for   generating   similar   rhGCSF.     Therefore,   it   is   essential   to   produce   new   GCSF   compounds   with   improved  properties  over  other  products.     The   aim   of   this   project   is   to   create   a   long   acting   GCSF   with   a   predictable   pharmacokinetic   profile   to   provide   a   more   effective   treatment   for   generating  HSCs  for  transplantation  purposes.       We   hypothesized   that   the   incorporation   of   variable   glycosylation   motifs   (2NAT,   4NAT   and   8NAT)   within   a   flexible   linker   (Gly4Ser)n   between   two   GCSF  ligands  will  Increase  the  molecular  weight  (MW)  of  GCSF  according  to   the   number   of   glycosylation   motifs   with   protection   from   proteolysis   resulting  in  reduced  clearance  with  out  blocking  bioactivity  (Figure  1.10  and   Table   1.3).   This   approach   also   alleviates   potential   problems   with   direct   glycosylation   of   the   ligand,   which   may   reduce   bioactivity   and   potentially   introduce  immunogenic  sites.       Main  Objectives   1-­‐ Design,  cloning  and  expression  of  GCSF  tandems.   2-­‐ Purification  of  GCSF  tandems  from  a  mammalian  cell  line.   3-­‐ Analyse   and   compare   the   biological   activity   of   GCSF   tandems   with   rhGCSF  using  an  in  vitro  proliferation  assay.     4-­‐ Determine   the   pharmacokinetic/pharmacodynamic   properties   of   rhGCSF  &  GCSF  tandems  in  a  normal  rat  model.             36         Figure  1-­‐10:    An  example  of  2NAT  glycosylation  motifs  and  its  control   2QAT  within  a  flexible  linker  (Gly4Ser)n  between  two  GCSF  ligands   (A)   The   glycosylation   motif   2NAT   inserted   to   the   linker   (glycosylated   linker).  (B)  Non-­‐glycosylation  motif  2QAT  control.       Table  1-­‐3:  List  of  GCSF  tandems   GCSF  tandems  linked  via  a  flexible  linker  incorporating  increasing  numbers   of   the   glycosylation   motif   NAT   (glycosylated   molecules)   or   control   motif   QAT  (non-­‐glycosylated  molecules).     Molecule  name   Number  of  NAT/QAT  motifs   Size  of  linker   GCSF8NAT_Hist   8  x  NAT   282bp   GCSF8QAT_Hist   8  x  QAT   282bp   GCSF4NAT_Hist   4  x  NAT   147bp   GCSF4QAT_Hist   4  x  QAT   147bp   GCSF2NAT_Hist   2  x  NAT   147bp   GCSF2QAT_Hist   2  x  QAT   147bp         37   2. Materials     2.1  Cell  Culture   Easy  filtered  flasks  75cm2   Nalgen  Nunc  intl   Easy  filtered  flasks  25cm2   Nalgen  Nunc  intl   100mm  tissue  culture  dishes   Iwaki  SLS   6  well  Cell  Culture  plates   Costar   24  well  Cell  Culture  plates   Costar   CHO  Flp-­‐In  cell  lines   (Invitrogen  Corp)  Gibco   Cryogenic  vials   Nalgen  Nunc  intl   Dimethylsulfoxide  (DMSO)   Sigma-­‐Aldrich   Dulbecco’s  Modified  Eagles  Medium   Sigma-­‐Aldrich   (DMEM)/F-­‐12       Fugene-­‐6   Roche  Diagnostics   Foetal  Calf  serum  (FCS)   Labtech   HyClone  SFM4CHO  Cell  Culture  Media                                                Thermo             Scientific   Hygromycin  B  Antibiotic  (50mg/ml)   (Invitrogen  Corp)  Gibco   Ham’s  F12   (Invitrogen  Corp)  Gibco   Labtech  chambers  coverglass   Nalgen  Nunc  intl   L-­‐Glutamine   (Invitrogen  Corp)  Gibco   Phosphate  Buffer  10X   (Invitrogen  Corp)  Gibco   Penicillin-­‐Streptomycin  (100µg/ml)   (Invitrogen  Corp)  Gibco   pOG44         (Invitrogen  Corp)  Gibco   Trypsin-­‐EDTA   (Invitrogen  Corp)  Gibco   Zeocin  Antibiotic  (100mg/ml)     (Invitrogen  Corp)  Gibco   Counting  chamber     Hawksley   Trypan  blue     Sigma         38   2.2 DNA  Manipulation   1  kb  ladder   (Biolabs)  New  England   100  bp  ladder     (Biolabs)  New  England   Agar   Sigma-­‐Aldrich   Agarose   Sigma-­‐Aldrich   CaCl2     Sigma-­‐Aldrich   Dithiothreitol  (DTT)   Sigma-­‐Aldrich   dNTPs     Upjohn  and  Pharmacia   DNA  loading  dye  (6X)   Promega   EDTA   BD  Biosciences   Expand  high  fidelity  PCR  system   Roche  Diagnostics   Ethidium  bromide  (0.5μg/ml)                                                                                  Sigma-­‐Aldrich                   Gen  EluteTM  Gel  Extraction  Kit   Sigma-­‐Aldrich   Glycerol   BD  Biosciences   KCl   BD  Biosciences   MgCl2   Sigma-­‐Aldrich   NaCl     BD  Biosciences   QIAGENprep  Spin  Miniprep  Kit   QIAGEN   QIAGENquick  PCR  Purification  Kit   QIAGEN   QIAGENquick  Gel  Extraction  Kit   QIAGEN   Restriction  enzymes   Promega   Synthesised  DNA  primers   MWG  Biotech   T4  DNA  ligase   Promega   Taq  polymerase   Promega   Tryptone   Melford   Tris  –base   Sigma-­‐Aldrich                                       39   2.2.1 Restriction  Endonucleases   2.2.1.1 Enzyme   Age1   New  England  Biolabs   BamH1   Promega   Nhe1         Promega   Xho1   Promega     2.2.2 Bacterial  Cell  Culture   2.2.2.1 Antibiotics   The  table  shows  the  antibiotics  used  and  their  final  concentrations:   Antibiotics   Final  concentration  of  antibiotic   Carbenicillin   100μg/ml     Kanamycin   10μg/ml   SURE  cells   Genotype;  e14-­‐(McrA-­‐)  Δ(mcrCB-­‐hsdSMR-­‐mrr)171   endA1  gyrA96  thi-­‐1  supE44  relA1  lac  recB  recJ  sbcC   umuC::Tn5  (Kanr)  uvrC  [F  ́  proAB  lacIqZΔM15  Tn10)     2.2.2.2 Media   The  table  explains  the  different  bacterial  growth  media  used  in  this  project:   Media   SOC  medium   Luria-­‐  Bertani  (LB)  medium   Formula   0.5%  (w/v)  yeast  extract   2.5mM  KCl   1mM  MgSO4·7H2O   1mM  MgCl2·6H2O   10mM  NaCl   2mM  glucose   2%  (w/v)  tryptone   Supplier         Invitrogen   1%  sodium  chloride,     1%  tryptone  and     0.5%  yeast     In-­‐house   40   2.3 Protein  Analysis     Acetic  acid   Fisher  chemical     Ammonium  persulfate   Merck               Ammonium  Sulphate   Sigma-­‐Aldrish   Avidin-­‐HRP   Biolegend   30  %  Acrylamide/Bis  solution  (Geneflow)   National  diagnostic   Benzamidine   Sigma-­‐Aldrish       Roller  bottle  (2L)   Greiner  Bio-­‐one   Bovine  gamma-­‐globulin   Sigma   Bovine  Serum  Albumin  (BSA)   Sigma   Bradford  reagent   Bio-­‐Rad     [BVD13-­‐3A5]  Purified  anti-­‐human  GCSF   Biolegend   [BVD11-­‐37G10]  Biotin  anti-­‐human  GCSF   Biolegend   Coomassie  Blue  reagent   Sigma-­‐Aldrich   Chemiluminescence  Blotting  substrate   Roche   Cuvettes  (polystyrol/polystyrene   SARSTEDT   EDTA   BDH   Ethanol   Fisher  Scientific       ECL  western  blotting  detection  system   Amersham-­‐Pharmacia   Imidazole       Sigma   2X  Laemmli  sample  buffer       Bio-­‐Rad   Fuji  Medical  X-­‐ray  Film   Fuji   Glacial  acetic  acid   Fisher   Glycine   Sigma-­‐Aldrich   Glycerol   Fisher  Scientific   Vortex  mixer   Stuart   Traceable  Nano  Timer   Fisher  Scientific   Methanol   Merck   Minispin  centrifuge   Eppendorf   Mixing  Table   Biotek  instrument  Inc   NaHCO3       Sigma   NaN3  [AnalR]   BDH   41   Nickel  Chloride                                                                                                                                            Sigma           PBS      Tablets                                                                                                                                                      OXOID                     Plate  Reader  (450nm/630nm)                                                                                  Yellowline           OS10  Basic   Polyvinylidene  diflouride  membranes                                                      Amersham-­‐Pharmacia                     96  well  ELISA  plates                                                                                                                        Costar             Rat  Serum                                                                                                                                                              Sigma             Recombinant  human  GCSF  @  200μg/ml                                              Biolegend           Sodium  acetate  (anhydrous)                                                                                          Fisher           Scientific   Sodium  hydroxide   BDH   Sodium  Lauryl  Sulfate  (SDS)   Sigma-­‐Aldrich   Sodium  phosphate  monobasic                                                                                    BDH           Sodium  phosphate  dibasic                                                                                                  BDH           Spectrophotometer   Unicam   Sprague  male  Dawley  Strain  rats  (~300  gram)                    Royan             Institute   (Esfahan)                           Sulphuric  Acid     Tetramethylethylenediamine  (TEMED)   Merck  BDH   TMB  Solution                                                                                                                                                  Sigma                                 TRIS    Base                                                                                                                                                              Sigma                                 Tris-­‐Hcl   Sigma-­‐Aldrich   Tween  20   Sigma-­‐Aldrich   Vivaflow  200  concentrator   Sartorius  Stedin           1ml  Blue  Sepharose  column                                                                                            PIERCE           1ml  IMAC  HP  column                                                                                                                    GE             Healthcare   3510  pH  Meter                                                                                                                                            JENWAY                                                                   42   2.3.1  Proliferation  Assay   AML-­‐193  cell  line                                                                (ATCC;           Cat.  #:  CRL-­‐9589;  Lot  #:  3475266)   Iscove’s  Modified  Dubellco  Media  (IMDM)                                        Gibco             L-­‐glutamine                                                                                                                                                            Gibco             Penicillin/Streptomycin                                                                                                              Gibco                                 Foetal  Calf  Serum                                                                                                                                      Labtech             Transferrin   Sigma   Insulin   Sigma   GMCSF  (5ng/ml)   Biosource   CellTitre  96  AQueous  Proliferation  Assay  Reagent          Promega             96-­‐well  Cell  Culture  Plate                                                                                                        Costar                                                                                                                                         43   3. General  Methods   3.1 Preparation  of  Luria-­‐Bertani  (LB)  Media   LB  media  comprised  of  1%  (w/v)  sodium  chloride,  1%  (w/v)  tryptone  and   0.5%   (w/v)   yeast   dissolved   in   MilliQ   water   and   then   autoclaved   to   fully   dissolve  these  components  as  well  as  sterilize  the  media.     3.2  Preparation  of  Agar  Plates   Agar   plates   were   made   up   by   adding   granulated   agar   to   LB   media   at   a   concentration   of   1.4%   (w/v)   and   autoclaved.     Appropriate   antibiotics   (carbencillin   at   100μg/ml)   were   added   to   the   agar   plates   when   cooled   to   approximately  60°C.  Agar  plates  were  allowed  to  set  and  dry.   3.3  Preparation  of  Chemically  Competent  Cells   Replicating   eukaryotic   DNA   in   E.   coli   can   be   difficult   because   eukaryotic   genes   might   incorporate   inverted   repeats   or   secondary   structures,   such   as   Z-­‐DNA,  that  can  be  deleted  or  rearranged  by  conventional  E.  coli.   The   SURE   (Stop   Unwanted   Rearrangement   Events)   cell   is   particular   strain   of   E.   coli   and  has  a  unique  advantage  over  conventional  E.  coli  commonly  used,  as  it  is   void   of   essential   components   of   the   pathways   that   hinders   cloning   of   eukaryotic   DNA   in   conventional   E.   coli   (largely   due   to   induction   of   DNA   rearrangement  and  deletion  of  nonstandard  DNA  fragments).    It  was  used  in   this  study  to  allow  for  cloning  of  multiple  repeat  DNA  sequences.   To  produce  competent  E.  coli  cells  with  the  ability  to  accept  foreign  DNA,  the   E.   coli   SURE   cells   were   plated   out   on   agar   plate   containing   selective   antibiotic   (Kanamycin   at   10μg/ml)   and   incubated   overnight   at   37°C.   The   next   day,   a   single   isolated   colony   was   selected   and   grown   in   LB   media   containing  Kanamycin  10μg/ml  during  the  day  in  an  orbital  shaker  at  37°C.   A  dilution  of  1/1000  was  seeded  into  50ml  LB  media  and  grown  continued   overnight  (37°C)  at  200rpm  in  an  orbital  shake.  The  following  day,  50ml  of   overnight   culture   was   transferred   into   500ml   warmed   LB   and   grown   until   an  optical  density  (measured  at  600nm)  of  approximately  0.9  was  reached.   44   Cells  were  centrifuged  at  2397  x  g  for  30  minutes  at  4°C  in  a  Beckman  Avanti   J201centrifuge.   The   supernatant   was   discarded   and   pellet   resuspended   in   200ml   ice   cold   sterile   0.1M   MgCl2   and   centrifuged   again   using   previous   parameters.   Following   removal   of   supernatant,   the   cell   pellet   was   resuspended   in   10ml   ice   cold   sterile   0.1M   CaCl2   followed   by   a   further   addition  of  90ml  0.1M  CaCl2  lower  case  and  left  on  ice  for  1  hour  (hr)  ,  then   centrifuged  as  before.  The  supernatant  was  discarded  and  cells  resuspended   by  gentle  swirling  or  pipetting  in  25ml  sterile  15%  glycerol/85mM  ice  cold   CaCl2.  Finally,  cells  were  aliquoted  into  1.5ml  pre-­‐chilled  sterile  eppendorfs   on  ice  and  immediately  flash  frozen  in  liquid  nitrogen  then  stored  at  80°C.   3.4 DNA  Cloning  of  GCSF  Tandems  for  Expression   3.4.1 Polymerase  Chain  Reaction  (PCR)   In   this   project,   the   PCR   was   used   for   DNA   fragment   amplification   and   to   introduce   restriction   enzymes   sites   into   the   expression   constructs   for   ease   of   cloning.   The   GCSF   molecule   was   PCRed   from   pGCSFSecTag4A1   (GCSF   fusion  protein  construct  containing  full  length  GCSF  and  its  signal  sequence   which  was  available  in  the  laboratory)  using  a  forward  primer  and  reverse   primer   giving   a   final   amplicon   of   ~600bp   (all   primers   used   in   this   study   are   listed   in   Appendix   A.1).   The   PCR   reactions   were   set   up   as   a   single   step   reaction   utilizing   two-­‐master   mixes   as   highlighted   in   the   Table   3.1.   A   negative   control   reaction   that   contained   primers   but   no   template   was   also   included.   Two   GCSF   molecules   were   PCRed   in   this   project,   GCSF   and   GCSF   with  signal  sequence.       45   Table  3-­‐1:  PCR  reaction  utilizing  two-­‐master  mixes          Master  Mix  1      Volume                  Master  Mix  2   Volume   Template   DNA   1μl  (100ng)   10x   polymerase   buffer   +   5μl   (plasmid)   MgCl2  (1.5mM)   Primer   1   (10pmol/µl   1μl   Sterile  water   19.15μl   Expand  Polymerase   0.85μl   =  10µM)   Primer   2   (10pmol/µl   1μl   =  10µM)   dNTPs  (10mM)   1.25μl   Sterile  water   20.75μl       Using   the   TC-­‐312   PCR   thermocycler,   master   mix   one   and   two   were   combined  and  cycled  as  follows:         Table  3-­‐2:  PCR  stages     Stage   Time   Temperature   Denaturation   2  min   94°C   25  cycles  @   30s   94°C   30s   54°C   30s   72°C   10min   72°C       Extension     Thereafter,  the  PCR  reactions  were  loaded  on  1%  (w/v)  TAE  agarose  gel.  A  1   Kb  DNA  ladder  runs  alongside  the  samples  during  electrophoresis  at  100  V   for   30   minutes   using   Bio-­‐Rad   PowerPac™   HC   and   gel   isolated   (see   section   3.1.5.2)  Gene  Snap  software  from  SYGENE  was  used  for  the  gel  Imaging.       46   3.4.2 DNA  Isolation  from  Agarose  Gel  Electrophoresis     A   1%   (w/v)   agarose   gel   was   prepared   by   dissolving   powdered   agarose   in   TAE   buffer   (1mM   EDTA,   20mM   acetic   acid   and   40mM   Tris   base,   pH   8.5)   upon   heating   for   1   minute   will   not   dissolve   otherwise.   0.5μg/ml   Ethidium   bromide  was  used  as  staining  agent.  1kb  DNA  ladder  (0.5μg/μl)  was  run  on   each   gel   as   standard.   The   agarose   gel   was   run   at   100V   using   the   Bio-­‐Rad   Power  PC.   To   prevent   damage   to   the   amplified   DNA   for   the   purification   step,   10%   of   the   samples   (PCR   reaction)   were   loaded   in   an   inner   well   for   visualization   with  ultraviolet  (UV)  light  using  the  Gene  Snap  software  from  SYGENE,  while   the   remaining   samples   were   loaded   in   a   peripheral   lane.   This   was   done   to   prevent   potential   damage   to   the   remaining   90%   DNA   samples   (in   the   peripheral  lane)  by  UV  light  during  visualization.  The  peripheral  lane  of  the   gel  was  separated  from  the  rest  of  the  gel  using  a  clean  razor  blade  prior  to   UV   exposure.   After   visualization   and   excision   of   the   DNA   fragment   of   interest   in   the   main   gel,   the   corresponding   region   on   the   peripheral   lane   was   then   cut   out   from   the   agarose   gel,   which   was   thereafter   extracted   using   GenElute  extraction  kit  as  per  manufacture’s  protocols  in  order  to  get  rid  of   the  PCR  contaminates.  DNA  samples  were  eluted  by  the  addition  of  50μl  of   buffer  EB  (10mM  Tris-­‐HCl,  pH  8.5)  and  quantified  at  a  wavelength  of  260nm   using   the   Nanodrop   spectrometer   ND100V.   Samples   were   then   ready   for   ligation  into  plasmid  or  stored  at  -­‐20°C  to  preserve  DNA.   3.4.3 Single  and  Double  Restriction  Enzyme  Digests   Restriction  enzymes  are  routinely  used  during  cloning  to  transfer  a  gene  of   interest  into  a  vector  plasmid.  Often,  both  the  vector  plasmid  and  the  target   gene   are   digested   with   the   same   restriction   enzymes   to   generate   similar   overhangs.   The   target   gene   to   be   expressed,   often   referred   to   as   insert,   is   thereafter  ligated  to  the  vector  plasmid  using  DNA  ligase  and  polymerase.         47   Restriction  Digestion  of  Insert  and  Vector  DNA     In  this  project,  single  or  double  restriction  digestions  (involving  use  of  one   or   two   restriction   enzymes   respectively)   were   set   up   for   both   the   insert   (PCR  product  of  target  DNA)  and  the  vector  DNA  (i.e.  the  expression  plasmid   DNA)  as  stated  below  in  Table  3.3  and  3.4  respectively.     1-­‐  Digestion  of  insert   For   the   digestion,   the   volume   of   DNA   used   was   dependent   on   the   sample   concentration.   The   volume   was   made   up   to   50μl   with   sterile   MilliQ   water.   The  reaction  mixture  was  incubated  for  2  hrs  in  a  thermostatic  water  bath  at   37°C   (all   restriction   enzymes   and   restriction   enzyme   buffers   used   in   this   study  are  given  in  Appendix  A.2).     Table  3-­‐3:  Single  step  double  digestion  of  Insert   Undigested  DNA  (PCR  fragment)   500ng   10x  restriction  Buffer     5μl   Restriction  enzyme  1    (10units  per  0.5-­‐1μg  of  DNA)   1μl   Restriction  enzyme  2  (10units  per  0.5-­‐1μg  of  DNA)   1μl   ac  BSA  (1mg/ml)   5μl   Sterile  MiliQ  water   Xμl   Total  volume   50μl     After   incubation,   samples   were   centrifuged   at   11,337   x   g   for   5   seconds   to   remove   condensation   and   then   analyzed   using   1%   agarose   gel   electrophoresis   Reaction   products   were   purified   from   agarose   gel   using   QIAquick   PCR   purification   kit   as   per   manufacture’s   protocol.   Purified   elutions   were   quantified   using   the   Nanodrop   spectrometer   ND   100   at   260nm.     48   2-­‐  Digestion  of  plasmid  DNA   The   plasmid   DNA   was   digested   in   a   two-­‐step   double   digestion.   The   double   digestion  was  carried  out  by  setting  up  single  restriction  digestions  as  stated   below   and   incubated   for   2   hrs   in   a   thermostatic   water   bath   at   37°C.   The   volume  was  made  up  to  10μl  with  sterile  MilliQ  water.   Table  3-­‐4:  Two-­‐step  double  digestion  of  Plasmid  DNA   Single   digest   with   Single   digest   with   enzyme  1   enzyme  2   Plasmid   1μl  (500ng)   1μl  (500ng)   10x  restriction  Buffer   1μl   1μl   Restriction   enzyme   1     1μl   -­‐   (10units  per  0.5-­‐1μg  of  DNA)   Restriction  enzyme  2   -­‐   1μl   ac  BSA  (1mg/ml)   1μl   1μl   Sterile  MiliQ  water   6μl   6μl   Total  volume   10μl   10μl    After   incubation,   2.5μl   aliquots   from   each   single   digest   were   taken   for   further  analysis  to  assess  the  integrity  of  the  restriction  enzymes  used.  The   remainder   volumes   in   the   single   digests   (7.5μl)   were   pooled   together   and   the   following   reagents   were   added   (2μl   acBSA,   2μl   10x   restriction   Buffer,   1μl   Restriction   enzyme   1,   1μl   Restriction   enzyme   2   and   14μl   Sterile   MilliQ   water)  to  give  a  final  total  volume  of  35μl.  The  reaction  mixture  containing   both   restriction   enzymes   was   then   incubated   for   1-­‐2   hrs   in   a   thermostatic   water  bath  at  37°C.   1%  agarose  gel  electrophoresis  was  used  to  analyse  samples  from  the  single   and   double   digestions   as   described   previously   in   section   3.4.2.   When   required,   reaction  products  were  purified  from  agarose  gel  using  QIAquick   PCR   purification   kit   as   per   manufacture’s   protocol.   Elutions   were   quantified   using  the  Nanodrop  spectrometer  ND  100  at  260nm.   49     3.4.4  General  DNA  Ligation     Fragments   and   linkers   were   ligated   into   the   pSecTag   plasmid   to   create   mammalian   expression   vectors.   Two   different   controls   (one   without   insert   and,  another  without  insert  and  DNA  ligase)  were  used  in  each  reaction  to   ascertain  the  success  of  DNA  ligation.  For  each  ligation,  the  amount  of  insert   (I)  used  was  calculated  using  the  formula  below  from  the  Promega  website.   A  molar  ratio  of  insert  to  plasmid  (P)  of  3:1  was  used.       ng  of  insert  (I)    =     (ng  of  P  x  Kb  size  of  I)  x  molar  ratio  of  I:P  (3/1)                    Kb  size  of  P                                             The   volume   of   MilliQ   water   was   added   to   compensate   for   the   variation   in   volume  of  fragment  (insert)  to  achieve  a  total  reaction  volume  of  10μl.  The   Plasmid  (P)  typically  used  in  this  method  was  around  50ng.   Ligation   reactions   were   set   up   as   shown   in   the   table   below   and   left   overnight  at  room  temperature.     Table  3-­‐5:  Preparation  of  ligation  reactions     (Experimental     ligation)   PCR   fragment   (insert)   Variable   or  linkers   Plasmid  vector   X  (~50ng)   Control  1   Control  2   -­‐   -­‐   Xμl   Xμl   1μl   1.5μl   1.5μl   Variable   X  μl   X  μl   T4  DNA  ligase  (3U/μl)   1μl   1μl   -­‐   Total  volume   10μl   10μl   10μl   10x  ligase  buffer   MilliQ  water     The   following   day,   the   ligation   reactions   were   subsequently   transformed   into   SURE   chemically   competent   E.coli   cells,   for   vector   amplification   and   initial  assessment  of  ligation  success.           50   3.4.5 Transformation  of  Plasmid  into  Chemically  Competent   E.coli   To   determine   if   the   target   gene   of   interest   or   linkers   has   been   integrated   into   the   vector   plasmid   DNA,   the   expression   plasmid   constructs   were   transformed   into   competent   E.   coli  cells.   To   transform   the   cells,   200μl   of   the   chemically   competent   E.coli   cells   were   added   into   each   eppendorf   and   a   suitable   volume   of   the   plasmid   of   interest   added   (no   more   than   10%   total   volume  of  cells),  all  this  was  carried  out  on  ice.  Cells  were  gently  mixed  and   incubated   on   ice   for   15   minutes.   Heat   shocking   at   42°C   for   1   minute   was   used   to   enhance   the   bacterial   cell   wall   permeability,   allowing   uptake   of   plasmid.  After  that,  cells  were  immediately  chilled  on  ice  for  an  additional  5   minutes.   Under   sterile   condition,   800μl   SOC   media   was   added   to   each   eppendorf  of  cells  and  incubated  for  a  further  40-­‐45  minutes  at  37°C.  After   incubation,   cells   were   centrifuged   for   5-­‐10   minutes   at   1073   x   g   using   the   bench  top  centrifuge  at  room  temperature.    The  supernatant  was  discarded   and   cell   pellet   resuspended   in   100μl   SOC   media.   Transformed   cells   were   plated   out   on   LB   agar   plates   containing   100μg/ml   carbenicillin   at   two   dilutions;  10%  (1  tenth  cells  diluted  in  90μl  SOC  media)  and  90%  (9/10  of   cells)   to   accommodate   any   potential   overgrowth   of   colonies,   and   allow   isolated  colony  selection.  Plates  were  incubated  overnight  at  37°C.  Next  day,   colony   counts   were   carried   out   for   each   experimental   ligation   and   both   controls  to  evaluate  whether  DNA  ligations  had  been  successful.   3.4.6 Plasmid  Preparation  and  Glycerol  Stocks   Small-­‐scale   copies   of   expression   plasmid   are   generally   made   to   propagate   transformed   cells   from   ligation   reactions   and   for   initial   analyses   of   expression  constructs  for  protein  expression  in  mammalian  cells.  A  number   of   colonies   from   the   previously   transformed   competent   E.   coli   cells   containing   the   expression   plasmid   were   selected   and   seeded   into   1ml   LB   broth   containing   1μl   of   100mg/ml   carbenicillin.   Colonies   were   then   incubated  during  the  day  at  37°C  in  orbital  shaker  (200rpm).  Thereafter,  a   1/1000  (5μl)  dilution  of  each  was  seeded  into  5ml  LB  broth  with  100μg/ml   51   stock  carbenicillin  and  incubated  overnight.  On  the  following  day,  200μl  of   50%   (v/v)   glycerol   (important   for   long-­‐term   storage   of   bacteria,   as   it   prevents   the   formation   of   ice   crystals   that   can   damage   the   cell   wall)   were   added   to   800μl   of   each   clone   and   stored   at   -­‐80°C.   Up   to   5ml   plasmid   DNA   from   overnight   cultures   were   purified   using   a   Qiagen   plasmid   mini   preparation  kit  as  per  the  manufacturer’s  procedure.  The  concentrations  of   purified   plasmid   were   quantified   using   a   Nanodrop   spectrometer   ND100V   at  a  wavelength  of  260nm.   3.4.7 Screening  of  Potential  Clones  from  Ligations   For  initial  screening  of  the  clones  so  as  to  identify  potential  positive  clones   with   insert   prior   to   sending   for   sequencing,   plasmid   samples   from   mini   preps  were  double-­‐digested  with  the  restriction  enzymes  that  were  used  for   the   cloning.     Plasmid   digestion   was   confirmed   using   agarose   gel   electrophoresis   as   outlined   before   at   section   3.4.3.   Positive   plasmids   were   sent  for  DNA  sequencing.     3.4.8 Plasmid  Sequence  Analysis   All   plasmid   samples   with   an   appropriate   forward   and   reverse   primers   were   sent   to   the   Department   of   Core   Genomic   within   the   Sheffield   University   School   of   Medicine   to   confirm   that   DNA   cloning   had   generated   a   vector   of   desired   nucleotide   sequence.   The   genomic   sequence   results   were   analyzed   using  SeqMan  (DNASTAR  Lasergene  software).     3.5  Cell  Culture  and  Protein  Expression   3.5.1 Growth  and  General  Maintenance  of  CHO  Flp-­‐In  Cells   Mammalian  CHO  Flp-­‐In   cells  were  used  for  protein  expression  of  the  tandem   GCSF   constructs,   which   are   the   subjects   of   our   studies.   From   the   liquid   nitrogen  stock  of  untransfected  Chinese  hamster  ovary  cells  (CHO  Flp-­‐In),  a   vial   (at   2-­‐3   x   106   cells/ml)   was   removed,   thawed   and   grown   in   growth   media   (Ham's/F12   DMEM   media   containing   10%   Fetal   Calf   Serum   (FCS),   2mM   L-­‐glutamine,   and   100µg/ml   Streptomycin   /   Penicillin   and   100μg/ml   52   Zeocin)  and  were  grown  in  a  5%  CO2  incubator  at  37°C.     For   passaging,   the   adherent   cells   were   removed   by   the   addition   of   trypsin/EDTA   and   incubated   2-­‐3   minutes.   The   resulting   suspension   was   then  diluted  with  complete  medium  transferred  to  a  sterile  30ml  universal   and   centrifuged   for   5   minutes   at   67   x   g   to   remove   supernatant.   The   cell   pellet  was  resuspended  in  the  appropriate  media.  Cells  were  maintained  at   approximately   80-­‐90%   confluence   in   a   T-­‐75   flask.   At   this   density   the   cells   are  routinely  passaged  at  1  in  10  dilution.     3.5.2 Trypan  Blue  Exclusion  Method   Trypan  blue  is  used  to  stain  cells  to  differentiate  between  live  and  dead  cells   during   cell   counting.   The   dead   cells   exclusively   take   up   the   blue   dye   and   thus   appear   blue   when   visualized   under   a   light   microscope.   This   distinguishes   them   from   the   live   cells,   which   appear   bright.   This   characteristic   of   dead   cells   differentiation   is   particularly   important   when   assessing   the   viability   of   the   cells   culture.   Using   heamocytometer   (used   to   count   cells   number   in   a   specific   volume   of   fluid   to   give   an   approximate   number  of  cells  in  the  whole  fluid),  Trypan  blue   was  applied  at  1:1  dilution   with   media   sample.   The   total   number   of   cells   was   calculated   per   ml   using   the  below  equation:                      Total  cell  number  per  ml=  total  cell  number  x  Dilution  factor  x  104                                                                                                        No  of  squares       3.5.3 Transient  Expression  of  GCSF  Tandems  in  CHO  Flp-­‐In     All   GCSF   constructs   (plasmids)   were   transiently   transfected   into   CHO   Flp-­‐In   cells   as   an   initial   screen   for   protein   expression   prior   to   development   of   a   stable  cell  line.   CHO   Flp-­‐In   cells   were   grown   in   antibiotic   free   media   (Ham's/F12   DMEM   media   containing   10%   Fetal   Calf   Serum   (FCS),   2mM   L-­‐glutamine,   and   100µg/ml   Streptomycin   /   Penicillin).   At   approximately   70%   confluence,   cells   were   removed   from   the   plate   by   the   addition   of   trypsin/EDTA   and   53   reseeded  at  a  density  of  0.2  x  106  cells/well  into  a  24  well  plate  on  the  day   prior   to   transfection.   Cells   were   allowed   to   grow   overnight   in   antibiotic   free   media.   The   following   day,   media   was   replaced   with   500μl   antibiotic   free   growth  media.    A   transfection   reaction   mix   was   prepared   using   a   Fugene-­‐6   transfection   reagent   (transfection   reagent   designed   to   transfect   plasmid   DNA   into   multiple   cell   lines   with   low   toxicity   and   high   efficiency)   to   experimental   DNA  ratio  of  3:2,  in  a  volume  of  100μl  serum  free  growth  media  (Ham's/F12   DMEM   media   containing   2mM   L-­‐glutamine,   and   100µg/ml   Streptomycin   /   Penicillin).  Prior  to  use,  Fugene-­‐6  was  equilibrated  to  room  temperature  and   vortexed.  Then,  6μl  of  Fugene-­‐6  was  added  to  the  reaction  mix,  being  careful   to   avoid   contact   with   the   plastic   sides   of   eppendorf,   and   followed   by   the   addition   of   4μg   of   plasmid.   Transfection   with   a   known   plasmid   that   expresses   well   (pGCSFSecTag4A1)   was   used   as   a   positive   control.   Non-­‐ transfected   cells   were   used   as   a   negative   control   in   each   experiment.   The   transfection   mixes   were   gently   flicked   to   mix   and   incubated   at   room   temperature  for  15  minutes.  Each  transfection  mix  was  pipetted  drop-­‐wise   into  its  corresponding  well  and  gently  rocked  to  ensure  an  even  distribution.   All  plates  were  incubated  at  37°C  with  5%  CO2  for  72  hrs  prior  to  harvesting   the   media.   Finally,   the   media   for   each   transfection   (containing   secreted   product)   was   centrifuged   and   supernatant   was   transferred   to   a   clean   universal   tube   for   protein   expression   analyses   using   any   available   convenient  method  (western  blot  or  Elisa).     3.5.4 Generation   o f   S table   C HO   F lp-­‐In   C ell   l ines   The   expression   for   GCSF   tandems   was   enabled   using   the   Invitrogen   CHO   Flp-­‐In   system.   The   CHO   Flp-­‐In   system   was   chosen   since   it   enabled   rapid   integration   of   a   GOI   into   a   specific   site   within   the   host   genome   for   high   expression.       The  CHO  Flp-­‐In  host  cell  line  has  a  single  Flp  recombinase  target  (FRT)  site   located   at   a   transcriptionally   active   genomic   locus   and   is   resistant   to   the   antibiotic   zeocin.   The   plasmid   (A   modified   version   of   pSecTag-­‐V5/FRT-­‐Hist:   54   see   Appendix   C)   contains   the   gene   of   interest   as   well   as   an   FRT   site   and   has   the  hygromycin  B  resistance  gene.     Stable   cells   are   generated   by   co-­‐transfection   of   the   plasmid   containing   the   gene   of   interest   (GOI)   and   another   plasmid,   pOG44   (a   5.8   kb   Flp-­‐ recombinase   expression   vector   responsible   for   Flp-­‐recombinase   expression)  into  the  Flp-­‐In  cell  line.     The  expression  of  Flp  recombinase  results  in  the  integration  of  the  GOI  into   the  genome  via  the  FRT  site.    Stable  cells  are  then  selected  using  hygromycin   B,   thus   the   need   for   clonal   selection   is   not   necessary   as   integration   of   the   DNA  is  directed.  The  process  of  culturing  the  Flp-­‐In  cell  line  was  conducted   as  per  manufacture’s  protocol  using  basic  cell  culture  techniques.     The   day   before   transfection   cells   were   removed   from   a   T75   flask   by   the   addition   of   trypsin/EDTA   and   reseeded   into   6   well   plates   at   a   concentration   of  0.25  x  106  cells  per  well  in  a  total  volume  of  2ml.  Cells  were  left  overnight   to  achieve  about  60-­‐70%  confluency.   Transfections   were   carried   out   the   next   day.   Briefly,   a   number   of   sterile   eppendorf   tubes   each   containing   92.5μl   serum   free   media   (without   antibiotics)  were  set  up.  To  each  tube,  7.5µl  of  Fugene-­‐6  was  slowly  dropped   onto  the  surface  of  media  and  flicked  to  mix.  In  separate  sterile  eppendorfs,   250ng  of  plasmid  of  interest  was  mixed  with  5μg  of  the  pOG44  plasmid  and   pipetted   into   the   tube   containing   Fugene-­‐6   mix   and   contents   mixed   gently   by   flicking.   Tubes   were   incubated   at   room   temperature   for   15   minutes.   Fugene-­‐6   only   was   used   as   a   negative   control   and   a   known   expression   plasmid  (pGCSFSecTag4A1)  was  used  as  a  positive  control.  All  transfection   mixes   were   carefully   pipetted   drop-­‐wise   onto   cells   in   each   labelled   individual  well  of  a  6  well  plate,  and  incubated  at  37  °C,  5%  CO2  for  24-­‐48   hrs.         55   At   24-­‐48   hrs   post-­‐transfection,   the   culture   medium   in   each   well   was   removed   and   replaced   with   2ml   growth   medium   (Ham's/F12   medium   containing  10%  FCS,  100µg/ml  Streptomycin  /100IU/ml  Penicillin,  2mM  L-­‐ glutamine)   containing   600μg/ml   Hygromycin   B   antibiotic   as   selective   reagent.   Cells   were   then   allowed   to   grow   until   desired   confluency   with   media  replacement  every  2  days.  Routine  observations  were  made  as  to  the   cells   appearance,   with   successfully   transfected   cells   appearing   fibroblastic   and   growing   out   in   clumps,   whereas,   dead   cells   appeared   spherical   in   solution.   Cells   were   not   allowed   to   get   too   confluent,   as   the   antibiotics   would  become  ineffective.    When  plates  had  a  number  of  individual  colonies   growing   out   they   were   removed   from   the   wells   by   the   addition   of   0.5ml   Trypsin-­‐EDTA   (T/E)   and   centrifuged   at   67   x   g   for   5   minutes   The   supernatant   was   discarded   and   cell   pellet   resuspended   in   5ml   transfected   growth  media  and  grown  in  T-­‐25  flasks  until  cells  became  nearly  confluent   with   media   change   every   2   days.   Once   confluent   cells   were   removed   and   transferred  to  T-­‐75  flasks  with  the  same  growth  media  in  a  total  volume  of   12-­‐15ml.  Cells  were  considered  stable  when  all  control  cells  were  dead  and   cells  were  dividing  normally.  This  took  normally  2-­‐4  weeks.     3.5.5 Analysis  of  Crude  Media  from  Transfected  Cells  Lines   For   protein   expression   analysis   of   stable   &   transiently   transfected   cells   lines,   it   was   required   to   serum   starve   cells   prior   to   analysis.     T-­‐75   flasks   containing   stable   cells   were   grown   until   an   appropriate   confluency   was   reached.  Thereafter,  cells  were  serum-­‐starved:  Media  was  removed  from  the   T-­‐75  flasks  and  replaced  with  12-­‐15ml  serum  free  media  and  incubated  for   2-­‐3  days  at  37  °C,  5%  CO2.  After  incubation  media  was  transferred  to  clean   30ml   universal   tube   and   centrifuged   to   remove   any   cellular   debris.   The   supernatant   of   each   sample   was   then   carefully   removed   to   another   sterile   universal   tube   and   stored   at   4°C   until   required   for   protein   expression   analyses  using  any  available  convenient  method  (western  blot  or  Elisa).     56   3.5.6 Storage  of  Stable  Cell  Lines  in  Liquid  Nitrogen   Stable  cell  lines  were  grown  to  confluency  in  T-­‐75  flasks  at  which  point  cells   were   removed   using   an   appropriate   volume   of   trypsin/EDTA.   Once   cells   were   detached   an   appropriate   volume   of   serum   growth   media   was   added   to   each   flask   to   neutralise   Trypsin/EDTA.     The   suspended   cells   for   all   flasks   were  then  pooled  together  in  a  clean  30ml  universal  tube  and  centrifuged  at   67  x  g  for  5  minutes  to  pellet.  The  supernatant  was  discarded  and  cells  were   resuspended   in   a   freezing   mixture   (foetal   calf   serum/DMSO   mixture   are   ratio   9/1)   to   make   a   final   concentration   between   2-­‐4   x   106   cells/ml.   Once   cells   were   resuspended,   immediately   1ml   portions   were   aliquoted   to   clean   cryogenic   freezing   vials   labelled   with   construct   name,   cell   line,   date   and   cell   number.   Cryogenic   vials   were   finally   placed   in   a   polystyrene   freezer   box   surrounded  with  cotton  wool  and  stored  at  -­‐80°C.  This  step  is  important  to   freeze  cells  gently  and,  prevent  ice  crystal  formation  within  the  stable  cells.   Later,  cells  were  transferred  to  liquid  nitrogen  for  long  time  storage.   3.5.7 Adaptation  of  Stable  CHO  Cells  to  Hyclone  Media   The  stable  cells  taken  from  liquid  nitrogen  were  thawed  and  grown  in  T75   flask   growth   medium   for   1-­‐2   days   until   an   appropriate   confluency.     Cells   were   resuspended   in   Hyclone   SMF4CHO   Utility   media   (no   Hygromycin   B)   for  adaption  and  media  was  change  every  2-­‐3  days.   3.5.8 Expression  of  GCSF  Tandems  in  Roller  Bottle  Culture     After   growing   to   the   maximum   level   in   T75   flask,   stable   cells   were   transferred   to   2   x   1   litre   roller   bottles   (maximum   volume   500ml)   at   a   starting   density   of   ~0.25x106/ml   (each   roller   bottle   contained   about   500ml).   Cells   were   grown   at   37°C   5%   CO2   with   mixing   (Cell   Roll   Cellspine   Control   Unit   was   used   to   mix   the   stable   cells   at   4rpm)   until   viability   was   ~30%,   with   samples   regularly   taken   every   2   or   3   days   to   assess   protein   expression  and  cell  viability.  Thereafter,  the  cells  were  centrifuged  at  14,981   x   g   (JLA   16.250)   for   30   minutes   at   4°C   to   clear   cellular   debris   and   10mM   final   concentration   of   Benzamidine   HCl   (serine   protease   inhibitor)   was   57   added   to   the   1L   media   sample   to   prevent   protein   degradation.   The   media   sample   was   then   concentrated   using   a   vivaflow   concentrator   (See   section   3.6)  and  stored  at  -­‐40°C  until  required.  Samples  were  analysed  by  western   blotting  and  Elisa.   3.6 Vivaflow  200  Concentrator   Vivaflow   200   concentrator   was   used   to   concentrate   media   sample   to   10x   less   volume   (i.e.   from   1L   to   100ml).   This   was   important   to   save   time   and   make   the   process   of   purification   easier.   The   Vivaflow   200   concentration   system  comprised  of  a  membrane  module  with  10kDa  molecular  weight  cut-­‐ off,  inlet  and  outlet  tubes,  and  pump  (Masterflex  pump).  Before  starting  the   process,  the  tubes  were  cleaned  with  0.5M  NaOH  followed  by  washing  with   500ml   of   deionized   water   with   the   filtrate   going   to   the   waste.   The   media   sample   was   then   circulated   through   the   system   for   approximately   3-­‐4   hrs   (the  media  sample  was  placed  on  ice  during  concentration  to  avoid  protein   degradation),   and   the   culture   volume   concentrated   down   to   10   times   less   volume   within   this   period,   followed   by   storage   at   -­‐40°C   until   ready   for   purification.   After   each   concentration   cycle,   deionized   water   was   flushed   through   the   system   with   the   filtrate   going   to   the   waste.   The   system   was   cleaned   by   recirculating   250ml   of   0.5M   NaOH   through   the   system   at   100ml/min   for   30-­‐40   minutes.   Finally,   the   system   was   drained   and   recirculated  with  500ml  deionized  water  for  5-­‐10  minutes  with  the  filtrate   going  to  the  waste.  For  storage,  the  module  was  filled  with  20%  ethanol  and   refrigerated  at  4°C.     3.7 Protein  Purification   3.7.1 Purification  of  GCSF  Tandems  Using  IMAC   Tandem  proteins  were  designed  with  a  His-­‐Tag  at  the  C-­‐terminus.  The  His-­‐ tag  in  each  tandem  facilitates  easy  purification  of  the  desired  protein  when   passed  over  a  metal  chelate  column.    The  high  affinity  of  histidine  for  nickel   ions   allows   the   protein   to   bind   to   the   Immobilized   Metal   Affinity   Chromatography   (IMAC)   column   while   most   of   other   proteins   remain   58   unbound.   The   column   was   washed   with   a   solution   of   moderate   ionic   strength   to   dissociate   proteins   that   may   have   bound   to   the   column   non-­‐ specifically.   The   target   protein   with   His-­‐tag   is   released   from   the   IMAC   by   adding   high   concentrations   of   histidine   analogues   (e.g.   imidazole)   to   compete  with  his-­‐tag  for  nickel  binding.  As  a  result,  the  eluted  target  protein   with   his-­‐tag   could   be   separated   from   any   protein   contaminants   (Block   et   al.,   2009).   The   separated   protein   fraction   undergoes   dialysis   in   PBS   at   4°C   for   2hrs  and  then  overnight  (after  PBS  buffer  change).  The  dialysed  protein  was   stored  at  -­‐80°C.    Step-­‐wise  detail  for  the  purification  step  is  provided  below.   Before  applying  the  media  sample  to  the  1  ml  IMAC  column  (GE  Healthcare),   all   tubing   was   cleaned   with   0.2M   NaOH   and   then   rinsed   with   H2O   prior   to   use.   A   fresh   IMAC   column   was   charged   with   4mg/ml   Nickel   chloride   and   equilibrated   with   equilibration   buffer   (20mM   NaP3,   pH   7.4,   0.5M   NaCl   and   10%   glycerol).   Media   sample   was   defrosted   and   centrifuged   at   30,910   x   g   (JA-­‐25.5)  for  30  minutes  at  4°C  to  clarify.  Media  sample  was  diluted  1:1  with   40mM   NaP3,   pH   7.4,   1M   NaCl,   20%   glycerol   and   20mM   imidazole.   Media   sample   was   then   filtered   (using   Millipore   0.22um   filters)   and   loaded   on   to   the   IMAC   column   at   2ml/minute   (sample   contained   10mM   imidazole)   at   room   temperature   or   sometimes   4°C   to   avoid   protein   degradation.   The   unbound  fraction  (flow  through)  was  collected  and  the  column  washed  with   10   column   volumes   (CV’s)   equilibration   buffer   with   the   addition   of   10mM   imidazole,  followed  by  10  CV’s  with  the  addition  of  10mM  imidazole  at  2ml/   minute.  These  two  washes  were  used  to  remove  contaminants.     Different   concentrations   of   imidazole   were   made   up   in   20mM   sodium   acetate,  0.5M  NaCl,  pH  6.0,  10%  glycerol  (pH  6.0  buffer)  as  shown  in  Table   3.6   were   used   to   elute   the   bound   protein   (Target   protein)   from   the   IMAC   column  using  a  step  elution  method.  1ml  fractions  were  collected  (3  x  1ml   per   concentration).   The   eluted   proteins   were   analysed   by   SDS-­‐ PAGE/Bradford   assay   and   the   relevant   fractions   were   pooled   and   dialysed   against   at   least   1L   PBS   buffer   at   4°C   for   1,   2   hrs   and   overnight   (A   fresh   change   of   PBS   buffer   was   used   per   dialyse)   to   remove   all   salts   and   59   imidazole.   At   the   end   of   each   purification,   the   IMAC   column   was   washed   with  0.2M  NaOH  followed  by  10  CV’s  of  water  and  stored  in  20%  ethanol.    Table  3-­‐6:    Concentrations  of  imidazole   Imidazole  concentration  (mM)   Imidazole  (ml)   pH  6.0  buffer  (ml)   20   0.2   4.8   50   0.5   4.5   100   1   4   200   2   3   350   3.5   1.5   500   5   -­‐   3.7.2 Purification  of  GCSF  Using  Cibacron  Blue  Sepharose     The   stable   cells   of   native   GCSF   (ligand   only)   were   taken   from   liquid   nitrogen,  thawed  and  grown  in  T75  flasks  as  described  in  section  3.5.7  and   3.5.8   and   were   thereafter   transferred   to   4   big   roller   bottles   (maximum   volume   1   liter)   at   a   starting   density   of   ~0.5x106/ml   (each   roller   bottle   contained   about   500ml).   Cells   were   grown   at   37°C   5%   CO2   until   viability   was  ~30%,  with  samples  regularly  taken  every  2  or  3  days  to  assess  protein   expression  and  cell  viability.  Because  native  GCSF  has  no  His-­‐tag,  a  Cibacron   Blue   Sepharose   (www.gelifesciences.com)   column   was   used   to   purify.   This   dye   ligand   chromatography   resembles   native   substrate   which   proteins   have   affinity  for.  When  proteins  pass  through  this  column,  the  goal  of  this  dye  is   to   bind   the   target   protein   and   expel   all   unbound   proteins.   The   bound   protein   could   then   be   eluted   by   changing   the   buffer   composition,   often   by   increasing  the  buffer  pH.   Media  sample  of  recombinant  human  GCSF  (rhGCSF)  was  placed  on  ice  and   precipitated   with   35%   (w/v)   (20.11g/100ml)   ammonium   sulphate   (AmSO4).    The  solution  was  then  incubated  at  4°C  with  mixing  for  1hr.  The   resulting  suspension  was  centrifuged  at  43589  x  g,  4°C  for  30  minutes  and   the   supernatant   was   kept   to   check   if   it   contains   rhGCSF   (1st   supernatant).   60   The  pellet  was  solubilized  in  10ml  20mM  TRIS  buffer  pH  7.0.  On  ice,  the  pH   of   the   media   was   dropped   from   7.0   to   5.0   by   adding   10ml   100mM   sodium   acetate   buffer   and   incubating   on   ice   at   4°C   for   30   minutes.   The   resulting   suspension   was   centrifuged   at   43589   x   g,   4°C   for   30   minutes   and   the   supernatant  was  kept  (2nd  supernatant).  A  1ml  Blue  Sepharose  column  (GE   Healthcare),  was  first  equilibrated  with  10  column  volumes  (CV’s)  of  20mM   sodium   acetate,   pH   5.0   and   then   supernatant   was   loaded   onto   the   1ml   column   at   room   temperature   and   unbound   fraction   collected   (Flow   through).  The  column  was  washed  with  5  CV’s  sodium  acetate  buffer,  pH  5.0   buffer  followed  by  3  ×  5ml  (3  ×  5  CV’s)  20mM  Tris  buffer  pH  7.0.  The  column   was   finally   washed   with   5   ×   2   ml   20mM   Tris,   1mM   EDTA   at   pH   8.0   (the   target   protein   was   observed   to   elute   at   a   high   concentration   in   this   step).   Eluted  proteins  were  stored  at  -­‐80°C.  Finally,  the  column  was  washed  with  5   ×   2ml   of   20mM   Tris,   1mM   EDTA   and   1M   NaCl,   pH   8.0   to   clean   the   column   and   stored   at   4°C.   When   required,   samples   were   analysed   by   coomassie   stain,  western  blotting  and  Elisa.   3.8 Analysis  of  Protein   3.8.1 Bradford  Protein  Assay     The   Bradford   protein   assay   was   routinely   used   to   measure   the   concentration   of   our   purified   tandem   proteins   and   rhGCSF   in   a   solution.   A   10mg/ml  solution  of  BSA  was  prepared  in  double  distilled  water  and  diluted   to   1mg/ml,   then   to   100μg/ml   by   10-­‐fold   dilution   (i.e.   1ml   BSA   +   9ml   ddH2O).   The   standards   below   were   prepared   from   this   100μg/ml   (Table   3.7).           61   Table  3-­‐7:  Standard  curve  preparation     Standard  BSA   Final  concentration   Dilution   (μg/ml)   in  assay  (μg/ml)   25   20   x4   2ml  100μg/ml  +  6ml  ddH2O   12.5   10   x2   2.5ml  25μg/ml  +  2.5ml  ddH2O   6.25   5   x2   2.5ml  12.5μg/ml  +  2.5ml  ddH2O   2.5   2   x2.5   2ml  6.25μg/ml  +  3ml  ddH2O   1.25   1   x2   2ml  2.5μg/ml  +  2ml  ddH2O     Standards   and   Unknown   samples   were   prepared   in   duplicate   as   follows:   0.8ml   of   each   standard   was   pipetted   into   separate1.5ml   eppendorf   tubes   and   0.2ml   dye   reagent   was   then   added,   mixed   gently   and   incubated   for   5mins  at  room  temperature.  Unknown  samples  were  diluted  appropriately   into   a   final   volume   of   0.8ml   ddH2O.   0.2ml   dye   reagent   was   then   added,   mixed   gently   and   incubated   for   5mins.   All   standard   and   unknown   samples   were   transferred   into   plastic   disposable   cuvettes   before   reading   in   spectrophotometer  at  595nm.     3.8.2 Analysis  of  Proteins  by  SDS-­‐  PAGE     3.8.2.1 Preparation   o f   S DS-­‐PAGE   G els   Bio-­‐Rad  electrophoresis  tank  and  gel  apparatus  were  used  for  the  analyses   of  expressed  tandem  proteins  as  per  manufacture’s  protocol.  A  1.0  mm  thick   discontinuous   10%   gel   was   made   by   mixing   3.3ml   of   30%   Acrylamide   mix   (0.8%   (w/v)   bis-­‐acrylamide   stock   solution   (37.5:1   ratio)),   3.6ml   sterile   MilliQ   water,   2.5ml   resolving   buffer   (1.5   M   Tris-­‐HCl.   O.4%   SDS   (w/v),   pH   8.8),   5μl   of   tetramethylethylenediamine   (TEMED),   100μl   of   10%   (w/v)   ammonium   persulfate   (APS).   The   mixture   was   poured   into   the   gel   plate   to   fill  about  4/5th  of  the  plate’s  height  while  200μl  isopropanol  was  poured  to   the  top  of  the  gel  to  create  a  smooth  level  and  bubble-­‐free  meniscus.  The  gel   was   allowed   to   set   at   room   temperature   for   approximately   30   minutes.   Afterwards,   the   isopropanol   was   poured   off   and   the   excess   isopropanol   was   62   removed  with  a  filter  paper.  A  5%  stacking  gel  was  prepared  for  the  second   layer   by   combining   830μl   of   30%   Acrylamide   (0.8%   (w/v)   bis-­‐acrylamide   stock  solution  (37.5:1  ratio)),  3.45ml  sterile  MilliQ  water,  630μl  of  stacking   buffer   (0.5   M   Tris-­‐HCl,   0.4%   SDS   (w/v),   pH   6.8),   5μl   of   TEMED,   50μl   of   10%   APS  and  the  mixture  was  poured  onto  the  top  of  the  resolving  gel  up  to  the   top  of  the  gel  plate,  and  a  1.0  mm  well  comb  was  immediately  inserted  into   the   top   of   the   stacking   gel,   avoiding   air   bubble   formation.   The   gel   was   allowed  to  set  at  room  temperature  for  15  minutes.   3.8.2.2 Preparation   o f   S amples   f or   S DS-­‐PAGE   Following   quantification   of   the   proteins,   0.5ml   eppendorf   tubes   were   labeled  and  the  final  volume  of  each  sample  was  added  to  equal  volume  of   2x   Laemmli   sample   buffer.   Sometimes,   dithiothreitol   (DDT),   a   reducing   agent   was   added   to   the   mixture   to   a   final   concentration   of   25mM   (to   minimize   dimerization   of   samples).   The   eppendorf   tubes   were   centrifuged   at   11,337g   for   4   seconds   and   incubated   at   95°C   for   5   minutes   using   the   Techne   Dri-­‐Block   BD-­‐2D.   Samples   were   carefully   pipetted   into   wells   of   the   10%  gel  in  1x  running  buffer  (0.25  M  Tris  HCl,  1.92  M  Glycine  and  1%  SDS   (w/v),   pH   8.3).   Samples   were   run   against   a   standard   molecular   weight   marker.  The  gel  was  initially  run  for  30  minutes  using  PS250-­‐2  power  pack   at  75v  until  samples  had  passed  through  the  stacking  gel  and  then  turned  up   to   100v   for   1   hour.     The   separated   proteins   were   visualized   by   coomassie   staining   (0.25%   Bromophenol   blue   R-­‐25   (w/v),   50%   methanol   (v/v),   10%   acetic   acid   (v/v)),   or   transferred   onto   a   PVDF   membrane   for   further   protein   analyses  by  western  blotting.     3.8.2.3 Visualized   P rotein   G els   w ith   C oomassie   B lue   Coomassie   blue   is   a   rapid   and   sensitive   technique   for   the   visualization   of   microgram   quantities   of   protein   using   the   principle   of   protein-­‐dye   binding   between   dye   sulfonic   acid   groups   and   positive   protein   amine   groups   through   ionic   interaction.   SDS-­‐PAGE   gel   was   incubated   in   the   coomassie   blue   staining   solution   at   room   temperature   with   gentle   shaking   for   30   minutes   on   an   orbital   shaker.   The   stain   was   decanted   and   rinsed   with   63   deionized   water.   The   gel   was   then   destained   (Destain   solution:   10%   methanol  (v/v)  and  5%  acetic  acid  (v/v)).  To  aid  the  destaining  process  the   solution   was   heated   in   a   microwave   oven   for   1   minute   at   full   power   and   gently   mixed   on   an   orbital   shaker   at   room   temperature   until   the   desired   background  was  achieved.     3.8.3 Western  Blotting   3.8.3.1 Transfer   o f   P roteins   t o   P VDF   M embrane   The   transfer   of   proteins   separated   by   SDS-­‐PAGE   onto   polyvinylidene   diflouride   (PVDF)   membrane   allows   specific   protein   analyses   by   western   blotting.   The   separated   proteins   were   routinely   transferred   from   the   gel   onto  PVDF  membrane  using  a  Mini-­‐Protean  3  blotting  apparatus  (Bio-­‐Rad).   As   PVDF   membrane   is   hydrophobic,   30   seconds   treatment   with   methanol   was   required   in   order   to   wet   the   membrane   before   equilibrating   in   the   transfer   buffer   (Transfer   buffer:   2g   glycine   and   5.8g   Tris   base   dissolved   in   1000  ml  of  MilliQ  water).  After  that,  2  gauze  pads,  2  filter  papers,  the  PVDF   membrane  and  the  gel  were  assembled  in  the  blotting  apparatus  as  shown   in  Figure  3.1.           64   .     Figure  3-­‐1:  Design  of  transferring  assembly   The   scheme   displays   the   preparation   of   the   variant   coats   in   the   transfer   buffer.   In   the   presence   of   high   pH   buffer,   the   protein   transfer   from   anode   towards  cathode,   thus  the  variant  coats  were  prearranged  to  permit  moving   of  proteins  into  the  PVDF  membrane.    Gauze  layer  (a),  filter  papers  (b),  the   gel  comprising  the  separated  proteins  (c)  and  the  PVDF  membrane  (d).     Using   the   PS250-­‐2   power   pack,   the   system   was   run   at   100v   for   1hr.   Following   completion   of   transfer   the   PVDF   membrane   was   blocked   overnight   at   4°C   in   100   ml   of   5%   milk   protein   prepared   in   PBS-­‐Tween-­‐20   (0.05%).   3.8.3.2 Western   B lotting   D etection   o f   G CSF     After   transfer   and   blocking,   the   PVDF   membrane   was   briefly   washed   with   50   ml   PBS/Tween-­‐20   (0.05%).   The   membrane   was   then   incubated   with   primary   rabbit   anti-­‐human   GCSF   antibody   at   a   dilution   of   1:5000   (2μl   antibody  in  10  ml  of  5%  blocking  buffer)  at  room  temperature  on  a  Stuart   Mini   Orbital   Shaker   at   115rpm   for   1.5   hrs.   The   membrane   was   then   washed   with   50   ml   of   PBS/T   for   15   minutes.   The   membrane   was   then   incubated   with   secondary   antibody   (Goat   anti-­‐Rabbit   antibody   linked   to   horseradish   peroxidase   (HRP)   at   a   dilution   of   1:10000   (1μl   antibody   into   10   ml   of   5%   blocking  buffer))  on  a  Mini  Orbital  Shaker  at  115rpm  for  35  minutes.  After   incubation,   the   membrane   was   rinsed   three   times   in   50   ml   of   PBS/T   on   a   Mini   Orbital   Shaker   at   65   rpm   for   15   minutes   per   wash.   Visualization   of   GCSF   constructs   was   carried   out   using   BM   chemiluminescence   blotting   65   substrate   solution   according   to   the   manufacturer’s   instructions.   Under   red   safety   light   in   a   dark   room,   a   sensitive   Fuji   film   was   placed   over   the   membrane   inside   a   hyper   cassette   and   exposed   for   approximately   10   seconds.   The   film   was   then   transferred   into   developing   solution   (Kodak)   until   bands   were   detected.   The   film   was   placed   into   water   to   remove   excess   developing   solution   and   then   placed   into   fixing   solution   (Kodak)   for   approximately   1   minute.   Different   lengths   of   exposure   were   used   to   optimise   the   quality   of   the   western   blot.   Molecular   weights   of   the   GCSF   constructs   were   determined   by   comparison   to   loaded   protein   standards   of   known  molecular  weights.     3.8.4 Enzyme  Linked  Immunosorbent  Assay  (Elisa)   Elisa  is  a  sensitive  and  specific  method  for  quantification  of  proteins.  It  was   used  in  this  project  to  measure  GCSF  tandem  protein  in  crude  and  purified   media,  and  also  in  rat  serum  (in  vivo  study).  The  method  involves  the  use  of   a   specific   monoclonal   antibody   (mAb),   used   to   coat   a   microtiter   plate   (Capture  antibody).  The  Ab  on  the  plate  will  capture  the  protein  of  interest   following   the   addition   of   the   sample.   The   addition   of   secondary   mAb   (Detection   antibody)   labelled   with   biotin,   will   also   bind   to   the   protein   of   interest.   The   biotin   labelled   antibody   allows   binding   of   a   streptavidin-­‐ conjugated   enzyme.   Washing   was   used   between   steps   to   remove   any   unbound  proteins.  A  substrate  is  added  to  the  reaction  to  produce  a  colour   reaction   that   is   directly   proportional   to   the   amount   of   bound   protein.   The   concentration   of   bound   protein   is   measured   by   comparison   with   a   standard   curve  of  known  protein  concentration.  A  standard  curve  ranging  from  0nM   to   5nM   was   generated   from   the   rhGCSF   working   concentration   (shaded   in   gray   in   Table   3.8).   Similar   standard   curves   were   generated   for   the   GCSF   tandems  (test  proteins  and  controls)  from  purified  proteins  by  diluting  each   individual  stock  concentrations  to  a  starting  of  5nM.       66   Table  3-­‐8:  Preparation  of  GCSF  standards   Sample  (μl)   LKC  buffer  (μl)   [GCSF]  nM   Dilution   Stock  @  16000nM   -­‐   16000   -­‐   5  of  16000nM   495   160   100x   25  of  160nM   400   10   16x   500  of  10nM   500   5   2x   500  of  5nM   500   2.5   2x   400  of  2.5nM   600   1   2.5x   500  of  1nM   500   0.5   2x   500  of  0.5nM   500   0.25   2x   500  of  0.25nM   500   0.125   2x   500  of  0.125nM   500   0.0625   2x   500  of  0.0625nM   500   0.03125   2x   500  of  0.03125M   500   0.0156   2x   500  of  0.156M   500   0.0078   2x     Details  of  the  initial  stock  concentrations  for  all  GCSF  (tandems  and  native)   used  for  the  standard  curve  are  tabulated  below  (Table  3.9).     Table  3-­‐9:  Initial  stock  concentrations  of  rhGCSF  and  GCSF  tandem   proteins   Construct   Commercial  rhGCSF   Purified  rhGCSF   GCSF2NAT   GCSF2QAT   GCSF4NAT   GCSF4QAT   GCSF8NAT   GCSF8QAT   Concentration  stock  (nM)   500   16000   10,800   1,0000   15,000   5,000   15,000   15,100     67   For   the   ELISA   assay:   96-­‐well   microtiter   plates   were   coated   with   100μl   of   capture   antibody   (BVD13-­‐3A5)   at   1μg/ml   in   coating   buffer   (0.1M   NaHCO3,   pH  9.2)  and  incubated  at  4°C  overnight.     The   next   day   plates   were   washed   3x   with   230μl   PBS-­‐Tween-­‐20   (0.05%),   patted  dry  and  blocked  for  1hr  at  room  temperature  with  200μl  3%  (w/v)   BSA  in  PBS-­‐Tween-­‐20.     Control   and   unknown   protein   samples   were   prepared   in   LKC   buffer   as   shown  in  Table  3.6.   (LKC:  50ml  of  0.5M  Tris,  15ml  of  0.5M  NaCl,  50μl  0.1%   Tween-­‐20,  0.25g  bovine  gamma  globulin,  0.25g  NaN3  &  2.5g  BSA  and  made   up  to  with  sterile  water  and  stored  at  4°C).     Plates   were   washed   as   previously   described   followed   by   the   addition   of   100μl  of  either  control  or  unknown  protein  samples  to  specified  wells  and   incubated  for  2  hrs  at  room  temperature  with  mixing.     Plates   were   washed   again   as   previously   described   and   100μl   of   detection   antibody   (BVD11-­‐37G10)   at   2μg/ml   in   LKC   buffer   was   then   added   to   all   wells  followed  by  incubation  for  2  hrs  at  room  temperature  with  mixing.   Plates   were   again   washed   as   described   followed   by   the   addition   of   100μl   streptavidin-­‐HRP   at   1μg/ml   made   up   in   0.5%   BSA/PBS-­‐T   to   all   wells   and   incubated   at   room   temperature   for   30   minutes   with   mixing.   Plates   were   finally   washed   6x   with   230μl   PBS-­‐T   before   the   addition   of   100μl   TMB   (3,3’5,5’-­‐tetramethylbenzide   liquid   substrate)   to   all   wells.   Once   a   good   colour   change   was   observed,   100μl   of   5%   sulphuric   acid   (H2SO4)   was   added   to   all   wells   to   stop   the   reaction.   The   plates   were   read   at   450nm   with   background  plate  correction  at  630nm  using  a  Biotech  FLX800  plate  reader   and   Gen5   software.   Results   were   analysed   using   Microsoft   Excel   and   GraphPad  Prism  6  used  for  curve  fit  analysis.       68   3.8.5 AML-­‐193  Proliferation  Assay   The  biological  activities  of  rhGCSF  and  GCSF  tandems  were  evaluated  using   an  AML-­‐193  cell-­‐based  proliferation  assay  (Human  acute  myeloid  leukemic   cell   line).   GCSF   stimulates   the   proliferation   of   the   AML-­‐193   cell   line.   The   main   objective   of   this   method   was   to   show   that   GCSF   tandems   could   stimulate   the   proliferation   of   AML-­‐193   cells   in   comparison   to   rhGCSF.     Details   of   this   assay   method   and   associated   cell   culture   development   are   provided  below.   3.8.5.1 Growth   o f   t he   A ML-­‐193   C ell   L ine   Cells   (ATCC,  Batch  No.  3475266)  were  removed  from  liquid  nitrogen  storage   and   defrosted   by   placing   into   a   37°C   water   bath   for   2   minutes.   The   contents   of  the  vial  were  then  transferred  to  a  T-­‐25  flask  containing  4  ml  of  culture   medium   (5%   FBS,   4mM   L-­‐glutamine,   100   U/ml   penicillin,   100   µg/ml   streptomycin,   5  µg/ml   transferrin,   5  µg/ml   insulin   and   5   ng/ml   GM-­‐CSF   in   Iscove’s   modified   Dulbecco’s   medium).   The   AML-­‐193   cells   were   routinely   cultured  to  a  density  of  2  x106  cells/ml  but  not  exceeding  2.5  x  106  cells/ml       (5%   CO2,   37°C).   Passages   were   performed   2   times   a   week   and   cell   density   and  viability  was  assessed  by  trypan  blue  exclusion  as  previously  described   at  section  3.5.2.     3.8.5.2 AML-­‐193   B ioassay   After  a  minimum  of  two  passages,  the  AML-­‐193  cells  were  ready  for  use  in   the   bioassay.   The   cells   were   prepared   for   the   assay   by   washing   3   times   with   10ml  PBS  and  the  washed  cells  were  recovered  by  centrifuging  at  168  x  g  for   5  minutes.  The  final  cell  pellet  was  then  reconstituted  in  the  assay  medium   (5%   FBS,   4mM   L-­‐glutamine,   100  U/ml   penicillin,   100  µg/ml   streptomycin,   5  µg/ml   insulin,   5  µg/ml   transferrin   in   Iscove’s   modified   Dulbecco’s   medium)  and  cell  density  adjusted  to  0.5  x  106  cells/ml.     A   commercial   GCSF   was   used   to   generate   a   standard   curve   for   the   assay.   0.2mg/ml   rhGCSF  (Biolegend)  was  reconstituted  in  PBS  and  1%  (w/v)  BSA   to   a   concentration   of   10   µg/ml   (500nM   stock),   divided   into   10  µl   aliquots   69   and   stored   at   -­‐80°C.   On   each   day   of   assay   1   vial   was   removed   from   the   frozen  stock  and  a  standard  curve  ranging  from  0nM  to  5nM  was  generated   as   shown   in   Table   3.10.   Similar   standard   curves   were   generated   for   the   GCSF  tandems  (test  proteins  and  controls)  from  purified  proteins  by  diluting   each  individual  stock  concentrations  to  a  starting  concentration  of  10nM.   Table  3-­‐10:  Preparation  of  GCSF  standard  curve   Sample  (μl)   Assay  media  (μl)   [GCSF]  nM   Dilution   5ul  of  Stock  @  500nM   495   5   100x   220  of  5nM   220   2.5   2x   200  of  2.5nM   300   1   2.5x   250  of  1nM   250   0.5   2x   250  of  0.5nM   250   0.25   2x   250  of  0.25nM   250   0.125   2x   250  of  125nM   250   0.06   2x   250  of  0.06nM   250   0.03   2x   250  of  0.03nM   250   0.015   2x   250  of  0.015nM   250   0.008   2x   150  of  0.008nM   300   0.003   3x   150  of  0.003nM   300   0.0015   3x   150  of  0.015nM   300   0.0008   3x     50µl  of  each  test  protein  was  added  to  the  wells  of  a  96-­‐well  microplate  in   triplicate.   50µl   of   AML-­‐193   cells   at   0.5   x   106   cells/ml   were   then   added   to   each   well   with   gentle   agitation   to   mix   the   contents   (i.e.   cells   suspension,   standard  and  samples).    Control  wells,  which  contained  only  assay  medium   and   cells   suspension   (50µl   +   50µl),   and   blank   wells,   which   contained   only   assay  medium  (100µl)  were  also  set  up.     For   3   days,   AML-­‐193   cells   were   exposed   to   the   different   concentrations   of   test   proteins   in   a   CO2   incubator   (5%   CO2,   37°C)   and   then   20µl   of   MTS   70   (Celltiter  96  Aqueous  One  Solution  from  Promega)  was  added  to  each  well.     Readings   were   taken   every   40   minutes   at   490nm   using   a   Biotech   FLX800   plate  reader  for  a  total  of  2  hrs.  The  results  were  displayed  using  Microsoft   Excel   and   analysed   with   Gen5   software.   GraphPad   Prism   6   was   used   for   curve  fit  analysis.   3.8.6 Short  Term  Stability  of  GCSF  Tandem  Molecules     The  protein  stability  was  assessed  by  testing  purified  samples  at  3  different   temperatures  (4°C,  RmT  and  -­‐80°C  freeze  thaw  (F/T)  cycles)  over  an  8  day   period.    Purified  tandem  proteins  were  taken  from  the  -­‐80°C  and  diluted  to   0.8   mg/ml   with   filter   sterile   PBS   and   kept   on   ice.   All   manipulations   were   carried  out  under  sterile  conditions.  Aliquots  of  protein  at  0.8  mg/ml  were   placed  at  RmT,  4°C  or  -­‐80°C  in  sterile  1.5  ml  eppendorf  tubes.    Samples  were   taken   on   days   0,   1,   4   and   8   (samples   taken   on   day   zero,   represent   untreated   sample   controls)   and   immediately   diluted   with   an   equal   volume   of   SDS-­‐ PAGE   buffer   to   0.4   mg/ml   (Laemmlli   buffer)   and   heated   at   95°C   for   5   minutes   to   denature.   Samples   were   analysed   by   10%   SDS-­‐PAGE   and   protein   bands  visualised  by  coomassie  staining  (A  total  of  7.5µg  protein  was  loaded   per   lane)   and   western   blotting   (a   total   of   100ng   protein   was   loaded   per   lane).   3.9 Experimental  Procedure  for  In  vivo  Study   The   aim   of   this   protocol   was   to   determine   the   pharmacokinetic   (PK)   and   pharmacodynamics   (PD)   properties   of   rhGCSF   and   GCSF   tandems   (GCSF2NAT,   4NAT,   8NAT   &   8QAT)   in   Sprague   Dawley   rats   following   intravenous  injection  and  look  at  the  effects  of  these  constructs  on  the  WBCs   and  neutrophils  population.   Pre-­‐Dose   (-­‐24   hr)   samples   of   300-­‐400µl   of   blood   sample   were   taken   from   each   rat   before   injection   with   test   protein.   These   served   as   control   blood   samples.   Next   day,   groups   of   six   male   rats   were   administered   a   single   intravenous   dose   of   rhGCSF,   GCSF   tandem,   or   vehicle   (PBS   only)   at   250µg   per  /kg  protein.  At  the  following  time  points  of  0.5,  1,  2,  4,  8,  12,  24,  48  &  72   71   hrs   post=injection,   ~300-­‐400µl   of   blood   samples   were   collected   from   each   rat   from   the   tail   vein   under   anaesthesia   using   isoflurane.   Blood   samples   were   centrifuged   (for   serum   preparation),   labelled   and   stored   at   -­‐80°C.   Counts  of  blood  cells  (CBCs)  were  performed  on  selected  samples  (-­‐24,  12,   24,  48  and  72  hrs)  using  an  automated  coulter  counter.  Blood  smears  were   fixed   and   stained   by   a   routine   laboratory   method   (H&E   or   Giemsa).   Thereafter,   all   samples   for   analysis   were   transferred   to   the   University   of   Sheffield   for   data   confirmation   and   further   analyses.   Elisa   was   used   to   measure  the  concentration  of  proteins  in  each  serum  sample  at  0.5,  1,  2,  4,  8,   12,  24,  48,  72  hrs  post-­‐dose.     3.9.1 Ethics   All  animal  experiments  were  approved  by  the  local  ethical  committee  of  the   University  of  Isfahan.   3.10  Statistical  Analysis     In  vitro  analyses  of  all  purified  GCSF  tandems  were  performed  on  GraphPad   Prism  6  using  a  Mann-­‐Whitney  test  (a  non-­‐parametric  test  that  is  often  used   to  compare  two  groups  that  come  from  the  same  population).  The  statistical   comparison  of  the  tandem  proteins  in  rat  plasma  for  the  pharmacokinetics   study   data   were   analysed   using   the   non-­‐compartmental   method   of   data   analysis.   This   involved   the   use   of   Winnonlin   6.3   PK   program   developed   by   Phoenix   Certara.   Significance   between   terminal   half-­‐life   data   of   GCSF   tandems   was   performed   with   GraphPad   Prism   6   using   one-­‐way   ANOVA   (used   to   determine   any   significant   differences   between   three   or   more   groups   of   sample   data).   The   pharmacodynamics   studies   of   the   GCSF   tandems  at  selected  sampling  time  points  (-­‐24,  12,  24,  48  and  72  hrs)   were   analysed  with  GraphPad  Prism  6  using  multiple  T-­‐test.       72   4. Results  1:  Cloning  and  Expression  of  GCSF  Tandems     4.1 Summary   Previous  studies  carried  out  by  the  Ross  Group  (UoS)  have  shown  that  the   use   of   glycosylated   linkers   between   two   GH   ligands   to   create   protein   tandems   results   in   their   glycosylation   and   an   increased   molecular   weight   whilst  maintaining  biological  activity.  This  technology  can  be  easily  applied   to  other  molecules  such  as  GCSF.  This  chapter  describes  cloning,  sequencing   and   expression   of   different   GCSF   constructs   to   produce   GCSF   tandems   linked  via  glycosylated  and  non-­‐glycosylated  flexible  linkers  (Gly4Ser)n.  The   initial  data  obtained  from  this  chapter  has  shown  that  it  is  possible  to  clone   and   express   two   GCSF   molecules   linked   by   a   flexible   linker   (Gly4Ser)n   in   mammalian   cell   lines.   Expressed   proteins   were   analysed   by   Elisa,   and   western   blotting   used   to   confirm   the   glycosylated   proteins   had   a   greater   molecular  weight  than  the  non-­‐glycosylated  proteins.                         73   4.2 Introduction   In   medicine,   any   protein   used   for   therapeutic   purposes   is   not   only   a   sequence   of   amino   acids   determined   by   a   particular   gene,   but   still   require   editing,   altering   of   amino   acids   or   addition   of   sugars.   These   modifications   following   the   preliminary   translation   of   the   protein   are   called   post-­‐ translational   processes   (Li   and   d'Anjou,   2009).     Glycosylation   refers   to   the   post-­‐translational   process   that   covalently   linking   oligosaccharide   to   polypeptides.  It  is  one  of  the  most  common  protein  modifications  and  more   than   50%   of   proteins   are   glycosylated   in   the   body,   which   are   mainly   secreted   or   part   of   cell   membrane   components   (Sola   et   al.,   2007).   Normally,   the   attachment   reaction   of   glycans   to   a   protein   begins   in   the   endoplasmic   reticulum   (ER)   and   is   completed   in   the   Golgi   apparatus.   The   two   major   forms   are   O-­‐linked   glycosylation   and   N-­‐linked   glycosylation   (Saint-­‐Jore-­‐ Dupas  et  al.,  2007).       N-­‐linked  glycosylation  is  attached  to  the  amide  nitrogen  of  asparagine  (Asn)   residues   within   the   common   consensus   sequence   Asn-­‐X-­‐Ser/Thr   where   X   can  be  any  amino  acid  except  proline  (Pro)  (Kornfeld  and  Kornfeld,  1985).     No   particular   consensus   sequences   are   recognized   for   O-­‐linked   glycosylation.   Thus,   N-­‐linked   glycosylation   is   preferentially   used   in   many   technologies  of  protein  modification  (Spiro,  2002).   Asterion   have   designed   different   strategies   of   protein   fusion   technologies   with  a  view  to  generating  longer  acting  therapies.  One  such  strategy  is  based   upon   tandem   proteins   linked   a   via   flexible   linker   (Gly4Ser)n   that   has   been   designed   to   contain   N-­‐linked   glycosylation   motifs   (glycosylation   consensus   sequences).   It   has   been   hypothesized   that   inserted   glycosylation   motifs   within   the   linker   region   rather   than   the   ligand   would   be   recognized   by   mammalian   cells   for   glycosylation.   This   would   result   in   an   increased   molecular   weight   without   interfering   with   the   biological   activity   of   the   ligand.       74   The   Linker   region   was   designed   to   incorporate   2,   4   and   8   glycosylation   motifs   (Asn-­‐Ala-­‐Thr   (NAT))   and   their   respective   controls   in   which   Glutamine  (or  Q)  replaces  N  in  NAT  sequence  motif  to  produce  QAT  which   would  not  be  recognized  by  mammalian  cells  for  glycosylation.  Up  to  eight   glycosylation   sites   were   introduced   to   assess   whether   more   efficient   glycosylation   could   increase   the   molecular   weight   and   subsequently   delay   renal  clearance.     The   table   below   summarizes   all   GCSF   tandem   structures   and   their   respective   controls   (Table   4.1).   A   full   description   of   the   amino   acid   sequences  of  these  tandems  can  be  seen  in  Appendix  A.2.       75   Table   4-­‐1:   The   structure   of   GCSF   tandems   with   modified   flexible   linkers  (Gly4Ser)n   The   glycosylation   consensus   sequences   are   highlighted   in   red   (NAT:   Asn-­‐ Ala-­‐Thr).   The   control   non-­‐glycosylation   consensus   sequences   are   highlighted   in   blue   (QAT:   Gln-­‐Ala-­‐Thr).   The   linkers   are   built   around   a   flexible   glycine-­‐serine   (Gly4Ser)n   sequence   (these   linkers   were   gene   synthesised   by   Eurofins   MWG).   Each   tandem   contains   a   C-­‐terminal   6   x   Histidine  tag  (Used  to  aid  purification  using  IMAC).     Molecule  name   pSecTagGCSF2NAT_Hist   Number  of   Size  of   NAT/QAT  motifs   linker   2  x  NAT   147bp   Summary  of  sequence   GCSF-­‐G4S-­‐  G2NAT-­‐  G4S-­‐G4S-­‐G4S-­‐G4S-­‐G2NAT-­‐ G4S-­‐GS-­‐GCSFx6H   pSecTagGCSF2QAT_Hist   2  x  QAT   147bp   GCSF-­‐G4S-­‐  G2QAT-­‐  G4S-­‐G4S-­‐G4S-­‐G4S-­‐G2QAT-­‐ G4S-­‐GS-­‐GCSFx6H   pSecTagGCSF4NAT_Hist   4  x  NAT   147bp   GCSF-­‐G4S-­‐  G2NAT-­‐  G4S-­‐G2-­‐NAT-­‐G4S-­‐G2-­‐  NAT-­‐ G4S-­‐G2NAT-­‐G4S-­‐GS-­‐GCSFx6H   pSecTagGCSF4QAT_Hist   4  x  QAT   147bp   GCSF-­‐G4S-­‐  G2QAT-­‐  G4S-­‐G2-­‐QAT-­‐G4S-­‐G2-­‐QAT-­‐ G4S-­‐G2QAT-­‐G4S-­‐GS-­‐GCSFx6H   pSecTagGCSF8NAT_Hist   8  x  NAT   282bp   GCSF-­‐G4S-­‐  G2-­‐NAT-­‐  G4S-­‐G2-­‐NAT-­‐G4S-­‐G2-­‐   NAT-­‐G4S-­‐G2NAT-­‐  G4S-­‐  G2-­‐NAT-­‐  G4S-­‐G2-­‐NAT-­‐ G4S-­‐G2-­‐  NAT-­‐G4S-­‐G2NAT-­‐G4S-­‐GS-­‐GCSFx6H   pSecTagGCSFT8QAT_Hist   8  x  QAT   282bp   GCSF-­‐G4S-­‐  G2QAT-­‐  G4S-­‐G2-­‐QAT-­‐G4S-­‐G2-­‐QAT-­‐ G4S-­‐G2-­‐QAT-­‐  G4S-­‐  G2-­‐QAT-­‐  G4S-­‐G2-­‐QAT-­‐G4S-­‐ G2-­‐QAT-­‐G4S-­‐G2-­‐QAT-­‐G4S-­‐GS-­‐GCSFx6H                 76   4.2.1 Aim  and  Hypotheses   As   mentioned   previously,   proteins   have   been   directly   modified   by   increasing   the   glycosylation   on   the   proteins   (Elliott   et   al.,   2003),   and   this   could   potentially   affect   protein   bioactivity   and   increase   immunogenicity.     Previously   a   tandem   of   two   GH   molecules   joined   by   a   flexible   (Gly4Ser)n   peptide  linker  containing  variable  numbers  of  glycosylation  motifs  from  2  to   8  was  created.    This  method  avoided   altering  or  modifying  the  protein  itself   directly.     These   molecules   were   shown   to   be   glycosylated   with   increased   MW   and   were   biologically   active.   In   this   chapter,   the   aim   is   to   replace   two   GH   ligands   in   a   tandem   with   two   GCSF   ligands.   The   initial   stage   will   be   to   construct   a   tandem   molecule   containing   two   GCSF   ligands   using   PCR   and   then   ligate   in   relevant   linker   regions.     As   GCSF   is   known   to   bind   to   its   receptor   at   2:2   stoichiometry,   we   hypothesized   that   the   resulting   tandem   GCSF   molecule   will   be   available   to   bind   to   the   two   GCSF-­‐receptors   and   induce  downstream  signal  transduction.   4.2.2 Objectives   1-­‐ Construct   a   tandem   molecule   containing   two   ligands   of   GCSF   using   PCR.   2-­‐ Insert  variable  linkers  between  a  tandem  GCSF  molecule  that  contain   increasing   numbers   of   glycosylation   motifs   and   control   non-­‐ glycosylation  motifs.   3-­‐ Express  and  analyse  GCSF  tandems  from  a  mammalian  cell  line.       77   4.3 Construction  of  GCSF  Tandems     To   initially   construct   the   GSCF   tandem   a   template   plasmid   was   used.   This   plasmid   contained   a   tandem   GH   molecule   linked   via   a   glycosylated   linker   containing   2   x   NAT   motifs   and   was   available   in   the   Ross   group   laboratory.   The   method   employed   will   be   to   replace   each   GH   molecule   with   a   GCSF   molecule  and  thus  produce  a  GCSF  tandem  containing  2  x  NAT  motifs  with  a   GCSF  secretion  signal.  Once  constructed,  it  will  be  a  simple  matter  to  replace   the   linker   region   in   the   GCSF   tandem   using   the   restriction   enzymes   XhoI/BamH1   with   other   suitable   linkers   containing   either   NAT   or   control   QAT   motifs   (Figure   4.1).   A   full   protein   sequence   of   GH,   GCSF   and   linker   region  constructs  in  Appendix  B.                                             78     Figure   4-­‐1:   The   diagram   summarizes   the   process   of   producing   GCSF   tandems  containing  variable  linkers    [A]   Both   GCSF-­‐L1   and   L2   (highlighted   in   green)   containing   restriction   enzyme  sites  (highlighted  in  yellow)  were  PCRed  from  pGCSFSecTag4A1.  [B]   Both   GH   molecules   (highlighted   in   green)   are   removed   from   pSecTagGH2NAT   plasmid   and   replaced   with   the   GCSF   (GCSF   L1   &   L2).   [C]   The   original   (Gly4Ser)n   linker   (highlighted   in   orange)   is   replaced   with   variable   glycosylated   and   non-­‐glycosylated   linkers.   The   plasmid   contains   a   CMV   promoter   region   (highlighted   in   white),   GCSF   signal   sequence   (highlighted   in   blue),   6x   Hist   tag   (highlighted   in   pink)   and   a   stop   codon   (highlighted   in   pink).   A   full   diagram   of   the   plasmid   used   in   this   study   is   provided  in  Appendix  C.         79   4.3.1 Construction  of  pSecTag  GCSF-­‐L1_Hist   The   first   full-­‐length   nucleotide   sequence   of   GCSF   with   signal   sequence   (GCSF-­‐L1)   was   PCRed   from   the   template   plasmid   pSecTagGCSF4A1   (See   Table   4.2   and   Figure   4.1.A   for   description)   using   a   forward   primer   (GCSF   Nhe1)   and   reverse   primer   (GCSF   XhoI   Rev)   (See   full   nucleotide   sequences   for   both   primers   in   Appendix   A.1)   and   digested   with   restriction   enzymes   NheI/XhoI.   This   produced   a   full-­‐length   GCSF   molecule,   GCSF-­‐L1   (612bp)   containing  the  signal  sequence  of  GCSF  with  a  5  prime  NheI  site  along  with  a   3  prime  XhoI  restriction  site  (Figure  4.2).       Table  4-­‐2:    Description  of  the  two  plasmids  that  were  used  to  produce   psecTagGCSF2NAT_Hist   Construct  name   N-­‐terminal   Linker   domain   C-­‐terminal   Expression   domain   vector   GCSF4A1   GCSF   (Gly4Ser)x6   GCSF-­‐R   pSecTag   GH2NAT_Hist   GH   (Gly4Ser)n  with  2   GH   pSecTag   glycosylated  motifs   The   pSecTagGH2NAT_Hist   was   digested   with   double   restriction   enzymes   Nhe1/Xho1   to   remove   GH   Ligand   1   (GH-­‐L1).   Single   enzyme   digests   were   used  to  confirm  enzyme  activity  (Figure  4.3).   Thereafter   GCSF-­‐L1   was   ligated   into   pSecTagGH2NAT_Hist   to   produce   the   plasmid   pSecTagGCSF2NAT_Hist-­‐L1   (This  contains  GCSF-­‐L1  and   GH-­‐L2   with   a   linker   region   containing   2   x   NAT   glycosylation   sites).   The   ligation   reaction   was  performed  as  described  at  section  3.4.4.     To   confirm   the   construction   of   pSecTagGCSF2NAT_Hist-­‐L1,   plasmid   mini   preparations   were   digested   using   Nhe1/Xho1   to   generate   a   GCSF-­‐L1   0.6kb   insert  (Figure  4.4).       80   1                    M                2 Lane  1:  PCR  GCSF-­‐L1  (~0.6kb).     Lane   M:   1kb   DNA   Ladder:   Molecular   weight   marker:   Bands   starting   at   the   bottom   (0.5,   1.0,  1.5,  2.0,  3.0,  4.0,  5.0,  6.0,  8.0,  10.0  Kb).     Lane  2:  Negative  control  (primers  only).     Figure  4-­‐2:    PCR  of  GCSF-­‐L1   PCR   of   GCSF-­‐L1   (DNA)   using   forward   primer   (GCSF   Nhe1)   and   reverse   primer   (GCSF   XhoI   Rev)   run   on   1%   agarose   gel   alongside   1kb   ladder   as   standard.       The   presence   of   GCSF-­‐L1   at   (~0.6kb)   which   is   the   expected   size   indicates   that  PCR  has  been  successful.         81     1                            2                        3                   M    1                    2                    3                    M       Lane   1:   The   upper   band   represents   the   pSecTagGH2NAT_Hist   (cut  plasmid),  lower  band   (~0.6kb)  represents  GH-­‐L1  insert  as  a  result  of   a   successful   double   digest   by   both   restriction   enzymes. Lane  2:  NheI  digest  of  pSecTagGH2NAT_Hist. Lane  3:  XhoI  digest  of  pSecTagGH2NAT_Hist.   Lane  M:  1kb  Ladder:  Molecular  weight  marker:   Bands   starting   at   the   bottom   (0.5,   1.0,   1.5,   2.0,   3.0,  4.0,  5.0,  6.0,  8.0,  10.0  Kb).   Figure  4-­‐3:  Double  digest  of  pSecTagGH2NAT_Hist   Double   digest   of   pGH2NAT_Hist   using   NheI   and   XhoI   separated   on   1%   agarose  gel  electrophoresis.     The  presence  of  an  insert  at  ~0.6kb  in  lane  1  indicates  a  successful  double   digest.     A   single   band   is   observed   in   both   lanes   2   and   3,   which   shows   that   plasmids   used   in   the   digestion   reaction   were   cut   by   the   individual   restriction   enzyme.     The   double   digested   plasmid   runs   at   a   slightly   lower   molecular  weight  compared  to  the  single  digested  plasmids,  which  is  due  to   the  excision  of  the  GH-­‐L1  from  the  starting  plasmid  pSecTagGH2NAT_Hist.   GCSF-­‐L1  was  then  ligated  into  pSecTagGH2NAT_Hist  to  produce  the  plasmid   pSecTagGCSF2NAT_Hist-­‐L1.  This  contains  GCSF-­‐L1  and  GH-­‐L2  with  a  linker   region  containing  2  x  NAT  glycosylation  sites.  For  potential  plasmid  clones   were  generated  from  the  ligation  plate  colonies  and  were  analysed  by  both   double  digestion  with  NheI/XhoI  (see  Figure  4.4)  and  DNA  sequencing.       82   1                    2                    3                4              M          Lane   1,   2   &   3:   The   upper   band   represents   the  cut  plasmid  and  the  small  fragment  band   represents   the   GCSF-­‐L1   (~0.6kb)   as   a   result   of   a   successful   double   digests   by   both   NheI/XhoI  restriction  enzymes.   Lane  4:  Unsuccessful  digest.   Lane   M:   1kb   Ladder:   Molecular   weight   marker:   Bands   starting   at   the   bottom   (0.5,   1.0,  1.5,  2.0,  3.0,  4.0,  5.0,  6.0,  8.0,  10.0  Kb).       Figure   4-­‐4:   Double   digest   of   pSecTagGCSF2NAT_Hist-­‐L1   potential   clones   Double   digest   of   pSecTagGCSF2NAT_Hist-­‐L1   using   NheI/XhoI   separated   on   1%  agarose  gel  electrophoresis.       The  presence  of  the  insert  (~0.6kb)  in  lanes  1,  2  and  3  indicates  a  positive   ligation.  Unsuccessful  digest  observed  at  lane  4  could  due  to  missing  one  or   both   NheI/XhoI   restriction   enzymes   or   that   this   clone   has   no   insert.   The   three   positive   clones   were   taken   forward   for   sequencing   using   CMVF   and   GHseq2Rev   primers   (see   full   nucleotide   sequences   for   both   primers   in   Appendix   A.1).   Sequencing   confirmed   the   ligation   of   GCSF   L1   to   pSecTagGH2NAT_Hist  to  form  pSecTagGCSF2NAT_Hist-­‐L1.             83   4.3.2 Construction  of  pSecTagGCSF-­‐L2_Hist     The   second   full-­‐length   nucleotide   sequence   of   GCSF-­‐L2   was   PCRed   using   a   forward   primer   (GCSF   BamH1)   and   reverse   primer   (GCSF   Age1   Rev)   (Appendix   A.1)   from   the   template   plasmid   pSecTagGCSF4A1   and   digested   with  BamHI/AgeI.  This  produced  a  full  length  GCSF  molecule  containing  a  5   prime   BamHI   site   along   with   a   3   prime   AgeI   site   (present   in   plasmid   just   prior   to   the   Hist   tag)   (Figure   4.5.).     The   PCR   fragment   was   then   digested   with  BamH1/Age1  and  gel  isolated.     pSecTagGCSF2NAT_Hist-­‐L1   was   then   digested   with   restriction   enzymes   BamHI/AgeI  to  remove  GH-­‐L2  and  the  digested  plasmid  gel  isolated.  Single   enzyme   digests   were   used   to   confirm   enzyme   activity   (Figure   4.6).   Thereafter   GCSF-­‐L2   was   ligated   into   pSecTagGCSF2NAT_Hist-­‐L1   to   form   the   new  plasmid,  pSecTagGCSF2NAT_Hist.  This  contains  a  GCSF  tandem  with  a   linker  region  containing  2  x  NAT  glycosylation  sites  (refer  to  Figure  4.1.B).   To  confirm  the  authenticity  of  the  new  construct,  plasmid  mini  preparations   were   digested   using   BamHI/AgeI   to   visualise   the   GCSF-­‐L2   insert   (Figure   4.7).     During  sequencing,  it  was  difficult  to  read  the  whole  gene  for  tandem  GCSF   (i.e.  GCSFL1-­‐linker-­‐GCSF-­‐L2)  using  CMVF  due  to  the  presence  of  GCSF-­‐L1  in   the   same   plasmid.   Therefore   we   decided   to   follow   the   protocol   below   to   remove   GCSF-­‐L2   using   BamHI/AgeI   (Figure   4.8)   and   ligate   to   the   plasmid   pSecTag_link  to  form  pSecTag_link_GCSF-­‐L2  (Figure  4.9).  Using  this  method   the  GCSF_L2  insert  was  successfully  sequenced  using  CMVF.     84   1                    M                                      2 Lane  1:  PCR  of  GCSF-­‐L2  (~0.6kb).   Lane   M:   1kb   DNA   Ladder:   Molecular   weight   marker:   Bands   starting   at   the   bottom  (0.5,  1.0,  1.5,  2.0,  3.0,  4.0,  5.0,  6.0,   8.0,  10.0  Kb).   Lane  2:  Negative  control  (primers  only).   Figure  4-­‐5:  Generation  of  PCR  fragment  GCSF-­‐L2   Generation   of   PCR   fragment   using   a   forward   primer   (GCSF   BmaH1)   and   reverse  primer  (GCSF  Age1  Rev)  separated  on  a  1%  agarose  gel.     The  result  in  lane  1  shows  that  GCSF-­‐L2  is  at  the  expected  size  of  ~  0.6kb,   indicating  a  successful  PCR.                   85   1                          2                      3                        M   Lane   1:   The   upper   band   represents   the   pSecTagGCSF2NAT_Hist_L1   (cut  plasmid)  lower   band   (~0.6kb)  represents  GH-­‐L2  insert   (~0.6kb)  as   a   result   of  a  successful  double  digest.       Lane2:  BamHI  digest  of  pSecTagGCSF2NAT_Hist_L1.       Lane3:  AgeI  digest  of  pSecTagGCSF2NAT_Hist_L1.   Age1.       Lane  M:  1kb  Ladder.  Molecular  weight   marker:  Bands   starting   at   the   bottom   (0.5,   1.0,   1.5,   2.0,   3.0,   4.0,   5.0,   6.0,  8.0,  10.0Kb).     Figure  4-­‐6:  Double  digestion  of  pSecTagGCSF2NAT_Hist_L1   Double  digestion  of  pSecTagGCSF2NAT_Hist_L1  using  BamHI  and  AgeI  RE’s   separated  on  1%  agarose  gel  electrophoresis.     The   presence   of       a   fragment   of   ~0.6kb   in   lane   1   indicates   a   successful   double   digest   by   both   restriction   enzymes   for   the   second   GH-­‐L2.   A   single   band  is  observed  in  both  2  and  3  lanes,  which  shows  that  the  plasmid  used   in  this  digestion  reaction  was  cut  by  the  individual  restriction  enzyme.  The   double  digested  plasmid  fragment  runs  at  a  slightly  lower  molecular  weight   compared  to  the  single  digested  plasmids,  which  is  due  to  the  excision  of  the   GH-­‐L2  fragment  from  pSecTagGCSF2NAT_Hist_L1.           86   1                    2                  3            4            M Lane  1:  Unsuccessful  digest.     Lane   2,   3   &   4:   The  upper  band  represents  the   pSecTagGCSF2NAT_Hist   cut   plasmid   and   the   below   small   fragment   band   represents   the   GCSF-­‐L2   (~0.6kb)   as   a   results   of   a   successful   double  digests  by  both   BamHI/AgeI  restriction   enzymes.      Lane   M:   1kb   Ladder:   Molecular   weight   marker:  Bands   starting   at  the  bottom  (0.5,  1.0,   1.5,  2.0,  3.0,  4.0,  5.0,  6.0,  8.0,  10.0  Kb).     Figure  4-­‐7:  Double  digest  of  pSecTagGCSF2NAT_Hist  potential  clones   Double   digest   of   pSecTagGCSF2NAT_Hist   using   BamHI/AgeI   separated   on   1%  agarose  gel  electrophoresis.     The   presence   of   an   insert   at   ~0.6kb   in   lanes   2,   3   and   4   indicates a positive ligation.   Unsuccessful   digest   observed   at   lane   1   probably   due   to   missing   one   or  both  BamHI/AgeI  restriction  enzymes  or that this clone has no insert.  The   three  positive  clones  take  forward  for  sequencing.             87     1                          2                      3                        M Lane   1:   The   upper   band   represents   the   pSecTagGCSF2NAT_Hist   cut   plasmid   and   the   small   fragment   represents   the   second   GCSF-­‐L2   (~0.6kb)   as   a   result   of   a   successful   double   digests   by   both   BamHI/AgeI  restriction  enzymes.       Lane2:  BamHI  digest  of  pSecTagGCSF2NAT_Hist.       Lane3:  AgeI  digest  of  pSecTagGCSF2NAT_Hist.       Lane   M:   1kb   Ladder.   Molecular   weight   marker:   Bands   starting   at   the   bottom   (0.5,   1.0,   1.5,   2.0,   3.0,   4.0,  5.0,  6.0,  8.0,  10.0Kb).     Figure  4-­‐8:  Removal  of  GCSF  L2  from  pGCSFsecTagGCSF2NAT_Hist     Double   digest   of   pSecTagGCSF2NAT_Hist   using   BamHI/AgeI   separated   on   1%  agarose  gel  electrophoresis.     The   presence   of   the   below   small   fragment   band   (~0.6kb)   at   lane   1   indicates   a   successful   double   digest   by   both   restriction   enzymes   for   the   second   GCSF-­‐ L2.  A  single  band  is  observed  in  both  2  and  3  lanes,  which  shows  the  plasmid   used  in  the  digestion  reaction  was  cut  by  the  individual  restriction  enzyme.   Also   the   double   digested   fragment   runs   at   a   slightly   lower   molecular   weight   compared  to  the  single  digestion  fragments,  which  is  due  to  excision  of  the   GCSF-­‐L2  fragment  from  the  pSecTagGCSF2NAT_Hist.  The  GCSF  L2  fragment   was  gel  isolated  and  ligated  to  pSecTag_link  (Figure  4.9).     88    1                    2                  3                4                  M           Lane   1,   2,   3   &   4:   The   upper   band   represents   the   pSecTag_link_GCSF-­‐L2   cut   plasmid   and   lower   band   represents   the   GCSF-­‐L2   (~0.6kb)   as   a   result   of   a   successful   double   digests   by   both   BamH1/Age1  restriction  enzymes.       Lane   M:   1kb   Ladder:   Molecular   weight   marker:  Bands  starting  at  the  bottom  (0.5,   1.0,  1.5,  2.0,  3.0,  4.0,  5.0,  6.0,  8.0,  10.0  Kb).         Figure  4-­‐9:  Double  digest  of  pSecTag_link_GCSF-­‐L2  potential  clones   Double  digest  of  pSecTag_link_GCSF-­‐L2  using  BamHI/AgeI  separated  on  1%   agarose  gel  electrophoresis.       The   presence   of   the   small   fragment   at   ~0.6kb   in   lanes   1,   2,   3   and   4   show   successful  double  digests  of  all  4  clones.  Positive  clones  were  confirmed  by   sequencing  using  CMVF.                         89     4.4 Generating  GCSF  Tandems  with  Variable  Linkers   Successful   construction   of   pSecTasgGCSF2NAT_Hist   produced   a   vector   containing   a   flexible   linker   with  two   glycosylation   motifs   between   two   GCSF   ligands  and  read  through  to  the  X6  Hist  purification  tag  present  within  this   plasmid.   As   a   result,   this   tandem   could   be   used   as   a   template   for   creating   multiple   GCSF   constructs   containing   variable   glycosylated   and   non-­‐ glycosylated  linkers   (refer   to   Table  4.1).  A   full   description   of   the   amino   acid   sequences  of  these  tandems  can  be  found  in  Appendix  B.   The   2NAT   linker   fragment   was   digested   from   pSecTagGCSF2NAT_Hist   using   Xho1/   BamH1   restriction   enzymes   and   replaced   with   other   linkers   containing   glycosylation   and   non-­‐glycosylation   motifs   (2QAT,   4NAT,   4QAT,   8NAT   and   8QAT)   (See   Figure   4.3.C).   The   linkers   were   already   available   as   BamHI/XhoI  digested  fragments.   The   ligation   of   each   linker   was   sequenced   using   CMVF   and/or   BGHRev   primers   to   confirm   the   successful   ligation   (A   full   description   of   the   amino   acid   sequences   of   all   linkers   can   be   found   in   Appendix   B.3)   and   double   digestion   with   XhoI   and   BamHI.   pSecTagGCSF4NAT_Hist   &   pSecTagGCSF8QAT_Hist   are   shown   as   examples   of   the   restriction   digests.   Single  enzyme  digests  were  used  to  confirm  enzyme  activity  (Figure  4.10  &   4.11).                 90    1              2              3                4                M       Lane   1:   Double   Xho1/BamH1   digests   of   pSecTagGCSF4NAT_Hist  to   produce  the   linker   4NAT.   The   presence   of   fragment   at   282bp   indicates   successful   cloning.     Lane  2:  Undigested  pSecTagGCSF4NAT  _Hist.   Lane  3:    XhoI  digest  of  pSecTagGCSF4NAT_Hist.     Lane  4:    BamHI  digest  of  pSecTagGCSF4NAT_Hist.     Lane   M:   1kb   Ladder:   Molecular   weight   marker:   Bands   starting   at   the   bottom  (0.5,   1.0,   1.5,   2.0,   3.0,   4.0,   5.0,   6.0,   8.0,  10.0  Kb).                                 Figure   4-­‐10:   Double   digest   potentially   positive   clone   of                         pSecTagGCSF4NAT_Hist.   1%  agarose  gel  electrophoresis  analysis  of  double  digest  potentially  positive   clone  of  GCSF4NAT_Hist  using  XhoI/BamHI  as  a  double  digest  for  the  linker   (4NAT).   Lane   1:   Double   XhoI/BamHI   digests   of        1                2              3            4        M1      M2   pSecTagGCSF8QAT_Hist  to  produce  the  linker  8QAT.  The   presence   of   fragment   at   282bp   indicates   successful   cloning.     Lane  2:  Undigested  pSecTagGCSF8QAT  _Hist.   Lane  3:  XhoI  digest  of  pSecTagGCSF8QAT_Hist     Lane  4:  BamHI  digest  of  pSecTagGCSF8QAT_Hist     Lane   M1:   100bp   ladder:   Molecular   weight   marker:   Bands   starting   at   the   bottom   (100,   200,   300,   400,   500,   600,  700,  800,  900,  1000,  1200,  1517  bp).   Lane  M2:   1kb  Ladder:  Molecular  weight  marker:  Bands   starting   at   the  bottom  (0.5,   1.0,  1.5,   2.0,  3.0,  4.0,   5.0,  6.0,   8.0,  10.0  Kb).     Figure   4-­‐11:   Double   digest   potentially   positive   clone   of   pSecTag   GCSF8QAT_Hist   1%  agarose  gel  electrophoresis  analysis  of  double  digest  potentially  positive   clone   of   pSecTagGCSF8QAT_Hist   using   XhoI/BamHI   as   a   double   digest   for   the  linker  (8QAT).   91   In   conclusion,   the   presence   of   a   fragment   at   282bp   in   lane   1   for   both   figures   as   resulted   of   double   digest   using  Xho1/BamH1   indicates   successful   cloning   for  both  linker  4NAT  and  8QAT.  A  supercoiled  band  is  observed  in  lane  2  for   both   figures   as   a   resulted   of   undigested   plasmid.   Also,   a   single   band   is   observed   in   both   3   and   4   lanes   for   both   figures,   which   shows   that   the   plasmid  used  in  this  digestion  reaction  was  cut  by  the  individual  restriction   enzyme.     4.5 Expression  and  Analysis  of  GCSF  Tandems   After   the   successful   construction   of   GCSF2NAT_Hist,   GCSF2QAT_Hist,   GCSF4NAT_Hist,   GCSF4QAT_Hist,   GCSF8NAT_Hist   and   GCSF8QAT_Hist,   all   plasmids  were  transiently  and  stably  transfected  into  CHO  Flp-­‐In  cells,  and   serum   free   media   harvested   for   analysis   (refer   to   section   3.5).   Media   samples  were  analysed  using  Elisa  and  western  blotting  to  detect  expression   and  an  increase  in  molecular  weight  of  proteins  due  to  glycosylation.   4.5.1 Transient  Transfection  of  CHO  Flp-­‐In  Cells   As   an   initial   assessment   of   protein   expression,   all   plasmid   constructs   were   transiently   transfected   into   CHO   Flp-­‐In   cells   and   serum   free   media   harvested  and  analysed  via  Elisa  and  western  blot.   4.5.1.1 Analysis   o f   E xpression   b y   E lisa   To   verify   whether   or   not   protein   tandems   were   successfully   expressed,   media   of   protein   samples   were   analysed   via   Elisa   using   the   procedure   described  in  section  3.8.4.  The  results  of  Elisa  are  given  in  Table  4.3.       92   Table   4-­‐3:   GCSF   sandwich   Elisa   analysis   of   transiently   expressed   tandem  proteins.  GCSF  was  used  as  standard   Protein   μg/ml   nM   SEM   %CV   Negative  control   0.07   0.0   0.0   2.80   GCSF2QAT_Hist   2.35   51.77   0.02   3.89   GCSF2NAT_Hist   1.82   40.19   0.01   2.09   GCSF4QAT_Hist   0.94   20.72   0.01   1.95   GCSF4NAT_Hist   2.07   45.81   0.01   3.09   GCSF8QAT_Hist   3.00   61.88   0.03   6.28   GCSF8NAT_Hist   5.64   116.20   0.02   2.81     It   can   be   seen   in   the   table   above   that   the   Elisa   has   detected   all   tandem   proteins,   which   indicates   a   successful   transient   expression.   However,   the   purpose   of   this   method   is   to   detect   expression   only,   as   true   level   of   expression  is  likely  to  be  inaccurate  since  the  GCSF  standard  curve  is  based   upon   monomeric   GCSF,   not   specific   tandem   molecules   (more   details   in   discussion  part).     A  maximum  expression  of  5.64μg/ml  for  GCSF8NAT_Hist  is  observed  and  a   minimum   of   0.94μg/ml   for   GCSF4QAT_Hist.   %CV   of   less   than   10%   can   be   seen   for   all   protein   samples   which   is   considered   an   acceptable   degree   of   variability   between   samples   (between   replicate   values)   in   many   commercially  used  Elisa’s.     4.5.1.2 Analysis   o f   E xpression   b y   W estern   B lotting   To   assess   whether   protein   tandems   have   undergone   successful   glycosylation,   10μl   of   each   media   sample   was   diluted   equally   with   2   x   Laemmli   buffer   and   separated   on   a   10%   SDS-­‐PAGE   gel.   The   samples   were   then   transferred   to   PVDF   membrane   and   western   blotted   as   outlined   in   methods   (Section   3.8.3).   The   results   of   the   western   blot   are   presented   in   Figure  4.12.     93     Figure   4-­‐12:   Western   blot   of   media   samples   from   transiently   transfected  CHO  Flp-­‐In  cells     Lane  1*;  GCSF2QAT_Hist  (2  x  QAT).  Lane  2*;  GCSF2NAT_Hist  (2  x  NAT).  Lane   3;   GCSF4QAT_Hist   (4   x   QAT).   Lane   4;   GCSF4NAT_Hist   (4   x   NAT).   Lane   5;   GCSF8QAT_Hist  (8  x  QAT).  Lane  6;  GCSF8NAT_Hist  (8  x  NAT).  Western  blot   shows   a   definite   increase   in   molecular   weight   for   glycosylated   molecules   (GCSF2NAT_Hist,   GCSF5NAT_Hist,   and   GCSF8NAT_Hist)   compared   to   non-­‐ glycosylated   controls   (GCSF2QAT_Hist,   GCSF4QAT_Hist,   GCSF8QAT_Hist).   *Lane   1   and   2   were   taken   from   another   gels   as   didn’t   run   correctly   with   the   above  gel.     As  can  be  observed  in  Figure  4.12  western  blotting  successfully  detected  all   GCSF   tandems.   A   very   clear   shift   can   be   seen   in   molecular   weight   for   glycosylated   molecules   (GCSF2NAT_Hist,   GCSF5NAT_Hist,   and   GCSF8NAT_Hist)   compared   to   non-­‐glycosylated   controls   (GCSF2QAT_Hist,   GCSF4QAT_Hist,  GCSF8QAT_Hist).  Glycosylated  molecules  show  an  increase   in  molecular  weight,  which  is  consistent  with  their  linkers’  being  successful   glycosylated.              In  addition,  it  can  also  be  seen  that  there  is  a  slight  difference  in  molecular   weight   between   GCSF8QAT_Hist   compared   to   GCSF4QAT_Hist   and   GCSF2QAT,  this  is  due  to  the  difference  in  linker  length.  GCSF8QATHist  has  a   94   linker   length   of   282bp   =   94   amino   acids   compared   to   147bp   =49   amino   acids  for  both  GCSF4QAT_Hist  and  GCSF2QAT_Hist.     4.5.2 Stable  Cell  Line  Development  in  CHO  Flp-­‐In  Cell  Lines   Since   all   tandems   were   shown   to   be   expressed   and   intact   from   transient   transfections,  all  GCSF  tandems  were  taken  forward  to  make  stable  cell  lines   in  CHO  FIp-­‐In  cells  as  described  in  section  3.5.4.  Non-­‐transfected  cells  were   used  as  a  negative  control.       4.5.2.1 Analysis   o f   G CSF   T andems   b y   E lisa   Stable  cells  were  incubated  in  serum  free  media  for  2-­‐3  days  then  harvested   and  analysed  (See  section  3.5.5).  Elisa  data  is  shown  in  Table  4.4.         Table  4-­‐4:  Results  of  GCSF  sandwich  Elisa  for  stably  expressed  GCSF   tandems   Construct   μg/ml   nM   SEM   %CV   GCSF2QAT   2.82   62.04   0.038   8.11   GCSF2NAT   3.97   87.81   0.021   3.46   GCSF4QAT   4.67   102.95   0.047   6.98   GCSF4NAT   3.13   69.20   0.018   3.48   GCSF8QAT   7.69   158.62   0.032   3.29   GCSF8NAT   3.89   80.14   0.047   8.01     From   the   above   data,   all   protein   samples   have   been   detected.   It   can   be   shown   that   GCSF8QAT_Hist   is   the   most   highly   expressed   protein   at   7.69μg/ml   and   GCSF2QAT_Hist   being   the   lower   expressed   at   2.82μg/ml.   %CV  is  below  10%  for  all  protein  samples,  which  indicates  good  consistency   between  triplicates.           95   4.5.2.2 Analysis   o f   G CSF   T andems   b y   W estern   B lotting     10μl   of   each   serum   free   media   sample   was   equally   diluted   in   2   x   Laemmli   buffer   and   separated   on   a   10%   SDS-­‐PAGE   gel.   The   samples   were   then   transferred   to   PVDF   membrane   and   analysed   by   western   blotting.   The   results  are  shown  in  Figure  4.13.       Figure   4-­‐13:   Western   blot   of   stable   CHO   Flp-­‐In   cell   media   expressing   GCSF  tandems   Lane  1;  GCSF2QAT_Hist  (2  x  QAT).  Lane  2;  GCSF2NAT_Hist  (2  x  NAT).  Lane   3;   GCSF4QAT_Hist   (4   x   QAT).   Lane   4;   GCSF4NAT_Hist   (4   x   NAT).   Lane   5;   GCSF8QAT_Hist  (8  x  QAT).  Lane  6;  GCSF8NAT_Hist  (8  x  NAT).  Western  bot   shows   an   increase   in   molecular   weight   for   glycosylated   molecules   (GCSF2NAT_Hist,   GCSF4NAT_Hist,   and   GCSF8NAT_Hist)   compared   to   non-­‐ glycosylated  controls  (GCSF2QAT_Hist,  GCSF4QAT_Hist,  GCSF8QAT_Hist).       Western   blotting   successfully   detected   all   GCSF   tandems.   All   glycosylated   and  non-­‐glycosylated   GCSF   tandems   are   running   at   approximately   the   same   molecular   weights   that   were   observed   previously   in   transient   transfections.   This   suggests   the   separation   by   SDS-­‐PAGE   has   run   appropriately   and   that   samples   have   successful   glycosylation   at   the   consensus   sequence   within   the   linker  region.  Glycosylated  constructs  exhibited  increased  molecular  weight   over  non-­‐glycosylated  molecules  and  rhGCSF  (Table  4.5).   96   Table  4-­‐5:  Determined  and  observed  MW’s  of  expressed  GCSF  tandems       Construct   Determined  MW   Observed  MW  (kDa)   (kDa)*   (approximation)   rhGCSF   18.8                                                _   GCSF2QAT_Hist   45.4   ~  45   GCSF2NAT_Hist   45.2   ~  52   GCSF4QAT_Hist   45.4   ~  45   GCSF4NAT_Hist   45.2   ~  60   GCSF8QAT_Hist   48.5   ~  49   GCSF8NAT_Hist   48.5   ~  70     *MW’s   calculated   using   DNASTAR   Lasergene   version   8   and   the   observed   MW  estimated  using  western  blot  analysis  (see  Figure  4-­‐13).     4.6 In  vitro  Biological  Activity  of  Crude  Media   Biological   activity   of   GCSF   tandems   was   measured   using   the   human   acute   myeloid   leukemic   cell   line   (AML-­‐193   cell   line),   which   proliferates   in   response  to  GCSF.  To  determine  if  the  expressed  GCSF  tandem  proteins  are   biologically   active,  stable   clone   crude   media   was   tested   for   bioactivity   using   the  AML  193  proliferation  assay  as  described  in  section  3.8.5.  As  expected,   all  crude  media  obtained  from  stable  cell  lines  are  biologically  active  above   that   of   the   negative   control,   growth   hormone   (GH).   The   assay   appears   to   reach   absorbance   saturation   of   around   0.25   for   the   rhGCSF   and   all   GCSF   tandems  (Figure  4.14.A  &  B).       97     Figure  4-­‐14:  In  vitro  biological  activity  for  rhGCSF  and  GCSF  tandems   (A)  Represents  a  good  GCSF  standard  curve  showing  progressive  increase  in   biological   activity   when   tested   using   an   in   house   AML-­‐193   proliferation   assay.  (B)  Represents  biological  activity  of  GCSF  tandems.  Results  are  given   as  standard  mean  of  error  SEM  for  triplicate  wells.     The  in  house  bioassay  indicates  a  good  standard  curve  with  gradual  increase   in  the  absorbance  with  increasing  GCSF  concentrations.    All  constructs  have   biological   activity   with   a   maximum   absorbance   of   about   0.25   observed   for   rhGCSF,   glycosylated   GCSF   tandems   and   their   respective   controls   (non-­‐ glycosylated   tandems).   Whether   the   glycosylations   affect   the   biological   activity   is   difficult   to   interpret   since   quantification   was   completed   using   monomeric   GCSF   and   not   the   specific   tandems:   there   could   well   be   differences   in   detection   between   the   tandems   perhaps   due   to   steric   hindrance  by  glycosylations.           98   4.7 Discussion   In  this  chapter,  the  cloning,  sequencing  and  expression  of  two  GCSF  ligands   linked   by   a   variably   glycosylated   and   non-­‐glycosylated   linkers   were   examined.     The   data   obtained   has   shown   that   it   is   possible   to   incorporate   2,   4   and   8   NAT   (Asn-­‐Ala-­‐Thr)   or   QAT   (Glu-­‐Ala-­‐Thr)   motifs       between   tandem   GCSF   ligands.   The   NAT   linker   containing   tandems   have   been   successfully   glycosylated   as   shown   by   an   increase   in   the   molecular   weight   of   the   constructs   upon   expression   in   mammalian   CHO   Flp-­‐In   cells   as   verified   by   western  blotting.   CHO  Flp-­‐In  cell  lines  were  used  to  express  all  GCSF  constructs  as  these  cell   lines   have   been   widely   used   to   express   recombinant   proteins.     The   main   reason  for  their  use  is  the  possession  of  glycosylation  mechanism  similar  to   those   in   humans.   The   cells   are   efficient   and   easy   to   culture   and   relatively   express   high   level   of   proteins   (Damiani   et   al.,   2009).   Additionally,   these   mammalian   cell   lines   are   preferred   due   to   their   ability   to   correctly   fold   proteins   during   biosynthesis   (Wurm,   2004).     Ordinarily,   CHO   cell   lines   were   not   capable   of   both   sialyation   types   observed   in   human   body   (alpha   2,3   and   2,6-­‐glycosidic   linkages   are   participate   in   glycan   structure)   as   it   lacks   the   enzyme   alpha   2,6-­‐   sialyltransferase   and   this   could   cause   immunological   response   and   thus   influence   the   therapeutic   application   of   the   GCSF   constructs.   However,   CHO   cells   that   are   capable   of   sialylation   have   been   genetically   engineered   by   insertion   of   alpha   2,6-­‐sialylation   expressing   into   the   DNA   sequence   of   the   cells   (Damiani   2009).   Furthermore,   the   CHO   cell   line   is   the   industry   standard   for   recombinant   therapeutic   proteins   production   and   many   protein   drugs   are   manufactured   in   this   system   (mostly  antibodies)  without  problems.       The   Elisa   has   detected   all   tandem   proteins,   indicating   successful   transient   and   stable   expression.   However,   the   purpose   of   this   method   was   to   detect   expression   only   as   quantifiable   levels   of   expression   are   likely   to   be   inaccurate.   This   may   be   due   to   overestimating   of   Elisa   since   the   standard   99   curve  is  based  upon  monomeric  GCSF  and  therefore  is  not  ideal  to  measure   these  tandems.  Purified  GCSF  tandem  was  not  obtainable  at  the  same  time  of   carrying  out  these  experiments  and  thus  the  monomeric  GCSF  was  the  best   alternative.  As  a  result,  the  number  of  GCSF  molecules  per  mole  of  tandem  is   double   the   number   of   rhGCSF   molecules.   Besides,   although   our   GCSF   tandems   are   not   directly   glycosylated,   it   is   difficult   to   understand   how   the   influence   of   protein   conformation   and   glycosylation   of   these   tandem   proteins   affects   the   binding   of   antibody   and   thus   influence   Elisa   results.   Byrne  et  al.,  (2007)  indicated  that  glycosylation  could  completely  eradicate   the   polypeptide   recognition   site   of   antibody.   In   contrast,   analysis   of   these   tandems   using   western   blotting   may   provide   a   truer   visual   of   expression   since   proteins   are   in   their   denatured   form   and   this   may   expose   their   recognition  sites  for  the  antibody  more.  Protein  denaturation  as  a  result  of   sample   heating   may   also   reduce   the   possible   steric   hindrance   caused   by   glycosylation  to  antibody  recognition  sites.  It  is  important  to  consider  these   factors   since   it   is   difficult   to   determine   whether   or   not   equivalent   concentrations   of   protein   sample   have   been   tested   in   the   GCSF   bioactivity   assay.     Previously   we   stated   that,   western   blotting   showed   an   increase   in   the   molecular  weight  of  glycosylated  tandems  (GCSF2NAT_Hist,  GCSF4NAT_Hist   and  GCSF8NAT_Hist).  The  only  different  in  these  tandems  compared  to  the   non-­‐glycosylated   control   is   the   presence   of   N-­‐linked   glycosylation   motifs   with  in  the  linker,  supporting  successful  glycosylation.     The   key   behind   successful   N-­‐linked   glycosylation   is   the   selection   of   a   consensus   sequence   (As   mentioned   previously,   the   common   consensus   sequence   for   N-­‐linked   glycosylation   is   Asn-­‐X-­‐Ser/Thr   where   X   can   be   any   amino   acid   except   proline   (Kornfeld   and   Kornfeld,   1985).   Thus,   efficiency   of   glycosylation   is   likely   to   be   determined   by   the   amino   acid   (Thr)   that   occupies   the   third   position.   It   has   been   shown   that   glycosylation   is   more   efficient   when   Threonine   is   occupying   the   third   position   as   opposed   to   Serine,   with   efficiency   shown   to   be   twice   as   high   and   in   other   cases   even   forty  fold  (Kasturi  et  al.,  1995,  Picard  et  al.,  1995).     100   The   crude   media   of   each   tandem   GCSF   stable   clones   was   tested   using   the   AML   193   proliferation   assay.   At   this   stage,   this   assay   would   give   a   preliminary   indication   whether   GCSF   tandems   could   maintain   biological   activity   since   their   biological   activity   showed   similar   data   to   rhGCSF.   Whether   the   glycosylations   affect   the   biological   activity   is   difficult   to   interpret   since   quantification   was   completed   using   monomeric   GCSF   and   not   the   specific   tandems:   there   could   well   be   differences   in   detection   between  the  tandems  perhaps  due  to  steric  hindrance  by  glycosylations.   It   has   been   reported   that   the   insertion   of   a   Histidine   purification   tag   in   to   recombinant   proteins   may   potentially   dislocate   the   three-­‐dimensional   structure  of  the  polypeptide  protein  (Block  et  al.,  2009),  this  may  eventually   impact   upon   receptor   binding   and   negatively   affect   protein   bioactivity.   All   tandems  were  Hist  tagged  and  the  initial  data  of  AML193  assay  from  crude   media   showed   that   it   is   possible   to   introduce   another   GCSF   and   X6His   purification  tag  into  the  C-­‐terminus  of  a  GCSF  (to  create  a  tandem  GCSF)  and   still  maintain  biological  activity.  Pure  proteins  would  give  better  indication   as   to   whether   a   X6His   purification   tag   or   glycosylation   could   negatively   affect   protein   bioactivity.     For   further   analysis   and   characterisation,   GCSF   tandem  proteins  were  purified  using  IMAC.  This  will  be  discussed  in  the  next   chapter.       101   5. Results  2:  Large-­‐scale  Production  and  Analysis  of   GCSF  Tandems   5.1 Summary   The   previous   chapter   showed   that   it   is   possible   to   clone,   sequence   and   express   two   GCSF   ligands   linked   by   variable   glycosylated   and   non-­‐ glycosylated   linkers.   Therefore,   for   further   PK/PD   analysis,   it   was   necessary   to  purify  these  GCSF  tandem  proteins  using  Nickel  IMAC.  This  chapter  shows   that   IMAC   is   an   appropriate   method   for   purifying   GCSF   tandems   with   histidine  tags.  IMAC  permits  a  fast  and  easy  way  for  purification  due  to  the   strong  affinity  of  a  nickel-­‐complex  (Ni2+-­‐NTA)  for  the  histidine  tag  present  in   the   GCSF   tandems.     All   GCSF   tandems   produced   high   levels   of   protein   (between   1   to   4mg   per   litre   as   assessed   by   Bradford   assay)   from   a   stable   CHO   cell   line   and   were   considered   to   be   90-­‐95%   pure   as   judged   by   SDS-­‐ PAGE.   Molecular   weights   and   the   integrity   of   each   GCSF   tandem   were   confirmed   by   SDS-­‐PAGE   and   western   blotting.   Purified   protein   was   then   passed  over  for  PK/PD  studies.         102   5.2 Introduction     For   decades,   cloning   and   the   following   recombinant   expression   of   protein   is   commonly   used   in   molecular   biology.   As   purification   of   recombinant   proteins   is   frequently   a   challenging   step,   a   short   affinity   epitope   tag   was   designed   to   aid   purification   process.   Usually,   these   short   amino   acid   sequences  are  added  either  to  the  N  or  C  terminal  ends  of  a  protein.  Then,   these   amino   acid   sequences   can   expose   epitopes   for   specific   binding   partners   like   antibodies.   The   X6   His   tag   (HHHHHH)   is   one   of   the   most   commonly   used   protein   tags   that   permits   a   fast   and   easy   way   for   purification   that   is   based   on   the   strong   affinity   of   histidine   for   divalent   cations   (such   as   Zn,   Cu,   Co   and   Ni).   The   electron   donor   groups   on   the   imidazole   ring   of   histidine   form   coordination   bonds   with   the   immobilized   metal  ion  matrices  under  native  conditions  in  high  or  low  salt  buffers.  In  our   purification  system,  immobilized  nickel-­‐complex  Ni2+-­‐NTA  beads  were  used.   In   the   Immobilized   Metal   Affinity   Chromatography   (IMAC)   purification   system,   the   bound   protein   is   eluted   off   the   column   using   imidazole,   which   is   a   histidine   analogue,   as   it   competes   with   his-­‐tagged   protein   for   binding   to   the   divalent-­‐metal   ions   matrix.     As   a   result,   the   eluted   target   protein   can   simply   be   separated   from   unwanted   contaminated   proteins   in   the   elution   mixture  (Block  et  al.,  2009,  Kreisig  et  al.,  2014).   5.2.1 Aim  and  Hypothesis   Following   the   successful   cloning   and   expression   of   the   tandem   proteins   in   the   previous   chapter,   it   was   necessary   to   purify   these   tandem   proteins   from   mammalian  cell  lines  (CHO  FIp-­‐In  cells)  stably  expressing  each  tandem.  The   tandem   GCSF   expressing   cells   were   grown   in   roller   bottles   and   target   protein  harvested  from  the  media  as  a  secreted  product,  purified  and  passed   over  for  in  vivo  PK/PD  analysis.  In  order  to  facilitate  the  purification  of  the   target   proteins,   the   expression   constructs   were   tagged   at   the   C-­‐terminal   with   a   6-­‐Histidine.   As   Nickel   IMAC   column   is   known   to   have   high   binding   affinity   for   histidine,   we   hypothesized   that   the   histidine   tag   in   our   GCSF   tandems  will  bind  to  nickel  column  and  facilitate  the  purification.   103   5.3 Results   5.3.1 Cell  Growth  and  Productivity   Hyclone   SMF4CHO   Utility   media   adapted   stable   CHO   Flp-­‐In   cell   lines   of   all   GCSF   tandems   were   thawed   from   liquid   nitrogen   stocks.   After   a   period   of   growth   and   expansion,   cells   were   transferred   to   two   non-­‐vented   roller   bottles  containing  Hyclone  SFM4CHO  Utility  media  and  grown  at  37°C  with   gentle  agitation  in  a  CO2-­‐free  incubator  at  37°C,  for  protein  production  (each   roller   bottle   contained   approximately   500ml   of   culture   volume).   Starting   with   a   density   of   ~0.25   x   106   cells/ml,   the   cells   were   grown   until   viability   declined   below   30%   (took   approximately   10   -­‐   11   days)   as   previously   described   in   section   3.5.8.   100µl   of   media   samples   were   routinely   taken   every   2   or   3   days   and   analysed   for   total   cell   number   and   viability   using   a   trypan   blue   exclusion   (see   section   3.5.2).     Elisa   was   also   used   to   analyse   protein  productivity  as  described  in  section  3.8.4.   There   were   cell   growth   and   maximal   viable   cell   population   differences   observed   for   the   GCSF   tandems   during   the   expression   studies.   While   all   GCSF   tandem   cells   had   approximately   a   similar   %   of   viable   cells   (Figure   5.1b),   they   however   had   varied   peak   viable   cells   number,   ranging   from   1.25   to   1.9   million   cells   per   ml   (1.25-­‐1.9   x   106/ml)   with   GCSF2NAT   and   GCSF4NAT  having  the  least  and  highest  peak  cells  densities,  respectively.  A   steady  exponential  growth  was  observed  for  the  GCSF2QAT  and  GCSF4NAT   up   to   a   maximal   viable   cell   density   at   day   4   (viable   densities   of   1.7   x   106/ml   &   1.14   x   106/ml   respectively).   This   is   in   contrast   to   other   cell   lines   in   which   maximal   viable   cell   densities   were   reached   on   day   7   for   2NAT   (1.24   x   106/ml),  GCSF8NAT  (1.68  x  106/ml)  and  GCSF8QAT  (1.34  x  106/ml),  and  at   day   9   for   GCSF4QAT   (1.62   x   106/ml)   (Figure   5.1a).   The   GCSF8NAT   culture   produces   product   at   an   earlier   time   point   than   other   cultures,   peaking   on   day   7   but   losses   viability   at   an   earlier   stage.   Most   of   the   tandem   GCSF   cultures   retained   a   high   viability   up   to   at   least   day   7   (day   6   for   8NAT   culture)   with   viabilities   at   ~60%   (Figure   5.1b).   Consequently,   8NAT-­‐ expressing  cells  were  harvested  for  protein  production  on  day  7  of  growth,   104   as   the   %viability   declined   rapidly   beyond   this   period,   whereas   all   other   GCSF  tandems  were  grown  up  to  day  9  before  culture  media  were  harvested.   Beyond   day   9   there   was   a   significant   loss   of   cell   viability   (Figure   5.1b;   4QAT   and   4NAT   expressing   cells).   This   was   expected   as   nutrient   depletion   and   host  cell  metabolic  waste  would  impact  cell  proliferation.  Therefore,  at  the   time  of  harvest,  majority  of  the  tandem  GCSF  expressing  cells  were  at  least   60%   viable.   Harvesting   the   culture   media   at   this   stage   helped   prevent   any   potential  protein  degradation.     Interestingly,  protein  production  was  detectable  as  early  as  day  2  of  culture   growth   with   a   consistent   increase   in   protein   productivity   as   the   cells   number   increases   (Figure   5.1c).   There   was   a   dramatic   increase   in   protein   productivity   for   most   GCSF   tandems   up   to   at   least   day   7   with   productivity   peaking   at   >50mg/L   except   2QAT   and   8QAT   which   maintain   a   steady   low   level   of   expression   showing   a   maximal   peak   expression   of   10mg/L   &   17mg/L   respectively.   The   significant   difference   in   protein   productivity   for   most  of  the  GCSF  tandems  expressing  cells  was  observed  to  occur  between   day   4   and   day   9,   during   which   2NAT   and   4NAT   protein   yield   rose   from   approximately   20mg/L   on   day   7   to   about   55mg/L   and   50mg/L   on   day   9   respectively,   and   30mg/L   on   day   4   to   50mg/L   on   day   7   for   8NAT.   While   the   underlying  reason  for  the  sudden  enhanced  productivity  was  unclear,  it  was   inferred  that  the  expressing  cells  might  have  become  better  adapted  to  the   conditions  of  growth.     105         Figure  5-­‐1:  GCSF  tandems  expressing  cells  growth  and  productivity   The   figure   shows   the   total   viable   cells   (a)   and   %   viability   of   the   tandem   expressing   cells   (b)   for   all   glycosylated   tandem   proteins   (GCSF2NAT,   GCSF4NAT   &   GCSF8NAT)   together   with   their   respective   controls,   non-­‐ glycosylated   tandem   proteins   (GCSF2QAT,   GCSF4QAT   &   GCSF8QAT),   as   well   as  the  productivity  of  the  cells  within  a  9-­‐day  growth  period  (c).   106   To   confirm   the   integrity   of   the   expressed   GCSF   tandem   proteins,   the   protein   samples  from  the  GCSF  tandems  expressing  cells  were  analysed  by  western   blotting.  The  results  revealed  a  consistent  increase  in  the  protein  expression   of  all  GCSF  tandems  with  variable  productivity  in  each  tandem  protein  until   day   9   (Figure   5.2).   A   significant   loss   of   protein   was   equally   observed   for   cells   grown   beyond   day   9   (i.e.   GCSF4NAT   and   GCSF4QAT),   similar   to   what   was   earlier   observed   in   the   Elisa   analyses.   Consequently,   culture   media   were   routinely   collected   on   day   9   (except   8NAT   which   was   collected   on   day   7)  for  protein  purification.                                   107       Figure  5-­‐2:  Western  blot  analysis  of  roller  bottle  media  samples   Stable   CHO   Flp-­‐In   cells   expressing   the   GCSF   tandems   were   grown   in   roller   bottle  cultures  and  samples  taken  every  2  days  up  to  a  total  of  11  days.    10µl   of   samples   from   culture   medium   for   each   tandem   GCSF   construct   were   analysed  by  western  blotting.         108   5.3.2 Purification  of  GCSF  Tandems  Using  IMAC   The   culture   media   for   each   GCSF   tandem   was   taken   from   roller   bottle   and   spun   down   to   pellet   the   cell   debris   by   centrifugation   (18000   g,  JLA-­‐16.250   Fixed-­‐Angle   Rotor   for   30   minutes   at   4°C).   10mM   final   concentration   of   Benzamidine   HCl   (serine   protease   inhibitor)   was   added   to   the   1L   culture   media   to   prevent   protein   degradation.   Using   a   Vivaflow   200   concentrator,   media   samples   were   concentrated   about   10-­‐fold   i.e.   100ml.   The   sample   was   diluted  1:1  with  equilibration  buffer  prior  to  IMAC  as  previously  described   at  section  3.7.1.       5.3.2.1 Purification   o f   G CSF2NAT   Culture  media  collected  from  GCSF2NAT  expressing  cells  was  concentrated   and  purified  on  a  Nickel  IMAC  column.  Western  blotting  results  showed  that   the   majority   of   GCSF2NAT   protein   bound   to   the   IMAC   column   with   negligible  amounts  observed  in  the  flow  through  (Figure  5.3.A).  High  purity   protein   (>90%   pure)   was   eluted   using   imidazole   containing   buffers   with   concentrations   ranging   between   200mM   and   350mM   (shown   in   bold   typeface,   elutions   6,7,8   &   9   in   figure   5.3.A).   Elutions   using   imidazole   concentrations  below  or  above  this  range  (100mM  or  500mM  respectively)   were   observed   to   be   of   low   purity   when   analysed   by   SDS-­‐PAGE   and   were   thus   excluded   from   the   study.   High   purity   elutions   were   collected   and   pooled   together   for   dialysis.   Dialysis   was   performed   to   remove   any   excess   imidazole   or   other   salts   (Figure   5.3.B).     Pre   and   post   dialysed   proteins   were   measured   by   Bradford   assay   (Table   5.1)   and   analysed   by   10%   SDS   PAGE   followed   by   western   blot   analysis   to   confirm   the   purification   and   size   of   GCSF2NAT.   Coomassie   stained   10%   SDS   -­‐PAGE   showed   that   the   post-­‐ dialysis   GCSF2NAT   protein   was   stable   with   minimal   protein   degradation   (visible   as   contaminated   bands   at   20-­‐25kda).   These   lower   molecular   weight   (LMW)   bands   may   have   resulted   from   handling   or   partial   cleavage   of   the   tandem   GCSF2NAT   linker   (the   observed   molecule   at   this   low   molecular   weight   was   predicted   to   be   an   equivalent   of   a   GCSF   ligand   and   non-­‐ glycosylated  linker).  Analysis  of  pellet  samples  post  dialysis  does  show  the   109   presence   of   a   protein   band   at   ~50kDa,   however   this   was   not   detected   by   western  blotting.   Western   blotting   also   did   not   pick   up   the   LMW   bands   of   pre   and   post   dialysed   samples   present   post   coomassie   staining.     It   was   presumed   that   these   LMW   bands   are   missing   the   antigenic   site   or   are   below   detection   limit   of   antibody.   Also,   the   blots   might   not   have   been   exposed   long   enough   to   see   the   contaminant   bands.   Nevertheless,   the   LMW   proteins   were   of   significantly   lower   intensity   when   compared   to   the   intact   molecule,   suggesting   an   appreciably   high   purity   yield   during   purification.     A   total   of   1.96mg  of  GCSF2NAT  was  recovered  of  >90%  purity  from  1  litre  a  media.                     110     Figure  5-­‐3:  purification  development  of  IMAC  for  GCSF2NAT   10μl   of   protein   samples   from   each   stage   of   the   GCSF2NAT   purification   process   were   separated   on   a   10%   SDS-­‐PAGE   and   stained   with   coomassie   blue   for   visualization   (A,   left   figure),   while   100ng   of   the   same   sample   was   analysed  by  western  blotting  (A,  right  figure)  (UL=Unfiltered  load,  L=Load,   FL=Flow   through,   W=Wash   pH   7.4   &   6.0,   M=   Markers,   5-­‐10=   Elutions)   Thereafter,  active  elution  fractions  were  pooled,  dialysed  and  analysed  on  a   10%   SDS-­‐PAGE   gel   by   coomassie   staining   (B,   left   figure)   and   western   blotting  (B,  right  figure)  (1=  Pre-­‐Dialysis,  2=  Post-­‐Dialysis  and  Post-­‐Dialysis   pellet).       111   Table  5-­‐1:  Concentrations  of  GCSF2NAT  during  the  purification  process   The   table   features   the   concentrations   of   GCSF2NAT   at   each   stage   of   the   purification  process.  The  elutions  in  bold  typeface  (6,  7,  8  &  9)  were  pooled   together   giving   3.39mg   of   >90%   pure   GCSF2NAT   tandem   protein,   which   equated  to  a  1.95%  recovery  from  total  protein.  However,  ~  42%  of  this  was   lost   during   dialyses   giving   a   final   amount   of   1.96mg,   which   equated   to   1.13%  recovery  from  total  protein.   Imidazole   Sample   Concentration   (nM)   Volume   Protein   (ml)   (µg/ml)   Total  Protein                   %Recovery   (mg)   Unfiltered  load   -­‐   220   680.46   149.70   -­‐   Load   -­‐   220   790.48   173.90   100   Unbound   -­‐   220   655.83   144.28   83   Wash  pH  7.4   -­‐   1   69.62   0.07   0.04   E5   200   1   839.74   0.84   0.48   E6   200   1   539.24   0.54   0.3   E7   350   1   1031.86   1.03   0.59   E8   350   1   1333.99   1.33   0.76   E9   350   1   570.44   0.57   0.32   E10   500   1   622.99   0.62   0.36   Pre-­‐Dialysis   -­‐   4   846.31   3.39   1.95   Post-­‐Dialysis   -­‐   4   491.07   1.96   1.13   Pellet   -­‐   0.1   780.62   0.08   0.05             112   5.3.2.2 Purification   o f   G CSF2QAT   Culture  media  collected  from  GCSF2QAT  expressing  cells  was  concentrated   and  purified  on  a  Nickel  IMAC  column.  Western  blotting  results  showed  that   the  majority  of  GCSF2QAT  protein  in  the  media  was  bound  to  the  IMAC  with   negligible  amounts  observed  in  the  flow  through  (Figure  5.4.A).  High  purity   protein   (>90%   pure)   was   eluted   using   imidazole   containing   buffers   with   concentrations   ranging   between   200mM   and   350mM   (shown   in   bold   typeface   7,8,9   &   10   in   figure   5.4.A).   Elutions   using   with   imidazole   concentrations  below  or  above  this  range  (100mM  or  500mM  respectively)   were   observed   to   be   of   low   purity   when   analysed   by   SDS-­‐PAGE   and   were   thus   excluded   from   the   study.   High   purity   elutions   were   collected   from   IMAC  purification  and  pooled  together  for  dialysis.  Dialysis  was  performed   to   remove   any   excess   imidazole   or   other   salts   (Figure   5.4.B).   Pre   and   post   dialysed   protein   were   then   measured   by   Bradford   assay   (Table   5.2)   and   analysed  by  10%  SDS  PAGE  followed  by  western  blot  analysis  to  confirm  the   purification  and  size  of  GCSF2QAT.  Coomassie  stained  10%  SDS  -­‐PAGE  post-­‐ dialysis   protein   analyses   showed   that   the   GCSF2QAT   protein   was   highly   stable   with   very   minimal   protein   degradation   (visible   as   contaminated   bands  at  20-­‐25kda).  These  LMW  bands  may  have  resulted  from  handling  or   partial  cleavage  of  the  tandem  GCSF2QAT  linker.  The  observed  molecule  at   this   LMW   was   predicted   to   be   equivalent   to   GCSF   ligand   and   non-­‐ glycosylated   linker   since   similar   bands   corresponding   to   these   were   observed   on   western   blot   at   elution   8.   Analysis   of   pellet   samples   post   dialysis  showed  the  presence  of  a  protein  band  at  ~40kDa  of  low  intensity,   however  this  was  not  detected  by  western  blotting.   Western   blotting   also   did   not   pick   up   the   LMW   bands   of   pre   and   post   dialysed  samples  present  post  coomassie  staining.    However,  it  did  pick  up   the  LMW  band  in  elution  8  and  this  might  confirm  that  LMW  bands  are  not   missing   the   antigenic   site   or   are   below   detection   limit   of   antibody,   but   the   blots   might   not   have   been   exposed   long   enough   to   see   the   contaminant   bands.   Nevertheless,   the   LMW   proteins   were   of   significantly   lower   intensity   when   compared   to   the   intact   molecule,   suggesting   an   appreciably   high   113   purity   yield   during   purification.     A   total   of   1.81mg   of   GCSF2QAT   was   recovered  of  >90%  purity  from  1  litre  a  media.       Figure  5-­‐4:  Purification  development  of  IMAC  for  GCSF2QAT   10μl   of   protein   samples   from   each   stage   of   the   GCSF2QAT   purification   process   were   separated   on   a   10%   SDS-­‐PAGE   and   stained   with   coomassie   blue   for   visualization   (A,   left   figure),   while   100ng   of   same   samples   were   analysed   by   Western   blot.   (A,   right   figure)   (UL=Unfiltered   load,   L=Load,   FL=Flow   through,   W=Wash   pH7.4   &   6.0,   M=Markers,   5-­‐11=   Elutions).   Thereafter,  active  elution  fractions  were  pooled,  dialysed  and  analysed  on  a   10%   SDS-­‐PAGE   gel   by   coomassie   staining   (B,   left   figure)   and   western   blotting  (B,  right  figure)  (1=  Pre-­‐Dialysis,  2=  Post-­‐Dialysis  and  Post-­‐Dialysis   pellet).       114   Table  5-­‐2:  Protein  concentrations  of  GCSF2QAT  during  the  purification   process   The   table   features   the   concentrations   of   GCSF2NAT   at   each   stage   of   the   purification  process  in  the  IMAC  column.  The  elutions  in  bold  typeface  (7,  8,   9   &   10)   were   pooled   together   giving   2.47mg   of   >90%   pure   GCSF2QAT   tandem   protein,   which   equated   to   a   2.4%   recovery   from   total   protein,   and   a   fraction   (27%)   of   this   was   also   lost   during   dialyses   to   give   a   final   total   amount  of  1.81mg,  which  equated  to  1.75%  recovery  from  total  protein.   Imidazole   Sample   Concentration   (nM)   Volume   Protein   (ml)   (µg/ml)   Total  Protein                   %Recovery   (mg)   Unfiltered  load   -­‐   230   503.12   115.72   -­‐   Load   -­‐   230   448.93   103.25   100   Unbound   -­‐   230   440.72   101.37   98   Wash  pH  7.4   -­‐   1   153.37   0.15   0.15   Wash  pH  6.0     1   26.93   0.03   0.03   E1   50   1   150.08   0.15   0.15   E2   100   1   486.70   0.49   0.5   E3   100   1   260.10   0.26   0.25   E4   100   1   376.68   0.38   0.37   E5   200   1   567.16   0.57   0.55   E6   200   1   386.54   0.39   0.38   E7   200   1   600.00   0.60   0.58   E8   350   1   987.52   0.99   0.96   E9   350   1   737.93   0.74   0.72   E10   350   1   476.85   0.48   0.46   E11   500   1   148.44   0.15   0.15   Pre-­‐Dialysis   -­‐   6   616.42   2.47   2.4   Post-­‐Dialysis   -­‐   6   452.22   1.81   1.75   Pellet   -­‐   0.1   56.49   0.01   0.01   115   5.3.2.3 Purification   o f   G CSF4NAT   Culture  media  collected  from  GCSF4NAT  expressing  cells  were  concentrated   and  purified  on  a  Nickel  IMAC  column.    Western  blot  results  showed  that  the   majority   of   GCSF4NAT   released   from   CHO   cells   were   bound   to   the   IMAC   column  with  negligible  amounts  observed  in  the  flow  through  (Figure  5.5.A).   High  purity  protein  (~95%  pure)  was  eluted  using  imidazole  elution  buffers   with   concentrations   ranging   between   200mM   and   350mM   (shown   in   bold   typeface   5,6,7,8   &   9   in   figure   5.5.A).   Elutions   using   with   imidazole   concentrations   below   or   above   this   range   (e.g.   100mM   and   500mM)   analysed  were  considered  to  be  of  low  purity  when  analysed  by  SDS-­‐PAGE   and  were  thus  excluded  from  the  study.  High  purity  elutions  were  collected   from   IMAC   and   pooled   together   for   dialysis.   Dialysis   was   performed   to   remove   any   excess   imidazole   or   other   salts   (Figure   5.5.B).   Pre   and   post   dialysed   proteins   were   measured   by   Bradford   assay   (Table   5.3)   and   analysed  by  10%  SDS  PAGE  followed  by  western  blot  analysis  to  confirm  the   integrity  of  purified  GCSF4NAT  protein  and  size.    In   addition,   A   few   other   LMW   bands   observed   by   10%   SDS-­‐PAGE   (A,   left   figure)  were  confirmed  to  be  contaminants  since  no  bands  corresponding  to   these  were  observed  after  dialysis  (B,  left  figure).  Western  blotting  also  did   not   pick   up   the   LMW   bands   of   elutions   samples   present   post   coomassie   staining.     Nevertheless,   the   LMW   proteins   were   of   significantly   lower   intensity  when  compared  to  the  intact  molecule,  suggesting  an  appreciably   high   purity   yield   during   purification   A   total   of   4.11mg   of   GCSF4NAT   was   recovered  of  >95%  purity  from  1  litre  a  media.         116     Figure  5-­‐5:  Purification  analysis  of  IMAC  for  GCSF4NAT   10μl   of   protein   samples   from   each   stage   of   the   GCSF2NAT   purification   process   were   separated   on   a   10%   SDS-­‐PAGE   and   stained   with   Coomassie   blue   for   visualization   (A,   left   figure),   while   100ng   of   same   samples   were   analysed   by   Western   blot.   (A,   right   figure)   (UL=Unfiltered   load,   L=Load,   FL=Flow   through,   W=Wash   pH7.4   &   6.0,   M=Markers,   4-­‐10=   Elutions).   Thereafter,  active  elution  fractions  were  pooled,  dialysed  and  analysed  on  a   10%   SDS-­‐PAGE   gel   by   coomassie   staining   (B,   left   figure)   and   western   blotting  (B,  right  figure)  (1=  Pre-­‐Dialysis,  2=  Post-­‐Dialysis  and  Post-­‐Dialysis   pellet).       117   Table  5-­‐3:  Protein  concentrations  of  GCSF4NAT  during  the  purification   process   The   table   features   the   concentrations   of   GCSF4NAT   at   each   stage   of   the   purification   process   in   the   IMAC   column.   The   elutions   in   bold   typeface   (5,6,7,8  &  9)   were   pooled   together   giving   5.41mg   of   >95%   pure   GCSF4NAT   tandem   protein,   which   equated   to   a   5.2%   recovery   from   total   protein,   and   a   fraction   (24%)   of   this   was   also   lost   during   dialyses   to   give   a   final   total   amount  of  4.11mg,  which  equated  to  4%  recovery  from  total  protein.   Imidazole   Sample   Concentration   (nM)   Volume   Protein   (ml)   (µg/ml)   Total  Protein                   Recovery   (mg)   Unfiltered  load   -­‐   228   411.17   93.75   -­‐   Load   -­‐   228   452.22   103.11   100   Unbound   -­‐   228   381.61   87.01   84   Wash  pH  7.4   -­‐   1   72.91   0.07   0.07   Wash  pH  6.0     1   22.00   0.02   0.02   E1   100   1   146.80   0.15   0.15   E2   100   1   360.26   0.36   0.35   E3   100   1   205.91   0.21   0.2   E4   200   1   629.56   0.63   0.61   E5   200   1   616.42   0.62   0.6   E6   200   1   856.16   0.86   0.83   E7   350   1   565.52   0.57   0.55   E8   350   1   1337.27   1.34   1.3   E9   350   1   1141.87   1.14   1.1   E10   500   1   595.07   0.60   0.58   Pre-­‐Dialysis   -­‐   6   902.13   5.41   5.2   Post-­‐Dialysis   -­‐   6   685.46   4.11   4   Pellet   -­‐   0.1   549.10   0.05   0.05         118   5.3.2.4 Purification   o f   G CSF4QAT   Culture  media  collected  from  GCSF4QAT  expressing  cells  were  concentrated   and  purified  on  Nickel  IMAC  column.  Western  blotting  results  showed  that   the   majority   of   GCSF4QAT   protein   bound   to   the   IMAC   column   with   negligible  amounts  observed  in  the  flow  through  (Figure  5.6.A).  High  purity   protein   (>90%   pure)   was   eluted   using   imidazole   elution   buffers   with   concentrations   ranging   between   200mM   and   350mM   (shown   in   bold   typeface   9,10,11   &   12   in   figure   5.6.A).   Eluting   using   with   imidazole   concentrations  below  or  above  this  range  (100mM  or  500mM  respectively)   were   observed   to   be   of   low   purity   when   analysed   by   SDS-­‐PAGE   and   were   thus   excluded   from   the   study.   High   purity   elutions   collected   from   IMAC   purification   were   pooled   together   for   dialysis.   Dialysis   was   performed   to   remove   any   excess   imidazole   or   other   salts   (Figure   5.6.B).   Pre   and   post   dialysed   proteins   were   measured   by   Bradford   assay   (Table   5.4)   and   analysed  by  10%  SDS  PAGE  followed  by  western  blot  analysis  to  confirm  the   purification  and  size  of  GCSF4QAT.     Western   blotting   showed   that   the   post-­‐dialysis   GCSF2NAT   protein   was   stable   with   very   minimal   protein   degradation   (visible   as   contaminated   bands  at  20-­‐25kda).  These  LMW  bands  may  have  resulted  from  handling  or   partial  cleavage  of  the  tandem  GCSF4QAT  linker  (the  observed  molecule  at   this   low   molecular   weight   was   predicted   to   be   an   equivalent   of   a   GCSF   ligand   and   non-­‐glycosylated   linker).   Nevertheless,   the   LMW   proteins   were   of   significantly   lower   intensity   when   compared   to   the   intact   molecule,   suggesting   an   appreciably   high   purity   yield   during   purification.     A   total   of   1.11mg  of  GCSF4QAT  was  recovered  of  >90%  purity  from  1  litre  a  media.       119      Figure  5-­‐6:  Purification  analysis  of  IMAC  for  GCSF4QAT   10μl   of   protein   samples   from   each   stage   of   the   GCSF4QAT   purification   process   were   separated   on   a   10%   SDS-­‐PAGE   and   stained   with   Coomassie   blue   for   visualization   (A,   left   figure),   while   100ng   of   same   samples   were   analysed   by   Western   blot.   (A,   right   figure)   (UL=Unfiltered   load,   L=Load,   FL=Flow  through,  W=Wash  pH7.4,  M=Markers,  1-­‐14=  Elutions).  Thereafter,   active  elution  fractions  were  pooled,  dialysed  and  analysed  on  a  10%  SDS-­‐ PAGE  gel  by  coomassie  staining  (B,  left  figure)  and  western  blotting  (B,  right   figure)  (1=  Pre-­‐Dialysis  and  2=  Post-­‐Dialysis).       120   Table  5-­‐4:  Concentrations  of  GCSF4QAT  during  the  purification  process   The   table   features   the   concentrations   of   GCSF4QAT   at   each   stage   of   the   purification   process   in   the   IMAC   column.   The   elutions   in   bold   typeface   (9,   10,   11   &12)  were  pooled  together  giving  1.83mg  of  >90%  pure  GCSF4QAT   tandem   protein,   which   equated   to   a   1.2%   recovery   from   total   protein,   and   a   fraction   (40%)   of   this   was   also   lost   during   dialyses   to   give   a   final   total   amount  of  1.11mg,  which  equated  to  0.7%  recovery  from  total  protein.   Imidazole   Sample   Concentration   (nM)   Volume   Protein   (ml)   (µg/ml)   Total  Protein                   Recovery   (mg)   Unfiltered  load   -­‐   217   1472.6   319.55   -­‐   Load   -­‐   217   700.82   152.08   100   Unbound   -­‐   217   638.42   138.54   91   Wash  pH  7.4   -­‐   1   247.62   0.25   0.16   E1   50   1   0.00   0.00   0   E2   50   1   0.00   0.00   0   E3   50   1   0.00   0.00   0   E4   100   1   0.00   0.00   0   E5   100   1   0.00   0.00   0   E6   100   1   208.21   0.21   0.14   E7   200   1   188.51   0.19   0.13   E8   200   1   165.52   0.17   0.11   E9   200   1   667.98   0.67   0.44   E10   350   1   671.26   0.67   0.44   E11   350   1   444.66   0.44   0.29   E12   350   1   214.78   0.21   0.14   E13   500   1   0.00   0.00   0   Pre-­‐Dialysis   -­‐   5   366.83   1.83   1.2   Post-­‐Dialysis   -­‐   5   222.33   1.11   0.7   Pellet   -­‐   0.1   41.71   0.00   0         121   5.3.2.5 Purification   o f   G CSF8NAT   Analyses   of   GCSF8NAT   protein   pre-­‐   and   post-­‐   purification   by   western   blotting  showed  that  the  GCSF8NAT  protein  in  the  culture  medium  (secreted   by   CHO   cells)   was   able   to   bind   to   the   IMAC   column.   However,   the   results   equally   showed   protein   degradation   for   this   particular   tandem   protein   in   both  the  crude  culture  media  sample  (Figure  5.7.A)  and  the  purified  elution   samples   (Figure   5.7.B).   For   the   culture   media   samples,   the   western   blot   results   (Figure   5.7.A)   showed   consistent   increase   in   expressed   GCSF8NAT   protein   by   CHO   cells   in   the   culture   media   during   the   9   days   incubation   period,   running   at   the   right   molecular   weight   of   ~70kDa.   However,   below   the   target   protein   bands,   another   band   running   at   a   LMW   (~37kDa)   was   also  observed.  The  degradation  increased  during  purification  (seen  as  triplet   bands   running   at   ~15-­‐25kDa   and   another   band   at   ~37kDa),   suggesting   that   the   tandem   GCSF8NAT   is   not   as   stable   as   other   tandems.   Therefore,   a   number   of   measures   were   put   in   place   to   improve   the   stability   of   the   protein.       Figure   5-­‐7:   Western   blot   analysis   of   roller   bottle   media   samples   for   GCSF8NAT   (A)   Results   of   western   blot   shows   increasing   protein   expression   from   day   0   to   day   9   for   GCSF8NAT.   (B)   Western   blot   analysis   showing   IMAC   elutions   and   dialysed   samples   for   GCSF8NAT   (M=Markers,   1-­‐4=   Elutions,   5=   Pre-­‐ Dialysis,  6=  Post-­‐Dialysis  &  7=  Post-­‐Dialyses  Pellet).         122   To   improve   the   stability   of   GCSF8NAT,   the   incubation   temperature   for   the   stable  cell  growth  and  the  culture  volume  were  optimised.  The  CHO  cells  of   GCSF8NAT   were   grown   in   2   litres   not   1   litre   of   Hyclone   media   at   31°C   (as   against  previous  37°C).  The  cells  were  seeded  at  0.25x106/ml,  and  allowed   to   grow   for   7   days   or   until   viability   reached   70%   (rather   than   9   days   or   allowing   the   variability   to   decline   just   below   70%,   respectively).   Western   blotting   analyses   of   the   protein   samples   in   the   culture   media   (Figure   5.8)   showed  a  consistent  increase  in  GCSF8NAT  protein  expression  during  the  7   days   of   cell   growth   (bands   running   at   ~50-­‐70kDa)   with   no   other   visible   bands  observed,  which  is  an  indication  that  this  adjustments  enhanced  the   stability  of  GCSF8NAT.       Figure  5-­‐8:  Western  blot  analysis  of  roller  bottle  media  samples  for   GCSF8NAT   10μl   samples   taken   at   2-­‐3   days   intervals   from   GCSF8NAT   culture   media   grown   for   7   days   were   separated   by   SDS-­‐PAGE   and   analysed   by   western   blotting.   The   western   blot   results   shows   a   consistent   increase   in   protein   expression  from  day  0  to  day  7  for  GCSF8NAT.         123   Due   to   earlier   challenges   encountered   with   GCSF8NAT   degradation   and   instability  at  37°C,  the  GCSF8NAT  protein  collected  from  the  culture  under   the   optimised   conditions   was   purified   at   4°C   to   avoid   potential   protein   degradation.     Western   blotting   showed   that   the   GCSF8NAT   protein   in   the   crude   culture   media   expressed   from   CHO   cells   was   able   to   bind   the   IMAC   column   and   are   stable   when   eluted   (Figure   5.9.A).   The   bound   protein   was   eluted   with   a   gradient   concentration   of   imidazole   and   majority   of   high   purity   protein   was   eluted   between   an   imidazole   concentration   of   200nM   and   350nM   (shown   in   bold   typeface   4,   5,   6   &   7   in   Figure   5.9.A).   Elutions   using  with  imidazole  concentrations  below  or  above  this  range  (100nM  and   500nM  respectively)  were  considered  to  be  of  low  purity  when  analysed  by   SDS-­‐PAGE   and   thus   excluded   from   the   study.   High   purity   elutions   were   collected   from   IMAC   and   pooled   together   and   dialysed   in   PBS   so   as   to   remove   any   excess   imidazole   or   other   salts   (Figure   5.9.B).   Pre   and   post   dialysed  protein  were  measured  by  Bradford  assay  (Table  5.5)  and  analysed   by  10%  SDS  PAGE  followed  by  coomassie  staining  and  western  blot  analysis   in   order   to   confirm   the   purification   and   size   of   the   GCSF8NAT.   Coomassie   stained   10%   SDS   -­‐PAGE   showed   that   the   post-­‐dialysis   GCSF8NAT   protein   was  stable  with  very  minimal  protein  degradation  (visible  as  contaminated   bands   at   ~40kda).     Analysis   of   pellet   samples   post   dialysis   does   show   the   presence   of   a   protein   band   at   ~70kDa,   however   this   was   not   detected   by   western  blotting.   Western   blotting   also   did   not   pick   up   the   LMW   bands   of   pre   and   post   dialysed   samples   present   post   coomasie   staining.     It   was   presumed   that   these   LMW   bands   are   missing   the   antigenic   site   or   are   below   detection   limit   of   antibody.   Also,   the   blots   might   not   have   been   exposed   long   enough   to   see   the   contaminant   bands.   Nevertheless,   the   LMW   proteins   were   of   significantly   lower   intensity   when   compared   to   the   intact   molecule,   suggesting   an   appreciably   high   purity   yield   during   purification.     A   total   of   2.91mg  of  GCSF8NAT  was  recovered  of  >95%  purity  from  2  litre  a  media.       124     Figure  5-­‐9:  Purification  analysis  of  IMAC  samples  for  GCSF8NAT   10μl   of   protein   samples   from   each   stage   of   the   GCSF8NAT   purification   process   were   separated   on   a   10%   SDS-­‐PAGE   and   stained   with   coomassie   blue   for   visualization   (A,   left   figure),   while   100ng   of   same   samples   were   analysed   by   western   blot.   (A,   right   figure)   (UL=Unfiltered   load,   L=Load,   FL=Flow   through,   W=Wash   pH7.4,   6.0   M=Markers,   1-­‐11=   Elutions).   Thereafter,  active  elution  fractions  were  pooled,  dialysed  and  analysed  on  a   10%  SDS-­‐PAGE  gel  by  coomasie  staining  (B,  left  figure)  and  western  blotting   (B,   right   figure)   (1=   Pre-­‐Dialysis   and   2=   Post-­‐Dialysis   &   3=   Post-­‐Dialysis   pellet).   125   Table  5-­‐5:  Concentrations  of  GCSF8NAT  during  the  purification  process   The   table   features   the   concentrations   of   GCSF4NAT   at   each   stage   of   the   purification  process  in  the  IMAC  column.  The  elutions  in  bold  typeface  (4,  5,   6   &   7)   were   pooled   together   giving   3.21mg   of   >95%   pure   GCSF8NAT   tandem  protein,  which  equated  to  a  0.67%  recovery  from  total  protein  with   a   fraction   (9%)   of   this   was   also   lost   during   dialyses   to   give   a   final   total   amount  of  2.91mg,  which  equated  to  0.61%  recovery  from  total  protein.   Imidazole   Sample   Concentration   (nM)   Volume   Protein   (ml)   (µg/ml)   Total  Protein                   Recovery   (mg)   Unfiltered  load   -­‐   220   3787.74   833.30   -­‐   Load   -­‐   220   2162.52   475.75   100   Unbound   -­‐   220   2074.96   456.49   96   Wash  pH  7.4   -­‐   1   3952.36   3.95   0.83   Wash  pH  6.0     1   22.42   0.02   0.004   E3   200   1   1129.25   1.13   0.24   E4   200   1   814.01   0.81   0.17   E5   200   1   1560.07   1.56   0.33   E6   350   1   1528.55   1.53   0.32   E7   350   1   589.84   0.59   0.33   E8   350   1   635.38   0.64   0.13   Pre-­‐Dialysis   -­‐   4   802.80   3.21   0.67   Post-­‐Dialysis   -­‐   4   727.50   2.91   0.61   Pellet   -­‐   0.1   601.40   0.06   0.01         126   5.3.2.6 Purification   o f   G CSF8QAT   Crude  culture  medium  from  stable  CHO  cells  expressing  tandem  GCSF8QAT   proteins   were   passed   over   an   IMAC   column.   Western   blotting   results   showed   that   the   majority   of   the   tandem   protein   was   bound   to   the   IMAC   column   and   only   very   little   was   observed   in   the   flow   through   (Figure   5.10.A   Wrong).  High  purity  protein  was  eluted  between  an  imidazole  concentration   of  200nM  and  350nM  (shown  in  bold  typeface  5,  6,  7  &  8  in  figure  5.10.A).   Other  elutions  below  or  above  this  range  (100nM  and  500nM  respectively)   were   observed   to   be   of   low   purity   when   analysed   by   SDS-­‐PAGE   and   thus   excluded   from   the   study.   High   purity   elutions   were   collected   and   pooled   together   for   dialysis   so   as   to   remove   any   excess   imidazole   or   other   salts   (Figure   5.10.B).   Pre   and   post   dialysed   protein   samples   were   measured   by   Bradford   assay   (Table   5.6)   and   analysed   by   10%   SDS   PAGE   followed   by   western  blot  analysis  to  confirm  the  purification  and  size  of  GCSF8QAT.     Coomassie   stained   10%   SDS   –PAGE   showed   the   post-­‐dialysis   GCSF8QAT   protein   was   stable   as   no   degradation   or   other   visible   contaminated   bands   were   observed   by   western   blot.   However,   a   very   minimal   amount   of   contaminated   bands   (presumed   to   be   degraded   tandem   GCSF8QAT   protein)   are   observed   on   the   coomassie   stained   gel,   which   are   insignificant   when   compared   to   the   amount   of   pure   protein   running   at   the   predicted   molecular   weight.  A  total  of  2.93mg  of  GCSF8QAT  was  recovered  of  >95%  purity  from   1  litre  a  media.         Also,   dimer   formation   was   observed   pre   and   post   dialysis   which   could   be   seen  as  bands  running  at  twice  the  predicted  molecular  weight  of  GCSF8QAT   in  the  western  blotting  image  (Figure  5.10.B,  right  picture).       127     Figure  5-­‐10:  Purification  development  of  IMAC  for  GCSF8QAT   10μl   of   protein   samples   from   each   stage   of   the   GCSF8QAT   purification   process   were   separated   on   a   10%   SDS-­‐PAGE   and   stained   with   coomassie   blue   for   visualization   (A,   left   figure),   while   100ng   of   same   samples   were   analysed   by   Western   blot.   (A,   right   figure)   (UL=Unfiltered   load,   L=Load,   FL=Flow   through,   W=Wash   pH7.4   &   6.0,   M=Markers,   1-­‐11=   Elutions).   Thereafter,  active  elution  fractions  were  pooled,  dialysed  and  analysed  on  a   10%  SDSPAGE  gel  by  coomassie  staining  (B,  left  figure)  and  western  blotting   (B,  right  figure)  (1=  Pre-­‐Dialysis  and  2=  Post-­‐Dialysis  &  3=  Post-­‐D-­‐Pellet).       128   Table  5-­‐6:  Protein  concentrations  of  GCSF8QAT  during  the  purification   process   The   table   features   the   concentrations   of   GCSF4QAT   at   each   stage   of   the   purification   process   in   the   IMAC   column.   Elutions   5   to   8   were   pooled   together   giving   3.96mg   of   >95%   pure   GCSF8QAT   tandem   protein,   which   equated   to   a   1.5%   recovery   from   total   protein   pre-­‐dialyses   and   a   fraction   (27%)   of   this   was   also   lost   during   dialyses   to   give   a   final   total   amount   of   2.93mg,  which  equated  to  1.1%  recovery  from  total  protein.           Imidazole   Sample   Concentration   (nM)   Volume   Protein   (ml)   (µg/ml)   Total  Protein                   Recovery   (mg)   Unfiltered  load   -­‐   230   1406.90   323.59   -­‐   Load   -­‐   230   1144.17   263.16   100   Unbound   -­‐   230   779.64   179.32   68   Wash  pH  7.4   -­‐   1   204.93   0.20   0.08   E1   50   1   116.26   0.12   0.05   E2   50   1   418.39   0.42   0.16   E3   200   1   707.39   0.71   0.27   E4   200   1   592.45   0.59   0.22   E5   200   1   1232.84   1.23   0.5   E6   350   1   1594.09   1.59   0.6   E7   350   1   1091.63   1.09   0.4   E8   350   1   1265.68   1.27   0.5   E9   500   1   648.28   0.65   0.24   Pre-­‐Dialysis   -­‐   4   990.80   3.96   1.5   Post-­‐Dialysis   -­‐   4   732.75   2.93   1.1   Pellet   -­‐   0.1   56.49   0.01   0.004         129   5.3.2.7 Summary   o f   G CSF   P rotein   T andems   P urification   It  was  possible  to  purify  GCSF  tandems  linked  by  a  flexible  linker  (Gly4Ser)n   from   a   mammalian   cell   line   (CHO   Flp-­‐In   cells).     The   glycosylated   tandems   showed   an   increase   in   molecular   weight   above   that   of   their   controls   (non-­‐ glycosylated   tandems)   as   assessed   by   SDS-­‐PAGE   and   western   blotting   (Figure  5.11).     Figure  5-­‐11:  Purified  GCSF  tandems  analysed  by  coomassie  blue  and   western  blot   (A)   Purified   GCSF   tandem   molecules   analysed   by   coomassie   blue.   Lane   1;   GCSF2QAT,   Lane   2;   GCSF2NAT,   Lane   3;   GCSF4QAT,   Lane   4;   GCSF4NAT,   Lane   5;   GCSF8QAT,   Lane   6;   GCSF8NAT,   Lane   M;   1kb   marker.   A   total   of   7.5µg  protein  was  loaded  per  lane   (B)   Purified   GCSF   tandem   molecules   analysed   by   western   blot.     Lane   1;   GCSF   2QAT,   Lane   2;   GCSF2NAT,   Lane   3;   GCSF   4QAT,   Lane   4;   GCSF4NAT,   Lane   5;   GCSF   8QAT,   Lane  6;  GCSF8NAT.   A   total  of   100ng  protein  was  loaded  per  lane       In   addition,   both   coomassie   blue   and   western   blotting   analyses   showed   a   large   smeared   band   for   each   individual   glycosylated   GCSF   tandem   when   compared   to   the   corresponding   non-­‐glycosylated   tandem   control,   which   could  be  attributed  to  a  large  heterogeneous  protein  population.  The  reason   for   the   observed   increasing   degree   of   population   heterogeneity   across   the   tandem  GCSF  proteins  will  be  discussed  later.       130   5.4 Discussion   The   expression   studies   of   the   GCSF   tandems   containing   a   C-­‐terminal   histidine   tag   showed   that   the   target   genes   could   be   stably   expressed   in   a   CHO   cell   line.   However,   glycosylated   tandems   showed   high   level   of   expression   compared   to   non-­‐glycosylated   tandems   (i.e.   both   2QAT   and   8QAT   showed   the   lowest   level   of   expression   during   culture   growth).   Since   there   is   no   difference   between   glycosylated   and   non-­‐glycosylated   tandems   except  for  the  linker,  it  seems  that  the  presence  of  NAT  motifs  is  enhancing   DNA   transcription   more   than   QAT   motifs,   which   in   turn   increased   the   productivity  of  protein.     The   purification   data   indicated   that   IMAC   was   an   appropriate   method   for   purifying  the  GCSF  tandems.  All  GCSF  tandems  were  easily  eluted  from  the   IMAC   by   adding   a   high   concentration   of   imidazole   (an   analogue   of   Histidine).   IMAC   permits   a   fast   and   easy   way   for   purification   due   to   the   strong  affinity  of  a  nickel-­‐complex  Ni2+-­‐NTA  to  the  histidine  sequences  of  the   GCSF  tandems.  Consequently,  all  constructs  produced  sufficient  quantities  of   pure   protein   (between   1   to   4mg   per   litre   as   assessed   by   Bradford   assay),   which   was   sufficient   for   the   PK/PD   studies,   and   was   considered   to   be   90-­‐ 95%  pure  as  assessed  by  SDS-­‐PAGE  followed  by  western  blotting  to  confirm   the  molecular  weight  and  integrity  of  each  construct.   Some  contaminating  bands  were  observed  for  all  GCSF  tandems  during  the   purification  process  as  evidenced  by  coomassie  staining.  While  these  bands   appeared  to  be  degraded  products,  the  amount  of  degradation  observed  was   negligible   when   compared   to   the   stable   non-­‐degraded   proteins.   Interestingly,   most   of   the   contaminated   or   degraded   GCSF   tandems   were   separated   from   the   stable   and   pure   proteins   by   dialysis   through   a   semipermeable   membrane   with   10kDa   molecular   weight   cut-­‐off   (e.g.   a   cellulose   membrane   with   pores).   Dialysis   technique   is   an   important   step   after   purification   for   removing   any   excess   imidazole,   salts   or   other   small   molecules.   Molecules   that   have   sizes   bigger   than   the   cellulose   membrane   pores   were   retained   inside   the   dialysis   bag,   while   other   smaller   molecules   or   salts   diffuse   through   the   pores   into   the   dialyses   buffer   Post-­‐dialysis   131   analysis  by  coomassie  staining  or  western  blotting  showed  all  GCSF  tandems   to  be  of  the  correct  molecular  weight  with  no  degradation  as  there  were  no   visible   contaminants   or   degraded   bands   observed,   suggesting   that   the   dialysis  was  efficient  for  removing  the  contaminants  for  most  of  the  purified   samples.     However,  for  the  majority  of  tandem  molecules  expression  and  purification   was   completed   without   difficulty,   however   for   one   tandem   in   particular   (GCSF8NAT)   we   experienced   difficulties   with   the   stability   during   the   expression   and   purification.   It   was   difficult   to   determine   what   caused   the   degradation,   as   other   tandems   were   stable   when   purified   using   the   same   purification   method.   To   investigate   the   reason   for   GCSF8NAT   instability,   samples   were   taken   during   cell   growth   in   roller   bottles   before,   during   and   after   the   purification.   The   protein   samples   were   analysed   by   western   blotting,   and   the   results   revealed   that   the   degradation   occurred   during   both   protein  expression  and  purification  processes.  Consequently,  the  conditions   for   expressions   and   purifications   were   modified   with   a   view   to   improving   the  stability  of  GCSF8NAT  during  the  entire  protein  production  processes.     Firstly  the  incubation  time  was  reduced  from  9  days  to  7  days  so  that  cells   could   be   harvested   at   a   higher   viability   (60%   or   higher)   as   it   was   earlier   noticed   that   growing   CHO   cells   for   longer   time   points   led   to   increased   protein  degradation,  probably  due  to  increased  protease  activity  associated   with  the  increased  cell  death.     The  temperature  at  which  growth  and  expression  was  also  reduced  to  31°C   from  37°C.  These  adjustments  made  to  the  conditions  of  protein  expression   were   necessary,   as   it   has   been   shown   that   while   longer   linkers   were   preferable   for   the   preservation   of   the   independent   folding   and   biological   activities   of   two   proteins,   they   could   also   be   easily   cleaved   by   proteases,   because   the   structures   and   the   adjacent   regions   of   these   linkers   are   more   loosely  connected  (Liu  et  al.,  2005).    A  similar  pattern  was  observed  with  the  GCSF8NAT,  which  has  a  long  linker   similar   to   GCSF8QAT   (non-­‐glycosylated   control).   In   the   initial   expression   132   studies   under   the   same   conditions,   GCSF8NAT   protein   was   degraded   whereas   GCSF8QAT   was   not.   Another   contributory   factor   to   GCSF8NAT   degradation   could   be   associated   with   the   presence   of   8   N-­‐linked   glycosylation   sites,   which   restricts   intramolecular   interaction   between   the   two  GCSF  ligands  and  thus  leaving  the  linker  exposed  to  proteolytic  attack.   However,   During   protein   expression,   there   is   a   probability   that   the   GCSF   molecules  in  GCSF8QAT  (non-­‐glycosylated  control)  for  instance  would  bind   to   each   other   due   to   the   formation   of   intra-­‐molecular   disulphide   bond   resulting   from   the   interaction   of   free   cysteine   residues   (Cys17)   that   were   present  on  the  individual  GCSF  molecule  (which  was  observed  to  be  the  case   as  dimer  formation  was  earlier  observed  for  the  tandem  GCSF8QAT  but  not   in   GCSF8NAT)   Consequently,   the   binding   of   the   GCSF   molecules   in   GCSF8QAT  would  protect  the  linker,  to  an  extent,  from  proteolytic  cleavage.     The   beneficiary   effect   of   decreasing   culture   temperature   was   seen   on   EPO   production   in   CHO   cell   line.   Consequently,   a   2.5-­‐fold   increase   in   the   maximum   concentration   of   EPO   was   achieved   by   decreasing   temperature   from   37°C   to   33°C   (Yoon   et   al.,   2003).   Lowering   the   culture   temperature   increase   protein   expression   by   slowing   down   protein   translation   and   thus   facilitates  correct  folding  of  protein.   Lowering   the   temperature   during   culture   of   GCSF8NAT   equally   enhanced   the  protein  production  when  compared  to  other  tandem  GCSF  proteins.  This   is  evidenced  by  a  49mg/L  productivity  with  GCSF8NAT  at  day  7  during  cell   growth  in  roller  bottles  compared  to  21mg/L,  19.3mg/L,  8.1mg/L,  34.5mg/L   and   16.3mg/L   yield   with   GCSF4NAT,   GCSF2NAT,   GCSF2QAT,   GCSF4QAT   and   GCSF8QAT  respectively.   Also,  the  adjustments  made  to  the  conditions  of  expression  cells  growth  for   the   tandem   GCSF8NAT   was   further   justified   by   the   abrogation   of   GCSF   protein   degradation   that   were   earlier   observed   during   initial   protein   expression   and   purification.   To   prevent   potential   protein   degradation   during   the   purification   process,   especially   for   GCSF8NAT,   all   the   tandems   GCSF  purifications  were  purified  at  4°C.    The  western  blotting  was  used  to   133   confirm   the   stability   since   degraded   proteins   were   not   easily   visible   in   the   gel   with   coomassie   staining   (~5-­‐10μg   purified   protein),   whereas   western   blot  offers  a  better  detection  with  ~0.1μg  of  protein  load.     Tandem  proteins  analysed  by  SDS-­‐PAGE  and  western  blotting  also  revealed   an   obvious   increase   in   the   population   heterogeneity   in   glycosylated   tandems  that  was  not  apparent  in  the  non-­‐glycosylated  tandem  controls.  A   possible   reason   for   the   observed   high   heterogeneity   of   the   population   could   be  due  to  both  macro-­‐  (glycosylation  site  occupancy)  and  micro-­‐  (structure   of  glycosylation  relating  length,  composition  and  branching  pattern),  which   is   frequently   present   in   the   final   population   of   recombinantly   expressed   glycosylated   proteins   (Sola   and   Griebenow,   2009   and   Sinclair   and   Elliott,   2005).   In   conclusion   the   glycosylated   GCSF   tandems   and   their   respective   non-­‐ glycosylated  controls  were  successfully  purified  by  IMAC  and  the  purity  was   90-­‐95%   as   assessed   by   SDS-­‐PAGE   and   western   blotting.   However,   before   being   passed   over   for   the   future   PK/PD   in   vivo   study   it   was   required   to   measure  their  biological  activity.  This  will  be  discussed  in  the  next  chapter.       134   6. Results  3:  In  vitro  Bioactivity  Evaluation  and   Temperature  Stability  of  GCSF  Tandems   6.1 Summary   Purification   of   glycosylated   and   non-­‐glycosylated   GCSF   tandem   molecules   was   achievable,   as   evidenced   by   the   previous   chapter.   The   purified   glycosylated   tandem   molecules   also   showed   increased   molecular   weight   above   that   of   controls   when   analysed   by   SDS-­‐PAGE   and   western   blotting.   However,  before  analysing  the  pharmacokinetics  and  pharmacodynamics  of   these  tandem  molecules,  it  is  imperative  to  analyse  their  in  vitro  bioactivity   and  stability.  In  this  study,  the  in  vitro  bioactivity  of  the  GCSF  tandems  was   tested   using   an   AML-­‐193   proliferation   assay.   These   cells   have   been   shown   to   proliferate   in   response   to   GCSF   treatment.   The   short-­‐term   stability   of   GCSF   tandems   was   investigated   by   two   different   ways.   First   by   testing   samples   from   the   stability   experiment   in   the   AML-­‐193   assay   at   37°C   for   3   days.   Second   by   testing   purified   samples   from   stock   at   3   different   temperatures   (4°C,   room   temperature   (RmT)   and   -­‐80°C   freeze   thaw   (F/T)   cycles)   over   an   8   day   period.   The   results   indicated   significant   increased   bioactivity  for  GCSF  tandems  compared  with  rhGCSF  (EC50  for  tandems  were   about   3-­‐fold   lower   than   that   for   rhGCSF).   All   GCSF   tandems   showed   good   stability   with   no   visible   signs   of   degradation   under   all   conditions   studied   (4°C,   RmT   and   -­‐80°C   freeze   thaw)   over   an   8   days   period   and   also   after   incubation  with  AML-­‐193  cells  at  37°C  for  3  days.         135   6.2 Introduction   Acute   myeloid   leukemia   (AML)   is   a   disorder   of   the   myeloid   lineage   characterized   by   the   quick   growth   of   abnormal  white   blood   cells  (WBCs)   that   accumulate   in   the   bone   marrow  and   result   of   exhibiting   a   new   morphological   and   immunophenotypic   features   to   the   myeloid   lineage   (Vardiman  et  al.,  2009).    AML-­‐193  is  one  of  8  cell  lines  that  were  established   from  50  patients  with  childhood  acute  leukemia  (Lange  et  al.,  1987,  Valtieri   et   al.,   1991).   From   those   eight   cell   lines,   AML-­‐193   was   derived   from   a   13-­‐ year-­‐old  female  with  acute  myeloid  leukemia,  which  was  then  the  only  cell   line   that   required   conditioned   media   to   grow   and   proliferate.   In   vitro,   cytokines   such   as   GM-­‐CSF   and   GCSF   have   been   identified   to   support   the   growth   and   proliferation   of   hematopoietic   stem   cells.   Interestingly,   these   cytokines  have  also  been  shown  to  support  the  growth  and  proliferation  of   AML-­‐193   cells   (Favreau   and   Sathyanarayana,   2012).   Based   on   these   findings,   we   hypothesized   that   AML-­‐193   cell   line   would   proliferate   in   response  to  GCSF  tandem  proteins  in  a  manner  comparable  to  the  available   rhGCSF.   Therefore   the   proliferation   bioassay   was   used   to   test   the   hypothesis.     The  basic  proliferation  bioassay  is  a  colorimetric  method  routinely  used  to   determine  the  number  of  viable  AML-­‐193  cells.  As  earlier  mentioned,  AML-­‐ 193   growth   is   stimulated   in   the   presence   of   an   active   GCSF   protein.   A   commercially   available   MTS   reagent   which   contains   a   tetrazolium   compound   and   an   electron   coupling   reagent   is   used   to   estimate   the   proliferation   of   cells.   Generally,   metabolically   active   cells   produce   dehydrogenase   enzymes   that   produce   NADPH   into   the   medium   as   a   by-­‐ product.   The   produced   NADPH   reduces   the   MTS   tetrazolium   compound   in   the   medium   to   give   a   yellow   coloration   (formazan   product   is   soluble   in   tissue   culture   media),   which   can   be   read   photometrically   at   490nm   (Berridge  and  Tan,  1993).  The  MTS  compound  is  hydrolyzed  by  the  NADPH   producing   cells   into   a   yellow   coloured   product.   The   intensity   of   the   colour   produced  is  directly  proportional  to  the  concentration  of  the  active  enzymes   present   in   the   medium,   which   in   turn   is   dependent   on   the   number   of   cells   136   present   (Cory   et   al.,   1991).   Therefore   cell   growth   is   directly   related   to   the   level  of  NADPH  in  the  media  and  hence  the  activity  of  GCSF.   In   the   modified   proliferation   assay   used   in   this   study,   the   AML-­‐193   cells   were   seeded   in   96   well   plates   to   the   rhGCSF,   GCSF   tandems   and   their   controls   separately   and   grown   for   3   days.   MTS   reagent   was   then   added   to   estimate  the  proliferation  of  AML-­‐193  cells  that  respond  to  the  presence  of   GCSF  in  the  medium.    Readings  were  taken  at  490nm  every  40  minutes  for  2   hrs  using  a  96  well  plate  reader.     6.2.1 Aim  and  Hypothesis     It  is  essential  that  the  biological  activity  of  a  protein  be  retained  following  its   purification.   Maintaining   the   activity   of   the   target   protein   is   especially   important  for  the  future  pharmacokinetic  (PK)  in   vivo  study.    Also,  proteins   are   best   known   to   perform   their   biological   functions   when   in   the   right   conformation.  Hence,  it  is  essential  to  investigate  the  biological  activities  of   the   GCSF   tandems,   at   different   temperature   conditions   to   determine   the   stability   of   the   molecules.   Therefore,   the   aim   of   this   chapter   is   to   measure   the  biological  activity  of  the  glycosylated  GCSF  tandems  using  AML-­‐193  cell   line,   and   to   determine   the   short-­‐term   stability   of   these   constructs   under   different   temperatures.   The   stability   was   assessed   at   3   different   temperatures   (4°C,   RmT   and   -­‐80°C   freeze   thaw   (F/T)   cycles)   over   8   days   period.   Also,   samples   of   GCSF   tandems   were   taken   after   incubation   with   AML-­‐193  cells  at  37°C  for  3  days  and  tested  for  stability.  We  hypothesized   that   our   GCSF   tandems   will   be   biologically   active   and   stable   similar   to   the   rhGCSF  post-­‐purification.         137   6.3 Results   6.3.1 In  vitro  Bioactivity  Evaluation   The   biological   activity   of   GCSF   tandems   was   evaluated   using   an   AML-­‐193   cell-­‐based   proliferation   assay.   AML-­‐193   cells   were   removed   from   liquid   nitrogen  storage  and  prepared  as  described  in  Section  3.8.5.1.  50µl  of  serial   dilutions   of   rhGCSF   and   GCSF   tandem   proteins   were   added   into   the   appropriate  well  of  a  96-­‐well  microplate.  50µl  of  AML-­‐193  cell  suspension  at   a   density   of   5x105   cells/ml   was   then   added   to   the   same   plate   and   shaken   gently   to   allow   mixing   of   cell   suspension   with   samples.     Control   wells   contained    assay  medium  and  AML-­‐193  cells    (50µl  +  50µl)  and  blank  wells   contained   only   assay   medium   (100µl)   as   previously   described   (Section   3.8.5.2).       Cells   were   exposed   to   different   concentrations   of   test   proteins   in   CO2   incubator   (5%   CO2,   37°C)   and   then   20µl   of   MTS   (Celltiter   96   aqueous   one   solution   cell   proliferation   assay   from   Promega)   was   added   to   each   well   to   give  a  colour  change  that  determines  the  number  of  viable  cells  in  the  assay.     Readings   were   taken   every   40   minutes   for   2   hrs   of   incubation   (5%   CO2,   37°C).   The   plates   were   finally   read   at   490nm   using   a   BioTek   FLX800   plate   reader   and   Gen5   software   as   previously   described   (Section   3.8.5.2).   The   results   from   different   concentrations   of   test   protein   were   subtracted   from   control  wells  (AML-­‐193  cells  only)  obtained  at  490nm.  Each  experiment  was   repeated  at  least  in  triplicate.       138   6.3.1.1 In   v itro   B iological   A ctivity   o f   G CSF2NAT   a nd   I ts   C ontrol   The   results   indicate   that   rhGCSF,   purified   GCSF2NAT   and   GCSF2QAT   tandems   can   stimulate   the   proliferation   of   the   AML-­‐193   cell   line.   Both   purified   GCSF2NAT   and   GCSF2QAT   tandems   show   a   significant   biological   activity   with   both   standard   curves   shifted   to   the   left   in   comparison   to   rhGCSF.   Besides   this,   both   tandems   show   a   greater   maximal   stimulation   of   AML193   cells   compared   to   rhGCSF   (Figure   6.1.A).   Based   on   the   findings   of   sigmoidal   dose   response   curves   (GraphPad   Prism),   the   values   of   EC50   (the   concentration   which   causes   50%   of   maximal   response)   for   each   protein   were  calculated  (Figure  6.1.B).         139     Figure  6-­‐1:  Proliferation  of  AML-­‐193  cells  in  the  presence  of  tandems   and  rhGCSF   (A)   AML-­‐193   cells   were   stimulated   with   GCSF2NAT,   GCSF2QAT   tandems   and   rhGCSF.   The   absorbance   values,   which   are   indicative   of   the   number   of   live   cells   present,   were   plotted   against   the   natural   logarithm   of   the   GCSF   concentrations   (rhGCSF   and   GCSF   tandems).   Sigmoidal   dose-­‐response   fit   (for   variable   slope)   was   used   for   the   data   analyses.   (B)   EC50   values   calculated   for   GCSF2NAT,   GCSF2QAT   tandem   &   rhGCSF   using   GraphPad   Prism   software.   Significance   between   EC50   values   of   GCSF   tandems   and   rhGCSF   was   performed   with   Graphpad   prism   using   Mann-­‐Whitney   test.   Results   are   given   as   standard   mean   of   error   (SEM)   for   triplicate   wells   in   graph  A  and  triplicate  EC50  in  graph  B.  (*  =  p  value  of  <0.05)         140   6.3.1.2 In   v itro   B iological   A ctivity   o f   G CSF4NAT   a nd   I ts   C ontrol   The   results   below   show   that   rhGCSF,   GCSF4NAT   and   GCSF4QAT   can   stimulate   the   proliferation   of   the   AML-­‐193   cell   line.   Both   tandems   show   significant  increased  bioactivity  with  both  standard  curves  shifted  to  the  left   in   comparison   to   rhGCSF   (Figure   6.2.A).       Based   on   the   findings   of   sigmoidal   dose   response   curves   (GraphPad  Prism),   the   values   of   EC50   were   shown   in   the  Figure  6.2.B.         141     Figure  6-­‐2:  Proliferation  of  AML-­‐193  cells  in  the  presence  of  tandems   and  rhGCSF   (A)   AML-­‐193   cells   were   stimulated   with   GCSF4NAT,   GCSF4QAT   tandems   and   rhGCSF.   (B)   Shows   EC50   values   calculated   for   GCSF4NAT,   GCSF4QAT   tandem   &   rhGCSF   Graph   using   Pad   Prism   software.   Significance   between   EC50   values   of   GCSF   tandems   and   rhGCSF   was   performed   with   GraphPad   Prism  using  Mann-­‐Whitney  test.  Results  are  given  as  SEM  for  triplicate  wells   in  graph  A  and  triplicate  EC50  in  graph  B.  (*  =  p  value  of  <0.05)         142   6.3.1.3 In   v itro   B iological   A ctivity   o f   G CSF8NAT   a nd   I ts   C ontrol   The   bioactivity   results   for   GCSF8NAT   &   GCSF8QAT   show   similar   results   to   previous   tandem   proteins   and   together   with   rhGCSF   can   stimulate   the   proliferation   of   the   AML-­‐193   cell   line.   Both   tandems   show   significant   increased  biological  activity  with  both  standard  curves  shifted  to  the  left  in   comparison  to  rhGCSF.  Both  tandems  also  show  a  greater  maximal  level  of   activity  than  rhGCSF  (Figure  6.3.A).  Based  on  the  findings  of  sigmoidal  dose   response   curves   (GraphPad   Prism),   the   values   of   EC50   were   calculated   (Figure  6.3.B).       143     Figure  6-­‐3:  Proliferation  of  AML-­‐193  cells  in  the  presence  of  tandems   and  rhGCSF   (A)   AML-­‐193   cells   were   stimulated   with   GCSF8NAT,   GCSF8QAT   tandems   and   rhGCSF.   (B)   Shows   EC50   values   calculated   for   GCSF4NAT,   GCSF4QAT   tandem   &   rhGCSF   using   GraphPad   Prism   software.   Significance   between   EC50   values   of   GCSF   tandems   and   rhGCSF   was   performed   with   GraphPad   Prism  using  Mann-­‐Whitney  test.  Results  are  given  as  SEM  for  triplicate  wells   in  graph  A  and  triplicate  EC50  in  graph  B.  (*  =  p  value  of  <0.05)         144   6.3.2 Short  Term  Stability  of  GCSF  Tandem  Molecules     The   protein   stability   was   assessed   by   two   different   ways.   First   by   testing   samples   from   the   stability   experiment   in   the   AML-­‐193   assay.   Second   by   testing  purified  samples  from  stock  at  3  different  temperatures  (4°C,  room   temperature  (RmT)  and  -­‐80°C  freeze  thaw  (F/T)  cycles)  over  8  days  period.     6.3.2.1 Protein   S amples   f rom   t he   S tability   E xperiment   i n   t he   AML-­‐193   A ssay     During  the  AML-­‐193  stability  experiment,  5μg  of  each  tandem  protein  were   incubated   with   AML-­‐193   cells   at   37°C   for   3   days.   Protein   samples   were   taken   and   analysed   by   10%   SDS-­‐PAGE   gel   and   protein   bands   visualised   by   western  blotting  (Figure  6.4).  This  step  is  important  to  determine  whether   GCSF  protein  tandems  are  degraded  or  not  during  incubation.       Figure  6-­‐4:  Stability  of  GCSF  tandems  after  incubation  with  AML-­‐193   cells   Western   blot   with   anti   GCSF   antibodies   show   no   visible   degradation   in   all   GCSF  tandems  (Lane  1;  GCSF2QAT,  Lane  2;  GCSF2NAT,  Lane  3;  GCSF4QAT,   Lane  4;  GCSF4NAT,  Lane  6;  GCSF8QAT,  Lane  7;  GCSF8NAT).  A  total  of  100ng   protein  was  loaded  per  lane.     145   6.3.2.2 Temperature   S tability   o f   G CSF   T andem   M olecules   Stock   purified   tandem   proteins   were   taken   from   -­‐80°C   and   diluted   to   0.8mg/ml   with   filter   sterile   PBS   and   kept   on   ice.   All   manipulations   were   carried   out   under   sterile   conditions.   Aliquots   of   protein   at   0.8mg/ml   were   placed  at  room  temperature  (RmT),  4°C  or  -­‐80°C  in  1.5ml  sterile  eppendorf   tubes.    Samples  were  taken  on  days  0,1,  4  and  8  (samples  taken  on  day  zero,   represent   untreated   sample   controls)   and   immediately   diluted   with   an   equal   volume   of   SDS-­‐PAGE   buffer   (Laemmli   buffer)   and   heated   at   95°C   for   5   minutes   to   denature.   Samples   were   analysed   by   10%   SDS-­‐PAGE   gel   and   protein  bands  visualised  by  coomassie  staining  and  western  blotting.  A  total   of  7.5μg  protein  was  loaded  per  lane  for  SDS-­‐PAGE  gel  and  100ng  per  lane   for  immunoblot.  Samples  were  analysed  by  SDS-­‐PAGE  gels  for  are  shown  in   Figures   6.5   to   6.10.     The   stability   western   blot   analysis   for   each   tandem   protein  is  also  shown  adjacent  to  the  respective  SDS-­‐PAGE  gel. 146   Figure  6-­‐5:  Temperature  stability  of  GCSF2NAT   Coomassie   blue   and   western   blot   with   anti   GCSF   antibodies   show   GCSF2NAT  under  different  temperatures  stability  conditions.  In  each  gel  the   lanes  contain  samples  taken  at  D0,  D1,  D4  and  D8  stored  at  RmT,  4°C,  and  -­‐ 80°C   thawed/refrozen   in   each   day.   No   visible   degradation   in   protein   tandems   in   any   sample   under   all   conditions   studied   (4°C,   RmT   and   -­‐80°C   freeze  thaw)  over  the  8  days  period. 147   Figure  6-­‐6:  Temperature  stability  of  GCSF2QAT   Coomassie  blue  and  western  blot  with  anti  GCSF  antibodies  show  GCSF2QAT   under   different   temperatures   stability   conditions.   In   each   gel   the   lanes   contain  samples  taken  at  D0,  D1,  D4  and  D8  stored  at  RmT,  4°C,  and  -­‐80°C   thawed/refrozen  in  each  day.  No  visible  degradation  in  protein  tandems  in   any  sample  under  all  conditions  studied  (4°C,  RmT  and  -­‐80°C  freeze  thaw)   over  the  8  days  period.       148   Figure  6-­‐7:  Temperature  stability  of  GCSF4NAT   Coomassie   blue   and   western   blot   with   anti   GCSF   antibodies   show   GCSF4NAT  under  different  temperatures  stability  conditions.  In  each  gel  the   lanes  contain  samples  taken  at  D0,  D1,  D4  and  D8  stored  at  RmT,  4°C,  and  -­‐ 80°C   thawed/refrozen   in   each   day.   No   visible   degradation   in   protein   tandems   in   any   sample   under   all   conditions   studied   (4°C,   RmT   and   -­‐80°C   freeze  thaw)  over  the  8  days  period. 149   Figure  6-­‐8:  Temperature  stability  of  GCSF4QAT Coomassie  blue  and  western  blot  with  anti  GCSF  antibodies  show  GCSF4QAT   under   different   temperatures   stability   conditions.   In   each   gel   the   lanes   contain  samples  taken  at  D0,  D1,  D4  and  D8  stored  at  RmT,  4°C,  and  -­‐80°C   thawed/refrozen  in  each  day.  No  visible  degradation  in  protein  tandems  in   any  sample  under  all  conditions  studied  (4°C,  RmT  and  -­‐80°C  freeze  thaw)   over   the   8   days   period.   Both   arrows   show   around   20%   potential   dimer   formation  as  judge  by  SDS-­‐PAGE  and  western  blot.         150   Figure  6-­‐9:  Temperature  stability  of  GCSF8NAT   Coomassie   blue   and   western   blot   with   anti   GCSF   antibodies   show   GCSF8NAT  under  different  temperatures  stability  conditions.  In  each  gel  the   lanes  contain  samples  taken  at  D0,  D1,  D4  and  D8  stored  at  RmT,  4°C,  and  -­‐ 80°C   thawed/refrozen   in   each   day.   No   visible   degradation   in   protein   tandems   in   any   sample   under   all   conditions   studied   (4°C,   RmT   and   -­‐80°C   freeze  thaw)  over  the  8  days  period.       151   Figure  6-­‐10:  Temperature  stability  of  GCSF8QAT   Coomassie  blue  and  western  blot  with  anti  GCSF  antibodies  show  GCSF8QAT   under   different   temperatures   stability   conditions.   In   each   gel   the   lanes   contain  samples  taken  at  D0,  D1,  D4  and  D8  stored  at  RmT,  4°C,  and  -­‐80°C   thawed/refrozen  in  each  day.  No  visible  degradation  in  protein  tandems  in   any  sample  under  all  conditions  studied  (4°C,  RmT  and  -­‐80°C  freeze  thaw)   over  the  8  days  period. In   summary,   all   samples   of   protein   tandems   show   no   visible   degradation   under   all   conditions   studied   (4°C,   RmT   and   -­‐80°C   freeze   thaw)   over   the   8   days   period,   which   implies   the   high   short   term   stability   of   these   tandems.   However,   GCSF4QAT   shows   around   20%   dimer   formation   upon   analysing   by  SDS-­‐PAGE  and  western  blot.  The  probable  reason  for  dimer  formation  in   one  of  the  constructs  (GCSF4QAT)  will  be  discussed  in  the  final  discussion.         152   6.4 Discussion In   this   chapter,   to   determine   the   suitability   of   the   purified   glycosylated   tandem  proteins  and  their  respective  non-­‐glycosylated  controls  for  the  PK  in   vivo   studies;   it   was   imperative   to   assess   their   biological   activities.   While   active  proteins  are  essential  for  use  in  the  PK  studies,  it  is  equally  important   that   the   proteins   be   stable   over   the   course   of   administration   in   an   animal   model   and   post-­‐administration   analyses   of   the   protein   samples.   Therefore,   AML-­‐193   cell   line   was   used   to   measure   the   bioactivity   of   these   tandem   proteins   due   to   the   ability   of   rhGCSF   to   proliferate   this   type   of   cell   line   in   vitro.     Furthermore,   stability   of   these   tandem   proteins   was   also   analysed   using   different   temperatures   (RmT,   4°C   and   -­‐80°C   F/T)   and   also   after   incubation  with  AML-­‐193  cells  at  37°C  for  3  days.     The   biological   activity   results   show   that   AML-­‐193   proliferation   assay   is   a   valid   quantitative   in  vitro   model   for   measuring   the   activity   of   glycosylated   GCSF   tandems   and   their   respective   non-­‐glycosylated   controls.   Also,   the   results   indicate   that   these   tandem   proteins   exhibit   agonistic   action   by   stimulating  the  proliferation  of  AML-­‐193  cells.  Initial  bioassay  experiments   showed  that  the  best  level  of  absorbance  achievable,  upon  the  induction  of   the   cells   to   MTS   reagent   at   37°C   for   2   hrs   is   around   0.25   at   490nm.     Extending   the   time   of   induction   beyond   2   hrs   resulted   in   no   significant   increase  in  the  colour  change.  It  is  apparent  from  the  results  that  the  tandem   proteins   exhibited   significant   increased   biological   activity   compared   to   the   rhGCSF  (EC50  for  tandems  were  about  3-­‐fold  lower  than  that  for  rhGCSF).     The   observed   increase   in   biological   activity   could   be   attributed   to   two   factors;  the  design  of  the  tandem  molecule  (two  GCSF  ligands  linked  with  a   flexible  linker)  and  also  the  linker  length.    Normally,  two  GCSF  ligands  bind   two   GCSF   receptors   forming   a   homodimer   (Larsen   et   al.,   1990).   The   binding   of   the   GCSF   ligand   to   the   receptor   occurs   at   a   2:2   ligand:receptor   subunit   stoichiometry,   forming   a   cross-­‐over   configuration   between   the   receptor   subunits.   This   structural   configuration   enables   the   trans-­‐phosphorylation   of   the  receptors  and  initiates  downstream  signal  transduction.  However,  a  lag   in  the  availability  or  binding  of  a  second  GCSF  ligand  to  the  receptor  would   153   result   in   delayed   activation   of   the   signaling   pathway.   This   potential   delay   may   be   absent   in   the   tandem   molecule,   as   there   are   two   covalently   linked   GCSF   molecules   in   tandem   that   are   readily   available   for   binding   to   the   receptors.     The   observed   potency   of   the   tandem   molecules   is   consistent   with   other   published  research  studies  where  two  ligands  linked  in  tandem  resulted  in  a   novel  molecule  with  higher  bioactivity  than  the  wild  type.  The  binding  of  the   EPO   ligand   to   the   receptor   occurs   at   a   2:2   ligand:receptor   subunit   stoichiometry   similar   to   that   in   GCSF.   Sytkowski   et   al.,   (1999)   generated   a   fusion   of   two   hematopoietic   growth   factor   erythropoietin   (Epo)   linked   in   tandem   using   a   17-­‐amino   acid   flexible   peptide   linker.   The   specific   biological   activity   of   the   Epo   tandem   at   1,007   IU/μg   was   found   to   be   significantly   greater   than   that   of   the   native   Epo   protein   at   352   IU/μg.   Furthermore,   the   comparative   studies   of   the   pharmacokinetics   of   the   Epo   tandem   and   the   native  Epo  proteins  in  mice  showed  the  tandem  molecule  to  be  more  potent,   causing   a   significant   increase   in   red   blood   cell   production   within   7   days   whereas   the   conventional   recombinant   Epo   control   had   no   obvious   stimulating  effect  (Sytkowski  et  al.,  1999).     Furthermore,   compared   to   the   fusion   constructs   in   the   aforementioned   research   studies,   in   our   GCSF   tandem   the   linker   length   is   a   significant   factor   for   consideration   in   the   design   of   a   fusion   expression   construct,   as   studies   have   shown   that   inappropriately   short   linkers   could   result   in   loss   of   biological   activities   or   potency   in   tandem   fusion   molecules.   Research   by   Qui   et  al.,  (1998),  showed  the  bioactivity  of  a  tandem  Epo  fusion  construct  with  a   shorter  linker  sequence  (glycine  (G)  3-­‐7  peptide  was  compared  with  the  Epo   tandem   with   a   longer   linker   sequence   17-­‐amino   acid   (A-­‐[G-­‐G-­‐G-­‐G-­‐S]3-­‐T) .The   results   revealed   that,   while   both   tandem   molecules   were   biologically   active   the   Epo   tandem   with   the   shorter   linker   showed   no   increase   in   bioactivity   over   the   native   Epo   whereas   the   Epo   tandem   with   the   longer   linker   was   significantly   more   biologically   active   (Qiu   et   al.,   1998).   This   suggests  that,  tandem  fusion  ligands  with  linker  length  long  enough  would   facilitate   simultaneous   binding   of   the   ligands   to   the   receptors,   which   in   154   effect  induces  the  homodimerization  and  activation  of  the  receptors,  which   in  turn  initiate  the  downstream  signal  transduction.    Whereas,  in  a  tandem   fusion   molecule   with   short   linker   sequence,   the   two   ligands   linker   could   potentially   restrict   the   binding   of   one   of   the   two   ligands   to   the   receptor,   which   inadvertently   would   exhibit   similar   activity   to   a   ‘free’   ligand   bound   receptor.   In   the   design   of   our   three   glycosylated   GCSF   tandems   and   their   controls,   linker   lengths   used   were   long   enough   and,   all   the   tandem   molecules   had   similar   biologically   activity.     This   suggests   that   the   linker   lengths   between   the   tandem   GCSF   ligands   in   the   six   expression   constructs   investigated   (49   amino   acids   for   GCSF2NAT/2QAT,   GCSF4NAT/4QAT   and   94-­‐amino   acids   for   GCSF8NAT/8QAT)   were   long   enough   to   allow   both   GCSF   ligands   in   the   tandem   molecule   to   bind   both   receptors.   This   equally   contributed   to   the   higher   biological   activities   observed   in   these   constructs   compared  to  the  rhGCSF.   The  data  also  indicate  that  the  biological  activity  of  glycosylated  tandems  is   very   similar   to   their   non-­‐glycosylated   controls.   For   instance,   the   tandem   molecule   with   2   glycosylation   sites   showed   similar   results   to   the   4   and   8   glycosylation   sites.     It   is   apparent   from   the   result   that   increasing   the   N-­‐ linked   glycosylation   sites   has   no   noticeable   negative   impact   on   the   in  vitro   biological  activity  of  GCSF.     The  GCSF  tandems  also  showed  a  high  degree  of  stability  in  vitro  during  the   temperature  stability  test.  The  results  (SDS-­‐PAGE  and  western  blot)  showed   no   visible   degradation   in   the   analysed   tandem   proteins   samples,   under   all   conditions  studied  (4°C,  RmT  and  -­‐80°C  freeze  thaw),  over  the  8  days  period   and   also   after   incubation   with   AML-­‐193   cells   at   37°C   for   3   days.   This   suggests   that   the   tandem   molecules   achieve   the   right   conformation   during   protein  synthesis  in  CHO  cells.     The   glycosylated   tandem   molecules   are   devoid   of   dimerisation,   whereas,   non-­‐glycosylated   tandems   showed   some   dimer   formation   that   was   more   obvious   in   GCSF4QAT   (a   control   for   GCSF4NAT   tandem   molecule).     GCSF4QAT   was   estimated   to   contain   approximately   20%   dimer   when   155   analysed   by   Coomassie   blue   and   western   blotting   (Refer   to   Figure   6.8).   However,   all   the   controls   showed   similar   biological   activity   in   the   proliferation  bioassay.  This  indicated  that  the  presence  of  a  dimer  does  not   affect  the  overall  bioactivity  of  the  GCSF4QAT  control  molecule.     Routinely,   proteins   are   stored   as   frozen   solutions   at   -­‐80°C   and   this   may   preserve  the  stability  of  proteins  for  longer  periods  of  time.  However,  it  was   observed   that   long   time   freezing   at   -­‐80°C   may   be   the   cause   of   dimer   formation   as   found   in   GCSF4QAT,   which   was   stored   at   -­‐80°C   for   about   a   year.   Prior   to   freezing   (i.e.   during   expression   and   purification)   no   dimerisation   was   observed   in   the   GCSF4QAT   just   like   the   other   test   molecules.     Two   factors   may   have   contributed   to   the   observed   dimerisation   of   GCSF4QAT   after   longer   storage   at   -­‐80°C.   First,   the   dimer   formation   could   be   attributed   to   the   presence   of   free   cysteine   residue   (Cys17)   in   the   GCSF   structure  coupled  with  the  absence  of  glycosylation  in  the  molecule.  Native   GCSF   contains   five   cysteine   residues,   two   internal   disulfide   bonds   at   positions   Cys36–   Cys42   and   Cys64–   Cys74   leaving   one   free   cysteine   residue   at   position   Cys17   with   a   free   sulfhydryl   group.   It   is   possible   that   during   longer  storage  in  -­‐80°C,  the  free  cysteine  (Cys17)  of  one  GCSF  tandem  may   form   inter-­‐molecular   disulfide   bonds   with   another   GCSF   tandem,   which   results  in  the  formation  of  GCSF  tandem  dimer.  This  view  is  supported  by  a   recent   study   where   the   Cys17   in   the   GCSF   molecule   was   substituted   with   Ala17   (alanine   instead   of   cysteine   in   wild-­‐type   GCSF)   using   direct   mutagenesis  and  recombinant  DNA  technology.  The  resulting  GCSF  mutant   exhibited  enhanced  in  vitro  stability  and  higher  activity  in  vivo  than  the  wild-­‐ type   GCSF,   perhaps   through   the   elimination   of   dimerisation   caused   by   the   formation  of  intermolecular  disulfide  bonds  (Jiang  et  al.,  2011).  While  other   GCSF   tandems   also   contain   GCSF   units   with   free   cysteine   (Cys17),   the   glycosylation   of   the   polypeptide   chains   may   have   induced   unique   conformational   changes   that   prevent   intermolecular   disulphide   bond   formation  with  another  tandem  or  with  GCSF  in  the  same  tandem.   156   Another   possible   reason   of   dimerisation   is   the   incorporation   of   glutamine   (Q)   into   the   linker   of   the   non-­‐glycosylated   control   GCSF   tandem   (i.e.   GCSF4QAT)  compared  to  asparagine  (N)  in  the  glycosylated  GCSF  tandems.   It   has   been   shown   that   incorporation   of   a   poly-­‐glutamine   into   a   protein   resulted  in  the  formation  of  dimers.  The  glutamine  repeats  in  proteins  form   hydrogen   bonds   as   a   polar   zipper   (two   paired   antiparallel   β-­‐sheet   strands   held   together   by   hydrogen   bonds   between   their   amide   groups)   (Perutz,   1995,  Hoffner  and  Djian,  2005).  However,  since  the  control  tandem  molecule   has   a   single   point   mutation   that   allows   for   glutamine   incorporation,   the   involvement  of  the  incorporated  glutamine  in  the  dimer  formation  is  rather   ambiguous.  Perhaps,  analysing  the  control  tandem  protein  on  a  reduced  gel   (i.e.  containing  DTT)  would  elucidate  the  type  of  bond  that  is  present  in  the   dimers,   whether   the   dimers   were   covalently   linked   or   not   through   disulphide  bridges.     Moreover,  the  control  tandem  has  no  N-­‐glycosylation  sites  at  the  linker  and   therefore   would   show   less   steric   hindrance   towards   the   formation   of   a   dimer   when   compared   to   other   glycosylated   tandems.   A   number   of   published   research   studies   have   highlighted   that   glycosylated   molecules   have   molecular   characteristics   that   significantly   affect   the   function   of   the   molecule.  For  instance,  it  was  recently  reported  that  N-­‐linked  glycosylation   modulates   dimerisation   of   human   protein   disulfide   isomerase   (PDIA2)   (important   enzyme   for   the   correct   maturation   and   folding   of   proteins   that   reside   or   transit   into   the   endoplasmic   reticulum   (ER)).     This   protein   was   shown   to   be   glycosylated   at   the   asparagine   residues   of   three   N-­‐linked   glycosylation  sites  (N127,  N284  and  N516).  The  finding  was  that  mutation  at   N284   led   to   an   increase   in   the   dimer   formation   of   PDIA2   (Walker   et   al.,   2013).   This   suggests   that,   the   site   directed   mutagenesis   abrogated   the   glycosylation  characteristics  in  the  mutant  molecule.   Glycosylation   in   the   tandem   molecules   possibly   creates   steric   hindrance.   The   presence   of   charge   repulsion   among   molecules   due   to   the   negative   charge   and   a   huge   relative   volume   occupied   by   sialic   acid   as   described   previously  in  the  introduction  chapter.   However,  the  biological  activity  data   157   of   non-­‐glycosylated   GCSF   tandems   were   comparable   to   that   observed   for   glycosylated  GCSF  tandems,  indicating  that  dimerisation  did  not  affect  the  in   vitro  biological  activity.     In   conclusion,   The   GCSF   tandems   demonstrated   better   biological   activity   compared   to   the   rhGCSF   and   showed   high   stability   at   4C,   RmT,   37°C   and   multiple  F/T  cycles  at  -­‐80°C.       158   7. Results  4:  Pharmacokinetic  &  Pharmacodynamic   Analysis  of  rhGCSF  and  GCSF  Tandems  in  a  Rat   Model   7.1 Summary   Glycoproteins   represent   a   major   value   for   the   next   marketed   and   clinical   generation  of  therapeutic  proteins.  A  full  understanding  of  the  function  and   nature   of   the   glycosylation   and   its   influence   on   pharmacology   properties   is   crucial   in   finding   and   developing   efficient   and   safe   glycoprotein   biopharmaceuticals.   In   the   previous   chapter,   glycosylated   GCSF   tandems   together   with   non-­‐glycosylated   controls   have   shown   better   bioactivity   compared   to   rhGCSF   and   high   stability.   In   vivo   pharmacokinetic   and   pharmacodynamics   properties   of   these   GCSF   tandem   proteins   were   measured   in   normal   Sprague   Dawley   strain   rats   with   full   ethical   approval.   GCSF2NAT,  GCSF4NAT  and  GCSF8NAT,  containing  2,  4  &  8  glycosylation  sites   respectively,  displayed  a  reduced  rate  of  clearance  compared  to  both  rhGCSF   and   non-­‐glycosylated   GCSF   tandem   controls.   Although   the   half-­‐life   of   GCSF8NAT   exhibited   no   further   enhancement   beyond   that   of   GCSF4NAT,   pharmacodynamics   (PD)   displayed   a   significant   increase   in   the   number   of   circulating   neutrophils   at   48   hrs   in   rats   compared   to   12   hrs   reported   for   GCSF4NAT   and   GCSF2NAT,   leading   us   to   hypothesize   that   GCSF8NAT   is   a   more  efficient  stimulator  of  neutrophils.         159   7.2 Introduction   Pharmacokinetics   (PK)   is   described   as   the   study   that   evaluates   the   time   course   of   drug   absorption,   distribution,   metabolism   and   excretion   in   living   organisms  such  as  rats.  It  is  sometimes  described  as  what  the  body  does  to  a   drug.  In  contrast,  pharmacodynamics  (PD)  is  described  as  what  a  drug  does   to   the   body,   for   example,   in   our   case   what   GCSF   does   to   neutrophil   population   in   the   organism.   Protein   drugs’   PK/PD   are   typically   affected   by   fast   elimination   in   intravenous   administration   from   the   human   body,   via   proteolytic,   renal,   hepatic,   and   receptor   mediated   clearance   mechanisms   (Tang   et   al.,   2004,   Mahmood   and   Green,   2005).   Thus,   PK/PD   data   that   are   produced  from  relevant  species  like  mouse  or  rat  support  the  prediction  of   PK/PD   in   humans   and   may   help   to   generate   safe   and   effective   therapeutic   applications  for  current  therapies.     In  humans,  GCSF  has  a  short  serum  in  vivo  half-­‐life  of  1.79  hrs  (Tanaka  et  al.,   1991).   Consequently,   individual   patients   with   neutropenia   require   daily   injections   to   increase   neutrophils   in   the   blood   circulation,   which   leads   to   poor  patient  compliance.  However,  the  current  product  on  the  market  that  is   a  long-­‐acting  form  of  GCSF  called  PEG-­‐rhGCSF  (with  terminal  half-­‐life  of  7.05   hrs)   which   is   administered   once   per   chemotherapy   cycle   to   enhance   the   number   of   neutrophils   (Tanaka   et   al.,   1991).   It   has   been   reported   that   modifications   of   GCSF   by   covalently   attaching   a   chemical,   polyethylene   glycol   (PEG),   can   change   the   PK   and   PD   properties   of   GCSF   to   significantly   increase   the   time   the   modified   native   GCSF   remains   effective   in   the   blood   circulation   (Delgado   et   al.,   1992).   Results   from   receptor   binding,   in   vitro   proliferation   and   neutrophil   function   studies   show   that   PEG-­‐rhGCSF   and   native   GCSF   have   a   similar   mechanism   of   action   in   the   circulation   (Lord   et   al.,   2001).     In   rat   models,   rhGCSF   is   mainly   eliminated   by   the   renal   route;   therefore,   the   presence   of   PEG   moiety   increases   the   molecular   weight   of   GCSF  and  reduces  its  renal  clearance  by  glomerular  filtration  (Jain  and  Jain,   2008).     Efforts   have   been   made   to   improve   the   pharmacokinetic   behaviour   of   therapeutic   proteins   by   the   addition   of   natural   carbohydrates.   In   the   160   previous   chapter,   the   addition   of   natural   carbohydrates   (via   N-­‐linked   glycosylations)   to   the   GCSF   tandem   has   improved   the   bioactivity   and   stability.   It   has   also   been   reported   that   hyperglycosylation   can   regulate   activity   and   the   pharmacokinetic   profile   of   GCSF   (using   the   mutant   hGCSF   (Phe140Asn))   leading   to   prolonged   GCSF   in   the   circulation   and   a   more   effective   molecule   than   native   GCSF   in   stimulating   differentiation   and   proliferation  of  hematopoietic  cells  (Chung  et  al.,  2011).    Hyperglycosylation   closely  resembles  PEGylation  but  has  an  added  advantage  over  PEGyaltion.   For   instance,   hyperglycosylation   involves   the   biodegradable   nature   of   carbohydrates,   while   in   contrast   the   whole   PEG   molecule   is   often   processed   for   excretion   in   the   human   body   without   undergoing   an   initial   biodegradation   which   could   be   toxic   to   the   body   (Patel   et   al.,   2014).   This   possible  toxicity  of  PEGylation  was  supported  by  the  detection  of  PEG  only   in   bile   (Caliceti   and   Veronese,   2003).   Beside   these   advantages,   this   project   used   hyperglycosylation   (N-­‐linked   glycosylated   linker)   to   increase   the   molecular   weight   of   GCSF   and   may   also   protect   it   from   proteolysis   due   to   the   presence   of   terminal   sialic   acid   that   shields   the   underlying   galactose   peptides  from  protease  recognition  and  cleavage  within  the  circulation  (Sola   and  Griebenow,  2009).    This  is  supported  by  a  study  carried  out  by  Raju  and   Scallon  (2007)  who  demonstrated  that  removal  of  terminal  sugars  from  Fc   antibody  fragments  resulted  in  increased  sensitivity  to  papain.     In  this  chapter,  the  construction  of  novel  recombinant  GCSF  tandems  results   in   molecules   with   reduced   clearance   while   retaining   bioactivity.   It   also   alleviates   potential   problems   with   direct   glycosylation   of   the   ligand,   which   may  reduce  bioactivity  and  potentially  introduce  immunogenic  sites.         161   7.3 Aim  and  Hypothesis   In   vivo   strength   of   therapeutic   proteins   is   often   strongly   associated   with   residence   time   of   blood   circulating.   This   is   a   function   of   the   drug’s   PK   behaviour   including,   serum   half-­‐life,   clearance   rate   and   the   minimum   and   maximum   concentrations.   In   the   previous   chapter,   glycosylated   GCSF   tandems   and   their   respective   controls   have   shown   better   bioactivity   in   comparison   to   rhGCSF   and   high   stability.     Thus,   the   aim   of   the   current   chapter   is   to   determine   the   pharmacokinetic   properties   of   rhGCSF   &   GCSF   tandems  in   normal   adult  Sprague  Dawley  strain  rats  following  intravenous   injections  and  the  effects  of  these  GCSF  tandems  on  white  blood  cell  (WBCs)   and   neutrophil   populations.   We   hypothesised   that   our   glycosylated   GCSF   tandems   will   have   longer   circulating   half-­‐lives   and   more   potent   (i.e.   mobilizing  more  neutrophils)  compared  to  the  rhGCSF.       162   7.4 Results   7.4.1 Preliminary  Test  for  the  Effect  of  Rat’s  Serum  on  Elisa   Assay   Prior   to   measuring   GCSF   tandems   in   rat   serum,   it   was   necessary   to   measure   the  sensitivity  of   the  Elisa  assay  by  testing  the  effect  of  rat  serum  on  rhGCSF   and   GCSF2NAT   standard   curves.   This   test   was   used   to   determine   the   specificity  of  the  Elisa  (i.e.  testing  if  Elisa  detects  rat  GCSF  from  serum)  and   to  assess  any  interference  of  rat  serum  with  the  sensitivity  of  the  Elisa.   Previous   studies   by   Asterion   on   human   growth   hormone   tandem   molecule   showed  that  2%  (v/v)  of  rat  serum  had  no  interference  with  the  sensitivity   of  the  growth  hormone  Elisa  standard  curve.  Hence,  this  concentration  of  rat   serum  (2%  v/v)  was  used  in  the  initial  trial  Elisa  experiments  to  determine   the  effect  of  rat  serum  on  rhGCSF  and  GCSF2NAT  Elisa’s.  The  results  of  the   experimental  studies  showed  that  2%  (v/v)  of  rat  serum  has  no  effect  on  the   Elisa   assay   using   rhGCSF   and   GCSF2NAT   as   standard   curve   controls.   Both   rhGCSF   and   GCSF2NAT   standard   curves   generated   in   2%   (v/v)   rat   serum   showed  no  OD  change  from  LKC  buffer  only  indicating  that  rat  serum  alone   has  no  interference  in  the  Elisa  at  this  concentration  (Figure  7.1).               163   rhGCSF  Elisa GCSF2NAT  Elisa B 4 4 3 3 Absorbance$$ Absorbance A 2 1 LKC buffer 2% Rat serum 0 -1.5 -1.0 -0.5 0.0 0.5 1.0 2 1 LKC buffer 2% Rat serum 0 -1.5 1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 Log$(nM) Log (nM) Figure  7-­‐1:  Effect  of  rat  serum  on  the  sensitivity  of  Elisa  assay   (A)   A   comparison       between   rhGCSF   diluted   in   2%   rat   serum   and   rhGCSF   diluted  in  LKC  buffer  only.  (B)  A  comparison  between  GCSF2NAT  diluted  in   2%  rat  serum  and  GCSF2NAT  diluted  in  LKC  buffer  only.  Results  are  given  as   SEM  for  triplicate  wells.     Sensitivity   of   the   assay   has   not   been   impaired   by   the   presence   of   rat   serum:   Therefore,   if   rat   serum   can   be   kept   to   2%   (v/v)   final   concentration   in   the   assay   then   it   will   be   acceptable   as   long   as   all   samples   receive   the   same   concentration  of  serum.                 164   7.4.2 Preliminary  Pharmacokinetic  Analysis  in  Normal   Sprague  Dawley  Rats   7.4.2.1   E lisa   R esults   The   pharmacokinetic   performance   of   glycosylated   GSCF   tandem   (GCSF2NAT,   4NAT,   8NAT   &   8QAT)   was   evaluated   in   Sprague   Dawley   rats   to   assess  the  longevity  of  exposure  in  comparison  to  rhGCSF.  250μg/kg  (75μg   per   rat)   of   rhGCSF,   GCSF2NAT,   GCSF4NAT,   GCSF8NAT,   GCSF8QAT   and   vehicle   (PBS   only)   were   given   by   intravenous   injection   (IV).   0.3   to   0.4ml   blood  sample  were  taken  at  specified  time  points  (-­‐24  pre-­‐injection,  0.5,  1,  2,   4,   8,   12,   24   &   48,   72   hrs)   and   centrifuged   for   serum   preparation   as   previously  described  in  section  3.9.                                                GCSF  Elisa  was  used  to  measure  the  concentration  of  proteins  in  each  serum   GCSF  tandems  samples  and  rhGCSF  at  -­‐24,  0.5,  1,  2,  4,  8,  12,  24,  48,  72  hrs   post-­‐dose.   The   data   analyses   of   these   samples   suggested   that   the   pharmacokinetic   profiles   for   GCSF   tandems   (GCSF2NAT,   4NAT,   8NAT   &   8QAT)   following   intravenous   dosing   at   250μg/kg   displayed   a   reduced   rate   of   clearance   compared   to   the   published   rhGCSF   (Figures   7.2   –   7.7).   However,   the   rhGCSF   molecule   in   this   study   appeared   to   be   mistakenly   injected  subcutaneously  as  against  intravenous  injection  of  other  test  GCSF   tandems.   This   however   was   not   confirmed   by   the   contractor   (more   details   provided  in  the  sections  below).  For  this  reason,  the  rhGCSF  samples  were   excluded   from   further   data   analyses,   so   as   to   avoid   inconsistencies   in   data   analyses.   165   7.4.2.1.1 Pharmacokinetics  Analysis  of  rhGCSF  in  Normal  Rats   Analysis   of   the   pattern   of   clearance   between   the   rats   shows   an   apparent   error   during   injection.   Rat   2   and   5   both   have   high   values   at   0.5   hr,   however,   Rat   5   again   show   an   unexpectedly   high   value   at   2hrs.   Rat   4   and   6   both   have   unexpectedly  low  values  at  all  time  points.  Rat  1  and  3  both  show  peaks  at   2hrs.     This   suggests   that   only   Rat   2   was   injected   intravenously   during   the   administration   while   other   rats   were   injected   subcutaneously   by   the   contractor.   Since   the   test   GCSF   tandems   were   injected   intravenously   these   data  were  excluded  in  this  research  to  avoid  inconsistency  in  data  analysis.   A  few  of  the  serum  samples  were  missing  as  the  contractor  who  did  the  in   vivo  studies  omitted  these  samples  (shaded  boxes  in  grey)  (Figure  7.2).                         166   A$ ! Time$(hr)$ 324$ 0.5$ 1$ 2$ 4$ 8$ 12$ 24$ 48$ 72$ Rat$ 1$ 0! 1.9! 4.6! 23.8! 0.7! 9.2! 1.4! 0! 0! 0! B$ 2$ 3.1! 2.3! 3.6! 3$ !! 21.7! 4! 4.1! 1.8! 5.9! !! 5.2! 2.6! !! !! 0! 0! 0! 0! 0! 0! 4$ 0! 2.5! 25.4! 45.8! 11.3! 4.5! 1.4! 0! 0! 0! 5$ 0! 0.6! 8.1! 5! 5.1! 3.1! 1.9! 6$ 2.6! 23.8! 7.1! 24.1! 8.3! !! 1.2! 0! 5! 0! !! 0! 0! 100 Rat Rat Rat Rat Rat Rat Log (nM) 10 1 2 3 4 5 6 1 0.1 0 5 10 15 Time (hours)   Figure  7-­‐2:  Elisa  analysis  of  rhGCSF  pharmacokinetics  in  normal  rat   (A)   Six   rats   were   injected   with   250μg/kg   of   native   purified   recombinant   GCSF   protein   intravenously.   Serum   samples   were   taken   24   hrs   before   the   injection   and   0.5,   1,   2,   4,   8,   12,   24,   48,   72   hrs   post   injection.     The   samples   were  analysed  by  ELISA  to  determine  the  concentration  (nM)  of  rhGCSF  in   the   serum   and   clearance   rate.   The   results   for   each   rat   at   each   individual   sampling  time  points  are  tabulated.  Missing  samples  shaded  in  grey.  (B)  The   concentrations  of  rhGCSF  were  plotted  against  the  time  of  sampling  for  each   individual  rat.         167   7.4.2.1.2 Pharmacokinetic  Analysis  of  GCSF2NAT  in  Normal  Rats   All   rats   in   this   group   exhibited   similar   patterns   of   protein   clearance   with   similar  an  early  peak  concentration  of  0.5  hrs,  which  suggests  that  the  rats   were  successfully  injected  intravenously.  Rat  6  serum  samples  and  a  few  of   other   serum   samples   were   missing   (shaded   boxes   in   grey),   as   these   were   not  received  from  the  contractor  who  did  the  IV  injection  and  sampling.  Two   samples   were   labelled   with   similar   information   (shaded   boxes   in   red)   and   were  therefore  omitted  from  the  study  (Figure  7.3).       168   A$ !! Time$(hr)$ 324$ 0.5$ 1$ 4$ !! 8$ 12$ 24$ 48$ 72$ Rat$ 1$ 2! 90.7! 45.5! 2$ 2.8! 42.1! 31.3! 12.4! 4.6! 6.8! 1.6! 0! 0! 12.7! 8.5! 3.9! 8.4! 0! 3$ 0! 20.5! 15.1! 10! 7.1! 6.2! 2! 0! 0! 4$ 0.7! 49! !! 5.8! 6.9! 5.7! 1.6! 1.9! 0! 5$ 0! 40! 11.9! 9.2! 8.2! 2.4! 0! 0! 0! 6$ !! !! !! !! !! !! !! !! !! B$ 100 Log (nM) Rat Rat Rat Rat Rat 1 2 3 4 5 10 1 0 10 20 30 Time (hours)                                           Figure  7-­‐3:  Elisa  analysis  of  GCSF2NAT  pharmacokinetics  in  normal  rat   (A)   Six   rats   were   injected   with   250μg/kg   of   purified   GCSF2NAT   intravenously.   Serum   samples   were   taken   24   hrs   before   the   injection   and   0.5,  1,  2,  4,  8,  12,  24,  48,  72  hrs  post  injection.    The  samples  were  analysed   by  an  Elisa  to  determine  the  concentration  (nM)  of  GCSF2NAT  in  the  serum   and   clearance.   The   results   for   each   rat   at   each   individual   sampling   time   point   were   tabulated.   Missing   samples   shaded   in   grey.   Omitted   samples   shaded  in  red.   (B)   The  concentration  of  GCSF2NAT  was  plotted  against  the   time  of  sampling,  for  each  individual  rat.       169   7.4.2.1.3 Pharmacokinetic  Analysis  of  GCSF4NAT  in  Normal  Rats   Analysis  of  the  pattern  of  clearance  between  the  rats  shows  that  rats  were   successfully   injected   intravenously.   Rat   6   serum   samples   and   a   few   of   the   other   samples   were   missing   (shaded   boxes   in   grey),   as   these   were   not   received  from  the  contractor  who  did  the  IV  injection  and  sampling  (Figure   7.4).   A$ !! Time$(hr)$ 324$ !! 0.5$ 1$ 2$ 4$ 8$ 12$ !! 24$ 48$ 72$ Rat$ 1$ 2$ 0! 63.1! 19.3! 14.6! 14.5! 81.6! 53! 36.7! 20.6! 12.3! !! 4.5! 1.7! 1.7! 1.8! 5.8! 0! 1.4! 3$ 0! !! 8.9! !! 7.8! 11! 8.2! 1.9! 0! 0! 4$ 3.3! 4.7! 14.6! 21.2! 11.6! 14.5! 8.2! 12! 1.9! 0.7! 5$ 25.8! 61.6! 75.9! 40.5! 35.6! 45.7! 29.8! 13.3! 7.8! 22.8! 6$ !! !! !! !! !! !! !! !! !! !! B$ 100 Rat Rat Rat Rat Rat 1 2 3 4 5 Log (nM) 10 1 0.1 0 20 40 60 80 Time (hours) Figure  7-­‐4:  Elisa  analysis  of  GCSF4NAT  pharmacokinetics  in  normal  rat (A)   Six   rats   were   injected   with   250μg/kg   of   purified   GCSF4NAT   intravenously.   Serum   samples   were   taken   24   hrs   before   the   injection   and   0.5,  1,  2,  4,  8,  12,  24,  48,  72  hrs  post  injection.    The  samples  were  analysed   by  Elisa  to  determine  the  concentration  (nM)  of  GCSF4NAT  in  the  serum  and   clearance.   Missing   samples   shaded   in   grey   (B).   The   results   for   each   rat   at   each   individual   sampling   time   point   were   tabulated.   The   concentration   of   GCSF4NAT  was  plotted  against  the  time  of  sampling,  for  each  individual  rat.                                                 170   7.4.2.1.4 Pharmacokinetics  Analysis  of  GCSF8NAT  in  Normal  Rats                                                Analysis  of  the  pattern  of  clearance  between  the  rats  shows  that  rats  were   successfully  injected  intravenously.  Rat  6  serum  samples  and  a  few  of  other   serum   samples   were   missing   (shaded   boxes   in   grey),   as   these   were   not   received   from   the   contractor   who   did   the   IV   injection   and   sampling.   In   addition,  Rat  1  showed  a  very  high  protein  concentration  in  all  time  points   (shaded  boxes  in  red)  compared  to  other  rats.  Rat  1  was  repeated  but  still   give   a   very   high   value.   Therefore,   it   was   decided   to   omit   from   the   current   study   after   discussions   with   Phoenix   Certara   (half-­‐life   =   270.7   hrs   as   analysed   by   Winnonlin   PK   program   provided   by   Phoenix   Certara)   (Figure   7.5).       171   A$ Rat$ !! Time$(hr)$ 524$ 0.5$ 1$ 2$ 4$ 8$ 12$ 24$ 48$ 72$ 1$ 22.6! 2$ 0! 86! 19.9! 17.5! 14.8! 13.1! 4.5! 4.3! 0.1! 1.3! !! 170.1! !! 16.4! !! 17.2! 9.1! !! 21.8! 3$ 2.1! 17.7! 11.1! 8.5! 7.7! 11.3! 5.2! 1.3! 0! 0! 4$ 5.7! 51.3! 37.6! 37.3! 27.6! 19.2! 13! 1.1! 0! 0! 5$ 0! 31.4! 18.1! 35.5! 18.6! 12.1! 5.1! 3.1! 0! 0! 6$ !! !! !! !! !! !! !! !! !! !! Rat$repeated$ 1$ 8.4! !! 144.8! !! 9.9! !! 15.8! 20.7! !! 16.8! B$ 100 Rat Rat Rat Rat Log (nM) 10 2 3 4 5 1 0.1 0.01 0 20 40 Time (hours) 60   Figure  7-­‐5:  Elisa  analysis  of  GCSF8NAT  pharmacokinetics  in  normal  rat   (A)   Six   rats   were   injected   with   250μg/kg   of   purified   GCSF8NAT   intravenously.   Serum   samples   were   taken   24   hrs   before   the   injection   and   0.5,  1,  2,  4,  8,  12,  24,  48,  72  hrs  post  injection.    The  samples  were  analysed   by  Elisa  to  determine  the  concentration  (nM)  of  GCSF8NAT  in  the  serum  and   clearance.   The   results   for   each   rat   at   each   individual   sampling   time   point   were   tabulated.   Missing   samples   shaded   in   grey.   Omitted   samples   shaded   in   red.   (B)   The   concentration   of   GCSF8NAT   was   plotted   against   the   time   of   sampling,  for  each  individual  rat.         172   7.4.2.1.5 Pharmacokinetic  Analysis  of  GCSF8QAT  in  Normal  Rats     Analysis  of  the  pattern  of  clearance  between  the  rats  shows  that  rats  were   successfully   injected   intravenously.   A   few   of   the   serum   samples   were   missing   (shaded   boxes   in   grey),   as   these   were   not   received   from   the   contractor  who  did  the  IV  injection  and  sampling.  Additionally,  Rat  5  and  6   showed   a   very   low   protein   concentration   in   all   time   points   compared   to   other   rats.     Both   rats   were   repeated   but   they   still   give   a   very   low   protein   concentration   (shaded   boxes   in   red),   therefore,   rat   5   &   6   were   omitted   from   this  study  (Figure  7.6).     A$ !! Time$(hr)$ 524$ !! 0.5$ 1$ 2$ 4$ 8$ 12$ 24$ 48$ 72$ Rat$ 1$ 66! 48.1! 16.6! 7.7! 5.6! 0! 0.4! 0! 0! 2$ 0! 3$ 0! 74! 46.2! 11.9! !! !! 9.6! 9.5! 7.4! 2.7! 1.6! 0! 0! B$ 4$ 0! 9.4! !! 10.5! 11.1! 8.3! 2.5! 1! 0! 0.8! !! 10.4! 3.9! 1.1! 0! 0! 5$ 0! 2.3! 0.8! 2.7! 7.7! 3.3! 2.7! 0.5! 0! 0! 6$ #0.2! 0.6! 0.1! 1.3! #0.5! #1.3! !! #1.9! #2! #1.2! Rat$repeated$ 5$ 6$ 0! #0.2! 4.1! 4.7! 0.8! 0.1! 0.2! 4.1! 0.65! #0.5! 0.794! #1.3! 0! !! 0.5! #1.9! 0! #2! 0! #1.2! 100 Rat Rat Rat Rat Log (nM) 10 1 2 3 4 1 0.1 0 10 20 Time (hours) 30   Figure  7-­‐6:  Elisa  analysis  of  GCSF8QAT  pharmacokinetics  in  normal  rat.   (A)   Six   rats   were   injected   with   250μg/kg   of   purified   GCSF8QAT   intravenously.   Serum   samples   were   taken   24   hrs   before   the   injection   and   0.5,  1,  2,  4,  8,  12,  24,  48,  72  hrs  post  injection.    The  samples  were  analysed   by  Elisa  to  determine  the  concentration  (nM)  of  GCSF8QAT  in  the  serum  and   clearance.   The   results   for   each   rat   at   each   individual   sampling   time   point   were   tabulated.   Missing   samples   shaded   in   grey.   Omitted   samples   shaded   in   red.   (B)   The   concentration   of   GCSF8QAT   was   plotted   against   the   time   of   sampling,  for  each  individual  rat.     173   7.4.2.1.6 Pharmacokinetic  Analysis  of  GCSF  Tandems  in  Normal  Rats       Figure  7-­‐7:  Elisa  analysis  of  GCSF  tandems  in  normal  rat  models                                                  (A)  250μg/kg  of  GCSF2NAT,  GCSF4NAT,  GCSF8NAT,  and  GCSF8QAT  proteins   were   given   intravenously   to   Sprague   Dawley   rats.   Serum   samples   were   taken  at  specified  time  points  (-­‐24,  0.5,  1,  2,  4,  8,  12,  24  &  48,  72  hrs).  The   results   of   the   mean   concentrations   for   each   tandem   at   each   individual   sampling   time   point   in   normal   rats   were   tabulated.  (B)   The   average   protein   concentrations   of   each   GCSF   tandem   were   plotted   against   the   time   of   sampling,   for   each   individual   rat.   Data   are   given   as   standard   error   of   the   mean   (SEM)   for   at   least   four   rats   per   group.   Significance   between   Elisa   results  of  GCSF  tandems  was  performed  with  GraphPad  Prism  using  Nonlin   fit  test.       174   From  the  Elisa  data  of  GCSF  tandems,  it  can  be  seen  that  for  IV  the  maximum   serum   concentration   was   reached   at   the   earliest   sampling   time-­‐point   of   0.5-­‐   hr  for  all  samples.  However,  the  rate  of  decline  thereafter  was  much  slower   for   GCSF   tandems   (GCSF2NAT,   4NAT,   8NAT   &   8QAT).   Interestingly,   all   the   GCSF   tandems   displayed   a   reduced   rate   of   clearance   compared   to   the   published  rhGCSF  when  the  half-­‐life  was  calculated  using  Winnonlin  6.3  PK   program   developed   by   Phoenix   Certara   (more   details   about   half-­‐life   in   the   next  section).       7.4.2.2 Terminal   H alf-­‐life   A nalyses   o f   G CSF   T andems   b y   t he   N on-­‐ Compartmental   M ethod                                              Comparative   analyses   of   the   Elisa   data   for   the   glycosylated   GCSF   tandems   (GCSF2NAT,   4NAT   and   8NAT)   and   their   non-­‐glycosylated   control   (GCSF8QAT)   suggested   that   all   the   GCSF   tandem   molecules   have   similar   clearance   with   GCSF4NAT   exhibiting   a   slightly   higher   serum   concentration   at  24  hrs  of  injection.  However,  the  increase  was  not  significantly  different   from   other   GCSF   tandems.   In   order   to   determine   in   vivo   terminal   half-­‐life   of   the   GCSF   tandems   and   control,   the   pharmacokinetic   data   were   analysed   using  the  non-­‐compartmental  method  of  data  analysis.  This  involved  the  use   of   Winnonlin   6.3   PK   program   developed   by   Phoenix   Certara   to   determine   the  terminal  half-­‐life  of  the  tandem  GCSF  proteins  in  each  individual  rat  used   in   this   study.   The   results   obtained   from   the   analyses   are   shown   below   (Table  7.1  and  Figure  7.8).       175   Table  7-­‐1:  Tandem  GCSF  proteins  terminal  half-­‐life  analyses  by  non-­‐ compartmental  method  for  each  rat   The   terminal   half-­‐life   of   the   tandem   GCSF   molecules   were   determined   using   a   Winnonlin   6.3   PK   program   developed   by   Phoenix   Certara.   The   results   of   the   groups   of   rats   used   for   each   individual   tandem   molecule   analyses   (B1-­‐ B5:   2NAT,   C1-­‐C4:   4NAT,   D2-­‐D5:   8NAT,   E1-­‐E4:   8QAT).   Cmax:   concentration   maximum  (nM);  Tmax:  time  maximum  (hr);  AUC0-­‐24:  Area  Under  the  Curve   during   24   hrs;   No   points:   Terminal   half-­‐life   points   (at   least   3   points);   Lambda   z:   Slope   of   the   drop;   Half-­‐life   (hr);   the   time   required   for   the   protein   concentration  to  fall  to  half  its  initial  amount.     Construct) Rat) !! !! !GCSF2NAT! !! !! B1! B2! B3! B4! B5! C1! C2! C3! C4! D2! D3! D4! D5! E1! E2! E3! E4! !GCSF4NAT! !GCSF8NAT! !GCSF8QAT!   Cmax)(nM)) Tmax)(hr)) AUC0624) 90.7! 49! 40! 42.1! 20.5! 81.6! 63.1! 21.2! 75.9! 86! 17.7! 51.3! 35.5! 66! 9.6! 74! 11.1! 0.5! 0.5! 0.5! 0.5! 0.5! 0.5! 0.5! 2! 1! 0.5! 0.5! 0.5! 2! 0.5! 2! 0.5! 4! 377.2! 202.6! 125! 201.6! 161.7! 366.8! 195.6! 154.3! 756.3! 242.8! 147.6! 380! 246.1! 141.9! 108.5! 203.1! 120.3!   176   No)points) Lambda)z) 3! 3! 4! 5! 5! 3! 5! 3! 3! 6! 3! 3! 4! 3! 4! 4! 3! 0.07! 0.09! 0.13! 0.11! 0.08! 0.06! 0.11! 0.04! 0.075! 0.07! 0.13! 0.18! 0.09! 0.15! 0.09! 0.12! 0.12! Half)life) (hr)) 9.66! 7.32! 5.14! 6.05! 8.26! 11.81! 6.56! 15.99! 9.26! 9.69! 5.31! 3.75! 7.85! 4.54! 7.62! 6.01! 5.81!   A) Construct) Rat)Group) Cmax)(hrs))Tmax)(hrs)) AUC0724) Published* GCSF* GCSF2NAT* GCSF4NAT* GCSF8NAT* GCSF8QAT* No)points) Lambda)z) Half)life) (hrs)) SEM) /* /* /* /* /* /* 1.79* /* B* C* D* E* 48.5* 49.8* 46.6* 49.8* 0.5* 0.5* 0.5* 0.5* 192.3* 257.4* 254.8* 162.8* 3* 6* 5* 5* 0.09* 0.07* 0.1* 0.12* 7.38* 10.74* 6.74* 5.87* 0.89* 2.31* 1.52* 0.73* B) Half Life (hours) 15 2NAT 4NAT 8NAT 8QAT 10 5 8Q A T T 8N A T A 4N 2N A T 0   Figure  7-­‐8:  Tandem  GCSF  proteins  terminal  half-­‐life  analyses  by  non-­‐ compartmental  method  for  each  rats’  group   (A)   The   terminal   half-­‐life   of   the   tandem   GCSF   molecules   were   determined   using   a   Winnonlin   6.3   PK   program   developed   by   Phoenix   Certara.   The   results   of   the   average   for   each   treatment   group   together   with   published   GCSF   were   tabulated.   (B)   The   average   terminal   half-­‐life   for   each   molecule   was  graphically  represented.  Data  are  given  as  standard  error  of  the  mean   (SEM)  for  at  least  four  rats  per  group.  Significance  between  terminal  half-­‐life   data  of  GCSF  tandems  was  performed  with  GraphPad  Prism  using  One-­‐Way   ANOVA.         177   7.4.2.3 Pharmacodynamics   o f   G CSF   T andems     Total   cell   blood   count   (CBC)   was   performed   on   selected   whole   blood   samples  (-­‐24,  12,  24,  48  and  72  hrs).  The  counts  of  white  blood  cells  (WBCs)   were   performed   using   a   coulter   counter   instrument.     Blood   smears   were   fixed   and   stained   using   Giemsa   stain   by   routine   laboratory   methods.   The   blood  neutrophil  count  was  performed  on  these  stained  slides  using  a  light   microscope  to  obtain  the  percentage  of  neutrophils.   Following   injection   with   GCSF2NAT,   4NAT   and   8NAT,   8QAT,   rhGCSF   and   vehicle  (PBS  only),  WBCs  count  showed  no  statistical  difference  in  number   of  WBCs  between  any  treatments  above  that  of  controls  (Vehicle,  rhGCSF  &   GCSF8QAT).   However,   WBC’s   peaked   at   24   hrs   post-­‐injection   for   GCSF8NAT   and  its  respective  control  GCSF8QAT  before  returning  to  baseline  values  at   72   hrs.   This   effect   was   not   significantly   different   from   the   rhGCSF   and   vehicle  control  (Figure  7.9E).     In   contrast,   all   glycosylated   GCSF   tandems   (GCSF2NAT,   4NAT   &   8NAT)   showed   increased   percentage   (%)   of   neutrophils   at   12   hrs   post   injection,   with  this  level  of  increase  being  higher  for  both  2NAT  and  4NAT  compared   to  controls  (Vehicle,  rhGCSF  &  GCSF8QAT).    However,  beyond  12  hrs  post-­‐ injection,   all   three   glycosylated   GCSF   tandems   showed   a   decline   in   their   neutrophil  levels,  which  are  more  pronounced  in  2NAT  and  4NAT,  up  to  72   hrs.    In  contrast,  GCSF8NAT  exhibited  a  higher  increase  in  neutrophil  levels   at  48  hrs  post  injection  following  a  marginal  decline  at  24  hrs,  compared  to   controls  which  may  imply  a  longer  duration  of  action  (Figure  7.9A-­‐D).     The   obvious   difference   in   the   number   of   neutrophils   mobilized   between   glycosylated  tandems  (at  12  hours  for  2NAT/4NAT  and  48  hours  for  8NAT)   is   significant   when   the   data   was   analysed   by   multiple   T-­‐test   (Figure   7.9).   However,   as   low   number   (~four)   of   rats   were   used   in   these   studies,   the   neutrophils   mobilization   differences   observed   were   not   significant.   Consequently,   using   more   rats   in   the   future   studies   is   required   to   assess   the   statistical   significance   of   the   number   of   neutrophils   mobilized   by   the   glycosylated  tandems  following  intravenous  injection  in  the  rats.   178       Figure   7-­‐9:   Percentage   change   in   blood   neutrophils   following   intravenous  administration  of  GCSF  tandems   (A)  GCSF2NAT  and  controls  (GCSF8QAT,  rhGCSF  &  vehicle),  (B)  GCSF4NAT   and   controls   (GCSF8QAT,   rhGCSF   &   vehicle),   (C)   GCSF8NAT   and   controls   (GCSF8QAT,   rhGCSF   &   vehicle),   (D)   GCSF2NAT,   GCSF4NAT   and   GCSF8NAT     &   (E)   showing   total   WBC   counts   following   intravenous   administration   of   GCSF   tandems   (GCSF2NAT,   4NAT,   8NAT   &   8QAT),   rhGCSF   and   vehicle   to   normal   rats.   Data   are   given   as   standard   error   of   the   mean   (SEM)   for   at   least   four  rats  per  group.  The  pharmacodynamics  studies  of  the  GCSF  tandems  at   these   selected   sampling   time   points   (-­‐24,   12,   24,   48   and   72   hrs)   were   analysed   with   GraphPad   Prism   using   multiple   T-­‐test.     Stars   indicate   neutrophil   values   that   are   significantly   different   between   rats   treated   with   GCSF  tandems,  rhGCSF  and  vehicle  (*  =  p  value  of  <0.05).   179   7.5 Discussion     In   the   previous   chapter,   the   GCSF   tandem   molecules   demonstrated   better   biological   activity   compared   to   the   rhGCSF   and   showed   high   stability   at   4°C,   RmT,  37°C  and  multiple  F/T  cycles.  Hence,  these  molecules  were  tested  for   the   pharmacokinetics   and   pharmacodynamics   (PK/PD)   properties   in   an   in   vivo   study.   The   pharmacokinetic   performance   of   selected   constructs   (GCSF2NAT,   4NAT,   8NAT   and   its   control   GCSF8QAT)   was   evaluated   in   normal   adult   Sprague   Dawley   rats.   Serum   samples   were   quantified   using   Elisa   technique   to   assess   the   delay   in   renal   clearance   (half-­‐life)   compared   to   the  reported  rhGCSF.  This  chapter  equally  evaluated  the  pharmacodynamics   (PD)   by   measuring   the   effect   of   these   GCSF   tandems   on   the   population   of   WBCs   and   neutrophils   since   GCSF   is   routinely   used   clinically   to   increase   the   number  of  neutrophils.   In   the   PK   analysis   the   detection   of   maximum   protein   levels   at   the   earliest   sampling  time-­‐point  of  0.5  hr,  and  decline  thereafter  for  the  rest  of  the  time   points,   verified   the   successful   IV   injection   of   rats   in   all   samples   except   rhGCSF.   The   results   for   rhGCSF   were   omitted   from   this   study   due   to   problems   with   the   mode   of   injection.   The   rats   should   have   been   injected   intravenously   through   the   neck   vein,   but   rats   showed   potential   subcutaneous   clearance   for   some   of   the   time   points   when   analysed   by   Elisa.   The   affected   sample   time   points   were   at   1   and   2   hrs,   which   showed   high   protein   levels   compared   to   samples   taken   at   time   point   0.5   hr.   It   was   presumed  that  the  contractor  might  have  missed  the  neck  vein  during  the  IV   injection  and  injected  the  rats’  muscle  instead  of  the  vein.  Out  of  the  5  rats   used   for   the   in   vivo   PK/PD   studies   of   rhGCSF,   only   1   rat   appeared   to   be   injected   intravenously.   To   annul   any   potential   inconsistencies   and   misinterpretation   of   data,   and   erroneous   data   analyses,   eliminating   the   rhGCSF   control   data   from   the   in   vivo   study   was   considered   appropriate.     Consequently,  the  half-­‐life  of  the  GCSF  tandems  proteins  were  compared  to   the  published  rhGCSF  due  to  time  limitation  of  this  project.       The   pharmacokinetic   profiles   for   all   GCSF   tandems   following   intravenous   dosing  at  250μg/kg  showed  an  approximately  3  fold  longer  circulating  half-­‐ 180   life   compared   to   that   reported   for   the   rhGCSF   (Tanaka   et   al.,   1991).   Additionally,   the   results   also   showed   that   GCSF4NAT   had   a   slower   rate   of   clearance   (10.74   hrs)   at   24   hrs   post   injection   compared   to   other   GCSF   tandems   (GCSF2NAT=7.38;   GCSF8NAT=6.74;   GCSF8QAT=5.87).   However,   this   increase   was   only   marginal,   and   not   significantly   different   from   other   GCSF  tandems.  Interestingly,  the  GCSF8NAT  with  more  glycosylation  sites  (8   sites)   had   a   slightly   lower   clearance   rate   compared   to   the   other   two   GCSF   tandems   with   lower   numbers   of   glycosylation   sites   (4   sites   in   GCSF4NAT   and   2   sites   in   GCSF2NAT).   This   implies   that   there   was   a   maximum   glycosylation  level  in  the  GCSF  tandems  beyond  which  further  glycosylation   provided   no   additional   benefit   towards   prolongation   of   the   half-­‐life   as   observed  in  GCSF8NAT.     The   GCSF8QAT   tandem   has   a   similar   clearance   rate   to   all   other   glycosylated   tandems,   implying   that   the   linker   glycosylations   are   not   required   for   the   improved   clearance   of   these   molecules.   It   therefore   appears   that   it   is   the   increase   in   Mw   of   the   tandem   by   virtue   of   containing   two   GCSF   molecules   that  is  responsible  for  the  observed  delayed  clearance.  Both  GCSF  molecules   also   have   the   potential   to   be   O-­‐link   glycosylated   which   would   also   contribute  to  the  increased  Mw  seen  over  that  of   native  GCSF  and  therefore   may   contribute   further   to   the   delayed   clearance.   Study   by   Marinaro   et   al.,   (2000)   showed   that   the   removal   of   O-­‐linked   glycosylated   from   human   IGFBP-­‐6  (Insulin-­‐like  growth  factor  binding  proteins-­‐6  act  as  inhibitor  of  IGF   actions)   decrease   the   circulating   half-­‐life   by   2.3   fold   compared   to   glycosylated-­‐IGFBP-­‐6  (Marinaro  et  al.,  2000).       This   theory   is   supported   by   a   study   that   was   performed   on   recombinant   human   Follicle-­‐Stimulating   Hormone   (rhFSH)   containing   different   N-­‐ glycosylated  linker  inserts  between  the  α-­‐  and  β-­‐  chains  in  the  same  rhFSH   ligand.   In   this   study,   the   terminal   half-­‐life   of   three   rhFSH   molecules   with   increasing  N-­‐linked  glycosylations:  rhFSH-­‐N1  (1  glycosylation  site),  rhFSH-­‐ N2   (2   glycosylation   sites)   and   rhFSH-­‐N4   (4   glycosylation   sites)   and   one   rhFSH  with  no  N-­‐linked  glycosylation  but  only  has  the  linker  insert  (rhFSH-­‐ N0),  were  assessed  after  intravenous  injection  into  Sprague  Dawley  rats. In   181   the   same   study,   two   other   molecules   with   O-­‐linked   glycosylation   were   equally   tested,   in   vivo,   to   determine   the   half-­‐life.   These   are   rhFSH-­‐CTP   (comprising  of  a  unique  carboxy-­‐terminal  peptide  (CTP)  that  contains  4  O-­‐   linked  glycosylation  sites  inserts  as  a  linker  between  the  α-­‐  and  β-­‐  chains  in   the  same  rhFSH  ligand)  and  rhFSH-­‐O1  (1  O-­‐linked  glycosylation  site  inserts   as  a  linker  between  the  α-­‐  and  β-­‐  chains  in  the  same  rhFSH  ligand).  The  half-­‐ lives   of   the   O-­‐linked   glycosylated   rhFSH   molecules   when   compared   to   those   of   the   N-­‐linked   showed   that   hyperglycosylation   of   either   O-­‐   or   N-­‐linked   of   the  rhFSH  ligand  increased  the  half-­‐life  of  rhFSH.  However,  the  increase  was   not  linearly  related  to  the  carbohydrate  sizes  and  numbers,  as  the  rhFSH-­‐2N,   rhFSH-­‐4N   and   rhFSH-­‐CTP   (4   O-­‐linked   glycosylation)   had   similar   half-­‐lives   which   were   2   fold   longer   compared   to   rhFSH,   rhFSH-­‐N0   and   rhFSH-­‐O1   (Klein  et  al.,  2002,  Weenen  et  al.,  2004).     Furthermore,   another   independent   study   also   described   an   increase   in   the   half-­‐life   of   rhFSH   by   introducing   two   N-­‐linked   glycosylation   sites   at   the   N   terminus   of   the   α-­‐subunit   (Perlman   et   al.,   2003),   which   is   different   from   the   linker   with   glycosylation   sites   in   the   earlier   mentioned   studies   on   rhFSH.   This  suggests  that  the  increase  in  the  number  of  glycosylation  sites  in  rhFSH,   irrespective   of   the   location   or   the   type   was   responsible   for   the   observed   increase  in  terminal  half-­‐life  of  rhFSH  in  these  studies.     The  importance  of  hyperglycosylation  to  a  longer  terminal  half-­‐life  is  further   highlighted   in   another   study   carried   out   to   investigate   the   effects   of   introducing  different  number  of  N-­‐linked  glycosylation  sites  on  the  half-­‐life   of   a   small   bispecific   single-­‐chain   diabody   (scDb   CEACD3)   (Used   for   the   retargeting   of   cytotoxic   T   cells   to   CEA-­‐expressing   tumor   cells).   Stock   et   al.   introduced  3,  6,  or  9  N-­‐glycosylation  sites  in  the  flanking  linker  of  the  scDb   molecule   and   a   C-­‐terminal   extension.   Interestingly,   the   results   showed   a   prolonged  circulating  half-­‐life  for  all  three  scDb  constructs  compared  to  the   unmodified   scDb.   However,   the   addition   of   3   N-­‐glycosylation   sites   is   adequate   to   prolong   circulation   time   and   not   significantly   different   from   6   or  9  N-­‐linked  glycosylation  (Stork  et  al.,  2008).     182   These  studies  revealed  that  the  addition  of  a  few  N  or  O-­‐glycans  either  in  a   linker   or   covalently   bound   to   a   protein   improved   the   pharmacokinetic   properties   of   the   protein,   thus   producing   a   novel   protein   with   moderately   prolonged   terminal   half-­‐life.   However,   there   is   a   maximum   benefit   of   glycosylation  regarding  the  half-­‐life,  and  additional  glycosylation  sites  might   not  extend  circulation  time.    This  view  is  consistent  with  our  findings  in  the   PK   studies,   where   all   the   proteins   showed   similar   terminal   half-­‐life   irrespective  of  the  numbers  of  glycosylation  sites  in  the  linker.     The  potency  of  each  GCSF  tandem  was  evaluated  in  vivo  using  white  blood   cell   (WBC)   and   neutrophil   counts.   WBC   counts   showed   no   statistical   difference   in   number   following   injection   with   either   vehicle   (PBS   only),   rhGCSF,  GCSF2NAT  and  GCSF4NAT.  However,  WBC  levels  peaked  at  24  hrs   post-­‐injection   for   GCSF8NAT   and   its   respective   control   GCSF8QAT   before   returning   to   baseline   values   at   72   hrs.   However,   this   effect   was   not   significantly  different  from  that  observed  in  the  rhGCSF  and  vehicle  controls.       In   contrast,   all   rats   injected   with   GCSF   tandems   and   rhGCSF   showed   an   increase   in   the   percentage   of   circulating   neutrophils   particularly   at   12   hrs   post  injection  for  GCSF2NAT  and  GCSF4NAT  compared  to  controls  (Vehicle,   rhGCSF  &  GCSF8QAT).  This  observed  increase  in  the  number  of  neutrophils   at   12   hrs   is   consistent   with   that   reported   by   Ulich   et   al.   (1988).   In   their   studies   in   mice,   they   observed   that   a   single   injection   of   GCSF   induced   a   temporary   neutropenia   in   the   mice.   However,   the   number   of   circulating   neutrophils  was  found  to  increase  significantly  by  5-­‐fold  within  30  minutes   and   peaked   at   12   hrs   post-­‐injection.   While   there   was   an   increase   in   the   number  of  circulating  neutrophils,  the  number  of  mature  neutrophils  in  the   bone  marrow  became  significantly  reduced  (Ulich  et  al.,  1988).     The   percentage   of   neutrophils   in   human   blood   circulation   is   around   60-­‐ 65%,  which  is  different  from  rats  and  mice.  The  percentage  of  neutrophils  in   normal  rats  is  between  22-­‐44.9%  (Sharma,  2013)  and  that  of  normal  mice  is   between   10-­‐40%   of   total   cell   counts   in   the   peripheral   blood   (www.ahc.umn.edu/rar/refvalues).  In  the  in  vivo  studies,  the  percentage  of   183   neutrophils   at   12   hrs   for   GCSF2NAT,   GCSF4NAT   and   GCSF8NAT   was   between   60-­‐70%   in   normal   rats   compared   to   rhGCSF,   GCSF8QAT   and   vehicle   (~45%).   The   presence   of   high   percentage   of   neutrophils,   but   very   low  percentage  of  other  cells  such  as  lymphocytes  (normal  range;  56-­‐78%),   monocytes,   eosinophils   or   basophils,   indicates   that   GCSF   is   selective   for   neutrophils.  This  observation  is  consistent  with  the  report  that  the  presence   of   mature   and   immature   (or   band)   neutrophils   in   circulation,   but   not   lymphocytes   or   eosinophils,   was   a   result   of   GCSF   induced   stimulation   of   neutrophils  from  the  bone  marrow  into  the  blood  (Semerad  et  al.,  2002).     In   the   current   study,   a   higher   increase   in   the   percentage   of   circulating   neutrophils   at   12   hrs   post   injection   was   observed   for   both   GCSF2NAT   and   GCSF4NAT   and   48   hrs   post   injection   for   GCSF8NAT   compared   to   controls   (Vehicle,   rhGCSF   &   GCSF8QAT).   This   confirms   the   positive   role   of   glycosylation   in   improving   the   pharmacodynamics   of   our   GCSF   tandems.   Although   the   half-­‐life   of   GCSF8NAT   exhibited   no   further   enhancement   beyond  either  the  GCSF2NAT,  GCSF4NAT  or  its  control  GCSF8QAT,  the  high   percentage   of   neutrophils   in   GCSF8NAT   compared   to   other   tandems   leads   us   to   hypothesize   that   GCSF8NAT   is   a   more   efficient   stimulator   of   neutrophils   as   it   has   a   longer   duration   of   action   evidenced   by   high   circulating   neutrophil   levels   at   48   hrs.   This   could   be   due   to   a   compound   effect  of  being  larger  in  size,  more  glycosylations  (more  terminal  sialisation)   or   the   presence   of   up   to   eight   glycosylations   sites   could   impede   the   interaction   of   these   GCSF   molecules   (impede   dimer   formation)   making   it   more  active  in  the  circulation.       The   number   of   N-­‐linked   glycosylation   sites   within   GCSF8NAT   tandem   linker   increased   its   size   to   ~70kDa,   which   is   bigger   than   both   GCSF4NAT   (~55kDa)  and  GCSF2NAT  (~40kDa).  We  can  hypothesis  that  GCSF4NAT  and   2NAT  are  cleared  more  efficiently  via  filtration  by  the  kidney,  the  increased   size  of  GCSF8NAT  delays  its  clearance  by  filtration,  as  kidney  is  reportedly   known  to  filter  molecules  smaller  than  the  human  serum  albumin  (67kDa)   (Dennis  et  al.,  2002).    Also  terminal  sialisation  of  GCSF8NAT  tandem  would   also  prevent  clearance  via  filtration  in  the  kidney  due  to  over-­‐expression  of   184   a   negatively   charged   barrier   alongside   the   glomerular   filtration   barrier,   preventing   the   passage   of   glycoproteins   through   charge   repulsion   (Varki,   2008).     Furthermore,  the  non-­‐glycosylated  GCSF  tandem  controls  were  observed  to   form   dimers   in   vitro,   which   could   be   attributed   to   the   formation   of   intermolecular   or   intramolecular   disulphide   bonds   resulting   from   the   interaction   of   the   free   cysteine   residues   (Cys17)   present   on   individual   GCSF   molecules.  Dimers  are  unable  to  induce  signal  transduction,  which  leads  to   less   stimulation   of   neutrophils.   Dimer   formation   is   potentially   easier   in   control  tandems  due  to  less  steric  hindrance  from  glycosylation  whereas  in   8NAT  (and  other  glycosylated  tandems)  the  presence  of  glycosylation  within   the  linker  could  be  inhibitory  to  the  formation  of  dimers.  For  instance,  2NAT   and   4NAT   stimulated   more   neutrophils   than   8QAT   at   12   hrs   but   8NAT   stimulated   more   neutrophils   than   8QAT   at   48   hrs   after   intravenous   injection.   The   presence   of   up   to   eight   glycosylation   sites   could   impede   the   interaction  of  these  GCSF  molecules  better  than  Both  GCSF2NAT  and  4NAT   (less   glycosylation   motifs).   This   was   evidenced   by   stability   results   in   the   previous  chapter,  showed  a  slight  dimer  formation  for  both  2NAT  and  4NAT,   whereas,   GCSF8NAT   showed   no   dimer   formation   at   all   (please   refer   to   Figure   6.5,   6.7   and   6.8).   This   may   also   be   a   contributing   factor   to   the   observed   increase   in   neutrophil   counts   with   these   proteins   and   therefore   increase  clearance  via  the  neutrophils-­‐mediated  pathway.   One   of   the   main   ways   in   which   GCSF   is   cleared   is   via   the   neutrophil   mediated  pathway.  It  is  this  pathway  that  we  hypothesise  is  the  main  route   of   clearance   of   our   glycosylated   GCSF   tandems.   This   effect   has   been   seen   with   Pegfilgratim,   which   was   developed   to   improve   PK   and   PD   of   GCSF   molecules   (i.e.   decreased   clearance   via   kidneys   therefore   increased   clearance   via   neutrophils).   For   example,   concentrations   of   Pegfilgrastim   in   patient   serums   following   administration,   remain   high   during   neutropenia,   but   reduced   when   the   numbers   of   neutrophils   increase   (Curran   and   Goa,   2002).     185   While   these   observations   were   different   from   our   initial   expectations   regarding   the   terminal   half-­‐life   of   the   GCSF   tandems   in   circulation,   as   we   hypothesized  that  increasing  the  glycosylation  sites  would  increase  the  size   and  thus  delay  the  clearance,  our  studies  showed  that  the  pharmacokinetic   and   pharmacodynamics   properties   of   a   therapeutic   protein   could   be   divergent.   These   observations   are   similar   to   the   findings   in   an   FSH   study   where  the  terminal  half-­‐life  of  different  analogues  of  FSH  with  two  (rhFSH-­‐ N2)   and   four   (rhFSH-­‐N4)   glycosylation   sites   were   found   to   have   similar   circulating   half-­‐lives   but   rhFSH-­‐N4   was   more   potent   in   the   stimulation   of   inhibin   A   and   follicles   than   the   rhFSH-­‐N2   due   to  the   addition   of   large,   highly   branched  carbohydrates  (Weenen  et  al.,  2004).     Therefore,  our  studies  of  the  tandem  proteins  revealed  that  all  glycosylated   tandem   GCSF   proteins   are   more   potent   than   rhGCSF   (containing   intramolecular   O-­‐linked   glycosylation)   and   GCSF8QAT   (containing   linker   and   O-­‐linked   glycosylation).   However,   despite   a   negligible   difference   in   circulating   half-­‐lives   of   the   tandem   GCSF   proteins,   GCSF8NAT   is   the   only   tandem  with  a  significant  potency  over  rhGCSF  and  GCSF8QAT,  being  a  more   potent  stimulator  of  neutrophils.           186   8. General  Discussion   GCSF  is  a  hormone  produced  by  different  tissues  to  stimulate  the  production   of  neutrophils  from  the  bone  marrow  into  the  blood  circulation.  The  rhGCSF   has   been   shown   to   stimulate   neutrophils   for   the   treatment   of   neutropenic   patients,   and   in   stem   cell   mobilization   in   the   circumstance   of   BM   transplantation.   The   commercial   products   of   rhGCSF   fall   under   the   category   of   either   short   acting   half-­‐life,   such   as,   Filgrastim   (NEUPOGEN®)   and   Lenograstim   (Granocyte®)   where   in   the   pharmacokinetic   properties   and   structural   homology   of   the   protein   are   similar   to   the   human   GCSF,   or   long   acting   half-­‐life   such   as,   Pegfilgrastim   (Neulasta),   where   in   chemical   alteration   with   PEGylation   has   been   applied   to   prolong   the   pharmacokinetics  by  reduced  renal  clearance.  Since  there  is  a  need  for  less   frequent   (daily)   dosing,   the   use   of   short-­‐acting   GCSF   products   have   been   limited   and   shifted   towards   the   use   of   the   long-­‐acting   products   (e.g.   Neulasta).   The   fast   growing   market   size   will   increase   demand   for   generating   similar   long-­‐acting   rhGCSF.     Therefore,   it   is   essential   to   produce   new   GCSF   compounds   with   improved   properties   over   other   products.   Asterion   is   employing  its  proprietary  protein  fusion  technology  (ProFuseTM)  to  produce   a   new   form   of   long-­‐acting   GCSF   molecule   that   retains   the   pharmacokinetic   and  efficiency  of  Neulasta,  but  with  a  more  competitive  manufacturing  and   cost-­‐of-­‐goods  benefits  over  Neulasta.   A   previous   study   by   Asterion   has   shown   that   the   use   of   glycosylated-­‐linkers   between   two   GH   ligands   to   create   protein-­‐tandems   resulted   in   their   glycosylation   and   an   increased   MW   whilst   maintaining   biological   activity.   This  technology  can  be  easily  applied  to  other  molecules  such  as,  GCSF.   The   data   obtained   in   this   study   have   shown   that   it   is   possible   to   clone,   express   and   purify   tandem   GCSF   molecules   containing   variably   glycosylated   linker   regions   from   a   CHO   cell   line.   The   purified   GCSF   tandems   are   stable   with   no   degradation.   The   apparent   stability   could   be   attributed   to   the   incorporated   Gly4Ser   flexible   linker.   This   had   been   demonstrated   in   187   recombinant  single-­‐chain  Fv  antibody  production  where  the  use  of  Gly4Ser   linker   offered   the   advantages   of   stability   and   lack   of   immunogenicity   (Huston   et   al.,   1993).   Our   data   (western   blots)   suggest   that   these   glycosylation   consensus   sequences   (motifs)   are   glycosylated   upon   expression   in   CHO   Flp-­‐In   cells   (mammalian   cell   lines)   as   the   MW   of   glycosylated   tandem   proteins   are   higher   than   their   corresponding   non-­‐ glycosylated   controls.   The   MW   of   the   non-­‐glycosylated   tandem   GCSF   controls   2QAT,   4QAT,   and   8QAT   were   45kDa,   45kDa   and   49kDa   respectively,   whereas   that   of   their   corresponding   glycosylated   GCSF   tandems   (2NAT,   4NAT,   and   8NAT)   were   approximately   52kDa,   60kDa   and   70kDa.     A   study   by   Mann   and   Jensen   (2003)   highlighted   that   each   unit   of   N-­‐ linked   glycosylation   contribute   a   minimum   of   approximately   800Da   to   the   construct  mass.  This  is  consistent  with  our  observation  of  increased  MWs  in   the   glycosylated   tandems   (~7kDa,   15kDa   and   ~21kDa)   compared   to   the   non-­‐glycosylated  controls.  However,  the  increased  MW  observed  was  higher   than  what  was  expected  using  Mann  and  Jensen’s  prediction,  which  could  be   attributed   to   the   unique   design   of   our   tandems   to   contain   flexible   linker.   Each  of  our  glycosylated  GCSF  tandem  contains  2,  4  or  8  glycosylation  motifs   within   the   flexible   linker   contribute   to   the   increase   in   MW.   Increasing   the   linker   length   to   accommodate   more   glycosylation   sites   (motifs)   further   increases   the   MW   of   the   molecule.   Also,   glycosylation   within   the   flexible   linker   region   increases   the   apparent   weight   of   the   tandems   not   just   by   weight   but   also   by   hydrodynamic   volume,   which   may   have   contributed   to   the   difference   in   MWs   between   our   study   and   Mann   and   Jensens’.   In   contrast,   the   non-­‐glycosylated   controls   used   in   this   work   have   shown   an   expected   MW   similar   to   those   determined   by   DNASTAR   laser   gene   (Table   3.2)  and  they  were  therefore  suitable  controls  in  this  perspective  of  proof  of   concept  testing.     All  GCSF  tandems  studied  showed  similar  in  vitro  biological  activities  in  an   AML-­‐193   proliferation   assay.   This   suggests   that   the   use   of   hyperglycosylated   linker   technology   is   applicable   for   GCSF   and   other   proteins  since  conjugation  is  not  impeding  or  shielding  residues  within  the   188   peptide   that   are   necessary   for   activity.   However   in   vivo,   although   all   tandems  showed  similar  terminal  half-­‐lives  (range:  5.87  –  10.74  hrs)  which   were   not   significantly   different,   studies   looking   at   neutrophil   mobilization   highlighted   striking   differences   between   these   tandems.   For   instance,   a   higher   increase   in   the   percentage   of   circulating   neutrophils   was   observed   for  both  GCSF2NAT  and  GCSF4NAT  at  12  hrs  post  injection  and  upto  48  hrs   post   injection   for   GCSF8NAT   compared   to   controls   (rhGCSF   &   GCSF8QAT).   While   this   confirms   the   positive   role   of   glycosylation   in   improving   the   pharmacodynamic  of  our  GCSF  tandems  it  equally  showed  GCSF8NAT  as  the   most  efficient  stimulator  of  neutrophils  as  it  has  a  longer  duration  of  action   evidenced  by  high  circulating  neutrophil  levels  at  48  hrs.  This  could  be  due   to   a   compound   effect   of   its   larger   in   size,   more   glycosylations   (more   terminal   sialylation)   which   limits   its   clearance   by   kidney   filtration   or   the   presence   of   8NAT   sites   in   the   linker   which   impedes   the   interaction   of   the   GCSF   molecules   (dimer   formation)   in   the   tandem   making   it   more   active   in   circulation.     Improving  the  circulating  half-­‐life  and  neutrophils  mobilization  has  been  the   mainstay  of  the  new  generation  of  GCSF.  GCSF  is  cleared  from  circulation  via   kidney   filtration   or   the   neutrophil   mediated   pathway.   Enhancing   the   clearance   via   neutrophil   mediated   pathway   increases   the   number   of   neutrophils   in   circulation,   a   characteristic   that   is   beneficial   for   treating   patients   with   neutropenia.   The   only   commercially   available   longer   acting   GCSF,   Pegfilgratim,   was   developed   to   improve   the   PK   and   PD   of   GCSF   molecule   by   PEGylation   (i.e.   decreased   clearance   via   kidneys   therefore   increased   clearance   via   neutrophils).   Similarly,   our   glycosylated   GCSF   tandems   were   designed   to   be   cleared   through   this   pathway.   We   hypothesized   that   increasing   the   glycosylation   sites   in   our   GCSF   tandems   would  increase  the  size  and  thus  delay  the  clearance.  However,  our  tandems   circulating   half-­‐lives   (5.87   –   10.74   hrs)   are   not   significantly   different   from   that   of   Pegfilgratim   (7.05   hrs)   even   though   the   molecular   weights   of   our   tandems  (~45kDa-­‐70kDa)  are  more  than  Pegfilgratim  (38.8kDa).     189   Despite   no   apparent   significant   improvement   in   circulating   half-­‐life   over   Pegfilgrastim,   our   GCSF   tandem   molecules   provide   some   physiological   advantages.   For   instance,   while   PEGylation   of   GCSF   in   Pegfilgrastim   has   been   shown   to   be   non-­‐biodegradable   and   toxic   (Patel   et   al.,   2014;   Caliceti   and   Veronese,   2003),   and   induce   vacuole   formation   in   renal   tubes   thereby   affecting   the   tissue   distribution   (Zhang   et   al.,   2014).   In   contrast,   our   GCSF   tandems   are   in   a   naturally   occurring   structure   being   hyperglycosylated,   and   therefore  can  undergo  degradation  within  the  human  body  with  no  potential   toxicity  of  PEGylation  was  supported  by  the  detection  of  PEG  in  bile  (Caliceti   and   Veronese,   2003).   Additionally,   the   cost   of   production   of   Pegfilgrastim   (involves  post-­‐expression  and  post-­‐purification  chemical  modification  of  the   starting   molecule   rhGCSF)   outweighs   that   of   our   GCSF   tandems,   which   are   easily  expressed  and  purified  using  simple  methods.   8.1 Future  Work   To  have  a  comprehensive  overview  of  our  tandems,  a  few  more  studies  are   required   which   were   not   covered   during   the   course   of   this   study.     For   instance,   in   this   study,   we   observed   high   variations   in   the   percentage   of   neutrophils   in   the   normal   rats   used   for   the   in   vivo   studies,   as   the   normal   range  of  neutrophils  is  22-­‐44%.  This  constituted  a  challenge  in  the  analysis   of   the   in   vivo   studies,   as   the   variations   in   the   neutrophil   counts   made   it   difficult   to   determine   the   exact   percentage   of   neutrophils   in   the   rats,   especially   since   rat’s   neutrophils   can   increase   even   with   stress   during   the   injection   or   due   to   infection   (Semerad   et   al.,   2002).   This   observation   was   seen  in  rats  injected  with  vehicle  (PBS  only)  that  exhibited  high  neutrophils   similar   to   the   test   rats   (please   refer   to   Figure   7.9).   Consequently,   neutropenic   rats   are   recommended   for   use   in   the   future   study,   as   we   need   a   base  line  for  the  evaluation  of  the  number  or  percentage  of  neutrophils.   Also,  it  was  not  conclusive  if  all  the  sites  of  glycosylation  were  occupied  by   glycan   moieties   during   protein   expression.   Therefore,   the   level   of   glycosylation   in   the   tandems   could   have   been   confirmed   by   Mass   spectrometry   (MS)   analysis   of   the   glycopeptides   fragments   following   190   treatment   of   glycosylated   tandems   with   the   PGNaseF   digestive   enzyme   (Gervais  et  al.,  2003).  Also,  the  effect  of  O-­‐linked  glycosylation  on  clearance   could   have   been   determined,   as   all   our   GCSF   tandems   (N-­‐linked   glycosylated)  and  their  respective  non-­‐glycosylated  controls  showed  similar   circulating   half-­‐life.   Additionally,   looking   at   the   structure   of   the   tandems   may   give   insight   as   to   whether   proteins   are   monomer,   dimer,   and   of   any   inter  or  intramolecular  associations.  A  number  of  analytical  methods  could   be   used   to   assess   these   structural   characteristics   such   as   Analytical   Ultracentrifugation   (monomer/dimer   formation),   Size   Exclusion   Chromatography  (monomer/Dimer),  Dynamic  Light  Scattering,  and  Small  X-­‐ Ray  Scattering.   8.1.1 Future  Work  to  Improve  GCSF  Tandems     Our   study   showed   that   tandem   GCSF   molecules   with   a   variable   N-­‐linked   glycosylated   flexible   linker   have   a   therapeutic   potential   of   stimulating   neutrophils   in   neutropenic   patients.   However,   improving   the   circulating   half-­‐life   of   our   tandems   over   the   commercially   available   long-­‐acting   GCSF   (Pegfilgrastim)  would  provide  novel  longer  acting  GCSF  products.  This  could   be   achieved   by   introducing   two   modifications:   removing   the   free   cysteine   residue   in   GCSF   molecule,   and   pH   switching   between   cell-­‐surface   and   endosome.   The  native  GCSF  contains  five  cysteine  residues,  two  internal  disulfide  bonds   at   positions   Cys36–   Cys42   and   Cys64–   Cys74   leaving   one   free   cysteine   residue  at  position  Cys17  with  a  free  sulfhydryl  group.  The  free  Cys17  of  one   GCSF  tandem  may  form  inter-­‐molecular  disulfide  bonds  with  another  GCSF   tandem,  which  results  in  the  formation  of  a  GCSF  tandem  dimer,  which  we   have  seen  in  non-­‐glycosylated  GCSF  tandems.  However,  it  has  been  reported   that   the   substitution   of   Cys17   with   alanine   (alanine   instead   of   cysteine   in   wild-­‐type   GCSF)   resulted   in   enhanced   stability   in   vitro   and   higher   bioavailability  in  vivo  than  wild-­‐type  GCSF,  possibly  through  the  elimination   of  dimerisation  caused  by  the  formation  of  inter-­‐molecular  disulfide  bonds   (Jiang  et  al.,  2011).  Therefore,  substituting  the  Cys17  with  alanine  for  both   191   GCSF   ligands   in   our   tandem   could   prevent   dimer   formation   and   thus   increase  their  bioavailability.     GCSF   binds   its   receptor   with   a   high   affinity   and   this   results   in   its   rapid   depletion   via   receptor-­‐mediated   endocytosis   by   circulating   neutrophils   expressing   GCSF-­‐R,   thus   diminishing   its   therapeutic   efficiency   at   a   pharmacokinetic   level.   It   has   been   demonstrated   that   substituting   ligand   residues   at   the   binding   site   of   GCSF   with   histidine   switches   protonation   states   between   cell-­‐surface   and   endosomal   pH.   Therefore,   decreasing   the   physiological   pH   from   ∼7   at   the   cell   surface   to   5   or   6   in   endosomes   by   selectively   mutating   amino   acids   at   the   GCSF   binding   site   will   deteriorate   interactions   at   endosomal   pH   whilst   maintaining   the   electrostatic   interactions   at   extracellular   pH,   and   consequently   increase   endocytic   GCSF   recycling   (Sarkar   et   al.,   2002).   Six   ligand   residues   located   at   the   binding   site   of  GCSF  (Glu20,  Gln21,  Asp110,  Asp113,  Thr117,  and  Gln120)  are  potential   targets  for  the  site-­‐directed  mutagenesis  with  6  histidine  residues  using  two   different   single-­‐histidine   mutants   (neutral   and   protonated   histidine).   The   suggested   mutation   is   based   on   the   principle   that   neutral   histidine   might   retain  relatively  tight  binding  on  the  cell  surface  while  protonated  histidine   might   lead   to   a   weaker   binding   in   endosomal   partitions.   Therefore,   substituting  the  amino  acids  of  each  GCSF  ligand  binding  site  in  our  tandems   with   histidine   would   reduce   its   receptor   binding   affinity   in   intracellular   endosomal  partitions  and  resultantly  leads  to  an  increased  recycling  of  GCSF   ligand   from   the   intracellular   cell   to   the   extracellular   medium   and   thereby   extend   the   circulating   half-­‐life   of   our   GCSF   tandems.   This   ultimately   facilitates   the   endocytic   GCSF   ligand   recycling   and   longer   half-­‐life   in   extracellular  circulation.             192   9. Conclusion   The  results  obtained  from  this  project  exhibited  that  it  is  possible  to  clone,   express   and   purify   GCSF   tandems.   It   also   appeared   that   the   use   of   glycosylated  motifs  (NAT)  within  a  flexible  linker  between  two  GCSF  ligands   to  generate  protein-­‐tandems  results  in  molecules  with  increased  molecular   weight   according   to   the   number   of   glycosylation   sites.   Tandems   of   GCSF   have   increased   in   vitro   bioactivity   compared   to   monomeric   GCSF   but   this   was   independent   of   glycosylation   and   glycosylation   did   not   inhibit   in  vitro   bioactivity.  Tandems  with  and  without  glycosylation  had  three-­‐fold  greater   half-­‐lives   than   rhGCSF   but   this   was   not   determined   by   the   number   of   glycosylation   sites.   There   was   evidence   that   GCSF8NAT   was   biologically   active   in   vivo.     The   results   confirm   the   hypothesis   that   it   is   possible   to   predictably   increase   the   molecular   weight   of   GCSF   tandems   and   retain   biological   activity   but   this   was   not   associated   with   a   predictable   prolongation  of  the  serum  half-­‐life.           193   Appendix  A   Appendix  A.1.  Nucleotide  Sequences  of  Primers   Primer   Nucleotide  Sequence   GCSF  Nhe1     5’-­‐AAATTTGGATCCGCTAGCCACCATGGCTGGACC-­‐3’   GCSF   Xho1   5’-­‐ATTCTCGAGGGGCTGGGCAAGGTGGCGTA-­‐3’   Rev   GCSF   5’-­‐AGGAGGGGATCCACCCCCCTGGG-­‐3’   BamH1   GCSF   Age1   5’-­‐AAGAAGACCGGTTCCACCGGTTCCACCTCCACCGGGCTGGGCAAGGTGGCG-­‐3’   Rev   CMVFor   5’-­‐TATTACCATGGTGATGCGGTTTTGG-­‐3’   BGHRev   5’-­‐TAGAAGGCACAGTCGAGG-­‐3’   GHseq2Rev   5’-­‐AAGGCCAGCTGGTGCAGACG-­‐3’     Appendix  A.2.  Restriction  Endonucleases  Cut  Sites     Enzyme   Restriction  site   Nhe1   5’  G/CTAGC  3’   3’  CGATC/G  5’   Xho1   5’  G/TCGAG  3’   3’  GAGCT/G  5’   BamH1   5’  G/GATCC  3’   3’  CCTAA/G  5’   Age1   5’  A/CCGGT  3’   3’  TGGCC/A  5’           194   Appendix  B   Appendix  B.1.  Nucleotide  and  Amino  Acid  Sequences  of  GH  Tandem   Nucleotide  sequence   GCTAGCcaccAtggctacaggctcccggacgtccctgctcctggcttttggcctgctctgcctgccctggct tcaagagggcagtgccTTCCCAACCATTCCCTTATCCAGGCTTTTTGACAACGCTATGCT CCGCGCCCATCGTCTGCACCAGCTGGCCTTTGACACCTACCAGGAGTTTGAAGAAGC CTATATCCCAAAGGAACAGAAGTATTCATTCCTGCAGAACCCCCAGACCTCCCTCTG TTTCTCAGAGTCTATTCCGACACCCTCCAACAGGGAGGAAACACAACAGAAATCCAA CCTAGAGCTGCTCCGCATCTCCCTGCTGCTCATCCAGTCGTGGCTGGAGCCCGTGCA GTTCCTCAGGAGTGTCTTCGCCAACAGCCTGGTGTACGGCGCCTCTGACAGCAACGT CTATGACCTCCTAAAGGACCTAGAGGAACGCATCCAAACGCTGATGGGGAGGCTGG AAGATGGCAGCCCCCGGACTGGGCAGATCTTCAAGCAGACCTACAGCAAGTTCGACA CAAACTCACACAACGATGACGCACTACTCAAGAACTACGGGCTGCTCTACTGCTTCA GGAAGGACATGGACAAGGTCGAGACATTCCTGCGCATCGTGCAGTGCCGCTCTGTGG AGGGCAGCTGTGGCTTCLinker:variableTTCCCAACCATTCCCTTATCCAGGCTTTT TGACAACGCTATGCTCCGCGCCCATCGTCTGCACCAGCTGGCCTTTGACACCTACCA GGAGTTTGAAGAAGCCTATATCCCAAAGGAACAGAAGTATTCATTCCTGCAGAACC CCCAGACCTCCCTCTGTTTCTCAGAGTCTATTCCGACACCCTCCAACAGGGAGGAAA CACAACAGAAATCCAACCTAGAGCTGCTCCGCATCTCCCTGCTGCTCATCCAGTCGT GGCTGGAGCCCGTGCAGTTCCTCAGGAGTGTCTTCGCCAACAGCCTGGTGTACGGCG CCTCTGACAGCAACGTCTATGACCTCCTAAAGGACCTAGAGGAAGGCATCCAAACGC TGATGGGGAGGCTGGAAGATGGCAGCCCCCGGACTGGGCAGATCTTCAAGCAGACCT ACAGCAAGTTCGACACAAACTCACACAACGATGACGCACTACTCAAGAACTACGGGC TGCTCTACTGCTTCAGGAAGGACATGGACAAGGTCGAGACATTCCTGCGCATCGTGC AGTGCCGCTCTGTGGAGGGCAGCTGTGGCTTC(ggtgga  ggtgga)  ACCGGT-­‐ catcatcaccatcaccat*   Amino  acid  Sequence:   matgsrtsIIIafgIIcIpwIqegsaFPTIPLSRLFDNAMLRAHRLHQLAFDTYQEFEEAYIP KEQKYSFLQNPQTSLCFSESIPTPSNREETQQKSNLELLRISLLLIQSWLEPVQFLRSVF ANSLVYGASDSNVYDLLKDLEEGIQTLMGRLEDGSPRTGQIFKQTYSKFDTNSHNDD ALLKNYGLLYCFRKDMDKVETFLRIVQCRSVEGSCGFLinker:variableFPTIPLSRLF DNAMLRAHRLHQLAFDTYQEFEEAYIPKEQKYSFLQNPQTSLCFSESIPTPSNREETQ QKSNLELLRISLLLIQSWLEPVQFLRSVFANSLVYGASDSNVYDLLKDLEEGIQTLMGR LEDGSPRTGQIFKQTYSKFDTNSHNDDALLKNYGLLYCFRKDMDKVETFLRIVQCRS VEGSCGFGGGGTGHHHHHH*   The  GH  tandem  with  a  4x  glycine  sequence  highlighted  in  purple  was  ligated   into  the  vector  pGHSecTag  between  Nhe1  restriction  site  highlighted  in  blue   and  Age1  restriction  site  highlighted  in  pink.  The  GH  tandem  containing  GH   signal   sequence   is   highlighted   in   bold   lower   case.   The   linker   between   two   GH  molecules  is  highlighted  in  red.    A  Hist  tag  is  highlighted  in  green.  This   molecule  will  be  used  as  the  template  to  replace  with  GCSF.       195   Appendix  B.2.  Nucleotide  and  Amino  Acid  Sequences  of  GCSF  Tandem   Nucleotide  sequence   GCTAGCcaccatggctggacctgccacccagagccccatgaagctgatggccctgcagctgctgctgtgg cacagtgcactctggacagtgcaggaagccACCCCCCTGGGCCCTGCCAGCTCCCTGCCCCAG AGCTTCCTGCTCAAGTGCTTAGAGCAAGTGAGGAAGATCCAGGGCGATGGCGCAGC GCTCCAGGAGAAGCTGTGTGCCACCTACAAGCTGTGCCACCCCGAGGAGCTGGTGCT GCTCGGACACTCTCTGGGCATCCCCTGGGCTCCCCTGAGCAGCTGCCCCAGCCAGGCC CTGCAGCTGGCAGGCTGCTTGAGCCAACTCCATAGCGGCCTTTTCCTCTACCAGGGG CTCCTGCAGGCCCTGGAAGGGATCTCCCCCGAGTTGGGTCCCACCTTGGACACACTG CAGCTGGACGTCGCCGACTTTGCCACCACCATCTGGCAGCAGATGGAAGAACTGGGA ATGGCCCCTGCCCTGCAGCCCACCCAGGGTGCCATGCCGGCCTTCGCCTCTGCTTTCC AGCGCCGGGCAGGAGGGGTCCTGGTTGCCTCCCATCTGCAGAGCTTCCTGGAGGTGT CGTACCGCGTTCTACGCCACCTTGCCCAGCCCLinker:variableACCCCCCTGGGCCCT GCCAGCTCCCTGCCCCAGAGCTTCCTGCTCAAGTGCTTAGAGCAAGTGAGGAAGATC CAGGGCGATGGCGCAGCGCTCCAGGAGAAGCTGTGTGCCACCTACAAGCTGTGCCAC CCCGAGGAGCTGGTGCTGCTCGGACACTCTCTGGGCATCCCCTGGGCTCCCCTGAGC AGCTGCCCCAGCCAGGCCCTGCAGCTGGCAGGCTGCTTGAGCCAACTCCATAGCGGC CTTTTCCTCTACCAGGGGCTCCTGCAGGCCCTGGAAGGGATCTCCCCCGAGTTGGGT CCCACCTTGGACACACTGCAGCTGGACGTCGCCGACTTTGCCACCACCATCTGGCAG CAGATGGAAGAACTGGGAATGGCCCCTGCCCTGCAGCCCACCCAGGGTGCCATGCCG GCCTTCGCCTCTGCTTTCCAGCGCCGGGCAGGAGGGGTCCTGGTTGCCTCCCATCTGC AGAGCTTCCTGGAGGTGTCGTACCGCGTTCTACGCCACCTTGCCCAGCCC(ggtgga   ggtgga)  ACCGGT-­‐catcatcaccatcaccat*   Amino  acid  Sequence:   MAGPATQSPMKLMALQLLLWHSALWTVQEATPLGPASSLPQSFLLKCLEQVRKIQ GDGAALQEKLCATYKLCHPEELVLLGHSLGIPWAPLSSCPSQALQLAGCLSQLHSGLF LYQGLLQALEGISPELGPTLDTLQLDVADFATTIWQQMEELGMAPALQPTQGAMPA FASAFQRRAGGVLVASHLQSFLEVSYRVLRHLAQPlinker:variableTPLGPASSLPQSF LLKCLEQVRKIQGDGAALQEKLCATYKLCHPEELVLLGHSLGIPWAPLSSCPSQALQL AGCLSQLHSGLFLYQGLLQALEGISPELGPTLDTLQLDVADFATTIWQQMEELGMAP ALQPTQGAMPAFASAFQRRAGGVLVASHLQSFLEVSYRVLRHLAQPGGGGTGHHHH HH*     The   GCSF   tandem   with   a   4x   glycine   sequence   highlighted   in   purple   was   ligated  into  the  vector  pGHSecTag  between  Nhe1  restriction  site   highlighted   in   blue   and   Age1   restriction   site   highlighted   in   pink.   The   GCSF   tandem   containing   GCSF   signal   sequence   is   highlighted   in   bold   lower   case.   The   linker   between   two   GCSF   molecules   is   highlighted   in   red.     A   Hist   tag   is   highlighted  in  green.  This  molecule  will  be  used  as  the  template  to  ligate  in   different  linker  constructions.       196   Appendix  B.3.  Nucleotide  and  Amino  Acid  Sequences  of  Linker  Regions   Linker   Nucleotide  sequence   GCSF2NAT   Amino  acid  sequence   CTCGAGGGTGGTGGAGGTAGTGGAGGAAACGCTA LEGGGGSGGNATGGGGS CAGGAGGTGGCGGGTCTGGTGGGGGGGGCTCTGG GGGGSGGGGSGGGGSGG AGGTGGAGGGTCAGGCGGGGGAGGATCAGGGGGA GGSGGNATGGGGSGS     GGCGGTTCCGGGGGCAACGCAACCGGGGGCGGAG GCTCCGGATCC   GCSF2QAT   CTCGAGGGTGGTGGAGGTAGTGGAGGACAGGCTA LEGGGGSGGQATGGGGS CAGGAGGTGGCGGGTCTGGTGGGGGGGGCTCTGG GGGGSGGGGSGGGGSGG AGGTGGAGGGTCAGGCGGGGGAGGATCAGGGGGA GGSGGQATGGGGSGS   GGCGGTTCCGGGGGCCAGGCAACCGGGGGCGGAG   GCTCCGGATCC   GCSF4NAT   CTCGAGGGTGGAGGAGGTTCTGGAGGTAATGCTA LEGGGGSGGNATGGGGS CTGGAGGTGGTGGCAGCGGAGGCAACGCAACAGG GGNATGGGGSGGNATG GGGTGGCGGATCTGGAGGAAACGCAACCGGTGGA GGGSGGNATGGGGSGS   GGGGGATCTGGTGGGAACGCTACCGGCGGAGGGG   GCTCTGGATCC   GCSF4QAT   CTCGAGGGCGGCGGTGGGTCCGGTGGCCAGGCTAC LEGGGGSGGQATGGGGS CGGAGGAGGCGGGAGTGGAGGCCAAGCCACAGGT GGQATGGGGSGGQATG GGCGGAGGGTCTGGCGGTCAGGCAACTGGCGGAG GGGSGGQATGGGGSGS   GAGGGTCAGGGGGGCAGGCCACGGGAGGTGGCGG   GAGCGGATCC     GCSF8NAT   CTCGAGGGCGGCGGAGGGAGTGGCGGTAACGCTA LEGGGGSGGNATGGGGS CGGGAGGAGGAGGCTCTGGCGGCAATGCAACCGG GGNATGGGGSGGNATG CGGTGGCGGGAGTGGCGGGAATGCCACAGGTGGG GGGSGGNATGGGGSGGG GGGGGTTCAGGCGGGAATGCTACTGGCGGCGGCG GSGGNATGGGGSGGNAT GTTCCGGAGGCGGAGGGTCTGGTGGGAACGCAAC GGGGSGGNATGGGGSGG CGGTGGTGGTGGAAGCGGAGGGAATGCTACCGGT NATGGGGSGS   GGCGGAGGAAGCGGTGGTAACGCCACTGGAGGCG   GCGGGTCCGGAGGCAACGCCACAGGGGGTGGAGG   GTCAGGATCC   GCSF8QAT   CTCGAGGGCGGCGGAGGGAGTGGCGGTCAAGCTA LEGGGGSGGQATGGGGS CGGGAGGAGGAGGCTCTGGCGGCCAGGCAACCGGC GGQATGGGGSGGQATG GGTGGCGGGAGTGGCGGGCAAGCCACAGGTGGGG GGGSGGQATGGGGSGGG GGGGTTCAGGCGGGCAGGCTACTGGCGGCGGCGGT GSGGQATGGGGSGGQAT TCCGGAGGCGGAGGGTCTGGTGGGCAAGCAACAG GGGGSGGQATGGGGSGG GTGGTGGTGGAAGCGGAGGGCAGGCTACTGGTGG QATGGGGSGS   CGGAGGAAGCGGTGGTCAAGCCACTGGAGGCGGC   GGGTCCGGAGGCCAGGCCACAGGGGGTGGAGGGT CAGGATCC   The  table  above  represents  the  linker  regions  that  were  ligated  between  two   GCSF  molecules  using  Xho1  restriction  site  highlighted  in  pink  and  BamH1   highlighted   in   green.   Glycosylation   motifs   highlighted   in   red   NAT   (AsnAlaThr)   are   amino   acids   in   which   N   is   recognized   by   cells   for   glycosylation.   While,   non-­‐glycosylation   motifs   (Controls)   highlighted   in   blue   QAT   (GlnAlaThr)   are   amino   acids   in   which   Q   is   not   recognized   by   cells   for   glycosylation.   197   Appendix  C.  pSecTag_Link-­‐Hist  Modulating  Vector:   Nhe1   GCSF  signal   GCSF1   sequence   Xho1   (Gly4Ser)n  Linker   BamH1   GCSF2     Age1           4x glycine/ 6x Hist/stop codon FRT site pSecTag_Link-Hist 5037 bp         198     Bibliography   AAPRO,  M.  S.,  BOHLIUS,  J.,  CAMERON,  D.  A.,  DAL  LAGO,  L.,  DONNELLY,  J.  P.,   KEARNEY,   N.,   LYMAN,   G.   H.,   PETTENGELL,   R.,   TJAN-­‐HEIJNEN,   V.   C.,   WALEWSKI,   J.,   WEBER,   D.   C.   &   ZIELINSKI,   C.   2011.   2010   update   of   EORTC   guidelines   for   the   use   of   granulocyte-­‐colony   stimulating   factor   to   reduce   the   incidence   of   chemotherapy-­‐induced   febrile   neutropenia  in  adult  patients  with  lymphoproliferative  disorders  and   solid  tumours.  Eur  J  Cancer,  47,  8-­‐32.   ABUCHOWSKI,   A.,   VAN   ES,   T.,   PALCZUK,   N.   C.   &   DAVIS,   F.   F.   1977.   Alteration   of   immunological   properties   of   bovine   serum   albumin   by   covalent   attachment  of  polyethylene  glycol.  J  Biol  Chem,  252,  3578-­‐81.   ANCLIFF,  P.  J.,  GALE,  R.  E.,  LIESNER,  R.,  HANN,  I.  &  LINCH,  D.  C.  2003.  Long-­‐ term   follow-­‐up   of   granulocyte   colony-­‐stimulating   factor   receptor   mutations   in   patients   with   severe   congenital   neutropenia:   implications   for   leukaemogenesis   and   therapy.   Br   J   Haematol,   120,   685-­‐690.   ANDERSEN,  J.  T.,  DABA,  M.  B.  &  SANDLIE,  I.  2010.  FcRn  binding  properties  of   an  abnormal  truncated  analbuminemic  albumin  variant.  Clin  Biochem,   43,  367-­‐72.   ANDERSEN,  J.  T.,  DEE  QIAN,  J.  &  SANDLIE,  I.  2006.  The  conserved  histidine   166  residue  of  the  human  neonatal  Fc  receptor  heavy  chain  is  critical   for  the  pH-­‐dependent  binding  to  albumin.  Eur  J  Immunol,  36,  3044-­‐51.   ANDERSEN,   J.   T.   &   SANDLIE,   I.   2009.   The   versatile   MHC   class   I-­‐related   FcRn   protects   IgG   and   albumin   from   degradation:   implications   for   development   of   new   diagnostics   and   therapeutics.   Drug   Metab   Pharmacokinet,  24,  318-­‐32.   ANDERSON,   C.   L.,   CHAUDHURY,   C.,   KIM,   J.,   BRONSON,   C.   L.,   WANI,   M.   A.   &   MOHANTY,   S.   2006.   Perspective-­‐-­‐   FcRn   transports   albumin:   relevance  to  immunology  and  medicine.  Trends  Immunol,  27,  343-­‐8.   ARVEDSON,   T.   &   GIFFIN,   M.   2012.   Structural   Biology   of   G-­‐CSF   and   Its   Receptor.   In:   MOLINEUX,   G.,   FOOTE,   M.   &   ARVEDSON,   T.   (eds.)   Twenty  Years  of  G-­‐CSF.  Springer  Basel.   ASANO,   S.,   URABE,   A.,   OKABE,   T.,   SATO,   N.   &   KONDO,   Y.   1977.   Demonstration  of  granulopoietic  factor(s)  in  the  plasma  of  nude  mice   transplanted  with  a  human  lung  cancer  and  in  the  tumor  tissue.  Blood,   49,  845-­‐52.   ATAERGIN,   S.,   ARPACI,   F.,   TURAN,   M.,   SOLCHAGA,   L.,   CETIN,   T.,   OZTURK,   M.,   OZET,   A.,   KOMURCU,   S.   &   OZTURK,   B.   2008.   Reduced   dose   of   lenograstim   is   as   efficacious   as   standard   dose   of   filgrastim   for   peripheral   blood   stem   cell   mobilization   and   transplantation:   a   randomized  study  in  patients  undergoing  autologous  peripheral  stem   cell  transplantation.  Am  J  Hematol,  83,  644-­‐8.   BAILON,   P.   &   WON,   C.   Y.   2009.   PEG-­‐modified   biopharmaceuticals.   Expert   Opin  Drug  Deliv,  6,  1-­‐16.   BAKER,   K.,   QIAO,   S.   W.,   KUO,   T.,   KOBAYASHI,   K.,   YOSHIDA,   M.,   LENCER,   W.   I.   &  BLUMBERG,  R.  S.  2009.  Immune  and  non-­‐immune  functions  of  the   (not  so)  neonatal  Fc  receptor,  FcRn.  Semin  Immunopathol,  31,  223-­‐36.   199   BASU,  S.,  HODGSON,  G.,  KATZ,  M.  &  DUNN,  A.  R.  2002.  Evaluation  of  role  of  G-­‐ CSF   in   the   production,   survival,   and   release   of   neutrophils   from   bone   marrow  into  circulation.  Blood,  100,  854-­‐61.   BEEKMAN,   R.   &   TOUW,   I.   P.   2010.   G-­‐CSF   and   its   receptor   in   myeloid   malignancy.  Blood,  115,  5131-­‐6.   BEEKMAN,   R.,   VALKHOF,   M.   G.,   SANDERS,   M.   A.,   VAN   STRIEN,   P.   M.,   HAANSTRA,   J.   R.,   BROEDERS,   L.,   GEERTSMA-­‐KLEINEKOORT,   W.   M.,   VEERMAN,   A.   J.,   VALK,   P.   J.,   VERHAAK,   R.   G.,   LOWENBERG,   B.   &   TOUW,   I.   P.   2012.   Sequential   gain   of   mutations   in   severe   congenital   neutropenia   progressing   to   acute   myeloid   leukemia.   Blood,   119,   5071-­‐7.   BENSINGER,   W.,   APPELBAUM,   F.,   ROWLEY,   S.,   STORB,   R.,   SANDERS,   J.,   LILLEBY,  K.,  GOOLEY,  T.,  DEMIRER,  T.,  SCHIFFMAN,  K.,  WEAVER,  C.  &   ET   AL.   1995.   Factors   that   influence   collection   and   engraftment   of   autologous  peripheral-­‐blood  stem  cells.  J  Clin  Oncol,  13,  2547-­‐55.   BERRIDGE,   M.   V.   &   TAN,   A.   S.   1993.   Characterization   of   the   cellular   reduction   of   3-­‐(4,5-­‐dimethylthiazol-­‐2-­‐yl)-­‐2,5-­‐diphenyltetrazolium   bromide   (MTT):   subcellular   localization,   substrate   dependence,   and   involvement   of   mitochondrial   electron   transport   in   MTT   reduction.   Arch  Biochem  Biophys,  303,  474-­‐82.   BLOCK,  H.,  MAERTENS,  B.,  SPRIESTERSBACH,  A.,  BRINKER,  N.,  KUBICEK,  J.,   FABIS,  R.,  LABAHN,  J.  &  SCHAFER,  F.  2009.  Immobilized-­‐metal  affinity   chromatography  (IMAC):  a  review.  Methods  Enzymol,  463,  439-­‐73.   BONEBERG,   E.   M.,   HARENG,   L.,   GANTNER,   F.,   WENDEL,   A.   &   HARTUNG,   T.   2000.   Human   monocytes   express   functional   receptors   for   granulocyte   colony-­‐stimulating   factor   that   mediate   suppression   of   monokines  and  interferon-­‐gamma.  Blood,  95,  270-­‐6.   BRADLEY,   T.   R.   &   METCALF,   D.   1966.   The   growth   of   mouse   bone   marrow   cells  in  vitro.  Aust  J  Exp  Biol  Med  Sci,  44,  287-­‐99.   BROXMEYER,   H.   E.,   ORSCHELL,   C.   M.,   CLAPP,   D.   W.,   HANGOC,   G.,   COOPER,   S.,   PLETT,  P.  A.,  LILES,  W.  C.,  LI,  X.,  GRAHAM-­‐EVANS,  B.,  CAMPBELL,  T.  B.,   CALANDRA,  G.,  BRIDGER,  G.,  DALE,  D.  C.  &  SROUR,  E.  F.  2005.  Rapid   mobilization   of   murine   and   human   hematopoietic   stem   and   progenitor  cells  with  AMD3100,  a  CXCR4  antagonist.  J   Exp   Med,  201,   1307-­‐18.   BUSSOLINO,   F.,   WANG,   J.   M.,   DEFILIPPI,   P.,   TURRINI,   F.,   SANAVIO,   F.,   EDGELL,   C.   J.,   AGLIETTA,   M.,   ARESE,   P.   &   MANTOVANI,   A.   1989.   Granulocyte-­‐   and   granulocyte-­‐macrophage-­‐colony   stimulating   factors   induce   human   endothelial   cells   to   migrate   and   proliferate.   Nature,  337,  471-­‐3.   BUTLER,  M.  &  SPEARMAN,  M.  2014.  The  choice  of  mammalian  cell  host  and   possibilities   for   glycosylation   engineering.   Curr   Opin   Biotechnol,   30,   107-­‐12.   BYRNE,   B.,   DONOHOE,   G.   G.   &   O'KENNEDY,   R.   2007.   Sialic   acids:   carbohydrate   moieties   that   influence   the   biological   and   physical   properties  of  biopharmaceutical  proteins  and  living  cells.  Drug  Discov   Today,  12,  319-­‐26.   200   CALICETI,  P.  &  VERONESE,  F.  M.  2003.  Pharmacokinetic  and  biodistribution   properties   of   poly(ethylene   glycol)-­‐protein   conjugates.   Adv   Drug   Deliv  Rev,  55,  1261-­‐77.   CHAUDHURY,   C.,   BROOKS,   C.   L.,   CARTER,   D.   C.,   ROBINSON,   J.   M.   &   ANDERSON,   C.   L.   2006.   Albumin   binding   to   FcRn:   distinct   from   the   FcRn-­‐IgG  interaction.  Biochemistry,  45,  4983-­‐90.   CHAUDHURY,  C.,  MEHNAZ,  S.,  ROBINSON,  J.  M.,  HAYTON,  W.  L.,  PEARL,  D.  K.,   ROOPENIAN,   D.   C.   &   ANDERSON,   C.   L.   2003.   The   major   histocompatibility   complex-­‐related   Fc   receptor   for   IgG   (FcRn)   binds   albumin  and  prolongs  its  lifespan.  J  Exp  Med,  197,  315-­‐22.   CHEERS,  C.,  HAIGH,  A.  M.,  KELSO,  A.,  METCALF,  D.,  STANLEY,  E.  R.  &  YOUNG,   A.   M.   1988.   Production   of   colony-­‐stimulating   factors   (CSFs)   during   infection:   separate   determinations   of   macrophage-­‐,   granulocyte-­‐,   granulocyte-­‐macrophage-­‐,  and  multi-­‐CSFs.  Infect  Immun,  56,  247-­‐51.   CHUNG,  H.  K.,  KIM,  S.  W.,  BYUN,  S.  J.,  KO,  E.  M.,  CHUNG,  H.  J.,  WOO,  J.  S.,  YOO,  J.   G.,  LEE,  H.  C.,  YANG,  B.  C.,  KWON,  M.,  PARK,  S.  B.,  PARK,  J.  K.  &  KIM,  K.   W.   2011.   Enhanced   biological   effects   of   Phe140Asn,   a   novel   human   granulocyte   colony-­‐stimulating   factor   mutant,   on   HL60   cells.   BMB   Rep,  44,  686-­‐91.   COOPER,  K.  L.,  MADAN,  J.,  WHYTE,  S.,  STEVENSON,  M.  D.  &  AKEHURST,  R.  L.   2011.  Granulocyte  colony-­‐stimulating  factors  for  febrile  neutropenia   prophylaxis   following   chemotherapy:   systematic   review   and   meta-­‐ analysis.  BMC  Cancer,  11,  404.   CORY,   A.   H.,   OWEN,   T.   C.,   BARLTROP,   J.   A.   &   CORY,   J.   G.   1991.   Use   of   an   aqueous   soluble   tetrazolium/formazan   assay   for   cell   growth   assays   in  culture.  Cancer  Commun,  3,  207-­‐12.   COX,   G.   N.,   CHLIPALA,   E.   A.,   SMITH,   D.   J.,   CARLSON,   S.   J.,   BELL,   S.   J.   &   DOHERTY,   D.   H.   2014.   Hematopoietic   properties   of   granulocyte   colony-­‐stimulating   factor/immunoglobulin   (G-­‐CSF/IgG-­‐Fc)   fusion   proteins  in  normal  and  neutropenic  rodents.  PLoS  One,  9,  e91990.   COX,   G.   N.,   SMITH,   D.   J.,   CARLSON,   S.   J.,   BENDELE,   A.   M.,   CHLIPALA,   E.   A.   &   DOHERTY,   D.   H.   2004.   Enhanced   circulating   half-­‐life   and   hematopoietic  properties  of  a  human  granulocyte  colony-­‐stimulating   factor/immunoglobulin  fusion  protein.  Exp  Hematol,  32,  441-­‐9.   CURRAN,  M.  P.  &  GOA,  K.  L.  2002.  Pegfilgrastim.  Drugs,  62,  1207-­‐1213.   D'SOUZA,   A.,   JAIYESIMI,   I.,   TRAINOR,   L.   &   VENUTURUMILI,   P.   2008.   Granulocyte  colony-­‐stimulating  factor  administration:  adverse  events.   Transfus  Med  Rev,  22,  280-­‐90.   DAGDAS,  S.,  OZET,  G.,  ALANOGLU,  G.,  AYLI,  M.,  GOKMEN  AKOZ,  A.  &  EREKUL,   S.  2006.  Unusual  extramedullary  hematopoiesis  in  a  patient  receiving   granulocyte  colony-­‐stimulating  factor.  Acta  Haematol,  116,  198-­‐202.   DALL'ACQUA,   W.   F.,   WOODS,   R.   M.,   WARD,   E.   S.,   PALASZYNSKI,   S.   R.,   PATEL,   N.  K.,  BREWAH,  Y.  A.,  WU,  H.,  KIENER,  P.  A.  &  LANGERMANN,  S.  2002.   Increasing  the  affinity  of  a  human  IgG1  for  the  neonatal  Fc  receptor:   biological  consequences.  J  Immunol,  169,  5171-­‐80.   DAMIANI,  R.,  OLIVEIRA,  J.  E.,  VORAUER-­‐UHL,  K.,  PERONI,  C.  N.,  VIANNA,  E.  G.,   BARTOLINI,  P.  &  RIBELA,  M.  T.  2009.  Stable  expression  of  a  human-­‐ like   sialylated   recombinant   thyrotropin   in   a   Chinese   hamster   ovary   201   cell  line  expressing  alpha2,6-­‐sialyltransferase.  Protein  Expr  Purif,  67,   7-­‐14.   DELGADO,  C.,  FRANCIS,  G.  E.  &  FISHER,  D.  1992.  The  uses  and  properties  of   PEG-­‐linked  proteins.  Crit  Rev  Ther  Drug  Carrier  Syst,  9,  249-­‐304.   DENNIS,   M.   S.,   ZHANG,   M.,   MENG,   Y.   G.,   KADKHODAYAN,   M.,   KIRCHHOFER,   D.,   COMBS,   D.   &   DAMICO,   L.   A.   2002.   Albumin   binding   as   a   general   strategy  for  improving  the  pharmacokinetics  of  proteins.  J  Biol  Chem,   277,  35035-­‐43.   DUMONT,  J.  A.,  LOW,  S.  C.,  PETERS,  R.  T.  &  BITONTI,  A.  J.  2006.  Monomeric   Fc   fusions:   impact   on   pharmacokinetic   and   biological   activity   of   protein  therapeutics.  BioDrugs,  20,  151-­‐60.   ELLIOTT,  S.,  EGRIE,  J.,  BROWNE,  J.,  LORENZINI,  T.,  BUSSE,  L.,  ROGERS,  N.  &   PONTING,   I.   2004.   Control   of   rHuEPO   biological   activity:   the   role   of   carbohydrate.  Exp  Hematol,  32,  1146-­‐55.   ELLIOTT,   S.,   LORENZINI,   T.,   ASHER,   S.,   AOKI,   K.,   BRANKOW,   D.,   BUCK,   L.,   BUSSE,  L.,  CHANG,  D.,  FULLER,  J.,  GRANT,  J.,  HERNDAY,  N.,  HOKUM,  M.,   HU,   S.,   KNUDTEN,   A.,   LEVIN,   N.,   KOMOROWSKI,   R.,   MARTIN,   F.,   NAVARRO,   R.,   OSSLUND,   T.,   ROGERS,   G.,   ROGERS,   N.,   TRAIL,   G.   &   EGRIE,  J.  2003.  Enhancement  of  therapeutic  protein  in  vivo  activities   through  glycoengineering.  Nat  Biotechnol,  21,  414-­‐21.   FAVREAU,   A.   J.   &   SATHYANARAYANA,   P.   2012.   miR-­‐590-­‐5p,   miR-­‐219-­‐5p,   miR-­‐15b   and   miR-­‐628-­‐5p   are   commonly   regulated   by   IL-­‐3,   GM-­‐CSF   and  G-­‐CSF  in  acute  myeloid  leukemia.  Leuk  Res,  36,  334-­‐41.   FOSSIEZ,  F.,  DJOSSOU,  O.,  CHOMARAT,  P.,  FLORES-­‐ROMO,  L.,  AIT-­‐YAHIA,  S.,   MAAT,   C.,   PIN,   J.   J.,   GARRONE,   P.,   GARCIA,   E.,   SAELAND,   S.,   BLANCHARD,   D.,   GAILLARD,   C.,   DAS   MAHAPATRA,   B.,   ROUVIER,   E.,   GOLSTEIN,   P.,   BANCHEREAU,   J.   &   LEBECQUE,   S.   1996.   T   cell   interleukin-­‐17  induces  stromal  cells  to  produce  proinflammatory  and   hematopoietic  cytokines.  J  Exp  Med,  183,  2593-­‐603.   FUKUNAGA,   R.,   ISHIZAKA-­‐IKEDA,   E.   &   NAGATA,   S.   1990a.   Purification   and   characterization   of   the   receptor   for   murine   granulocyte   colony-­‐ stimulating  factor.  J  Biol  Chem,  265,  14008-­‐15.   FUKUNAGA,   R.,   SETO,   Y.,   MIZUSHIMA,   S.   &   NAGATA,   S.   1990b.   3   DIFFERENT   MESSENGER-­‐RNAS   ENCODING   HUMAN   GRANULOCYTE   COLONY-­‐ STIMULATING   FACTOR   RECEPTOR.   Proceedings   of   the   National   Academy  of  Sciences  of  the  United  States  of  America,  87,  8702-­‐8706.   FURMENTO,   V.   A.,   MARINO,   J.,   BLANK,   V.   C.   &   ROGUIN,   L.   P.   2014.   The   granulocyte   colony-­‐stimulating   factor   (G-­‐CSF)   upregulates   metalloproteinase-­‐2   and   VEGF   through   PI3K/Akt   and   Erk1/2   activation  in  human  trophoblast  Swan  71  cells.  Placenta,  35,  937-­‐46.   GABERC-­‐POREKAR,   V.,   ZORE,   I.,   PODOBNIK,   B.   &   MENART,   V.   2008.   Obstacles  and  pitfalls  in  the  PEGylation  of  therapeutic  proteins.  Curr   Opin  Drug  Discov  Devel,  11,  242-­‐50.   GAERTNER,   H.   F.   &   OFFORD,   R.   E.   1996.   Site-­‐specific   attachment   of   functionalized   poly(ethylene   glycol)   to   the   amino   terminus   of   proteins.  Bioconjug  Chem,  7,  38-­‐44.   GASCON,   P.   2012.   Presently   available   biosimilars   in   hematology-­‐oncology:   G-­‐CSF.  Target  Oncol,  7  Suppl  1,  S29-­‐34.   202   GHETIE,   V.,   HUBBARD,   J.   G.,   KIM,   J.   K.,   TSEN,   M.   F.,   LEE,   Y.   &   WARD,   E.   S.   1996.   Abnormally   short   serum   half-­‐lives   of   IgG   in   beta   2-­‐ microglobulin-­‐deficient  mice.  Eur  J  Immunol,  26,  690-­‐6.   HAMIDI,  M.,  AZADI,  A.  &  RAFIEI,  P.  2006.  Pharmacokinetic  consequences  of   pegylation.  Drug  Delivery,  13,  399-­‐409.   HAMILTON,   J.   A.   2008.   Colony-­‐stimulating   factors   in   inflammation   and   autoimmunity.  Nat  Rev  Immunol,  8,  533-­‐44.   HARADA,   M.,   QIN,   Y.,   TAKANO,   H.,   MINAMINO,   T.,   ZOU,   Y.,   TOKO,   H.,   OHTSUKA,   M.,   MATSUURA,   K.,   SANO,   M.,   NISHI,   J.,   IWANAGA,   K.,   AKAZAWA,  H.,  KUNIEDA,  T.,  ZHU,  W.,  HASEGAWA,  H.,  KUNISADA,  K.,   NAGAI,   T.,   NAKAYA,   H.,   YAMAUCHI-­‐TAKIHARA,   K.   &   KOMURO,   I.   2005.  G-­‐CSF  prevents  cardiac  remodeling  after  myocardial  infarction   by   activating   the   Jak-­‐Stat   pathway   in   cardiomyocytes.   Nat   Med,   11,   305-­‐11.   HARTUNG,  T.,  DOECKE,  W.  D.,  BUNDSCHUH,  D.,  FOOTE,  M.  A.,  GANTNER,  F.,   HERMANN,   C.,   LENZ,   A.,   MILWEE,   S.,   RICH,   B.,   SIMON,   B.,   VOLK,   H.   D.,   VON   AULOCK,   S.   &   WENDEL,   A.   1999.   Effect   of   filgrastim   treatment   on  inflammatory  cytokines  and  lymphocyte  functions.  Clin  Pharmacol   Ther,  66,  415-­‐24.   HASEGAWA,   M.   1993.   A   thermodynamic   model   for   denaturation   of   granulocyte   colony-­‐stimulating   factor:   O-­‐linked   sugar   chain   suppresses   not   the   triggering   deprotonation   but   the   succeeding   denaturation.  Biochim  Biophys  Acta,  1203,  295-­‐7.   HERMANS,   M.   H.,   VAN   DE   GEIJN,   G.   J.,   ANTONISSEN,   C.,   GITS,   J.,   VAN   LEEUWEN,  D.,  WARD,  A.  C.  &  TOUW,  I.  P.  2003.  Signaling  mechanisms   coupled   to   tyrosines   in   the   granulocyte   colony-­‐stimulating   factor   receptor  orchestrate  G-­‐CSF-­‐induced  expansion  of  myeloid  progenitor   cells.  Blood,  101,  2584-­‐90.   HIROKAWA,  M.,  LEE,  M.,  MOTEGI,  M.  &  MIURA,  A.  B.  1996.  Reversible  renal   impairment   during   leukocytosis   induced   by   G-­‐CSF   in   non-­‐Hodgkin's   lymphoma.  Am  J  Hematol,  51,  328-­‐9.   HOFFNER,  G.  &  DJIAN,  P.  2005.  Transglutaminase  and  diseases  of  the  central   nervous  system.  Front  Biosci,  10,  3078-­‐92.   HOGGATT,   J.   &   PELUS,   L.   M.   2014.   New   G-­‐CSF   agonists   for   neutropenia   therapy.  Expert  Opin  Investig  Drugs,  23,  21-­‐35.   HÖLIG,   K.   2013.   G-­‐CSF   in   Healthy   Allogeneic   Stem   Cell   Donors.   Transfusion   Medicine  and  Hemotherapy,  40,  225-­‐235.   HUSTON,  J.  S.,  TAI,  M.  S.,  MCCARTNEY,  J.,  KECK,  P.  &  OPPERMANN,  H.  1993.   Antigen   recognition   and   targeted   delivery   by   the   single-­‐chain   Fv.   Cell   Biophys,  22,  189-­‐224.   ICHIKAWA,   Y.,   PLUZNIK,   D.   H.   &   SACHS,   L.   1966.   In   vitro   control   of   the   development   of   macrophage   and   granulocyte   colonies.   Proc   Natl   Acad  Sci  U  S  A,  56,  488-­‐95.   INOKUCHI,   R.,   MANABE,   H.,   OHTA,   F.,   NAKAMURA,   K.,   NAKAJIMA,   S.   &   YAHAGI,   N.   2015.   Granulocyte   colony-­‐stimulating   factor-­‐producing   lung  cancer  and  acute  respiratory  distress  syndrome.  Clin  Respir  J,  9,   250-­‐2.   ISRAEL,  E.  J.,  WILSKER,  D.  F.,  HAYES,  K.  C.,  SCHOENFELD,  D.  &  SIMISTER,  N.   E.   1996.   Increased   clearance   of   IgG   in   mice   that   lack   beta   2-­‐ 203   microglobulin:  possible  protective  role  of  FcRn.  Immunology,  89,  573-­‐ 8.   JAIN,   A.   &   JAIN,   S.   K.   2008.   PEGylation:   an   approach   for   drug   delivery.   A   review.  Crit  Rev  Ther  Drug  Carrier  Syst,  25,  403-­‐47.   JIANG,   Y.,   JIANG,   W.,   QIU,   Y.   &   DAI,   W.   2011.   Effect   of   a   structurally   modified   human  granulocyte  colony  stimulating  factor,  G-­‐CSFa,  on  leukopenia   in  mice  and  monkeys.  J  Hematol  Oncol,  4,  28.   KASTURI,  L.,  ESHLEMAN,  J.  R.,  WUNNER,  W.  H.  &  SHAKIN-­‐ESHLEMAN,  S.  H.   1995.   The   hydroxy   amino   acid   in   an   Asn-­‐X-­‐Ser/Thr   sequon   can   influence   N-­‐linked   core   glycosylation   efficiency   and   the   level   of   expression  of  a  cell  surface  glycoprotein.  J  Biol  Chem,  270,  14756-­‐61.   KAWAKAMI,   M.,   TSUTSUMI,   H.,   KUMAKAWA,   T.,   ABE,   H.,   HIRAI,   M.,   KUROSAWA,   S.,   MORI,   M.   &   FUKUSHIMA,   M.   1990.   Levels   of   serum   granulocyte   colony-­‐stimulating   factor   in   patients   with   infections.   Blood,  76,  1962-­‐4.   KIM,   P.   J.,   LEE,   D.   Y.   &   JEONG,   H.   2009.   Centralized   modularity   of   N-­‐linked   glycosylation  pathways  in  mammalian  cells.  PLoS  One,  4,  e7317.   KINSTLER,  O.  B.,  BREMS,  D.  N.,  LAUREN,  S.  L.,  PAIGE,  A.  G.,  HAMBURGER,  J.  B.   &   TREUHEIT,   M.   J.   1996.   Characterization   and   stability   of   N-­‐ terminally  PEGylated  rhG-­‐CSF.  Pharm  Res,  13,  996-­‐1002.   KLEIN,   J.,   LOBEL,   L.,   POLLAK,   S.,   FERIN,   M.,   XIAO,   E.,   SAUER,   M.   &   LUSTBADER,  J.  W.  2002.  Pharmacokinetics  and  pharmacodynamics  of   single-­‐chain   recombinant   human   follicle-­‐stimulating   hormone   containing   the   human   chorionic   gonadotropin   carboxyterminal   peptide  in  the  rhesus  monkey.  Fertil  Steril,  77,  1248-­‐55.   KORNFELD,   R.   &   KORNFELD,   S.   1985.   Assembly   of   asparagine-­‐linked   oligosaccharides.  Annu  Rev  Biochem,  54,  631-­‐64.   KREISIG,  T.,  PRASSE,  A.  A.,  ZSCHARNACK,  K.,  VOLKE,  D.  &  ZUCHNER,  T.  2014.   His-­‐tag   protein   monitoring   by   a   fast   mix-­‐and-­‐measure   immunoassay.   Sci  Rep,  4,  5613.   LAMBERTINI,  M.,  DEL  MASTRO,  L.,  BELLODI,  A.  &  PRONZATO,  P.  2014.  The   five   "Ws"   for   bone   pain   due   to   the   administration   of   granulocyte-­‐ colony  stimulating  factors  (G-­‐CSFs).  Crit   Rev   Oncol   Hematol,  89,  112-­‐ 28.   LANGE,   B.,   VALTIERI,   M.,   SANTOLI,   D.,   CARACCIOLO,   D.,   MAVILIO,   F.,   GEMPERLEIN,  I.,  GRIFFIN,  C.,  EMANUEL,  B.,  FINAN,  J.,  NOWELL,  P.  &   ET   AL.   1987.   Growth   factor   requirements   of   childhood   acute   leukemia:   establishment   of   GM-­‐CSF-­‐dependent   cell   lines.   Blood,   70,   192-­‐9.   LANGRISH,  C.  L.,  CHEN,  Y.,  BLUMENSCHEIN,  W.  M.,  MATTSON,  J.,  BASHAM,  B.,   SEDGWICK,   J.   D.,   MCCLANAHAN,   T.,   KASTELEIN,   R.   A.   &   CUA,   D.   J.   2005.   IL-­‐23   drives   a   pathogenic   T   cell   population   that   induces   autoimmune  inflammation.  J  Exp  Med,  201,  233-­‐40.   LARSEN,   A.,   DAVIS,   T.,   CURTIS,   B.   M.,   GIMPEL,   S.,   SIMS,   J.   E.,   COSMAN,   D.,   PARK,  L.,  SORENSEN,  E.,  MARCH,  C.  J.  &  SMITH,  C.  A.  1990.  Expression   cloning  of  a  human  granulocyte  colony-­‐stimulating  factor  receptor:  a   structural   mosaic   of   hematopoietin   receptor,   immunoglobulin,   and   fibronectin  domains.  J  Exp  Med,  172,  1559-­‐70.   204   LI,  H.  &  D'ANJOU,  M.  2009.  Pharmacological  significance  of  glycosylation  in   therapeutic  proteins.  Curr  Opin  Biotechnol,  20,  678-­‐84.   LIESCHKE,  G.  J.,  GRAIL,  D.,  HODGSON,  G.,  METCALF,  D.,  STANLEY,  E.,  CHEERS,   C.,   FOWLER,   K.   J.,   BASU,   S.,   ZHAN,   Y.   F.   &   DUNN,   A.   R.   1994.   Mice   lacking   granulocyte   colony-­‐stimulating   factor   have   chronic   neutropenia,  granulocyte  and  macrophage  progenitor  cell  deficiency,   and  impaired  neutrophil  mobilization.  Blood,  84,  1737-­‐46.   LIONGUE,   C.   &   WARD,   A.   C.   2014.   Granulocyte   colony-­‐stimulating   factor   receptor  mutations  in  myeloid  malignancy.  Front  Oncol,  4,  93.   LITAM,   P.   P.,   FRIEDMAN,   H.   D.   &   LOUGHRAN,   T.   P.,   JR.   1993.   Splenic   extramedullary   hematopoiesis   in   a   patient   receiving   intermittently   administered   granulocyte   colony-­‐stimulating   factor.   Ann  Intern  Med,   118,  954-­‐5.   LIU,   F.,   WU,   H.   Y.,   WESSELSCHMIDT,   R.,   KORNAGA,   T.   &   LINK,   D.   C.   1996.   Impaired   production   and   increased   apoptosis   of   neutrophils   in   granulocyte   colony-­‐stimulating   factor   receptor-­‐deficient   mice.   Immunity,  5,  491-­‐501.   LLAMAS-­‐VELASCO,  M.,  GARCIA-­‐MARTIN,  P.,  SANCHEZ-­‐PEREZ,  J.,  FRAGA,  J.  &   GARCIA-­‐DIEZ,   A.   2013.   Sweet's   syndrome   with   subcutaneous   involvement   associated   with   pegfilgrastim   treatment:   first   reported   case.  J  Cutan  Pathol,  40,  46-­‐9.   LORD,  B.  I.,  MOLINEUX,  G.,  POJDA,  Z.,  SOUZA,  L.  M.,  MERMOD,  J.  J.  &  DEXTER,   T.   M.   1991.   Myeloid   cell   kinetics   in   mice   treated   with   recombinant   interleukin-­‐3,   granulocyte   colony-­‐stimulating   factor   (CSF),   or   granulocyte-­‐macrophage  CSF  in  vivo.  Blood,  77,  2154-­‐9.   LORD,  B.  I.,  WOOLFORD,  L.  B.  &  MOLINEUX,  G.  2001.  Kinetics  of  neutrophil   production   in   normal   and   neutropenic   animals   during   the   response   to   filgrastim   (r-­‐metHu   G-­‐CSF)   or   filgrastim   SD/01   (PEG-­‐r-­‐metHu   G-­‐ CSF).  Clin  Cancer  Res,  7,  2085-­‐90.   LÖWENBERG,  B.,  VAN  PUTTEN,  W.,  THEOBALD,  M.,  GMÜR,  J.,  VERDONCK,  L.,   SONNEVELD,  P.,  FEY,  M.,  SCHOUTEN,  H.,  DE  GREEF,  G.,  FERRANT,  A.,   KOVACSOVICS,   T.,   GRATWOHL,   A.,   DAENEN,   S.,   HUIJGENS,   P.   &   BOOGAERTS,   M.   2003.   Effect   of   Priming   with   Granulocyte   Colony-­‐ Stimulating   Factor   on   the   Outcome   of   Chemotherapy   for   Acute   Myeloid  Leukemia.  New  England  Journal  of  Medicine,  349,  743-­‐752.   MAHMOOD,   I.   &   GREEN,   M.   D.   2005.   Pharmacokinetic   and   pharmacodynamic  considerations  in  the  development  of  therapeutic   proteins.  Clin  Pharmacokinet,  44,  331-­‐47.   MANZ,   M.   G.,   MIYAMOTO,   T.,   AKASHI,   K.   &   WEISSMAN,   I.   L.   2002.   Prospective   isolation   of   human   clonogenic   common   myeloid   progenitors.  Proc  Natl  Acad  Sci  U  S  A,  99,  11872-­‐7.   MARINARO,  J.  A.,  CASLEY,  D.  J.  &  BACH,  L.  A.  2000.  O-­‐glycosylation  delays  the   clearance  of  human  IGF-­‐binding  protein-­‐6  from  the  circulation.  Eur   J   Endocrinol,  142,  512-­‐6.   MARTINO,   M.,   LASZLO,   D.   &   LANZA,   F.   2014.   Long-­‐active   granulocyte   colony-­‐stimulating   factor   for   peripheral   blood   hematopoietic   progenitor  cell  mobilization.  Expert  Opin  Biol  Ther,  14,  757-­‐72.   MCCRACKEN,   S.,   LAYTON,   J.   E.,   SHORTER,   S.   C.,   STARKEY,   P.   M.,   BARLOW,   D.   H.   &   MARDON,   H.   J.   1996.   Expression   of   granulocyte-­‐colony   205   stimulating   factor   and   its   receptor   is   regulated   during   the   development  of  the  human  placenta.  J  Endocrinol,  149,  249-­‐58.   METCALF,   D.   2010.   The   colony-­‐stimulating   factors   and   cancer.   Nat   Rev   Cancer,  10,  425-­‐34.   MESHKIBAF,  S.  2015.  Granulocyte  colony-­‐stimulating  factor:  Its  role  in  gut-­‐ homing   macrophage   generation   and   colitis,   and   production   by   probiotics.  Electronic  Thesis  and  Dissertation  Repository.  Paper  2867.   MILLA,   P.,   DOSIO,   F.   &   CATTEL,   L.   2012.   PEGylation   of   proteins   and   liposomes:   a   powerful   and   flexible   strategy   to   improve   the   drug   delivery.  Curr  Drug  Metab,  13,  105-­‐19.   MOLINEUX,   G.   2004.   The   design   and   development   of   pegfilgrastim   (PEG-­‐ rmetHuG-­‐CSF,   neulasta((R))).   Current   Pharmaceutical   Design,   10,   1235-­‐1244.   MOLINEUX G.,   FOOTE,   M.,   ARVEDSON,   T.   2012.   Twenty   years   of   G-­‐CSF:   clinical   and   nonclinical   discoveries.  Springer  Science  &  Business  Media.   MORIKAWA,   K.,   MORIKAWA,   S.,   NAKAMURA,   M.   &   MIYAWAKI,   T.   2002.   Characterization   of   granulocyte   colony-­‐stimulating   factor   receptor   expressed  on  human  lymphocytes.  Br  J  Haematol,  118,  296-­‐304.   MORRIS,   K.   T.,   KHAN,   H.,   AHMAD,   A.,   WESTON,   L.   L.,   NOFCHISSEY,   R.   A.,   PINCHUK,   I.   V.   &   BESWICK,   E.   J.   2014.   G-­‐CSF   and   G-­‐CSFR   are   highly   expressed   in   human   gastric   and   colon   cancers   and   promote   carcinoma  cell  proliferation  and  migration.  Br  J  Cancer,  110,  1211-­‐20.   NAGATA,   S.,   TSUCHIYA,   M.,   ASANO,   S.,   KAZIRO,   Y.,   YAMAZAKI,   T.,   YAMAMOTO,  O.,  HIRATA,  Y.,  KUBOTA,  N.,  OHEDA,  M.,  NOMURA,  H.  &   ET   AL.   1986.   Molecular   cloning   and   expression   of   cDNA   for   human   granulocyte  colony-­‐stimulating  factor.  Nature,  319,  415-­‐8.   NATALELLO,  A.,  AMI,  D.,  COLLINI,  M.,  D'ALFONSO,  L.,  CHIRICO,  G.,  TONON,  G.,   SCARAMUZZA,  S.,  SCHREPFER,  R.  &  DOGLIA,  S.  M.  2012.  Biophysical   characterization   of   Met-­‐G-­‐CSF:   effects   of   different   site-­‐specific   mono-­‐ pegylations  on  protein  stability  and  aggregation.  PLoS  One,  7,  e42511.   NGUYEN-­‐JACKSON,   H.,   PANOPOULOS,   A.   D.,   ZHANG,   H.,   LI,   H.   S.   &   WATOWICH,   S.   S.   2010.   STAT3   controls   the   neutrophil   migratory   response   to   CXCR2   ligands   by   direct   activation   of   G-­‐CSF-­‐induced   CXCR2  expression  and  via  modulation  of  CXCR2  signal  transduction.   Blood,  115,  3354-­‐63.   NGUYEN-­‐JACKSON,  H.  T.,  LI,  H.  S.,  ZHANG,  H.,  OHASHI,  E.  &  WATOWICH,  S.  S.   2012.  G-­‐CSF-­‐activated  STAT3  enhances  production  of  the  chemokine   MIP-­‐2  in  bone  marrow  neutrophils.  J  Leukoc  Biol,  92,  1215-­‐25.   NICOLA,   N.   A.   &   METCALF,   D.   1984.   Binding   of   the   differentiation-­‐inducer,   granulocyte-­‐colony-­‐stimulating   factor,   to   responsive   but   not   unresponsive  leukemic  cell  lines.  Proc  Natl  Acad  Sci  U  S  A,  81,  3765-­‐9.   NICOLA,   N.   A.,   METCALF,   D.,   MATSUMOTO,   M.   &   JOHNSON,   G.   R.   1983.   Purification   of   a   factor   inducing   differentiation   in   murine   myelomonocytic  leukemia  cells.  Identification  as  granulocyte  colony-­‐ stimulating  factor.  J  Biol  Chem,  258,  9017-­‐23.   NIMUBONA,   S.,   GRULOIS,   I.,   BERNARD,   M.,   DRENOU,   B.,   GODARD,   M.,   FAUCHET,   R.   &   LAMY,   T.   2002.   Complete   remission   in   hypoplastic   acute   myeloid   leukemia   induced   by   G-­‐CSF   without   chemotherapy:   report  on  three  cases.  Leukemia,  16,  1871-­‐3.   206   NOMIYAMA,   J.,   SHINOHARA,   K.   &   INOUE,   H.   1994.   Erythema   nodosum   caused   by   the   administration   of   granulocyte   colony-­‐stimulating   factor  in  a  patient  with  refractory  anemia.  Am  J  Hematol,  47,  333.   O'MALLEY,   D.   P.,   WHALEN,   M.   &   BANKS,   P.   M.   2003.   Spontaneous   splenic   rupture   with   fatal   outcome   following   G-­‐CSF   administration   for   myelodysplastic  syndrome.  Am  J  Hematol,  73,  294-­‐5.   OBER,  R.  J.,  MARTINEZ,  C.,  LAI,  X.,  ZHOU,  J.  &  WARD,  E.  S.  2004.  Exocytosis  of   IgG   as   mediated   by   the   receptor,   FcRn:   an   analysis   at   the   single-­‐ molecule  level.  Proc  Natl  Acad  Sci  U  S  A,  101,  11076-­‐81.   PANOPOULOS,   A.   D.,   BARTOS,   D.,   ZHANG,   L.   &   WATOWICH,   S.   S.   2002.   Control   of   myeloid-­‐specific   integrin   alpha   Mbeta   2   (CD11b/CD18)   expression  by  cytokines  is  regulated  by  Stat3-­‐dependent  activation  of   PU.1.  J  Biol  Chem,  277,  19001-­‐7.   PANOPOULOS,   A.   D.   &   WATOWICH,   S.   S.   2008.   Granulocyte   colony-­‐ stimulating   factor:   molecular   mechanisms   of   action   during   steady   state  and  'emergency'  hematopoiesis.  Cytokine,  42,  277-­‐88.   PATEL,   A.,   CHOLKAR,   K.   &   MITRA,   A.   K.   2014.   Recent   developments   in   protein   and   peptide   parenteral   delivery   approaches.   Ther   Deliv,   5,   337-­‐65.   PELUS,   L.   M.,   HOROWITZ,   D.,   COOPER,   S.   C.   &   KING,   A.   G.   2002.   Peripheral   blood   stem   cell   mobilization.   A   role   for   CXC   chemokines.   Crit   Rev   Oncol  Hematol,  43,  257-­‐75.   PERLMAN,   S.,   VAN   DEN   HAZEL,   B.,   CHRISTIANSEN,   J.,   GRAM-­‐NIELSEN,   S.,   JEPPESEN,   C.   B.,   ANDERSEN,   K.   V.,   HALKIER,   T.,   OKKELS,   S.   &   SCHAMBYE,   H.   T.   2003.   Glycosylation   of   an   N-­‐terminal   extension   prolongs   the   half-­‐life   and   increases   the   in   vivo   activity   of   follicle   stimulating  hormone.  J  Clin  Endocrinol  Metab,  88,  3227-­‐35.   PERUTZ,   M.   F.   1995.   Glutamine   repeats   as   polar   zippers:   their   role   in   inherited  neurodegenerative  disease.  Mol  Med,  1,  718-­‐21.   PICARD,   V.,   ERSDAL-­‐BADJU,   E.   &   BOCK,   S.   C.   1995.   Partial   glycosylation   of   antithrombin  III  asparagine-­‐135  is  caused  by  the  serine  in  the  third   position  of  its  N-­‐glycosylation  consensus  sequence  and  is  responsible   for   production   of   the   beta-­‐antithrombin   III   isoform   with   enhanced   heparin  affinity.  Biochemistry,  34,  8433-­‐40.   PISAL,   D.   S.,   KOSLOSKI,   M.   P.   &   BALU-­‐IYER,   S.   V.   2010.   Delivery   of   therapeutic   proteins.   Journal   of   Pharmaceutical   Sciences,   99,   2557-­‐ 2575.   PRENDIVILLE,   J.,   THIESSEN,   P.   &   MALLORY,   S.   B.   2001.   Neutrophilic   dermatoses  in  two  children  with  idiopathic  neutropenia:  association   with   granulocyte   colony-­‐stimulating   factor   (G-­‐CSF)   therapy.   Pediatr   Dermatol,  18,  417-­‐21.   PUSIC,  I.  &  DIPERSIO,  J.  F.  2008.  The  use  of  growth  factors  in  hematopoietic   stem   cell   transplantation.   Current   Pharmaceutical   Design,   14,   1950-­‐ 1961.   QIU,  H.,  BELANGER,  A.,  YOON,  H.  W.  &  BUNN,  H.  F.  1998.  Homodimerization   restores  biological  activity  to  an  inactive  erythropoietin  mutant.  J  Biol   Chem,  273,  11173-­‐6.   ROBERTS,  A.  W.  2005.  G-­‐CSF:  a  key  regulator  of  neutrophil  production,  but   that's  not  all!  Growth  Factors,  23,  33-­‐41.   207   SAINT-­‐JORE-­‐DUPAS,   C.,   FAYE,   L.   &   GOMORD,   V.   2007.   From   planta   to   pharma   with   glycosylation   in   the   toolbox.   Trends  Biotechnol,   25,   317-­‐ 23.   SAMPSON,  M.,  ZHU,  Q.  S.  &  COREY,  S.  J.  2007.  Src  kinases  in  G-­‐CSF  receptor   signaling.  Front  Biosci,  12,  1463-­‐74.   SCHNEIDER,  A.,  KRUGER,  C.,  STEIGLEDER,  T.,  WEBER,  D.,  PITZER,  C.,  LAAGE,   R.,   ARONOWSKI,   J.,   MAURER,   M.   H.,   GASSLER,   N.,   MIER,   W.,   HASSELBLATT,   M.,   KOLLMAR,   R.,   SCHWAB,   S.,   SOMMER,   C.,   BACH,   A.,   KUHN,  H.  G.  &  SCHABITZ,  W.  R.  2005.  The  hematopoietic  factor  G-­‐CSF   is   a   neuronal   ligand   that   counteracts   programmed   cell   death   and   drives  neurogenesis.  J  Clin  Invest,  115,  2083-­‐98.   SEMERAD,  C.  L.,  LIU,  F.,  GREGORY,  A.  D.,  STUMPF,  K.  &  LINK,  D.  C.  2002.  G-­‐ CSF  is  an  essential  regulator  of  neutrophil  trafficking  from  the  bone   marrow  to  the  blood.  Immunity,  17,  413-­‐23.   SHAH,   J.   &   WELSH,   S.   J.   2014.   The   clinical   use   of   granulocyte-­‐colony   stimulating  factor.  Br  J  Hosp  Med  (Lond),  75,  C29-­‐32.   SINCLAIR,   A.   M.   &   ELLIOTT,   S.   2005.   Glycoengineering:   the   effect   of   glycosylation   on   the   properties   of   therapeutic   proteins.   J  Pharm  Sci,   94,  1626-­‐35.   SIVAKUMAR,   R.,   ATKINSON,   M.   A.,   MATHEWS,   C.   E.   &   MORE,   L.   2015.   G-­‐CSF:   A  Friend  or  Foe?  Immunome  Research,  2015.   SOLA,  R.  J.  &  GRIEBENOW,  K.  2009.  Effects  of  glycosylation  on  the  stability  of   protein  pharmaceuticals.  J  Pharm  Sci,  98,  1223-­‐45.   SOLA,  R.  J.  &  GRIEBENOW,  K.  2010.  Glycosylation  of  therapeutic  proteins:  an   effective  strategy  to  optimize  efficacy.  BioDrugs,  24,  9-­‐21.   SOLA,   R.   J.,   RODRIGUEZ-­‐MARTINEZ,   J.   A.   &   GRIEBENOW,   K.   2007.   Modulation   of   protein   biophysical   properties   by   chemical   glycosylation:  biochemical  insights  and  biomedical  implications.  Cell   Mol  Life  Sci,  64,  2133-­‐52.   SOUZA,   L.   M.,   BOONE,   T.   C.,   GABRILOVE,   J.,   LAI,   P.   H.,   ZSEBO,   K.   M.,   MURDOCK,  D.  C.,  CHAZIN,  V.  R.,  BRUSZEWSKI,  J.,  LU,  H.,  CHEN,  K.  K.  &   ET   AL.   1986.   Recombinant   human   granulocyte   colony-­‐stimulating   factor:  effects  on  normal  and  leukemic  myeloid  cells.  Science,  232,  61-­‐ 5.   SPIRO,   R.   G.   2002.   Protein   glycosylation:   nature,   distribution,   enzymatic   formation,   and   disease   implications   of   glycopeptide   bonds.   Glycobiology,  12,  43R-­‐56R.   STORK,   R.,   ZETTLITZ,   K.   A.,   MULLER,   D.,   RETHER,   M.,   HANISCH,   F.   G.   &   KONTERMANN,   R.   E.   2008.   N-­‐glycosylation   as   novel   strategy   to   improve   pharmacokinetic   properties   of   bispecific   single-­‐chain   diabodies.  J  Biol  Chem,  283,  7804-­‐12.   SYTKOWSKI,   A.   J.,   LUNN,   E.   D.,   RISINGER,   M.   A.   &   DAVIS,   K.   L.   1999.   An   erythropoietin   fusion   protein   comprised   of   identical   repeating   domains   exhibits   enhanced   biological   properties.   J   Biol   Chem,   274,   24773-­‐8.   TAMADA,   T.,   HONJO,   E.,   MAEDA,   Y.,   OKAMOTO,   T.,   ISHIBASHI,   M.,   TOKUNAGA,   M.   &   KUROKI,   R.   2006.   Homodimeric   cross-­‐over   structure  of  the  human  granulocyte  colony-­‐stimulating  factor  (GCSF)   receptor  signaling  complex.  Proc  Natl  Acad  Sci  U  S  A,  103,  3135-­‐40.   208   TANAKA,  H.,  SATAKE-­‐ISHIKAWA,  R.,  ISHIKAWA,  M.,  MATSUKI,  S.  &  ASANO,   K.   1991.   Pharmacokinetics   of   recombinant   human   granulocyte   colony-­‐stimulating   factor   conjugated   to   polyethylene   glycol   in   rats.   Cancer  Res,  51,  3710-­‐4.   TANG,   L.,   PERSKY,   A.   M.,   HOCHHAUS,   G.   &   MEIBOHM,   B.   2004.   Pharmacokinetic   aspects   of   biotechnology   products.   J  Pharm  Sci,   93,   2184-­‐204.   TANHEHCO,  Y.  C.,  ADAMSKI,  J.,  SELL,  M.,  CUNNINGHAM,  K.,  EISENMANN,  C.,   MAGEE,   D.,   STADTMAUER,   E.   A.   &   O'DOHERTY,   U.   2010.   Plerixafor   mobilization  leads  to  a  lower  ratio  of  CD34+  cells  to  total  nucleated   cells   which   results   in   greater   storage   costs.   Journal   of   Clinical   Apheresis,  25,  202-­‐208.   TOUW,   I.   P.   &   VAN   DE   GEIJN,   G.   J.   2007.   Granulocyte   colony-­‐stimulating   factor  and  its  receptor  in  normal  myeloid  cell  development,  leukemia   and  related  blood  cell  disorders.  Front  Biosci,  12,  800-­‐15.   TRAINER,   P.   J.,   DRAKE,   W.   M.,   KATZNELSON,   L.,   FREDA,   P.   U.,   HERMAN-­‐ BONERT,  V.,  VAN  DER  LELY,  A.  J.,  DIMARAKI,  E.  V.,  STEWART,  P.  M.,   FRIEND,   K.   E.,   VANCE,   M.   L.,   BESSER,   G.   M.,   THORNER,   M.   O.,   PARKINSON,   C.,   KLIBANSKI,   A.,   POWELL,   J.   S.,   BARKAN,   A.   L.,   SHEPPARD,   M.   C.,   MALDONADO,   M.,   ROSE,   D.   R.,   CLEMMONS,   D.   R.,   JOHANNSSON,   G.,   BENGT-­‐ÅKE   BENGTSSON,   B.-­‐Å.,   STAVROU,   S.,   KLEINBERG,   D.   L.,   COOK,   D.   M.,   PHILLIPS,   L.   S.,   BIDLINGMAIER,   M.,   STRASBURGER,  C.  J.,  HACKETT,  S.,  ZIB,  K.,  BENNETT,  W.  F.,  DAVIS,  R.  J.   &   SCARLETT,   J.   A.   2000.   Treatment   of   Acromegaly   with   the   Growth   Hormone–Receptor  Antagonist  Pegvisomant.  New   England   Journal   of   Medicine,  342,  1171-­‐1177.   ULICH,  T.  R.,  DEL  CASTILLO,  J.  &  SOUZA,  L.  1988.  Kinetics  and  mechanisms   of   recombinant   human   granulocyte-­‐colony   stimulating   factor-­‐ induced  neutrophilia.  Am  J  Pathol,  133,  630-­‐8.   VALTIERI,   M.,   BOCCOLI,   G.,   TESTA,   U.,   BARLETTA,   C.   &   PESCHLE,   C.   1991.   Two-­‐step   differentiation   of   AML-­‐193   leukemic   line:   terminal   maturation   is   induced   by   positive   interaction   of   retinoic   acid   with   granulocyte   colony-­‐stimulating   factor   (CSF)   and   vitamin   D3   with   monocyte  CSF.  Blood,  77,  1804-­‐12.   VARDIMAN,   J.   W.,   THIELE,   J.,   ARBER,   D.   A.,   BRUNNING,   R.   D.,   BOROWITZ,   M.   J.,   PORWIT,   A.,   HARRIS,   N.   L.,   LE   BEAU,   M.   M.,   HELLSTROM-­‐ LINDBERG,   E.,   TEFFERI,   A.   &   BLOOMFIELD,   C.   D.   2009.   The   2008   revision   of   the   World   Health   Organization   (WHO)   classification   of   myeloid   neoplasms   and   acute   leukemia:   rationale   and   important   changes.  Blood,  114,  937-­‐51.   VARKI,  A.  2008.  Sialic  acids  in  human  health  and  disease.  Trends  Mol  Med,  14,   351-­‐60.   WADA,  H.,  YOSHIDA,  S.,  ISHIBASHI,  F.,  MIZOBUCHI,  T.,  MORIYA,  Y.,  HOSHINO,   H.,   OKAMOTO,   T.,   SUZUKI,   M.,   SHIBUYA,   K.   &   YOSHINO,   I.   2011.   Pleomorphic   carcinoma   of   the   lung   producing   granulocyte   colony-­‐ stimulating  factor:  report  of  a  case.  Surg  Today,  41,  1161-­‐5.   WALKER,  A.  K.,  SOO,  K.  Y.,  LEVINA,  V.,  TALBO,  G.  H.  &  ATKIN,  J.  D.  2013.  N-­‐ linked   glycosylation   modulates   dimerization   of   protein   disulfide   isomerase  family  A  member  2  (PDIA2).  FEBS  J,  280,  233-­‐43.   209   WEENEN,  C.,  PENA,  J.  E.,  POLLAK,  S.  V.,  KLEIN,  J.,  LOBEL,  L.,  TROUSDALE,  R.   K.,  PALMER,  S.,  LUSTBADER,  E.  G.,  OGDEN,  R.  T.  &  LUSTBADER,  J.  W.   2004.  Long-­‐acting  follicle-­‐stimulating  hormone  analogs  containing  N-­‐ linked  glycosylation  exhibited  increased  bioactivity  compared  with  o-­‐ linked  analogs  in  female  rats.  J  Clin  Endocrinol  Metab,  89,  5204-­‐12.   WELTE,  K.,  PLATZER,  E.,  LU,  L.,  GABRILOVE,  J.  L.,  LEVI,  E.,  MERTELSMANN,  R.   &  MOORE,  M.  A.  1985.  Purification  and  biochemical  characterization   of   human   pluripotent   hematopoietic   colony-­‐stimulating   factor.   Proc   Natl  Acad  Sci  U  S  A,  82,  1526-­‐30.   WERNER,  J.  M.,  BREEZE,  A.  L.,  KARA,  B.,  ROSENBROCK,  G.,  BOYD,  J.,  SOFFE,  N.   &   CAMPBELL,   I.   D.   1994.   Secondary   structure   and   backbone   dynamics  of  human  granulocyte  colony-­‐stimulating  factor  in  solution.   Biochemistry,  33,  7184-­‐92.   WHITE,   S.   M.,   ALARCON,   M.   H.   &   TWEARDY,   D.   J.   2000.   Inhibition   of   granulocyte   colony-­‐stimulating   factor-­‐mediated   myeloid   maturation   by   low   level   expression   of   the   differentiation-­‐defective   class   IV   granulocyte   colony-­‐stimulating   factor   receptor   isoform.   Blood,   95,   3335-­‐40.   WHITE,  S.  M.,  BALL,  E.  D.,  EHMANN,  W.  C.,  RAO,  A.  S.  &  TWEARDY,  D.  J.  1998.   Increased   expression   of   the   differentiation-­‐defective   granulocyte   colony-­‐stimulating   factor   receptor   mRNA   isoform   in   acute   myelogenous  leukemia.  Leukemia,  12,  899-­‐906.   WILKINSON,  I.  R.,  FERRANDIS,  E.,  ARTYMIUK,  P.  J.,  TEILLOT,  M.,  SOULARD,   C.,  TOUVAY,  C.,  PRADHANANGA,  S.  L.,  JUSTICE,  S.,  WU,  Z.,  LEUNG,  K.  C.,   STRASBURGER,   C.   J.,   SAYERS,   J.   R.   &   ROSS,   R.   J.   2007.   A   ligand-­‐ receptor   fusion   of   growth   hormone   forms   a   dimer   and   is   a   potent   long-­‐acting  agonist.  Nat  Med,  13,  1108-­‐13.   WONGTRAKUL-­‐KISH,  K.,  KOLARICH,  D.,  PASCOVICI,  D.,  JOSS,  J.  L.,  DEANE,  E.   &   PACKER,   N.   H.   2012.   Characterization   of   N-­‐   and   O-­‐linked   glycosylation   changes   in   milk   of   the   tammar   wallaby   (Macropus   eugenii)  over  lactation.  Glycoconj  J.   WURM,   F.   M.   2004.   Production   of   recombinant   protein   therapeutics   in   cultivated  mammalian  cells.  Nat  Biotechnol,  22,  1393-­‐8.   YAMAGUCHI,   T.,   MIYATA,   Y.,   HAYAMIZU,   K.,   HASHIZUME,   J.,   MATSUMOTO,   T.,   TASHIRO,   H.   &   OHDAN,   H.   2012.   Preventive   effect   of   G-­‐CSF   on   acute   lung   injury   via   alveolar   macrophage   regulation.   J  Surg  Res,   178,   378-­‐84.   YE,   P.,   RODRIGUEZ,   F.   H.,   KANALY,   S.,   STOCKING,   K.   L.,   SCHURR,   J.,   SCHWARZENBERGER,  P.,  OLIVER,  P.,  HUANG,  W.,  ZHANG,  P.,  ZHANG,   J.,  SHELLITO,  J.  E.,  BAGBY,  G.  J.,  NELSON,  S.,  CHARRIER,  K.,  PESCHON,  J.   J.   &   KOLLS,   J.   K.   2001.   Requirement   of   interleukin   17   receptor   signaling  for  lung  CXC  chemokine  and  granulocyte  colony-­‐stimulating   factor   expression,   neutrophil   recruitment,   and   host   defense.   J   Exp   Med,  194,  519-­‐27.   YOON,  S.  K.,  SONG,  J.  Y.  &  LEE,  G.  M.  2003.  Effect  of  low  culture  temperature   on   specific   productivity,   transcription   level,   and   heterogeneity   of   erythropoietin  in  Chinese  hamster  ovary  cells.  Biotechnol  Bioeng,  82,   289-­‐98.   210   YOWELL,   S.   L.   &   BLACKWELL,   S.   2002.   Novel   effects   with   polyethylene   glycol  modified  pharmaceuticals.  Cancer  treatment  reviews,  28,  3-­‐6.   ZHANG,  F.,  LIU,  M.  R.  &  WAN,  H.  T.  2014.  Discussion  about  several  potential   drawbacks   of   PEGylated   therapeutic   proteins.   Biol   Pharm   Bull,   37,   335-­‐9.   ZHAO,   S.,   ZHANG,   Y.,   TIAN,   H.,   CHEN,   X.,   CAI,   D.,   YAO,   W.   &   GAO,   X.   2013.   Extending  the  serum  half-­‐life  of  G-­‐CSF  via  fusion  with  the  domain  III   of  human  serum  albumin.  Biomed  Res  Int,  2013,  107238.       211