Graduate Theses and Dissertations Graduate School 11-4 ...

100
University of South Florida Scholar Commons Graduate eses and Dissertations Graduate School 11-4-2014 Cloning and Characterization of IL-1β, IL-8, IL-10, and TNFα from Golden Tilefish (Lopholatilus chamaeleonticeps) and Red Snapper (Lutjanus campechanus) Kristina L. Deak University of South Florida, [email protected] Follow this and additional works at: hps://scholarcommons.usf.edu/etd Part of the Genetics Commons , and the Oceanography Commons is esis is brought to you for free and open access by the Graduate School at Scholar Commons. It has been accepted for inclusion in Graduate eses and Dissertations by an authorized administrator of Scholar Commons. For more information, please contact [email protected]. Scholar Commons Citation Deak, Kristina L., "Cloning and Characterization of IL-1β, IL-8, IL-10, and TNFα from Golden Tilefish (Lopholatilus chamaeleonticeps) and Red Snapper (Lutjanus campechanus)" (2014). Graduate eses and Dissertations. hps://scholarcommons.usf.edu/etd/5416

Transcript of Graduate Theses and Dissertations Graduate School 11-4 ...

Page 1: Graduate Theses and Dissertations Graduate School 11-4 ...

University of South FloridaScholar Commons

Graduate Theses and Dissertations Graduate School

11-4-2014

Cloning and Characterization of IL-1β, IL-8, IL-10,and TNFα from Golden Tilefish (Lopholatiluschamaeleonticeps) and Red Snapper (Lutjanuscampechanus)Kristina L. DeakUniversity of South Florida, [email protected]

Follow this and additional works at: https://scholarcommons.usf.edu/etd

Part of the Genetics Commons, and the Oceanography Commons

This Thesis is brought to you for free and open access by the Graduate School at Scholar Commons. It has been accepted for inclusion in GraduateTheses and Dissertations by an authorized administrator of Scholar Commons. For more information, please contact [email protected].

Scholar Commons CitationDeak, Kristina L., "Cloning and Characterization of IL-1β, IL-8, IL-10, and TNFα from Golden Tilefish (Lopholatiluschamaeleonticeps) and Red Snapper (Lutjanus campechanus)" (2014). Graduate Theses and Dissertations.https://scholarcommons.usf.edu/etd/5416

Page 2: Graduate Theses and Dissertations Graduate School 11-4 ...

         

Cloning  and  Characterization  of  IL-­‐1β,  IL-­‐8,  IL-­‐10,  and  TNFα  from  Golden  Tilefish  (Lopholatilus      

chamaeleonticeps)  and  Red  Snapper  (Lutjanus  campechanus)        by        

Kristina  L.  Deak    

     

A  thesis  submitted  in  partial  fulfillment    of  the  requirements  for  the  degree  of  

Master  of  Science  with  a  concentration  in  Biological  Oceanography  

Department  of  Marine  Science  College  of  Marine  Science  University  of  South  Florida  

     

Major  Professor:  Steve  A.  Murawski,  Ph.D.  Cathy  J.  Walsh,  Ph.D.  Dana  L.  Wetzel,  Ph.D.  

   

Date  of  Approval:  November  4,  2014  

     

Keywords:  Genetics,  Immunology,  Oil  Spill,  Deepwater  Horizon      

Copyright  ©  2014,  Kristina  L.  Deak            

Page 3: Graduate Theses and Dissertations Graduate School 11-4 ...

     

   

DEDICATION      

  This  thesis  is  dedicated  to  all  those  affected  by  the  Deepwater  Horizon  explosion  of  April  20,  

2010,  including  the  11  men  who  lost  their  lives  that  day.    I  hope  that  we  will  continue  to  learn  from  this  

tragedy  and  work  towards  efforts  to  mitigate  the  ecological  impact  of  future  spills.    

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 4: Graduate Theses and Dissertations Graduate School 11-4 ...

         

ACKNOWLEDGMENTS    

  Funding  for  this  work  was  provided  by  a  grant  from  the  BP/Gulf  of  Mexico  Research  Initiative,  C-­‐

IMAGE,  the  Fish  Florida  scholarship,  and  the  Guy  Harvey  Ocean  Foundation  scholarship.    

  I  would  like  to  express  my  most  sincere  gratitude  to  my  advisor  and  mentor,  Dr.  Steve  

Murawski.    Working  with  Dr.  Murawski  has  been  an  incredible  experience,  and  has  taught  me  a  great  

deal  about  fish,  fishing,  professional  development,  and  how  to  effectively  communicate  science.    I  look  

forward  to  many  more  summers  at  sea,  especially  catching  that  obligatory  last  set  tiger.  

  Thank  you  to  many  members  of  Mote  Marine  Laboratory,  including  my  committee  members  Dr.  

Cathy  Walsh  and  Dr.  Dana  Wetzel.    Dr.  Walsh’s  years  of  marine  immunology  and  biochemistry  

experience  were  an  incredible  asset  to  my  research,  and  I  appreciate  her  assistance  with  

troubleshooting  and  data  analysis.    Dr.  Wetzel  provided  endless  amounts  of  lab  space,  reagents,  and  

flexibility  as  I  juggled  my  commitments  as  a  chemist  in  her  lab  and  a  student  at  USF.      Dr.  Tracy  

Sherwood  was  a  valuable  resource  for  method  optimization  and  data  interpretation.    Finally,  I  thank  Dr.  

Carl  Luer,  for  use  of  his  laboratory  instrumentation  throughout  this  project.  

  A  world  of  credit  is  due  to  (doctoral  candidate!)  Erin  Pulster.    She  has  been  absolutely  

instrumental  to  my  development  as  a  scientist,  and  working  with  her,  first  as  an  intern,  then  as  a  co-­‐

worker  at  Mote,  has  opened  so  many  doors.    Her  daily  assistance  has  significantly  and  positively  laid  the  

foundation  for  my  career.      Thank  you,  for  the  thousands  of  pieces  of  advice,  millions  of  questions  

answered,  and  countless  laughs  over  the  years.    I  am  excited  for  all  of  our  future  collaborations.  

  Finally,  I  am  grateful  for  the  unwavering  support  from  Jesse  Hoemann,  Amy  Wallace,  the  ladies  

of  the  Murawski  lab,  and  my  parents.    I  truly  appreciate  all  of  your  assistance  throughout  this  thesis…  

now  let’s  rally  for  the  PhD!

Page 5: Graduate Theses and Dissertations Graduate School 11-4 ...

  i  

         

TABLE  OF  CONTENTS    List  of  Tables   iii    List  of  Figures     iv    Abstract   viii    Chapter  One:  Introduction     1  

1.1 Deepwater  Horizon,  PAH  immunotoxcity,  and  traditional  biomarkers  for  PAH    exposure   1  

  1.2  Teleost  cytokines  and  the  immune  response   2     1.3  Characteristics  of  vertebrate  IL-­‐1β,  IL-­‐8,  IL-­‐10,  and  TNFα   4     1.4  Objectives   7    Chapter  Two:  Methods   10     2.1  Isolation  of  total  RNA   10     2.2  cDNA  production   11     2.3  Initial  identification  and  cloning  of  IL-­‐1β,  IL-­‐8,  IL-­‐10,  and  TNFα  fragments   11     2.4  5’-­‐  and  3’-­‐RACE   13     2.5  Sequence  confirmation   14     2.6  Analysis  of  confirmed  protein  coding  sequence   14    Chapter  Three:  Results  and  Discussion   16     3.1  Cloning  and  sequence  analysis  of  golden  tilefish  IL-­‐1β   16     3.2  Cloning  and  sequence  analysis  of  red  snapper  IL-­‐1β   19     3.3  Cloning  and  sequence  analysis  of  golden  tilefish  IL-­‐8   19     3.4  Cloning  and  sequence  analysis  of  red  snapper  IL-­‐8   21     3.5  Cloning  and  sequence  analysis  of  golden  tilefish  IL-­‐10   22     3.6  Cloning  and  sequence  analysis  of  red  snapper  IL-­‐10   23     3.7  Cloning  and  sequence  analysis  of  golden  tilefish  TNFα   25     3.8  Cloning  and  sequence  analysis  of  red  snapper  TNFα   26    Chapter  Four:  Conclusions  and  Future  Work   51     4.1  Conclusions   51     4.2  Future  work   52    References     53    Appendix  A:  Map  of  2013  Tissue  Sampling  Stations   63    Appendix  B:  2013  Tissue  Sampling  Location  Data   64  

Page 6: Graduate Theses and Dissertations Graduate School 11-4 ...

  ii  

Appendix  C:  Golden  Tilefish  and  Red  Snapper  Biometrics  Data   65    Appendix  D:  Extended  PCR  Parameters   66    Appendix  E:  Experimentally  Obtained  Nucleotide  Sequences   67    Appendix  F:  Experimentally  Derived  Cytokine  Amino  Acid  Sequences   73        Appendix  G:  Sequence  Confirmation     74    Appendix  H:  Table  of  Common  Names  and  Their  Associated  Scientific  Names   87                                        

                                                                   

Page 7: Graduate Theses and Dissertations Graduate School 11-4 ...

  iii  

     

   

LIST  OF  TABLES    Table  1:   Oligonucleotide  primers  used  throughout  this  study,  including  their  purpose,   15     annealing  temperature,  expected  fragment  band  size  in  number  of  base  pairs  (bp),       and  primer  source.        Table  2:   Percent  amino  acid  identity  between  confirmed,  mature  amino  acid  sequence  and   31     the  top  five  most  homologous  fish  species,  according  to  BLASTp  analysis.            Table  3:   Percent  amino  acid  identity  among  vertebrate  groups.   32    Table  B1:   2013  longline  tissue  sampling  location  data.   64    Table  C1:   Golden  tilefish  and  red  snapper  biometrics  data,  specific  to  fish  whose  spleens  were     65     used  in  this  study.    Table  D1:   Extended  PCR  parameters  for  primers  used.   66    Table  F1:   Amino  acid  sequences  for  each  cytokine,  translated  from  the  experimentally     73     obtained  nucleotide  sequence  by  the  ExPASy  Translate  tool.      Table  H1:   List  of  all  common  names  used  in  this  thesis  with  the  associated  scientific  name.   87                                            

Page 8: Graduate Theses and Dissertations Graduate School 11-4 ...

  iv  

         

LIST  OF  FIGURES    

Figure  1:   Schematic  of  PCR  results  including  locations  of  the  coding  region,  5’  UTR,  3’  UTR,     28     locations  at  which  primers  bound,  and  the  size  and  position  of  the  resulting       fragments.          Figure  2:   Clustal  W  multiple  sequence  alignment  of  representative  vertebrate  IL-­‐1β,  shaded   29     with  BOXSHADE.        Figure  3:   Multiple  sequence  alignment  of  golden  tilefish  IL-­‐1β  with  the  top  five  most     33     homologous  species,  according  to  BLASTP.        Figure  4:     Rooted  phylogenetic  tree  of  vertebrate  IL-­‐1β  proteins,  produced  in  MEGA  6  with  a     34     NJ  method.        Figure  5:     Clustal  Omega  alignment  between  golden  tilefish  and  red  snapper  IL-­‐1β.       35    Figure  6:   Clustal  W  multiple  sequence  alignment  of  representative  vertebrate  IL-­‐8,  shaded   36     with  BOXSHADE.    Figure  7:     Multiple  sequence  alignment  of  golden  tilefish  IL-­‐8  with  the  top  five  most     37     homologous  species,  according  to  BLASTP.        Figure  8:   Rooted  phylogenetic  tree  of  vertebrate  IL-­‐8  proteins,  produced  in  MEGA  6  with     38     a  NJ  method.      Figure  9:   Multiple  sequence  alignment  of  red  snapper  IL-­‐8  with  the  top  five  most       39     homologous  species,  according  to  BLASTP.        Figure  10:   Clustal  Omega  alignment  between  golden  tilefish  and  red  snapper  IL-­‐8.       40    Figure  11:     Clustal  W  multiple  sequence  alignment  of  representative  vertebrate  IL-­‐10,  shaded   41     with  BOXSHADE.    Figure  12:     Clustal  Omega  alignment  between  red  snapper  and  golden  tilefish  IL-­‐10.   43    Figure  13:     Multiple  sequence  alignment  of  red  snapper  IL-­‐10  with  the  top  five  most     44     homologous  species,  according  to  BLASTP.        Figure  14:     Rooted  phylogenetic  tree  of  vertebrate  IL-­‐10  proteins,  produced  in  MEGA  6  with  a     45     NJ  method.      

Page 9: Graduate Theses and Dissertations Graduate School 11-4 ...

  v  

Figure  15:     Multiple  sequence  alignment  of  representative  vertebrate  TNFα,  shaded  with   46     BOXSHADE.    Figure  16:     Clustal  Omega  alignment  between  red  snapper  and  golden  tilefish  TNFα   48    Figure  17:     Multiple  sequence  alignment  of  red  snapper  TNFα  with  the  top  five  most     49     homologous  species,  according  to  BLASTP.        Figure  18:   Rooted  phylogenetic  tree  of  vertebrate  TNFα  proteins,  produced  in  MEGA  6  with     50     a  NJ  method.    Figure  A1:     2013  longline  stations  aboard  the  R/V  Weatherbird  II  yielding  tissue  samples.     63    Figure  E1:    Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  golden  tilefish  IL-­‐1β,     67     derived  from  the  assembled  5’-­‐  and  3’-­‐RACE  products.    Figure  E2:   Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  golden  tilefish     68     IL-­‐1β,  obtained  via  cloning  from  two  different  fish.          Figure  E3:   Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐1β,     68     derived  from  the  partial  cloned  fragment.        Figure  E4:   Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  golden  tilefish  IL-­‐8,     68     derived  from  the  3’-­‐RACE  product.        Figure  E5:   Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  golden  tilefish  IL-­‐8,     69     obtained  via  cloning  from  two  different  fish.        Figure  E6:     Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐8,  derived     69     from  the  assembled  5’-­‐  and  3’-­‐RACE  products.        Figure  E7:   Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  red  snapper  IL-­‐8,     69     obtained  via  cloning  from  two  different  fish.        Figure  E8:   Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  golden  tilefish  IL-­‐10,     70     derived  from  the  assembled  5’-­‐RACE  and  initial  fragment  products.        Figure  E9:   Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐10,  derived     70     from  the  assembled  5’-­‐  and  3’-­‐RACE  products.        Figure  E10:   Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  red  snapper  IL-­‐10,     71     obtained  via  cloning  from  two  different  fish.        Figure  E11:   Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  golden  tilefish  TNFα,     71     derived  from  the  assembled  3’-­‐RACE  and  initial  fragment  products.          

Page 10: Graduate Theses and Dissertations Graduate School 11-4 ...

  vi  

Figure  E12:   Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  red  snapper  TNFα,     72     derived  from  the  assembled  5’-­‐  and  3’-­‐RACE  products.        Figure  E13:   Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  red  snapper  TNFα,   72      obtained  via  cloning  from  two  different  fish.        Figure  G1:   5’  to  3’  nucleotide  sequence  for  the  translated  region  of  golden  tilefish  IL-­‐1β,     75     cloned  from  fish  F.    Figure  G2:   5’  to  3’  nucleotide  sequence  for  the  translated  region  of  golden  tilefish  IL-­‐1β,     75     cloned  from  fish  H.    Figure  G3:     Clustal  Omega  alignment  of  the  nucleotide  sequences  for  golden  tilefish  IL-­‐1β   76     from  fish  F  and  H.    Figure  G4:     Clustal  Omega  alignment  of  the  protein  sequences  for  golden  tilefish  IL-­‐1β  from     77     fish  F  and  H.    Figure  G5:   Clustal  Omega  alignment  of  golden  tilefish  IL-­‐1β  from  fishes  F  and  H  and  the  top     77     five  homologous  species,  according  to  BLASTp.    Figure  G6:   5’  to  3’  nucleotide  sequence  for  the  translated  region  of  golden  tilefish  IL-­‐8,     78     obtained  from  fish  F.    Figure  G7:   5’  to  3’  nucleotide  sequence  for  the  translated  region  of  golden  tilefish  IL-­‐8,   78     obtained  from  fish  H.    Figure  G8:   Clustal  Omega  alignment  of  the  nucleotide  sequences  for  golden  tilefish  IL-­‐8  from   79     fish  F  and  H.    Figure  G9:   Clustal  Omega  alignment  of  the  protein  sequences  for  golden  tilefish  IL-­‐8  from  fish     79     F  and  H.    Figure  G10:  5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐8,  obtained  from  fish  C.     80    Figure  G11:  5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐8,  obtained  from  fish  E.   80    Figure  G12:  Clustal  Omega  alignment  of  the  nucleotide  sequences  for  red  snapper  IL-­‐8  from     80     fish  C  and  E.    Figure  G13:  Clustal  Omega  alignment  of  the  protein  sequences  for  red  snapper  IL-­‐8  from  fish   81     C  and  E.    Figure  G14:  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  snapper  IL-­‐10,  cloned       81     from  fish  C.      Figure  G15:  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  snapper  IL-­‐10,  cloned   81     from  fish  E.  

Page 11: Graduate Theses and Dissertations Graduate School 11-4 ...

  vii  

 Figure  G16:    Clustal  Omega  alignment  of  the  nucleotide  sequences  for  red  snapper  IL-­‐10   82     from  fish  C  and  E.    Figure  G17:  Clustal  Omega  alignment  of  the  protein  sequences  for  red  snapper  IL-­‐10  from     83     fish  C  and  E.  

 Figure  G18:  Clustal  Omega  alignment  of  red  snapper  IL-­‐10  from  fish  C  and  E  and  the  top     83     five  homologous  species,  according  to  BLASTp.    Figure  G19:  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  red  snapper  TNFα,     84     cloned  from  fish  C.    Figure  G20:  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  red  snapper  TNFα,     84     cloned  from  fish  E.    Figure  G21:  Clustal  Omega  alignment  of  the  nucleotide  sequences  for  red  snapper  TNFα   85     from  fish  C  and  E.    Figure  G22:  Clustal  Omega  alignment  of  the  protein  sequences  for  red  snapper  TNFα  from   86     fish  C  and  E.        

     

                                         

Page 12: Graduate Theses and Dissertations Graduate School 11-4 ...

  viii  

         

ABSTRACT    

  Cytokines  are  pleiotropic  and  redundant  signaling  molecules  that  govern  the  inflammatory  

response  and  immunity,  a  critical  ecological  parameter  for  organism  success  and  population  growth.    

Produced  at  the  site  of  injury  or  pathogen  intrusion  by  a  variety  of  cell  types,  cytokines  mediate  cell-­‐

signaling  in  either  an  autocrine  or  paracrine  manner.    The  type  and  magnitude  of  the  cytokine  milieu  

produced  subsequently  dictates  the  strength  and  form  of  immune  response.    As  the  most  diverse  

vertebrate  group,  with  a  high  sensitivity  to  contaminants,  fish  represent  an  important  foci  for  the  

evaluation  of  immune  system  evolution,  function,  and  alteration  upon  toxicant  exposure.    While  many  

cytokines  have  been  identified  in  teleosts,  primary  study  has  been  limited  to  model  species  (e.g.  

zebrafish  and  fugu).    However,  evidence  exists  for  several  variations  of  cytokine  genes  within  taxa,  

underscoring  the  need  for  species-­‐specific  evaluation.      

In  this  study,  two  pro-­‐inflammatory  cytokines  (IL-­‐1β  and  TNFα),  one  chemokine  (IL-­‐8),  and  one  

anti-­‐inflammatory  cytokine  (IL-­‐10)  were  cloned,  sequenced,  and  characterized  for  the  first  time  in  two  

commercially  relevant  Perciformes  in  the  Gulf  of  Mexico,  golden  tilefish  (Lopholatilus  chamaeleonticeps)  

and  red  snapper  (Lutjanus  campechanus).    The  complete  amino  acid  sequence  was  obtained  and  

confirmed  for  IL-­‐1β  and  IL-­‐8  from  golden  tilefish  and  for  IL-­‐8,  IL-­‐10,  and  TNFα  from  red  snapper,  with  

partial  sequences  obtained  for  the  remaining  proteins.    The  results  indicate  high  homology  among  

Perciformes  for  all  cytokines  studied,  but  divergence  with  other  teleost  orders,  and  low  conservation  

when  compared  to  birds,  amphibians,  and  mammals.      

The  sequences  will  be  used  to  create  a  multi-­‐plexed  antibody-­‐based  assay  for  the  routine  

detection  of  cytokines  in  teleost  serum.    This  would  allow  the  biochemical  response  to  fish  health  

Page 13: Graduate Theses and Dissertations Graduate School 11-4 ...

  ix  

challenges,  such  as  oil  spills  and  other  contamination  events,  to  be  monitored  at  the  protein  level,  

building  upon  the  current  regime  of  genetic  biomarkers.    Thus,  this  work  will  aid  in  the  understanding  of  

how  oil  spills  and  other  contamination  events  may  alter  the  immune  response  in  fishes.  

Page 14: Graduate Theses and Dissertations Graduate School 11-4 ...

  1  

         

CHAPTER  ONE:  

INTRODUCTION  

 

1.1 Deepwater  Horizon,  PAH  immunotoxicity,  and  traditional  biomarkers  for  PAH  exposure  

  The  Deepwater  Horizon  (DWH)  explosion  on  April  20th,  2010  and  subsequent  oil  spill  resulted  in  

the  second  largest  environmental  disaster  and  response  effort  in  human  history  [1,  2].    Before  final  

cessation  on  July  15,  2010,  780  million  liters  (L)  of  crude  oil  (3.9%  polycyclic  aromatic  hydrocarbons  

(PAHs)  by  weight)  were  released  into  one  of  the  most  productive  fisheries  in  the  United  States,  resulting  

in  a  36.6%  offshore  fishery  closure  [2-­‐4].    In  addition  to  a  surface  sheen,  four  temporally  persistent  

subsurface  plumes  were  noted  at  25  meters  (m),  265  m,  875  m,  and  1175  m  [2].    A  total  of  7  million  L  of  

dispersant  were  utilized  in  an  attempt  to  mitigate  oil  reaching  the  shoreline,  including  4  million  L  applied  

at  the  surface  and  2.9  million  L  applied  at  the  wellhead  [1].    This  application  of  dispersant  at  depth  had  

never  before  been  attempted,  and  its  resulting  longevity  and  effect  in  the  water  column  were  uncertain,  

though  recent  sampling  indicates  dispersal  of  petroleum  far  beyond  the  scope  of  the  1175  m  plume  [2].    

In  addition,  there  is  evidence  of  dispersant-­‐enhanced  extraction  of  toxic  compounds,  like  benzene  and  

PAHs,  to  the  entire  span  of  the  water  column,  increasing  overall  aquatic  toxicity  [2,  4].    These  soluble  

enhanced  constituents  were  bioavailable  and  free  to  diffuse  across  biological  membranes,  linking  

dispersant  application  to  an  increased  PAH  uptake  by  fish  [5].  

Metabolism  of  PAHs  induces  their  immunotoxic  effects  and  triggers  cytokine  expression  in  

characteristic  ways.    When  PAHs  cross  biological  membranes  they  activate  the  aryl  hydrocarbon  

receptor  (AhR)  pathway,  causing  AhR  to  dissociate  from  cofactors,  including  heat  shock  protein  90  [6,  7].  

The  AhR  and  PAH  complex  then  enters  the  nucleus  and  upregulates  cytochrome  P450  (CYP1A)  

Page 15: Graduate Theses and Dissertations Graduate School 11-4 ...

  2  

expression,  a  hydroxylase  that  aids  in  PAH  metabolism  [6,  8,  9].    This  induction  of  CYP1A  generally  

increases  the  intracellular  concentration  of  reactive  oxygen  species  [10].    ROS  are  produced  naturally,  as  

a  product  of  anaerobic  respiration,  and  cells  contain  antioxidant  systems  to  neutralize  them  [11,  12].  

However,  if  these  antioxidant  systems  are  overwhelmed,  the  NF-­‐кB  pathway  can  be  initiated,  a  stress  

response  nuclear  factor  network  that  triggers  the  release  of  many  pro-­‐inflammatory  cytokines  [10].    This  

release  of  pro-­‐inflammatory  cytokines  is  a  clear  indication  of  the  immune  response.      

Traditional  biomarkers  utilized  to  understand  the  effect  of  oil  contamination  on  marine  

organisms  focus  on  the  earlier  portion  of  the  PAH  biochemical  pathway,  particularly  through  

measurement  of  AhR  and  Cyp1A  gene  expression  [13,  14].  While  these  targets  have  utility,  they  

primarily  indicate  the  exposure  to  and  metabolism  of  oil,  not  necessarily  the  resulting  biological  impact  

[15,  16].    A  need  exists  for  a  novel  suite  of  biomarkers  that  can  quantify  the  alterations  in  the  

biochemistry  of  marine  organisms  that  lead  to  physiological  response  and  aberrant  pathologies,  such  as  

the  noted  increase  in  lesions  in  fishes  found  in  contaminated  waters  [7,  17,  18].    This  includes  the  

identification  and  quantification  of  a  link  between  PAH  exposure  and  subsequent  alterations  of  immune  

system  function  [19-­‐21].  

 

1.2  Teleost  cytokines  and  the  immune  response  

  Immunity  is  an  ecologically  critical  biological  function  that  allows  an  organism  to  fight  pathogens  

while  continuing  to  proliferate,  thus  aiding  in  population  growth  [22].    Cytokines  are  pleiotropic  

signaling  peptides,  proteins,  and  glycoproteins  that  mediate  inflammation  and  influence  both  innate  and  

adaptive  immunity  [23-­‐25].    Secreted  by  a  variety  of  cell  types  in  response  to  a  stressor,  cytokines  have  

pro-­‐  and  anti-­‐inflammatory  roles  and  control  the  ratio  of  various  T  helper  (Th)  cell  subsets,  which  in  turn  

helps  determine  the  type  of  immune  response  and  maintains  homeostasis  in  vertebrates  [26,  27].    

Cytokines  can  regulate  the  ability  of  phagocytes  to  destroy  foreign  pathogens,  as  well  as  control  antigen  

Page 16: Graduate Theses and Dissertations Graduate School 11-4 ...

  3  

presentation  on  T  cells,  and  stimulate  lymphocyte  migration  throughout  the  body  [25].    They  can  

function  through  either  autocrine  or  paracrine  signaling  and  their  expression  can  either  be  altered  in  

quick  bursts,  such  as  during  infection,  or  their  aberrant  expression  can  be  prolonged,  as  is  the  case  in  

well-­‐studied  human  diseases  like  rheumatoid  arthritis  [28-­‐30].  

  Fish  are  one  of  the  earliest  vertebrate  groups  to  possess  hallmarks  of  both  innate  and  adaptive  

immunity  and  are  frequently  used  as  models  in  evolutionary  immunological  studies  [31,  32].    While  

some  of  the  main  immune  organs  are  different  (in  addition  to  the  thymus  and  spleen,  teleost  immunity  

is  also  dependent  upon  the  head  kidney)  many  of  the  cell  types  involved  in  the  immune  response  are  

conserved  between  fish  and  mammals,  including  monocytes/macrophages,  non-­‐specific  cytotoxic  cells  

(NCC),  natural  killer  cells  (NK),  and  granulocytes  [33].      In  addition,  critical  complexes  and  receptors  to  

the  mammalian  immune  system  are  conserved  in  teleosts,  including  immunoglobulins,  T  cell  receptors  

(TCR),  and  major  histocompatibility  complexes  (MHC)  [33-­‐38].    Given  their  continual  dermal  exposure  to  

aquatic  viruses,  bacteria,  contaminants,  and  other  threats,  it  is  not  surprising  that  fish  would  have  also  

developed  unique  immune  strategies,  including  the  presence  of  specialized  toll-­‐like  receptors  (TLR)-­‐22  

and  TLR3  for  the  detection  of  viral  RNA  on  the  cell  surface  and  intracellularly,  respectively  [34]  

    The  major  secretory  cells  of  cytokines,  Th  cells,  are  produced  upon  naive  CD4+  cell  recognition  

of  a  foreign  peptide  and  subsequent  differentiation  [25,  35].    Genomic  work  in  teleosts  confirmed  the  

presence  of  critical  T  cell  function  and  signaling-­‐related  molecules  indicative  of  all  four  major  Th  subsets,  

including  Th1,  Th2,  Th17,  and  Treg  cells,  and  functional  studies  suggest  similar  lymphocyte  activity  to  

mammals  [25,  31,  35].    However,  the  definitive  presence  of  these  Th  cell  subsets  cannot  yet  be  

substantiated,  given  the  lack  of  teleost-­‐specific  antibodies  [25,  34].    In  addition,  unique  T  cell  molecules  

have  been  identified  in  fish.  For  example,  teleost  cells  appear  to  have  relevant  Th1  cytokines,  including  

IFNγ,  although  they  also  have  a  teleost-­‐specific  molecule,  IFNγ-­‐rel  [35].    

Page 17: Graduate Theses and Dissertations Graduate School 11-4 ...

  4  

Many  important  mammalian  cytokines  have  been  identified  in  fish,  including  at  least  19  

interleukins,  and  members  of  the  tumor  necrosis  factor,  chemokine,  and  interferon  families.    However,  

teleost  functional  studies  are  significantly  lagging  behind  mammalian  studies  and  gene  examination  is  

limited  to  a  few  model  species  [32,  36-­‐38].    Teleost  cytokine  identification  is  complicated  by  the  fish  

specific  genome  duplication  event  (3R  or  FSGD),  which  lead  to  fish-­‐specific  copies  of  several  known  

cytokines  (potentially  partitioning  functions  of  ancestral  genes  among  them),  along  with  a  likely  fourth  

duplication  event  in  Salmoniformes  [32,  39,  40].    In  addition  to  possible  differential  function  and  

expression  between  isoforms  of  the  same  signaling  protein,  certain  species  of  fish  may  have  multiple  

variations  of  a  single  gene,  while  other  species  do  not,  underscoring  the  need  for  species-­‐specific  

examination  [24,  25,  40-­‐42].  

 

1.3  Characteristics  of  vertebrate  IL-­‐1β ,  IL-­‐8,  IL-­‐10,  and  TNFα  

In  this  study,  four  cytokines  were  explored  in  two  important  fishery  species  in  the  Gulf  of  

Mexico,  golden  tilefish  (Lopholatilus  chamaeleonticeps)  and  red  snapper  (Lutjanus  campechanus).    The  

proteins  of  interest  include  two  proinflammatory  cytokines  (interleukin  (IL)-­‐1β  and  tumor  necrosis  factor  

alpha  (TNFα)),  one  chemokine  (IL-­‐8),  and  one  anti-­‐inflammatory  cytokine  (IL-­‐10).    These  targets  were  

selected  given  their  importance  in  the  inflammatory  and  immune  response  as  well  as  their  extensive  

prior  evaluation  in  model  fish  species,  primarily  in  relation  to  toxicant  exposure  and  with  regards  to  

parasite  infection  in  an  aquaculture  setting  [19,  21,  37,  38,  43].      

Interleukin-­‐1β  is  one  of  the  earliest  expressed  pro-­‐inflammatory  cytokines,  induced  by  tissue  

injury,  microbial  invasion,  and  autoimmune  disease  [37,  44].    Produced  by  a  variety  of  cell  types,  

primarily  macrophages  and  neutrophilic  granulocytes  in  teleosts,  IL-­‐1β  stimulates  T  cells  and  initiates  a  

cascade  of  potent  effects  by  subsequent  up-­‐  or  down-­‐regulation  of  other  cytokines,  resulting  in  

inflammation  [39,  45].    The  IL-­‐1β  gene  has  been  identified  in  a  number  of  teleost  species,  including  

Page 18: Graduate Theses and Dissertations Graduate School 11-4 ...

  5  

Cypriniformes,  Perciformes,  Salmoniformes,  and  Gadiformes,  and  similarities  to  its  mammalian  

counterpart  are  observed,  though  the  teleost  gene  clusters  independently  of  birds,  mammals  and  

amphibians  on  phylogenetic  trees  [38,  44,  46-­‐51].        

Differential  phylogenetic  clustering  could,  in  part,  be  due  to  fundamental  sequence  

discrepancies  between  the  mammalian  and  fish  orthologues.    For  example,  while  mammalian  IL-­‐1β  

exhibits  a  seven  exon  structure,  the  Perciform  orthologue  has  four  to  five  exons,  and  Salmoniformes  

have  six,  though  the  classic  β  trefoil  shape  with  12  β-­‐sheet  structure  is  conserved  among  all  studied  

fishes  [38,  47].    In  addition,  mammalian  IL-­‐1β  is  produced  as  an  inactive  peptide  that  must  be  

proteolytically  cleaved  by  caspase-­‐1  at  the  IL-­‐1β  converting  enzyme  (ICE)  cut  site  [47,  52,  53].    

Interestingly,  teleosts  typically  lack  this  ICE  cut  site,  though  experimental  evidence  suggests  the  fish  

orthologue  is  also  produced  in  an  inactive  form  and  later  activated  via  a  cleavage  mechanism  [47].    

Multiple  isoforms  of  the  gene  exist  in  several  species,  including  carp,  trout,  salmon,  catfish,  and  

platyfish,  possibly  due  to  duplication  of  the  IL-­‐1  locus  during  the  FSGD  event  [47,  53].      

Despite  these  differences,  functional  studies  in  rainbow  trout,  haddock,  and  olive  flounder,  

suggest  that  teleost  IL-­‐1β  is  involved  with  regulation  of  immune  genes,  lymphocyte  activation,  migration  

of  leucocytes,  phagocytosis,  and  bactericidal  activity  [37,  54].    Recombinant  IL-­‐1β  from  sea  bass  and  

rainbow  trout  can  induce  the  release  of  calcium  (Ca2+)  in  teleost  leukocytes,  an  activity  not  clearly  

demonstrated  in  mammalian  species  [53].    An  increase  in  intracellular  [Ca2+]  has  also  been  noted  in  fish  

cells  upon  exposure  to  benzo(a)pyrene  and  other  PAH’s,  potentially  linking  the  activities  of  this  cytokine  

to  PAH  exposure  [7].    Upregulation  of  the  IL-­‐1β  gene  has  been  noted  after  exposure  to  heavy  crude  oil  in  

rockfish  liver  [19].    Suggested  functions  from  mammalian  studies  include  response  to  tissue  injury  [44].    

Chemokines,  like  IL-­‐8,  are  chemoattractant  cytokines  that  primarily  assist  in  cell  migration,  

principally  by  invoking  leukocyte  mobilization  to  sites  of  infection  or  injury  [55].    Specifically,  IL-­‐8  

belongs  to  the  CXC  family  of  chemokines  and  attracts  neutrophils,  T  lymphocytes,  and  basophils  to  

Page 19: Graduate Theses and Dissertations Graduate School 11-4 ...

  6  

inflamed  regions  [37].      While  current  functional  assays  of  IL-­‐8  in  teleosts  display  similar  activity  and  

basal  levels  of  expression  as  the  mammalian  orthologue,  teleost  sequences  have  only  30-­‐40%  amino  

acid  identity  with  the  mammalian  protein  [55,  56].      

Recent  evidence  suggests  that  teleost  CXC  is  a  novel  lineage  from  other  vertebrates,  with  up  to  

three  distinct  strains  evolving  within  fish  families  [25,  42,  56,  57].    The  hallmark  of  IL-­‐8  proteins  in  

mammals  is  the  presence  of  an  ELR  motif  before  the  first  cysteine,  which  most  fish  homologues  (except  

Gadoids)  lack.    This  motif  is  necessary  for  neutrophil  chemotaxis,  therefore,  the  function  of  IL-­‐8  in  

teleosts  may  be  slightly  different  from  that  in  mammals  [42,  48,  56].    However,  a  recent  study  in  

flounder  (which  has  a  SLH  motif,  rather  than  the  mammalian  ELR  motif)  designated  a  leucine  residue  at  

the  N-­‐terminus  of  the  IL-­‐8  gene  as  responsible  for  induction  of  chemotaxis  in  the  fish  [57].    In  addition,  

in  vitro  functional  studies  in  rainbow  trout  demonstrate  the  ability  of  IL-­‐8  to  recruit  monocyte  cells,  as  

well  as  induce  expression  of  pro-­‐inflammatory  cytokines,  such  as  IL-­‐1β  and  TNFα  [42,  49].    Upon  

exposure  to  benzo(a)pyrene,  a  toxic  PAH,  IL-­‐8  was  up-­‐regulated  in  flounder  macrophage  cells  [21,  58]  

One  of  the  most  well  studied  pro-­‐inflammatory  cytokines  in  fish,  TNFα  has  been  shown  to  have  

numerous,  pleiotropic  functions  [38,  39].    Some  of  these  functions  include  upregulation  of  adhesion  

molecules,  induction  of  chemokine  release,  induction  of  the  NF-­‐кb  pathway,  regulation  of  apoptosis,  

and  stimulation  of  IL-­‐6  production.    TNFα  has  also  been  shown  to  upregulate  the  production  of  IL-­‐1β  and  

IL-­‐8  in  rainbow  trout  [37,  59,  60].    Therefore,  TNFα  is  critical  in  antimicrobial  defense  and  response  to  

injury,  and  helps  potentiate  the  inflammatory  response  [38,  61].    Its  function  in  fish  mirrors  that  noted  in  

mammals,  although  multiple  forms  of  the  gene  have  been  observed  in  teleosts,  including  two  variants  in  

bluefin  tuna  [38,  39,  62].    Structural  information  for  TNFα  among  cloned  species  has  indicated  high  

levels  of  divergence,  indicating  a  need  for  species-­‐specific  evaluation  of  the  molecule  [62].      However,  

the  majority  of  the  TNFα  family  signature  and  two  cysteine  residues  are  conserved  among  all  studied  

vertebrates  [60].  

Page 20: Graduate Theses and Dissertations Graduate School 11-4 ...

  7  

Finally,  IL-­‐10  is  one  of  the  most  important  and  potent  anti-­‐inflammatory  cytokines  [41].    

Through  induction  of  Treg  cells,  IL-­‐10  inhibits  production  of  pro-­‐inflammatory  cytokines  by  macrophages  

and  dendritic  cells,  including  the  regulation  of  TNFα  [41,  63,  64].    Produced  after  proinflammatory  

mediators,  IL-­‐10  can  inhibit  Th1  and  Th17  immunity,  partially  through  presentation  of  antigen  

presenting  cells,  and  helps  prevent  a  run-­‐away  immune  response.    Essentially,  IL-­‐10  is  responsible  for  

limiting  the  ability  of  monocytes/macrophages  to  partake  positively  in  the  immune  response  while  

simultaneously  aiding  in  the  production  of  anti-­‐inflammatory  mediators  [41].    In  addition,  in  mammalian  

systems,  IL-­‐10  has  been  implicated  in  the  down-­‐regulation  of  ROS,  which  can  be  produced  as  a  by-­‐

product  of  PAH  metabolism  [10,  65].    This  interleukin  can  regulate  the  function  of  many  leukocyte  

classes,  including  the  activation  of  B  cells,  the  proliferation  of  NK  cells,  and  the  differentiation  of  

cytotoxic  T  cells  [66,  67].      

To  date,  IL-­‐10  has  been  identified  in  several  teleosts,  including  European  seabass,  rainbow  trout,  

cod,  fugu,  common  carp,  silver  carp,  zebrafish,  and  goldfish,  and  functional  studies  suggests  analogous  

function  to  that  of  its  mammalian  counterpart  [25,  64,  66,  68-­‐70].    For  example,  recombinant  goldfish  IL-­‐

10  can  regulate  multiple  isoforms  of  TNFα,  IL-­‐1β1,  and  IL-­‐8  in  fish  monocytes  [64].    Furthermore,  the  IL-­‐

10  receptor  subunit  1  has  been  identified  in  grass  carp,  indicating  at  least  partially  conserved  signaling  

pathways  for  the  molecule  throughout  evolution  [67].    To  date,  rainbow  trout  are  the  only  teleost  

species  identified  with  more  than  one  gene  encoding  IL-­‐10,  though  both  trout  genes  have  92%  identity  

in  the  translated  region  [25,  71].      In  addition  to  four  disulfide  bond-­‐forming  cysteine  residues  conserved  

in  all  vertebrates,  fish  IL-­‐10  contains  two  cysteine  residues  that  appear  to  be  teleost-­‐specific  [70].  

 

1.4  Objectives  

In  this  study,  the  above-­‐mentioned  cytokines  (IL-­‐1β,  IL-­‐8,  IL-­‐10,  and  TNFα)  were  isolated,  cloned,  

sequenced,  and  characterized  for  the  first  time  from  two  important  fishery  species  in  the  Gulf  of  Mexico,  

Page 21: Graduate Theses and Dissertations Graduate School 11-4 ...

  8  

red  snapper  and  golden  tilefish.    Red  snapper  are  reef-­‐associated,  offshore  Perciformes  with  high  

fecundity  [72,  73].    They  represent  the  most  economically  important  snapper  species  to  the  Gulf  of  

Mexico  commercial  fishery,  whose  3.24  million  pound  yield  in  2011  produced  $13.8  million  in  revenue  

[72,  73].    They  are  also  one  of  the  most  targeted  fishes  for  the  enormous  Gulf  of  Mexico  recreational  

fishery,  with  their  annual  landings  almost  always  exceeding  the  Annual  Catch  Limit  set  by  NOAA,  

sometimes  exceeding  the  limit  by  as  much  as  49.6%  [74].    Golden  tilefish  represent  a  smaller  

commercial  fishery,  but  their  habitat  selection  method  of  burrow  formation  in  deep  sediments  may  

enhance  their  exposure  to  settled  hydrocarbons  and  other  contaminants,  making  them  valuable  for  

environmental  health  assessment  studies  [75,  76].      

It  was  hypothesized  that  field-­‐caught  golden  tilefish  and  red  snapper  cytokine  sequences  would  

share  high  homology  with  the  corresponding  Perciform  genes.    Anomalies  were  anticipated  as  

comparisons  across  other  evolutionary  lineages  ensued,  including  comparisons  between  other  

Actinopterygii,  as  well  as  between  cartilaginous  fish,  amphibians,  birds,  and  mammals.    The  results  of  

this  study  will  aid  in  the  body  of  literature  identifying  conserved  genetic  regions  within  fishes  and  among  

vertebrate  taxa  and  strengthen  the  phylogenetic  analysis  of  cytokine  structure,  potentially  aiding  in  

predictions  concerning  the  evolution  of  the  immune  system.    

The  resulting  characterized  sequences  and  phylogenetic  correlations  from  this  study  will  

eventually  be  used  to  create  antibodies  for  the  generation  of  a  highly  cross-­‐reactive  magnetic  bead-­‐

based  assay  for  detection  of  cytokines  at  the  protein  level  in  Perciform  serum,  plasma,  and  cell  culture  

supernatant.    The  kit  would  likely  be  validated  utilizing  a  dosing  study  and  correlated  with  an  evaluation  

of  gene  expression  patterns  in  vivo  and  in  vitro  in  response  to  a  water  accommodated  fraction  (WAF)  of  

oil  (DWH  crude  oil)  and  a  chemically  enhanced  WAF  (CEWAF;  DWH  crude  oil  +  Corexit  9500).    In  

addition,  gene  expression  patterns  and  protein  production  levels  will  likely  be  compared  to  those  found  

in  fish  samples  with  known  PAH  body  burdens  collected  from  the  Gulf  of  Mexico  in  the  vicinity  of  the  

Page 22: Graduate Theses and Dissertations Graduate School 11-4 ...

  9  

DWH  explosion  from  2012-­‐2014.    This  work  will  form  the  basis  of  the  PhD  dissertation  to  be  prepared  by  

the  author  from  2015  through  2018  and  will  be  the  first  to  produce  a  thoroughly  validated,  Perciform-­‐

specific  antibody-­‐based  kit  for  cytokine  detection.    Given  the  amount  of  variability  in  cytokine  genes  

across  teleosts,  such  a  thorough  investigation  first  required  an  analysis  of  the  target  genes  in  each  

species  of  interest.        

 

                                                               

       

Page 23: Graduate Theses and Dissertations Graduate School 11-4 ...

  10  

         

CHAPTER  TWO:  

METHODS  

 

2.1  Isolation  of  total  RNA  

  Golden  tilefish  (n=39)  and  red  snapper  (n=38)  spleen  samples  were  obtained  from  freshly  caught  

fish  from  the  Gulf  of  Mexico  in  August  2013  aboard  the  R/V  Weatherbird  II  (a  map  of  sampling  locations  

can  be  found  in  Appendix  A,  with  location  coordinates  and  depths  in  Appendix  B).    For  this  study,  four  

golden  tilefish  spleen  samples  were  used,  from  fish  weighing  2.630-­‐4.054  kg,  caught  at  stations  SL11-­‐

150  (labeled  F  and  G)  and  SL14-­‐100  (labeled  T8  and  T9;  all  biometrics  data  in  Appendix  C).    In  addition,  

two  red  snapper  spleen  samples  were  used,  labeled  C  and  E,  from  fish  weighing  4.508-­‐5.274  kg,  caught  

at  station  SL4-­‐40.    Fish  were  caught  via  demersal  longline  fishing  from  35  to  225  m  depth  for  red  

snapper  and  from  156  to  508  m  depth  for  golden  tilefish,  depending  upon  the  site,  using  a  5  mile  line  

with  approximately  500  baited  hooks  and  an  average  soak  time  of  two  hours.      

Fish  were  landed  on  deck,  bled,  and  tissue  samples  were  immediately  taken.    All  spleen  tissue  

was  subsampled  into  5  mm  x  5  mm  x  5  mm  cubes,  immediately  stored  in  1  mL  RNAlater  (Life  

Technologies,  Grand  Island,  New  York),  and  placed  in  a  -­‐20oC  freezer,  until  transport  to  a  final  storage  

location  at  -­‐80oC.    Samples  were  thawed  at  room  temperature  and  subsampled  if  necessary  to  50-­‐100  

mg  pieces.    Total  RNA  was  isolated  using  TRI  Reagent  (Molecular  Bioproducts,  San  Diego,  California)  

according  to  manufacturer’s  protocol  and  dissolved  in  25  µL  nuclease-­‐free  water  (Life  Technologies,  

Grand  Island,  New  York).    Samples  were  DNAse  treated  using  the  Turbo  DNA-­‐free  kit,  according  to  

manufacturer’s  instructions  (Ambion,  Austin,  Texas).      RNA  was  quantified  using  a  Nanodrop  

Page 24: Graduate Theses and Dissertations Graduate School 11-4 ...

  11  

spectrophotometer  (Nanodrop  Technologies,  Wilmington,  Delaware)  and  sample  concentration  was  

adjusted  to  1  µg/µL  with  nuclease-­‐free  water.    Samples  were  stored  at  -­‐80oC.  

 

2.2  cDNA  production  

  Total  RNA  (1  µL)  was  reverse  transcribed  to  cDNA  using  the  Reverse  Transcription  System,  per  

manufacturer’s  instructions  (Promega,  Madison,  Wisconsin).    Proper  performance  of  the  reverse  

transcription  reaction  was  assessed  via  sufficient  amplification  of  β-­‐actin,  using  reverse-­‐transcription  

polymerase  chain  reaction  (RT-­‐PCR),  GoTaq  Green  Master  Mix  (Promega,  Madison,  Wisconsin)  and  

European  seabass  β-­‐actin  primers  and  parameters  [50].    Each  PCR  product  was  visualized  on  a  1.2%  

Ultrapure  agarose  (Life  Technologies,  Grand  Island,  New  York)  gel  with  1x  TAE  buffer  (Life  Technologies,  

Grand  Island,  New  York),  run  at  70  V  for  45  minutes  and  viewed  on  a  Chemidoc  gel  imager  (BioRad,  

Hercules,  California).      Rapid  Amplification  of  cDNA  Ends  (RACE)-­‐ready  cDNA  was  also  produced  using  1  

µL  Total  RNA,  but  with  the  SMARTer  RACE  5’/3’  kit  according  to  the  manufacturer’s  protocol  (CloneTech,  

Mountain  View,  California).      

 

2.3  Initial  identification  and  cloning  of  IL-­‐1β ,  IL-­‐8,  IL-­‐10,  and  TNFα  fragments  

  Degenerate  primers  were  taken  from  the  teleost  literature  or  designed  from  available  teleost  

FASTA  sequences  in  GenBank,  using  PriFi  (cgi-­‐www.daimi.au.dk/cgi-­‐chili/PRiFi/main;  Table  1).    Primer  

dimer  plausibility  was  screened  for  using  the  Multiple  Primer  Analyzer  Tool  

(www.thermoscientificbio.com/webtools/multipleprimer;  Thermo  Scientific).    All  oligonucleotides  were  

purchased  from  Invitrogen  and  adjusted  to  10  µM  with  nuclease-­‐free  water  (Life  Technologies,  Grand  

Island,  New  York).    Gradient  RT-­‐PCR  experiments  were  utilized  to  obtain  fragments  of  the  expected  base  

pair  (bp)  size  as  predicted  in  the  literature  or  using  PriFi  (Table  1).      All  RT-­‐PCR  experiments  began  with  a  

94oC  initialization  for  five  minutes,  followed  by  35  cycles  of  denaturation  at  94oC  for  30-­‐45  seconds,  

Page 25: Graduate Theses and Dissertations Graduate School 11-4 ...

  12  

primer  annealing  at  48-­‐60oC  for  30-­‐45  seconds,  and  primer  extension  at  72oC  for  45-­‐60  seconds,  ending  

with  a  final  extension  at  72oC  for  5-­‐10  minutes  (for  extended  PCR  parameter  information,  please  see  

Appendix  D).    Initial  25  µL  reactions  were  performed  and  products  were  viewed  on  a  1.2%  Ultrapure  

agarose  gel  (Life  Technologies,  Grand  Island,  New  York),  as  previously  described.  

  Reaction  parameters  producing  a  single  band  of  the  expected  size  (as  predicted  in  the  literature  

or  primer  design  software),  were  repeated  to  produce  100  µL  of  PCR  product,  and  subsequently  run  on  a  

1.2%  Ultrapure  agarose  gel  (Grand  Island,  New  York)  for  isolation.    Bands  were  excised  from  the  gel  and  

DNA  fragments  were  purified  using  the  Wizard  SV  Gel  and  PCR  Clean-­‐Up  System  by  centrifugation,  

according  to  the  manufacturer’s  protocol  (Promega,  Madison,  Wisconsin).    Purified  DNA  was  quantified  

using  a  Nanodrop  spectrophotometer  (Nanodrop  Technologies,  Wilmington,  Delaware)  and  cloned  using  

the  pGEM  T  Easy  Vector  System,  per  manufacturer’s  instructions  (Promega,  Madison,  Wisconsin).    

Ligation  reactions  (10  µL)  were  prepared  with  a  1:3  insert:vector  ratio.    Following  an  overnight  

incubation  at  4oC,  the  reaction  was  transformed  into  JM109  High  Efficiency  Competent  Cells  (Promega,  

Madison,  Wisconsin)  and  grown  with  SOC  medium  (Fisher,  Waltham,  Massachusetts)  for  90  minutes  at  

37oC  on  an  orbital  shaker  plate,  rotating  at  550  rpm.    Following  incubation,  100  µL  aliquots  of  the  

transformation  reactions  were  transferred  to  two  agar  plates  (Sigma-­‐Aldrich,  St.  Louis,  Missouri),  one  

undiluted  and  one  diluted  1:10  in  SOC  medium  (Fisher,  Waltham,  Massachusetts).    Colonies  were  grown  

at  37oC  for  18  hours.    

  Ten  colonies  were  selected  from  each  plate  and  grown  in  2  mL  LB  broth  (5  g  Tryptone  (EMD,  

Rockland,  Massachusetts),  2.5  g  yeast  (EMD,  Rockland,  Massachusetts),  2.5  g  NaCl  (Fisher,  Waltham,  

Massachusetts),  500  µL  100  mg/mL  Ampicillin  (Fisher,  Waltham,  Massachusetts),  to  500  mL  DI  H2O)  in  

15  mL  conical  vials.    Conical  vials  were  placed  on  an  orbital  shaker  at  550  rpm  at  37oC  for  20  hours.    

Plasmid  purification  was  performed  using  the  Manual  FastPlasmid  Mini  Kit  (5Prime,  Hamburg,  

Germany).    Colonies  were  screened  for  insert  using  RT-­‐PCR  with  universal  T7  and  SP6  primers,  ordered  

Page 26: Graduate Theses and Dissertations Graduate School 11-4 ...

  13  

from  Life  Technologies  (Grand  Island,  New  York;  Table  1).    Purified  plasmid  DNA  was  quantified  as  

before,  adjusted  to  50-­‐100  ng/uL  and  two  samples  with  sufficient  260/280  values  and  concentrations  

were  sent  to  Eurofin  Genomics  for  sequencing  (Eurofins  MWG  Operon,  Huntsville,  Alabama).    The  vector  

was  removed  from  sequencing  results  using  VecScreen  (www.ncbi.nlm.nih.gov/tools/vecscreen)  and  the  

remaining  fragment  was  input  to  the  nucleotide  Blast  Local  Alignment  Tool  (BLASTn;  

www.ncbi.nlm.nih.gov/blast).      

 

2.4  5’-­‐  and  3’-­‐RACE  

If  sufficient  homology  existed  between  the  obtained  fragment  and  other  available  teleost  

nucleotide  sequences  for  the  same  cytokine,  5’/3’-­‐RACE  was  performed  using  the  SMARTer  RACE  5’/3’  

Kit  and  the  Advantage  2  Polymerase  Mix  (Clonetech,  Mountain  View,  California).    Gene-­‐specific  

oligonucleotide  primers  designed  for  5’/3’-­‐RACE  (GSP1  for  3’-­‐RACE  and  GSP2  for  5’-­‐RACE)  using  

Primer3Plus  (www.bioinformatics.nl/cgi-­‐bin/primer3plus)  and  the  parameters  dictated  in  the  SMARTer  

5’/3’-­‐RACE  protocol.    Primer  dimer  plausibility  between  designed  primers  and  kit  provided  long  and  

short  universal  primers  (Clonetech,  Mountain  View,  California)  was  again  screened  using  the  Multiple  

Primer  Analyzer  Tool.  

Gradient  RT-­‐PCR  was  performed  to  optimize  the  annealing  temperature  using  RACE-­‐ready  cDNA  

and  PCR  products  were  visualized  on  a  1.2%  Ultrapure  agarose  gel  (Life  Technologies,  Grand  Island,  New  

York).    When  a  bright,  single  band  was  obtained,  a  100  µL  reaction  was  repeated  under  the  same  

conditions,  extracted  and  purified,  as  above.    Products  were  screened  for  inserts  using  via  RT-­‐PCR  with  

the  respective  GSP1  or  GSP2  primer.    After  quantification,  samples  were  sent  for  sequencing,  as  

previously  described.    The  resulting  5’/3’  RACE  sequences  were  assembled  and  translated  into  amino  

acids  using  the  ExPASy  translate  tool  (web.expasy.org/translate).    Homology  with  teleost  proteins  was  

confirmed  using  BLASTp  (www.ncbi.nlm.nih.gov/blast).  

Page 27: Graduate Theses and Dissertations Graduate School 11-4 ...

  14  

2.5  Sequence  confirmation  

Gene-­‐specific  primers  were  designed  with  Primer3Plus  to  isolate  the  coding  sequence  for  the  

protein  of  interest.    They  were  applied  to  the  cDNA  of  two  different  fish  from  each  species,  in  order  to  

validate  the  sequence  of  the  coding  region  and  eliminate  the  possibility  of  single  nucleotide  

polymorphisms  skewing  sequence  data.    The  isolation,  amplifying,  cloning,  and  sequencing  were  

performed  according  to  previously  described  methods.    The  resulting  sequences  were  compared  using  

Clustal  Omega  (www.ebi.ac.uk/Tools/msa/clustalo),  BLASTn,  and  BLASTp  and  any  discrepancies  were  

resolved  by  comparison  to  other  teleost  sequences  and  experimental  data.      

 

2.6  Analysis  of  confirmed  protein  coding  sequence    

  All  confirmed  protein  sequences  were  screened  for  the  presence  of  the  family  signature,  

according  to  PROSITE  (prosite.expasy.org)  and  the  relevant  teleost  literature  [38,  44,  48,  66,  68].  N  

glycosylation  sites  were  identified  by  the  NetNGlyc  1.0  server  (www.cbs.dtu.dk/services/NetNGlyc),  

putative  signal  peptide  cut  sites  were  identified  using  SignalP  4.1  (www.cbs.dtu.dk/services/SignalP),  

and  transmembrane  domain  regions  were  identified  using  TMPred  

(www.ch.embnet.org/software/TMPred_form.html).    Percent  identity  was  calculated  by  comparing  

amino  acid  sequences  using  BLASTp.    Phylogenetic  trees  were  generated  using  complete  sequences  by  

the  neighbor-­‐joining  (NJ)  method  in  MEGA6  (www.megasoftware.net),  validated  by  bootstrap  analysis  

with  10,000  repetitions  using  the  p-­‐distance  method  with  partial  deletions  and  an  80%  cutoff  [48,  77].      

 

 

 

       

Page 28: Graduate Theses and Dissertations Graduate School 11-4 ...

  15  

Table  1:  Oligonucleotide  primers  used  throughout  this  study,  including  their  purpose,  annealing  temperature,  expected  fragment  band  size  in  number  of  base  pairs  (bp),  and  primer  source.        

 

Name Nucleotide,Sequence,(5'33>,3') Use

Annealing,temp.,(C)

Fragment,size,(bp)

SourceBA_F ATCGTGGGGCGCCCCAGGCACA

BA_R CTCCTTAATGTCACGCACGATTTC

gtIL-1b_F GGGAAAGAATCTRTACCTGTCYTG

gtIL-1b_R GTGCTGATGAACCAGT

gtIL-8_F GCAGCTGAACMAAGAAAAGGAAAG

gtIL-8_R CAGCGGCAGTGCAKCTCCAYTCCCA

gtIL-10_F TGCTGCTCCTTCGTGGAGGGCTTCCC

gtIL-10_R CCAGCTCCCCCATGGCTTTATA

gtTNFa_F AGGCGTCSTTCAGAGTCTCC

gtTNFa_R GGSCTACTACAATCTCAGGTCA

rsIL-1b_F GCTGGAGAGCGTAGTYGAAGA

rsIL-1b_R CGGAGTCCTGTTTGTAGAAGAGAR

rsIL-8_F GCAGCTGAACMAAGAAAAGGAAAG

rsIL-8_R CAGCGGCAGTGCAKCTCCAYTCCCA

rsIL-10_F TGCTGCTCCTTCGTGGAGGGCTTCCC

rsIL-10_R CCAGCTCCCCCATGGCTTTATA

rsTNFa_F ATCAGCAGCAAAGCCAAGGCAGC

rsTNFa_R GAAGGTCTTGCCCTCGTCTGTCTCCAG

T7 TAATACGACTCACTAAGGG

SP6 ATTTAGGTGACACTATAGAA

gtIL-1b_GSP1 CCTGCCACAAGGATGGTGAAGAGCC 64 775

gtIL-1b_GSP2 CCAGTCACTGTAGGGGACGGACACG 64 750

gtIL-8_GSP1 TGAACAGCAGAGTCTTCGTCGCCACT 70 634

gtIL-10_GSP2 GGCTTGAGGTCTCTGGTGTCCTCCGT 62 513

gtTNFa_GSP1 AAACCGGCCTCTACTTCGTCTAC 60 779

rsIL-8_GSP1 GAGGAGAGAGGGAACTGAGCAAAAGA 68 784

rsIL-8_GSP2 CACCACAATAGAGGCGATGA 66 279

rsIL-10_GSP1 GAAGCAAACGACGACTTGGACAGAGC 68 711

rsIL-10_GSP2 TCCATGTGAGGCTTGAGGTCTCTGGT 68 519

rsTNFa_GSP1 TCTACAGCCAGGCGTCGTTCAGG 70 777

rsTNFa_GSP2 TACCAGCCGTGTCCGTCTCTGTAGC 70 744

gtf_IL-1b_F TGGGGAGAACAACACTGAGA

gtf_IL-1b_R TCCACACTCCACTTTGGTTG

gtf_IL-8_F CACAGTGCTTCACCAGAAGG

gtf_IL-8_R AAGGACATTAAGCCGTCCTCT

rsf_IL-8_F TTCAGGCATCTTTTTGAAGGA

rsf_IL-8_R GGCCTCAGTTGGATGAATGT

rsf_IL-10_F GCCAGCAACATCCTCTTCAT

rsf_IL-10_R TCTTCAAGCAGAGGCCACAT

rsf_TNFa_F GCCAAGTGCTCGACTGGTAG

rsf_TNFa_R GCAAACACGCCAAAGAAAGT56

48

54

54

54

48

48

50

55

50

54

Scapigliati,H2001RTHqualityH

control

Scapigliati,H2001

Buonocore,H2007

Xie,H2008

Lu,H2008

55

48

50

54

500

199

279

441

572

474 Covello,H2009

UniversalIsolatingH

fromHvector

430

277

406

Fragment

+H160Hbp

Deak,HPriFi

Buonocore,H2007

ObtainingH

intitialH

fragments

Deak,HPriFi

ObtainingH

finalH

proteinH

sequence

Deak,HHHHHHHHHH

Primer3Plus

881

513

453

647

884

5'-HandH3'-H

RACEH

Deak,H

Primer3Plus

Page 29: Graduate Theses and Dissertations Graduate School 11-4 ...

  16  

         

CHAPTER  THREE:  

RESULTS  AND  DISCUSSION  

 

3.1  Cloning  and  sequence  analysis  of  golden  tilefish  IL-­‐1β    

  Degenerate  primers  gtIL-­‐1b_F  and  gtIL-­‐1b_R  were  used  to  amplify,  via  PCR,  an  initial  199  base  

pair  (bp)  fragment  from  cDNA  isolated  from  golden  tilefish  spleen  F  (Table  1).    The  fragment  was  cloned,  

sequenced,  and  utilized  for  the  design  of  gene-­‐specific  oligonucleotides  to  perform  5’-­‐  and  3’-­‐RACE  

(Table  1).    The  5’-­‐RACE  reaction  yielded  a  750  bp  product  and  3’-­‐RACE  resulted  in  a  775  bp  product,  with  

a  153  bp  overlap  between  the  two  fragments.    This  resulted  in  an  assembled  RACE  product  of  1,396  bp  

with  a  756  bp  translated  region  (Figure  1a).    The  547  bp  3’  untranslated  region  (UTR)  contained  three  

instability  motifs  (ATTTA),  five  endotoxin  responsive  motifs  (TTATTTAT),  two  of  which  were  overlapping,  

and  a  polyadenylation  signal  (AATAAA)  ending  14  bp  upstream  of  the  poly-­‐A  tail.    The  5’  UTR  was  93  bp  

in  length  and  likely  incomplete  (all  experimentally  determined  nucleotide  sequences  may  be  found  in  

Appendix  E).    The  assembled  RACE  product  was  utilized  to  design  gene-­‐specific  primers  to  isolate  the  

coding  sequence  for  subsequent  cloning  and  sequence  confirmation  through  analysis  of  clones  from  

different  fish  (Table  1).      

  The  resulting  sequences,  cloned  from  separate  fish,  shared  high  consensus,  yielding  a  final,  

confirmed  756  bp  coding  region,  which  was  translated  into  251  amino  acids  and  contained  the  family  

signature,  [FCL]-­‐x-­‐S-­‐[ASLV]-­‐x(2)-­‐[PRS]-­‐x(2)-­‐[FYLIV]-­‐[LI]-­‐[SCAT]-­‐T-­‐x(7]-­‐[LIVMK],  as  

FvSVpySdwYISTagddnkpV  (Figure  2;  all  experimentally  derived  amino  acid  sequences  may  be  found  in  

Appendix  F,  all  confirmed  sequence  comparisons  may  be  found  in  Appendix  G).    No  conservation  was  

observed  at  the  aspartic  acid  residue  marking  the  ICE  cleavage  site  in  mammals  (Figure  2).  This  residue  is  

Page 30: Graduate Theses and Dissertations Graduate School 11-4 ...

  17  

typically  absent  in  non-­‐mammalian  vertebrates,  though  processing  by  other  proteases  has  been  

postulated  [78,  79].    No  signal  peptide  was  predicted  by  SignalP,  which  suggests  secretion  by  a  non-­‐

canonical  pathway,  avoiding  the  golgi-­‐endoplasmic  reticulum  excretion  mechanism,  another  vertebrate  

characteristic  for  IL-­‐1β  [78,  80].  

Five  N-­‐glycosylation  sites  were  predicted  using  the  NetNGlyc  1.0  server,  at  golden  tilefish  (gt)N8,  

gtN48,  gtN103,  gtN132,  and  gtN209  (Figure  3).    The  gtN8  motif  was  shared  with  other  Perciformes  and  

Pleuronectiformes  and  the  gtN132  motif  was  shared  with  all  other  screened  type  II  IL-­‐1β  proteins  in  

teleosts  (except  red  seabream).    The  gtN48  and  gtN209  motifs,  however,  were  shared  exclusively  with  red  

seabream  (Figures  3).    Two  cysteine  residues  were  observed,  at  gtC166  and  gtC235  (Figure  2).      The  first  

was  conserved  among  all  vertebrates,  while  the  latter  was  conserved  among  several  teleosts  (Figure  2;  

Figure  3).    There  was  high  conservation  among  the  12  β-­‐sheets  across  all  vertebrates  screened  in  the  

multiple  alignment,  particularly  in  sheets  3-­‐11  (Figure  2).    This  might  indicate  that  similar  folding  

patterns  and  tertiary  structure  have  been  upheld  throughout  evolutionary  lineages  [38].    Interestingly,  

there  appears  to  be  a  golden  tilefish-­‐specific  sequence  addition  between  sheets  5  and  6,  not  found  in  

any  other  screened  vertebrate  (Figure  2).  

However,  of  the  seven  residues  cited  as  being  essential  for  binding  to  the  IL-­‐1  receptor  in  

humans,  tilefish  shared  only  two  with  humans,  residues  human  (h)R4  and  hK103,  numbered  according  to  

the  location  of  the  mature  human  peptide  [81].    The  hK103  residue  was  highly  conserved  in  many  

Perciformes,  chicken,  and  the  mammals  screened  in  the  multiple  alignment,  but  substituted  with  an  

arginine  in  the  Salmoniformes,  Gadiformes,  and  Cypriniformes  (Figure  2).    Two  specific  sites  were  

previously  characterized  as  being  critical  for  human  IL-­‐10  ligand-­‐receptor  binding,  Site  A  and  Site  B  [82].    

Of  the  25  residues  involved  with  Van  der  Waals  force-­‐dependent  Site  A,  golden  tilefish  demonstrated  

conservation  with  11  of  them,  including  identity  with  7  amino  acids  (hS13,  hL31,  hQ32,  hG33,  hN129,  hP131,  

and  hQ142;  Figure  2).    Of  the  25  residues  important  for  Site  A  binding,  four  are  critical,  hR11,  hQ15,  hH30,  

Page 31: Graduate Theses and Dissertations Graduate School 11-4 ...

  18  

and  hQ32  [80].    Golden  tilefish  IL-­‐1β  had  a  similar  amino  acid  substitution  (asparagine)  at  hQ15,  and  

shared  identity  with  hQ32  (Figure  2).    Golden  tilefish  IL-­‐1β  showed  weaker  conservation  at  site  B,  

demonstrating  identity  with  four  of  the  important  hydrophilic  residues  (hR4,  hK103,  hN108,  hS152),  but  no  

conservation  with  the  remaining  17  residues  (Figure  2).  

The  lack  of  conservation  among  the  ligand-­‐receptor  binding  sites  may  indicate  high  diversity  in  

binding  mechanisms  and  sequences  and  limits  the  likelihood  of  cross-­‐reactivity  between  non-­‐placental  

mammals  and  other  vertebrates  [38,  83].    In  human  mutagenic  studies,  even  conserved  amino  acid  

substitutions  were  sufficient  to  interrupt  30-­‐50%  of  receptor  binding  activity,  with  a  complete  loss  of  

activity  when  the  critical  Site  A  hQ15  residue  was  mutated  [80].    Therefore,  even  though  the  hQ15  residue  

was  substituted  with  the  structurally  similar  arginine  in  tilefish,  this  might  be  enough  to  disrupt  similar  

receptor  binding  mechanisms  between  mammalian  and  teleost  IL-­‐1β  molecules.      

Golden  tilefish  IL-­‐1β  shared  high  amino  acid  percent  identity  with  other  Perciformes  studied,  

including  70%  amino  acid  identity  with  European  seabass,  70%  with  Mandarin  fish,  and  68%  identity  

with  barred  knifejaw  (Table  2;  all  common  names  used  and  their  associated  scientific  names  may  be  

found  in  Appendix  H).    Identity  was  significantly  less  well-­‐correlated  with  Cypriniformes  (33%),  

Elasmobranchs  (28%),  amphibians  (33%),  birds  (31%),  and  mammals  (29-­‐32%;  Table  3).    Low  identity  

with  the  smaller  spotted  catshark  was  anticipated,  as  it  is  one  of  the  most  ancestral  IL-­‐1β  molecules  

identified  [78].      

On  the  constructed  phylogenetic  tree,  teleost  fishes  clustered  separately  from  remaining  

vertebrates,  with  additional  subdivisions  between  Cypriniformes,  Gadiformes,  Salmoniformes,  and  the  

remaining  Perciformes,  which  clustered  with  Plueronectiformes  (Figure  4).      These  branchings  could  be  a  

result  of  the  Type  I  and  Type  II  division  of  IL-­‐1β  molecules.    Perciformes  typically  contain  Type  II  IL-­‐1β,  

while  Cypriniformes,  Elasmobranchs,  amphibians,  birds,  and  mammals  contain  Type  I  and  

Salmoniformes  and  Gadiformes  display  both  types  [78].    Type  II  IL-­‐1β  molecules  tend  to  be  shorter,  

Page 32: Graduate Theses and Dissertations Graduate School 11-4 ...

  19  

owing  to  their  lack  of  an  additional  exon  at  the  5’  end  of  the  protein,  and  generally  share  higher  

homology  among  one  another  as  compared  to  their  identity  with  known  Type  I  molecules  [76].    There  

was  an  apparent  split  in  the  teleost  portion  of  the  tree,  dividing  the  Perciformes,  Pleuronectiformes,  

Gadiformes,  and  Salmoniformes  Type  II  IL-­‐1  β  from  the  zebrafish,  amphibian,  bird,  and  mammal  Type  I  

IL-­‐1  β  [49].    Golden  tilefish  IL-­‐1β  exhibited  stronger  homology  with  Type  II  molecules,  is  short,  at  only  

251  amino  acids  in  length,  and  therefore,  likely  belonged  to  the  Type  II  group.      

 

3.2  Cloning  and  sequence  analysis  of  red  snapper  IL-­‐1β    

  Degenerate  primers  rsIL-­‐1b_F  and  rsIL-­‐1b_R  were  utilized  to  amplify  an  initial  430  bp  fragment  

from  red  snapper  spleen  cDNA,  according  to  PCR  conditions  given  (Table  1).    This  fragment  was  cloned,  

sequenced,  and  its  297  bp  coding  region  was  translated  into  a  99  amino  acid  sequence.    The  partial  

amino  acid  sequence  for  red  snapper  IL-­‐1β  had  good  homology  with  other  teleost  IL-­‐1β  proteins,  

including  European  seabass  (67%)  and  southern  Bluefin  tuna  (67%;  Table  3).    It  shared  64%  identity  with  

the  43%  of  golden  tilefish  IL-­‐1β  it  covered,  although  it  did  not  share  any  N-­‐glycosylation  sites  (Table  3;  

Figure  5).    The  partial  sequence  did  not  include  the  family  signature,  however,  it  did  include  β-­‐sheets  3-­‐

9,  the  majority  of  which  showed  significant  levels  of  conservation  with  other  vertebrate  species  (Figure  

2).    The  final  sequence  was  not  cloned  and  confirmed  from  multiple  fish,  as  there  was  sufficient  

information  to  provide  a  comparison  between  golden  tilefish  and  red  snapper  IL-­‐1β  for  the  purpose  of  

antibody  production  (Figure  5).  

 

3.3  Cloning  and  sequence  analysis  of  golden  tilefish  IL-­‐8  

  Degenerate  primers  gtIL-­‐8_F  and  gtIL-­‐8_R  were  designed  with  PriFi  and  used  to  amplify,  clone,  

and  sequence  an  initial  279  fragment  from  golden  tilefish  spleen  cDNA  (Table  1).    This  sequence  was  

then  used  to  design  primers  for  5’-­‐  and  3’-­‐RACE,  gtIL-­‐8_GSP2  and  gtIL-­‐8_GSP1,  respectively  (Table  1).    

Page 33: Graduate Theses and Dissertations Graduate School 11-4 ...

  20  

The  resulting  3’-­‐RACE  product  was  634  bp  in  length  and  provided  sufficient  sequence  information  for  

analysis,  without  attaching  a  5’-­‐RACE  product  (Figures  1b).    Instead,  the  3’-­‐RACE  product  was  assembled  

with  the  initial  fragment.    This  assembled  sequence  contained  a  172  bp  5’  UTR,  297  bp  translated  region,  

and  381  bp  3’  UTR,  which  contained  2  instability  motifs  (ATTTA),  two  endotoxin  motifs  (one  TTATTTAT  

and  one  ATATTTAT),  and  a  polyadenylation  signal  (ATTAAA),  ending  18  bp  upstream  of  the  poly-­‐A  tail.    

Gene-­‐specific  primers  were  designed  (gtfIL-­‐8_F  and  gtfIL-­‐8_R)  to  clone  and  confirm  the  IL-­‐8  coding  

sequence  from  spleen  cDNA  of  two  different  golden  tilefish  fish  according  to  the  PCR  parameters  given  

(Table  1).    This  resulted  in  a  479  bp  product  with  a  297  bp  coding  region,  which  was  translated  into  98  

amino  acids.    

  The  translated  portion  included  the  IL-­‐8  family  signature,  CXC-­‐[LIVMS]-­‐x(4,6)-­‐[LIVMFY]-­‐x(2)-­‐

[RKSEQNA]-­‐x-­‐[LIVM]-­‐x(2)-­‐[LIVMA]-­‐x(5)-­‐[STAG]-­‐x(2)-­‐C-­‐x(3)-­‐[EQ]-­‐[LIVM](2)-­‐x(9,10)-­‐C-­‐[LV]-­‐[DN],  as  

CrCIqteskpigRhIakVeliipnshCeetEIIatlkktgqevCLD  (Figure  6).    A  putative  cut  site  for  the  signal  peptide  

was  predicted  between  gtG22  and  gtM23  (Figure  6).    This  site  aligned  with  the  end  of  the  signal  peptide  

and  start  of  mature  protein  among  all  other  vertebrates  as  well  [48,  49].    No  N  glycosylation  sites  were  

found.    

The  sequence  contained  the  CXC  motif  but  lacked  the  ELR  motif,  the  latter  of  which  is  critical  for  

neutrophil  chemotaxis  and  angiogenesis  in  mammals  [45].    The  lack  of  a  complete  ELR  sequence  is  a  

characteristic  pattern  shared  by  many  fish  homologues  [42,  57,  84].    Golden  tilefish  had  ELH  in  place  of  

ELR,  as  did  striped  trumpeter,  European  seabass,  hapuku,  and  many  other  fishes  (Figure  7).      A  study  

utilizing  recombinant  proteins  for  the  ELR  region  in  black  seabream  concluded  that  the  ELR  motif  could  

not  be  the  only  residues  stimulating  chemotaxis  and  that  fish  species  lacking  this  motif  may  express  an  

ancestral  version  of  the  chemokine  [84].      Given  the  high  conservation  of  the  SLGV  motif  in  fishes  prior  

to  the  ELR-­‐CXC  motif,  L6  has  been  proposed  as  the  responsible  residue  for  inducing  chemotaxis  in  

Page 34: Graduate Theses and Dissertations Graduate School 11-4 ...

  21  

flounder  [57].    This  residue  was  conserved  in  golden  tilefish,  so  the  L6  functional  hypothesis  may  hold  for  

this  species  as  well  (Figure  7).        

  The  mature  golden  tilefish  IL-­‐8  sequence  shared  high  homology  with  other  Perciformes,  

including  striped  trumpeter  (86%  amino  acid  identity),  Mandarin  fish  (84%),  and  European  seabass  (84%;  

Table  2).    High  amino  acid  identity  was  seen  with  other  teleosts,  however,  lower  identities  were  shared  

with  Elasmobranchs  (43%),  rodents  (36%)  and  humans  (41%;  Table  3).    Teleosts  grouped  separately  from  

other  vertebrates  in  the  phylogenetic  tree,  with  a  division  noted  between  the  two  predicted  types  of  IL-­‐

8  protein  in  teleosts,  designated  by  branches  separating  Salmoniformes  and  Cypriniformes  from  other  

bony  fishes  (Figure  8).    As  anticipated,  golden  tilefish  clustered  with  other  Perciformes,  with  high  

bootstrapping  statistical  support  (Figure  9).  

 

3.4  Cloning  and  sequence  analysis  of  red  snapper  IL-­‐8  

  PriFi-­‐designed  degenerate  primers  rsIL-­‐8_F  and  rsIL-­‐8_R  were  used  to  amplify,  clone,  and  

sequence  an  initial  277  bp  fragment  from  red  snapper  spleen  cDNA  according  to  the  PCR  conditions  

listed  (Table  1).    Gene-­‐specific  primers,  rsIL-­‐8_GSP2  and  rsIL-­‐8_GSP1,  were  designed  for  5’-­‐  and  3’-­‐RACE  

based  upon  this  fragment  yielding  a  5’-­‐RACE  product  of  279  bp  and  a  784  bp  3’-­‐RACE  product  with  a  127  

bp  overlap  between  the  two  sequences  (Table  1;  Figure  1c).    When  aligned,  this  produced  a  937  bp  

sequence  with  a  249  bp  5’  UTR,  300  bp  translated  region,  and  a  388  bp  3’  UTR  containing  3  instability  

motifs  (ATTTA),  two  endotoxin  responsive  motifs  (ATATTTAT)  and  one  polyadenylation  signal,  ending  11  

bp  upstream  of  the  poly-­‐A  tail.    Gene-­‐specific  primers  rsfIL-­‐8_F  and  rsfIL-­‐8_R  were  designed  to  amplify,  

clone,  and  confirm  the  coding  region  from  two  different  red  snapper  fish  samples  (Table  1).    This  

confirmed  the  sequence  for  mature  red  snapper  IL-­‐8,  with  a  translated  region  of  300  bp,  encoding  99  

amino  acids.      

Page 35: Graduate Theses and Dissertations Graduate School 11-4 ...

  22  

  The  protein  sequence  contained  the  IL-­‐8  family  signature  in  the  form:  

CrCIqtesrpigRhIgkVelippnshCeetEIIatlkktgqevCLD  (Figure  6).    The  sequence  contained  the  CXC  motif  but  

again  lacked  the  ELR  motif,  which  was  substituted  with  ELH  (Figure  9).    Signal  P  4.1  predicted  a  cut  site  

for  the  signal  peptide  at  red  snapper  (rs)G22  and  rsM23,  which  aligned  well  with  the  cut  site  in  other  

vertebrates  (Figures  6).    As  with  the  tilefish  sequence,  no  N  glycosylation  sites  were  predicted.    

  The  mature  red  snapper  IL-­‐8  protein  shared  the  highest  homology  with  humphead  snapper  (92%  

amino  acid  identity),  European  seabass  (88%),  and  Mandarin  fish  (88%;  Table  2).    It  was  also  84%  

identical  to  the  mature  golden  tilefish  IL-­‐8  sequence,  though  golden  tilefish  was  slightly  more  

homologous  to  mammals  than  red  snapper  (Table  3;  Figure  10).    The  red  snapper  partial  sequence  

grouped  with  other  Perciform  IL-­‐8  proteins  on  the  rooted  phylogenetic  tree,  and  branched  before  the  

European  seabass  and  red  seabream  node  (Figure  8).  

 

3.5  Cloning  and  sequence  analysis  of  golden  tilefish  IL-­‐10    

  Degenerate  primers  gtIL-­‐10_F  and  gtIL-­‐10_R  were  used  to  amplify  a  441  bp  fragment  from  

golden  tilefish  spleen  cDNA  using  the  PCR  conditions  listed  (Table  1).    This  product  was  cloned,  

sequenced,  and  used  to  design  primers  for  5’-­‐  and  3’-­‐RACE,  though  3’-­‐RACE  did  not  yield  sufficient  

sequence  data,  and  therefore,  was  omitted  from  the  table  (Table  1).    5’-­‐RACE  resulted  in  513  bp  

sequence,  which  had  a  197  bp  overlap  with  the  originally  cloned  fragment  (Figure  1d).    The  assembled  

sequence  consisted  of  714  bp  and  a  493  bp  coding  region  that  translated  into  164  amino  acids.      

The  first  teleost  IL-­‐10  family  signature  motif  was  conserved  as  follows:  L-­‐[FILMV]-­‐x(3)-­‐[ILV]-­‐x(3)-­‐

[FILMV]-­‐x(5)-­‐C-­‐x(5)-­‐[LIMV]-­‐[LIMV]-­‐x(3)-­‐L-­‐x(2)-­‐[IV]-­‐[FILMV]  as  LLdqtIehsLktpfaChainsILgfyLgtVL  (Figure  

11).    Only  a  portion  of  the  second  teleost  IL-­‐10  family  signature  was  present  (Figure  12).    The  partial  

sequence  shared  high  homology  with  other  teleost  IL-­‐10,  including  European  seabass  (76%)  amino  acid  

identity,  gilthead  seabream  (76%),  and  green  spotted  puffer  (66%;  Table  3).      The  final  sequence  was  

Page 36: Graduate Theses and Dissertations Graduate School 11-4 ...

  23  

confirmed  via  sequencing  of  the  IL-­‐10  PCR  fragment  from  two  additional  golden  tilefish  spleen  samples  

(Figure  12).    

 

3.6  Cloning  and  sequence  analysis  of  red  snapper  IL-­‐10  

  An  initial  406  bp  fragment  was  amplified  from  red  snapper  spleen  cDNA  by  degenerate  primers  

rsIL-­‐10_F  and  rsIL-­‐10_R  under  the  PCR  conditions  listed  (Table  1).    The  fragment  was  cloned,  sequenced,  

and  utilized  to  design  gene-­‐specific  primers  for  5’-­‐  and  3’-­‐RACE  (Table  1).    5’-­‐RACE  yielded  a  519  bp  

fragment  and  3’-­‐RACE  yielded  a  711  bp  fragment,  with  a  71  bp  overlap  between  the  two  products  

(Figure  1d).    The  assembled  RACE  sequence  contained  1108  bp  with  a  564  bp  translated  region,  180  bp  

5’  UTR,  and  a  364  bp  3’  UTR.    The  3’  UTR  contained  three  instability  motifs  (ATTTA),  one  endotoxin  

responsive  motif  (ATATTTAT),  and  one  polyadenylation  motif  (AATAAA),  ending  15  bp  upstream  of  the  

poly-­‐A  tail.    Gene-­‐specific  primers  were  designed  to  confirm  the  IL-­‐10  coding  sequence  from  two  

different  red  snapper  spleen  cDNA  samples,  using  the  PCR  conditions  given  (Table  1).      The  amplified  

sequence  was  647  bp  in  length  with  a  564  bp  coding  region  that  translated  into  187  amino  acids.  

  The  translated  red  snapper  IL-­‐10  sequence  contained  two  teleost  IL-­‐10  family  signatures  in  the  

following  forms:  L-­‐[FILMV]-­‐x(3)-­‐[ILV]-­‐x(3)-­‐[FILMV]-­‐x(5)-­‐C-­‐x(5)-­‐[LIMV]-­‐[LIMV]-­‐x(3)-­‐L-­‐x(2)-­‐[IV]-­‐[FILMV]  as  

LLdqtIehsLktpfaChainsILgfyLgtVL  and  KA-­‐x(2)-­‐E-­‐x-­‐D-­‐[ILV]-­‐[FLY]-­‐[FILMV]-­‐x(2)-­‐[ILMV]-­‐[EKQZ]  as  

KAmgElnLLFnyIE  [48,  66,  68].    Two  potential  N-­‐glysoylation  sites  were  predicted  using  the  NetNGlyc  1.0  

sever  at  rsN13  and  rsN146,  the  latter  of  which  was  shared  among  all  Perciformes  analyzed  (Figure  13).    Of  

note,  however,  is  rsN146  as  it  was  within  alpha  helix  E,  and  therefore,  unlikely  to  be  functional.    Mouse  

and  human  IL-­‐10  have  an  N  glycosylation  site  between  the  D  and  E  chains,  on  the  surface  of  the  

molecule,  where  it  may  be  readily  available  for  use  [85].  

A  putative  cut  site  for  the  signal  peptide  was  predicted  between  residues  rsC22  and  rsT23  using  

SignalP  4.1,  which  occurred  after  the  human  cut  site  in  the  multiple  alignment  (Figure  11).    Six  conserved  

Page 37: Graduate Theses and Dissertations Graduate School 11-4 ...

  24  

cysteine  residues  were  observed,  including  two  teleost-­‐specific  cysteines  (rsC26  and  rsC31)  and  four  

vertebrate  conserved  cysteines  responsible  for  forming  intra-­‐chain  disulfide  bonds  in  humans  (rsC30,  

rsC80,  rsC130,  and  rsC136;  Figure  11;  Figure  13).    Within  fish  there  was  good  conservation  among  all  six  

alpha  helical  chains,  however,  discrepancies  arose,  particularly  in  chain  D  and  E,  when  the  comparison  

was  extended  to  chicken,  frog,  and  mammals  (Figure  11;  Figure  13).  

  There  are  28  conserved  hydrophobic  residues  in  humans,  critical  for  stabilizing  the  molecule  

[86].      Of  these,  10  were  identical  in  red  snapper  IL-­‐10  and  16  were  substituted  for  other  hydrophobic  

amino  acids  (Figure  11).    There  are  six  conserved  amino  acids  between  viral  and  human  IL-­‐10,  which  

implies  their  importance,  four  of  which  were  identical  in  red  snapper  and  one  of  which  was  similarly  

substituted  by  another  amino  acid  [86].    In  addition,  red  snapper  IL-­‐10  contained  all  five  amino  acid  

residues  conserved  between  human  IL-­‐10  and  IFNγ,  residues  also  implicated  in  core  stabilization  [87].    

The  critical  core-­‐stabilizing  surface  ion  pair  hR27-­‐hE151  was  perfectly  conserved  in  red  snapper  [85].    The  

hI69  residue,  responsible  for  the  molecules  immunostimulatory  properties  in  humans,  was  not  conserved  

directly  in  the  multiple  sequence  alignment,  however,  all  teleosts,  including  red  snapper,  had  an  

isoleucine  residue  preceding  hI69  by  one  amino  acid  [88].  

    In  humans,  IL-­‐10  cellular  action  requires  two  receptor  chain  assemblies  to  occur  sequentially  on  

the  cell  surface,  at  Site  1  and  Site  2  [89].    There  are  nine  crucial  residues  for  binding  to  IL-­‐10R1  in  the  Site  

1  region  [88].    Red  snapper  shared  four  of  them,  hP20,  hR27,  hK138,  and  hE151  [89].    However,  of  the  ten  

important  residues  for  binding  IL-­‐10/IL-­‐10R2  at  Site  2,  only  hH90  was  conserved  in  red  snapper  [89].    The  

hR24  residue  has  a  shared  importance  in  both  Site  1  and  Site  2  binding  [89].    While  snapper  did  not  have  

this  residue,  it  contained  an  arginine  one  amino  acid  further  along  the  sequence  (Figure  11).    

Importantly,  IL-­‐10R1  has  been  confirmed  in  several  fish  species,  including  grass  carp,  goldfish,  and  

zebrafish,  definitive  teleost  orthologues  to  human  IL-­‐10R2  have  not  been  confirmed  [67]  

Page 38: Graduate Theses and Dissertations Graduate School 11-4 ...

  25  

  Red  snapper  IL-­‐10  shared  the  highest  amino  acid  identity  with  European  seabass  (77%),  gilthead  

seabream  (76%),  and  cod  (66%;  Table  2).    Lower  homology  was  observed  with  Cypriniformes  (45-­‐50%),  

frog  (37%),  chicken  (33%),  and  mammals  (29%;  Table  3).    Red  snapper  IL-­‐10  clustered  with  European  

seabass  after  phylogenetic  analysis  and  there  was  a  clear  division  between  Cypriniformes  and  other  

teleosts,  as  well  as  between  fishes,  chicken,  and  mammals  (Figure  14).    There  was  high  homology  

between  red  snapper  and  golden  tilefish  IL-­‐10  molecules,  with  99%  identity  over  the  sequenced  portion  

of  the  protein  (Table  3;  Figure  12).    The  high  identity  is  unusual,  given  the  typical  range  of  amino  acid  

homology  in  fishes,  however,  the  nucleotide  sequences  for  golden  tilefish  and  red  snapper  were  

confirmed  from  at  least  two  fish,  each.    The  high  identity  could  be  an  artifact  of  the  Fish  Genome  

Duplication  event,  where  similar  gene  copies  have  been  conserved  in  these  fish,  or,  it  is  possible  the  

molecule  is  highly  conserved  among  Perciformes,  which  may  become  more  apparent  as  more  fish  IL-­‐10  

proteins  are  sequenced.    

 

3.7  Cloning  and  sequence  analysis  of  golden  tilefish  TNFα    

  Degenerate  primers  were  used  to  obtain  a  572  bp  fragment  from  golden  tilefish  spleen  cDNA  

isolated  from  spleen  under  the  PCR  conditions  given  (Table  1).    This  fragment  was  cloned,  sequenced,  

and  used  to  create  gene-­‐specific  primers  for  5’-­‐/3’-­‐RACE,  although  the  5’-­‐RACE  primer  failed  to  provide  

sufficient  data  for  inclusion  (Table  1).    The  3’-­‐RACE  product  was  779  bp  in  length,  and  contained  the  

entirety  of  the  initially  isolated  fragment.    The  assembled  product  was  795  bp  in  length  with  a  290  bp  

coding  region  that  was  translated  into  96  amino  acids.    The  resulting  amino  acid  sequence  corresponded  

with  the  C  terminal  of  the  TNFα  protein,  and  while  it  was  not  cloned  and  confirmed,  sufficient  data  was  

obtained  to  indicate  significant  homology  with  a  portion  of  red  snapper  TNFα  and  that  of  other  

Perciformes  (Table  3;  Figure  15;  Figure  16).  

 

Page 39: Graduate Theses and Dissertations Graduate School 11-4 ...

  26  

3.8  Cloning  and  sequence  analysis  of  red  snapper  TNFα      

  An  initial  474  bp  fragment  was  obtained  from  red  snapper  spleen  cDNA  using  degenerate  

primers  rsTNFa_F  and  TNFa_R  under  the  PCR  conditions  given  (Table  1).    This  fragment  was  cloned,  

sequenced,  and  used  to  design  primers  for  5’-­‐/3’-­‐RACE  (Table  1).    The  5’-­‐RACE  reaction  resulted  in  a  744  

bp  product,  the  3’-­‐RACE  reaction  yielded  a  777  bp  product,  and  there  was  a  91  bp  overlap  between  the  

two  fragments  (Figure  1e).    The  assembled  product  was  1431  bp  in  length,  with  a  764  bp  translated  

region,  156  bp  5’  UTR,  and  a  510  bp  3’  UTR.      The  3’  UTR  contained  3  instability  motifs  (ATTTA),  two  

endotoxin  responsive  motifs  (TTATTTAT  and  ATATTTAT)  and  one  polyadenylation  signal    (ATTAAA)  

ending  4  bp  upstream  from  the  poly-­‐A  tail.    This  product  was  used  to  design  primers  to  amplify,  clone,  

and  confirm  the  coding  sequence  of  red  snapper  TNFα  sequence  from  the  spleen  cDNA  from  two  

different  fish  (Table  1).    The  resulting  fragment  was  884  bp  in  length  with  a  768  bp  coding  region,  which  

translated  into  a  256  amino  acid  product.      

  The  red  snapper  TNFα  sequence  contained  the  TNFα  family  signature,  [LV]-­‐x-­‐[LIVM]-­‐x(3)-­‐G-­‐

[LIVMF]-­‐Y-­‐[LIVMFY](2)-­‐x(2)-­‐[QEKHL],  as  ivIpqtGLYFVysQ  (Figure  15).    Two  cysteine  residues  were  

observed,  at  rsC156  and  rsC196  (Figure  17).    These  appear  to  correspond  to  the  two  human  cysteine  

residues  responsible  for  proper  folding  of  mature  TNFα.    In  the  multiple  sequence  alignment,  red  

snapper  C156  appeared  one  residue  before  the  mammalian  cysteine,  while  the  position  of  rsC196  was  

conserved  among  all  vertebrates  examined  (Figure  15).    A  transmembrane  domain  was  predicted  by  

TMPred  from  amino  acid  rsV35  through  rsW54  (Figure  15).    This  corresponded  perfectly  with  the  

predicted  transmembrane  domain  in  other  Perciformes,  suggesting  that  red  snapper  TNFα  may  be  

membrane-­‐bound  (Figures  17).    The  TACE  cleavage  site  was  identified  at  rsT86-­‐rsL87,  a  few  residues  later  

than  where  it  occurs  in  Cypriniformes  and  mammals  (Figure  17).    In  having  both  a  predicted  

transmembrane  domain  and  TACE  cleavage  site,  it  is  possible  that  red  snapper  TNFα  is  secreted  in  a  

manner  similar  to  its  mammalian  orthologue,  with  the  mature  molecule  freely  secreted  from  its  

Page 40: Graduate Theses and Dissertations Graduate School 11-4 ...

  27  

membrane-­‐bound  state  after  cleavage  at  the  TACE  site  [38].    No  glycosylation  sites  were  predicted.    High  

conservation  in  the  C  terminal  region  of  the  protein  was  observed  among  all  analyzed  vertebrates,  

common  to  the  TNF  homology  domain  (THD;  Figure  15;  Figure  17).  

  Red  snapper  TNFα  shared  the  highest  homology  with  barred  knifejaw  (81%  amino  acid  identity),  

striped  trumpeter  (80%),  and  orange-­‐spotted  grouper  (78%;  Table  2).    Homology  was  reduced  between  

red  snapper  and  Salmoniformes  (40%),  Cypriniformes  (38-­‐46%),  frog  (27%),  chicken  (42%,  over  a  partial  

sequence),  and  mammals  (35-­‐36%;  Table  3).      This  was  not  surprising,  given  the  high  degree  of  genetic  

divergence  noted  in  TNFα  genes  [62].    Carp,  rainbow  trout,  and  salmon  all  exhibit  multiple  isoforms  of  

the  gene,  likely  as  a  result  of  a  tetraploidation  event  which  occurred  along  these  lineages,  while  the  two  

forms  discovered  in  bluefin  tuna  are  indicative  of  gene  duplication  events  [60,  62,  90,  91].    Phylogenetic  

analysis  suggested  clustering  of  red  snapper  with  red  seabream,  and  this  pair  clustered  with  other  

Perciformes  (Figure  18).    While  the  analysis  displayed  vague  distinctions  among  Perciformes,  there  was  

clear  clustering  between  the  Perciform  and  Pleuronectiform  node,  and  Tetraodontiformes,  

Salmoniformes,  Cypriniformes,  and  tetrapods  (Figure  18).  Red  snapper  TNFα  shared  80%  amino  acid  

identity  with  the  partial  TNFα  sequence  from  golden  tilefish,  as  may  be  expected  as  much  of  this  

covered  the  THD  region  (Table  3;  Figure  16).    

 

 

                             

Page 41: Graduate Theses and Dissertations Graduate School 11-4 ...

  28  

   

   

   

     

     Figure  1:  Schematic  of  PCR  results  including  the  locations  of  the  coding  region,  5’  UTR,  3’  UTR,  locations  at  which  primers  bound,  and  the  size  and  position  of  the  resulting  fragments.      Genes  shown  are:  a)  golden  tilefish  IL-­‐1β,  b)  golden  tilefish  IL-­‐8,  c)  red  snapper  IL-­‐8,  d)  red  snapper  IL-­‐10,  and  e)  red  snapper  TNFα.  

Page 42: Graduate Theses and Dissertations Graduate School 11-4 ...

  29  

     

 

 

   

   

Page 43: Graduate Theses and Dissertations Graduate School 11-4 ...

  30  

   

     Figure  2.  Clustal  W  multiple  sequence  alignment  of  representative  vertebrate  IL-­‐1β,  shaded  with  BOXSHADE.    The  family  signature  is  boxed,  the  mammalian  ICE  cleavage  site  is  marked  with  an  arrow,  and  the  vertebrate  conserved  cysteine  is  marked  with  an  asterisk  (*).    The  12  β-­‐sheets  are  underlined  and  numbered.    Residues  involved  with  ligand-­‐receptor  binding  at  Site  A  are  labeled  with  an  “a”  above  the  alignment  and  those  involved  with  binding  at  Site  B  are  labeled  with  a  “b”.    The  GenBank  accession  codes  used  were:  European  seabass  (CAC41006.1),  red  seabream  (AAP33156.1),  southern  bluefin  tuna  (AGB24759.1),  olive  flounder  (BAB8682.1),  Atlantic  salmon  (NP_001117054.1),  rainbow  trout  (CAA11684.1),  Atlantic  cod  (ABV59377.1),  haddock  (CAJ58294.1),  zebrafish  (AAH98597.1),  smaller  spotted  catshark  (CAC09453.1),  African  clawed  frog  (CAB53499.1),  chicken  (CAA75239.1),  mouse  (AAA39276.1),  and  human  (AAA59135.1).                                  

Page 44: Graduate Theses and Dissertations Graduate School 11-4 ...

  31  

Table  2.  Percent  amino  acid  identity  between  confirmed,  mature  amino  acid  sequence  and  the  top  five  most  homologous  fish  species,  according  to  BLASTp  analysis.      

   

                       

Target Species Highest.scoring.teleosts Order Query.cover E.valueAmino.acid.identity.(%)

European)seabass)|)CAC41006.1 Perciformes 98% 5.00E=119 70%

Mandarin)fish)|)AAV65041.1 Perciformes 100% 1.00E=117 70%

Barred)knifejaw)|)ACH87392.1 Perciformes 100% 3.00E=117 68%

Striped)trumpeter)|)ACQ99510.1 Perciformes 98% 2.00E=116 70%

Red)seabream)|)AAP33156.1 Perciformes 97% 1.00E=108 67%

European)seabass)|)CAM32186.1 Perciformes 100% 8.00E=53 84%

Striped)trumpeter)|)ACQ99511.1 Perciformes 98% 3.00E=52 86%

Mandarin)fish)|)AFK65606.1 Perciformes 98% 7.00E=52 84%

Humphead)snapper)|)AGV99968.1 Perciformes 96% 7.00E=51 83%

Red)seabream)|)ADK35756.1 Perciformes 100% 8.00E=51 84%

European)seabass)|)CAM32186.1 Perciformes 98% 8.00E=53 88%

Humphead)snapper)|)AGV99968.1 Perciformes 95% 2.00E=55 92%

Mandarin)fish)|)AFK65606.1 Perciformes 97% 4.00E=55 88%

Red)seabream)|)ADK35756.1 Perciformes 100% 5.00E=55 87%

Striped)trumpeter)|)ACQ99511.1 Perciformes 97% 2.00E=54 87%

European)seabass)|)ABH09454.1 Perciformes 99% 2.00E=96 77%

Gilthead)seabream)|)AGS55345.1 Perciformes 92% 8.00E=93 76%

Atlantic)cod)|)ABV64720.1 Gadiformes 84% 7.00E=74 66%

Green)spotted)puffer)|)CAD67773.1 Tetraodontiformes 87% 2.00E=73 65%

Rainbow)trout)|)NP_001232028.1 Salmoniformes 83% 2.00E=57 57%

Barred)knifejaw)|)ACM69339.1 Perciformes 98% 2.00E=158 81%

Striped)trumpeter)|)ACQ99509.1 Perciformes 98% 1.00E=155 80%

Orange=spotted)grouper)|)AEH59794.1 Perciformes 98% 1.00E=153 78%

Large)yellow)croaker)|)ABK62876.1 Perciformes 98% 2.00E=153 79%

European)seabass)|)AAZ20770.1 Perciformes 98% 8.00E=152 77%

IL=1β

IL=10)

TNFα

IL=8

Golden)tilefish)

Golden)tilefish)

Red)snapper

Red)snapper

Red)snapper

Page 45: Graduate Theses and Dissertations Graduate School 11-4 ...

  32  

Table  3:  Percent  amino  acid  identity  among  vertebrate  groups.    Asterisks  (*)  denote  partial  sequences.      

Snapper TilefishGolden  tilefish  (F)   Bony  fishes Perciformes 64 xRed  snapper  (C)  * Bony  fishes Perciformes x 64

European  seabass  |  CAC80553.1 Bony  fishes Perciformes 67 70Red  seabream  |  AAP33156.1 Bony  fishes Perciformes 66 67

Southern  bluefin  tuna  |  AGB24759.1 Bony  fishes Perciformes 67 63Olive  flounder  |  BAB8682.1 Bony  fishes Pleuronectiformes 63 56

Atlantic  salmon  |  NP_001117054.1 Bony  fishes Salmoniformes 61 51Rainbow  trout  |  CAA1164.1 Bony  fishes Salmoniformes 59 51Atlantic  cod  |  ABV59377.1 Bony  fishes Gadiformes 61 51Haddock  |  CAJ58294.1 Bony  fishes Gadiformes 61 51Zebrafish  |  AAH98597.1 Bony  fishes Cypriniformes 45 33

Smaller  spotted  catshark  |  CAC09435.1 Cartilaginous  fishes Carcharhiniformes 40 28African  clawed  frog  |  CAB53499.1 Amphibians Anura 38 33

Chicken  |  CAA75239.1 Birds Galiformes 38 31Mouse  |  AAA39276.1 Mammals Rodentia 42 32Human  |  AAA59135.1 Mammals Primates 47 29Golden  tilefish  (H)   Bony  fishes Perciformes 84 xRed  snapper  (C)   Bony  fishes Perciformes x 84

European  seabass  |  CAM32186.1 Bony  fishes Perciformes 88 84Red  seabream  |  ADK35756.1 Bony  fishes Perciformes 87 84Olive  flounder  |  AAL05442.1 Bony  fishes Pleuronectiformes 69 70Atlantic  salmon  |  ADK11046.1 Bony  fishes Salmoniformes 58 57Rainbow  trout  |  AAO25646.1 Bony  fishes Salmoniformes 61 59

Haddock  |  CAD97422.2 Bony  fishes Gadiformes 68 74Silver  carp  |  AEX32919.1 Bony  fishes Cypriniformes 63 59

Zebrafish  |  XP_001342606.2 Bony  fishes Cypriniformes 59 58Smaller  spotted  catshark  |  CAD91126.3 Cartilaginous  fishes Carcharhiniformes 45 43

African  clawed  frog  |  AEB96252.1 Amphibians Anura 54 46Chicken  |  ABD49203.2 Birds Galiformes 47 49Marmot  |  ABY67262.1 Mammals Rodentia 35 36Human  |  AAH13615.1 Mammals Primates 40 41Golden  tilefish  (T8)* Bony  fishes Perciformes 99 xRed  snapper  (C)   Bony  fishes Perciformes x 99

European  seabass  |  CAK29522.1 Bony  fishes Perciformes 77 78Gilthead  seabream  |  AGS55345.1 Bony  fishes Perciformes 76 76Green  spotted  puffer  |  CAD67773.1 Bony  fishes Tetraodontiformes 65 66Atlantic  salmon  |  ABM46994.1* Bony  fishes Salmoniformes 50 50

Atlantic  cod  |  ABV64720.1 Bony  fishes Gadiformes 66 66Grass  carp  |  AEA50953.1 Bony  fishes Cypriniformes 45 43Zebrafish  |  AAW78362.1 Bony  fishes Cypriniformes 50 47

African  clawed  frog  |  CAE92388* Amphibians Anura 37 35Chicken  |  CAF18432.1 Birds Galiformes 33 32Rat  |  AAA41425.1 Mammals Rodentia 29 28

Human  |  AAI04254.1 Mammals Primates 29 28Golden  tilefish  (H)* Bony  fishes Perciformes 80 xRed  snapper  (C)   Bony  fishes Perciformes x 80

European  seabass  |  AAZ20770.1 Bony  fishes Perciformes 77 90Red  seabream  |  AAP76392.1 Bony  fishes Perciformes 77 79

Pacific  bluefin  tuna  |  BAG72141.1 Bony  fishes Perciformes 68 85Fugu  |  NP_001033074.1 Bony  fishes Tetraodontiformes 64 69

Olive  flounder  |  BAA94970.1 Bony  fishes Pleuronectiformes 65 72Rainbow  trout  |  CAB92316.1 Bony  fishes Salmoniformes 48 61Grass  carp  |  ADY80577.1 Bony  fishes Cypriniformes 46 45Zebrafish  |  NP_998024.2 Bony  fishes Cypriniformes 38 45

African  clawed  frog  |  BAF95747.1 Amphibians Anura 27 26Chicken  |  AEW10516.1* Birds Galiformes 42 36Mouse  |  BAA19513.1 Mammals Rodentia 35 43Human  |  NP_000585.2 Mammals Primates 36 43

IL-­‐10

TNFα

Cytokine Species Amino  acid  identity  (%)

IL-­‐1β

IL-­‐8

OrderVertebrate  group

 

Page 46: Graduate Theses and Dissertations Graduate School 11-4 ...

  33  

 

   Figure  3.  Multiple  sequence  alignment  of  golden  tilefish  IL-­‐1β  with  the  top  five  most  homologous  species,  according  to  BLASTP.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified  by  Clustal  Omega.    Predicted  N-­‐glycosylation  sites  are  boxed  and  both  the  family  signature  motif  and  conserved  cysteine  residues  are  shaded  in  grey.    The  locations  corresponding  to  the  12  β-­‐sheets  are  underlined,  with  the  corresponding  human  sequence  below  along  with  the  identifying  sheet  number.    The  mammalian  ICE  cleavage  site  is  marked  with  an  arrow.    The  GenBank  accession  codes  used  were:  European  seabass  (CAC41006.1),  mandarin  fish  (AAV65041.1),  barred  knifejaw  (ACH87392.1),  striped  trumpeter  (ACQ99510.1),  and  red  seabream  (AAP33156.1).    

Page 47: Graduate Theses and Dissertations Graduate School 11-4 ...

  34  

   

Figure  4.  Rooted  phylogenetic  tree  of  vertebrate  IL-­‐1β  proteins,  produced  in  MEGA  6  with  a  NJ  method.    The  p-­‐distance  method  was  used,  bootstrapped  10,000  times,  with  results  reflected  at  the  branch  nodes.    Partial  sequences  were  excluded  and  human  IFNγ  was  used  as  an  outgroup.    The  Genbank  Accession  codes  used  were:  European  seabass  (CAC80553.1),  red  seabream  (AAP33156.1),  southern  bluefin  tuna  (AGB24759.1),  olive  flounder  (BAB8682.1),  Atlantic  salmon  (NP_01117054.1),  rainbow  trout  (CAA11684.1),  Atlantic  cod  (ABV59377.1),  haddock  (CAJ58294.1),  zebrafish  (AAH98597.1),  smaller  spotted  catshark  (CAC09435.1),  African  clawed  frog  (CAB53499.1),  chicken  (CAA75239.1),  mouse  (AAA39276.1),  human  (AAA59135.1)  and  human  IFNγ  (AAB59534.1).                                                    

Page 48: Graduate Theses and Dissertations Graduate School 11-4 ...

  35  

   

Figure  5.  Clustal  Omega  alignment  between  golden  tilefish  and  red  snapper  IL-­‐1β.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified  and  the  locations  of  the  β-­‐sheets  covered  by  both  sequences  are  underlined  and  numerated,  corresponding  to  their  sheet  number  in  the  complete  protein  (1-­‐12).                        

Page 49: Graduate Theses and Dissertations Graduate School 11-4 ...

  36  

   

Figure  6.  Clustal  W  multiple  sequence  alignment  of  representative  vertebrate  IL-­‐8,  shaded  with  BOXSHADE.    The  family  signature  is  boxed,  the  ELR  motif  is  marked  by  carrots  (^)  and  the  CXC  motif  is  marked  using  asterisks  (*).    The  putative  cut  site  for  the  signal  peptide  is  marked  with  an  arrow  above  the  sequence.    The  GenBank  accession  codes  used  were:  European  seabass  (CAM32186.1),  red  seabream  (ADK35756.1),  olive  flounder  (AAL05442.1),  Atlantic  salmon  (ADK11046.1),  rainbow  trout  (AAO25646.1),  haddock  (CAD97422.2),  silver  carp  (AEX32919.1),    zebrafish  (XP_001342606.2),  smaller  spotted  catshark  (CAD91126.3),  African  clawed  frog  (AEB896252.1),  chicken  (ABD49203.2),  marmot  (ABY67262.1),  and  human  (AAH13615.1).                                        

Page 50: Graduate Theses and Dissertations Graduate School 11-4 ...

  37  

 

   Figure  7.    Multiple  sequence  alignment  of  golden  tilefish  IL-­‐8  with  the  top  five  most  homologous  species,  according  to  BLASTP.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified  by  Clustal  Omega.    The  ELR  motif  is  boxed,  the  CXC  motif  is  underlined,  the  putative  cut  site  for  the  signal  peptide  is  marked  with  an  arrow,  and  the  family  signature  is  shaded  in  grey.    The  GenBank  accession  codes  used  were:  European  seabass  (CAM32186.1),  Mandarin  fish  (AFK65606.1),  humphead  snapper  (AGV99968.1),  red  seabream  (ADK35756.1),  and  striped  trumpeter  (ACQ99511.1).                                    

Page 51: Graduate Theses and Dissertations Graduate School 11-4 ...

  38  

   Figure  8.  Rooted  phylogenetic  tree  of  vertebrate  IL-­‐8  proteins,  produced  in  MEGA  6  with  a  NJ  method.    The  p-­‐distance  method  was  used,  bootstrapped  10,000  times,  with  results  reflected  at  the  branch  nodes.    Partial  sequences  were  excluded  and  human  IFNγ  was  used  as  an  outgroup.    The  Genbank  accession  codes  used  were:  European  seabass  (CAM32186.1),  red  seabream  (ADK35756.1),olive  flounder  (AAL05442.1),  Atlantic  salmon  (ADK11046.1),  rainbow  trout  (AAO25646.1),  haddock  (CAD97422.2),  silver  carp  (AEX32919.1),  zebrafish  (XP_001342606.2),  smaller  spotted  catshark  (CAD91126.3),  African  clawed  frog  (AEB96252.1),  chicken  (ABD49203.2),  marmot  (ABY67262.1),  human  (AAH13615.1)  and  human  IFNγ  (AAB59534.1).                                              

Page 52: Graduate Theses and Dissertations Graduate School 11-4 ...

  39  

   Figure  9.    Multiple  sequence  alignment  of  red  snapper  IL-­‐8  with  the  top  five  most  homologous  species,  according  to  BLASTP.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified  by  Clustal  Omega.    The  ELR  motif  is  boxed,  the  CXC  motif  is  underlined,  the  putative  cut  site  for  the  signal  peptide  is  marked  with  an  arrow,  and  the  family  signature  is  shaded  in  grey.    The  GenBank  accession  codes  used  were:  European  seabass  (CAM32186.1),  Mandarin  fish  (AFK65606.1),  humphead  snapper  (AGV99968.1),  red  seabream  (ADK35756.1),  and  striped  trumpeter  (ACQ99511.1).                                                                        

Page 53: Graduate Theses and Dissertations Graduate School 11-4 ...

  40  

   Figure  10.  Clustal  Omega  alignment  between  golden  tilefish  and  red  snapper  IL-­‐8.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified  and  the  position  of  the  ELR  and  CRC  motifs  are  underlined.                                                                        

Page 54: Graduate Theses and Dissertations Graduate School 11-4 ...

  41  

 

   

   

   Figure  11.  Clustal  W  multiple  sequence  alignment  of  representative  vertebrate  IL-­‐10,  shaded  with  BOXSHADE.    The  family  signature  is  boxed,  the  four  vertebrate  conserved  cysteines  are  marked  with  asterisks  (*),  and  the  alpha  helical  domains  are  underlined  and  lettered.    The  putative  cut  site  for  the  human  signal  peptide  is  marked  with  an  arrow  below  the  alignment  and  the  putative  cut  site  for  the  snapper  signal  peptide  is  marked  with  an  arrow  above  the  alignment.    Below  the  alignment,  residues  critical  for  ligand-­‐receptor  binding  at  Site  1  are  marked  with  a  “1”  and  those  critical  for  ligand-­‐receptor  binding  at  Site  2  are  labeled  with  a  “2.”    The  residue  responsible  for  hIL-­‐10  immunostimulatory  properties  is  marked  with  an  “I”  below  the  sequence.    Above  the  sequence,  hydrophobic  residues  are  

Page 55: Graduate Theses and Dissertations Graduate School 11-4 ...

  42  

marked  with  an  “h,”  critical  core-­‐stabilizing  amino  acids  for  IL-­‐10  and  IFNγ  are  marked  with  a  “c”  and  the  viral  conserved  amino  acids  are  marked  with  a  “v.”    The  GenBank  accession  codes  used  were:    European  seabass  (CAK29522.1),  gilthead  seabream  (AGS55345.1),  green  spotted  puffer  (CAD67773.1),  Atlantic  salmon  (ABM46994.1),  Atlantic  cod  (ABV64720.1),  grass  carp  (AEA50953.1),  zebrafish  (AAW78362.1),  African  clawed  frog  (CAE92388.1),  chicken  (CAF18432.1),  rat  (AAA41425.1)  and  human  (AAI04254.1).                                                                                                  

Page 56: Graduate Theses and Dissertations Graduate School 11-4 ...

  43  

   Figure  12.  Clustal  Omega  alignment  between  red  snapper  and  golden  tilefish  IL-­‐10.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified  and  the  positions  of  the  six  alpha  helical  domains  are  underlined  and  labeled.                                                                    

Page 57: Graduate Theses and Dissertations Graduate School 11-4 ...

  44  

   Figure  13.    Multiple  sequence  alignment  of  red  snapper  IL-­‐10  with  the  top  five  most  homologous  species,  according  to  BLASTP.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified  by  Clustal  Omega.      Two  predicted  N-­‐glycosylation  site  are  boxed  and  the  two  IL-­‐10  family  signatures  are  highlighted  in  grey.    Six  conserved  cysteine  residues  are  highlighted  in  grey,  including  two  fish-­‐specific  residues  at  rsC26  and  rsC31.    The  six  alpha  helical  domains  are  underlined  and  lettered  “a-­‐f.”    The  putative  cut  site  for  the  signal  peptide  is  marked  with  an  arrow.    The  GenBank  accession  codes  used  were:  European  seabass  (CAK29522.1),  gilthead  seabream  (AGS55345.1),  green  spotted  puffer  (CAD67773.1),  Atlantic  cod  (ABV64720.1),  and  rainbow  trout  (NP_001232028.1).        

Page 58: Graduate Theses and Dissertations Graduate School 11-4 ...

  45  

   Figure  14.  Rooted  phylogenetic  tree  of  vertebrate  IL-­‐10  proteins,  produced  in  MEGA  6  with  a  NJ  method.    The  p-­‐distance  method  was  used,  bootstrapped  10,000  times,  with  results  reflected  at  the  branch  nodes.    Partial  sequences  were  excluded  and  human  IFNγ  was  used  as  an  outgroup.    The  GenBank  accession  codes  used  were:  European  seabass  (CAK29522.1),  gilthead  seabream  (AGS55345.1),  green  spotted  puffer  (CAD67773.1),  Atlantic  cod  (ABV64720.1),  grass  carp  (AEA50953.1),  zebrafish  (AAW78362.1),  chicken  (CAF18432.1),  rat  (AAA41425.1),  human  (AAI04254.1),  and  human  IFNγ  (AAB59534.1).                                                        

Page 59: Graduate Theses and Dissertations Graduate School 11-4 ...

  46  

   

     

     

   

Page 60: Graduate Theses and Dissertations Graduate School 11-4 ...

  47  

   

Figure  15.    Clustal  W  multiple  sequence  alignment  of  representative  vertebrate  TNFα,  shaded  with  BOXSHADE.    The  family  signature  is  boxed,  the  predicted  Perciform  transmembrane  domain  is  underlined,  the  snapper  TACE  cleavage  site  is  marked  with  an  arrow  above  the  alignment,  and  the  human  TACE  cleavage  site  is  marked  with  an  arrow  below  the  alignment.    The  partially  conserved  cysteine  residue  is  notated  with  a  colon  (:)  and  the  fully  conserved  cysteine  is  marked  with  an  asterisk  (*).    The  GenBank  accession  codes  used  were:  European  seabass  (AAZ20770.1)  ,  red  seabream  (AAP76392.1),  Pacific  bluefin  tuna  (BAG72141.1),  fugu  (NP_001033074.1),  olive  flounder  (BAA94970.1),  rainbow  trout  (CAB92316.1),  grass  carp  (ADY80577.1),  zebrafish  (NP_998024.2),  African  clawed  frog  (BAF95747.1),  chicken  (AEW10516.1),  mouse  (BAA19513.1),  and  human  (NP_00585.2).                                                                                    

Page 61: Graduate Theses and Dissertations Graduate School 11-4 ...

  48  

 Figure  16.  Clustal  Omega  alignment  between  red  snapper  and  golden  tilefish  TNFα.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified.                                                                      

Page 62: Graduate Theses and Dissertations Graduate School 11-4 ...

  49  

   Figure  17.    Multiple  sequence  alignment  of  red  snapper  TNFα  with  the  top  five  most  homologous  species,  according  to  BLASTP.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified  by  Clustal  Omega.    The  predicted  transmembrane  domain  is  underlined,  the  TACE  cleavage  site  is  boxed,  and  the  family  signature  and  two  conserved  cysteine  residues  are  highlighted  in  grey.    The  GenBank  accession  codes  used  were:  barred  knifejaw  (ACM69339.1),  striped  trumpeter  (ACQ99509.1),  orange-­‐spotted  grouper  (AEH59794.1),  large  yellow  croaker  (ABK62876.1),  and  European  seabass  (AAZ20770.1).          

Page 63: Graduate Theses and Dissertations Graduate School 11-4 ...

  50  

   

Figure  18.  Rooted  phylogenetic  tree  of  vertebrate  TNFα  proteins,  produced  in  MEGA  6  with  a  NJ  method.    The  p-­‐distance  method  was  used,  bootstrapped  10,000  times,  with  results  reflected  at  the  branch  nodes.    Partial  sequences  were  excluded  and  human  IFNγ  was  used  as  an  outgroup.    The  GenBank  accession  codes  used  were:  European  seabass  (AAZ20770.1)  ,  red  seabream  (AAP76392.1),  Pacific  bluefin  tuna  (BAG72141.1),  fugu  (NP_001033074.1),  olive  flounder  (BAA94970.1),  rainbow  trout  (CAB92316.1),  grass  carp  (ADY80577.1),  zebrafish  (NP_998024.2),  African  clawed  frog  (BAF95747.1),  mouse  (BAA19513.1),  human  (NP_00585.2),  and  human  IFNγ  (AAB59534.1).            

                                     

Page 64: Graduate Theses and Dissertations Graduate School 11-4 ...

  51  

         

CHAPTER  FOUR:  

CONCLUSIONS  AND  FUTURE  WORK  

 

4.1  Conclusions  

  In  this  study,  IL-­‐1β,  IL-­‐8,  IL-­‐10,  and  TNFα  were,  for  the  first  time,  isolated,  cloned,  sequenced,  

and  characterized  from  golden  tilefish  and  red  snapper.    The  full  amino  acid  sequences  were  obtained  

for  golden  tilefish  IL-­‐1β  and  IL-­‐8,  as  well  as  for  red  snapper  IL-­‐8,  IL-­‐10  and  TNFα.    Partial  sequences  were  

obtained  for  golden  tilefish  IL-­‐10  and  TNFα,  as  well  as  red  snapper  IL-­‐1β.    In  comparing  these  sequences,  

it  was  evident  that  there  was  a  high  level  of  conservation  across  all  four  genes  in  Perciformes,  however,  

there  was  divergence  among  other  teleost  lineages,  particularly  when  Perciformes  were  compared  to  

Salmoniformes  and  Cypriniformes.    Homology  was  low  when  golden  tilefish  and  red  snapper  cytokine  

sequence  were  compared  with  other  vertebrate  groups,  including  cartilaginous  fishes,  amphibians,  

birds,  and  mammals.      

There  was  conservation  among  secondary  structural  elements  across  vertebrates,  evidenced  in  

both  the  β-­‐sheets  of  IL-­‐1β  and  the  alpha  helices  of  IL-­‐10.    There  also  appeared  to  be  similar  cleavage  

sites  present  in  red  snapper  and  golden  tilefish  for  the  mammalian  orthologues  known  to  be  

proteolytically  cleaved  prior  to  generation  of  the  mature  peptide,  as  predicted  in  the  IL-­‐8,  IL-­‐10,  and  

TNFα  teleost  sequences.    However,  there  was  low  conservation  at  known  receptor  binding  sites  

between  mammals  and  teleosts,  particularly  in  the  IL-­‐1β  and  IL-­‐10  sequences  presented  here.    This  

would  suggest  diversity  among  receptor  sites  and  binding  mechanisms  among  vertebrate  groups.    

Coupled  with  the  low  overall  amino  acid  identity  observed  between  mammals  and  the  teleosts,  the  

Page 65: Graduate Theses and Dissertations Graduate School 11-4 ...

  52  

likelihood  that  cross-­‐reactivity  would  occur  between  mammalian  cytokine  antibodies  and  red  snapper  or  

golden  tilefish  is  limited.    Importantly,  this  suggests  that  commercially  available  detection  kits  for  

mammalian  cytokines  probably  would  not  work  well  for  accurately  detecting  and  quantifying  cytokine  

expression  in  Perciformes.  The  results  of  this  study  unequivocally  substantiate  the  need  for  species-­‐

specific  evaluation  of  these  critical  immune  system  regulators.    

 

4.2  Future  work  

  Building  upon  the  information  gathered  here,  antibodies  can  be  designed  for  Perciform-­‐specific  

IL-­‐1β,  IL-­‐8,  IL-­‐10,  and  TNFα.      These  antibodies  can  then  be  multiplexed  into  a  magnetic  bead-­‐based  

assay  for  detection  of  the  above-­‐mentioned  cytokines  in  fish  serum,  plasma,  and  cell  culture  

supernatant.    Given  the  high  level  of  nucleotide  and  amino  acid  similarity  between  red  snapper  and  

golden  tilefish,  as  well  as  among  the  available  Percifromes,  it  is  possible  that  this  kit  will  be  cross-­‐

reactive  with  many  families  in  this  large  teleost  order.    

  A  suite  of  in  vivo  and  in  vitro  gene  expression  studies  would  aid  in  our  understanding  of  cytokine  

expression  by  evaluating  the  timing  and  magnitude  of  cytokine  release  upon  exposure  to  contaminants,  

such  as  DWH  crude  and  Corexit  9500.    Correlations  could  be  investigated  between  gene  expression  in  

field  collected  golden  tilefish  and  red  snapper  samples  caught  in  the  vicinity  of  the  DWH  explosion  from  

2013  and  2014,  and  the  resulting  circulating  serum  cytokine  concentration.    Additional  study  goals  could  

include  an  analysis  of  the  timing  and  magnitude  of  the  expression  of  different  markers  along  the  PAH  

metabolic  pathway  leading  to  the  induction  of  an  immune  response.  

  The  thoroughly  validated  kit  would  provide  a  novel  tool  for  the  rapid  assessment  of  immune  

system  status  in  Perciformes.    Not  only  would  this  be  useful  to  aid  in  fishery  management  decisions  

during  future  contamination  events,  but  also  for  daily  use  in  the  burgeoning  aquaculture  field.    This  

work  will  form  the  basis  of  the  author’s  PhD  dissertation,  beginning  January  2015.    

Page 66: Graduate Theses and Dissertations Graduate School 11-4 ...

  53  

         

REFERENCES    

1.   Barron,  M.G.,  Ecological  impacts  of  the  deepwater  horizon  oil  spill:  implications  for  

immunotoxicity.  Toxicologic  Pathology,  2012.  40(2):  p.  315-­‐20.  

2.   Spier,  C.,  et  al.,  Distribution  of  hydrocarbons  released  during  the  2010  MC252  oil  spill  in  deep  

offshore  waters.  Environmental  Pollution,  2013.  173(0):  p.  224-­‐230.  

3.   Rotkin-­‐Ellman,  M.,  K.K.  Wong,  and  G.M.  Solomon,  Seafood  contamination  after  the  BP  Gulf  oil  

spill  and  risks  to  vulnerable  populations:  a  crtitique  of  the  FDA  risk  assessment.  Environmental  

Health  Perspectives,  2012.  120(2):  p.  157-­‐161.  

4.   Allan,  S.E.,  B.W.  Smith,  and  K.A.  Anderson,  Impact  of  the  Deepwater  Horizon  Oil  Spill  on  

Bioavailable  Polycyclic  Aromatic  Hydrocarbons  in  Gulf  of  Mexico  Coastal  Waters.  Environmental  

Science  &  Technology,  2012.  46(4):  p.  2033-­‐2039.  

5.   Ramachandran,  S.D.,  et  al.,  Oil  dispersant  increases  PAH  uptake  by  fish  exposed  to  crude  oil.  

Ecotoxicology  and  Environmental  Safety,  2004.  59(3):  p.  300-­‐308.  

6.   Lu,  M.,  et  al.,  Molecular  characterization  of  the  aryl  hydrocarbon  receptor  (AhR)  pathway  in  

goldfish  (Carassius  auratus)  exposure  to  TCDD:  The  mRNA  and  protein  levels.  Fish  &  Shellfish  

Immunology,  2013.  35(2):  p.  469-­‐475.  

7.   Reynaud,  S.  and  P.  Deschaux,  The  effects  of  polycyclic  aromatic  hydrocarbons  on  the  immune  

system  of  fish:  A  review.  Aquatic  Toxicology,  2006.  77(2):  p.  229-­‐238.  

8.   Sarasquete,  C.  and  H.  Segner,  Cytochrome  P4501A  (CYP1A)  in  teleostean  fishes.  A  review  of  

immunohistochemical  studies.  Science  of  The  Total  Environment,  2000.  247(2–3):  p.  313-­‐332.  

Page 67: Graduate Theses and Dissertations Graduate School 11-4 ...

  54  

9.   Hahn,  M.E.,  Aryl  hydrocarbon  receptors:  diversity  and  evolution.  Chemico-­‐Biological  Interactions,  

2002.  141(1–2):  p.  131-­‐160.  

10.   Regoli,  F.  and  M.E.  Giuliani,  Oxidative  pathways  of  chemical  toxicity  and  oxidative  stress  

biomarkers  in  marine  organisms.  Marine  Environmental  Research,  2014.  93(0):  p.  106-­‐117.  

11.   Farmen,  E.,  et  al.,  Oxidative  stress  responses  in  rainbow  trout  (Oncorhynchus  mykiss)  

hepatocytes  exposed  to  pro-­‐oxidants  and  a  complex  environmental  sample.  Comparative  

Biochemistry  and  Physiology  Part  C:  Toxicology  &  Pharmacology,  2010.  151(4):  p.  431-­‐438.  

12.   Harvey,  H.R.,  et  al.,  Polycyclic  aromatic  and  aliphatic  hydrocarbons  in  Chukchi  Sea  biota  and  

sediments  and  their  toxicological  response  in  the  Arctic  cod,  Boreogadus  saida.  Deep  Sea  

Research  Part  II:  Topical  Studies  in  Oceanography,  2014.  102(0):  p.  32-­‐55.  

13.   Tairova,  Z.M.,  et  al.,  PAH  biomarkers  in  common  eelpout  (Zoarces  viviparus)  from  Danish  waters.  

Marine  Environmental  Research,  2012.  75(0):  p.  45-­‐53.  

14.   Wang,  L.,  et  al.,  Expression  of  CYP1C1  and  CYP1A  in  Fundulus  heteroclitus  during  PAH-­‐induced  

carcinogenesis.  Aquatic  Toxicology,  2010.  99(4):  p.  439-­‐447.  

15.   Billiard,  S.M.,  et  al.,  Binding  of  polycyclic  aromatic  hydrocarbons  (PAHs)  to  teleost  aryl  

hydrocarbon  receptors  (AHRs).  Comparative  Biochemistry  and  Physiology  Part  B:  Biochemistry  

and  Molecular  Biology,  2002.  133(1):  p.  55-­‐68.  

16.   Holth,  T.F.,  et  al.,  Effects  of  water  accommodated  fractions  of  crude  oils  and  diesel  on  a  suite  of  

biomarkers  in  Atlantic  cod  (Gadus  morhua).  Aquatic  Toxicology,  2014.  154(0):  p.  240-­‐252.  

17.   Murawski,  S.A.,  et  al.,  Prevalence  of  external  skin  lesions  and  polycyclic  aromatic  hydrocarbon  

concentrations  in  the  Gulf  of  Mexico  fishes,  post-­‐Deepwater  Horizon.  Transactions  of  the  

American  Fisheries  Society,  2014.  143(4):  p.  1084-­‐1097.  

18.   Lang,  T.,  et  al.,  Liver  histopathology  in  Baltic  flounder  (Platichthys  flesus)  as  indicator  of  

biological  effects  of  contaminants.  Marine  Pollution  Bulletin,  2006.  53(8–9):  p.  488-­‐496.  

Page 68: Graduate Theses and Dissertations Graduate School 11-4 ...

  55  

19.   Kim,  H.N.,  et  al.,  Acute  toxic  responses  of  the  rockfish  (Sebastes  schlegeli)  to  Iranian  heavy  crude  

oil:  Feeding  disrupts  the  biotransformation  and  innate  immune  systems.  Fish  &  Shellfish  

Immunology,  2013.  35(2):  p.  357-­‐365.  

20.   Carlson,  E.A.,  Y.  Li,  and  J.T.  Zelikoff,  Exposure  of  Japanese  medaka  (Oryzias  latipes)  to  

benzo[a]pyrene  suppresses  immune  function  and  host  resistance  against  bacterial  challenge.  

Aquatic  Toxicology,  2002.  56(4):  p.  289-­‐301.  

21.   Song,  J.Y.,  et  al.,  Immune  gene  expression  levels  correlate  with  the  phenotype  of  Japanese  

flounder  exposed  to  heavy  oil.  Interdisciplinary  Studies  on  Environmental  Chemistry,  2008:  p.  

111-­‐222.  

22.   Segner,  H.,  et  al.,  Fish  immunotoxicology:  Research  at  the  crossroads  of  immunology,  ecology,  

and  toxicology.  Interdisciplinary  Studies  on  Environmental  Chemistry,  2012.  6:  p.  1-­‐12.  

23.   Leonard,  W.,  Type  I  cytokines  and  their  interferons,  and  their  receptors,  in  Fundamental  

Immunology,  W.E.  Paul,  Editor.  2012,  Lippincott  Williams  &  Wilkins:  New  York.  p.  601-­‐638.  

24.   Savan,  R.  and  M.  Sakai,  Genomics  of  fish  cytokines.  Comparative  Biochemistry  and  Physiology  

Part  D:  Genomics  and  Proteomics,  2006.  1(1):  p.  89-­‐101.  

25.   Wang,  T.  and  C.J.  Secombes,  The  cytokine  networks  of  adaptive  immunity  in  fish.  Fish  &  Shellfish  

Immunology,  2013.  35(6):  p.  1703-­‐1718.  

26.   Feghali,  C.A.  and  T.M.  Wright,  Cytokines  in  acute  and  chronic  Inflammation  Frontiers  in  

Bioscience,  2007.  2:  p.  d12-­‐26.  

27.   Moser,  M.  and  O.  Leo,  Key  concepts  in  immunology.  Vaccine,  2010.  28,  Supplement  3(0):  p.  C2-­‐

C13.  

28.   Radke,  L.,  et  al.,  Induced  cytokine  response  of  human  PMBC-­‐cultures:  Correlation  of  gene  

expression  and  secretion  profiling  and  the  effect  of  cryopreservation.  Cellular  Immunology,  2012.  

272(2):  p.  144-­‐153.  

Page 69: Graduate Theses and Dissertations Graduate School 11-4 ...

  56  

29.   Holloway,  A.F.,  S.  Rao,  and  M.F.  Shannon,  Regulation  of  cytokine  gene  transcription  in  the  

immune  system.  Molecular  Immunology,  2002.  38(8):  p.  567-­‐580.  

30.   Hafler,  D.A.,  Cytokines  and  interventional  immunology.  Nat  Rev  Immunol,  2007.  7(6):  p.  423-­‐423.  

31.   Zhu,  L.-­‐y.,  et  al.,  Advances  in  research  of  fish  immune-­‐relevant  genes:  A  comparative  overview  of  

innate  and  adaptive  immunity  in  teleosts.  Developmental  &  Comparative  Immunology,  2013.  

39(1–2):  p.  39-­‐62.  

32.   Sarropoulou,  E.  and  J.M.O.  Fernandes,  Comparative  genomics  in  teleost  species:  Knowledge  

transfer  by  linking  the  genomes  of  model  and  non-­‐model  fish  species.  Comparative  Biochemistry  

and  Physiology  Part  D:  Genomics  and  Proteomics,  2011.  6(1):  p.  92-­‐102.  

33.   Zapata,  A.,  et  al.,  Ontogeny  of  the  immune  system  of  fish.  Fish  &  Shellfish  Immunology,  2006.  

20(2):  p.  126-­‐136.  

34.   Oshiumi,  H.,  et  al.,  Pan-­‐vertebrate  toll-­‐like  receptors  during  evolution.  Current  Genomics,  2008.  

9(7):  p.  488-­‐493.  

35.   Laing,  K.J.  and  J.D.  Hansen,  Fish  T  cells:  Recent  advances  through  genomics.  Developmental  &  

Comparative  Immunology,  2011.  35(12):  p.  1282-­‐1295.  

36.   Kono,  T.,  et  al.,  Establishment  of  a  multiplex  RT-­‐PCR  assay  for  the  rapid  detection  of  fish  

cytokines.  Veterinary  Immunology  and  Immunopathology,  2013.  151(1–2):  p.  90-­‐101.  

37.   Reyes-­‐Cerpa,  S.,  et  al.,  Fish  cytokines  and  immune  response,  in  New  Advances  and  Contributions  

in  Fish  Biology,  H.  Turker,  Editor.  2012,  InTech:  Online.  

38.   Lepen  Pleić,  I.,  et  al.,  Characterization  of  three  pro-­‐inflammatory  cytokines,  TNFα1,  TNFα2  and  

IL-­‐1β,  in  cage-­‐reared  Atlantic  bluefin  tuna  Thunnus  thynnus.  Fish  &  Shellfish  Immunology,  2014.  

36(1):  p.  98-­‐112.  

39.   Plouffe,  D.A.,  et  al.,  Comparison  of  select  innate  immune  mechanisms  of  fish  and  mammals.  

Xenotransplantation,  2005.  12(4):  p.  266-­‐277.  

Page 70: Graduate Theses and Dissertations Graduate School 11-4 ...

  57  

40.   Lieschke,  G.J.  and  N.S.  Trede,  Fish  immunology.  Current  Biology,  2009.  19(16):  p.  R678-­‐R682.  

41.   Sabat,  R.,  et  al.,  Biology  of  interleukin-­‐10.  Cytokine  &  Growth  Factor  Reviews,  2010.  21(5):  p.  

331-­‐344.  

42.   Abdelkhalek,  N.K.,  et  al.,  Molecular  evidence  for  the  existence  of  two  distinct  IL-­‐8  lineages  of  

teleost  CXC-­‐chemokines.  Fish  &  Shellfish  Immunology,  2009.  27(6):  p.  763-­‐767.  

43.   Román,  L.,  et  al.,  Cytokine  expression  in  head-­‐kidney  leucocytes  of  European  sea  bass  

(Dicentrarchus  labrax  L.)  after  incubation  with  the  probiotic  Vagococcus  fluvialis  L-­‐21.  Fish  &  

Shellfish  Immunology,  2013.  35(4):  p.  1329-­‐1332.  

44.   Covello,  J.M.,  et  al.,  Cloning  and  expression  analysis  of  three  striped  trumpeter  (Latris  lineata)  

pro-­‐inflammatory  cytokines,  TNF-­‐α,  IL-­‐1β  and  IL-­‐8,  in  response  to  infection  by  the  ectoparasitic,  

Chondracanthus  goldsmidi.  Fish  &  Shellfish  Immunology,  2009.  26(5):  p.  773-­‐786.  

45.   Whyte,  S.K.,  The  innate  immune  response  of  finfish  –  A  review  of  current  knowledge.  Fish  &  

Shellfish  Immunology,  2007.  23(6):  p.  1127-­‐1151.  

46.   Batista,  C.R.,  et  al.,  Impairment  of  the  immune  system  in  GH-­‐overexpressing  transgenic  zebrafish  

(Danio  rerio).  Fish  &  Shellfish  Immunology,  2014.  36(2):  p.  519-­‐524.  

47.   Husain,  M.,  et  al.,  Cloning  of  the  IL-­‐1β3  gene  and  IL-­‐1β4  pseudogene  in  salmonids  uncovers  a  

second  type  of  IL-­‐1β  gene  in  teleost  fish.  Developmental  &  Comparative  Immunology,  2012.  

38(3):  p.  431-­‐446.  

48.   Seppola,  M.,  et  al.,  Characterisation  and  expression  analysis  of  the  interleukin  genes,  IL-­‐1β,  IL-­‐8  

and  IL-­‐10,  in  Atlantic  cod  (Gadus  morhua  L.).  Molecular  Immunology,  2008.  45(4):  p.  887-­‐897.  

49.   Corripio-­‐Miyar,  Y.,  et  al.,  Cloning  and  expression  analysis  of  two  pro-­‐inflammatory  cytokines,  IL-­‐

1β  and  IL-­‐8,  in  haddock  (Melanogrammus  aeglefinus).  Molecular  Immunology,  2007.  44(6):  p.  

1361-­‐1373.  

Page 71: Graduate Theses and Dissertations Graduate School 11-4 ...

  58  

50.   Scapigliati,  G.,  et  al.,  Phylogeny  of  cytokines:  molecular  cloning  and  expression  analysis  of  sea  

bass  Dicentrarchus  labrax  interleukin-­‐1β.  Fish  &  Shellfish  Immunology,  2001.  11(8):  p.  711-­‐726.  

51.   Lu,  D.-­‐Q.,  et  al.,  Interleukin-­‐1β  gene  in  orange-­‐spotted  grouper,  Epinephelus  coioides:  Molecular  

cloning,  expression,  biological  activities  and  signal  transduction.  Molecular  Immunology,  2008.  

45(4):  p.  857-­‐867.  

52.   Benedetti,  S.,  et  al.,  Evolution  of  cytokine  responses:  IL-­‐1β  directly  affects  intracellular  Ca2+  

concentration  of  teleost  fish  leukocytes  through  a  receptor-­‐mediated  mechanism.  Cytokine,  

2006.  34(1–2):  p.  9-­‐16.  

53.   Ogryzko,  N.V.,  S.A.  Renshaw,  and  H.L.  Wilson,  The  IL-­‐1  family  in  fish:  Swimming  through  the  

muddy  waters  of  inflammasome  evolution.  Developmental  &  Comparative  Immunology,  2014.  

46(1):  p.  53-­‐62.  

54.   Secombes,  C.J.,  T.  Wang,  and  S.  Bird,  The  interleukins  of  fish.  Developmental  &  Comparative  

Immunology,  2011.  35(12):  p.  1336-­‐1345.  

55.   Alejo,  A.  and  C.  Tafalla,  Chemokines  in  teleost  fish  species.  Developmental  &  Comparative  

Immunology,  2011.  35(12):  p.  1215-­‐1222.  

56.   Chen,  J.,  et  al.,  Phylogenetic  analysis  of  vertebrate  CXC  chemokines  reveals  novel  lineage  specific  

groups  in  teleost  fish.  Developmental  &  Comparative  Immunology,  2013.  41(2):  p.  137-­‐152.  

57.   Kurata,  O.,  et  al.,  N-­‐terminal  region  is  responsible  for  chemotaxis-­‐inducing  activity  of  flounder  IL-­‐

8.  Fish  &  Shellfish  Immunology,  2014.  38(2):  p.  361-­‐366.  

58.   Hur,  D.,  J.-­‐K.  Jeon,  and  S.  Hong,  Analysis  of  immune  gene  expression  modulated  by  

benzo[a]pyrene  in  head  kidney  of  olive  flounder  (Paralichthys  olivaceus).  Comparative  

Biochemistry  and  Physiology  Part  B:  Biochemistry  and  Molecular  Biology,  2013.  165(1):  p.  49-­‐57.  

59.   Feldman,  M.  and  J.  Saklatvala,  Proinflammatory  cytokines,  in  Cytokine  Reference,  J.J.  Oppenheim  

and  M.  Feldman,  Editors.  2000,  Academic  Press:  New  York.  p.  291-­‐305.  

Page 72: Graduate Theses and Dissertations Graduate School 11-4 ...

  59  

60.   Zhang,  A.,  et  al.,  Functional  characterization  of  TNF-­‐α  in  grass  carp  head  kidney  leukocytes:  

Induction  and  involvement  in  the  regulation  of  NF-­‐κB  signaling.  Fish  &  Shellfish  Immunology,  

2012.  33(5):  p.  1123-­‐1132.  

61.   Morrison,  R.N.,  et  al.,  Molecular  cloning  and  expression  analysis  of  tumour  necrosis  factor-­‐α  in  

amoebic  gill  disease  (AGD)-­‐affected  Atlantic  salmon  (Salmo  salar  L.).  Fish  &  Shellfish  

Immunology,  2007.  23(5):  p.  1015-­‐1031.  

62.   Kadowaki,  T.,  et  al.,  Two  types  of  tumor  necrosis  factor-­‐α  in  bluefin  tuna  (Thunnus  orientalis)  

genes:  Molecular  cloning  and  expression  profile  in  response  to  several  immunological  stimulants.  

Fish  &  Shellfish  Immunology,  2009.  27(5):  p.  585-­‐594.  

63.   O'Garra,  A.  and  P.  Vieira,  TH1  cells  control  themselves  by  producing  interleukin-­‐10.  Nat  Rev  

Immunol,  2007.  7(6):  p.  425-­‐428.  

64.   Grayfer,  L.,  et  al.,  Characterization  and  functional  analysis  of  goldfish  (Carassius  auratus  L.)  

interleukin-­‐10.  Molecular  Immunology,  2011.  48(4):  p.  563-­‐571.  

65.   Bogdan,  C.,  Y.  Vodovotz,  and  C.  Nathan,  Macrophage  deactivation  by  interleukin  10.  The  Journal  

of  Experimental  Medicine,  1991.  174(6):  p.  1549-­‐1555.  

66.   Wei,  H.,  et  al.,  Functional  expression  and  characterization  of  grass  carp  IL-­‐10:  An  essential  

mediator  of  TGF-­‐β1  immune  regulation  in  peripheral  blood  lymphocytes.  Molecular  Immunology,  

2013.  53(4):  p.  313-­‐320.  

67.   Wei,  H.,  et  al.,  Identification  of  grass  carp  IL-­‐10  receptor  subunits:  Functional  evidence  for  IL-­‐10  

signaling  in  teleost  immunity.  Developmental  &  Comparative  Immunology,  2014.  45(2):  p.  259-­‐

268.  

68.   Pinto,  R.D.,  et  al.,  Molecular  characterization,  3D  modelling  and  expression  analysis  of  sea  bass  

(Dicentrarchus  labrax  L.)  interleukin-­‐10.  Molecular  Immunology,  2007.  44(8):  p.  2056-­‐2065.  

Page 73: Graduate Theses and Dissertations Graduate School 11-4 ...

  60  

69.   Zhang,  D.C.,  et  al.,  Cloning,  characterization  and  expression  analysis  of  interleukin-­‐10  from  the  

zebrafish  (Danio  rerio).  Journal  of  Biochemistry  and  Molecular  Biology,  2005.  38(5):  p.  571-­‐576.  

70.   Buonocore,  F.,  et  al.,  Interleukin-­‐10  expression  by  real-­‐time  PCR  and  homology  modelling  

analysis  in  the  European  sea  bass  (Dicentrarchus  Labrax  L.).  Aquaculture,  2007.  270(1–4):  p.  512-­‐

522.  

71.   Harun,  N.O.,  et  al.,  Sequencing  of  a  second  interleukin-­‐10  gene  in  rainbow  trout  Oncorhynchus  

mykiss  and  comparative  investigation  of  the  expression  and  modulation  of  the  paralogues  

in  vitro  and  in  vivo.  Fish  &  Shellfish  Immunology,  2011.  31(1):  p.  107-­‐117.  

72.   Solís,  D.,  et  al.,  Evaluating  the  impact  of  individual  fishing  quotas  (IFQs)  on  the  technical  

efficiency  and  composition  of  the  US  Gulf  of  Mexico  red  snapper  commercial  fishing  fleet.  Food  

Policy,  2014.  46(0):  p.  74-­‐83.  

73.   Forrest,  R.E.,  et  al.,  Modelling  the  effects  of  density-­‐dependent  mortality  in  juvenile  red  snapper  

caught  as  bycatch  in  Gulf  of  Mexico  shrimp  fisheries:  Implications  for  management.  Fisheries  

Research,  2013.  146(0):  p.  102-­‐120.  

74.   Doerpinghaus,  J.,  et  al.,  An  assessment  of  sector  separation  on  the  Gulf  of  Mexico  recreational  

red  snapper  fishery.  Marine  Policy,  2014.  50,  Part  A(0):  p.  309-­‐317.  

75.   Kitts,  A.,  P.  Pinto  da  Silva,  and  B.  Rountree,  The  evolution  of  collaborative  management  in  the  

Northeast  USA  tilefish  fishery.  Marine  Policy,  2007.  31(2):  p.  192-­‐200.  

76.   Steimle,  F.W.,  V.S.  Zdanowicz,  and  D.F.  Gadbois,  Metals  and  organic  contaminants  in  northwest  

Atlantic  deep-­‐sea  tilefish  tissues.  Marine  Pollution  Bulletin,  1990.  21(11):  p.  530-­‐535.  

77.   Pereiro,  P.,  et  al.,  The  first  characterization  of  two  type  1  interferons  in  turbot  (Scophthalmus  

maximus)  reveals  their  differential  role,  expression  pattern  and  gene  induction.  Developmental  

&  Comparative  Immunology,  2014.  45:  p.  233-­‐244.  

Page 74: Graduate Theses and Dissertations Graduate School 11-4 ...

  61  

78.   Bird,  S.,  et  al.,  Evolution  of  interleukin-­‐1β.  Cytokine  &  Growth  Factor  Reviews,  2002.  13(6):  p.  

483-­‐502.  

79.   López-­‐Castejón,  G.,  et  al.,  Characterization  of  ATP-­‐gated  P2X7  receptors  in  fish  provides  new  

insights  into  the  mechanism  of  release  of  the  leaderless  cytokine  interleukin-­‐1β.  Molecular  

Immunology,  2007.  44(6):  p.  1286-­‐1299.  

80.   Lopez-­‐Castejon,  G.  and  D.  Brough,  Understanding  the  mechanism  of  IL-­‐1b  secretion.  Cytokine  &  

Growth  Factor  Reviews,  2011.  22(4):  p.  189-­‐195.  

81.   Labriola-­‐Tompkins,  E.,  et  al.,  Identification  of  the  discontinuous  binding  site  in  human  interleukin  

1  beta  for  the  type  I  interleukin  1  receptor.  Proceedings  of  the  National  Academy  of  Sciences,  

1991.  88(24):  p.  11182-­‐11186.  

82.   Vigers,  G.P.A.,  et  al.,  Crystal  structure  of  the  type-­‐I  interleukin-­‐1  receptor  complexed  with  

interleukin-­‐1[beta].  Nature,  1997.  386(6621):  p.  190-­‐194.  

83.   Wedlock,  D.N.,  et  al.,  Molecular  cloning  and  physiological  effects  of  brushtail  possum  interleukin-­‐

1β.  Veterinary  Immunology  and  Immunopathology,  1999.  67(4):  p.  359-­‐372.  

84.   Cai,  Z.,  et  al.,  Functional  Characterization  of  the  ELR  Motif  in  Piscine  ELR+CXC-­‐Like  Chemokine.  

Marine  Biotechnology,  2009.  11(4):  p.  505-­‐512.  

85.   Zdanov,  A.,  et  al.,  Crystal  structure  of  interleukin-­‐10  reveals  the  functional  dimer  with  an  

unexpected  topological  similarity  to  interferon  γ.  Structure,  1995.  3(6):  p.  591-­‐601.  

86.   Fickenscher,  H.,  et  al.,  The  interleukin-­‐10  family  of  cytokines.  Trends  in  Immunology,  2002.  23(2):  

p.  89-­‐96.  

87.   Walter,  M.R.  and  T.L.  Nagabhushan,  Crystal  Structure  of  Interleukin  10  Reveals  an  Interferon  

.gamma.-­‐like  Fold.  Biochemistry,  1995.  34(38):  p.  12118-­‐12125.  

88.   Ding,  Y.,  et  al.,  A  Single  Amino  Acid  Determines  the  Immunostimulatory  Activity  of  Interleukin  10.  

The  Journal  of  Experimental  Medicine,  2000.  191(2):  p.  213-­‐224.  

Page 75: Graduate Theses and Dissertations Graduate School 11-4 ...

  62  

89.   Josephson,  K.,  N.J.  Logsdon,  and  M.  Walter,  Crystal  structure  of  the  IL-­‐10/IL-­‐10R1  complex  

reveals  a  shared  receptor  binding  site.  Immunity,  2001.  14:  p.  35-­‐46.  

90.   Zou,  J.,  et  al.,  Differential  expression  of  two  tumor  necrosis  factor  genes  in  rainbow  trout,  

Oncorhynchus  mykiss.  Developmental  &  Comparative  Immunology,  2002.  26(2):  p.  161-­‐172.  

91.   Haugland,  Ø.,  et  al.,  Differential  expression  profiles  and  gene  structure  of  two  tumor  necrosis  

factor-­‐α  variants  in  Atlantic  salmon  (Salmo  salar  L.).  Molecular  Immunology,  2007.  44(7):  p.  

1652-­‐1663.  

                                                                   

Page 76: Graduate Theses and Dissertations Graduate School 11-4 ...

  63  

         

APPENDIX  A:    

MAP  OF  2013  TISSUE  SAMPLING  STATIONS        

 Figure  A1:  2013  longline  stations  aboard  the  R/V  Weatherbird  II  yielding  tissue  samples.  Yellow-­‐circled  stations  indicate  the  locations  from  which  golden  tilefish  spleen  samples  were  used  in  this  study.  Red-­‐circled  stations  indicate  the  locations  from  which  red  snapper  spleen  samples  were  used  in  this  study.                                            

Page 77: Graduate Theses and Dissertations Graduate School 11-4 ...

  64  

         

APPENDIX  B:    

2013  TISSUE  SAMPLING  LOCATION  DATA      

Table  B1:  2013  longline  tissue  sampling  location  data.  Latitude  and  longitude  data  are  provided  in  degrees,  minutes,  decimal  minutes  format  from  shipboard  GPS  data  at  the  start  and  end  of  setting  out  gear.    The  mean  depth  is  the  average  of  the  start  and  end  shipboard  depth  data  obtained  while  setting  out  the  gear.        

Station  ID Sampling  dateLatitude                      

at  Set  StartLongitude            at  Set  Start

Latitude                      at  Set  End

Longitude                  at  Set  End

Mean  depth                      (m)

PCB03-­‐100 8/15/2013 29o  41.360 86o  24.080 29o  44.101 86o  19.080 101

SL7-­‐150 8/15/2013 29o  29.226 86o  40.371 29o  26.437 86o  44.838 329

SL14-­‐60 8/16/2013 29o  27.619 87o  26.448 29o  30.645 87o  21.768 235

SL14-­‐100 8/16/2013 29o  24.292 87o  29.536 29o  19.455 87o  29.499 362

SL8-­‐40 8/19/2013 29o  53.263 87o  17.663 29o  49.051 87o  20.614 54

SL8-­‐100 8/19/2013 29o  44.666 87o  11.552 29o  42.253 87o  07.648 211

SL9-­‐80 8/20/2013 29o  17.746 88o  00.589 29o  16.038 87o  56.303 191

SL9-­‐150 8/20/2013 29o  14.984 87o  59.256 29o  13.356 87o  54.006 328

SL10-­‐40 8/21/2013 29o  12.305 88o  52.125 29o  08.116 88o  53.573 59

HE-­‐365 8/25/2013 28o  22.609 90o  30.839 28o  18.876 90o  27.504 49

SL16-­‐150 8/26/2013 28o  37.780 90o  00.397 28o  34.138 89o  56.721 204

SL12-­‐100 8/26/2013 28o  30.962 89o  36.251 28o  33.833 89o  33.017 381

SL11-­‐150 8/27/2013 29o  02.685 88o  36.926 29o  03.926 88o  32.267 328

SL5-­‐200 8/28/2013 28o  28.897 86o  07.870 28o  29.553 86o  02.774 371

SL5-­‐100 8/29/2013 28o  33.130 85o  22.370 28o  32.686 85o  27.755 179

SL4-­‐40 8/29/2013 28o  05.265 84o  26.300 28o  07.896 84o  22.086 61                            

Page 78: Graduate Theses and Dissertations Graduate School 11-4 ...

  65  

         

APPENDIX  C:    

GOLDEN  TILEFISH  AND  RED  SNAPPER  BIOMETRICS  DATA    

 Table  C1:  Golden  tilefish  and  red  snapper  biometrics  data,  specific  to  fish  whose  spleens  were  used  in  this  study.      

Species cDNA  Sample  ID Station  ID Sex Standard  Length  (cm) Fork  Length  (cm) Weight  (kg)

F SL11-­‐150 F 58 66 3.672

G SL11-­‐150 M 58 68 4.054

T8 SL14-­‐100 U 57 63 2.630

T9 SL14-­‐100 U 56 63 2.974

C SL4-­‐40 M 57 65 4.508

E SL4-­‐40 F 62 70 5.274

Golden  tilefish

Red  snapper  

                                                     

Page 79: Graduate Theses and Dissertations Graduate School 11-4 ...

  66  

         

APPENDIX  D:  

EXTENDED  PCR  PARAMETERS      

Table  D1.  Extended  PCR  parameters  for  primers  used.  

 

Temp%(C)% Time%(min) Temp%(C)% Time%(min) Temp%(C)% Time%(min) Temp%(C)% Time%(min) Temp%(C)% Time%(min)

BA 94 5 94 0.75 55 0.75 72 0.75 72 5 35

rs_IL:1b 94 5 94 0.75 48 0.75 72 0.75 72 10 35

rs_IL:8 94 5 94 0.5 50 0.5 72 1 72 10 35

rs_IL:10 94 5 94 0.75 55 0.75 72 0.75 72 10 35

rs_TNFa 94 5 94 0.5 50 0.5 72 1 72 10 35

gtIL:1b 94 5 94 0.75 48 0.75 72 0.75 72 10 35

gtIL:8 94 5 94 0.5 50 0.5 72 1 72 10 35

gtIL_10 94 5 94 0.75 54 0.75 72 0.75 72 10 35

gt_TNFa 94 5 94 0.5 48 0.5 72 1 72 10 35

T7/SP6 94 5 94 0.5 48 0.5 72 0.75 72 5 35

rsIL:8_GSP1 94 5 94 0.5 68 0.5 72 1 72 10 35

rsIL:8_GSP2 94 5 94 0.5 66 0.5 72 1 72 10 35

rsIL:10_GSP1 94 5 94 0.5 68 0.5 72 1 72 10 35

rsIL:10_GSP2 94 5 94 0.5 68 0.5 72 1 72 10 35

rsTNFa_GSP1 94 5 94 0.5 70 0.5 72 1 72 10 35

rsTNFa_GSP2 94 5 94 0.5 70 0.5 72 1 72 10 35

gtIL:1b_GSP1 94 5 94 0.5 64 0.5 72 1 72 10 35

gtIL:1b_GSP2 94 5 94 0.5 64 0.5 72 1 72 10 35

gtIL:8_GSP1 94 5 94 0.5 70 0.5 72 1 72 10 35

gtIL:10_GSP2 94 5 94 0.5 62 0.5 72 1 72 10 35

gtTNFa_GSP1 94 5 94 0.5 60 0.5 72 1 72 10 35

rsf_IL:8_F 94 5 94 0.5 54 0.5 72 1 72 10 35

rsf_IL:8_R 94 5 94 0.5 54 0.5 72 1 72 10 35

rsf_IL:10_F 94 5 94 0.5 54 0.5 72 1 72 10 35

rsf_IL:10_R 94 5 94 0.5 54 0.5 72 1 72 10 35

rsf_TNFa_F 94 5 94 0.5 56 0.5 72 1 72 10 35

rsf_TNFa_R 94 5 94 0.5 56 0.5 72 1 72 10 35

gtf_IL:1b_F 94 5 94 0.5 54 0.5 72 1 72 10 35

gtf_IL:1b_R 94 5 94 0.5 54 0.5 72 1 72 10 35

gtf_IL:8_F 94 5 94 0.5 54 0.5 72 1 72 10 35

gtf_IL:8_R 94 5 94 0.5 54 0.5 72 1 72 10 35

#"of"CyclesReaction

Initital"Denaturation" Denaturation Annealing Extension Final"Extension"

Page 80: Graduate Theses and Dissertations Graduate School 11-4 ...

  67  

         

APPENDIX  E:      

EXPERIMENTALLY  OBTAINED  CYTOKINE  NUCLEOTIDE  SEQUENCES    

 

   Figure  E1.  Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  golden  tilefish  IL-­‐1β,  derived  from  the  assembled  5’-­‐  and  3’-­‐RACE  products.    Black  text  reflects  the  5’  and  3’  UTR,  the  red  text  indicates  the  start  (ATG)  and  stop  (TAA)  codons,  the  continuous  green  and  blue  text  indicate  the  coding  region  for  the  golden  tilefish  IL-­‐1β  protein,  and  the  blue  text  represents  the  overlap  between  the  5’  and  3’-­‐RACE  products.        

Page 81: Graduate Theses and Dissertations Graduate School 11-4 ...

  68  

   Figure  E2.  Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  golden  tilefish  IL-­‐1β,  obtained  via  cloning  from  two  different  fish.    Black  text  represents  the  5’  and  3’  UTR,  red  text  indicates  the  start  (ATG)  and  stop  (TAA)  codon,  and  green  text  designates  the  coding  region.      

   Figure  E3.  Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐1β,  derived  from  the  partial  cloned  fragment.    Black  text  designates  the  5’  UTR,  red  text  reflects  the  start  codon  (ATG)  and  green  text  indicates  the  coding  region  for  the  red  snapper  IL-­‐1β  protein.        

   Figure  E4.  Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  golden  tilefish  IL-­‐8,  derived  from  the  3’  RACE  product.    Black  text  reflects  the  5’  and  3’  UTR,  the  red  text  indicates  the  start  (ATG)  and  stop  (TGA)  codons,  the  continuous  green  text  indicates  the  coding  region  for  the  golden  tilefish  IL-­‐8  protein.          

Page 82: Graduate Theses and Dissertations Graduate School 11-4 ...

  69  

   Figure  E5.  Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  golden  tilefish  IL-­‐8,  obtained  via  cloning  from  two  different  fish.    Black  text  represents  the  3’  UTR,  red  text  indicates  the  start  (ATG)  and  stop  (TGA)  codon,  and  green  text  designates  the  coding  region.    

   

 Figure  E6.  Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐8,  derived  from  the  assembled  5’-­‐  and  3’-­‐RACE  products.    Black  text  reflects  the  5’  and  3’  UTR,  the  red  text  indicates  the  start  (ATG)  and  stop  (TGA)  codons,  the  green  text  marks  the  coding  region  for  the  red  snapper  IL-­‐8  protein,  and  the  blue  text  represents  the  overlap  between  the  5’-­‐  and  3’-­‐RACE  products.      

   Figure  E7.  Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  red  snapper  IL-­‐8,  obtained  via  cloning  from  two  different  fish.    Black  text  represents  the  5’  and  3’  UTR,  red  text  indicates  the  start  (ATG)  and  stop  (TGA)  codon,  and  green  text  designates  the  coding  region.              

Page 83: Graduate Theses and Dissertations Graduate School 11-4 ...

  70  

 

   Figure  E8.  Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  golden  tilefish  IL-­‐10,  derived  from  the  assembled  5’  RACE  and  initial  fragment  products.    Black  text  reflects  the  5’  UTR,  the  red  text  indicates  the  start  (ATG)  codon,  the  green  text  marks  the  coding  region  for  the  golden  tilefish  IL-­‐10  protein,  and  the  blue  text  represents  the  overlap  between  the  5’-­‐RACE  and  initial  fragment  products.    The  initial  fragment  accounts  for  the  3’  portion  of  the  gene.      

   Figure  E9.  Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐10,  derived  from  the  assembled  5’  and  3’  RACE  products.    Black  text  reflects  the  5’  and  3’  UTR,  the  red  text  indicates  the  start  (ATG)  and  stop  (TGA)  codons,  the  continuous  green  and  blue  text  indicate  the  coding  region  for  the  red  snapper  IL-­‐10  protein,  and  the  blue  text  represents  the  overlap  between  the  5’-­‐  and  3’-­‐RACE  products.          

Page 84: Graduate Theses and Dissertations Graduate School 11-4 ...

  71  

   Figure  E10.  Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  red  snapper  IL-­‐10,  obtained  via  cloning  from  two  different  fish.    Black  text  represents  the  5’  and  3’  UTR,  red  text  indicates  the  start  (ATG)  and  stop  (TGA)  codon,  and  green  text  designates  the  coding  region.      

   Figure  E11.  Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  golden  tilefish  TNFα,  derived  from  the  assembled  3’-­‐RACE  and  initial  fragment  products.    Red  text  indicates  the  start  (ATG)  codon,  the  green  text  marks  the  coding  region  for  the  golden  tilefish  TNFα,  protein,  and  the  blue  text  represents  the  overlap  between  the  3’-­‐RACE  and  initial  fragment  products.    The  initial  fragment  accounts  for  the  5’  portion  of  the  gene.      

Page 85: Graduate Theses and Dissertations Graduate School 11-4 ...

  72  

   Figure  E12.  Experimentally  obtained  5’  to  3’  nucleotide  sequence  for  red  snapper  TNFα,  derived  from  the  assembled  5’-­‐  and  3’-­‐RACE  products.    Black  text  reflects  the  5’  and  3’  UTR,  the  red  text  indicates  the  start  (ATG)  and  possible  stop  (AAA)  codons,  the  continuous  green  and  blue  text  indicate  the  coding  region  for  the  red  snapper  TNFα  protein,  and  the  blue  text  represents  the  overlap  between  the  5’-­‐  and  3’-­‐RACE  products.    The  stop  codon  was  not  observed  in  the  cloned  and  confirmed  sequence  and  may  be  a  sequencing  artifact.      

   Figure  E13.  Confirmed  5’  to  3’  nucleotide  sequence  for  the  coding  region  of  red  snapper  TNFα,  obtained  via  cloning  from  two  different  fish.    Black  text  represents  the  5’  and  3’  UTR,  red  text  indicates  the  start  (ATG)  codon,  and  green  text  designates  the  coding  region.    The  stop  codon  is  absent.    

Page 86: Graduate Theses and Dissertations Graduate School 11-4 ...

  73  

         

APPENDIX  F:    

EXPERIMENTALLY  DERIVED  CYTOKINE  AMINO  ACID  SEQUENCES      

Table  F1.  Amino  acid  sequences  for  each  cytokine,  translated  from  the  experimentally  obtained  nucleotide  sequence  by  the  ExPASy  Translate  tool.      

     

Complete( Partial(

Golden'tilefish

MESKMKCNMTKMWSPKMPEGLDFEITHHPLTMRSVVNLIVAMEKFKGNRSESPLSTEFRDENLLNIMLDSMEEEETVFECFSAPSVQYSRTGEHQCRVTDSENRSLVLVPDSMELHAVMLQGGSENCKVHLNMSTYLDLTPSAGARTVALGITGKTLEDKNLYLSCHKDGEEPTLHLEAVANKDSLLRIGSSSDMERFLFYKHVTGLNNTTFVSVPYSDWYISTAGDDNKPVEMCLENARRHRTFNIQRQS

Red''''''snapper

MELHAVRLQGGSEESNEVHLNMATYAHPTPTTDALPVALGIKGTNLYLSCHKEGDKPTLHLEAVENNSSLSNISSTSDMVRSLFYKQDSESNSRGRHGG

Golden'tilefish

MNSRVFVATIVVLLAFSAISEGMTLRSLGVELHCRCIQTESKPIGRHIAKVELIIPNSHCEETEIIATLKVTGQEVCLDPEAPWVKRVINRIMSNRKR

Red''''''snapper

MSSRVFIASIVVVLAFLATSEGMALRSLGVELHCRCIQTESRPIGRHIGKVELIPPNSHCEETEIIATLKKTGQEVCLDPEAPWVKKVIEKIMSNSSRR

Golden'tilefish

MTPRSLLLCVLVNLSVLCTVWCTPLCNNDCCRFVERFPVRLERLRADYSQIRGYYEANDDLDRALLDQTIEHSLKTPFACHAINSILGFYLGTVLPEALAEVTEDTRDLKPHMESIQTIFDQLKTDVTRCKHYFSCKEQFDITTLNSTYTQMESKGLYKAMGEL

Red''''''snapper

MTPRSLLLCVLVNLSVLRTVWCTPLCNNDCCRFVERFPVRLERLRADYSQIRGYYEANDDLDRALLDQTIEHSLKTPFACHAINSILGFYLGTVLPEALAEVTEDTRDLKPHMESIQTIFDQLKTDVTRCKHYFSCKEQFDITTLNSTYTQMESKGLYKAMGELNLLFNYIETYLASKRPKNHVASA

Golden'tilefish

DGDREGAGKRLTPLSHRIWSYSDSVGNKVSLMSAVRSACQNTAQEDSYKVGQGWYNAIYLGAVFQLNRGDKLWTETNQLSELDTDEGKTFFGVFAL

Red''''''snapper

MVAYTTAPGDVEMGVDKRTVVLVEKKSSTQRMWKVSAALLFVAICVGGVMLFAWYWTRRPEMQTQSGQTGAQTGREATEKTDPHYTLRRISSKAKAAIHLEGSYDEGENSNHQLEWKNGQGQAFAQGGFRLVNNEIIIPQTGLYFVYSQASFRVSCSDGDAEGAGKGLMPLSHRIWRLSDSIGSRASLMSAVRSACQNTAQDDSYRDGHGWYNAIYLGAVFQLNKGDRLWTETNQLSELETEDGKTFFGVFAIEFP

Sequence(data(

ILC1β

ILC8

ILC10

TNFα

Cytokine Species Amino(acid(sequence

Page 87: Graduate Theses and Dissertations Graduate School 11-4 ...

  74  

         

APPENDIX  G:    

SEQUENCE  CONFIRMATION      

G.1  Golden  tilefish  IL-­‐1β  sequence  confirmation       To  confirm  the  coding  region  sequence  for  golden  tilefish  IL-­‐1β,  the  fragment  was  isolated  from  

two  different  fish,  designated  fish  F  and  fish  H,  and  compared  (Figures  D1-­‐5).    Discrepancies  were  

resolved  by  first  confirming  that  the  nucleotide  was  properly  recognized  by  visual  inspection  of  the  

chromatogram  file,  using  Chromas  lite.    Next,  nucleotide  sequences  were  compared  to  one  another  

using  BLASTn  (Figure  D3).    The  sequences  were  translated  into  amino  acids  using  the  Expasy  Translate  

tool  and  compared  using  BLASTp,  to  determine  whether  or  not  single  nucleotide  polymorphisms  altered  

the  final  protein  sequence  (Figure  D4).  

  There  were  two  discrepancies  between  fish  F  and  fish  H,  at  amino  acid  positions  50  and  240.    

Fish  F  has  S50,  encoded  for  by  TCA,  while  fish  H  has  P50,  encoded  for  by  CCA.    The  difference  therefore  

lies  in  the  presence  of  a  cysteine,  rather  than  a  threonine,  though  both  chromatogram  peaks  were  clear  

at  this  point.    A  multiple  sequence  alignment  was  performed  with  the  five  most  homologous  fish  and  the  

serine  residue  was  most  commonly  found  at  position  50  (Figure  D5).      

  The  second  discrepancy  arose  at  R240  in  fish  F  and  K240  in  fish  H,  both  similar  amino  acids  with  

positively  charged  R  groups.    R240  is  encoded  for  by  AGA  while  K240  is  encoded  by  AAA,  therefore,  the  

difference  between  the  two  is  the  A  versus  G  at  nucleotide  position  811.    Arginine  was  more  conserved  

across  the  Perciformes  screened,  therefore,  fish  F’s  sequence  was  selected  for  the  final  analysis.    

 

Page 88: Graduate Theses and Dissertations Graduate School 11-4 ...

  75  

   Figure  G1.  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  golden  tilefish  IL-­‐1β,  cloned  from  fish  F.    Red  text  indicates  the  stop  and  start  codons,  green  text  reflects  the  translated  region,  and  black  text  designates  the  untranslated  region.    Underlined  codons  are  those  that  had  single  nucleotide  differences  between  fish  F  and  fish  H.      

   Figure  G2.  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  golden  tilefish  IL-­‐1β,  cloned  from  fish  H.    Red  text  indicates  the  stop  and  start  codons,  green  text  reflects  the  translated  region,  and  black  text  designates  the  untranslated  region.    Underlined  codons  are  those  that  had  single  nucleotide  differences  between  fish  F  and  fish  H.      

Page 89: Graduate Theses and Dissertations Graduate School 11-4 ...

  76  

   

   Figure  G3.  Clustal  Omega  alignment  of  the  nucleotide  sequences  for  golden  tilefish  IL-­‐1β  from  fish  F  and  H.    Identical  nucleotides  are  indicated  by  an  asterisk  (*).    

Page 90: Graduate Theses and Dissertations Graduate School 11-4 ...

  77  

   Figure  G4.  Clustal  omega  alignment  of  the  protein  sequences  for  golden  tilefish  IL-­‐1β  from  fish  F  and  H.    Identical  (*)  and  similar  (:)  amino  acids  are  indicated.      

   

   Figure  G5.    Clustal  omega  alignment  of  golden  tilefish  IL-­‐1β  from  fish  F  and  H  and  the  top  five  homologous  species,  according  to  BLASTp.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified.  The  two  amino  acids  of  interest  are  shaded  in  grey.    The  GenBank  accession  codes  used  were:  

Page 91: Graduate Theses and Dissertations Graduate School 11-4 ...

  78  

European  seabass  (CAC41006.1),  mandarin  fish  (AAV65041.1),  barred  knifejaw  (ACH87392.1),  striped  trumpeter  (ACQ99510.1),  and  red  seabream  (AAP33156.1).      G.2  Golden  tilefish  IL-­‐8  sequence  confirmation       The  mature  protein  was  cloned  and  sequenced  from  two  different  fish,  fish  F  and  H,  and  the  

resulting  sequence  data  was  identical  between  the  two.      Comparing  direct  PCR  sequencing  results  from  

golden  tilefish  F  and  fish  H  confirmed  the  complete  protein,  including  the  signal  peptide  sequence  

(Figures  D6,  D7).    There  was  a  single  nucleotide  difference,  as  fish  F  had  an  adenine  at  position  102  and  

fish  H  had  a  guanine  (Figure  D8).    However,  this  did  not  change  the  translated  protein  sequence  (Figure  

D9).  

 

Figure  G6.    5’  to  3’  nucleotide  sequence  for  the  translated  region  of  golden  tilefish  IL-­‐8,  obtained  from  fish  F.    Black  text  indicates  the  untranslated  regions,  red  indicates  the  start  and  stop  codons,  and  green  text  shows  the  span  of  the  translated  region.          

 

Figure  G7.    5’  to  3’  nucleotide  sequence  for  the  translated  region  of  golden  tilefish  IL-­‐8,  obtained  from  fish  H.    Black  text  indicates  the  untranslated  regions,  red  indicates  the  start  and  stop  codons,  and  green  text  shows  the  span  of  the  translated  region.          

Page 92: Graduate Theses and Dissertations Graduate School 11-4 ...

  79  

 

Figure  G8.    Clustal  Omega  alignment  of  the  nucleotide  sequences  for  golden  tilefish  IL-­‐8  from  fish  F  and  H.    Identical  nucleotides  are  indicated  by  an  asterisk  (*).      

   Figure  G9.  Clustal  Omega  alignment  of  the  protein  sequences  for  golden  tilefish  IL-­‐8  from  fish  F  and  H.    Identical  (*)  and  amino  acids  are  indicated.      G.3  Red  snapper  IL-­‐8  sequence  confirmation       The  mature  protein  was  cloned  and  sequenced  from  two  different  fish,  fish  C  and  E,  and  the  

resulting  sequence  data  was  identical  between  the  two.      Comparing  direct  PCR  sequencing  results  from  

golden  tilefish  C  and  fish  E  confirmed  complete  protein,  including  the  signal  peptide  sequence  (Figures  

D10,  D11).    There  was  one  discrepancy,  at  nucleotide  141,  where  fish  C  had  a  cytosine  and  fish  E  had  a  

thymine,  however,  this  did  not  change  the  overall  protein  sequence  (Figures  D12,  D13).  

Page 93: Graduate Theses and Dissertations Graduate School 11-4 ...

  80  

 

Figure  G10.    5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐8,  obtained  from  fish  C.    Black  text  indicates  the  untranslated  regions,  red  indicates  the  start  and  stop  codons,  and  green  text  shows  the  span  of  the  translated  region.          

 

Figure  G11.    5’  to  3’  nucleotide  sequence  for  red  snapper  IL-­‐8,  obtained  from  fish  E.    Black  text  indicates  the  untranslated  regions,  red  indicates  the  start  and  stop  codons,  and  green  text  shows  the  span  of  the  translated  region.          

 

Figure  G12.    Clustal  Omega  alignment  of  the  nucleotide  sequences  for  red  snapper  IL-­‐8  from  fish  C  and  E.    Identical  nucleotides  are  indicated  by  an  asterisk  (*).    

       

Page 94: Graduate Theses and Dissertations Graduate School 11-4 ...

  81  

 

 Figure  G13.  Clustal  Omega  alignment  of  the  protein  sequences  for  red  snapper  IL-­‐8  from  fish  C  and  E.    Identical  (*)  amino  acids  are  indicated.      G.4  Red  snapper  IL-­‐10  sequence  confirmation       To  confirm  the  coding  region  sequence  for  red  snapper  IL-­‐10  the  fragment  was  isolated  from  

two  different  fish,  designated  fish  C  and  fish  E,  and  compared  (Figures  D14-­‐18).  Sequences  were  

analyzed  as  described  above.      One  discrepancy  was  noted.    Fish  C  had  residue  M151,  encoded  for  by  

ATG,  where  fish  E  had  residue  V151,  encoded  by  GTG.    When  compared  against  the  top  five  most  

homologous  species,  according  to  BLASTp,  the  methionine  was  more  conserved  (Figure  D18).    

Therefore,  sequence  from  fish  C  was  utilized.    

 

 Figure  G14.  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  red  snapper  IL-­‐10,  cloned  from  fish  C.    Black  text  indicates  the  untranslated  regions,  red  indicates  the  start  and  stop  codons,  and  green  text  shows  the  span  of  the  translated  region.      The  underlined  text  highlights  the  codon  that  differs  between  fish  C  and  fish  E.      

 Figure  G15.  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  red  snapper  IL-­‐10,  cloned  from  fish  E.    Black  text  indicates  the  untranslated  regions,  red  indicates  the  start  and  stop  codons,  and  green  text  shows  the  span  of  the  translated  region.      The  underlined  text  highlights  the  codon  that  differs  between  fish  C  and  fish  E.    

Page 95: Graduate Theses and Dissertations Graduate School 11-4 ...

  82  

 

   

   Figure  G16.    Clustal  Omega  alignment  of  the  nucleotide  sequences  for  red  snapper  IL-­‐10  from  fish  C  and  E.    Identical  nucleotides  are  indicated  by  an  asterisk  (*).          

Page 96: Graduate Theses and Dissertations Graduate School 11-4 ...

  83  

   Figure  G17.  Clustal  Omega  alignment  of  the  protein  sequences  for  red  snapper  IL-­‐10  from  fish  C  and  E.    Identical  (*)  and  similar  (:)  amino  acids  are  indicated.      

 Figure  G18.    Clustal  Omega  alignment  red  snapper  IL-­‐10  from  fish  C  and  E  and  the  top  five  homologous  species,  according  to  BLASTp.    Identical  (*)  or  similar  (:  or  .)  amino  acid  sequences  were  identified.  The  amino  acids  of  interest  is  shaded  in  grey.    The  GenBank  accession  codes  used  were:  European  seabass  (ABH09454.1),    gilthead  seabream  (AGS55345.1),  Atlantic  cod  (ABV64720.1),  green  spotted  puffer  (CAD67773.1),  and  rainbow  trout  (NP_001232028.1).                

Page 97: Graduate Theses and Dissertations Graduate School 11-4 ...

  84  

G.5  Red  snapper  TNFα  sequence  confirmation       To  confirm  the  coding  region  sequence  for  red  snapper  TNFα  the  fragment  was  isolated  from  

two  different  fish,  designated  fish  C  and  fish  E,  and  compared  as  described  above  (Figures  D19-­‐D22).  The  

only  difference  occurred  at  nucleotide  position  85.    Fish  C  had  a  guanine  residue  and  fish  E  had  an  

adenine  residue  at  this  location.    However,  as  this  was  in  the  5’  UTR,  it  was  not  considered  a  significant  

difference  (Figure  D21).    The  protein  sequences  were  identical  (Figure  D22).        

 

 

Figure  G19.  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  red  snapper  TNFα,  cloned  from  fish  C.    Black  text  indicates  the  untranslated  region,  red  indicates  the  stop  codons,  and  green  text  shows  the  span  of  the  translated  region.      The  underlined  text  highlights  the  nucleotide  that  differs  between  fish  C  and  fish  E.    

 

Figure  G20.  5’  to  3’  nucleotide  sequence  for  the  translated  region  of  red  snapper  TNFα,  cloned  from  fish  E.    Black  text  indicates  the  untranslated  regions,  red  indicates  the  start  codons,  and  green  text  shows  the  span  of  the  translated  region.      The  underlined  text  highlights  the  nucleotide  that  differs  between  fish  C  and  fish  E.    

Page 98: Graduate Theses and Dissertations Graduate School 11-4 ...

  85  

   

   Figure  G21.    Clustal  Omega  alignment  of  the  nucleotide  sequences  for  red  snapper  TNFα  from  fish  C  and  E.    Identical  nucleotides  are  indicated  by  an  asterisk  (*).      

Page 99: Graduate Theses and Dissertations Graduate School 11-4 ...

  86  

 

   Figure  G22.    Clustal  Omega  alignment  of  protein  sequences  for  red  snapper  TNFα  from  fish  C  and  E.    Identical  amino  acids  are  indicated  by  an  asterisk  (*).            

 

 

 

 

 

 

 

 

 

 

 

   

Page 100: Graduate Theses and Dissertations Graduate School 11-4 ...

  87  

         

APPENDIX  H:    

TABLE  OF  COMMON  NAMES  USED  AND  THEIR  ASSOCIATED  SCIENTIFIC  NAMES      

Table  H1:  List  of  all  common  names  used  in  this  thesis  with  the  associated  scientific  name.    

Common  name Scientific  name

African  clawed  frog Xenopus  laevis

Atlantic  cod Gadus  morhua

Atlantic  salmon Salmo  salar

Barred  knifejaw Oplegnathus  fasciatus

Channel  catfish   Ictalurus  punctatus  

Chicken Gallus  gallus

Common  carp Cyprinus  carpio

European  seabass Dicentrarchus  labrax

Fugu Takifugu  rubripes

Gilthead  seabream Sparus  aurata

Golden  tilefish Lopholatilus  chamaeleonticeps

Goldfish Carassius  auratus

Grass  carp Ctenopharyngodon  idella

Green  spotted  puffer Tetraodon  nigroviridis  

Haddock Melanogrammus  aeglefinus

Hapuku Polyprion  oxygeneios

Human Homo  sapiens

Humphead  snapper Lutjanus  sanguineus

Mandarin  fish Siniperca  chuatsi

Marmot Marmota  monax

Mouse Mus  musculus

Olive  flounder   Paralichthyes  olivaceus

Orange-­‐spotted  grouper Epinephalus  coioides

Platyfish Xiphophorous  maculatus

Rainbow  trout Oncorhynchus  mykiss

Rat Rattus  norvegicus  

Red  seabream Pagrus  major

Red  snapper Lutjanus  campechanus

Rockfish Sebastes  schlegeli

Silver  carp Hypophthalmichthys  molitrix

Smaller  spotted  catshark Scyliorhinus  canicula

Southern  bluefin  tuna Thunnus  maccoyii

Striped  trumpeyer Latris  lineata

Zebrafish Danio  rerio