CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I ·...

10
1 CHAPTER 12: TAMARISK ECOLOGY IN THE GRAND CANYONINVASION, CONTROL, AND FUTURE MANAGEMENT CATHERINE ESKRA INTRODUCTION An essential counterpart to the preceding discussion of native riparian plants (see M. Kelso in this volume) is an examination of invasive plants, and the interaction between hydrologic flow regime and invasive populations. Grand Canyon riparian areas are now home to numerous invasive plant species, many of which are characterized by prolific seed output, rapid germination and growth, robust vegetative spread, and/or seed dispersal over long distances (Makarick 2008). Invasive innercanyon species that have garnered the most attention from Grand Canyon managers and visitors include Russian olive (Elaeagnus angustifolia), camelthorn (Alhagi maurorum), Ravenna grass (Saccharum ravennae), cheatgrass (Bromus tectorum), sowthistles (Sonchus spp.), whitetop (Cardaria draba), and several knapweed species (Centaurea spp. and Acroptilon repens). Perhaps the most notorious riparian invasive in the Grand Canyon and other areas throughout the western United States is known as tamarisk or saltcedar, which is a hybridization of Tamarix species native to Eurasia (Gaskin & Schaal 2002). While tamarisk has been associated with negative impacts on native species and ecosystems, in many areas it is now incorporated into the riparian habitat. Riparian areas make up less than 2% of the land in the U.S. Southwest, but over 65% of wildlife in the region depend on this habitat (Makarick et al 2002). Management of these crucial areas requires a deep understanding of key actors such as tamarisk; many factors play into the interesting and complex question of how to treat this well established invader. HISTORY OF TAMARISK IN THE GRAND CANYON As is the case with many currently invasive plant species, tamarisk was first introduced and spread in the U.S. deliberately by private and public land managers (Wyman 2007). Brought in as early as 1805, tamarisk was planted widely as an ornamental, a wind break, and for bank stabilization purposes (see Figure 1) (Zouhar 2003, Stevens 2009). Essentially naturalized by the 1870s, saltcedar was planted through the 1930s and now occupies between 1 and 1.6 million acres from northern Mexico to Montana, and from Kansas to central California (Shafroth et al 2005). Efforts to control this wildly successful exotic species began in the 1960s and continue today. Tamarisk arrived in the Grand Canyon between 1922 and 1938, where it occupied high terraces and tributaries in the years before the 1963 completion of the Glen Canyon Dam (Stevens 2009). The dam drastically decreased downstream sediment loads, greatly reducing material for new terraces and sand bars (see papers by N. Figure 1. Tamarisk growing along the Colorado River in the Grand Canyon (NPS.gov/grca)

Transcript of CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I ·...

Page 1: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  1  

                       CHAPTER  12:  TAMARISK  ECOLOGY  IN  THE    

GRAND  CANYON—INVASION,  CONTROL,  AND  FUTURE  MANAGEMENT  

CATHERINE  ESKRA  

INTRODUCTION  An  essential  counterpart  to  the  preceding  discussion  of  native  riparian  plants  (see  M.  

Kelso  in  this  volume)  is  an  examination  of  invasive  plants,  and  the  interaction  between  hydrologic  flow  regime  and  invasive  populations.  Grand  Canyon  riparian  areas  are  now  home  to  numerous  invasive  plant  species,  many  of  which  are  characterized  by  prolific  seed  output,  rapid  germination  and  growth,  robust  vegetative  spread,  and/or  seed  dispersal  over  long  distances  (Makarick  2008).  Invasive  inner-­‐canyon  species  that  have  garnered  the  most  attention  from  Grand  Canyon  managers  and  visitors  include  Russian  olive  (Elaeagnus  angustifolia),  camelthorn  (Alhagi  maurorum),  Ravenna  grass  (Saccharum  ravennae),  cheatgrass  (Bromus  tectorum),  sowthistles  (Sonchus  spp.),  whitetop  (Cardaria  draba),  and  several  knapweed  species  (Centaurea  spp.  and  Acroptilon  repens).  Perhaps  the  most  notorious  riparian  invasive  in  the  Grand  Canyon  and  other  areas  throughout  the  western  United  States  is  known  as  tamarisk  or  saltcedar,  which  is  a  hybridization  of  Tamarix  species  native  to  Eurasia  (Gaskin  &  Schaal  2002).  While  tamarisk  has  been  associated  with  negative  impacts  on  native  species  and  ecosystems,  in  many  areas  it  is  now  incorporated  into  the  riparian  habitat.  Riparian  areas  make  up  less  than  2%  of  the  land  in  the  U.S.  Southwest,  but  over  65%  of  wildlife  in  the  region  depend  on  this  habitat  (Makarick  et  al  2002).  Management  of  these  crucial  areas  requires  a  deep  understanding  of  key  actors  such  as  tamarisk;  many  factors  play  into  the  interesting  and  complex  question  of  how  to  treat  this  well-­‐established  invader.    HISTORY  OF  TAMARISK  IN  THE  GRAND  CANYON  

As  is  the  case  with  many  currently  invasive  plant  species,  tamarisk  was  first  introduced  and  spread  in  the  U.S.  deliberately  by  private  and  public  land  managers  (Wyman  2007).  Brought  in  as  early  as  1805,  tamarisk  was  planted  widely  as  an  ornamental,  a  wind  break,  and  for  bank  stabilization  purposes  (see  Figure  1)  (Zouhar  2003,  Stevens  2009).  Essentially  naturalized  by  the  1870s,  saltcedar  was  planted  through  the  1930s  and  now  occupies  between  1  and  1.6  million  acres  from  northern  Mexico  to  Montana,  and  from  Kansas  to  central  California  (Shafroth  et  al  2005).  Efforts  to  control  this  wildly  successful  exotic  species  began  in  the  1960s  and  continue  today.    

 Tamarisk  arrived  in  the  Grand  

Canyon  between  1922  and  1938,  where  it  occupied  high  terraces  and  tributaries  in  the  years  before  the  1963  completion  of  the  Glen  Canyon  Dam  (Stevens  2009).  The  dam  drastically  decreased  downstream  sediment  loads,  greatly  reducing  material  for  new  terraces  and  sand  bars  (see  papers  by  N.    

Figure  1.  Tamarisk  growing  along  the  Colorado  River  in  the  Grand  Canyon  (NPS.gov/grca)  

Page 2: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  2  

Burley,  E.  Bartolomeo,  and  S.  Gibson  in  this  volume).  At  the  same  time,  the  reduction  in  peak  flood  magnitude  and  increase  in  base  flow  associated  with  hydropower  operations  resulted  in  a  more  stable  riparian  zone,  with  more  consistent  moisture  throughout  the  year  (Stevens  et  al  1995,  Ralston  2010).  The  riparian  plant  community  expanded  under  the  new  hydrologic  regime,  with  increases  in  cover  and  density  of  native  species  such  as  sandbar  willow  and  cattails,  and  a  dramatic  expansion  of  tamarisk  in  the  Grand  Canyon  (see  Figure  2)  (Turner  &  Karpiscak  1980,  Stevens  2002).  Perhaps  due  to  priority  effects,  superior  competitive  ability  in  post-­‐dam  conditions,  or  both,  tamarisk  has  become  a  riparian  dominant  and  in  some  areas  has  been  able  to  establish  a  vigorous  monoculture  (Stevens  2002,  Makarick  et  al  2002,  Zouhar  2003).    

 

   Figure  2.  Photographs  taken  in  1889/1890  (left)  and  replicated  in  the  late  1900s  (right)  show  the  expansion  of  Grand  Canyon  riparian  vegetation  over  the  century  (Webb  1996).    WHY  IS  TAMARISK  SO  SUCCESSFUL  IN  POST-­‐DAM  CONDITIONS?  

While  tamarisk  was  present  in  the  Grand  Canyon  prior  to  the  completion  of  the  Glen  Canyon  Dam,  there  is  a  general  consensus  that  this  species  expanded  rapidly  just  after  the  hydrologic  regime  change  (Turner&  Karpiscak  1980,  Stevens  2002,  Makarick  et  al  2002).  Factors  in  the  success  of  tamarisk  relative  to  many  native  species  in  the  more  stable  post-­‐dam  riparian  areas  can  be  associated  with  two  general  advantages:  the  ability  to  spread  quickly  and  the  capacity  to  grow  in  conditions  unfamiliar  to  many  natives.    Rapid  Spread  

Stevens  (2002)  states  that  a  mature  tamarisk  plant  can  produce  2.5x108  seeds  each  year;  while  other  sources  do  not  all  report  such  high  numbers,  the  production  of  hundreds  of  thousands  of  seeds  by  a  single  plant  may  be  common  (Makarick  2011).  Zouhar  (2003)  presents  a  comprehensive  review  of  tamarisk  regeneration  and  spread  mechanisms.  Tamarisk  plants  can  flower  in  their  first  year,  and  reproduce  sexually  through  much  of  the  growing  season,  which  contributes  to  prolific  seed  production.  Seeds  are  very  small  and  are  readily  dispersed  by  wind  and  water,  which  enhances  spread  especially  in  riparian  areas  such  as  the  inner  Grand  Canyon.  With  no  dormancy  requirements,  tamarisk  seeds  can  germinate  within  24  hours  in  favorable  conditions,  with  a  germination  rate  between  approximately  20%  and  50%  depending  on  the  time  of  year.  Tamarisk  can  also  produce  new  plants  vegetatively  from  stem  fragments,  which  may  be  transported  by  water  as  well.  The  production  of  numerous  propagules  capable  of  traveling  long  distances  likely  contributed  to  tamarisk  arriving  first  in  many  newly  stabilized  riparian  areas.  In  this  case,  inhibitory  priority  effects  (Young  et  al  2001)  associated  with  reduction  of  space  or  nutrients  for  later-­‐arriving  species  may  have  promoted  development  of  tamarisk  monoculture.  

Page 3: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  3  

 Advantage  Relative  to  Natives  in  Altered  Hydrologic  Conditions  

Tamarisk  is  often  found  in  areas  common  to  Fremont  cottonwood  (Populus  fremontii),  sandbar  willow  (Salix  exigua),  and  Goodding's  willow  (Salix  gooddingii).  Altered  stream  flows  across  the  U.S.  Southwest  have  been  associated  with  a  relative  increase  in  tamarisk  over  these  native  species  (Stromberg  et  al  2007,  Merritt  &  Poff  2010).  This  comparative  effect  may  be  due  at  least  as  much  to  the  generally  poor  performance  of  native  plants  under  altered  hydrology  as  to  tamarisk  doing  particularly  well  under  these  conditions  (Merritt  &  Poff  2010).  For  example,  native  species  such  as  cottonwood  are  adapted  to  release  seeds  in  synchrony  with  natural  flow  regimes,  taking  advantage  of  the  natural  scouring  floods  that  open  and  moisten  new  sites  for  germination  (see  M.  Kelso  in  this  volume).  With  altered  flow,  germination  sites  are  often  not  available  and  cottonwood  populations  decline  (Merritt  &  Poff  2010).  Unlike  many  natives,  tamarisk  produces  seeds  throughout  the  summer,  and  germination  is  favored  by  the  unnatural  consistent  base  flow  imposed  by  dam  operations  (Zouhar  2003,  Stromberg  et  al  2007).  Sources  such  as  Zouhar  (2003)  state  that  tamarisk  out-­‐performs  many  native  riparian  plants  in  "harsh"  environmental  extremes,  such  as  drought  and  high  salinity.  While  altered  flow  regimes  have  increased  base  flows  in  the  Grand  Canyon,  such  that  low-­‐lying  areas  are  in  fact  more  consistently  wet  throughout  the  year,  tamarisk  may  be  able  to  persist  in  the  slightly  higher,  drier  areas  and  thereby  produce  more  seeds  that  swamp  out  nearby  native  species.  Tamarisk  is  more  drought  tolerant  than  many  natives  because  it  is  a  facultative  phreatophyte  once  established  (meaning  that  it  can  use  surface  water  from  unsaturated  soil  as  well  as  ground  water);  it  also  has  a  deep  and  extensive  root  system  (DiTomaso  1998,  Zouhar  2003).  Salinity  in  Grand  Canyon  soils  has  increased  with  the  altered  hydrology  (see  A.  Oliver  in  this  volume),  which  would  favor  tamarisk  over  most  native  plants  that  do  not  tolerate  salinity.  Both  of  these  harsh  conditions  are  exacerbated  by  tamarisk  itself,  as  discussed  in  the  following  section.  

 TAMARISK  IMPACTS  ON  NATIVE  SPECIES  AND  ECOSYSTEM  FUNCTION  

One  reason  that  the  tamarisk  invasion  has  received  so  much  attention  is  its  demonstrated  direct  negative  impacts  on  native  species  and  ecosystems.  Once  established,  tamarisk  may  form  dense  monoculture  stands  through  vegetative  reproduction  that  exclude  other  plant  species  (Makarick  et  al  2002).  The  high  capacity  to  stabilize  sediments—associated  with  its  deep  roots—for  which  tamarisk  was  originally  planted  by  land  managers  combines  with  the  lack  of  scouring  floods  in  the  altered  hydrologic  regime  to  maintain  these  stands  for  decades,  which  would  not  have  been  possible  under  natural  hydrologic  conditions  (see  Figure  3)  (Lovich  &  Hoddle  2011).  Not  only  are  native  plants  unable  to  germinate  in  tamarisk  stands,  but  the  associated  sediment  impoundment  (Zouhar  2003)  may  reduce  materials  that  would  otherwise  create  new  open  sand  bars  for  native  germination.  Tamarisk  has  a  higher  evapotranspiration  rate  than  many  native  plants  under  dry  or  saline  conditions,  which  decreases  water  availability  and  exacerbates  these  conditions  under  which  tamarisk  outperforms  native  species  (see  Figure  3)  (Nagler  et  al  2003).  

Page 4: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  4  

             Figure  3.  Left:  Deep  tamarisk  roots  retain  sediment  and  keep  tamarisk  stands  in  place  in  the  absence  of  scouring  floods  (co.laplata.co.us).  Right:  Tamarisk  growing  near  water  sources  may  reduce  water  availability  to  other  plants  and  animals  (NPS.gov/grca).  

 According  to  some  sources,  tamarisk  also  increases  soil  salinity  via  salt  deposition  

excreted  from  the  leaves,  though  the  lack  of  flooding  associated  with  altered  hydrology  likely  plays  a  larger  role  in  salt  accumulation  in  the  soil  (Zouhar  2003).  High  salinity  prevents  the  germination  of  many  native  plants,  while  tamarisk  germination  is  not  negatively  affected  (Zouhar  2003).  Tamarisk  also  increases  fire  frequency  and  sprouts  rapidly  after  fire,  unlike  most  native  riparian  plants  (Lovich  &  Hoddle  2011).  In  some  sites  it  is  only  those  native  plants  that  are  also  halophytic  and  fire-­‐tolerant,  such  as  arrowweed  (Pluchea  sericea),  that  codominate  with  tamarisk;  in  many  cases  these  plants  are  ultimately  outcompeted  (Zouhar  2003).  

 In  addition  to  impacting  native  plants,  tamarisk  has  also  been  associated  with  negative  

effects  on  wildlife  and  ecosystem  species  diversity.  Some  studies  suggest  that  tamarisk  offers  little  value  to  most  native  reptiles,  amphibians,  mammals,  and  birds  (Lovich  &  de  Gouvenain  1998).  In  addition  to  decreasing  water  available  to  other  plants,  tamarisk  may  compromise  water  sources  for  bighorn  sheep,  deer,  salamanders,  toads,  and  other  wildlife  (Stephenson  &  Calcarone  1999,  Makarick  2002).  Some  comparisons  of  tamarisk-­‐dominated  areas  with  those  occupied  with  mature  native  cottonwood  and  willow  have  found  lower  abundance  and  diversity  of  various  vertebrates  and  macroinvertebrates  in  invaded  habitat  (Zouhar  2003,  Bailey  et  al  2001).  Counterexamples  have  also  been  found,  as  discussed  in  Section  VI  of  this  chapter.  However,  the  negative  impacts  on  ecosystems—both  real  and  perceived—and  negative  consequences  for  rafters  and  other  human  visitors  have  driven  efforts  to  reduce  or  eradicate  tamarisk  using  various  control  methods  (Makarick  et  al  2002).      TAMARISK  CONTROL  METHODS  

In  Grand  Canyon  National  Park,  managers  are  pursuing  chemical  and  mechanical  control  of  tamarisk  in  tributaries  and  side  canyons,  developed  areas,  and  springs  that  are  located  above  the  pre-­‐dam  water  level  of  the  river  (Makarick  2011).  Given  the  association  between  altered  hydrology  and  tamarisk  spread,  another  approach  is  to  determine  whether  the  High  Flow  Experiments  (discussed  by  N.  Burley  in  this  volume)  that  may  provide  some  benefits  of  the  

Page 5: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  5  

former  natural  flow  regime  might  also  help  to  control  tamarisk  (Ralston  2010).  In  addition,  the  tamarisk  leaf  beetle  is  a  biocontrol  that  was  approved  for  release  in  2001;  while  not  released  in  Grand  Canyon  NP,  it  was  found  there  in  2009  (Oltrogge  &  Makarick  2009).  As  with  many  invasive  species,  the  most  effective  control  is  gained  through  monitoring,  prevention,  early  detection,  and  local  eradication  (Zouhar  2003).  The  efficacy,  advantages,  and  disadvantages  of  control  methods  associated  with  Grand  Canyon  NP  are  discussed  below.  

 Chemical  and  Mechanical  Control  

Tamarisk  may  be  difficult  to  kill  using  only  chemical  or  mechanical  treatments  alone,  but  combined  approaches  may  be  more  effective  (Zouhar  2003).  Mechanical  control  can  be  appropriate  in  early  or  small  invasions  where  plants  are  easier  to  uproot  (Lovich  2000).  Zouhar  (2003)  discusses  root  plowing  and  cutting  as  initial  effective  approaches  to  heavy  invasions,  and  herbicides  such  as  imazapyr,  triclopyr,  and  glyphosate  to  enhance  mortality.  One  of  the  most  common  and  effective  treatments  is  to  remove  the  aboveground  portion  of  the  plant  and  apply  herbicide  to  the  remaining  cut  stump  (Lovich  2000).  Tamarisk  readily  resprouts  if  treated  with  fire  alone,  but  summer  fire  treatment  may  be  used  in  combination  with  chemical  and  mechanical  methods  to  enhance  tamarisk  mortality  (Zouhar  2003).  The  broad  scale  tamarisk  removal  project  Grand  Canyon  National  Park  involves  hand  crews  using  combination  chemical-­‐mechanical  treatments  best  suited  to  each  target  site  (Makarick  2011).  By  2011,  approximately  275,000  tamarisk  plants  had  been  removed  from  over  6,000  acres  in  the  park,  with  only  12  percent  of  treated  trees  requiring  additional  control;  the  project  is  now  in  a  cyclic  maintenance  phase  (Makarick  2011).  

 Effects  of  HFEs  on  Tamarisk    

Ralston  (2010)  discusses  riparian  vegetation  changes  in  response  to  the  HFEs,  and  the  March  2008  HFE  in  particular.  The  1996  and  2008  experimental  flows  primarily  buried  vegetation  rather  than  scouring  it,  so  these  would  not  have  played  an  important  role  in  removing  well-­‐established  tamarisk  stands.  Even  scouring  floods  do  not  necessarily  favor  natives,  however—following  the  uncontrolled  scouring  floods  in  the  mid-­‐1980s,  tamarisk  seedlings  dominated  newly  exposed  sediment  deposits.  By  comparison  to  these  uncontrolled  floods  and  a  Low  Summer  Steady  Flow  experiment  in  2000,  the  March  2008  HFE  resulted  in  relatively  lower  subsequent  tamarisk  seedling  establishment.  This  suggests  that,  even  if  the  HFEs  cannot  remove  most  established  tamarisk,  at  least  these  flows  will  not  contribute  significantly  to  its  spread  if  they  are  timed  before  the  April-­‐September  tamarisk  seed  production.  A  March  flow  can  provide  a  chance  for  native  plants  to  establish  in  open  areas,  potentially  reducing  site  availability  for  tamarisk  seedlings.    Tamarisk  Leaf  Beetle  Biocontrol    

In  some  cases,  introducing  herbivores  from  an  invasive  plant's  native  range  can  help  to  control  the  plant  populations;  the  tamarisk  leaf  beetle  (Diorhabda  spp.)  has  been  investigated  as  a  biocontrol  for  tamarisk  (Zouhar  2003).  After  study  to  determine  the  predicted  range  limits  and  targets  of  this  species,  tamarisk  beetles  were  released  in  some  western  areas  in  2001  (Oltrogge  &  Makarick  2009).  Apprehension  about  releasing  these  beetles  relates  to  the  potential  for  damage  to  non-­‐target  plants,  losing  the  ecological  or  economic  benefits  of  tamarisk,  leaving  open  niches  for  other  invasive  plants,  and  the  fact  that  native  vegetation  may  not  be  able  to  recover  or  survive  in  areas  where  tamarisk  is  removed  (Zouhar  2003).  The  beetle  was  not  to  be  released  within  200  miles  of  habitat  for  the  southwestern  willow  flycatcher,  an  endangered  bird  that  now  nests  in  tamarisk;  it  was  not  released  in  Grand  Canyon  National  Park  (Oltrogge  &  

Page 6: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  6  

Makarick  2009).  The  tamarisk  beetle  was  found  in  the  park  and  other  unanticipated  locations  in  2009,  however,  and  the  USDA  halted  beetle  release  at  that  time  in  response  to  a  lawsuit  to  protect  willow  flycatcher  habitat  (Sands  2009,  Makarick  2011).  Beetle  populations  are  currently  widespread  in  the  Colorado  Plateau  (see  Figure  4),  and  obvious  signs  of  defoliation  are  visible  in  the  Grand  Canyon  (Makarick  2011).  Studies  have  not  yet  shown  that  the  beetle  contributes  to  tamarisk  mortality,  but  episodic  defoliation  does  reduce  tamarisk  water  use,  which  is  a  desired  outcome  of  tamarisk  control  (Hultine  et  al  2010a).        IMPACTS  OF  TAMARISK  REMOVAL  ON  NATIVE  ECOSYSTEMS  

Tamarisk  may  not  be  ideal  habitat  for  wildlife,  but  especially  in  heavily  invaded  areas  it  may  be  in  use  as  the  best  habitat  available  (Zouhar  2003).  In  the  Grand  Canyon,  many  bird  species  do  nest  in  tamarisk,  including  mourning  dove,  Southwestern  willow  flycatcher,  Bell's  vireo,  yellow  warbler,  and  blue  grosbeak  (Makarick  et  al  2002).  Some  insect  species  use  tamarisk,  and  these  provide  food  for  foraging  bats;  peregrine  falcons  may  in  turn  prey  on  bats  or  birds  foraging  around  tamarisk  patches  (Makarick  et  al  2002).  Anderson  (1998)  suggests  that  at  higher  elevations,  the  diversity  of  communities  associated  with  tamarisk  stands  is  not  necessarily  lower  than  diversity  in  native  vegetation.    

 The  relative  use  of  tamarisk  can  vary  dramatically  by  geographic  location  and  species.  If  

this  invasive  plant  has  occupied  an  area  for  a  substantial  length  of  time  there  may  be  other  species  reliant  on  its  presence.  The  southwestern  willow  flycatcher  is  one  example:  some  studies  on  this  endangered  bird  have  found  no  negative  effects  from  breeding  in  tamarisk  (Sogge  et  al  2008),  and  the  bird's  reliance  on  this  habitat  in  some  areas  has  led  to  a  halt  in  tamarisk  biocontrol  (as  discussed  above).  Another  example  is  described  by  Converse  et  al  (1998),  which  found  that  densities  of  subadult  humpback  chub—an  endangered  fish—were  almost  twice  as  high  near  vegetated  shorelines  (occupied  largely  by  tamarisk)  as  the  chub  densities  associated  with  natural  talus  and  debris  fan  shorelines.  In  this  case,  the  chub  may  be  making  use  of  increased  cover  and  shade  associated  with  these  novel  riparian  shorelines  to  avoid  predation  from  non-­‐native  fish.  In  both  of  these  examples,  and  in  other  cases  of  wildlife  utilizing  tamarisk,  the  question  is  whether  tamarisk  control  and  removal  can  be  followed  by  replacement  with  native  species  that  will  provide  at  least  the  same  habitat  value.  Stromberg  (1998)  compared  30  traits  between  locations  occupied  by  tamarisk  and  by  Fremont  cottonwood  in  riparian  areas,  and  found  that  the  functional  role  of  these  two  species  was  equivalent  for  

 Figure  4.  Distribution  of  the  tamarisk  leaf  beetle  in  the  Colorado  Plateau  as  of  2010,  including  locations  in  Grand  Canyon  National  Park  (http://www.fronterasdesk.org/news/2011/sep/05/tamarisk-­‐river-­‐environment-­‐non-­‐native-­‐plants)  

Page 7: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  7  

about  half  of  the  traits  that  indicate  riparian  ecosystem  function.  Can  the  system  that  emerges  after  tamarisk  control  permanently  exceed  this  level  of  function?    REMOVAL  VERSUS  RECONCILIATION  

While  negative  effects  have  been  associated  with  the  presence  of  invasive  tamarisk,  there  may  also  be  negative  consequences  to  its  removal  in  riparian  areas  in  the  Grand  Canyon  and  throughout  the  west.    Management  of  this  species  requires  a  site-­‐specific  evaluation  of  these  factors,  incorporating  a  consideration  of  the  economic  cost  of  removal  and  the  likelihood  of  its  efficacy  as  well  (Shafroth  et  al  2005).  In  the  Grand  Canyon,  tamarisk  occupies  a  novel  niche  space  that  was  not  present  before  the  completion  of  the  Glen  Canyon  dam.  Native  plants  are  not  necessarily  well-­‐suited  to  this  entire  niche  even  in  the  "best  case"  scenario  of  complete  tamarisk  eradication.  Tamarisk  removal  without  native  colonization  can  leave  open  space  to  facilitate  the  spread  of  many  other  opportunistic  invasive  plants,  and  result  in  net  habitat  degradation  (Sogge  et  al  2008,  Hultine  et  al  2010b).    

 Ecologists  have  suggested  that  incorporating  some  level  of  native  vegetation  into  

tamarisk-­‐dominated  areas  may  bring  the  habitat  benefits  of  natives  without  the  higher  cost  and  risk  associated  with  full-­‐scale  tamarisk  removal  efforts  (van  Riper  III  et  al  2008).  Here  the  question  is  whether  this  mixed  community  can  be  self-­‐sustaining  under  altered  hydrologic  conditions,  even  with  the  potential  aid  of  experimental  flows,  or  if  a  tamarisk  monoculture  would  soon  return.  Examples  of  successful  establishment  of  native  communities  in  formerly  tamarisk-­‐dominated  areas  are  few  and  far  between,  even  with  more  comprehensive  removal  methods;  long-­‐term  persistence  of  these  communities  is  still  more  elusive  (Zouhar  2003).  In  the  Grand  Canyon,  where  there  is  little  flexibility  to  reincorporate  a  significant  degree  of  natural  flow  characteristics  into  the  regime,  replacing  tamarisk  with  a  sustainable  native  plant  community  seems  unlikely  if  not  impossible  in  many  areas.  Grand  Canyon  National  Park  has  pursued  a  program  of  tamarisk  control  in  areas  that  are  newly  invaded  or  where  removal  potentially  has  a  high  value  for  wildlife  or  humans,  which  is  a  well-­‐reasoned  approach  to  this  high-­‐profile  invasion.  The  composition  and  value  of  what  ultimately  persists  in  these  treatment  areas,  however,  remains  to  be  seen.          SOURCES  CITED    Anderson  B.  1998.  The  case  for  salt  cedar.  Restoration  and  Management  Notes  16:  130-­‐134.    Bailey  JK,  JA  Schweitzer,  &  TG  Whitham.  2001.  Salt  cedar  negatively  affects  biodiversity  of  

aquatic  macroinvertebrates.  Wetlands  21:  442-­‐447.    Converse,  Yvette  K.;  Hawkins,  Charles  P.;  Valdez,  Richard  A.  1998.  Habitat  relationships  of  

subadult  humpback  chub  in  the  Colorado  River  through  Grand  Canyon:  spatial  variability  and  implications  of  flow  regulation.  Regulated  Rivers:  Research  and  Management.  14(3):  267-­‐284.  

 Di  Tomaso  JM.  1998.  Impact,  biology,  and  ecology  of  saltcedar  (Tamarix  spp.)  in  the  

Southwestern  United  States.  Weed  Technology  12:  326-­‐336.  

Page 8: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  8  

 Gaskin  JF  &  BA  Schaal.  2002  Hybird  Tamarix  widespread  in  U.S.  invasion  and  undetected  in  

native  Asian  range.  Proceedings  of  the  National  Academy  of  Sciences  99:  11256-­‐11259.    Hultine  KR,  PL  Nagler,  K  Morine,  SE  Bush,  KG  Burtch,  PE  Dennison,  EP  Glenn,  &  JR  Ehleringer.  

2010a.  Sap  flux-­‐scaled  transpiration  by  tamarisk  (Tamarix  spp.)  before,  during  and  after  episodic  defoliation  by  the  saltcedar  leaf  beetle  (Diorhabda  carinulata).  Agricultural  and  Forest  Meteorology  150:  1467-­‐1475.  

 Hultine  KR,  J  Belnap,  C  van  Riper  III,  JR  Ehleringer,  PE  Dennison,  ME  Lee,  PL  Nagler,  KA  Snyder,  

SM  Uselman,  &  JB  West.  2010b.  Tamarisk  biocontrol  in  the  western  United  States:  ecological  and  societal  implications.  Frontiers  in  Ecology  and  the  Environment  8:  467-­‐474.  

 Lovich  JE  &  RC  de  Gouvenain.  1998.  Saltcedar  invasion  in  desert  wetlands  of  the  southwestern  

United  States:  ecological  and  political  implications.  Pp.  447-­‐467  in:     Majumdar  SK,  EW  Miller,  FJ  Brenner,  eds.  Ecology  of  wetlands  and  associated  systems.  

The  Pennsylvania  Academy  of  Science.    Lovich  J.  2000.  Tamarix  ramosissima/Tamarix  chinensis/Tamarix  gallica/Tamarix  parviflora.  Pp.  

312-­‐317  in:  Bossard  CC,  JM  Randall,  &  MC  Hoshovsky,  eds.  Invasive  plants  of  California's  wildlands.  Berkeley,  CA:  University  of  California  Press.  

 Lovich  J  &  M  Hoddle.  2011.  Saltcedar.  Center  for  Invasive  Species  Research,  University  of  

California  Riverside.  Accessed  online  2/2012.     <http://cisr.ucr.edu/saltcedar.html>    Makarick  L,  L  Leap,  G  Chesher,  T  Felger,  &  S  White.  2002.  Environmental  assessment/assessment  

of  effect:  tamarisk  management  and  tributary  restoration,  Grand  Canyon  National  Park,  Arizona.  Accessed  online  2/2012.  

  <http://www.nps.gov/grca/naturescience/exotic-­‐tamarisk.htm>    Makarick  L.  2008.  Vegetation  Management—Exotic  Plant  Species.  Grand  Canyon  National  Park  

site  bulletin.  Accessed  online  2/2012.     <http://www.nps.gov/grca/naturescience/exotic-­‐tamarisk.htm>    Makarick  L.  2011.  Tamarisk  management  and  tributary  restoration.  Grand  Canyon  National  Park  

site  bulletin.  Accessed  online  2/2012.     <http://www.nps.gov/grca/naturescience/exotic-­‐tamarisk.htm>    Merritt  DM  &  NL  Poff.  2010.  Shifting  dominance  of  riparian  Populus  and  Tamarix  along  gradients  

of  flow  alteration  in  western  North  American  rivers.  Ecological  Applications  20:  135-­‐152.    Nagler  PL,  EP  Glenn,  TL  Thompson.  2003.  Comparison  of  transpiration  rates  among  saltcedar,  

cottonwood,  and  willow  trees  by  sap  flow  and  canopy  temperature  methods.  Agricultural  and  Forest  Meteorology  116:  73-­‐89.  

   

Page 9: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  9  

Oltrogge  M  &  L  Makarick.  2009.  Tamarisk  beetle  (Diorhabda  spp.)  found  along  Colorado  River  within  Grand  Canyon  National  Park.  Grand  Canyon  National  park  news  release  10/21/2009.  Accessed  online  2/2012.    

  <http://www.nps.gov/grca/parknews/tamarisk-­‐beetle.htm>    Ralston  BE.  2010.  Riparian  vegetation  response  to  the  March  2008  short-­‐duration,  high-­‐flow  

experiment—implications  of  timing  and  frequency  of  flood  disturbance  on  nonnative  plant  establishment  along  the  Colorado  River  below  Glen  Canyon  Dam.  U.S.  Geological  Survey  Open-­‐File  Report  2010–1022.  Accessed  online  1/2012.  <http://pubs.usgs.gov/of/2010/1022/>  

 Sands  W.  2009.  Tamarisk  beetle  program  put  on  hold.  The  Durango  Telegraph  8/20/2009.  

Accessed  online  1/2012.     <http://www.durangotelegraph.com/index.cfm/archives/2009/august-­‐20-­‐2009>    Shafroth  PB,  JR  Cleverly,  TL  Dudley,  JP  Taylor,  C  Van  Riper  III,  EP  Weeks,  &  JN  Stuart.  2005.  

Control  of  Tamarix  in  the  Western  United  States:  implications  for  water  salvage,  wildlife  use,  and  riparian  restoration.  Environmental  Management  35:  231-­‐246.  

 Sogge  MK,  SJ  Sferra,  &  EH  Paxton.  2008.  Tamarix  as  habitat  for  birds:  implications  for  riparian  

restoration  in  the  Southwestern  United  States.  Restoration  Ecology  16:  146-­‐154.    Stephenson  JR  &  GM  Calcarone.  1999.  Mountain  and  foothills  ecosystems:  habitat  and  species  

conservation  issues.  Pp.  15-­‐60  in:  Southern  California  mountains  and  foothills  assessment.  Gen.  Tech.  Rep.  PSW-­‐GTR-­‐172.  Albany,  CA:  U.S.  Department  of  Agriculture,  Forest  Service,  Pacific  Southwest  Research  Station.  

 Stevens  LE,  JC  Schmidt,  TJ  Ayers,  &  BT  Brown.  1995.  Flow  regulation,  geomorphology,  and  

Colorado  River  marsh  development  in  the  Grand  Canyon,  Arizona.  Ecological  Applications  5:  1025–1039.  

 Stevens  LE.  2002.  Exotic  tamarisk  on  the  Colorado  Plateau.  In:  Canyons,  cultures,  and  

environmental  change:  an  introduction  to  the  land-­‐use  history  of  the  Colorado  Plateau.  Ed.  JD  Grahame  &  TD  Sisk.  Accessed  online  1/2012.  

  <http://cpluhna.nau.edu/Biota/tamarisk.htm>    Stevens  L.  2009.  Scourge  of  the  west:  the  natural  history  of  tamarisk  in  the  Grand  Canyon.  Grand  

Canyon  River  Guides.  Accessed  online  1/2012.    <http://www.gcrg.org/bqr/6-­‐2/scourge.htm>  

   Stromberg,  Juliet  C.  1998.  Functional  equivalency  of  saltcedar  (Tamarix  chinensis)  and  Fremont  

cottonwood  (Populus  fremontii)  along  a  free-­‐flowing  river.  Wetlands.  18:  675-­‐686.    Stromberg  JC,  S  J  Lite,  R  Marler,  C  Paradzick,  PB  Shafroth,  D  Shorrock,  JM  White,  &  MS  White.  

2007.  Altered  stream-­‐flow  regimes  and  invasive  plant  species:  the  Tamarix  case.  Global  Ecology  and  Biogeography  16:  381-­‐393.  

   

Page 10: CHAPTER’12: TAMARISK ECOLOGY&IN&THE& GRAND ANYON I · acres#from#northern#Mexico#to#Montana,#and#from#Kansas#to#central#California(Shafroth#et#al# 2005).#Efforts#to#control#this#wildly#successful#exotic#species#began#in#the1960s#and#continue#

  10  

Turner  RM  &  MM  Karpiscak.  1980.  Recent  vegetation  changes  along  the     Colorado  River  between  Glen  Canyon  Dam  and  Lake  Mead,  Arizona.  U.S.     Geological  Survey  Professional  Paper  No.  1132.  USGS,  Washington,  D.C.    Van  Riper  III  C,  KL  Paxton,  C  O'Brien,  PB  Shafroth,  LJ  McGrath.  2008.  Rethinking  avian  response  

to  Tamarix  on  the  Lower  Colorado  River:  a  threshold  hypothesis.  Restoration  Ecology  16:  155-­‐167.  

 Webb  RH.  1996.  Grand  Canyon,  a  century  of  change.  Tuczon:  University  of  Arizona  Press.    Wyman  S.  2007.  Saltcedar  (Tamarix).  Briefing  paper  for  Bureau  of  Land  Management,  National  

Riparian  Service  Team.  Accessed  online  2/2012.  <www.blm.gov/or/programs/nrst/files/tamarisk_paper.pdf>  

 Young  TP,  JM  Chase,  RT  Huddleston.  2001.  Community  succession  and  assembly:  comparing,  

contrasting,  and  combining  paradigms  in  the  context  of  ecological  restoration.  Ecological  Restoration  19:  5-­‐18.  

 Zouhar,  Kris.  2003.  Tamarix  spp.  In:  Fire  Effects  Information  System.  U.S.  Department  of  

Agriculture,  Forest  Service,  Rocky  Mountain  Research  Station,  Fire  Sciences  Laboratory.  Accessed  online  1/2012.  

  <  http://www.fs.fed.us/database/feis/>