Results 1 to 4 of 4

Thread: Human land Use and Amphibian Communities.

  1. #1
    Iratus ranunculus
    Guest

    Default Human land Use and Amphibian Communities.

    The Effect of Wetland Hydroperiod and Upland Habitat Structure on Amphibian Metacommunities: A Review
    By Iratus Ranunculus

    Amphibians which use aquatic breeding modes, particularly the Ranidae, have been used as model systems in ecology since the 1960s (McDairmid and Altig 1999). Their complex life cycles allow them to be used to study evolutionary and ecological processes such as predator-prey interactions (Formanowicz. 1978; Formanowicz 1986; Brodie and Formanowicz 1987; Pearl, Adams et al. 2003), competition (Griffiths, Edgar et al. 1991; Bardsley and Beebee 2000; Relyea 2002), and metapopulation dynamics (Van Buskirk and McCollum 2000; Marsh 2001; Petranka 2007). It is also this complex life cycle that makes them ideal organisms for examining the interdependency of wetlands and the surrounding upland areas. This is because aquatically breeding amphibians are linked directly to both due to their terrestrial habitat use and dispersal, and their aquatic reproduction (Petranka 2007). Unfortunately amphibians are currently in a world-wide decline (Collins and Storfer 2003). Understanding amphibian metapopulations and the conditions that structure amphibian communities is of paramount importance if we are to conserve the species which remain. Their fate may also serve as harbinger for what will occur to wetland ecosystems and the organisms which dwell in them as a whole. The following review details the importance of upland habitat and wetland hydroperiod in determining the persistence and metapopulation dynamics of species within wetland anuran communities.

    Amphibian metapopulation dynamics have often been viewed as a “ponds as patches” model (Marsh 2001) where breeding locations are used to delineate sub-populations that exchange migrants and define the units in which local extinction and colonization occur. However, many amphibians have short maximum dispersal distances (Smith and Green 2005) and the actual distribution of amphibian dispersal is log-normal. This means that most amphibians are site loyal but there are occasional long range dispersers. These individuals would be heavily impacted by the terrestrial habitat between wetland sites. It is possible to assess how upland habitat structure affects species composition and community structure. This can be done in several ways. The first is to use observed or experimentally manipulated changes in the makeup of terrestrial habitat to determine the effect on particular populations. The other is to examine species richness and composition across a gradient of habitat fragmentation, or habitat composition.

    Lithobates sylvaticus, the wood frog, is a common subject for both observational studies of populations and in experimental manipulation (Patrick, Harper et al. 2008; Windmiller, Homan et al. 2008) A particularly good example is the work done by Windmiller et al (2008). They evaluated the effects of urban development on a series of vernal pools in Massachusetts on wood frogs (L. sylvaticus) and Ambystomatid salamanders (Windmiller, Homan et al. 2008). They found massive and sudden population declines in all three species, including the near-extirpation of L. sylvaticus from one of the ponds. This decline persisted for at least a two year period post construction and may have been mediated by direct mortality of adults and the resulting lack of recruitment.

    This mechanism has experimental support as well. In work done by Patrick et. al. (2008), newly metamorphosed frogs selected experimentally manipulated habitat at a coarse grain of 2.2 ha forest treatments. They showed no preference for fine grained patches of suitable habitat during dispersal, but displayed it once they were settled into area at a scale of 2-4 square meters. This, combined with a lack of density-dependent habitat selection, resulted in small patches of habitat with very high numbers of frogs. The high densities lead to extremely high mortality in the population. This is particularly important because it has been shown in some Ranids, particularly Lithobates castesbeianus, that the most important life history stage for structuring amphibian populations are the number of metamorphs that leave the pond successfully (Govindarajulu, Altwegg et al. 2005). High mortality in the metamorph stage has the potential to cause population collapses and in the more terrestrial amphibians such as L. sylvaticus, result in extirpation in what once may have been a source habitat within the regional metacommunity.

    So what does this do when applied across a region with multiple species? The results vary somewhat depending on what measures are used to assess amphibian diversity (Knutson, Sauer et al. 1999; Guerry and Hunter 2002; Pillsbury and Miller 2008; Babbitt, Baber et al. 2009). Knutson et al (1999) found that amphibian species richness was positively associated with both forested and agricultural habitat, as well as with areas of forest-agricultural edge. This work is in agreement with Guerry and Hunter (2002) who showed that amphibians have particular preferences for open habitat, forested habitat, and to some degree on the adjacency of ponds to forested habitat. In this context an area which has a patchy mosaic of open and forested habitat will contain more species than a more homogenous landscape.
    However, Guerry and Hunter (2002) did not classify agricultural habitat and naturally occurring habitat differently. Certain processes that occur on agricultural lands have been shown to have devastating impacts on amphibian recruitment (Davidson 2004; Relyea 2009; Relyea 2009; Relyea and Jones 2009). Pesticides often act as endocrine disruptors (though others as neurotoxins) and have been found to sex-reverse genetically male frogs (Hayes, Collins et al. 2002; Hayes 2004). They have also been shown to the following effects: increase the time to metamorphosis and thus increases desiccation risk (Hayes, Case et al. 2006; Marquez-Garcia, Correa-Solis et al. 2009), decrease size at metamorphosis (Hayes, Case et al. 2006), depress the immune system (Rohr, Raffel et al. 2008) and increase the risk of mortality by predation by disrupting their ability to effectively respond to predators(Relyea 2009; Relyea 2009; Relyea and Jones 2009). In addition to pesticides, nitrogen fertilizers have also been shown to negatively impact larval amphibians (Griffis-Kyle and Ritchie 2007). Given these results, one would not expect to see amphibians present, let alone positively associated with these locations.

    The discrepancy may be explained by sampling method. Both of these studies used amphibian call surveys to detect the presence of amphibians. These surveys indicate that adults are present and attempting to breed, however they do nothing to indicate the success of those reproductive attempts. These agricultural areas may act as population sinks (Rowe, Hopkins et al. 2001), where adults migrate through suitable terrestrial habitat from suitable breeding habitat and fail to locally recruit new individuals into the population sufficient to replace losses. This makes the population continually reliant on new immigrants. This however, would not be detected by a call survey.

    In a larval survey amphibian species richness is similar in wetlands surrounded by woodlands, prairies, and rangeland, however it is significantly reduced in pasture(Babbitt, Baber et al. 2009). Additionally, mean abundance is highest in wetlands surrounded by forest, and lower elsewhere, particularly in range and pasture land, but in natural prairie as well. These results indicate that simply measuring presence is insufficient evidence to conclude that a population is doing well, but rather the abundance of species which are present needs to be taken into account as well. This is especially true with animals that may disperse between possible breeding areas. It also shows a distinct difference between the use of larval surveys and the call surveys found in Knutson et al (1999). In all of these cases, a call survey would not have detected the decreased larval abundance and thus recruitment in these populations and it may have detected non-breeding but present species.

    For urban-rural gradients the results are unequivocal. Conversion of upland habitat to human habitation is negatively associated with amphibian species richness (Parris 2006; Gagne and Fahrig 2007; Egan and Paton 2008; Pillsbury and Miller 2008; Randhir and Ekness 2009). In cases where this relationship is examined more closely, not only is the upland habitat affected, but so are the wetlands themselves. Hydroperiods often shift(Rubbo and Kiesecker 2005) and wetland hydroperiods act as a filter for amphibian species(Skelly 2001). This species filter is not only a direct effect of the hydroperiod through desiccation risk (Rowe and Dunson 1995; Bridges 2002) but also indirectly through mediating predation risk (Lardner 2000).

    In unmodified systems, if hydroperiod is short species composition will be biased in favor of species which exhibit the ability to metamorphose quickly. They need to get out of the pond before it dries. One would expect that both exploitative and interference competition should be very strong in these communities, and that this should be done in preference to exhibiting anti-predator mechanisms which pose a cost to development time (Anholt, Skelly et al. 1996; Lardner 2000; Relyea 2002) . This has been found in the genus Pseudacris (Skelly 1995), where Pseudacris trisceriata inhabits temporary ponds and under experimental and field conditions reaches metamorphosis much faster than Pseudacris crucifer which breeds in more permanent water. In this same system it was found that this difference is mediated by foraging behavior and activity level (Skelly 1995) and is the result of a trade-off between anti-predator behavior and growth. Reduced activity reduces the risk of predation, but also limits the amount of time the tadpole spends foraging. Activity level in this system is known to have this effect because it has been experimentally manipulated by tricaine anesthesia (Skelly 1994). Lithobates sylvaticus tadopoles were either left as controls or anesthetized with Tricaine and both groups were exposed to larval dragonflies in the genus Anax. The sharp decrease in activity was associated with an equivalently sharp drop in predation-mediated mortality.

    Relative activity level has also been shown to mediate interspecific exploitative competition in a Lithobates sylvaticus-Lithobates pipiens competitive system (wood frogs and leopard frogs respectively). It was found that under field conditions, Lithobates sylvaticus fares better in competition with L pipiens. So well in fact that they depress growth in the leopard frogs without themselves having their growth depressed (Relyea 2000). With the addition of predator cues from caged invertebrate and fish predators however the effect was reversed and the leopard frogs asymmetrically depressed wood frog growth. In the laboratory component of this experiment it was shown that wood frogs were consistently more active than leopard frogs, even when exposed to invertebrate predators. They were no differences between species when exposed to fish, because in the field neither of them typically breeds in water containing fish. This was enough for leopard frogs, perhaps with the addition of predator or competitor-induced morphological (Relyea 2002)shifts to reverse the asymmetry of competitive interactions. It was also found that changes in morphology that grant anti-predator advantages in wood frogs are the opposite of changes that enhance competitive ability (Relyea 2002). This probably mediates the interaction between wood frogs and leopard frogs observed.

    These anti-predator mechanisms are often phenotypically plastic, where one genotype dictates multiple phenotypes that are dependent on environmental context. Plastic phenotypes tend to evolve when an organism is subject to a temporally or spatially variable environment (Relyea ; Van Buskirk, McCollum et al. 1997; Schoeppner and Relyea 2009).
    Anti--predator mechanisms have the opposite relationship to hydroperiod that competitive mechanisms do, just as it has inverse effects on morphological phenotype (Lardner 2000). Amphibians which breed in temporary water where predation risk is predictably low not only tend to lack constitutive morphological defenses, they also do not exhibit morphological defenses induced by predation. Those that breed in spatiotemporally variable water bodies tended to display plastic phenotypes, while those which breed exclusively in permanent water exhibit stronger constitutive defenses (Lardner 2000; Jara and Perotti 2009). It has been shown the behavioral phenotypes tend to be more plastic than morphology (Relyea 2001). When this was subject to an exhaustive survey in North American and Eurasian anurans, it was found that while morphological plasticity was significantly related to the spatiotemporal variability in hydroperiod and thus predation risk, behavioral plasticity was not (Van Buskirk 2002). This may indicate that the costs of behavioral plasticity may be lower than the costs of morphological plasticity, probably because behavioral changes can respond to moment-by-moment changes in predation risk(Van Buskirk and Arioli 2002; Peacor 2006), while morphological changes are not as easily reversed (Van Buskirk 2002). Amphibians also appear to evaluate predation risk and respond in a way which is proportionate to that risk, rather than the responses being determined based on some sort of risk-threshold (Anholt, Skelly et al. 1996; Murray, Roth et al. 2004).

    Given that this is the case, do these mechanisms which are seen in the lab have the predicted effect in the field? What would be predicted is that amphibians which have short larval periods would be excluded from areas where hydroperiods are more permanent, because they lack the ability to deal with both the predators present in permanent water bodies and cannot compete with those that can. This has been shown repeatedly in the literature (Skelly, Werner et al. 1999; Baber 2004; Cunningham, Calhoun et al. 2007). The converse is also true, but for different reasons. Short hydroperiods exclude amphibians with long development times due to whole clutch mortality prior to metamorphosis, though temporary water can be used as foraging habitat or as a corridor for dispersal (Cook and Jennings 2007). However, some of these effects may be obscured due to differences in sampling techniques, some of which are unsuitable for detecting successful reproduction (Knutson, Sauer et al. 1999; Guerry and Hunter 2002; Babbitt, Baber et al. 2009).

    It should be clear from this review that both upland habitat and wetland habitat are essential for maintaining amphibian species richness and abundance within a region(Cunningham, Calhoun et al. 2007; Babbitt, Baber et al. 2009). Maintaining a complex mosaic of open and wooded upland habitat, as well as maintaining natural spatiotemporal variability in wetland hydroperiods will allow more species to persist within a region than would if degraded, only be able to support a subset of the regional species pool(Herrmann, Babbitt et al. 2005). Upland habitat provides dispersal corridors as well as foraging and overwintering sites for metamorphosed amphibians (Funk, Greene et al. 2005; Patrick, Harper et al. 2008). It is essential for maintaining an age structure that is conducive to species persistence(Govindarajulu, Altwegg et al. 2005), and for maintaining the interconnectivity of metapopulations (Greenwald, Purrenhage et al. 2009). The effects of hydroperiod are mediated by the interaction between interspecific competition and predation risk(Relyea 2004; Werner, Skelly et al. 2007), and directly impacts recruitment into the population. Not only is this useful for understanding ecological and evolutionary processes, but it also serves as a cautionary tale. Humans have been modifying the habitat upon which amphibians depend (Egan and Paton 2008; Windmiller, Homan et al. 2008; Baldwin and Demaynadier 2009). It has been done for housing, agriculture, mineral extraction, flood control, storm water retention and many other purposes. However this is having a negative impact on amphibian populations, and due to the status of amphibians as bioindicators (Beebee and Griffiths 2005), and their current mass-decline (Collins and Storfer 2003) we have legitimate cause for concern.


    Citations
    Anholt, B. R., D. K. Skelly, et al. (1996). "Factors modifying antipredator behavior in larval toads." Herpetologica 52(3): 301-313.

    Babbitt, K. J., M. J. Baber, et al. (2009). "INFLUENCE OF AGRICULTURAL UPLAND HABITAT TYPE ON LARVAL ANURAN ASSEMBLAGES IN SEASONALLY INUNDATED WETLANDS." Wetlands 29(1): 294-301.

    Baber, M. J. J. (2004). "The relationship between wetland hydroperiod and nestedness patterns in assemblages of larval amphibians and predatory macroinvertebrates." Oikos 107(1): 16-27.

    Baldwin, R. F. and P. G. Demaynadier (2009). "Assessing threats to pool-breeding amphibian habitat in an urbanizing landscape." Biological Conservation 142(8): 1628-1638.

    Bardsley, L. and T. J. C. Beebee (2000). "Competition between Bufo larvae in a eutrophic pond." Oecologia 124(1): 33-39.

    Beebee, T. J. C. and R. A. Griffiths (2005). "The amphibian decline crisis: A watershed for conservation biology?" Biological Conservation 125(3): 271-285.

    Bridges, C. M. (2002). "Tadpoles balance foraging and predator avoidance: Effects of predation, pond drying, and hunger." Journal of Herpetology 36(4): 627-634.

    Brodie, E. D. and D. R. Formanowicz (1987). "ANTIPREDATOR MECHANISMS OF LARVAL ANURANS - PROTECTION OF PALATABLE INDIVIDUALS." Herpetologica 43(3): 369-373.

    Collins, J. P. and A. Storfer (2003). "Global amphibian declines: sorting the hypotheses." Diversity and Distributions 9(2): 89-98.

    Cook, D. G. and M. R. Jennings (2007). "Microhabitat use of the California red-legged frog and introduced bullfrog in a seasonal marsh." Herpetologica 63(4): 430-440.

    Cunningham, J. M., A. J. K. Calhoun, et al. (2007). "Pond-breeding amphibian species richness and habitat selection in a beaver-modified landscape." Journal of Wildlife Management 71(8): 2517-2526.

    Davidson, C. (2004). "Declining downwind: Amphibian population declines in california and historical pesticide use." Ecological Applications 14(6): 1892-1902.

    Egan, R. and P. W. Paton (2008). Multiple Scale habitat Characteristics of Pond-Breeding Amphibians Across a Rural-Urban Gradient. Urban Herpetology. J. C. Mitchell, R. E. J. Brown and B. Bartholemew. Salt lake, SSAR: 53-67.

    Formanowicz, D. R. (1986). "ANURAN TADPOLE AQUATIC INSECT PREDATOR PREY INTERACTIONS - TADPOLE SIZE AND PREDATOR CAPTURE SUCCESS." Herpetologica 42(3): 367-373.

    Formanowicz., D. R. (1978). Palatability of a Community of Amphibian Larvae to Aquatic Invertebrate Predators. Biology, Adelphi University. Master of Science: 33.

    Funk, W. C., A. E. Greene, et al. (2005). "High dispersal in a frog species suggests that it is vulnerable to habitat fragmentation." Biology Letters 1(1): 13-16.

    Gagne, S. A. and L. Fahrig (2007). "Effect of landscape context on anuran communities in breeding ponds in the National Capital Region, Canada." Landscape Ecology 22(2): 205-215.

    Govindarajulu, P., R. Altwegg, et al. (2005). "Matrix model investigation of invasive species control: Bullfrogs on Vancouver Island." Ecological Applications 15(6): 2161-2170.

    Greenwald, K. R., J. L. Purrenhage, et al. (2009). "Landcover predicts isolation in Ambystoma salamanders across region and species." Biological Conservation 142(11): 2493-2500.

    Griffis-Kyle, K. L. and M. E. Ritchie (2007). "Amphibian survival, growth and development in response to mineral nitrogen exposure and predator cues in the field: an experimental approach." Oecologia 152(4): 633-642.

    Griffiths, R. A., P. W. Edgar, et al. (1991). "INTERSPECIFIC COMPETITION IN TADPOLES - GROWTH-INHIBITION AND GROWTH RETRIEVAL IN NATTERJACK TOADS, BUFO-CALAMITA." Journal of Animal Ecology 60(3): 1065-1076.

    Guerry, A. D. and M. L. Hunter (2002). "Amphibian distributions in a landscape of forests and agriculture: An examination of landscape composition and configuration." Conservation Biology 16(3): 745-754.

    Hayes, T. B. (2004). "There is no denying this: Defusing the confusion about atrazine." Bioscience 54(12): 1138-1149.

    Hayes, T. B., P. Case, et al. (2006). "Pesticide mixtures, endocrine disruption, and amphibian declines: are we underestimating the impact?" Environ Health Perspect 114 Suppl 1: 40-50.

    Hayes, T. B., A. Collins, et al. (2002). "Hermaphroditic, demasculinized frogs after exposure to the herbicide atrazine at low ecologically relevant doses." Proceedings of the National Academy of Sciences of the United States of America 99(8): 5476-5480.

    Herrmann, H. L., K. J. Babbitt, et al. (2005). "Effects of landscape characteristics on amphibian distribution in a forest-dominated landscape." Biological Conservation 123(2): 139-149.

    Jara, F. G. and M. G. Perotti (2009). "Toad Tadpole Responses to Predator Risk: Ontogenetic Change between Constitutive and Inducible Defenses." Journal of Herpetology 43(1): 82-88.

    Knutson, M. G., J. R. Sauer, et al. (1999). "Effects of landscape composition and wetland fragmentation on frog and toad abundance and species richness in Iowa and Wisconsin, USA." Conservation Biology 13(6): 1437-1446.

    Lardner, B. (2000). "Morphological and life history responses to predators in larvae of seven anurans." Oikos 88(1): 169-180.

    Marquez-Garcia, M., M. Correa-Solis, et al. (2009). "Effects of pond drying on morphological and life-history traits in the anuran Rhinella spinulosa (Anura: Bufonidae)." Evolutionary Ecology Research 11(5): 803-815.

    Marsh, P. C. M. (2001). "Metapopulation Dynamics and Amphibian Conservation." Conservation Biology 15(1): 40-49.

    McDairmid, R. W. and R. Altig, Eds. (1999). Tadpoles The Biology of Anuran Larvae.

    Murray, D. L., J. D. Roth, et al. (2004). "Predation risk avoidance by terrestrial amphibians: The role of prey experience and vulnerability to native and exotic predators." Ethology 110(8): 635-647.

    Parris, K. M. (2006). "Urban amphibian assemblages as metacommunities." Journal of Animal Ecology 75(3): 757-764.

    Patrick, D. A., E. B. Harper, et al. (2008). "Terrestrial habitat selection and strong density-dependent mortality in recently metamorphosed amphibians." Ecology 89(9): 2563-2574.

    Peacor, S. D. (2006). "Behavioural response of bullfrog tadpoles to chemical cues of predation risk are affected by cue age and water source." Hydrobiologia 573: 39-44.

    Pearl, C. A., M. J. Adams, et al. (2003). "Behavioral responses of anuran larvae to chemical cues of native and introduced predators in the Pacific Northwestern United States." Journal of Herpetology 37(3): 572-576.

    Petranka, J. W. (2007). "Evolution of complex life cycles of amphibians: bridging the gap between metapopulation dynamics and life history evolution." Evolutionary Ecology 21(6): 751-764.

    Pillsbury, F. C. and J. R. Miller (2008). "Habitat and landscape characteristics underlying anuran community structure along an urban-rural gradient." Ecological Applications 18(5): 1107-1118.

    Randhir, T. and P. Ekness (2009). "Urbanization effects on watershed habitat potential: a multivariate assessment of thresholds and interactions." Ecohydrology 2(1): 88-101.

    Relyea, R. A. The heritability of inducible defenses in tadpoles.

    Relyea, R. A. (2000). "Trait-mediated indirect effects in larval anurans: Reversing competition with the threat of predation." Ecology 81(8): 2278-2289.

    Relyea, R. A. (2001). "Morphological and behavioral plasticity of larval anurans in response to different predators." Ecology 82(2): 523-540.

    Relyea, R. A. (2002). "Competitor-induced plasticity in tadpoles: Consequences, cues, and connections to predator-induced plasticity." Ecological Monographs 72(4): 523-540.

    Relyea, R. A. (2002). "Costs of phenotypic plasticity." American Naturalist 159(3): 272-282.

    Relyea, R. A. (2004). "Fine-tuned phenotypes: Tadpole plasticity under 16 combinations of predators and competitors." Ecology 85(1): 172-179.

    Relyea, R. A. (2009). "A cocktail of contaminants: how mixtures of pesticides at low concentrations affect aquatic communities." Oecologia 159(2): 363-376.

    Relyea, R. A. (2009). "The lethal impact of Roundup on aquatic and terrestrial amphibians (vol 15, pg 1118, 2005)." Ecological Applications 19(1): 276-276.

    Relyea, R. A. and D. K. Jones (2009). "THE TOXICITY OF ROUNDUP ORIGINAL MAX (R) TO 13 SPECIES OF LARVAL AMPHIBIANS." Environmental Toxicology and Chemistry 28(9): 2004-2008.

    Rohr, J. R., T. R. Raffel, et al. (2008). "Understanding the net effects of pesticides on amphibian trematode infections." Ecological Applications 18(7): 1743-1753.

    Rowe, C. L. and W. A. Dunson (1995). "IMPACTS OF HYDROPERIOD ON GROWTH AND SURVIVAL OF LARVAL AMPHIBIANS IN TEMPORARY PONDS OF CENTRAL PENNSYLVANIA, USA." Oecologia 102(4): 397-403.

    Rowe, C. L., W. A. Hopkins, et al. (2001). "Failed recruitment of southern toads (Bufo terrestris) in a trace element-contaminated breeding habitat: Direct and indirect effects that may lead to a local population sink." Archives of Environmental Contamination and Toxicology 40(3): 399-405.

    Rubbo, M. J. and J. M. Kiesecker (2005). "Amphibian breeding distribution in an urbanized landscape." Conservation Biology 19(2): 504-511.

    Schoeppner, N. M. and R. A. Relyea (2009). "Phenotypic plasticity in response to fine-grained environmental variation in predation." Functional Ecology 23(3): 587-594.

    Skelly, D. K. (1994). "ACTIVITY LEVEL AND THE SUSCEPTIBILITY OF ANURAN LARVAE TO PREDATION." Animal Behaviour 47(2): 465-468.

    Skelly, D. K. (1995). "A BEHAVIORAL TRADE-OFF AND ITS CONSEQUENCES FOR THE DISTRIBUTION OF PSEUDACRIS TREEFROG LARVAE." Ecology 76(1): 150-164.

    Skelly, D. K. (1995). "COMPETITION AND THE DISTRIBUTION OF SPRING PEEPER LARVAE." Oecologia 103(2): 203-207.

    Skelly, D. K. (2001). "Distributions of pond-breeding anurans: An overview of mechanisms." Israel Journal of Zoology 47(4): 313-332.

    Skelly, D. K., E. E. Werner, et al. (1999). "Long-term distributional dynamics of a Michigan amphibian assemblage." Ecology 80(7): 2326-2337.

    Smith, M. A. and D. M. Green (2005). "Dispersal and the metapopulation paradigm in amphibian ecology and conservation: are all amphibian populations metapopulations?" Ecography 28(1): 110-128.

    Van Buskirk, J. (2002). "A comparative test of the adaptive plasticity hypothesis: Relationships between habitat and phenotype in anuran larvae." American Naturalist 160(1): 87-102.

    Van Buskirk, J. (2002). "Phenotypic lability and the evolution of predator-induced plasticity in tadpoles." Evolution 56(2): 361-370.

    Van Buskirk, J. and M. Arioli (2002). "Dosage response of an induced defense: How sensitive are tadpoles to predation risk?" Ecology 83(6): 1580-1585.

    Van Buskirk, J. and S. A. McCollum (2000). "Influence of tail shape on tadpole swimming performance." Journal of Experimental Biology 203(14): 2149-2158.

    Van Buskirk, J., S. A. McCollum, et al. (1997). "Natural selection for environmentally induced phenotypes in tadpoles." Evolution 51(6): 1983-1992.

    Werner, E. E., D. K. Skelly, et al. (2007). "Amphibian species richness across environmental gradients." Oikos 116(10): 1697-1712.

    Windmiller, B. S., R. N. Homan, et al. (2008). Breeding Amphibian Population Declines Following Loss of Upland Forest Habitat Around Vernal Pools in Massachusetts, USA. Urban
    Herpetology. J. C. Mitchell, R. E. J. Brown and B. Bartholemew. Salt Lake, SSAR: 41-53.

  2. # ADS
    Circuit advertisement
     

  3. #2
    bshmerlie
    Guest

    Default Re: Human land Use and Amphibian Communities.

    Is this a school paper and are we supposed to correct it for you?
    Sorry, in all seriousness...my first question to you is why a "review" and not an original study on the subject. Reminds me of my Uncle who is a Pysics professor at a college and his "work" was always a "review" or comparing and contrasting other peoples studys or experiments. Keep it original...it makes for much more interesting reading. Besides, someone with your "Star Trek" and "Star Wars" background has got to be able to find ways to spruce it up a bit. You lost me serveral times while reading your article. Although some science nerd my find it facinating you're missing out on the other 99% of the population. IMO

  4. #3
    Iratus ranunculus
    Guest

    Default Re: Human land Use and Amphibian Communities.

    Quote Originally Posted by bshmerlie View Post
    Is this a school paper and are we supposed to correct it for you?
    Sorry, in all seriousness...my first question to you is why a "review" and not an original study on the subject. Reminds me of my Uncle who is a Pysics professor at a college and his "work" was always a "review" or comparing and contrasting other peoples studys or experiments. Keep it original...it makes for much more interesting reading. Besides, someone with your "Star Trek" and "Star Wars" background has got to be able to find ways to spruce it up a bit. You lost me serveral times while reading your article. Although some science nerd my find it facinating you're missing out on the other 99% of the population. IMO
    No. Term paper from a couple semesters ago. It is one of those things that I will eventually clean up and publish when I have a dry spell in data collection. Sometimes it is useful for someone to summarize current literature. Reviews serve as good general references to allow someone to get the broad overview of the subject material.

    As for original work... I am currently collecting data, or rather, getting frog eggs sufficient to get the thousands of tadpoles I need to collect data. Next week I start collecting data on water snake foraging strategies. Once that is done, you guys get an original work

  5. #4
    Founder John's Avatar
    Join Date
    Aug 2008
    Nationality
    [Ireland]
    Location
    United States
    Age
    46
    Posts
    5,963
    Picture Albums: Member Photo Albums

    Default Re: Human land Use and Amphibian Communities.

    Thanks for posting this. I've moved it to the general discussion area.
    Founder of Frogforum.net (2008) and Caudata.org (2001)

Thread Information

Users Browsing this Thread

There are currently 1 users browsing this thread. (0 members and 1 guests)

Similar Threads

  1. Possible Amphibian Ban?; article
    By findiviglio in forum General Discussion & News
    Replies: 2
    Last Post: June 28th, 2010, 07:53 PM
  2. GBR Press: Benefits for communities and caecilians
    By Frog News in forum Amphibian News Feeds
    Replies: 0
    Last Post: December 27th, 2009, 01:28 AM
  3. Amphibian ID help
    By sebastianbawn in forum Other Amphibians
    Replies: 16
    Last Post: October 7th, 2009, 10:34 AM
  4. All Amphibian book
    By Tom Highum in forum Book Reviews
    Replies: 4
    Last Post: September 25th, 2009, 08:27 PM
  5. New in the Land of Oz
    By Marilu in forum Introductions Area
    Replies: 6
    Last Post: April 7th, 2009, 02:56 PM

Tags for this Thread

Posting Permissions

  • You may not post new threads
  • You may not post replies
  • You may not post attachments
  • You may not edit your posts
  •