nena masthead
NENA Home Staff & Editors For Readers For Authors

Assessment of Genetic Divergence between Lacustrine and Riverine Smallmouth Bass in Lake Erie and Four Tributaries
W. Calvin Borden

Northeastern Naturalist, Volume 15, Issue 3 (2008): 335–348

Full-text pdf (Accessible only to subscribers.To subscribe click here.)

 

Access Journal Content

Open access browsing of table of contents and abstract pages. Full text pdfs available for download for subscribers.



Current Issue: Vol. 30 (3)
NENA 30(3)

Check out NENA's latest Monograph:

Monograph 22
NENA monograph 22

All Regular Issues

Monographs

Special Issues

 

submit

 

subscribe

 

JSTOR logoClarivate logoWeb of science logoBioOne logo EbscoHOST logoProQuest logo

2008 NORTHEASTERN NATURALIST 15(3):335–348 Assessment of Genetic Divergence between Lacustrine and Riverine Smallmouth Bass in Lake Erie and Four Tributaries W. Calvin Borden* Abstract - Diverse freshwater lacustrine fishes enter tributaries to spawn, but resident riverine members may also occupy these same tributaries. While mark-recapture and biotelemetry studies suggest reproductive isolation between such populations, the assertion has rarely been tested genetically. To address this question, Micropterus dolomieu (Smallmouth Bass) from the southern shoreline of Lake Erie were compared genetically to bass in adjacent tributaries. Results from mitochondrial DNA sequences support the hypothesis that lacustrine and riverine populations segregate. Furthermore, divergences among tributary populations were often as large as those divergences between lacustrine and riverine bass, suggesting that each river population may become genetically distinct. Introduction Migration in fishes is a common phenomenon generally related to food distribution, habitat preferences, and reproductive requirements (Lucas and Baras 2001). While diadromous migrations (those occurring between fresh and salt waters) are often the most spectacular, potamodromous migrations (those occurring solely within freshwaters) appear to be more common than previously recognized. Fishes display seasonal migrations in search of spawning sites, optimal water temperature, or flow (Bunt et al. 2002). Consequently, discrete populations may have contact over relatively small spatial and temporal scales and the genetic consequences of this contact are usually unknown. Black basses are well-recognized members of both riverine and lacustrine warmwater fish communities and one of the top predators in aquatic ecosystems (Sowa and Rabeni 1995). Micropterus dolomieu Lacépède (Smallmouth Bass) are indigenous to Lake Erie and all of its tributaries (Trautman 1981) and colonized these waters from the Mississippian refugium (Mandrak and Crossman 1992) following retreat of the last glacial advance ≈14,000 years ago. Adult lacustrine Smallmouth Bass follow a structured migratory pattern among spawning sites and seasonal home ranges (Ridgeway et al. 2002). Each spring, more than 70% of spawning males return to within 100 m of their previous year’s nest (Ridgway et al. 1991). Thus, philopatry is probable (Gross et al. 1994), though not yet demonstrated. The preferred nest habitat is sand or gravel in the shal- *Department of Biological, Geological, and Environmental Sciences, Cleveland State University, Cleveland, OH 44115. Current address - School of Biological Sciences, University of Nebraska-Lincoln, 348 Manter Hall, Lincoln, NE 68588-0118; cal. borden@gmail.com. 336 Northeastern Naturalist Vol. 15, No. 3 lows of lake margins and river mouths (Scott and Crossman 1973). Nestsite fidelity is arguably adaptive since the nonrandom location of nests is important for egg and early larval survival rates, all critical components of recruitment rates (Rejwan et al. 1997). Following courtship display, males defend the eggs and fry. Even with good nest sites and paternal care, survival to a free-swimming larval stage ranges from 21–96% due to predators, fungal infection, storms, water level, and temperature fluctuations (Steinhart et al. 2005, Webster 1954). Consequently, there is an enormous discrepancy in male fitness, as larger males procure more matings (Wiegmann et al. 1992), support larger broods (Ridgway and Friesen 1992), defend nests more successfully (Philipp et al. 1997), and thereby contribute more offspring to the fall young-of-the-year class (Knotek and Orth 1998). In fact, Gross and Kapuscinski (1997) determined that only 5.4% of spawning males contribute to 54.7% of an average young-of-the-year class. This age-0 class remains within 200 m of the nest and overwinters nearby before dispersing from their natal area as juveniles (Ridgway et al. 2002). Lyons and Kanehl (2002) reviewed the migration patterns of native Smallmouth Bass noting that some lake individuals move into tributaries in the spring for spawning but subsequently return to the lake, a potamodromous migratory pattern described as lacustrine-adfluvial (Varley and Gresswell 1988). Such populations of Smallmouth Bass have been described from Cayuga Lake, NY (Webster 1954), eastern Lake Ontario– St. Lawrence River (Stone et al. 1954), Lake Simcoe, ON (Robbins and MacCrimmons 1977), and Lake Ontario, NY (Gerber and Haynes 1988). Thus, several groups of Smallmouth Bass may co-exist within a single lake system: lacustrine populations residing and spawning within the lake, potamodromous populations residing in the lake but spawning in tributaries, and resident tributary populations. Therefore, genetic variation in two regions of the mitochondrial genome (control region and cytochrome-b) was quantified and used to test this assertion with respect to lake and river groups. Two hypotheses were evaluated in the context of Smallmouth Bass behavior and migration patterns: (1) have riverine and lacustrine Smallmouth Bass diverged genetically, and (2) are Smallmouth Bass from adjacent rivers more closely related to each other than they are to lacustrine fish? Material and Methods Smallmouth Bass were collected from the central and western basins of Lake Erie and from four tributaries along the southcentral and southeastern shoreline by trawling, seining, or electroshocking (Fig. 1). Detailed sampling site locations and dates are given in the caption of Figure 1. Site names within Lake Erie (Fairport Harbor, Perry, Ashtabula, Conneaut, OH, and Van Buren Bay, NY; listed west to east) were delineated by the Ohio Division of Natural Resources, but largely represent the nearest city on the southern shore. Predominantly pre-spawning adults (>300 mm TL) were 2008 W.C. Borden 337 sampled from lacustrine sites, and they represent fish previously described in Borden and Stepien (2006) with the addition of one fish from “Ashtabula.” Lacustrine fish from the western basin of Lake Erie and Long Point Bay, ON, were excluded from this analysis because lacustrine fish sampled closest to the river mouths were assumed to be most relevant for contrasting potential lacustrine-riverine differences. The Chagrin, Cuyahoga, and Grand rivers (all in OH) were tributaries in the central basin of Lake Erie, and Cattaraugus Creek (NY) is a tributary in the eastern basin (Fig. 1). Juvenile and adult Figure 1. Collection sites of Smallmouth Bass from Lake Erie and four tributaries are indicated with a circle. Coastal collections (west to east) were made at Fairport Harbor, OH; Perry, OH; Ashtabula, OH; Conneaut, OH; and Van Buren Bay, NY. Dashed lines indicate watershed boundaries. Vertical bars on rivers indicate the location of falls or dams: Cuyahoga Falls on the Cuyahoga River, Chagrin Falls on the Chagrin River, and Harpersfield Dam on the Grand River. Inset: Great Lakes highlighting Lake Erie. Riverine sampling sites listed from mouth to headwater: Cattaraugus Creek (n total = 11): Erie County, NY, below bridge of state highways 5 and 20, 14–29 April 2003. Chagrin River (n total = 19): Cuyahoga County, OH, 500 feet south of Cedar Road off Chagrin River Road, 18 June 2003, (n = 11); Cuyahoga County, OH, mouth of Griswold Creek, 27 June 2005, (n = 4); Geauga County, OH, Fairmount Road and state highway 306, 27 June 2005, (n = 4). Cuyahoga River (n total = 10): Summit County, OH, at Brust Park in Munroe Falls, OH, 08 July 2003, (n = 5); Portage County, OH, at John Brown Tannery Park in Kent, OH, 08 July 2003, (n = 5). Grand River (n total = 12): Lake County, OH, mouth of Grand River, April 2005, (n = 1); Lake County, OH, 100 yards downstream of state highway 528 bridge, 18 June 2003, (n = 1); Ashtabula/Lake County line in OH, 14 September 2001, (n = 1); Ashtabula County, OH, 200 yards downstream of Harpersfield Bridge on old state highway 534, 18 June 2003, (n = 4); Ashtabula County, OH, Riverdale Bridge, 10 October 2005, (n = 3); Ashtabula County, OH, Rock Creek, tributary of Grand River near state highway 45, 14 September 2001, (n = 2). 338 Northeastern Naturalist Vol. 15, No. 3 Smallmouth Bass were sampled from the Cuyahoga River (n = 10, range = 92–385 mm total length [TL], median total length [η] = 169 mm), Chagrin River (n = 19, 35–235 mm TL, η = 100 mm), Grand River (n = 12, 60–447 mm TL, η = 107 mm), and Cattaraugus Creek (n = 11, 313–456 mm TL, η = 381 mm) spanning the months of April to October from 2001 to 2005. The median length of riverine fish was small, indicating a predominately juvenile collection. The only exception was Cattaraugus Creek, where all fish were of spawning size and age. Smallmouth Bass have not been stocked within Lake Erie in recent history (K. Kayle, Ohio Department of Natural Resources, Division of Wildlife [ODW], Fairport Harbor, OH, pers. comm.). The collection locations, dates, and median size of the fish sampled within the rivers lend confidence that stream-resident Smallmouth Bass were collected. Multiple collection sites within each river mitigated the effects of sampling the same “family” or cohort of Smallmouth Bass. Collection sites in the Chagrin, Cuyahoga, and Grand Rivers were located upstream of lowhead dams or natural falls. Summer sampling of river populations reduced the confounding effects of sampling lacustrine fish during either a spring or fall migration (Lyons and Kanehl 2002). The two exceptions were collections made in April near the mouth (less than 5 km from Lake Erie) of the Grand River (n = 1) and Cattaraugus Creek (n = 11). Despite attempts to mitigate the effects of non-random sampling, small sample sizes may not have removed the potential effects of random chance due to sampling error. Fin clips were taken and stored either on ice in the field and then at -80 °C in the laboratory or placed directly in 95% ethyl alcohol in the field. DNA extraction, amplification, and sequencing details, including haplotype designations, were described previously (Borden and Stepien 2006). Concatenated sequences (840 total bp) were constructed from 540 bp of the 3’ end of cytochrome-b and 300 bp of the 5’ end of the control region and analyzed as such. Mitochondrial DNA variation was described using several diversity indices. Population structure between riverine and lacustrine sites and among riverine sites was evaluated using statistical analyses of differentiation (Raymond and Rousset 1995), estimates of divergence (ΦST of Weir and Cockerham 1984), and average pairwise distances corrected for within population variation (dA = dxy – {dx + dy}/2 [Nei 1987]). In these analyses of population structure, lacustrine samples were pooled as a single lake collection based on the findings of Borden and Stepien (2006) and Stepien et al. (2007). An analysis of molecular variance (AMOVA; Excoffier et al. 1992) using haplotype pairwise differences was performed to partition genetic variation. The goal was to quantify the contribution from (1) lake versus river samples and (2) differences among sampling sites within a habitat (lake or river). All calculations were performed with Arlequin 3.11 (Excoffier et al. 2006), and p-values were evaluated using sequential Bonferroni correction (Rice 1989). 2008 W.C. Borden 339 A mitochondrial gene tree was constructed using maximum parsimony under a branch-and-bound search and neighbor-joining with mean uncorrected p-distances (PAUP*v.4.0b10; Swofford 1998). Phylogenetic analyses of intraspecific haplotypes violate assumptions of bifurcating lineages and the retention of ancestral haplotypes in the population (Posada and Crandall 2001). Consequently, a haplotype network was constructed using the median- joining (Bandelt et al. 1999) and post-processing Steiner algorithms (Polzin and Daneschmand 2003) in Network v. 4.1.1.2 (Rohl 2004). Nested-clade analysis (Templeton et al. 1995) was employed to differentiate population structure from population history. Haplotypes in the resulting network were nested using ANeCA (Automated Nested Clade Analysis; Panchal 2007, Panchal and Beaumont 2007) following the criteria outlined by Templeton et al. (1987), Templeton and Sing (1993), Crandall (1996), and Templeton (2002). The spatial arrangement among sampling sites was best described by linear distances along rivers and the shoreline of Lake Erie. Thus, a matrix of pairwise distances among the nine sampling sites and the nesting structure of haplotypes were tested for nonrandom associations between haplotypes and geography using Geodis v2.5 (Posada et al. 2000). Two summary statistics, Dc and Dn, describing the spatial relationships of haplotypes in a clade and among nested clades, respectively, were then interpreted using an inference key (14 July 2004) in ANeCA. Results Sixteen concatenated haplotypes were found in 91 Smallmouth Bass from four Lake Erie tributaries and adjacent lake populations (Table 1). From the 10 haplotypes present in multiple copies, one was restricted to Lake Erie, four were restricted to rivers, and five were shared between samples from the rivers and the lake. Cytochrome-b and control region sequences were submitted to Genbank (Table 1). Analyses on pooled data for fish collected in rivers and Lake Erie indicated that these populations have diverged (fixation index [ΦST = 0.14, P < 0.01], genetic distance [dA = 0.04, P = < 0.01], and population differentiation test [P < 0.01]). Of the four lake-to-river pairwise comparisons, the average ΦST ± SE was 0.26 ± 0.10 and the average pairwise distance was 0.65 ± 0.34 (Table 2). However, a lake-to-river partition (AMOVA) explained only 9.3% of the variation (FCT = 0.09, P = 0.14), in part because among-sampling-site variation within habitat type (lake or river) explained 19.4% of the variation (FSC = 0.21, P < 0.01). The largest source of variation occurred within sampling sites (71.4%, FST = 0.29, P < 0.01). Even fish from Cattaraugus Creek were divergent from fish in Lake Erie, suggesting that these two groups do not belong to the same population despite their close proximity. Samples from each river generally differed significantly from other rivers (Table 2). The greatest divergence occurred between the Chagrin and Grand Rivers, while the lowest divergence occurred between the Chagrin and Cuyahoga Rivers. The average river-to-river divergence (average ΦST = 340 Northeastern Naturalist Vol. 15, No. 3 0.27 ± 0.15) and average river-to-river genetic distance (dA = 0.89 ± 0.53) were similar in magnitude to pairwise lake-to-river differences. Haplotype divergences ranged from 0.1% to 0.6%, indicative of the relatively young age of this system (<14,000 y). These small divergences were reflected in a poorly resolved gene tree (not shown) and a cycle within the haplotype network (Fig. 2A). Yet, some clade divergences could be attributed to intrinsic and extrinsic population mechanisms. Nested-clade analysis identified three 2-step clades and six 1-step clades (Fig. 2B). Clade II-1 was predominantly composed of riverine haplotypes, lacustrine Table 2. Population divergence (ΦST above diagonal) and genetic distance (dA below diagonal) estimates among riverine and between riverine–lacustrine Smallmouth Bass based on concatenated mtDNA sequences of cytochrome-b and the control region. Negative values are converted to zero. “*” indicates statistical significance following sequential Bonferroni correction. Cuyahoga R. Chagrin R. Grand R. Cattaraugus C. Lake Erie Cuyahoga River <0.01 0.34* 0.20* 0.29* Chagrin River 0.00 0.45* 0.31* 0.38* Grand River 1.21* 1.56* 0.32* 0.18* Cattaraugus Creek 0.66* 0.92* 0.93* 0.19* Lake Erie 0.74* 1.08* 0.36* 0.41* Table 1. Distribution of mtDNA concatenated haplotypes and nested-clade delineations. Haplotype numbers of cytochrome b (Cytb) and control region (Ctrl) sequences are consistent with those in Borden and Stepien (2006). LE = Lake Erie; Cu = Cuyahoga River; Ch = Chagrin River; GR = Grand River; Ca = Cattaraugus Creek. Cytb GenBank Ctrl GenBank Clade Hap LE Cu Ch GR Ca n hap # accession hap # accession II-1 I-1 A 4 4 1 11 EU267709 B 1 1 2 10 EU267708 I-2 C 4 10 14 6 EU267711 3 D 4 3 1 8 1 3 DQ354377 E 2 2 7 EU267712 3 II-2 I-3 F 6 6 1 4 DQ354378 I-4 G 6 7 1 14 1 2 DQ354376 II-3 I-5 H 7 6 2 1 16 1 DQ354383 1 DQ354375 I 3 2 5 1 7 DQ354381 J 12 1 1 14 2 DQ354384 1 K 1 1 4 DQ354386 1 L 1 1 1 12 EU267710 M 1 1 8 EU267713 1 I-6 N 2 2 5 DQ354387 1 O 1 1 5 9 EU267707 P 1 1 5 5 DQ354379 Totals 39 10 19 12 11 91 h 0.83 0.82 0.63 0.67 0.87 SE < 0.01 0.03 0.02 0.04 0.03 2008 W.C. Borden 341 haplotypes dominated clade II-2, and clade II-3 was comprised of riverine and lacustrine haplotypes. Contiguous range expansion was the most-likely mechanism identified producing clade II-1, which itself was composed of a Figure 2. Relationships of 16 concatenated mtDNA haplotype as determined by (A) a haplotype network and (B) nested-clade analysis. Circles represent haplotypes identified by letters as in Table 1. (A) This network was produced from the median joining and Steiner algorithms in Network v. 4.1.1.2. Black fill indicates lacustrine origin of collection; white fill indicates river origin. Circle size indicates the relative proportion of each haplotype; the smallest circles are singletons. Haplotypes are separated by one base-pair change except haplotypes D and C, which are separated by two base changes both in the cytochrome-b gene. (B) One-step clades are identified by an “Iclade number” and enclosed in rounded boxes. Two-step clades are identified by a “II-clade number” and enclosed in dashed boxes. Clades I-2 and II-1 rejected the null hypothesis of a random association between geography and haplotype distribution. 342 Northeastern Naturalist Vol. 15, No. 3 western clade (I-2 in the Chagrin and Cuyahoga rivers or Lake Erie) and an eastern clade (I-1 in Cattaraugus Creek). Restricted gene flow with isolation by distance was the most likely mechanism identified producing clade I-2. While the remaining clades failed to reject the null hypothesis of a random association between the distribution of sampling sites and haplotypes, many of the clades were characterized by haplotypes restricted to, or dominated by, a single habitat. Discussion Evidence for discrete lacustrine and riverine populations Genetic evidence is consistent with the hypothesis that riverine and lacustrine Smallmouth Bass represent reproductively isolated groups, though contemporary gene flow between them cannot be ruled out. Riverine fish shared only 31% of the mtDNA haplotypes with lacustrine fish, and most private haplotypes occurred in riverine fish. While the lake/river partition of Smallmouth Bass accounts for only 9.3% of the total variation, inferences from F-statistics and genetic distances support a genetic distinction. The large variation among and within sampling sites, combined with modest sample sizes, may have contributed to the statistical non-significance and therefore limits conclusions that can be drawn in this study. However, the relatively weaker estimates of population divergence among geographically distant lacustrine sites within Lake Erie, as observed by Borden and Stepien (2006) and Stepien et al. (2007), suggest that the lake/river dichotomy is an important biological element shaping the genetic structure of Smallmouth Bass populations in the watershed. While distinct lacustrine populations can be associated with large bays (e.g., Sandusky Bay, OH; Long Point Bay, ON) or shoals and reefs, particularly among the islands in the western basin of Lake Erie (Borden and Stepien 2006, Kelso 1973, Stepien et al. 2007), each tributary appears to be characterized by its own divergent population of Smallmouth Bass. Most haplotype clades are indicative of either a river or lake habitat. Moreover, riverine clades (I-1, I-2, I-6, II-1) occur at the periphery of the haplotype network, while lacustrine-dominated clades are generally more centrally located, suggesting that riverine Smallmouth Bass were founded from lacustrine populations. As a consequence, restricted gene flow between Smallmouth Bass in the central basin of Lake Erie and those in the Cuyahoga and Chagrin Rivers is a possible mechanism explaining their divergence as indicated by the NCA (Fig. 2B, clade I-2). Though small sample sizes limit firm conclusions, the data are consistent with a colonization scenario of dispersal from ancestral waterways into the tributaries, followed by restricted gene flow, which ultimately resulted in divergent populations. Such lake/ river divergence of Smallmouth Bass could be maintained by seasonal migratory habits (Lyons and Kanehl 2002, Ridgway et al. 2002) and nest-site fidelity (Ridgway et al. 1991). Alternatively, the possibility that Smallmouth 2008 W.C. Borden 343 Bass colonizing Lake Erie and its tributaries originated from multiple and distinct gene pools is not addressed here and cannot be eliminated as a potential contributor to their genetic divergence. Differentiation of lacustrine and riverine fish populations is taxonomically and geographically widespread (Lucas and Baras 2001). For example, Trautman (1981) observed morphological and meristic differences in lacustrine and riverine populations of Etheostoma blennioides Miller (Greenside Darter) and E. nigrum eulepsis (Hubbs and Greene) (Johnny Darter) (Percidae) and Campostoma pullum (Agassiz) Central Stoneroller (Cyprinidae) within these same tributaries of Lake Erie. Even in a relatively young lake (≈14,000 years) such as Lake Erie, genetic evidence is consistent with reproductive segregation of lacustrine and riverine populations of Smallmouth Bass. A genetic evaluation of putative potamodromous populations (Gerber and Haynes 1988, Robbins and MacCrimmon 1977, Stone et al. 1954, Webster 1954) co-occurring with stream and lake populations of Smallmouth Bass would be worthy of future study. Evidence for divergence among tributary populations Perhaps as important as the apparent separation of riverine-lacustrine populations is that many of the relatively small Lake Erie tributaries seem to possess genetically distinct populations of Smallmouth Bass. Minimal levels of divergence among sites in Lake Erie (Borden and Stepien 2006, Stepien et al. 2007) suggest that the among-site variation (19.4%) is due largely to haplotype variation among the tributary populations. Divergence among the tributary populations is an equally important element that shapes the genetic structure of Smallmouth Bass populations across the Lake Erie watershed and it logically arises as a by-product of a river/ lake dichotomy. Among these proximate rivers, three apparent geographical patterns were observed: (1) the Chagrin and Cuyahoga River fish are characterized by a small genetic divergence; (2) this pairing differs from neighboring Grand River bass and more distant Cattaraugus Creek bass; and (3) the Grand River population, while significantly divergent from bass in Lake Erie, are more similar to lake populations than they are to bass from other tributaries. The glacial history provides clues that are consistent with these patterns. The Cuyahoga River bounds the Chagrin River on three sides (Fig. 1) and both rivers possess naturally occurring falls and rapids that currently limit upstream migration of fishes. Chagrin Falls is a naturally occurring 6-m waterfall, and Cuyahoga Falls is a series of rapids and falls through which the river drops 67 m; both of these natural barriers have been supplemented with small dams. A small genetic divergence between bass in these rivers is consistent with recent mixing of headwater populations, which may have been facilitated by river capture of the headwaters during isostatic rebound (Bishop 1995). 344 Northeastern Naturalist Vol. 15, No. 3 The Grand River may have had a more recent and extended connection with the lacustrine habitat of primordial Lake Erie not shared by the Chagrin and Cuyahoga Rivers. During the last ice advance in the late Woodfordian period (≈15,000 years ago), two glacial lakes (Rock Creek Lake and Grand River Lake) occupied the present day Grand River valley (White 1982, White and Totten 1979). This extended lacustrine connection with the lowland Grand River may explain the lower divergence between the current Smallmouth Bass in the Grand River and Lake Erie. According to results from the nested-clade analysis, the apparent divergence of bass in Cattaraugus Creek may be due to contiguous range expansion. Consistent with this interpretation is that the Wisconsinin sheet retreated from the Erie basin in a southwest to northeast direction, possibly allowing Smallmouth Bass from the central basin to disperse more than 200 km along the southern shore in an easterly direction toward Cattaraugus Creek. This scenario is supported by an eastern clade (I-1) restricted to Cattaraugus Creek that is derived from a western clade (I-2) found in the central basin of Lake Erie and the adjacent Chagrin and Cuyahoga Rivers. While results indicate that river populations may be distinct from one another, specific relationships among river populations remain uncertain. For example, the small mtDNA divergence of Smallmouth Bass between the Cuyahoga and Chagrin Rivers is contradicted by a very large and statistically significant nDNA microsatellite divergence (Stepien et al. 2007). However, discordance between mitochondrial and nuclear markers is not uncommon (e.g., Brown et al. 2005, Canestrelli et al. 2007). The distribution of genetic variation within a single river was not evaluated, but resident-stream Smallmouth Bass possess variable migratory behaviors within and among rivers (Lyons and Kanehl 2002, Paragamian and Coble 1975, Reynolds 1965), which may have contributed to this discrepancy. Small samples sizes might also have played a role in their discordance, necessitating that conclusions drawn from these results should be tempered. As anticipated for other aquatic organisms in these rivers, Chagrin and Grand River populations of Allocapnia recta (Claassen) (winter stonefly) are also genetically divergent (Yasick et al. 2007), potentially indicating a shared genetic response by diverse organsimal groups to common geographical events. Regardless of their interrelationships, all riverine populations appear to have diverged from each other, even those within close geographical proximity. The small geographical scale of the four tributaries and modest sample sizes from those populations soften the results of this study. However, the suggestions that riverine populations are divergent, and distinct from lacustrine Smallmouth Bass, warrant a more thorough evaluation of their genetic variation. Similarly, future research to assess the presence of potamodrous Smallmouth Bass populations in this system would be well served. 2008 W.C. Borden 345 Acknowledgments This study was made possible only through the generosity of numerous field hands. They include D. Einhouse and staff (NY State Department of Environment and Conservation); K. Kayle, J. Deller, C. Knight, and T. Bader (ODW); B. Zawiski and S. Taylor (Ohio Environmental Protection Agency); M. Coburn and L. Kousa (John Carroll University); J. Giboney, M. Jedlicka, and T. Marth (National Science Foundation sponsored Research Experiences for Undergraduates program, DBI 0243878 to B. Walton and C. Stepien, and Cleveland State University [CSU] President’s Fellowships program); A. Ford, R. Krebs, O. Lockhart (CSU); and C. Stepien (University of Toledo). M. Blum sequenced samples at the CSU DNA Facility. The Ohio Sea Grant (Project Number R/LR-5 to C. Stepien), as well as a CSU Doctoral Dissertation Research Expense Award and a research grant from the CSU DNA Analysis Facility, both to C. Borden, funded this project. Page charges were offset by the College of Science. The manuscript benefited from critiques by M. Coburn, P. Doerder, C. Stepien, and in particular R. Krebs, who performed yeoman’s work. E. Carson and two anonymous reviewers provided many valuable and constructive comments. Literature Cited Bandelt, H.–J., P. Forster, and A. Röhl. 1999. Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution 16:37–48. Bishop, P. 1995. Drainage rearrangement by river capture, beheading, and diversion. Progress in Physcial Geography 19:449–473. Borden, W.C., and C.A. Stepien. 2006. Discordant population genetic structuring of Smallmouth Bass, Micropterus dolomieu Lacepède, in Lake Erie based on mitochondrial DNA sequences and nuclear DNA microsatellites. Journal of Great Lakes Research 32:242–257. Brown, K.M., G.A. Baltazar, and M.B. Hamilton. 2005. Reconciling nuclear microsatellite and mitochondrial marker estimates of population structure: Breeding-population structure of Chesapeake Bay Striped Bass (Morone saxatilis). Heredity 94:606–615. Bunt, C.M., S.J. Cooke, and D.P. Philipp. 2002. Mobility of riverine Smallmouth Bass related to tournament displacement and seasonal habitat use. Pp. 545–552, In D.P. Philipp and M.S. Ridgway (Eds.). Black Bass: Ecology, Conservation, and Management. American Fisheries Society, Symposium 31, Bethesda, MD. 724 pp. Canestrelli, D., A. Verardi, and G. Nascetti. 2007 Genetic differentiation and history of populations of the Italian Treefrog Hyla intermedia: Lack of concordance between mitochondrial and nuclear markers. Genetica 130:241–255. Crandall, K.A. 1996. Multiple interspecies transmissions of human and simian T-cell leukaemia/lymphoma virus type I sequences. Molecular Biology and Evolution 13:115–131. Excoffier, L., P.E. Smouse, and J.M. Quattro. 1992. Analysis of molecular variance inferred from metric distances among DNA haplotypes: Application to human mitochondrial DNA restriction data. Genetics 131:479–491. Excoffier, L., G. Laval, and S. Schneider. 2006. Arlequin ver. 3.1: An integrated software package for population genetics data analysis. Evolutionary Bioinformatics Online 1:47–50. 346 Northeastern Naturalist Vol. 15, No. 3 Gerber, G.P., and J.M. Haynes. 1988. Movements and behavior of Smallmouth Bass, Micropterus dolomieu, and Rock Bass, Ambloplites rupestris, in southcentral Lake Ontario and two tributaries. Journal of Freshwater Ecology 4:425–440. Gross, M.L., and Kapuscinski, A.R. 1997. Reproductive success of Smallmouth Bass estimated and evaluated from family–specific DNA fingerprints. Ecology 78:1424–1430. Gross, M.L., A.R. Kapuscinski, and A.J. Faras. 1994. Nest-specific DNA fingerprints of Smallmouth Bass in Lake Opeongo, Ontario. Transactions of the American Fisheries Society 123:449–459. Kelso, R.M. 1973. Movement of Yellow Perch (Perca flavescens), Smallmouth Bass (Micropterus dolomieu) and White Bass (Morone chrysops) released in Long Point Bay, Lake Erie, during 1971 and 1972. Technical Report No. 386. Fisheries Research Board of Canada. 29 pp. Knotek, W.L., and D.J. Orth. 1998. Survival for specific life intervals of Smallmouth Bass, Micropterus dolomieu, during parental care. Environmental Biology of Fishes 51:285–296. Lucas, M.C., and E. Baras. 2001. Migration of Freshwater Fishes. Blackwell Science, Oxford, UK. Lyons, J., and P. Kanehl. 2002. Seasonal movements of Smallmouth Bass in streams. Pp. 149–160, In D.P. Philipp and M.S. Ridgway (Eds.). Black Bass: Ecology, Conservation, and Management. American Fisheries Society, Symposium 31, Bethesda, MD. 724 pp. Mandrak, N.E., and E.J. Crossman. 1992. Postglacial dispersal of freshwater fishes into Ontario. Canadian Journal of Zoology 70:2247–2259. Nei, M. 1987. Molecular Evolutionary Genetics. Columbia University Press, New York, NY. 512 pp. Panchal, M. 2007. The automation of nested-clade phylogenetic analysis. Bioinformatics 23:509–510. Panchal, M., and M.A. Beaumont. 2007. The automation and evaluation of nestedclade phylogeographic analysis. Evolution 61:1466–1480. Paragamian, V.L., and D.W. Coble. 1975. Vital statistics of Smallmouth Bass in two Wisconsin Rivers, and other waters. Journal of Wildlife Management 39:201–210. Philipp, D.P., C.A. Toline, M.F. Kubacki, D.B.F. Philipp, and F.J.S. Phelan. 1997. The impact of catch-and-release angling on the reproductive success of Smallmouth Bass and Largemouth Bass. North American Journal of Fisheries Management 17:557–567. Polzin, T., and S.V. Daneschmand. 2003. On Steiner trees and minimum spanning trees in hypergraphs. Operations Research Letters 31:12–20. Posada, D., and K.A. Crandall. 2001. Intraspecific gene genealogies: Trees grafting into networks. Trends in Ecology and Evolution 16:37–45. Posada, D., K.A. Crandall, and A.R. Templeton. 2000. GeoDis: A program for the cladistic nested analysis of the geographical distribution of genetic haplotypes. Molecular Ecology 9:487–488. Raymond, M., and F. Rousset. 1995. An exact test for population differentiation. Evolution 49:1280–1283. 2008 W.C. Borden 347 Rejwan, C., B.J. Shuter, M.S. Ridgway, and N.C. Shuter. 1997. Spatial and temporal distributions of Smallmouth Bass (Micropterus dolomieu) nests in Lake Opeongo, Ontario. Canadian Journal of Fisheries and Aquatic Sciences 54:2007–2013. Reynolds, J.B. 1965. Life history of Smallmouth Bass, Micropterus dolomieu Lacepede, in the Des Moines River, Boone County, Iowa. Iowa State Journal of Science 39:417–436. Rice, W.R. 1989. Analyzing tables of statistical tests. Evolution 43:223–225. Ridgway, M.S., B.J. Shuter, and E.E. Post. 1991. The relative influence of body size and territorial behaviour on nesting asynchrony in male Smallmouth Bass, Micropterus dolomieu (Pisces: Centrarchidae). Journal of Animal Ecology 60:665–681. Ridgway, M.S., B.J. Shuter, T.A. Middel, and M.L. Gross. 2002. Spatial ecology and density-dependent processes in Smallmouth Bass: The juvenile transition hypothesis. Pp. 47–60, In D.P. Philipp and M.S. Ridgway (Eds.). Black Bass: Ecology, Conservation, and Management. American Fisheries Society, Symposium 31, Bethesda, MD. 724 pp. Ridgway, M.S., and T.G. Friesen. 1992. Annual variation in parental care in Smallmouth Bass, Micropterus dolomieu. Environmental Biology of Fishes 35:243–255. Robbins, W.H., and H.R. MacCrimmon. 1977. Vital statistics and migratory patterns of a potamodromous stock of Smallmouth Bass, Micropterus dolomieu. Journal of the Fisheries Research Board of Canada 34:142–147. Rohl, A. 2004. Network v. 4.1.1.2. (Shareware Phylogenetic Network Software). Presented by Fluxus Technology Ltd. Available online at www.fluxus-engineering. com. Accessed January 2006. Scott, W.B., and E.J. Crossman. 1973. Freshwater Fishes of Canada. Bulletin 184. Fisheries Research Board of Canada, Ottawa, Canada. 966 pp. Sowa, S.P., and C.F. Rabeni. 1995. Regional evaluation of the relation of habitat to distribution and abundance of Smallmouth Bass and Largemouth Bass in Missouri streams. Transactions of the American Fisheries Society 124:240–251. Steinhart, G.B., N.J. Leonard, R.A. Stein, and E.A. Marschall. 2005. Effects of storms, angling, and nest predation during angling on Smallmouth Bass (Micropterus dolomieu) nest success. Canadian Journal of Fisheries and Aquatic Sciences 62:2649–2660. Stepien, C.A., D.J. Murphy, and R.M. Strange. 2007. Broad- to fine-scale population genetic patterning in the Smallmouth Bass Micropterus dolomieu across the Laurentian Great Lakes and beyond: An interplay of behaviour and geography. Molecular Ecology 16:1605–1624. Stone, U.B., D.G. Pasko, and R.M. Roecker. 1954. A study of Lake Ontario–St. Lawrence River Smallmouth Bass. New York Fish and Game Journal 1:1–26. Swofford, D.L. 1998. PAUP*v4b10. Phylogenetic Analysis Using Parsimony (*and Other Methods). Sinauer Associates, Sunderland, MA. Templeton, A.R. 2002. Out of Africa again and again. Nature 416:45–51. Templeton, A.R., and C.F. Sing. 1993. A cladistic analysis of phenotypic associations with haplotypes inferred from restriction endonuclease mapping. 4. Nested analyses with cladogram uncertainty and recombination. Genetics 134:659–669. 348 Northeastern Naturalist Vol. 15, No. 3 Templeton, A.R., E. Boerwinkle, and C.F. Sing. 1987. A cladistic analysis of phenotypic associations with haplotypes inferred from restriction endonuclease mapping. 1. Basic theory and an analysis of alcohol dehydrogenase activity in Drosophila. Genetics 117:343–351. Templeton, A.R., E. Routman, and C.A. Phillips. 1995. Separating population structure from population history: A cladistic analysis of the geographical distribution of mitochondrial DNA haplotypes in the Tiger Salamander, Ambystoma tigrinum. Genetics 140:767–782. Trautman, M.B. 1981. The Fishes of Ohio. Ohio State University Press, Columbus, OH. 782 pp. Varley, J.D., and R.E. Gresswell. 1988. Ecology, status, and management of the Yellowstone Cutthroat Trout. American Fisheries Society Symposium 4:13–24. Webster, D.A. 1954. Smallmouth Bass, Micropterus dolomieui, in Cayuga Lake. Part I. Life history and environment. Cornell Experiment Station Memoir 327:1–39. Weir, B.S., and C.C. Cockerham. 1984. Estimating F-statistics for the analysis of population structure. Evolution 38:1358–1370. White, G.W. 1982. Glacial geology of northeastern Ohio (including a chapter on Pleistocene beaches and strandlines by S.M. Totten). Ohio Division of Geological Survey Bulletin 68. Ohio Department of Natural Resources. Columbus, OH. White, G.W., and S.M. Totten. 1979. Glacial geology of Ashtabula County, Ohio. Ohio Division of Geological Survey Report of Investigations 112. Ohio Department of Natural Resources. Columbus, OH. Wiegmann, D.D., J.R. Baylis, and M.H. Hoff. 1992. Sexual selection and fitness variation in a population of Smallmouth Bass (Micropterus dolomieu). Evolution 46:1740–1753. Yasick, A.L., R.A. Krebs, and J.A. Wolin. 2007. The effect of dispersal ability in winter and summer stoneflies on their genetic differentiation. Ecological Entomology 32:399–404.