ArticlePDF Available

Population genetic structure of three freshwater mussel (Unionidae) species within a small stream system: Significant variation at local spatial scales

Authors:

Abstract and Figures

1. Unionid mussels are highly threatened, but little is known about genetic structure in populations of these organisms. We used allozyme electrophoresis to examine partitioning of genetic variation in three locally abundant and widely distributed species of mussels from a catchment in Ohio. 2. Within-population variation was similar to that previously reported for freshwater mussels, but genotype frequencies exhibited heterozygote deficiencies in many instances. All three species exhibited significant among-population variation. Evidence of isolation-by-distance was found in Elliptio dilatata and Ptychobranchus fasciolaris, while Lampsilis siliquoidea showed no geographical pattern of among-population variation. 3. Our results suggest that the isolating effects of genetic drift were greater in L. siliquoidea than in the other species. Differentiation of populations occurred at a much smaller spatial scale than has previously been found in freshwater mussels. Differences among species may reflect differences in the dispersal abilities of fishes that serve as hosts for the glochidia larvae of mussels. 4. Based on our results, we hypothesise that species of mussels that are common to large rivers exhibit relatively large amounts of within-population genetic variation and little differentiation over large geographical distances. Conversely, species typical of small streams show lower within-population genetic variation and populations will be more isolated. If this hypothesis can be supported, it may prove useful in the design of conservation strategies that maintain the genetic structure of target species.
Content may be subject to copyright.
Population genetic structure of three freshwater mussel
(Unionidae) species within a small stream system:
significant variation at local spatial scales
DAVID J. BERG*, ALAN D. CHRISTIAN
AND SHELDON I. GUTTMAN
*Department of Zoology, Miami University, Hamilton, OH, U.S.A.
Department of Zoology, Miami University, Oxford, OH, U.S.A.
SUMMARY
1. Unionid mussels are highly threatened, but little is known about genetic structure in
populations of these organisms. We used allozyme electrophoresis to examine partitioning
of genetic variation in three locally abundant and widely distributed species of mussels
from a catchment in Ohio.
2. Within-population variation was similar to that previously reported for freshwater
mussels, but genotype frequencies exhibited heterozygote deficiencies in many instances.
All three species exhibited significant among-population variation. Evidence of isolation-
by-distance was found in Elliptio dilatata and Ptychobranchus fasciolaris, while Lampsilis
siliquoidea showed no geographical pattern of among-population variation.
3. Our results suggest that the isolating effects of genetic drift were greater in L. siliquoidea
than in the other species. Differentiation of populations occurred at a much smaller spatial
scale than has previously been found in freshwater mussels. Differences among species
may reflect differences in the dispersal abilities of fishes that serve as hosts for the
glochidia larvae of mussels.
4. Based on our results, we hypothesise that species of mussels that are common to large
rivers exhibit relatively large amounts of within-population genetic variation and little
differentiation over large geographical distances. Conversely, species typical of small
streams show lower within-population genetic variation and populations will be more
isolated. If this hypothesis can be supported, it may prove useful in the design of
conservation strategies that maintain
the genetic structure of target species.
Keywords: allozyme, dispersal, Elliptio dilatata,Lampsilis siliquoidea,Ptychobranchus fasciolaris
Introduction
Maintenance of genetic diversity has been recognised
as a key component in minimising extinction proba-
bilities of populations (Saccheri et al., 1998) and
ensuring survival of threatened species (Spielman,
Brook & Frankham, 2004). This genetic diversity is
partitioned into two components. Diversity within
populations is reflected in factors such as selection,
inbreeding probability, genetic drift and the loss or
maintenance of rare alleles. Diversity among popula-
tions reflects the geographical structuring of popula-
tions, with variation among populations due to local
selection or adaptation, and to reduced gene flow or
isolation. Both these components of total genetic
variation must be quantified in order to develop
Correspondence: David J. Berg, Department of Zoology,
Miami University, 1601 University Boulevard,
Hamilton, OH 45011, U.S.A.
E-mail: bergdj@muohio.edu
Present address: Alan D. Christian, Department of Biological
Sciences, Arkansas State University, State University,
AR 72467, U.S.A.
Freshwater Biology (2007) 52, 1427–1439 doi:10.1111/j.1365-2427.2007.01756.x
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd 1427
strategies for adequate maintenance of genetic varia-
tion in species of conservation interest.
Freshwater mussels of the families Unionidae and
Margaritiferidae (Unionoidea) are among the largest
benthic invertebrates in freshwater ecosystems, and
they can make up a significant proportion of total
benthic biomass (Strayer et al., 2004). As a result, they
can be key elements in energy flow and nutrient
cycling within freshwater ecosystems. Large mussel
beds with densities greater than 100 individuals per
m
2
and total numbers of individuals greater than
440 000 have been recently reported in North Ameri-
can streams and rivers (Christian, 1995; Harris &
Christian, 2000). Thus, these organisms are often key
components of freshwater benthic communities, but
they are also among the most threatened faunal
elements in North America (Lydeard et al., 2004).
In fact, a large proportion of North American
freshwater mussels currently have some form of
conservation status at national or regional levels.
Declining abundance and richness of freshwater
mussels are due to combinations of habitat modifica-
tion (impoundment of rivers, alterations in land use),
pollution, commercial harvesting and the relatively
recent invasion of North America by dreissenid
mussels (Bogan, 1993). While over 70%of North
American mussel taxa are of conservation concern
(Williams et al., 1993), there is little evidence that any
mussel taxa are currently stable. Even relatively
common species have shown declines in abundance
over time (Anthony & Downing, 2001). In this sense,
all species of freshwater mussels appear to be at
elevated risk of extinction. The precipitous decline of
North America’s freshwater mussel fauna has attrac-
ted the attention of researchers and management
agencies, but effective conservation planning is still
hampered by lack of basic biological knowledge about
these organisms.
Little is known of the population genetic structure
of freshwater mussels. While a number of studies
have described genetic variation in these organisms
(Mulvey et al., 1997; Berg et al., 1998; Berg & Berg,
2000; Curole, Foltz & Brown, 2004), most have focused
on species that are already highly endangered, and in
which much of the genetic structure has presumably
been lost due to the small size of individual popula-
tions and/or the greatly reduced number of popula-
tions remaining (for examples, see Stiven &
Alderman, 1992; Machordom et al., 2003). Because
we maintain that all freshwater mussels are at
elevated risk of extinction, we believe it is important
to study species that remain relatively common across
most of their ranges and hence might still retain their
original genetic structuring.
In this paper, we use allozyme electrophoresis to
describe genetic structure of three of the most abun-
dant species of freshwater mussels in the Darby Creek
ecosystem of central Ohio. Darby Creek is a small
tributary of the Scioto River, which itself is a major
tributary of the Ohio River. All three of these species
range widely throughout the Mississippi River basin
(Parmalee & Bogan, 1998). We quantify within-
population and among-population genetic variation
of these species at multiple sites in the Darby system
and compare results among species. Based on an
earlier population genetic study of the unionid
Quadrula quadrula (Rafinesque, 1820) which found
almost no geographical pattern in allozyme variation
along approximately 1000 km of the Ohio River (Berg
et al., 1998), we predicted that species along the
<200 km of Darby Creek that we sampled would
show little or no geographical structure. By combining
our results quantifying genetic variation of small-
stream populations with the earlier results from the
large Ohio River, we develop a hypothesis to explain
genetic structure of freshwater mussel species based
on river size and the associated fish faunas.
Methods
Study areas, sample collection and electrophoresis
Freshwater mussels were collected from 15 sites on
Big Darby Creek and its largest tributary, Little Darby
Creek, in central Ohio, from 1994 to 1998. The Darby
Creek ecosystem is of considerable conservation
interest because of its diversity of freshwater mussels
(41 species reported; Watters, 1996) and fishes (>90
species; http://bigdarby.org). These numbers are
remarkable for small Midwestern streams, and have
earned the Darby Creek ecosystem a listing as one of
The Nature Conservancy’s Last Great Places. How-
ever, this outstanding aquatic resource is being
threatened by land use changes due to suburban
sprawl associated with the city of Columbus (Watters,
1996).
Mussels were collected from a 60-km stretch of Big
Darby Creek in Union and Madison counties up-
1428 D.J. Berg et al.
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
stream of the confluence with Little Darby Creek,
from a 28-km stretch of Little Darby Creek in Madison
County, and from a single site below the confluence of
these streams in Franklin County (Fig. 1). Elliptio
dilatata (Rafinesque) were collected from five sites on
Big Darby Creek (river kilometres 92.1, 103.9, 112.9,
116.5, 123.1) and five sites on Little Darby Creek (river
kilometres 12.4, 19.4, 28.1, 31.9, 40.5). Lampsilis siliq-
uoidea (Barnes) were collected at three sites on Big
Darby Creek (river kilometres 63.5, 100.8, 106.0) and
two sites on Little Darby Creek (river kilometres 12.4,
19.4). Ptychobranchus fasciolaris (Rafinesque) were col-
lected from three sites on Big Darby Creek (river
kilometres 53.2, 100.8, 106.0) and two sites on Little
Darby Creek (river kilometres 12.4, 31.8).
At each site, 20–50 individuals per species were
collected by snorkeling in water <1.5 m deep. Most
mussels were non-destructively sampled using a
mantle biopsy (Berg et al., 1995), while a small
number of individuals (<5 per site) were destructively
sampled. Tissues were flash-frozen in liquid nitrogen
and stored at )80 C until analysed. Allozyme
electrophoresis using starch and cellulose acetate gels
was performed for a variety of enzyme systems using
standard recipes (Harris & Hopkinson, 1976; Hebert &
Beaton, 1989) and buffer systems (Selander et al., 1971;
Clayton & Tretiak, 1972; Hebert & Beaton, 1989) with
modifications. Fifteen loci were resolved for E. dilatata,
12 for L. siliquoidea and 14 for P. fasciolaris (Table 1).
Tissue samples were homogenised in 28 lLof2%
2-phenoxyethanol as a grinding buffer and centri-
fuged at 14 000 g. Supernatant liquids were then
loaded onto filter paper wicks (for starch gels) or
cellulose acetate plates. Gels were run between 40 min
and 5 h, depending on the gel and buffer systems
used, with starch gels always running much longer
than cellulose acetate gels. Alleles were identified by
assigning a 1 to the allele that migrated furthest
anodally (the fastest allele) and 2, 3, etc. to the second
fastest, third fastest, etc. When two loci were present,
Fig. 1 Map of the Darby Creek drainage system (a tributary of
the Scioto River, itself a tributary of the Ohio River) in Ohio,
showing approximate locations of sample sites as river kilo-
meters from the mouths of Big Darby (BD) and Little Darby (LD)
creeks.
Table 1 Enzyme systems and buffers used for allozyme analysis
of freshwater mussels from the Darby Creek system
System
Enzyme
number
No.
loci Gel type Buffer Species
LAP 1 Cellulose acetate CAM E, L, P
CK 2.7.3.2 1 Cellulose acetate CAM E
MDH 1.1.1.37 2 Cellulose acetate TEB E, P
2 Starch TC 6.7 L
GPI 5.3.1.9 1 Cellulose acetate TEB E
1 Cellulose acetate TG L, P
PGM 2.7.5.1 2 Cellulose acetate TEB E
2 Cellulose acetate TEM L, P
PEP* 3.4.11 1 Cellulose acetate TEB E
1 Cellulose acetate TG P
ENOL 4.2.1.11 1 Cellulose acetate TEM E
MPI 5.3.1.8 2 Cellulose acetate TEB E
1 Cellulose acetate TEB L, P
SOD 1.15.1.1 1 Cellulose acetate TG E, P
EST 3.1.1.1 2 Cellulose acetate TG E
2 Cellulose acetate TEM L, P
HEX 2.7.1.1 1 Cellulose acetate TG E
1 Cellulose acetate TEB L, P
AAT 2.6.1.1 1 Starch TC 6.7 L
1 Cellulose acetate TEB P
IDH 1.1.1.42 1 Starch TC 6.7 L, P
E¼Elliptio dilatata,L¼Lampsilis siliquoidea,P¼Ptychobranchus
fasciolaris.
*Leucine-valine as dipeptide.
Freshwater mussel population genetics 1429
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
the fastest (most anodal) was designated 1. The
number of loci analysed was limited by the small
amounts of tissue available for many individuals. At
least 20 individuals were analysed for all loci for each
population; additional individuals were analysed for
each variable locus. Loci were considered ‘polymor-
phic’ when the most common allele was present at
a frequency £0.95. Because the amount of tissue
available from each individual was often limited and
occasionally gels did not produce results, sample size
varied from 9 to 50 for each locus-by-population
combination (Tables 2–4). Sample size per locus
across all population-by-locus combinations averaged
36.4 individuals (±0.7 SE, n¼280 combinations).
Table 2 Allele frequencies at variable loci for Elliptio dilatata from Big and Little Darby Creeks. Population codes follow Figure 1
Locus Allele
Populations
BD 92.1 BD 103.9 BD 112.9 BD 116.5 BD 123.1 LD 12.4 LD 19.4 LD 28.1 LD 31.9 LD 40.5
LAP n ¼48* n¼49* n¼40* n¼18 n¼48* n¼50 n¼47 n¼39 n¼28* n¼50*
1 0.219 0.296 0.288 0.250 0.208 0.120 0.085 0.103 0.143 0.180
2 0.292 0.184 0.250 0.333 0.219 0.150 0.117 0.205 0.054 0.160
3 0.438 0.388 0.300 0.306 0.313 0.620 0.277 0.590 0.679 0.490
4 0.052 0.133 0.162 0.111 0.260 0.110 0.521 0.103 0.125 0.170
CK n ¼50 n¼50 n¼50 n¼47 n¼50 n¼50 n¼49 n¼15 n¼28 n¼50
1 0.033
2 1.000 1.000 1.000 1.000 1.000 1.000 1.000 0.967 1.000 1.000
MDH-1 n ¼49 n¼50 n¼50 n¼37 n¼50 n¼21 n¼49 n¼25 n¼28 n¼49
1 0.010 0.024 0.082 0.054 0.071
2 0.796 0.800 0.770 0.851 0.750 0.810 0.878 0.920 0.893 0.867
3 0.204 0.200 0.220 0.149 0.250 0.167 0.041 0.080 0.054 0.061
GPI n ¼50 n¼50 n¼50* n¼47 n¼50 n¼50 n¼49 n¼40 n¼28 n¼50
1 0.980 1.000 0.820 1.000 1.000 1.000 1.000 1.000 1.000 1.000
2 0.020 0.180
PGM-1 n ¼39 n¼39 n¼49 n¼22 n¼37 n¼23 n¼49* n¼39 n¼28 n¼40
1 0.031
2 0.923 0.808 0.898 0.909 0.878 0.696 0.837 0.962 0.982 0.963
3 0.077 0.192 0.102 0.091 0.122 0.304 0.133 0.039 0.018 0.038
PGM-2 n ¼50* n¼50* n¼49 n¼47 n¼50 n¼45 n¼49 n¼40 n¼28 n¼40
1 0.400 0.290 0.429 0.394 0.340 0.133 0.143 0.025 0.054 0.063
2 0.180 0.100 0.031 0.032 0.080 0.133 0.020 0.013 0.071 0.038
3 0.420 0.610 0.541 0.575 0.580 0.733 0.827 0.963 0.875 0.900
4 0.010
PEP n ¼34 n¼50* n¼35 n¼22 n¼40 n¼33 n¼46 n¼35 n¼28 n¼44*
1 0.941 0.790 0.957 0.886 0.950 0.909 0.663 0.829 0.839 0.568
2 0.059 0.210 0.043 0.114 0.050 0.091 0.337 0.171 0.161 0.432
SOD n ¼41 n¼44 n¼29 n¼34 n¼30 n¼50 n¼48 n¼20 n¼20 n¼50
1 1.000 1.000 1.000 1.000 0.967 1.000 1.000 1.000 1.000 1.000
2 0.033
EST-1 n ¼49 n¼50 n¼50 n¼17 n¼50 n¼50 n¼49 n¼25 n¼28 n¼50
1 1.000 1.000 1.000 1.000 1.000 1.000 1.000 0.940 1.000 1.000
2 0.060
Mean sample
size per locus
41.9 (2.8) 46.1 (1.9) 46.8 (1.7) 34.5 (3.0) 47.0 (1.6) 44.1 (2.6) 48.3 (0.2) 27.3 (2.8) 26.8 (0.8) 45.2 (2.4)
Mean number
of alleles per
locus
1.6 (0.2) 1.5 (0.2) 1.7 (0.3) 1.5 (0.2) 1.6 (0.2) 1.6 (0.3) 1.7 (0.3) 1.7 (0.2) 1.6 (0.3) 1.6 (0.3)
%P(95%
criterion)
33.3 33.3 33.3 33.3 33.3 33.3 33.3 26.7 26.7 26.7
Mean
heterozygosity
0.090 0.10 0.11 0.11 0.12 0.12 0.10 0.08 0.08 0.08
Numbers in parentheses are standard errors. n¼number of individuals.
*Locus-by-population combinations with heterozygote deficiencies.
1430 D.J. Berg et al.
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
Table 3 Allele frequencies at variable loci for Lampsilis siliquo-
idea from Big and Little Darby Creeks. Population codes follow
Fig. 1
Locus Allele
Populations
BD 63.5 BD 100.8 BD 106.0 LD 12.4 LD 19.4
LAP n ¼27* n¼37** n¼28** n¼42** n¼20
1 0.333 0.054 0.518 0.357 0.175
2 0.333 0.378 0.333 0.550
3 0.259 0.162 0.482 0.298 0.050
4 0.074 0.297 0.012 0.200
5 0.108 0.025
MDH-1 n ¼33 n¼37 n¼33 n¼45 n¼20
1 1.000 1.000 1.000 0.978 1.000
2 0.022
GPI n ¼34 n¼37 n¼33 n¼45 n¼20
1 0.118 0.041 0.152 0.133 0.325
2 0.882 0.960 0.849 0.867 0.675
PGM-1 n ¼30* n¼35* n¼31* n¼31* n¼18
1 0.283 0.471 0.484 0.613 0.417
2 0.183 0.429 0.435 0.323 0.444
3 0.433 0.100 0.081 0.065 0.111
4 0.100 0.028
PGM-2 n ¼32* n¼36 n¼34* n¼45* n¼20
1 0.078 0.153 0.118 0.056 0.050
2 0.781 0.847 0.809 0.900 0.950
3 0.141 0.074 0.044
MPI n ¼34* n¼24 n¼33 n¼45 n¼13
1 0.941 1.000 1.000 1.000 1.000
2 0.059
EST-1 n ¼28* n¼37 n¼25 n¼36* n¼20
1 0.071 0.054 0.100 0.056 0.050
2 0.500 0.608 0.740 0.458 0.600
3 0.429 0.311 0.120 0.458 0.350
4 0.027 0.040 0.028
HEX n ¼26 n¼26* n¼27 n¼45 n¼10
1 0.962 0.769 1.000 1.000 0.050
2 0.039 0.231 0.950
AAT n ¼15 n¼9n¼33 n¼45* n¼20
1 1.000 1.000 1.000 0.600 1.000
2 0.400
IDH n ¼33 n¼37 n¼33 n¼27 n¼20
1 1.000 1.000 0.985 1.000 1.000
2 0.015
Mean sample
size per locus
30.0 (1.6) 32.4 (2.5) 31.3 (0.9) 41.3 (1.8) 18.4 (1.0)
Mean number
of alleles per
locus
2.1 (0.3) 2.0 (0.4) 1.8 (0.3) 2.1 (0.3) 2.0 (0.4)
%P(95%cri-
terion)
50.0 41.7 41.7 50.0 50.0
Mean hetero-
zygosity
0.11 0.16 0.17 0.12 0.16
Numbers in parentheses are standard errors. n¼number of
individuals.
*Locus-by-population combinations with heterozygote defici-
encies.
**Locus-by-population combinations with heterozygote excesses.
Table 4 Allele frequencies at variable loci for Ptychobranchus
fasciolaris from Big and Little Darby Creeks. Population codes
follow Fig. 1
Locus Allele
Populations
BD 53.2 BD 100.8 BD 106.0 LD 12.4 LD 31.8
LAP n ¼28* n¼46* n¼29* n¼40 n¼25
1 0.089 0.120 0.069 0.100 0.360
2 0.768 0.533 0.379 0.900 0.600
3 0.143 0.348 0.552 0.040
MDH-1 n ¼39 n¼49 n¼31 n¼36 n¼45
1 1.000 1.000 0.936 1.000 1.000
2 0.065
MDH-2 n ¼39 n¼39 n¼31 n¼31 n¼33
1 1.000 1.000 0.968 1.000 1.000
2 0.032
GPI n ¼37 n¼49 n¼30 n¼40 n¼45
1 0.068 0.112 0.050 0.100 0.133
2 0.162 0.163 0.150 0.113 0.133
3 0.419 0.388 0.400 0.488 0.578
4 0.324 0.265 0.317 0.300 0.156
5 0.027 0.071 0.067
6 0.017
PGM-1 n ¼38* n¼46* n¼29* n¼39 n¼32
1 0.342 0.261 0.431 0.449 0.563
2 0.461 0.609 0.431 0.551 0.422
3 0.197 0.120 0.138 0.016
4 0.011
PGM-2 n ¼38* n¼48* n¼30* n¼41 n¼38*
1 0.487 0.458 0.533 0.500 0.500
2 0.382 0.427 0.350 0.500 0.487
3 0.118 0.115 0.067 0.013
4 0.013 0.050
PEP n ¼37* n¼49 n¼31* n¼40 n¼26
1 0.865 1.000 0.903 1.000 0.923
2 0.135 0.065 0.077
3 0.016
4 0.016
MPI n ¼35 n¼49 n¼31 n¼40 n¼31
1 0.914 0.888 0.952 1.000 1.000
2 0.086 0.112 0.048
EST-1 n ¼39 n¼49* n¼31 n¼41 n¼45
1 1.000 0.949 0.968 0.951 1.000
2 0.051 0.032 0.049
EST-2 n ¼31* n¼29 n¼27* n¼33 n¼36
1 0.855 1.000 0.926 1.000 1.000
2 0.145 0.074
HEX n ¼35 n¼47 n¼31* n¼40 n¼32
1 1.000 0.957 0.742 1.000 1.000
2 0.043 0.258
IDH n ¼38 n¼39 n¼31 n¼32 n¼36
1 0.013
2 0.987 1.000 1.000 1.000 1.000
Mean sample
size per locus
36.0 (1.0) 44.7 (1.6) 30.3 (0.3) 36.9 (1.1) 35.4 (1.7)
Mean number
of alleles per
locus
2.1 (0.3) 2.0 (0.3) 2.5 (0.4) 1.5 (0.2) 1.7 (0.3)
Freshwater mussel population genetics 1431
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
Statistical analyses
Electrophoretic data were analysed using standard
population genetic techniques contained in Tools for
Population Genetic Analysis (TFPGA; Miller, 1997).
Descriptive statistics calculated for each population
included mean number of alleles per locus, percent-
age polymorphic loci (95%level), and average direct-
count heterozygosity. Comparisons of measured
genotype frequencies with Hardy–Weinberg (H-W)
expectations were evaluated using the Monte Carlo
simulation method of TFPGA. To minimise the
probability of a type 1 error, we adjusted significance
levels to an experimentwise error rate of a¼0.05
using the ‘sharper’ sequential-comparison Bonferroni
(s-cB) technique described by Lessios (1992).
For polymorphic loci, allele frequencies among
populations were compared using a Monte Carlo
simulation of Fisher’s Exact Test for RxC contingency
tables in TFPGA, followed by s-cB to hold the exper-
imentwise error rate at a¼0.05. Among-population
genetic variation was further analysed by performing
an analysis of molecular variance (
AMOVAAMOVA
) using
the standard F
ST
values estimation procedure in
GENALEX6GENALEX6
(Peakall & Smouse, 2005) with interpolation
of missing data and by calculating hierarchical values
of h, the F
ST
analogue of Weir & Cockerham (1984), for
polymorphic loci. These values of hwere bootstrapped
(1000 replicates) to obtain 95%confidence intervals.
Allele frequency differences were integrated across loci
by calculating unbiased genetic distances for all pairs
of populations within a species (Nei, 1978). Genetic
similarity of populations was determined by construc-
tion of dendrograms using genetic distance and the
unweighted-pair-group method of analysis (UPGMA)
to cluster populations (Sokal & Sneath, 1963). Bootstrap
values for the nodes of the dendrogram were based on
1000 replicates. Mantel tests were performed between
matrices of hand geographical distance (calculated as
river kilometres from the mouth of Big Darby or from
the mouth of Little Darby) among pairs of populations
to evaluate the likelihood of isolation-by-distance and
between matrices of residuals of hand geographical
distance to estimate the relative contributions of gene
flow and genetic drift in determining population
genetic structure (Hutchison & Templeton, 1999).
Estimates of hfor this analysis were obtained using
the coancestry distances available from TFPGA (Miller,
Blinn & Keim, 2002).
Results
Within-population variation
Five of the 15 loci sampled were variable in all
populations of E. dilatata (Table 2). Four loci had one
or two variable populations, and six loci were fixed
for a single allele each. Twelve of 47 locus-by-
population combinations, representing five different
loci, showed genotype frequencies that did not match
H-W expectation because of heterozygote deficits.
Of the 12 loci sampled for L. siliquoidea, five were
polymorphic in all populations (Table 3), five were
variable in 1–3 populations, and two were fixed.
Sixteen of 29 locus-by-population combinations
showed genotype frequencies that were different
from H-W expectation. Thirteen of these were hetero-
zygote deficits. Seven different loci exhibited a hetero-
zygote deficit in at least one population, while all
heterozygote excesses were at the LAP locus.
Variation was detected at 12 of 14 loci sampled for
P. fasciolaris (Table 4). Four loci were polymorphic in
all five populations, three loci had three variable
populations, and five loci had 1–2 variable popula-
tions. Two loci were fixed. Sixteen of 30 locus-by-
population combinations had genotype frequencies
that were different from H-W expectation. These
combinations involved six different loci and all differ-
ences were due to heterozygote deficits. Fifteen of these
differences were found in Big Darby populations.
Among-population variation
For E. dilatata populations, pairwise differences in
allele frequencies at variable loci were low between
Table 4 (Continued)
Locus
Populations
BD 53.2 BD 100.8 BD 106.0 LD 12.4 LD 31.8
%P(95%
criterion)
50.0 42.9 57.1 28.6 35.7
Mean
hetero-
zygosity
0.12 0.13 0.11 0.11 0.12
Numbers in parentheses are standard errors. n¼number of
individuals.
*Locus-by-population combinations with heterozygote defici-
encies.
1432 D.J. Berg et al.
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
populations from the same creek, but high when Big
Darby populations were compared with those from
Little Darby. For Big Darby Creek, nine of 90 (10%)
locus-by-population-pair combinations showed signi-
ficant differences in allele frequencies. Among Little
Darby Creek comparisons, 13 of 90 (14%) were
significant. Seventy-two of 225 (32%) comparisons of
Big Darby populations with Little Darby populations
showed significant shifts in allele frequencies. Indi-
vidual locus values (not shown) for hwere low but
variable (all £0.181), while overall values were signi-
ficant (Table 5). The
AMOVAAMOVA
revealed that 90%of
variation was found within populations, while 4%
was among populations within streams, and 6%was
among streams. All these estimates were significant
(P¼0.001, 999 permutations). Unbiased genetic dis-
tances averaged 0.0118 (n¼45 pairs; range 0.0005–
0.0278) among all pairs of populations. Average
distances among populations within streams were
0.0024 (n¼10) for Big Darby and 0.0095 (n¼10) for
Little Darby, while average distance for pairs of
populations consisting of one population from each
stream was 0.0165 (n¼25). A dendrogram (Fig. 2)
based on genetic distance showed two major bran-
ches, one containing all Big Darby populations and
the other containing the Little Darby populations.
However, bootstrap support for the best tree was
minimal, with only two of eight nodes supported in
more than half of all permutations. For pairs of
populations, hand geographical distance were signi-
ficantly correlated (Fig. 3; r¼0.672, P¼0.0040, 999
permutations), but residuals of hwere not correlated
with geographical distance (r¼)0.233, P¼0.9320,
999 permutations).
For L. siliquoidea, an average of 27%(±2.1%SE,
range ¼20–40%,n¼100 comparisons) of variable
loci had significant differences in allele frequencies for
pairs of populations. Individual locus values (not
shown) for hwere highly variable, ranging from 0.000
to 0.558. While h
PS
was significantly different
from zero, h
S
was not (Table 5). Within-population
variation accounted for 83%of total variation, while
variation among populations within streams com-
prised the remaining 17%. Both of these values were
significant (
AMOVAAMOVA
,P¼0.001, 999 permutations),
while variation among streams was not significantly
different from zero (P¼1.000). Over all pairs of
populations, average unbiased genetic distance was
0.0642 (n¼10 pairs; range ¼0.0234–0.1392). Genetic
distance within stream was greater for Little Darby
(0.1345, n¼1) than Big Darby (average ¼0.0289, n¼
3), while comparisons between Big and Little Darby
populations were intermediate (average ¼0.0701,
n¼6). A dendrogram based on these distances
showed no pattern and had little bootstrap support
(tree not shown; 1 node 62%support, all others <27%
support). Geographical distance was not correlated
with hamong pairs of populations (Fig. 3; r¼0.1146,
P¼0.2980, 999 permutations), nor with residuals of h
(r¼0.1304, P¼0.4220, 999 permutations).
Patterns of allele frequencies among populations of
P. fasciolaris were similar to that of E. dilatata.
Comparisons between Big Darby populations and
between Little Darby populations rarely showed
Table 5 Hierarchical F-statistics for freshwater mussels from Darby Creek. Values of Fcorrespond to Wright’s F
IT
;q
PS
is differenti-
ation among populations within a stream, q
S
is differentiation among streams (both are components of Wright’s F
ST
); values of f are
inbreeding coefficients (Wright’s F
IS
). Table entries are mean values over all loci (lower, upper bounds of 95%confidence intervals)
Species Fh
PS
h
S
f
Elliptio dilatata 0.2971 (0.1762, 0.3917) 0.1005 (0.0507, 0.1571) 0.0644 (0.0161, 0.1332) 0.2186 (0.1000, 0.3253)
Lampsilis siliquoidea 0.4214 (0.1991, 0.7245) 0.1244 (0.0541, 0.2587) )0.1004 ()0.4003, 0.0411) 0.3392 (0.1021, 0.6559)
Ptychobranchus fasciolaris 0.4249 (0.2219, 0.7053) 0.0592 (0.0141, 0.1404) 0.0321 (0.0073, 0.0785) 0.3887 (0.2014, 0.6542)
Fig. 2 Nei unbiased genetic distances (Nei, 1978) for popula-
tions of Elliptio dilatata from the Darby Creek catchment. Popu-
lation codes follow Fig. 1. Distances were calculated using 15
allozyme loci and a tree was constructed using UPGMA. Values
at the nodes represent the percentages of 1000 bootstrapped
trees that produced these nodes. Unlabelled nodes had values
<50%.
Freshwater mussel population genetics 1433
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
significant differences in allele frequencies (2 of 12
variable loci for BD, 1 of 12 for LD). Population BD
53.2 (the site below the confluence of the two streams)
was similar to the other Big Darby populations
(differences at 2 of 12 loci for each comparison).
Allele frequencies were much more variable when
comparing Little Darby populations with Big Darby
populations; of 72 variable-locus-by-population pairs,
22 (31%) had significant differences in allele frequen-
cies. Individual locus values for hwere all <0.196; over
all loci, both components of hwere significantly
different than zero (Table 5). Analysis of molecular
variation revealed that 94%of total variation was
found within populations, 3%was among popula-
tions within streams, and 3%was among streams. All
these values were significant (
AMOVAAMOVA
,P¼0.001, 999
permutations). Average genetic distance was 0.0144
(n¼10 pairs; range ¼0.0061–0.0312) when compar-
ing all pairs of populations. Within stream distances
were similar (0.0088 for Big Darby and 0.0074 for Little
Darby), as were values for the lower Big Darby
compared with each of these (average ¼0.0144 versus
Big Darby and 0.0087 versus Little Darby; n¼2
comparisons each). Comparison of Big Darby versus
Little Darby showed the greatest distances (aver-
age ¼0.0212, n¼4). No pattern was seen in the
dendrogram based on genetic distances and bootstrap
support was very low (tree not shown; 1 node 50%
support, all others <47%support). The correlation of
geographical distance with hfor pairs of populations
was marginally insignificant with P-values varying
between 0.0477 and 0.0535 (Fig. 3; r¼0.675; we ran
five sets of 10 000 permutations each because of the
‘borderline’ values). Geographical distance was not
significantly correlated with residuals of h(r¼0.4369,
P¼0.0910, 999 permutations).
Discussion
Our measures of within-population genetic variation
are similar to those previously reported for freshwa-
ter mussels. For E. dilatata, our measure of average
heterozygosity (H)¼0.10 is similar to a value of 0.11
for a single population (Davis, 1984). Lampsilis
siliquoidea average heterozygosities of 0.14 for our
study were similar to those of 0.11 previously
reported by Davis (1984). Our average values for
both these species and for P. fasciolaris (H¼0.12) are
in the middle of the ranges reported from several
surveys of heterozygosity in unionoid bivalves (Berg
et al., 1998; Curole et al., 2004). Deviations of geno-
type frequencies from H-W expectations occurred in
a significant proportion (25–55%) of polymorphic
(a)
(b)
(c)
Fig. 3 Population differentiation (h) as a function of geograph-
ical distance for three species of freshwater mussels from the
Darby Creek catchment (a) Elliptio dilatata; (b) Lampsilis siliquo-
idea; (c) Ptychobranchus fasciolaris. Correlations (r¼0.67, 0.11 and
0.68, respectively) were significant for E. dilatata and P. fasciolaris
(Mantel test, P£0.05).
1434 D.J. Berg et al.
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
locus-by-population combinations. Reported values
for freshwater mussels have been quite variable, with
anywhere from 0%to 75%of locus-by-population
combinations exhibiting deviation from expected
heterozygote frequencies (summarised in Berg et al.,
1998 and Johnson, Liang & Farris, 1998). Like our
results, most of these deviations are caused by
heterozygote deficiencies.
Such heterozygote deficiencies have been noted in
many studies of marine bivalves utilising both
allozyme (reviewed by Berger, 1983) and microsatel-
lite markers (Launey & Hedgecock, 2001). In our
study, mutation is an unlikely factor because allo-
zymes are relatively conserved genetic markers.
Selection cannot be ruled out; in particular, the LAP
locus seems to show consistent deviation from H-W
expectation in all three species, although these devi-
ations include heterozygous excesses in three popu-
lations of L. siliquoidea, but a deficit in a fourth.
Selection against deleterious recessive alleles has been
shown to cause heterozygote deficits in Pacific oysters
(Launey & Hedgecock, 2001) and blue mussels
(Myrand, Tremblay & Se
´vigny, 2002). Genetic drift
coupled with restricted gene flow and inbreeding is
another potential cause of heterozygote deficiencies.
Studies of genetic variation within multiple popula-
tions of Q. quadrula and Amblema plicata (Berg et al.,
1998; Elderkin et al., 2007) showed few or no devia-
tions from H-W expectation. Given that genetically
effective population size is likely to be larger in large
river systems, mussel populations in these large rivers
should be less susceptible to the effects of genetic drift
and inbreeding. This is consistent with our observa-
tion of much greater occurrence of deviation from
H-W expectation in Darby Creek mussels than in
mussels that are typical of larger rivers. Of the three
species we sampled, Darby Creek populations of L.
siliquoidea (4–65 individuals representing 1–27%of all
mussels collected at a site) were generally smaller
than those of E. dilatata (45–400 individuals or 18–73%
of all mussels) or P. fasciolaris (32–70 individuals or 6–
27%of all mussels) (A.D. Christian, unpublished),
and our analysis indicated that the former was
furthest from gene flow–genetic drift equilibrium
(see discussion below). Combinations of restricted
gene flow, inbreeding, and other factors have been
implicated as the cause for heterozygote deficits in
coot clams (Gaffney et al., 1990). Overall, Darby Creek
populations of all three species of freshwater mussels
contain moderate levels of within-population genetic
variation, with many occurrences of heterozygote
deficiencies.
The pattern of among-population genetic variation
was different for L. siliquoidea than for the other two
species. While
AMOVAAMOVA
revealed significant variation
among populations for all three species, both E. dilatata
and P. fasciolaris also showed significant variation
among streams. This pattern was more pronounced in
E. dilatata, but we found some evidence of isolation-
by-distance (indicated by the correlations of geo-
graphical distance and h) in both species. Conversely,
we found no evidence for any sort of geographical
pattern in among-population variation for L. siliquo-
idea. Residuals of hwere larger for L. siliquoidea than
for E. dilatata and P. fasciolaris, and they were not
correlated with geographical distance, indicating that
random effects of genetic drift were greater than the
homogenising effects of gene flow (Hutchison &
Templeton, 1999). For P. fasciolaris, residuals tend to
increase with geographical distance so that the lack of
a significant correlation may be due to the small
sample size (10 comparisons). In this case, isolation-
by-distance and gene flow–genetic drift equilibrium
may be established at the within-basin spatial scale we
examined (Hutchison & Templeton, 1999). Isolation-
by-distance has been reported for other freshwater
mussel species at large geographical scales (>1000 km;
Berg et al., 1998; Elderkin et al., 2007). Geographical
distance among populations was a significant factor in
determining among-population genetic variation
within drainage systems at allozyme loci in a
stream-dwelling fish from Australia (McGlashan,
Hughes & Bunn, 2001). A study using amplified
fragment length polymorphisms to estimate isolation-
by-distance in four species of aquatic insects found
that dispersal ability was negatively correlated with
degree of isolation of populations (Miller et al., 2002).
Given that mussel distributions are not continuous
(they live in ‘beds’), an isolation-by-distance pattern is
best explained by a stepping-stone model of dispersal.
While isolation-by-distance and presence along a
linear stretch of stream appear to be important factors
for E. dilatata and P. fasciolaris, differences among
populations of L. siliquoidea appear to occur more
randomly.
Few other studies have examined among-popula-
tion variation of freshwater mussels within individual
stream systems; most studies have examined such
Freshwater mussel population genetics 1435
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
variation across wider geographical ranges and mul-
tiple drainage basins. While all three species in the
Darby Creek system exhibited significant among-
population variation at distances no greater than 200
river km, populations of the freshwater mussel
Q. quadrula showed no such differentiation along a
stretch of the Ohio and Tennessee rivers greater than
1000 river km (Berg et al., 1998). In fact, these latter
populations showed no significant differences in allele
frequencies among populations along this length of
river, and values of hwere very low, even among
populations separated by >2500 river km (overall h¼
0.031). Amblema plicata, a mussel species common to
both large and small rivers, exhibits even lower
variation among populations within large rivers
(h¼0.017; Elderkin et al., 2007). However, for both
of these species, geographical distances were signifi-
cantly correlated with genetic distances. Based on our
study and these others, it appears that the spatial scale
at which significant allozyme variation occurs among
freshwater mussel populations is quite variable.
While Darby Creek populations are differentiated
within short distances in a small stream system and
this differentiation may reflect effects of genetic drift
and inbreeding, Q. quadrula and A. plicata populations
are similar over much greater distances in larger river
systems.
The complex life cycle of unionoid mussels includes
a stage as glochidia larvae that are most often parasitic
on freshwater fishes. As adult mussels are essentially
sessile, dispersal among populations of freshwater
mussels is a function of host movement and these
movements can determine mussel distributions (Lee,
DeAngelis & Koh, 1998). Because dispersal promotes
gene flow among populations, host fish vagility
should be negatively correlated with measures of
differentiation (h, genetic distance) among mussel
populations. We propose that the lack of genetic
differentiation seen in a large-river mussel species
such as Q. quadrula is probably because of high
vagility of hosts; mussels from large rivers are likely
to use large, highly mobile fishes as hosts – flathead
catfish, Pylodictus olivaris (Rafinesque), in the case
of Q. quadrula [Ohio State University Museum of
Biological Diversity (MBD) Mussel/Host Database;
http://www.biosci.ohio-state.edu/molluscs/OSUM2/].
Conversely, populations of mussels from small
streams, such as those from Darby Creek, would tend
to rely on host fishes that generally move shorter
distances or are unlikely to survive passage in a
mainstem river when moving from one tributary to
another. Lampsilis siliquoidea utilises centrarchids and
minnows (MBD database), while mussels of the genus
Ptychobranchus (the database has no entries for P.
fasciolaris) parasitise darters (MBD database). Purpor-
ted hosts for E. dilatata include both centrarchids and
some fishes that are typical of larger rivers, but the
nature of the evidence in the database casts some
doubt as to which fishes are the relevant hosts for this
species. Other species of the genus Elliptio have been
reported to use darters and other percids (MBD
database), which are more likely to be found in our
study area than are the centrarchids reported in the
MBD database. Individual flathead catfish show
movements of tens or hundreds of kilometers (Vok-
oun & Rabeni, 2005), while both darters and sunfishes
move <1 km (Freeman, 1995). Smallmouth bass,
Micropterus dolomieui (Lacepe
´de), another centrarchid
host for L. siliquoidea, disperse an average of 6.5 km in
streams and individuals occasionally migrate up to
75 km during severe winters (Lyons & Kanehl, 2002).
Darters are typically considered a minimally disper-
sing group with movements of metres (Scalet, 1973),
and limited dispersal by darter hosts has been
implicated as a factor leading to patchy distribution
and poor colonisation ability of the endangered
mussel Alasmidonta heterodon (I. Lea) (McLain & Ross,
2005). Moreover, variation in species richness and
composition of freshwater mussel communities is
much greater in small streams than in larger streams
or across catchments (Watters, 1992; Haag & Warren,
1998; Vaughn & Taylor, 2000). These patterns in which
dispersal characteristics appear to be a major deter-
minant of among-population genetic structure are
consistent with those found for a variety of aquatic
invertebrates (Bilton, Freeland & Okamura, 2001).
We hypothesise that mussels from large rivers
should show relatively little among-population gen-
etic variation, while populations from headwaters and
other small streams should show genetic differenti-
ation that is randomly distributed across small spatial
scales unless gene flow (which is likely a function of
host fish vagility) is very high. At the same time, we
predict that the combination of gene flow among
large-river populations, larger genetically effective
population sizes, and more stable flow regimes
should maintain higher levels of within-population
genetic variation in large rivers than in small streams.
1436 D.J. Berg et al.
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
Conversely, genetic drift and inbreeding, resulting
from factors such as smaller genetically effective
population sizes and/or less stable environmental
conditions, should lead to lower within-population
genetic variation in small streams and more frequent
deviation of genotype frequencies from H-W expec-
tation. Given that dispersal ability is a key factor
structuring freshwater invertebrate communities (Bil-
ton et al., 2001), these same forces are likely to be at
play in shaping patterns of genetic diversity in other
aquatic macroinvertebrates. If so, our hypothesis may
explain patterns of genetic variation for many differ-
ent taxa within stream systems of different sizes.
If our hypothesis does accurately explain patterns
of genetic variation in freshwater mussels, it can
inform the development of conservation strategies for
these organisms. In general, among-population gen-
etic variation will be high in species typical of small
streams. Therefore, any given population is likely to
contain a relatively small proportion of total genetic
variation. In order to preserve much of the total
variation, a large number of populations that are
relatively near to each other will need to be protected
across the entire range of the species. Conversely,
single populations of large-river species should con-
tain much of the total genetic variation present in the
species and protection of a comparatively small
number of populations should occur over large
geographical ranges. Because these differences are
likely to result from differences in host-fish charac-
teristics, conservation of unionoid mussels is inevit-
ably tied to the presence of such fishes. Further testing
of this hypothesis should include use of additional
genetic markers, more species of mussels including
those that utilise a variety of strategies for host
selection (Haag & Warren, 2003), and a variety of
geographical areas.
Acknowledgments
We thank Jan Trybula, Vivianaluxa Gervasio, Chad
Garland, and a number of Miami University under-
graduate students for assisting with field work and
performing electrophoretic analyses. Curt Elderkin
has engaged in discussions with us throughout this
project. Todd Levine, Emy Monroe, Makiri Sei, Rich-
ard Seidel, and several anonymous reviewers greatly
improved the quality of this manuscript. This research
was funded by grants from the Darby Creek program
of the Ohio Chapter of The Nature Conservancy and
the Ohio Sea Grant Development Fund.
References
Anthony J.L. & Downing J.A. (2001) Exploitation
trajectory of a declining fauna: a century of freshwater
mussel fisheries in North America. Canadian Journal of
Fisheries and Aquatic Sciences,58, 2071–2090.
Berg D.J. & Berg P.H. (2000) Conservation genetics of
freshwater mussels: comments on Mulvey et al. Con-
servation Biology,14, 1920–1923.
Berg D.J., Cantonwine E.G., Hoeh W.R. & Guttman S.I.
(1998) Genetic structure of Quadrula quadrula (Bivalvia:
Unionidae): little variation across large distances.
Journal of Shellfish Research,17, 1365–1373.
Berg D.J., Haag W.R., Guttman S.I. & Sickel J.B. (1995)
Mantle biopsy: a technique for nondestructive tissue-
sampling of freshwater mussels. Journal of the North
American Benthological Society,14, 577–581.
Berger E.M. (1983) Population genetics of marine gastro-
pods and bivalves. In The Mollusca, Volume 6: Ecology
(Ed. Russell-Hunter W.D.), pp. 563–596. Academic
Press: Orlando, FL.
Bilton D.T., Freeland J.R. & Okamura B. (2001) Dispersal
in freshwater invertebrates. Annual Review of Ecology
and Systematics,32, 159–181.
Bogan A.E. (1993) Freshwater bivalve extinctions (Mol-
lusca: Unionoida): a search for causes. American Zoolo-
gist,33, 599–609.
Christian, A.D. (1995). Analysis of the commercial mus-
sel beds in the Cache and White rivers in Arkansas. MS
thesis, Arkansas State University, State University, AR,
197pp.
Clayton J.W. & Tretiak D.N. (1972) Amine citrate buffers
in starch gel electrophoresis. Journal of the Fisheries
Research Board of Canada,29, 1169–1172.
Curole J.P., Foltz D.W. & Brown K.A. (2004) Extensive
allozyme monomorphism in a threatened species of
freshwater mussel, Margaritifera hembeli Conrad (Bi-
valvia: Margaritiferidae). Conservation Genetics,5, 271–
278.
Davis G.M. (1984) Genetic relationships among some
North American Unionidae (Bivalvia): sibling species,
convergence, and cladistic relationships. Malacologia,
25, 629–648.
Elderkin C.L., Christian A.D., Vaughn C.C., Metcalfe-
Smith J.L. & Berg D.J. (2006) Population genetics of
the freshwater mussel, Amblema plicata (Say 1817)
(Bivalvia: Unionidae): evidence of high dispersal
and post-glacial colonization. Conservation Genetics,8,
355–372.
Freshwater mussel population genetics 1437
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
Freeman M.C. (1995) Movements by two small fishes in a
large stream. Copeia,1995, 361–367.
Gaffney P.M., Scott T.M., Koehn R.K. & Diehl W.J. (1990)
Interrelationships of heterozygosity, growth rate and
heterozygote deficiencies in the coot clam, Mulinia
lateralis.Genetics,124, 687–699.
Haag W.R. & Warren M.L., Jr (1998) Role of ecological
factors and reproductive strategies in structuring
freshwater mussel communities. Canadian Journal of
Fisheries and Aquatic Sciences,55, 297–306.
Haag W.R. & Warren, M.L., Jr (2003) Host fishes and
infection strategies of freshwater mussels in large
Mobile Basin streams, USA. Journal of the North
American Benthological Society,22, 78–91.
Harris J.L. & Christian A.D. (2000) Current Status of the
Freshwater Mussel Fauna of the White River, Arkansas,
River Miles 10-255. Final Report, Memphis District, U.S.
Army Corps of Engineers, Memphis, TN.
Harris H. & Hopkinson D.A. (1976) Handbook of Enzyme
Electrophoresis in Human Genetics. American Elsevier
Publishing Company, New York.
Hebert P.D.N. & Beaton M.J. (1989) Methodologies for
Allozyme Analysis Using Cellulose Acetate Electrophoresis.
Helena Laboratories, Beaumont, TX.
Hutchison D.W. & Templeton A.R. (1999) Correlation of
pairwise genetic and geographic distance measures:
inferring the relative influences of gene flow and drift
on the distribution of genetic variability. Evolution,53,
1898–1914.
Johnson R.L., Liang F.Q. & Farris J.L. (1998) Genetic
diversity among four Amblemini species (Bivalvia:
Unionidae) in the Cache and White Rivers, Arkansas.
The Southwestern Naturalist,43, 321–332.
Launey S. & Hedgecock D. (2001) High genetic load in the
Pacific oyster Crassostrea gigas.Genetics,159, 255–265.
Lee H.-L., DeAngelis D. & Koh H.-L. (1998) Modeling
spatial distribution of the unionid mussels and the
core-satellite hypothesis. Water Science and Technology,
38, 73–79.
Lessios H.A. (1992) Testing electrophoretic data for
agreement with Hardy-Weinberg expectations. Marine
Biology,112, 517–523.
Lydeard, C., Cowie R.H., Ponder W.F. et al. (2004) The
global decline of nonmarine mollusks. Bioscience,54,
321–330.
Lyons J. & Kanehl P. (2002) Seasonal movements of
smallmouth bass in streams. In: Black Bass: Ecology,
Conservation, and Management (Ed. Philipp D.P. &
Ridgeway M.S.), pp. 149–160. Symposium 31, Ameri-
can Fisheries Society, Bethesda, MD.
Machordom A., Araujo R., Erpenbeck D. & Ramos M.A.
(2003) Phylogeography and conservation genetics of
endangered European Margaritiferidae (Bivalvia:
Unionoidea). Biological Journal of the Linnean Society,
78, 235–252.
McGlashan D.J., Hughes J.M. & Bunn S.E. (2001) Within-
drainage population genetic structure of the fresh-
water fish Pseudomugil signifer (Pseudomugilidae) in
northern Australia. Canadian Journal of Fisheries and
Aquatic Sciences,58, 1842–1852.
McLain D.C. & Ross M.R. (2005) Reproduction based on
local patch size of Alasmidonta heterodon and dispersal
by its darter host in the Mill River, Massachusetts,
USA. Journal of the North American Benthological Society,
24, 139–147.
Miller M.P. (1997) Tools for Population Genetic Analysis
(TFPGA), Ver. 1.31. http://www.marksgeneticsoftware.
net.
Miller M.P., Blinn D.W. & Keim P. (2002) Correlations
between observed dispersal capabilities and patterns
of genetic differentiation in populations of four aquatic
insect species from the Arizona White Mountains,
U.S.A. Freshwater Biology,47, 1660–1673.
Mulvey M.C., Lydeard C., Pyer D.L., Hicks K.M., Brim-
Box J., Williams J.D. & Butler R.S. (1997) Conservation
genetics of North American freshwater mussels
Amblema and Megalonaias.Conservation Biology,11,
868–878.
Myrand B., Tremblay R. & Se
´vigny J.-M. (2002) Selection
against blue mussel (Mytilus edulis L.) homozygotes
under stressful conditions. The Journal of Heredity,93,
238–248.
Nei M. (1978) Estimation of average heterozygosity and
genetic distance from a small number of individuals.
Genetics,89, 583–590.
Parmalee P.W. & Bogan A.E. (1998) The Freshwater
Mussels of Tennessee. The University of Tennessee
Press, TN, Knoxville.
Peakall R. & Smouse P.E. (2005) GenAlEx 6: Genetic
Analysis in Excel. Population Genetic Software for Teaching
and Research. Australian National University, Canberra,
Australia. http://www.anu.edu.au/BoZo/GenAlEx/.
Saccheri I., Kuussaari M., Kankare M., Vikman P.,
Fortelius W. & Hanski I. (1998) Inbreeding and
extinction in a butterfly metapopulation. Nature,392,
491–494.
Scalet, C.G. (1973) Stream movements and population
density of the orangebelly darter, Etheostoma radiosum
cyanorum (Osteichthyes: Percidae). Southwestern Natur-
alist,17, 381–387.
Selander R.K., Smith M.H., Yang S.Y., Johnson W.E. &
Gentry J.B. (1971) Biochemical polymorphism and
systematics in the genus Peromyscus. 1. Variation in
the old field mouse (Peromyscus polionotus). Studies
in Genetics VI. University of Texas Publications,7103,
49–90.
1438 D.J. Berg et al.
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
Sokal R.R. & Sneath P.H.A. (1963) Principles of Numerical
Taxonomy. W.H. Freeman, San Francisco, CA.
Spielman D., Brook B.W. & Frankham R. (2004) Most
species are not driven to extinction before genetic
factors impact them. Proceedings of the National Academy
of Sciences of the United States of America,101, 15261–
15264.
Stiven A.E. & Alderman J. (1992) Genetic similarities
among certain freshwater mussel populations of the
Lampsilis genus in North Carolina. Malacologia,34, 355–
369.
Strayer D.L., Downing J.A., Haag W.R., King T.L., Layzer
J.B., Newton T.J. & Nichols S.J. (2004) Changing
perspectives on pearly mussels, North America’s most
imperiled animals. Bioscience,54, 429–439.
Vaughn C.C. & Taylor C.M. (2000) Macroecology of a
host-parasite relationship. Ecography,23, 11–20.
Vokoun J.C. & Rabeni C.E. (2005) Variation in annual
movement cycle of flathead catfish within and between
two Missouri watersheds. North American Journal of
Fisheries Management,25, 563–572.
Watters G.T. (1992) Unionids, fishes, and the species-area
curve. Journal of Biogeography,19, 481–490.
Watters G.T. (1996) 1996 Freshwater Mussel Survey of Big
Darby Creek. Report to The Ohio Chapter of The Nature
Conservancy. Columbus, OH.
Weir B.S. & Cockerham C.C. (1984) Estimating F-
statistics for the analysis of population structure.
Evolution,38, 1358–1370.
Williams J.D., Warren, M.L., Jr, Cummings K.S., Harris
J.L. & Neves R.J. (1993) Conservation status of fresh-
water mussels of the United States and Canada.
Fisheries,18, 6–22.
(Manuscript accepted 13 February 2007)
Freshwater mussel population genetics 1439
2007 The Authors, Journal compilation 2007 Blackwell Publishing Ltd, Freshwater Biology,52, 1427–1439
... Active research and management of imperilled freshwater mussels increased in the USA after the development of a national strategy for mussel conservation in the late 1990s (National Native Mussel Conservation Committee, 1998;Freshwater Mollusk Conservation Society, 2016). However, population genetic studies of freshwater mussels have been limited, despite the importance of such research for informing conservation efforts (but see, for example, Berg, Christian & Guttman, 2007;Elderkin et al., 2008;Inoue et al., 2014). Nearly all population genetic studies of freshwater mussels have used allozyme markers or microsatellites, which can display higher error rates during genotyping, have limited capacity to detect fine-scale geographical patterns of genetic differentiation and possess increased susceptibility to homoplasy compared with nextgeneration sequencing approaches (Pasqualotto, Denning & Anderson, 2007;Jeffries et al., 2016;Garrison, Johnson & Whelan, 2021). ...
... The stepping-stone model of dispersal (Kimura & Weiss, 1964) has been shown to provide the best explanation of population structure in other freshwater mussels, including in E. triquetra (Berg, Christian & Guttman, 2007;Beaver, Woolnough & Zanatta, 2019 could be biasing migration inference (Beerli, 2004;Delser et al., 2019) and also obscuring a historical pattern of IBD. Nevertheless, both analyses of genetic structure and migration analyses indicate population fragmentation, and, if migration between sites is currently occurring, it must be rare. ...
Article
Full-text available
Freshwater mussels are an important element of freshwater biodiversity and provide essential ecosystem services. However, mussels are among the most imperilled groups of organisms in the world. Although research has increased in recent years, information about range‐wide genetic diversity and historical demography of most species is lacking. One such species is Cumberlandian combshell Epioblasma brevidens , which is listed as endangered under the US Endangered Species Act. Genetic diversity of E. brevidens was analysed using a high‐resolution RADseq approach and included the previously overlooked Bear Creek population. Hypotheses were tested about population decline, comparative genetic diversity and population structure with model‐based approaches enabled by a genome‐scale dataset. Estimates of genetic differentiation among populations of E. brevidens were lower than past analyses, suggesting higher historical population connectivity than previously known. Demographic analyses indicate relatively recent splits among E. brevidens populations in the late Pleistocene to early Holocene, with clear founder effects in two populations. The Clinch River population has the highest genetic diversity and effective population size, despite demographic analyses revealing decline of this population since the Pleistocene. Analyses of both population structure and migration show evidence of past gene flow, but all populations are currently isolated by artificial barriers. Analyses indicate that populations began to decline before industrialization, but fragmentation and population extirpation has been exacerbated by modern habitat destruction. Relatively high genetic diversity in the Bear Creek population indicates that water quality improvements in the last 20 years have had a positive impact on population viability, offering promise for targeted management actions. In contrast, the Big South Fork population that has been presumed stable showed low genetic diversity and effective population size. Furthermore, genetic structure among sampled populations indicates that reintroduction efforts should use broodstock from as close to the reintroduction site as possible.
... Genetic connectivity of mussel populations likely depends on many factors such as mussel habitat suitability and spatial configuration and host fish identity, abundance and behavior Roe et al., 2001;Kelly, 2005;Berg et al., 2007). ...
Article
Full-text available
Unionid drift dispersal
... Relationships among populations of freshwater organisms have been formed by major geological changes in aquatic systems [13,14]. Genetic relationships within and among populations are not only formed by historic geological events, but also affected by intrinsic effects [15], including habitat preferences and dispersal abilities among habitats [16,17,18]. Dispersal ability appears to be major determinant of population genetics of freshwater organisms [17,19]. ...
Article
Full-text available
This is an Open Access Journal / article distributed under the terms of the Creative Commons Attribution License (CC BY-NC-ND 3.0) which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. All rights reserved. Freshwater mussels are bivalve inhabitants of inland waters represented on all continents except Antarctica. They are among the most endangered animal taxa. In the present study population estimation and phylogenetic study of freshwater mussels (Unionidae) was conducted. Totally 3019 animals in 5 different species of freshwater mussels were recorded in Cauvery river, Srirangam, Tiruchirapplli district, Tamil Nadu, India. They were belonging to two genera, Lamellidens and Parreysia and one family Unionidae. Parreysia favidens was recorded in maximum number and minimum numbers of Lamellidens corrianus were recorded in Cauvery river. Phylogenetic study result revealed that there was close relationship between Lamellidens corrianus and Lamellidens marginalis and distantly related species were Lamellidens corrianus and Parreysia corrugata. ABSTRACT RESEARCH ARTICLE
... These shortterm climate fluctuations and associated changes in water bodies could have favoured the faunal dispersal between basins (Suárez & Lombardo, 2004) including the coypu and other species that can only disperse across the watercourses (e.g. fishes, juvenile phases of insects, mussels, etc.) (Wilcock et al., 2003;Berg et al., 2007). Related to this, Byrne et al. (2020) observed the same patterns for Cnesterodon decemmaculatus, a freshwater native fish, in the Luján River and a neighbouring river basin. ...
Article
Myocastor coypus is a rodent native to South America that is strongly linked to river systems. Past studies on the coypu in its native distribution range were aimed mainly at its ecology and parasitology and genetic studies are very limited. Here, we used sequences of the mitochondrial D-loop control region to study the genetic diversity, population genetic structure and some aspects of historical population dynamics of coypu at the centre of its native range. Our results showed moderate to high levels of genetic diversity and an absence of genetic structure in the study area. Bayesian analysis of population structure (BAPS) showed the existence of two haplogroups distributed in most sampling sites. These results suggest that movement of coypu is independent of the structure of current known river networks and its dispersal patterns are related to flooding events and the presence of lagoons connecting river basins. The demographic expansion patterns in these populations and those of other South American mammals during the Late Pleistocene support the hypothesis that demographic changes in wild populations are related to fluctuations in climate and ecology. The mitochondrial data obtained in this study constitute the first record of these types of sequences in the native range of M. coypus. Myocastor coypus es un roedor nativo de Sudamérica fuertemente ligado a los sistemas fluviales. En su área de distribución nativa, los estudios llevados a cabo en la especie estuvieron dirigidos principalmente a estudiar su ecología y zoonosis, sin embargo, los estudios genéticos son muy limitados. En el presente trabajo, utilizamos secuencias del d-loop de la Región Control del ADN mitocondrial para estudiar la diversidad genética, la estructura genética de la población y algunos aspectos de la dinámica poblacional histórica del coipo en el centro de su área de distribución. Nuestros resultados develaron niveles moderados y altos de diversidad genética, y ausencia de estructuración genética en el área de estudio. El BAPS exhibió la existencia de dos haplogrupos distribuidos en la mayoría de los sitios de muestreo. Estos resultados sugieren que el movimiento de los coipos sería independiente de la estructura de las redes fluviales actuales conocidas y que sus patrones de dispersión estarían relacionados con eventos de inundación y con la presencia de lagunas que conectan las cuencas fluviales. Los patrones de expansión demográfica observados en esta población de coipos y en otras poblaciones de mamíferos sudamericanos durante el Pleistoceno tardío, apoyan la hipótesis de que las fluctuaciones climáticas y los cambios ecológicos están relacionados con cambios demográficos en las poblaciones silvestres. Los datos mitocondriales obtenidos en este estudio constituyen el primer registro de este tipo de secuencias en el área de distribución nativa de esta especie.
... Habitat disturbances and fragmentation can influence the genetic structure of mussel beds by blocking gene flow among beds (Watters 1996;Strayer et al. 2004;Newton et al. 2008;Schwalb et al. 2011;Galbraith et al. 2015). Multiple studies have evaluated genetic structure in unionid mussels (Berg et al. 1998;Kelly and Rhymer 2005;Elderkin et al. 2006;Berg et al. 2007;Zanatta and Wilson 2011;Galbraith et al. 2015;Jones et al. 2015;Inoue and Berg 2017;Hoffman et al. 2018), but few studies have estimated effective populations sizes of mussel beds and compared these estimates to total bed populations (Inoue et al. 2015). ...
Article
Myriad anthropogenic factors have led to substantial declines in North America's freshwater mussel populations over the last century. A greater understanding of mussel dispersal abilities, genetic structure, and effective population sizes is imperative to improve conservation strategies. This study used microsatellites to investigate genetic structure among mussel beds and estimate effective population sizes of a common North American mussel species, Amblema plicata, in the Little River, Oklahoma. We used five microsatellite loci to genotype 270 individuals from nine mussel beds distributed throughout the river and one of its tributaries, the Glover River. Our results indicate that subpopulations of A. plicata in the Little River are genetically similar. Upstream subpopulations had less genetic diversity than sites located downstream of the confluence of the Glover and Little rivers. Downstream subpopulations were primarily assigned to the same genetic group as upstream subpopulations, but they were admixed with a second genetic group. Low flows during droughts likely influenced the observed genetic structuring in A. plicata populations in the Little River. Additionally, downstream subpopulations may be admixed with a genetically distinct population of A. plicata, which may account for the increased genetic diversity. Estimates of effective population sizes (Ne) of large mussel beds were low compared to the total abundance (N) of A. plicata. While our data have limitations, they provide important information on the spatial scale at which conservation plans should focus and the population sizes that should be sustained through relocation and restocking programs.
Article
Full-text available
ecological and evolutionary consequences. In lotic freshwater systems, landscape features such as barriers or connectors, can affect dispersal and thus gene flow. It is of special interest to characterize population genetic structure in the presence of impassable barriers because they can restrict dispersal and thus isolate populations. On the other hand, connectors such as man-made canals connect watersheds that would otherwise be isolated, thereby enhancing gene flow and potentially genetically homogenizing populations. Freshwater mussels (Order: Unionida), many of which are highly threatened, have a unique life history involving parasitism on host fish species for larval development and dispersal. Landscape features that affect fish passage will consequently affect mussel dispersal and the maintenance of populations and metapopulations. This manuscript explores how landscape features affect the population genetic structure and genetic diversity of the freshwater mussel species Lampsilis siliquoidea, Fatmucket, in the southwestern Lake Ontario drainage. Data from seven microsatellite loci revealed that 1) waterfalls act as barriers for gene flow between freshwater mussel populations within the same river, 2) populations above the barrier had lower genetic diversity, 3) there was no detectable effect of manmade canals on genetic diversity and, 4) populations that were not hydrologically connected by a canal were marginally more differentiated than the populations connected by a canal. Landscape features can alter the connectivity within and between drainages thus changing the scale at which management will be effective.
Article
Full-text available
Abstract The shell morphologies of the freshwater mussel species Pleurobema clava (federally endangered) and Pleurobema oviforme (species of concern) are similar, causing considerable taxonomic confusion between the two species over the last 100 years. While P. clava was historically widespread throughout the Ohio River basin and tributaries to the lower Laurentian Great Lakes, P. oviforme was confined to the Tennessee and the upper Cumberland River basins. We used two mitochondrial DNA (mtDNA) genes, 13 novel nuclear DNA microsatellite markers, and shell morphometrics to help resolve this taxonomic confusion. Evidence for a single species was apparent in phylogenetic analyses of each mtDNA gene, revealing monophyletic relationships with minimal differentiation and shared haplotypes. Analyses of microsatellites showed significant genetic structuring, with four main genetic clusters detected, respectively, in the upper Ohio River basin, the lower Ohio River and Great Lakes, and upper Tennessee River basin, and a fourth genetic cluster, which included geographically intermediate populations in the Ohio and Tennessee river basins. While principal components analysis (PCA) of morphometric variables (i.e., length, height, width, and weight) showed significant differences in shell shape, only 3% of the variance in shell shape was explained by nominal species. Using Linear Discriminant and Random Forest (RF) analyses, correct classification rates for the two species' shell forms were 65.5% and 83.2%, respectively. Random Forest classification rates for some populations were higher; for example, for North Fork Holston (HOLS), it was >90%. While nuclear DNA and shell morphology indicate that the HOLS population is strongly differentiated, perhaps indicative of cryptic biodiversity, we consider the presence of a single widespread species the most likely biological scenario for many of the investigated populations based on our mtDNA dataset. However, additional sampling of P. oviforme populations at nuclear loci is needed to corroborate this finding.
Article
Full-text available
The tribes Contradentini and Rectidentini (Unionidae) comprise a diverse clade of freshwater mussels endemic to South-east Asia. Our understanding of the diversity and phylogeny of this radiation has improved dramatically in recent years, but this systematic transformation has not yet benefited from comprehensive museum sampling or phylogenomic methods. A synthetic taxonomic revision of the Contradentini+Rectidentini that leverages these useful and accessible methods is needed. We set out to (1) generate a phylogenomic reconstruction of the supraspecific relationships of the Contradentini+Rectidentini using anchored hybrid enrichment, (2) revise the taxonomy and geographic boundaries of the generic and species-level diversity of the radiation, and (3) identify patterns of freshwater mussel diversity and distribution in this clade and discuss the processes that may have precipitated them. Our phylogenomic reconstruction using over 1600 loci, with a total alignment length of over a half a million nucleotides, recovers a well supported phylogeny of the clade that resolves four independent multispecies radiations endemic to the Mekong drainage. We examined, digitised, and imaged 1837 records from 15 natural history museums that provided the necessary data to document the morphological variation and geographic distributions of the focal taxa. We also analysed 860 COI sequences, 519 of which were generated in this study, to better understand the species boundaries and geographic distributions of the recovered clades. We recognise 54 valid species in the tribes Contradentini and Rectidentini, including 9 described herein as new to science. Out of this revision emerged several interesting biogeographic patterns that appear to have resulted from recent stream capture, historical confluence, and intradrainage barriers to dispersal. We hypothesise that these phenomena shaped the diversity and distribution of the Contradentini+Rectidentini, contributing to the formation of several characteristic freshwater mussel provinces in South-east Asia.
Article
Full-text available
Many landscape and biotic processes shape the genetic structure of populations. The genetic structure of species with parasitic stages may also depend on the life history and ecology of their host. We investigated population genetic structure of the mussel Margaritifera margaritifera in Southern Sweden, and in relation to the population size and life history of its hosts, Salmo trutta and S. salar. Mussel populations were genetically differentiated into two clusters, further subdivided into four clusters and distinct conservation units. Regardless of host species, the genetic differentiation was lower among mussel populations sustained by sea-migrating than by resident hosts, while the genetic diversity was higher in mussel populations sustained by sea-migrating than by resident hosts. Genetic diversity of mussel populations was positively related to host abundance. Mussel population size was positively related to high genetic diversity of mussels sustained by resident hosts, while low mussel population size sustained by sea-migrating hosts had a high genetic diversity. The results of our study suggest a combined influence of mussels and host fish on genetic structure of unionoid mussels. We suggest to conserve not only mussel population sizes and host fish species, but also consider host migratory/resident behaviour and abundance when designing conservation programs.
Article
Full-text available
South-western Australia is isolated from other forested regions of Australia by desert and bounded on southern and western sides by the Southern and Indian Oceans, respectively, with Westralunio carteri (Iredale, 1934) as the sole endemic freshwater mussel. Its conservation status is vulnerable. This species has a history of nomenclatural change and its systematic placement and population genetic history are largely unknown. We sampled 46 individuals from 13 sites across W. carteri’s distribution and sequenced two mitochondrial genes (16S rDNA and cytochrome c oxidase subunit I) and one nuclear gene (28S rDNA). The mitochondrial haplotype networks and COI phylogenies revealed three evolutionarily significant units (ESUs): “W. carteri” I including the west coast populations, “W. carteri” II from the south and south-eastern range, and “W. carteri” III only occurring in the south-western tip of Australia. Four species delimitation methods identified two molecular operational taxonomic units supporting two distinct species (“W. carteri” I and “W. carteri” II + III). Phylogeographic patterns revealed herein confirm the historical separation of Western and Southern paleo-basins, also highlighting the isolation of the southwestern extremity of the region. This underlines the need for taxonomic revision and will require a reevaluation of W. carteri’s conservation status.
Article
Full-text available
We investigated host fishes, timing and modes of glochidial release, and host-attraction strategies for 7 species of freshwater mussels from the Buttahatchee and Sipsey rivers (Mobile Basin), Alabama and Mississippi, USA. We determined hosts as fish species that produced juvenile mussels from laboratory-induced glochidial infections. We established the following primary mussel/host relationships: Elliptio arca with Etheostoma artesiae and Percina nigrofasciata; Fusconaia cerina with 6 species of minnows (Cyprinidae); Lampsilis ornata with Micropterus salmoides; Medionidus acutissimus with 8 species of darters (Percidae); Obovaria unicolor with Ammocrypta beani, A. meridiana, and Etheostoma artesiae; Pleurobema decisum with Cyprinella venusta; and Quadrula asperata with Ictalurus punctatus. For most mussel species, host use was similar to that of closely related species, indicating that, in general, this trait is highly conserved at the generic level. Four mussel species used host-attraction strategies that targeted their specific host fish. Fusconaia cerina and P. decisum released glochidia in conglutinates that elicited feeding responses from fishes in the field and in the laboratory. Gravid female Lampsilis ornata and M. acutissimus displayed mantle lures. Host-attraction strategies were less apparent for E. arca and Q. asperata, but these species released glochidia in association with copious mucous secretions, which may serve to entangle fishes, facilitating host infection. No host-attraction strategy was apparent for O. unicolor.
Article
Attempts to relate estimates of regional FST to gene flow and drift via Wright's (1931) equation FST ≈ 1/ (4Nm + 1) are often inappropriate because most natural sets of populations probably are not at equilibrium (McCauley 1993), as assumed by the island model upon which the equation is based, or ineffective because the influences of gene flow and drift are confounded in the product Nm. Evaluations of the association between genetic (FST ) and geographic distances separating all pairwise populations combinations in a region allows one to test for regional equilibrium, to evaluate the relative influences of gene flow and drift on population structure both within and between regions, and to visualize the behavior of the association across all degrees of geographic separation. Tests of the model using microsatellite data from 51 populations of eastern collared lizards (Crotaphytus collaris collaris) collected from four distinct geographical regions gave results highly consistent with predicted patterns of association based on regional differences in various historical and ecological factors that affect the amount of drift and gene flow. The model provides a prerequisite for and an alternative to regional FST analyses, which often simply assume regional equilibrium, thus potentially leading to erroneous and misleading inferences regarding regional population structure.
Article
North American freshwater bivalves of the families Unionidae and Margaritiferidae represent one of the endangered faunas of the world. Effective management of threatened and endangered species requires knowledge not only of abundances of these species but also the degree of variation within species and the geographic distribution of this intraspecific variation. We used allozyme electrophoresis to examine the genetic structure of seven Quadrula quadrula populations from the Ohio, Tennessee, and Tensas Rivers. We then considered the implications of our results for the development of effective bivalve conservation strategies. Descriptive measures of genetic variation within populations are quite high (2.1 ± 0.1(se) alleles per locus; 61.4 ± 2.6% polymorphic loci; 0.24 ± 0.01 heterozygosity) relative to other unionids. Genotype frequencies met Hardy-Weinberg expectations at all polymorphic loci. Among-population variation was low and mostly confined to differences between the Tensas River population (lower Mississippi River basin) and the Ohio River basin populations. Significant differences in allele frequencies among populations were only detected at 3 of 10 loci; no differences in allele frequencies were found among Ohio River basin populations. Genetic distances, though all small, were significantly correlated with geographic distance. Estimated gene flow was high among populations, but variation among populations did tend to follow the predictions of an isolation-by-distance model of dispersal. The low levels of among-population genetic variation are remarkable given that these populations are separated by distances as great as 2,500+ river kilometers. High levels of gene flow may ensure that within-population variation remains high and that populations do not become differentiated due to genetic drift. An optimum conservation strategy for this species in the mainstem of the Ohio River would center on the protection of a number of large populations and maintenance of corridors for dispersal of host fishes. Successful protection of threatened and endangered species requires conservation of both abundance and genetic diversity of unionids. Further work is needed to characterize general patterns of genetic structure within freshwater bivalve species.
Article
We reviewed the literature and gathered new radio telemetry data from the Otter Creek/Pecatonica River system in southwestern Wisconsin to examine general patterns in seasonal migrations by stream-dwelling smallmouth bass Micropterus dolomieu. During summer, smallmouth bass typically remain in localized areas, with net movements less than 1 km. In Otter Creek and the Pecatonica River movements were less than 200 m. In the fall, some smallmouth bass populations remain sedentary, while others migrate more than 75 km to reach winter habitat. Migration distance is correlated with winter severity; smallmouth bass move little in streams that do not freeze, but often travel more than 5 km in systems with ice. In the fall, most smallmouth bass left Otter Creek and entered the much larger Pecatonica River where they overwintered. Mean net movement between summer and winter habitats was 6.5 km (range 3.1-20.8 km). No smallmouth bass population has been observed to make long-distance movements during the winter. Winter habitat and activity levels vary among populations, and have little obvious relation with winter severity. In some systems, smallmouth bass occupy the deepest water available or areas of cover (boulders, logs), where they remain inactive for the winter. However, in other systems, smallmouth bass avoid the deepest holes, and are often active in areas with little cover. In the Otter/Pecatonica system, smallmouth bass occupied slow-moving runs (maximum depth 0.9-1.8 m) with limited cover during the winter and had home ranges that averaged 299 m in length (range 130-710 m). Many smallmouth bass populations undertake spawning migrations in the spring, usually to smaller tributaries. Migration distances vary, but may be greater than 10 km. Most smallmouth bass from Otter Creek that overwintered in the Pecatonica River, returned to Otter Creek during the spring. Mean net distance traveled was 6.5 km (range 0-11.5 km). Many smallmouth bass populations exhibit homing tendencies towards particular habitats. However, stocked smallmouth bass usually show little site fidelity, and quickly disperse long distances from the stocking location. Overall, movement tendencies differ dramatically among stream-dwelling smallmouth bass populations, suggesting that population-specific management strategies are needed.
Article
Movements by adult Percina nigrofasciata and juvenile Lepomis auritus were examined in a large Coastal Plain stream in the southeastern United States. I marked fishes with subcutaneous injections of acrylic paints to indicate capture location within a 550-m long study site. Recaptures over an 18-month period primarily occurred within 33 m of the original capture location, suggesting longterm residence in relatively small areas. However, 11 Percina and three Lepomis moved at least 100 m (maximum distance moved = 200 m for Lepomis, 420 m for Percina), and individuals of both species shifted between distinctly different mesohabitats (boulder riffle, sand pool, and gravel riffle). Distance moved did not strongly relate to time between captures. Long-distance movements (> 33 m) by Percina mostly occurred from Nov. through June. The winter and spring-time movements by darters occurred in upstream and downstream directions and into all three mesohabitats, and at least 40% of these movements were unassociated with periods of extreme high flow. Periodic long-range movements may allow small fishes to respond to variation in resources over a large area and across a variety of stream habitats.