Download Chapter 4 The remedy to genetic erosion problems

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Hologenome theory of evolution wikipedia , lookup

Inclusive fitness in humans wikipedia , lookup

Evolution wikipedia , lookup

Evolutionary landscape wikipedia , lookup

Inclusive fitness wikipedia , lookup

Genetic drift wikipedia , lookup

Adaptation wikipedia , lookup

Philopatry wikipedia , lookup

Genetics and the Origin of Species wikipedia , lookup

Introduction to evolution wikipedia , lookup

Transcript
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
Supporting information
Appendix S1 Factors expected to affect the magnitude of genetic rescues for fitness
The magnitude of genetic rescue due to outcrossing within species is expected to
equal the extent of inbreeding depression if inbreeding is totally eliminated and there
is no outbreeding depression. Empirical findings from several studies supports this
prediction (see main text).
The magnitude of inbreeding depression (ID) depends upon the increase in
inbreeding coefficient (∆F), and the genetic load in the population, as indicated by the
following equation (after Falconer & Mackay 1996):
# loci
ID = ∑2piqidi∆F
(eqn S1)
i = 1,
where 2piqi is the heterozygosity for each polymorphic locus and di its dominance
deviation (deviation of the heterozygote mean from the mean of the homozygotes)
and the summation is over all polymorphic loci with genotypes differing in their
impacts on the quantitative traits (reproductive fitness in the current context). The
2piqi term is reduced by a history of small population sizes or for selfing species. For
habitually inbreeding species, this term will be very small, such that there will be at
best a weak relationship between inbreeding depression and F while for mixed
mating species the relationship is likely to be highly variable. Thus, tests for impacts
of ∆F are best restricted to outbreeding species (as I did).
As expected from this equation (with the effects of loci combining additively on
average), inbreeding depression is usually approximately linearly related to the
inbreeding coefficient (see Falconer & Mackay 1996; Lynch & Walsh 1998). An
alternative formulation is provided by Morton et al. (1956) for the ratio of survival in
inbred (SI) and outbred (So) populations where the effects of loci are expected to
combine multiplicatively:
SI/SO = eA – BF
(eqn S2)
where B is the number of haploid lethal equivalents (a measure of the genetic load in
the population) and eA is the survival of the non-inbred population. In this case, the
natural logarithm of survival is expected to decline linearly with F, as is usually found
empirically (Ralls et al. 1988). Thus, by extension the magnitude of genetic rescue
effects should be positively correlated with the difference in inbreeding coefficient
(ΔF) between the inbred and outcrossed populations.
The amount of benefit (heterosis) on fitness from crossing two populations is
predicted to depend on the genetic distinctiveness (y - difference in allele frequencies
between the target and rescuing populations summed across loci affecting fitness),
as illustrated by the following equation for F1 heterosis for traits determined by
zygotic genotypes (Falconer & Mackay 1996) (the F2 heterosis is half that in the F1):
# loci
F1 heterosis = ∑diyi2
(eqn S3)
i = 1,
The extent of inbreeding depression is expected to differ with ploidy level
(diploid ≥ polyploid > haplodiploid > haploid: Frankham et al. 2010, 2014), but there
1
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
were too few data to make meaningful comparisons of genetic rescue effects among
ploidy levels in my analyses.
The predictions about genetic rescue being tested are:
1. The outcrossing of inbred populations to isolated populations of the same
sexually reproducing species (with the same karyotype, adapted to the same
environment and isolated for ≤ 500 years) will be beneficial, especially in
naturally outbreeding species (Frankham et al. 2011).
2. Genetic rescue effects will be positively correlated with ΔF for naturally
outbreeding species (see above). Relationships are expected with both the
maternal and the zygotic ΔF, as discussed in Appendix S2 below.
3. The magnitude of genetic rescue effects will be larger in wild or stressful
environments than in benign/captive ones (Dudash 1990; Fox & Reed 2011;
Enders & Nunney 2012).
4. The benefits of outcrossing will be greater in naturally outbreeding species
than inbreeding (selfing or mixed mating) ones (Byers & Waller 1999).
5. Outbred immigrants will yield larger average benefits than inbred ones (Pickup
et al. 2013). This assumes that the populations being crossed have been
isolated for sufficient generations to have differences in their contents of
deleterious alleles. No rescue is expected if the crossed populations are
recent derivatives from the same source, such as source-sink populations.
6. The magnitude of rescue effects will be larger for composite fitness than for
individual fitness components (Frankham et al. 2010)
7. There will be little difference in the magnitude of genetic rescue among major
eukaryotic taxa.
8. Genetic rescue effects will be greater in F1 than in F2 and later generations
for zygotically determined traits, but F2 > F1 for maternally determined traits
(see Appendix S2).
2
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
Appendix S2 Inbreeding depression and genetic rescue for maternally and
zygotically determined traits
As some traits are determined by maternal genotypes, others by zygotic genotype
and total and composite fitness by a combination of these (Wright 1977; Roach &
Wulff 1987; Wolfe 1993; Montalvo 1994; Falconer & Mackay 1996), changes in both
maternal and zygotic inbreeding coefficients need to be considered both during
inbreeding and following crossing of populations (Falconer & Mackay 1996). In
general, the maternal inbreeding coefficient lags the zygotic one by a generation. For
example inbreeding depression in litter size for mice is modest until the mothers are
inbred, whilst adult survival typically exhibits inbreeding depression as soon as the
zygotes are inbred (Bowman & Falconer 1960). Similarly in plants, seed number and
weight show little inbreeding depression when maternal plants are outbred and
zygotes are inbred, but substantial inbreeding depression when maternal plants are
inbred (Wolfe 1993; Montalvo 1994).
In genetic rescue experiments, the reversal of inbreeding depression is
likewise delayed for maternally determined compared to zygotically determined traits
(Table A1) and the two inbreeding levels are not equal until the F3 generation.
Consequently, genetic rescue effects for total and composite fitness are expected to
be dependent on both maternal and zygotic ΔF values. In the F1, the maternal ΔF
(ΔFm) is zero, as the inbreeding coefficient is that of the mothers and is not changed
by crossing. Conversely the zygotic ΔF (ΔFz) is the difference between the maternal
F [½ (Fa + Fb)] and the zygotic F (0) = ½ (Fa + Fb).
Table A1 Maternal and zygotic inbreeding coefficients in F1, F2 and F3 crosses
(involving multiple individuals in each case) between (a) different inbred populations,
or (b) between inbred females and outbred males from different populations (with
random mating of the parents of each generation) and ΔFm and ΔFz values for
comparisons between crosses and the inbred parent population.
__________________________________________________________________
Maternal F
Zygotic F
ΔFm
ΔFz
__________________________________________________________________
(a) Reciprocal crosses between different inbred parental populations with inbreeding
coefficients of Fa and Fb with comparisons being made between crossed populations
and the mean of the two inbred parents
__________________________________________________________________
F1
½ (Fa + Fb)
0
0
½ (Fa + Fb)
F2
0
¼ (Fa + Fb) ½ (Fa + Fb) ¼ (Fa + Fb)
F3
¼ (Fa + Fb)
¼ (Fa + Fb) ¼ (Fa + Fb) ¼ (Fa + Fb)
(b) Crosses between inbred female and outbred male parent populations with
inbreeding coefficients of Fa and 0 with comparison being made between crossed
populations and the inbred parent
___________________________________________________________________
F1
Fa
0
0
Fa
3
125
126
127
F2
F3
0
¼ Fa
¼ Fa
¼ Fa
4
Fa
¾ Fa
¾ Fa
¾ Fa
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
Appendix S3 Additional details of data selection criteria
Within studies of the same populations, I preferentially used data on the most
inclusive fitness characters I could find or construct. Where there were data only on
multiple separate fitness components, I computed composite fitness from the
component data, as is commonly done in plant studies (e.g. Hereford 2009a).
Species with permanent translocation heterozygosity across much of their
genomes were excluded, as they exhibit little or no homozygosity or inbreeding
depression (Heiser & Shaw 2006).
I did not exclude any outbred/inbred comparisons purely on the basis of small
sample size and low statistical power to minimize bias. However, I did exclude the
F2/inbred parent comparison in the song sparrow data of Marr et al. (2002) as the
inbreeding coefficient in the F2 (0.076) was higher than that in the inbred parent
population (0.066).
In assessing the risk of outbreeding depression, adaptive differentiation,
where not directly known, was inferred from differences in biotic and/or physical
environmental between the populations (e.g. temperature, day length, soils, large
geographic distances, different host plants, and diets; (Turesson 1922; Antonovics
1976; Snaydon & Davies 1982; Bradshaw 1984; Hoffmann & Weeks 2007).
Where information was unclear, I emailed authors in an attempt to clarify details.
Assessments of the above characteristics were made on the balance of probabilities
to include as many studies as possible. Where information on one characteristic in
the outbreeding depression screen was not available (often the case for
chromosomes) the risk assessments were based on the remaining information
(potentially increasing the risk of outbreeding depression).
In minimizing pseudoreplication, I chose traditional measures of mean
populations fitness (composite fitness, or survival or fecundity), rather than
demographic measures such as population growth rate or increase in populations
size (recommended by Whiteley et al. 2015), as fitness differences may not be
detectable under conditions of density regulation when populations are at carrying
capacity (Frankham et al. 2010 p. 292), as observed, for example by Adams et al.
(2011). However, genetic rescue in traditional measures of population fitness are
expected when populations are recovering from low population sizes (Frankham et
al. 2010 p. 292).
Since there were often uncertainties in information for screening against a
high risk of outbreeding depression, especially in inbreeding levels, isolation of
populations and likelihood of adaptive differentiation of populations, data were
categorised as higher versus lower certainty of assessments, based on strength of
evidence relating to risk of outbreeding depression and the statistical power of the
data, and the two categories compared. As the two had similar characteristics and
did not differ significantly in the fitness benefits of outcrossing, analyses reported
here were completed on the combined data.
Information on variables expected to affect the magnitude of genetic rescue
effects were recorded, namely items 2-7 in Appendix S1, with the addition of number
of inbred and outcrossed populations (for computing the weighting factor used in
tests for publication bias).
Inbreeding coefficients
5
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
The inbreeding coefficients (F) in the inbred and crossed populations were based on
the probability that two alleles at a locus in an individual are identical by descent
(Malécot 1969; Falconer & Mackay 1996). F values were zygotic inbreeding
coefficients, except were indicated otherwise. When inbreeding coefficients were not
reported (often), indirect estimates were obtained where possible from allozyme or
microsatellite heterozygosity data using the equation (Frankham et al. 2010):
F = 1 – (HI/HO)
(eqn S4)
where HI is the heterozygosity of the inbred populations and Ho that of the non-inbred
comparator population. Only estimates from outbreeding species were used in
analyses.
The inbreeding coefficients for admixed populations (after two or more
generations) formed from populations with inbreeding coefficients of Fi with
proportions pi of alleles from each population were computed using the following
equation (Margan et al 1998):
Fpooled = ∑pi2Fi
(eqn S5)
This assumes that there is no selective difference between immigrant and resident
alleles, as do pedigree inbreeding coefficients.
6
198
199
Table S1. Genetic rescue (GR) data set 1 for fitness (uploaded as a separate Supporting Information Excel file).
7
200
201
202
203
204
205
206
Table S2 Genetic rescue data set for evolutionary potential (GREvP) for fitness traits
Generations
Finbredb
1.179
3.55
0.405
0.203
1.34073
Holleley et al. (2011)
1.064
12
0.165
0.03
1.16168
Invertebrate
Holleley et al. (2011)
1.036
26
0.165
0.115
1.05988
mysid shrimp
Invertebrate
Market et al. (2010)
1.270
3
0.313
0.156
1.22727
Americamysis bahia
mysid shrimp
Invertebrate
Market et al. (2010)
1.861
3
0.313
0.052
1.37879
Americamysis bahia
mysid shrimp
Invertebrate
Market et al. (2010)
1.847
3
0.313
0.039
1.39773
Species
Common name
Major taxa
Reference
Drosophila melanogaster
fruit fly
Invertebrate
Margan et al. (1998)
Drosophila melanogaster
fruit fly
Invertebrate
Drosophila melanogaster
fruit fly
Americamysis bahia
GREvP/Gena
aGenetic
Fcrossc
GDX/GDId
Comments
Comparison of Ne =
50 vs 2x50
treatments
Comparison of N =
50 migration rates of
0.04 vs 0.0025
Comparison of N =
50 migration rates of
0.01 vs 0.0025
Comparison of 1X
vs 2X treatments
Comparison of 1x vs
6X treatments
Comparison of 1x vs
8X treatments
rescue ratio per generation.
coefficient of inbred parent population(s).
cInbreeding coefficient of zygotes in crossed population.
dRatio of genetic diversity in outcrossed population to that in the inbred parent population, estimated from (1 - F
cross)/(1 – Finbred).
bInbreeding
8
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
Appendix S4 Effect size for evolutionary potential and variables expected to
affect it
Effect size
The response ratio was used as the effect size for evolutionary potential, but
was expressed on a per generation basis, as different studies had different
durations in generations (t). Selection responses (selected – control) per
generation in the outcrossed/inbred populations was converted to a per
generation basis by raising the response ratio to the power of 1/t. Response
ratios for Margan et al. (1998) data were obtained by subtracting the NaCl
concentrations at extinction for the fully inbred control populations from that of
the populations of interest. For the remaining studies where population sizes
under selection were reported after different numbers of generations (t, and t
– x), the ratio of Nt/Nt-x for the crossed populations was divided by the
corresponding ratio for inbred parent populations, and the resulting ratio
raised to the power of 1/x.
Variables affecting genetic rescue for evolutionary potential
The extent of evolutionary genetic adaptation (GA) in the short term is
predicted by the breeders’ equation GA = h2S, where h2 is the heritability and
S the selection differential (Falconer & Mackay 1996). This provides a basis
for determining the variables expected to predict genetic rescue for
evolutionary potential (GREvP), as follows:
GREvP = hX2 Sx
hI2 SI
(eqn S6)
where the subscripts I and X stand for the inbred parent and the crossed
population in F2 and later generations. The effect of crossing on heritability is
predicted to be closely related to the increase in genetic diversity, as follows:
hX2 = VAX/VPX = VAX (VPI) = ∑2pixqixaix2(VPI)
hI2 VAI/VPI
VAI (VPX) ∑2piIqiIaiI2 (VPX)
(eqn S7)
where VA is the additive genetic variation, VP the phenotypic variation, 2pq is
the heterozygosity for alleles affecting fitness, and a is half the difference in
mean between the two homozygotes at the locus. Thus, genetic rescue for
fitness will depend on the increase in heterozygosity in the crossed population
after Hardy-Weinberg equilibrium has been established, compared to that in
the inbred parent population. While VA for fitness traits often increases initially
before decreasing with inbreeding in the environment to which the population
is adapted (Willi et al. 2006), this is unlikely in the context of adaptation to new
environments, as then genetic variation for fitness is primarily additive, rather
than non-additive (Frankham et al. 1999; England et al. 2003; Frankham et al.
2011). The benefits of gene flow on evolutionary potential should depend
upon the proportionate increase in genetic diversity for fitness, if there is no
outbreeding depression. In the current context this proportion should typically
be similar to that for neutral markers (Briscoe et al. 1992; Frankham et al.
9
257
258
259
260
261
1999; England et al. 2003; Gilligan et al. 2005). It there is heterozygosity data,
then GREvP should be related to the ratio of
heterozygositycross/heterozygosityparent (GDX/GDI) while if there are only
inbreeding coefficient, the ratio can be estimated as (1- FX)/(1 - FP). The latter
expression has been used to determine the ratios in this paper.
10
262
263
Table S3 Characteristics of the nine studies that reported deleterious effects of outcrossing.
Species
Mating
system
GR
Caenorhabditis
elegans
Selfing
0.90
Inbreeding
difference
ΔF
≥ 0.75
Environment
Fitness
measure
Statistical
power
Gen
Details
Benign
Total
Good
F1
Two F1 crosses of
populations were mildly
deleterious, compared to the
inbred parent strains
collected from <10cm and
15m apart. Convincing
evidence for mild
outbreeding depression
A panmictic population
formed from four isolated Ne
= 100 populations (at an
advanced generation) was
mildly deleterious compared
to the mean of the parent
populations, when assessed
in benign conditions, but
strongly beneficial under
stressful conditions (GR =
2.14). Other related crosses
of replicates of different Ne
treatments were also
beneficial.
Five lines subject to a single
generation of full-sib mating
had a slightly higher
average fitness than their 10
F1’s. Two other related
treatments from the same
base population with higher
F’s showed beneficial
Drosophila
melanogaster
Outbreeding
0.91
0.18
Benign
Composite
Low
>F2
Drosophila
melanogaster
Outbreeding
(but inbred
base
population)
0.98
0.25
(0.0375)a
Benign
Egg-adult
survival
Low
F1
11
References
Dolgin et al.
(2007)
Woodworth et al.
(2002)
Tantawy (1957);
Falconer (1954)
Drosophila
melanogaster
Outbreeding
(but inbred
base
population)
1.00
0.51
(0.0765)a
Benign
Egg-adult
survival
Low
F1
Lymnaea
stagnalis
Mixed mating
0.86
0.0625
Benign
Composite
Low
F1
12
effects of crossing. The
base population was highly
inbred (F > 0.85) before this
experiment commenced, as
it originated from a single
inseminated wild females
and was maintained with a
small population size (~ 20
pairs/generation) for ~ 6
years (~ 150 generations).
Thus, the ΔF with respect to
similar inbreeding from an
outbred base population is
effectively only ~ 0.04).
Five lines inbred by double
first-cousin mating had a
slightly lower fitness than
their 10 F1’s under benign
conditions. Another related
treatment with higher F
showed beneficial effects of
outcrossing. The base
population was already
highly inbred (F > 0.85)
before this experiment
commenced (see above).
The effective ΔF is only ~
0.08.
Six F1 crosses of four Ne = 8
bottlenecked populations
were mildly deleterious
compared to their parent
lines, but crosses of a
related treatment with Ne = 5
were beneficial compared to
Tantawy (1957);
Falconer (1954)
Coutellec &
Caquet (2011)
Musca
domestica
Outbreeding
0.95
< 0.20
Benign
Egg-adult
survival
Low
>F2
Diodia teres
Mixed mating
0.94
≥ 0.76 (FIS)
Benign
Pollination
success
Low
F1
Echinacea
angustifolia
Selfincompatible
0.89
NAb
Field
Composite
Low
F1
Viola stagnina
Mixed mating
0.91
NA
Benign
Seeds/capsule
Very low
F1
13
their parent populations.
A comparison of isolated
inbred populations versus
the same sized populations
with low migration (0.025)
every generation exhibited a
slightly lower fitness in the
migration treatment at
generation 5, but a strongly
beneficial effect at
generation 24, as were both
assessments for a treatment
with a high immigration rate.
Pollination success in F1
crosses was variable, but
mildly deleterious on
average, compared to
parent populations
(variable), but composite
fitness in F2 crosses in the
field were beneficial
compared to inbred parents.
Inbred parents had slightly
higher average fitness than
their F1 but the comparison
was strongly affected by one
major outlier in within
population cross with about
double the achene count of
the next highest. Many
plants were alive but had not
yet flowered (much of the
life-cycle was not included).
Cross of two populations
had slightly lower fitness
Backus et al.
(1995); Bryant et
al. (1999)
Hereford
(2009b)
Wagenius et al.
(2010)
Eckstein & Otte
(2005)
than the F1, but this was a
very small study (9 plants),
with no information
presented on the habitats of
the populations.
264
265
266
aThe
numbers in brackets are the equivalent inbreeding coefficients from similar inbreeding in an outbred base population.
not available
bInformation
14
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
Appendix S5 Additional consideration of results
Mating system effect on genetic rescue
When mating system was tested at a finer scale, the fitness benefit was 78%
in self-incompatible species, 59% in other outbreeders, 16% in mixed mating
and 39% in selfers (Kruskall-Wallis H = 10.76, df = 3, P = 0.013). While the
latter two are in reverse order compared to expectations, the numbers of data
points were only 6 for selfing species.
Difference between F1 and F2 genetic rescue effects
An alternative explanation for F2 > F1 genetic rescue effects is that the
inbreeding levels and ΔF values might be higher for studies reporting F2 than
F1 data. This can be excluded, as differences are in the opposite direction to
this. Mean F values in inbred parental populations were 0.615 and 0.440 for
studies with F1 and F2 data, respectively, while corresponding ΔFz values
were 0.567 and 0.261.
ΔFzygotic and ΔFmaternal effects
Conclusion about the predictive power of ΔFz depended on whether
regression models fitted an intercept or not, while ∆Fm was significantly
supported irrespective of whether the single factor regressions were fitted with
an intercept, or through the origin. For both simple regressions and both AICc
model selection analyses the intercept terms were significant, so I
preferentially reported analyses with intercepts fitted, despite tests of
inbreeding effects being more powerful when regressed through the origin
(e.g. Montgomery et al. 2010).
In simple linear regressions using data set 2 with an intercept fitted, ∆Fm
was a significant predictor of ln GR for outbreeding species (Fig. 1: b = 0.563
+ 0.298, P = 0.032, n = 69), whilst ΔFz was not (b = - 0.109 + 0.177, P =
0.770, n = 68). There is much lower power associated with ∆Fm, compared to
∆Fz, as it is only F2 and later generation data (26 of 69) that have non-zero
values of ∆Fm, while all ∆Fz are > 0.
Single fitness component versus composite fitness effect
The prediction that genetic rescue effects would be greater for composite
fitness than its components was not supported in either single variables tests
(Table 1) or AICc model selection with multiple variables (Table 2). This is, at
least partially, an artefact due to a low proportion of self-incompatible species
having composite fitness data, but predominantly showing large genetic
rescue effects.
Best fitting AICc model
The best fitting model as defined by Akaike model selection in Table 2 was:
ln GR = 0.4468 + 0.587 Environment
15
317
318
319
320
321
322
323
324
325
326
327
328
329
where environments are coded 0 benign and 1 stressful.
Effect of outbred versus inbred immigrants on genetic rescue
Gene flow from both outbred and inbred donors were beneficial (GR of 2.136
and 1.519, respectively), but the benefits were 40.6% greater with outbred
than inbred immigrants. The test for effects of outbred versus inbred
immigrants was supported by the non-parametric single variable test, but not
by either of the AICc model selections. These tests had low statistical power
as differences are only expected in F2 and later generations and there were
very few such data points for outbred immigrants beyond the F1.
16
330
331
332
333
334
Fig. S1 Histograms of natural logarithm of genetic rescue ratio (ln GR) for
composite fitness in outbreeding species in benign versus stressful
environments. Medians for ln GR and de-transformed GR values are also
presented.
Benign
Median = 0.412
GR = 1.509
40
Percent
30
20
10
0
0
1
2
3
4
5
ln GR
335
40
Stressful
Median 0.954
GR 2.625
Percent
30
20
10
0
0
1
2
3
ln GR
336
337
338
17
4
5
339
340
341
342
Table S4 Variables affecting the magnitude of genetic rescue for composite
fitness based on model selection statistics using the Akaike AICc procedure
on dataset 5 (n = 29). The best fitting model is bolded.
Model
C, environment,
σ2
C, ΔFm,
environment, σ2
C, ΔFm,
environment,
immigrant source,
σ2
C, ΔFm, ΔFz,
environment,
immigrant source,
σ2
C, σ2
Ki
AICc
3 - 48.090
Δi
0
wi
0.6580
4
-46.092
1.998
0.2423
5
- 43.884
4.206
0.0803
6
- 40.674
7.416
0.0161
2
- 37.439
10.661
0.0032
Ki = number of parameters estimated, AICc = Akaike’s information criteria
adjusted for small sample size, Δi = deviation of the model from the best fitting
model, and wi = Akaike weights (approximate probability that the model is the
best information theoretic one), C is the constant (intercept), ΔFm and ΔFz are
the difference in maternal and zygotic inbreeding coefficient between crossed
and inbred populations.
.
343
344
345
18
346
347
348
349
350
Fig. S2 Funnel plot for all genetic rescue data for fitness (ln GR) from data set
1 against logarithm of the sample size weighting factor (log w) with line of best
fit inserted (the slope is non-significant).
351
352
19
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
Table S5 Mean genetic rescue ratios (GR: F1/inbred parents) for outcrosses
of inbred parental populations for fitness traits in several domestic plant and
animal species
Species
Plants
Maize
Sorghum
Cotton
Wheat
Barley
Tomato
Animals
Poultry
Swine
GR
Trait
Mating system
Reference
2.90
2.00
1.48
1.29
1.32
1.45
grain yield
grain yield
grain yield
grain yield
grain yield
fruit yield
outcrossing
mainly selfing
selfer
selfer
selfer
selfer
Sinha & Khanna (1975)
Sinha & Khanna (1975)
Sinha & Khanna (1975)
Sinha & Khanna (1975)
Sinha & Khanna (1975)
Williams & Gilbert (1960)
1.22
eggs to
500 days
litter weight
at 154 days
outbreeding
Shoffner et al. (1966)
outbreeding
Dickerson et al. (1946)
1.79
20
373
374
375
376
377
378
379
380
381
382
383
Appendix S6 Additional Discussion
On the basis that genetic rescue is the reversal of inbreeding depression (see
Introduction), the proportion of studies reporting beneficial effects of
outcrossing should equal the proportion of studies reporting inbreeding
depression (given similar sized studies and environmental regimes). This
prediction was supported, as 92.9% of studies herein exhibited beneficial
effects following outcrossing of inbred populations, compared to inbreeding
depression for fitness traits in 93.2% of captive mammal populations (Ralls
and Ballou 1983) and 91.2% for a broad array of animal and plant taxa in wild
environments (Crnokrak & Roff 1999).
21
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
Appendix S7 Beneficial effects of outcrossing on evolutionary potential for
traits peripherally related to fitness
Benefits of crossing on evolutionary potential for traits peripherally related to
fitness have also been reported for sternopleural bristle number in Drosophila
(Robertson 1969; Swindell & Bouzat 2006), body weight in mice (Falconer &
King 1953; Roberts 1967 [only in a line selected for high, not in a line selected
for low weight]; Eisen 1975) and nesting-building behavior in mice (Bult &
Lynch 2000). In the Robertson (1969) study (the largest one), response to
selection in crossed lines exceeded that of the mean of the parent lines in
27/30 cases. The crosses that failed to exceed the parents in response or only
gave limited response typically involved conditions where there was little
differentiation between crossed populations and little effective augmentation
of genetic diversity, as expected.
22
399
400
401
402
403
404
Appendix S8 Additional discussion of genetic rescue guidelines
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
(i)
Jonathan D. Ballou and Robert C. Lacy (pers. comm.) are currently working
on means for managing gene flow in fragmented populations.
My guidelines for genetic rescue are less stringent, than those of
Edmands (2007) and Hedrick and Fredrickson (2010), due to:
the urgency to act promptly, as conservation biology is a crisis
discipline (Soulé 1985) where inaction is often a recipe for further
decline and extinction of threatened populations for genetic and nongenetic reasons,
(ii)
the ubiquity of inbreeding depression in adequately studied naturally
outbreeding diploid species, due to either recent matings between
relatives, and/or longer term genetic drift effects due to increase in
frequency and fixation of deleterious alleles plus fixation of alleles at
loci exhibiting heterozygote advantage (the latter issues are also
referred to as increased genetic load) (Lacy 1997; Frankham et al.
2014),
(iii)
the low power of most experiments on threatened populations to
detect inbreeding or outbreeding depression, even if it truly exists (e.g.
Lacy 1997; Kalinowski & Hedrick 1999), while meta-analyses will
usually give a more reliable overview of likely effects. For example,
Ralls & Ballou (1983) reported that 41/44 mammal populations had
high juvenile mortality for inbred than outbred progeny, but only 34%
of the comparisons were significantly different. Further, 90% of the
valid data sets in the Crnokrak & Roff (1999) meta-analysis indicated
that inbreeding reduced means, but only 53% were significant.
(iv) the low power of experiments on genetic rescue of threatened,
populations in captivity, the long-time involved and possible difficulties
in getting permission to do experiments on them. First, an experiment
in captivity would likely involve only a single replicate with low sample
sizes and would likely yield a lower benefit or disadvantage than in
wild conditions. Second, this approach is not feasible for species that
have not been successfully bred in captivity (e.g. northern hairy-nosed
wombats). Third, to be convincing the experiment would need to go to
at least the F3 generation (Appendix S2) and to encompass the full life
cycle for that generation, leading to substantial delays during which
the wild population may well be extirpated. This would not be feasible
in large long-lived species such as elephants, or long-lived trees
where > 100 years would be required,
(v)
the development of effective means to predict the risk of outbreeding
depression (Frankham et al. 2011), and
(vi) evidence of consistent and highly beneficial effects of outcrossing on
fitness and evolutionary potential documented herein.
Edmands (2007) recommended that genetic rescue only proceed when there
is clear evidence of inbreeding depression and where evaluations have
already been done of crosses over two generations (wherever possible). I
disagree with these recommendations for all five of the reasons above.
Edmands (2007) also recommended that the population(s) chosen for
outcrossing be as similar genetically and adaptively as possible. I agree with
23
448
449
450
451
452
453
454
455
456
457
458
459
460
the need for adaptive similarity (Frankham et al. 2011), but query the
requirement for genetic similarity, as theory predicts that the benefits of
outcrossing increase with differences in allele frequencies between target and
donor populations (Appendix S1).
My guidelines are closer to those of Hedrick & Fredrickson (2010) but
also less stringent than theirs. For example, I have stronger views about the
urgency of action to avoid population extinction, and do not require
experimental data from captive populations prior to proceeding, as
outcrossing of inbred populations is consistently beneficial in such species as
shown herein. I also place more emphasis on the mating system and modes
of inheritance of the species involved.
24
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
References for Supporting Information
Adams JR, Vucetich LM, Hedrick PW, Peterson RO, Vucetich JA (2011)
Genomic sweep and potential genetic rescue during limiting
environmental conditions in an isolated wolf population. Proceedings of
the Royal Society B-Biological Sciences 278, 3336-3344.
Antonovics J (1976) The nature of limits to natural selection. Annals of the
Missouri Botanic Garden 63, 224-227.
Asa C, Miller P, Agnew M, et al. (2007) Relationship of inbreeding with sperm
quality and reproductive success in Mexican gray wolves. Animal
Conservation 10, 326-331.
Backus VL, Bryant EH, Hughes CR, Meffert LM (1995) Effect of migration or
inbreeding followed by selection on low-founder-number populations:
Implications for captive breeding. Conservation Biology 9, 1216-1224.
Bailey MF, McCauley DE (2006) The effects of inbreeding, outbreeding and
long-distance gene flow on survivorship in North American populations
of Silene vulgaris. Journal of Ecology 94, 98-109.
Barrett SC, Charlesworth D (1991) Effect of a change in the level of
inbreeding on the genetic load. Nature 352, 522-524.
Benson JF, Hostetler JA, Onorato DP, et al. (2011) Intentional genetic
introgression influences survival of adults and subadults in a small,
inbred felid population. Journal of Animal Ecology 80, 958-967.
Bijlsma R, Westerhof MDD, Roekx LP, Pen I (2010) Dynamics of genetic
rescue in inbred Drosophila melanogaster populations. Conservation
Genetics 11, 449-462.
Bilde T, Maklakov AA, Schilling N (2007) Inbreeding avoidance in spiders:
evidence for rescue effect in fecundity of female spiders with
outbreeding opportunity. Journal of Evolutionary Biology 20, 12371242.
Boessenkool S, Taylor SS, Tepolt CK, Komdeur J, Jamieson IG (2007) Large
mainland populations of South Island robins retain greater genetic
diversity than offshore island refuges. Conservation Genetics 8, 705714.
Bonnier G (1961) Experiments on hybrid superiority in Drosophila
melanogaster. I. Egg laying capacity and larval survival. Genetics 46,
9-24.
Bossuyt B (2007) Genetic rescue in an isolated metapopulation of a naturally
fragmented plant species, Parnassia palustris. Conservation Biology
21, 832-841.
Bouzat JL, Johnson JA, Toepfer JE, et al. (2009) Beyond the beneficial effects
of translocations as an effective tool for the genetic restoration of
isolated populations. Conservation Genetics 10, 191-201.
Bowman JC, Falconer DS (1960) Inbreeding depression and heterosis of litter
size in mice. Genetical Research 1, 262-274.
Bradshaw AD (1984) The importance of evolutionary ideas in ecology - and
vice versa. In: Evolutionary Ecology (ed. Shorrocks B), pp. 1-25.
Blackwell, Oxford.
Briscoe DA, Malpica JM, Robertson A, et al. (1992) Rapid loss of genetic
variation in large captive populations of Drosophila flies: Implications
25
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
for the genetic management of captive populations. Conservation
Biology 6, 416-425.
Bryant EH, Backus VL, Clark ME, Reed DH (1999) Experimental tests of
captive breeding for endangered species. Conservation Biology 13,
1487-1496.
Bult A, Lynch C (2000) Breaking through artificial selection limits of an
adaptive behavior in mice and the consequences for correlated
responses. Behavior Genetics 30, 193-206.
Busch JW (2006) Heterosis in an isolated, effectively small, and self-fertilizing
population of the flowering plant Leavenworthia alabamica. Evolution
60, 184-191.
Byers DL (1998) Effect of cross proximity on progeny fitness in a rare and a
common species of Eupatorium (Asteraceae). American Journal of
Botany 85, 644-653.
Byers DL, Waller DM (1999) Do plant populations purge their genetic load?
Effects of population size and mating history on inbreeding depression.
Annual Review of Ecology and Systematics 30, 479-513.
Charpentier A, Grillas P, Thompson JD (2000) The effects of population size
limitation on fecundity in mosaic populations of the clonal macrophyte
Scirpus maritimus (Cyperaceae). American Journal of Botany 87, 502507.
Chaston JM, Dillman AR, Shapiro-Ilan DI, et al. (2011) Outcrossing and
crossbreeding recovers deteriorated traits in laboratory cultured
Steinernema carpocapsae nematodes. International Journal for
Parasitology 41, 801-809.
Coutellec MA, Caquet T (2011) Heterosis and inbreeding depression in
bottlenecked populations: a test in the hermaphroditic freshwater snail
Lymnaea stagnalis. Journal of Evolutionary Biology 24, 2248-2257.
Crnokrak P, Roff DA (1999) Inbreeding depression in the wild. Heredity 83,
260-270.
Darwin C (1876) The Effects of Cross and Self Fertilisation in the Vegetable
Kingdom John Murray, London.
Demauro MM (1993) Relationship of breeding system to rarity in the Lakeside
daisy (Hymenoxys acaulis var. glabra). Conservation Biology 7, 542550.
Dickerson GE, Lush JL, Culbertson CC (1946) Hybrid vigor in single crosses
between inbred lines of Poland China swine. Journal of Animal Science
5, 16-24.
Dolgin ES, Charlesworth B, Baird SE, Cutter AD (2007) Inbreeding and
outbreeding depression in Caenorhabditis nematodes. Evolution 61,
1339-1352.
Driscoll CA, Menotti-Raymond M, Nelson G, Goldstein D, O'Brien SJ (2002)
Genomic microsatellites as evolutionary chronometers: A test in wild
cats. Genome Research 12, 414-423.
Dudash MR (1990) Relative fitness of selfed and outcrossed progeny in a selfcompatible, protandrous species, Sabatia angularis L. (Gentianaceae):
A comparison of three environments. Evolution 44, 1129-1139.
Dudash MR, Carr DE, Fenster CB (1997) Five generations of enforced selfing
and outcrossing in Mimulus guttatus: Inbreeding depression variation at
the population and family level. Evolution 51, 54-65.
26
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
Eckstein RL, Otte A (2005) Effects of cleistogamy and pollen source on seed
production and offspring performance in three endangered violets.
Basic and Applied Ecology 6, 339-350.
Edmands S (2007) Between a rock and a hard place: evaluating the relative
risks of inbreeding and outbreeding depression for conservation and
management. Molecular Ecology 16, 463-475.
Ehiobu NG, Goddard ME, Taylor JF (1990) Predicting heterosis in crosses
between inbred lines of Drosophila melanogaster. Theoretical and
Applied Genetics 80, 321-325.
Eisen EJ (1975) Population size and selection intensity effects on long-term
selection response in mice. Genetics 79, 305-323.
Enders LS, Nunney L (2012) Seasonal stress drives predictable changes in
inbreeding depression in field-tested captive populations of Drosophila
melanogaster. Proceedings of the Royal Society B 279, 3756-3764.
England PR, Osler GHR, Woodworth LM, et al. (2003) Effect of intense versus
diffuse population bottlenecks on microsatellite genetic diversity and
evolutionary potential. Conservation Genetics 4, 595-604.
Eriksson O, Bremer B (1993) Genet dynamics of the clonal plant Rubus
saxatilis. Journal of Ecology 81, 533-542.
Escobar JS, Nicot A, David P (2008) The different sources of variation in
inbreeding depression, heterosis and outbreeding depression in a
metapopulation of Physa acuta. Genetics 180, 1593-1608.
Falconer DS (1954) Selection for sex ratio in mice and Drosophila. American
Naturalist 88, 385-397.
Falconer DS, King JWB (1953) A study of selection limits in the mouse.
Journal of Genetics 51, 561-579.
Falconer DS, Mackay TFC (1996) Introduction to Quantitative Genetics,
Fourth edn. Longman, Harlow, England.
Fenster CB, Galloway LF (2000) Inbreeding and outbreeding depression in
natural populations of Chamaecrista fasciculata (Fabaceae).
Conservation Biology 14, 1406-1412.
Finger A, Kettle CJ, Kaiser-Bunbury CN, et al. (2011) Back from the brink:
potential for genetic rescue in a critically endangered tree. Molecular
Ecology 20, 3773-3784.
Forney KA, Gilpin ME (1989) Spatial structure and population extinction: A
study with Drosophila flies. Conservation Biology 3, 45-51.
Fox CW, Reed DH (2011) Inbreeding depression increases with
environmental stress: An experimental study and meta-analysis.
Evolution 65, 246-258.
Frankham R, Lees K, Montgomery ME, et al. (1999) Do population size
bottlenecks reduce evolutionary potential? Animal Conservation 2, 255260.
Frankham R, Ballou JD, Briscoe DA (2010) Introduction to Conservation
Genetics, 2nd edition Cambridge University Press, Cambridge, U.K.
Frankham R, Ballou JD, Eldridge MDB, et al. (2011) Predicting the probability
of outbreeding depression. Conservation Biology 25, 465-475.
Frankham R, Bradshaw CJA, Brook BW (2014) Genetics in conservation
management: Revised recommendations for the 50/500 rules, Red List
criteria and population viability analyses. Biological Conservation 170,
56–63.
27
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
Fredrickson RJ, Siminski P, Woolf M, Hedrick PW (2007) Genetic rescue and
inbreeding depression in Mexican wolves. Proceedings of the Royal
Society B-Biological Sciences 274, 2365-2371.
Galloway LF, Fenster CB (2000) Population differentiation in an annual
legume: Local adaptation. Evolution 54, 1173-1181.
Garcia-Moreno J, Matocq MD, Roy MS, Geffen E, Wayne RK (1996)
Relationships and genetic purity of the endangered Mexican wolf based
on analysis of microsatellite loci. Conservation Biology 10, 376-389.
Gilligan DM, Briscoe DA, Frankham R (2005) Comparative losses of
quantitative and molecular genetic variation in finite populations of
Drosophila melanogaster. Genetical Research 85, 47-55.
Gitzendanner MA, Weekley CW, Germain-Aubrey CC, Soltis DE, Soltis PS
(2012) Microsatellite evidence for high clonality and limited genetic
diversity in Ziziphus celata (Rhamnaceae), an endangered, selfincompatible shrub endemic to the Lake Wales Ridge, Florida, USA.
Conservation Genetics 13, 223-234.
Groom MJ, Preuninger TE (2000) Population type can influence the
magnitude of inbreeding depression in Clarkia concinna (Onagraceae).
Evolutionary Ecology 14, 155-180.
Hauser TP, Loeschcke V (1994) Inbreeding depression and mating-distance
dependent offspring fitness in large and small populations of Lychnis
flos-cuculi (Caryophyllaceae). Journal of Evolutionary Biology 7, 609622.
Heber S, Briskie JV, Apiolaza LA (2012) A test of the 'genetic rescue'
technique using bottlenecked donor populations of Drosophila
melanogaster. PLoS One 7, e43113.
Heber S, Varsani A, Kuhn S, et al. (2013) The genetic rescue of two
bottlenecked South Island robin populations using translocations of
inbred donors. Proceedings of the Royal Society B-Biological Sciences
280, 20122228.
Hedgecock D, McGoldrick DJ, Bayne BL (1995) Hybrid vigor in Pacific
oysters: An experimental approach using crosses among inbred lines.
Aquaculture 137, 285-298.
Hedrick PW, Fredrickson RJ (2008) Captive breeding and the reintroduction of
Mexican and red wolves. Molecular Ecology 17, 344-350.
Hedrick PW, Frederickson R (2010) Genetic rescue guidelines with examples
from Mexican wolves and Florida panthers. Conservation Genetics 11,
615-626.
Heiser DA, Shaw RG (2006) The fitness effects of outcrossing in Calylophus
serrulatus, a permanent translocation heterozygote. Evolution 60, 6476.
Heliyanto B, Veneklaas EJ, Lambers H, Krauss SL (2005) Preferential
outcrossing in Banksia ilicifolia (Proteaceae). Australian Journal of
Botany 53, 163-170.
Hereford J (2009a) A quantitative survey of local adaptation and fitness tradeoffs. American Naturalist 173, 579-588.
Hereford J (2009b) Postmating/prezygotic isolation, heterosis, and
outbreeding depression in crosses within and between populations of
Diodia teres (Rubiaceae) Walt. International Journal of Plant Sciences
170, 301-310.
28
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
Heschel MS, Paige KN (1995) Inbreeding depression, environmental stress,
and population size variation in scarlet gilia (Ipomopsis aggregata).
Conservation Biology 9, 126-133.
Hoffmann AA, Weeks AR (2007) Climatic selection on genes and traits after a
100 year-old invasion: a critical look at the temperate-tropical clines in
Drosophila melanogaster from eastern Australia. Genetica 129, 133147.
Hogg JT, Forbes SH, Steele BM, Luikart G (2006) Genetic rescue of an
insular population of large mammals. Proceedings of the Royal Society
of London B: Biological Sciences 273, 1491-1499.
Holleley CE, Nichols RA, Whitehead MR, et al. (2011) Induced dispersal in
wildlife management: experimental evaluation of the risk of hybrid
breakdown and the benefit of hybrid vigor in the F1 generation.
Conservation Genetics 12, 31-40.
Honnay O, Jacquemyn H, Roldan-Ruiz I, Hermy M (2006) Consequences of
prolonged clonal growth on local and regional genetic structure and
fruiting success of the forest perennial Maianthemum bifolium. Oikos
112, 21-30.
Hostetler JA, Onorato DP, Bolker BM, et al. (2012) Does genetic introgression
improve female reproductive performance? A test on the endangered
Florida panther. OecologiaM. 168, 289-300.
Hostetler JA, Onorato DP, Nichols JD, et al. (2010) Genetic introgression and
the survival of Florida panther kittens. Biological Conservation 143,
2789-2796.
Johnson WE, Onorato DP, Roelke ME, et al. (2010) Genetic Restoration of
the Florida Panther. Science 329, 1641-1645.
Kalinowski ST, Hedrick PW (1999) Detecting inbreeding depression is difficult
in captive endangered species. Animal Conservation 2, 131-136.
Lacy RC (1997) Importance of genetic variation to the viability of mammalian
populations. Journal of Mammalogy 78, 320-335.
La Haye MJJ, Koelewijn HP, Siepel H, Verwimp N, Windig JJ (2012) Genetic
rescue and the increase of litter size in the recovery breeding program
of the common hamster (Cricetus cricetus) in the Netherlands.
Relatedness, inbreeding and heritability of litter size in a breeding
program of an endangered rodent. Hereditas 149, 207-216.
Latter BDH, Mulley JC (1995) Genetic adaptation to captivity and inbreeding
depression in small laboratory populations of Drosophila melanogaster.
Genetics 139, 255-266.
Laverty TM, Plowright RC (1988) Fruit and seed set in mayapple
(Podophyllum peltatum): influence of intraspecific factors and local
enhancement near Pedicularis canadensis. Canadian Journal of
Botany 66, 173-178.
Lynch M, Walsh B (1998) Genetics and Analysis of Quantitative Traits
Sinauer, Sunderland, MA.
Madsen T, Stille B, Shine R (1996) Inbreeding depression in an isolated
population of adders Vipera berus. Biological Conservation 75, 113118.
Madsen T, Shine R, Olsson M, Wittzell H (1999) Restoration of an inbred
adder population. Nature 402, 34-35.
29
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
Madsen T, Ujvari B, Olsson M (2004) Novel genes continue to enhance
population growth in adders (Vipera berus). Biological Conservation
120, 145-147.
Malécot G (1969) The Mathematics of Heredity W.H. Freeman and Company,
San Francisco.
Malick LE, Kidwell JF (1966) The effect of mating status, sex and genotype on
longevity in Drosophila melanogaster. Genetics 54, 203-209.
Margan SH, Nurthen RK, Montgomery ME, et al. (1998) Single large or
several small? Population fragmentation in the captive management of
endangered species. Zoo Biology 17, 467-480.
Markert JA, Champlin DM, Gutjahr-Gobell R, et al. (2010) Population genetic
diversity and fitness in multiple environments. BMC Evolutionary
Biology 10, 205.
Marr AB, Keller LF, Arcese P (2002) Heterosis and outbreeding depression in
descendants of natural immigrants to an inbred population of song
sparrows (Melospiza melodia). Evolution 56, 131-142.
Mattila ALK, Duplouy A, Kirjokangas M, et al. (2012) High genetic load in an
old isolated butterfly population. Proceedings of the National Academy
of Sciences of the United States of America 109, E2496-E2505.
Montalvo AM (1994) Inbreeding depression and maternal effects in Aquilegia
caerulea, a partially selfing plant. Ecology 75, 2395-2409.
Montgomery ME, Woodworth LM, England PR, Briscoe DA, Frankham R
(2010) Widespread selective sweeps affecting microsatellites in
Drosophila populations adapting to captivity: Implications for captive
breeding programs. Biological Conservation 143, 1842-1849.
Morton NE, Crow JF, Muller HJ (1956) An estimate of the mutational damage
in man from data on consanguineous marriages. Proceedings of the
National Academy of Sciences, USA 42, 855-863.
Mullarkey AA, Byers DL, Anderson RC (2013) Inbreeding depression and
partitioning of genetic load in the invasive biennial Alliaria petiolata
(Brassicaceae). American Journal of Botany 100, 509-518.
Newman D, Tallmon DA (2001) Experimental evidence for beneficial fitness
effects of gene flow in recently isolated populations. Conservation
Biology 15, 1054-1063.
Oakley CG, Winn AA (2012) Effects of population size and isolation on
heterosis, mean fitness, and inbreeding depression in a perennial plant.
New Phytologist 196, 261-270.
Oostermeijer JGB, Altenburg RGM, Den Nijs HCM (1995) Effects of
outcrossing distance and selfing on fitness components in the rare
Gentiana pneumonanthe (Gentianaceae). Acta Botanica Neerlandica
44, 257-268.
Ouborg NJ, Van Treuren R (1994) The significance of genetic erosion in the
process of extinction. IV. Inbreeding load and heterosis in relation to
population-size in the mint Salvia pratensis. Evolution 48, 996-1008.
Paland S, Schmid B (2003) Population size and the nature of genetic load in
Gentianella germanica. Evolution 57, 2242-2251.
Parsons PA (1964) A diallel cross for mating speeds in Drosophila
melanogaster. Genetica 35, 141-151.
30
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
Peer K, Taborsky M (2005) Outbreeding depression, but no inbreeding
depression in haplodiploid ambrosia beetles with regular sib mating.
Evolution 59, 317-323.
Pekkala N, Knott KE, Kotiaho JS, Nissinen K, Puurtinen M (2012) The
benefits of interpopulation hybridization diminish with increasing
divergence of small populations. Journal of Evolutionary Biology 25,
2181-2193.
Pickup M (2008) Local Adaptation and Outbreeding Depression in
Fragmented Populations of Rutidosis leptorrhynchoides (Asteraceae)
PhD thesis, Australian National University.
Pickup M, Young AG (2008) Population size, self-incompatibility and genetic
rescue in diploid and tetraploid races of Rutidosis leptorrhynchoides
(Asteraceae). Heredity 100, 268-274.
Pickup M, Field DL, Rowell DM, Young AG (2013) Source population
characteristics affect heterosis following genetic rescue of fragmented
plant populations. Proceedings of the Royal Society B: Biological
Sciences 280.
Pisanu PC, Gross CL, Flood L (2009) Reproduction in wild populations of the
threatened tree Macadamia tetraphylla: Interpopulation pollen enriches
fecundity in a declining species. Biotropica 41, 391-398.
Polans NO, Allard RW (1989) An experimental evaluation of the recovery
potential of ryegrass populations from genetic stress resulting from
restriction of population size. Evolution 43, 1320-1324.
Raijmann LEL, Van Leeuwan NC, Kersten R, et al. (1994) Genetic variation
and outcrossing rate in relation to population size in Gentiana
pneumonanthe L. Conservation Biology 8, 1014-1026.
Ralls K, Ballou J (1983) Extinction: lessons from zoos. In: Genetics and
Conservation: A Reference for Managing Wild Animal and Plant
Populations (eds. Schonewald-Cox CM, Chambers SM, MacBryde B,
Thomas L), pp. 164-184. Benjamin/Cummings, Menlo Park, CA.
Ralls K, Ballou JD, Templeton A (1988) Estimates of lethal equivalents and
the cost of inbreeding in mammals. Conservation Biology 2, 185-193.
Richards CM (2000) Inbreeding depression and genetic rescue in a plant
metapopulation. American Naturalist 155, 383-394.
Roach DA, Wulff RD (1987) Maternal effects in plants. Annual Review of
Ecology and Systematics 18, 209-235.
Roberts RC (1967) The limits to artificial selection for body weight in the
mouse: III. Selection from crosses between previously selected lines.
Genetics Research 9, 73-85.
Robertson DE (1969) Selection in small populations. PhD thesis, University of
New South Wales, Sydney.
Robertson FW, Reeve ECR (1955) Studies in quantitative inheritance. VIII.
Further analyses of heterosis in crosses between inbred lines.
Zeitschrift fur indukt. Abstammungs- und Vererbungslehre 86, 439-458.
Roff DA (2002) Inbreeding depression: Tests of the overdominance and
partial dominance hypotheses. Evolution 56, 768-775.
Saccheri IJ, Lloyd HD, Helyar SJ, Brakefield PM (2005) Inbreeding uncovers
fundamental differences in the genetic load affecting male and female
fertility in a butterfly. Proceedings of the Royal Society B-Biological
Sciences 272, 39-46.
31
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
Santiago E, Dominguez A, Albornoz J, Pineiro R, Izquierdo JI (1989)
Environmental sensitivity and heterosis for egg-laying in Drosophila
melanogaster. Theoretical and Applied Genetics 78, 243-248.
Schwartz MK, Mills LS (2005) Gene flow after inbreeding leads to higher
survival in deer mice. Biological Conservation 123, 413-420.
Scobie AR, Wilcock CC (2009) Limited mate availability decreases
reproductive success of fragmented populations of Linnaea borealis, a
rare, clonal self-incompatible plant. Annals of Botany 103, 835-846.
Severns P (2003) Inbreeding and small population size reduce seed set in a
threatened and fragmented plant species, Lupinus sulphureus ssp
kincaidii (Fabaceae). Biological Conservation 110, 221-229.
Sexton JP, Strauss SY, Rice KJ (2011) Gene flow increases fitness at the
warm edge of a species' range. Proceedings of the National Academy
of Sciences of the United States of America 108, 11704-11709.
Sheridan PM, Karowe DN (2000) Inbreeding, outbreeding, and heterosis in
the yellow pitcher plant, Sarracenia flava (Sarraceniaceae), in Virginia.
American Journal of Botany 87, 1628-1633.
Shoffner RN, Narayanan S, Pilkey AM (1966) The potential of the advanced
generation cross. Proceedings of the 13th World's Poultry Congress,
Section Papers, pp. 36-41.
Sinha SK, Khanna R (1975) Physiological, biochemical, and genetic basis of
heterosis. Advances in Agronomy 27, 123-174.
Snaydon RW, Davies TM (1982) Rapid divergence of plant populations in
response to recent changes in soil conditions. Evolution 36, 289-297.
Soulé ME (1985) What is conservation biology? Bioscience 35, 727-734.
Spielman D, Frankham R (1992) Modeling problems in conservation genetics
using captive Drosophila populations: Improvement in reproductive
fitness due to immigration of one individual into small partially inbred
populations. Zoo Biology 11, 343-351.
Swindell W, Bouzat J (2006) Gene flow and adaptive potential in Drosophila
melanogaster. Conservation Genetics 7, 79-89.
Tantawy AO (1957) Heterosis and genetic variance in hybrids between inbred
lines of Drosophila melanogaster in relation to the level of
homozygosity. Genetics 42, 535-543.
Trinkel M, Ferguson N, Reid A, et al. (2008) Translocating lions into an inbred
lion population in the Hluhluwe-iMfolozi Park, South Africa. Animal
Conservation 11, 138-143.
Turesson G (1922) The genotypical response of the plant species to the
habitat. Hereditas 3, 211-350.
Vandepitte K, Honnay O, Jacquemyn H, Roldan-Ruiz I (2010) Effects of
outcrossing in fragmented populations of the primarily selfing forest
herb Geum urbanum. Evolutionary Ecology 24, 1353-1364.
van Oosterhout C, Zijlstra WG, van Heuven MK, Brakefield PM (2000)
Inbreeding depression and genetic load in laboratory metapopulations
of the butterfly Bicyclus anynana. Evolution 54, 218-225.
van Oosterhout C, Van Heuven MK, Brakefield PM (2004) On the neutrality of
molecular genetic markers: pedigree analysis of genetic variation in
fragmented populations. Molecular Ecology 13, 1025-1034.
van Treuren R, Bilsma R, van Delden W, Ouborg NJ (1991) The significance
of genetic erosion in the process of extinction. I. Genetic differentiation
32
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
in Salvia pratensis and Scabiosa columbaria in relation to population
size. Heredity 66, 181-189.
van Treuren R, Bijlsma R, Ouborg NJ, van Delden W (1993) The significance
of genetic erosion in the process of extinction. IV. Inbreeding
depression and heterosis effects caused by selfing and outcrossing in
Scabiosa columbaria. Evolution 47, 1669-1680.
Vilà C, Sundqvist A-K, Flagstad Ø, et al. (2003) Rescue of a severely
bottlenecked wolf (Canis lupus) population by a single immigrant.
Proceedings of the Royal Society of London B: Biological Sciences
270, 91-97.
Vrijenhoek RC (1994) Genetic diversity and fitness in small populations. In:
Conservation Genetics (eds. Loeschcke V, Tomiuk J, Jain SK), pp. 3753. Birkhäuser Verlag, Basel, Switzerland.
Vrijenhoek RC, Lerman S (1982) Heterozygosity and developmental stability
under sexual and asexual breeding systems. Evolution 36, 768-776.
Wabakken P, Sand H, Liberg O, Bjarvall A (2001) The recovery, distribution,
and population dynamics of wolves on the Scandinavian peninsula,
1978-1998. Canadian Journal of Zoology-Revue Canadienne De
Zoologie 79, 710-725.
Wagenius S, Hangelbroek HH, Ridley CE, Shaw RG (2010) Biparental
inbreeding and interremnant mating in a perennial prairie plant: Fitness
consequences for progeny in their first eight years. Evolution 64, 761771.
Waite TA, Vucetich J, Saurer T, et al. (2005) Minimizing extinction risk through
genetic rescue. Animal Biodiversity and Conservation 28, 121–130.
Weekley CW, Kubisiak TL, Race TM (2002) Genetic impoverishment and
cross-incompatibility in remnant genotypes of Ziziphus celata
(Rhamnaceae), a rare shrub endemic to the Lake Wales Ridge,
Florida. Biodiversity and Conservation 11, 2027-2046.
Weller SG, Sakai AK, Thai DA, Tom J, Rankin AE (2005) Inbreeding
depression and heterosis in populations of Schiedea viscosa, a highly
selfing species. Journal of Evolutionary Biology 18, 1434-1444.
Westemeier RL, Brawn JD, Simpson SA, et al. (1998) Tracking the long-term
decline and recovery of an isolated population. Science 282, 16951698.
Whiteley AR, Fitzpatrick SW, Funk WC, Tallmon DA (2015) Genetic rescue to
the rescue. Trends in Ecology & Evolution 30, 42-49.
Willi Y, Fischer M (2005) Genetic rescue in interconnected populations of
small and large size of the self-incompatible Ranunculus reptans.
Heredity 95, 437-443.
Willi Y, Van Buskirk J, Hoffmann AA (2006) Limits of adaptive potential of
small populations. Annual Review of Ecology, Evolution and
Systematics 37, 433-458.
Willi Y, van Kleunen M, Dietrich S, Fischer M (2007) Genetic rescue persists
beyond first-generation outbreeding in small populations of a rare plant.
Proceedings of the Royal Society B 274, 2357-2364.
Williams W, Gilbert N (1960) Heterosis and the inheritance of yield in the
tomato. Heredity 14, 133-145.
33
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
Wolfe LM (1993) Inbreeding depression in Hydrophyllum appendiculatum: role
of maternal effects, crowding, and parental mating history. Evolution
47, 374-386.
Woodworth LM, Montgomery ME, Briscoe DA, Frankham R (2002) Rapid
genetic deterioration in captive populations: Causes and conservation
implications. Conservation Genetics 3, 277-288.
Worthen WB, Stiles EW (1986) Phenotypic and demographic variability
among patches of Maianthemum canadense (Desf.) in central New
Jersey, and the use of self-incompatibility for clone discrimination.
Bulletin of the Torrey Botanical Club 113, 398-405.
Wright LI, Tregenza T, Hosken DJ (2008) Inbreeding, inbreeding depression
and extinction. Conservation Genetics 9, 833-843.
Wright S (1977) Evolution and the Genetics of Populations. Volume 3.
Experimental Results and Evolutionary Deductions University of
Chicago Press, Chicago.
Wright S, Lewis P (1921) Factors in the resistance of guinea-pigs to
tuberculosis with especial regard to inbreeding and heredity. American
Naturalist 55, 20-50.
Zajitschek SRK, Zajitschek F, Brooks RC (2009) Demographic costs of
inbreeding revealed by sex-specific genetic rescue effects. BMC
Evolutionary Biology 9, 289.
926
927
34