24.06.2014 Views

Beta vulgaris - Save Our Seeds

Beta vulgaris - Save Our Seeds

Beta vulgaris - Save Our Seeds

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

3<br />

45<br />

12345678910 11<br />

13 14 15 21 31 32 33<br />

Received 6 March 2003<br />

FirstCite®<br />

Accepted 19 March 2003<br />

34<br />

e-publishing<br />

Published online<br />

38 46<br />

47 48 49<br />

50 Evidence for gene flow via seed dispersal from crop<br />

51 to wild relatives in <strong>Beta</strong> <strong>vulgaris</strong> (Chenopodiaceae):<br />

52 consequences for the release of genetically modified<br />

53 crop species with weedy lineages<br />

54 J.-F. Arnaud * , F. Viard†, M. Delescluse and J. Cuguen<br />

55 UMR CNRS 8016, Laboratoire de Génétique et Evolution des Populations Végétales, Bâtiment SN2, Université de Lille 1,<br />

56 59655 Villeneuve d’Ascq cedex, France<br />

57 Gene flow and introgression from cultivated to wild plant populations have important evolutionary and<br />

58 ecological consequences and require detailed investigations for risk assessments of transgene escape into<br />

59 natural ecosystems. Sugar beets (<strong>Beta</strong> <strong>vulgaris</strong> ssp. <strong>vulgaris</strong>) are of particular concern because: (i) they are<br />

60 cross-compatible with their wild relatives (the sea beet, B. <strong>vulgaris</strong> ssp. maritima); (ii) crop-to-wild gene<br />

61 flow is likely to occur via weedy lineages resulting from hybridization events and locally infesting fields.<br />

62 Using a chloroplastic marker and a set of nuclear microsatellite loci, the occurrence of crop-to-wild gene<br />

63 flow was investigated in the French sugar beet production area within a ‘contact-zone’ in between coastal<br />

64 wild populations and sugar beet fields. The results did not reveal large pollen dispersal from weed to wild<br />

65 beets. However, several pieces of evidence clearly show an escape of weedy lineages from fields via seed<br />

66 flow. Since most studies involving the assessment of transgene escape from crops to wild outcrossing<br />

67 relatives generally focused only on pollen dispersal, this last result was unexpected: it points out the key<br />

68 role of a long-lived seed bank and highlights support for transgene escape via man-mediated long-distance<br />

69 dispersal events.<br />

70 Keywords: <strong>Beta</strong> <strong>vulgaris</strong>; contact zone; cpDNA; crop–wild hybrids; GMO; microsatellites;<br />

71 risk assessment; weed beets1<br />

72<br />

73<br />

74 1. INTRODUCTION<br />

75 Over the last decade, the development of transgenic plants<br />

76 has heightened concerns about the potential negative<br />

77 effects of the wide-scale commercial release of genetically<br />

78 engineered crops (Darmency 1994; Raybould & Gray<br />

79 1994; Gray & Raybould 1998; Hails 2000). Among the<br />

80 commonly listed risks associated with environmental<br />

81 release of genetically engineered cultivars is the hybridiz-<br />

82 ation of transgenic plants with wild relatives, and the sub-<br />

83 sequent introgression of transgenic traits into the gene<br />

84 pool of wild plant populations in natural ecosystems<br />

85 (Ellstrand et al. 1999). Furthermore, the risk of a trans-<br />

86 genic crop escaping cultivation is likely to be higher in<br />

87 crops where non-transgenic varieties have weedy tend-<br />

88 encies (Raybould & Gray 1994; Bartsch et al. 1999). Even<br />

89 if gene flow between wild and cultivated relatives results<br />

90 in a transgene escape, its spread will obviously depend on<br />

91 the relative fitness enhancement attributable to the nature<br />

92 of the engineered trait, as well as on other factors such as<br />

93 hybrid fertility, particularly during seedling establishment<br />

94 (Van Raamsdonk & Schouten 1997; Bartsch et al. 1999;<br />

95 Hails 2000).<br />

96 With regard to crop/wild gene flow and its potential<br />

97 consequences, the <strong>Beta</strong> <strong>vulgaris</strong> complex is of particular<br />

98 interest as crop, wild and weedy forms can be found in<br />

99 parapatry or sympatry in Europe with overlapping flower-<br />

1<br />

*<br />

28 Author for correspondence (jean-francois.arnaud@univ-lille1.fr).<br />

29 † Present address: UMR 7127, EGPM, Station Biologique, Place<br />

30 Georges-Teissier, BP 74, 29682 Roscoff cedex, France.<br />

1<br />

2 Proc. R. Soc. Lond. B 03PB0216.1 © 2003 The Royal Society<br />

3 DOI 10.1098/rspb.2003.2407<br />

6<br />

ing periods, are wind-pollinated and all appear to be interfertile<br />

(Bartsch et al. 1999; Saeglitz et al. 2000). In<br />

northern France, the sugar beet (<strong>Beta</strong> <strong>vulgaris</strong> ssp.<br />

<strong>vulgaris</strong>) is an extensively cultivated plant. Numerous<br />

fields can be found close to the coastline where the common<br />

wild form (B. <strong>vulgaris</strong> ssp. maritima) also occurs.<br />

Conspecific weed beets, which germinate from the seed<br />

bank, are also present within the sugar beet fields. Previous<br />

studies demonstrated that these weeds result, primarily,<br />

from hybridization events between cultivated and<br />

wild inland beets within seed production areas for sugar<br />

beet (Boudry et al. 1993; Bartsch et al. 1999; Van Dijk &<br />

Desplanque 1999). Various concepts coexist as to which<br />

plants are weeds (Baker 1974). In the present study, the<br />

term ‘weed’ is used to define: (i) the in-row and out-row<br />

bolters found within crops and resulting from hybridization<br />

events; and (ii) beets encountered within man-disturbed<br />

habitats in the close vicinity of the agrosystem,<br />

such as roadsides, recently built slopes, etc. (Baker 1974;<br />

Hornsey & Arnold 1979; Desplanque et al. 2002). Sugar<br />

beets are biennial, in contrast to weed beets which<br />

inherited the possibility of first-year flowering from wild<br />

forms and bolt late in the growing season without the<br />

requirement of vernalization (dominant B allele; see Boudry<br />

et al. 1993). Sugar beet seed grown for commercial<br />

purposes in northern France contains a fraction of hybrid<br />

seeds and the resulting crop–wild F 1 hybrid plants are<br />

annuals, bolting, flowering and setting large amounts of<br />

seeds. These contaminant hybrids have given rise to<br />

weedy lineages; their crucial character ‘earliness of flowering’<br />

preadapts them for invasive success in cultivated beet<br />

100<br />

101<br />

102<br />

103<br />

104<br />

105<br />

106<br />

107<br />

108<br />

109<br />

110<br />

111<br />

112<br />

113<br />

114<br />

115<br />

116<br />

117<br />

118<br />

119<br />

120<br />

121<br />

122<br />

123<br />

124<br />

125<br />

126<br />

127<br />

128<br />

129<br />

130<br />

12 PROCB: proceedings of the royal society<br />

08-05-03 14:12:57 Rev16.04x PROCBD216P<br />

3


123<br />

1 03PB0216.2 J.-F. Arnaud and others Weed–wild interactions in <strong>Beta</strong> <strong>vulgaris</strong><br />

2<br />

3<br />

131 fields (Van Dijk & Desplanque 1999). Therefore, these<br />

132 characteristics place beets in a high-risk category in terms<br />

133 of the likelihood of transgene escape, since herbicide-<br />

134 resistant transgenic sugar beet lines already exist (Boudry<br />

135 et al. 1993; Van Raamsdonk & Schouten 1997;<br />

136 Desplanque et al. 2002; Viard et al. 2002).<br />

137 However, at the current time, there is no evidence for<br />

138 gene exchange between weedy lineages and wild coastal<br />

139 relatives in sugar beet production areas. Hybridizations<br />

140 are not necessarily confined to narrow zones of close para-<br />

141 patry, and long-distance pollen dispersal up to hundreds<br />

142 of metres away from the cultivated source could be sus-<br />

143 pected in this outcrossing species (Lavigne et al. 2002). In<br />

144 this context, we attempted to evaluate the genetic features<br />

145 of nearby populations located within the French area of<br />

146 sugar beet production (northern France) by analysing: (i)<br />

147 a typical wild coastal population; (ii) a neighbouring weed<br />

148 population situated within a sugar beet field; and (iii) an<br />

149 intermediate population located between the two latter.<br />

150 Such a linear sampling scheme reflects the geography of<br />

151 a potential contact zone and provides a unique opport-<br />

152 unity to evaluate gene exchange from sugar beet crops to<br />

153 coastal wild sea beet populations, and the distance at<br />

154 which pollen and seed dispersal occurs. In this study, the<br />

155 likelihood of weed seed migration was traced back using<br />

156 a maternally inherited chloroplastic marker diagnostic of<br />

157 cultivated lines (Ran & Michaelis 1995). Possible introgr-<br />

158 ession via pollen and/or seed flow was also investigated<br />

159 using nuclear microsatellite markers (Mörchen et al. 1996;<br />

160 Viard et al. 2002).<br />

161 2. MATERIAL AND METHODS<br />

162 (a) Sampling<br />

163 To investigate fine-scale crop–wild gene flow, three locations<br />

164 were sampled: (i) a natural coastal population of wild beets<br />

165 (named COAST); (ii) a weed population (named FIELD)<br />

166 located within a sugar beet field situated 1.5 km apart; and (iii)<br />

167 a contact zone with a possible mixture of wild and weedy beets,<br />

168 located along the river of Wimereux (named RIVER) (figure 1).<br />

169 The present studied area is characterized by the high incidence<br />

170 of sugar beet farming for more than 100 years and is typical of<br />

171 what can be found in northern France where sugar beet crops<br />

172 border the coast. From an intensive field survey, this study site<br />

173 was chosen as representative of a case of close contact between<br />

174 coastal sea beet populations (together with individuals sampled<br />

175 in the river embankments) and sugar beet fields found in close<br />

176 proximity.<br />

177 The COAST population is a typical sea beet population found<br />

178 along the seashore, with individuals located at the upper level<br />

179 of the high tide. The FIELD population is characteristic of the<br />

180 infestation of sugar beet fields by weed beets. Only weed plants<br />

181 bolting outside the sowing lines (‘out-row weeds’ following the<br />

182 terminology of Viard et al. (2002)) were sampled. Weedy forms<br />

183 found outside the rows of cultivation, were derived from crosses<br />

184 of either physiological bolters or in-row weeds found in the<br />

185 vicinity (Desplanque et al. 2002; Viard et al. 2002). The RIVER<br />

186 population represents a situation analogous to an estuarial zone,<br />

187 as there is an inland tide stream that can explain the colonization<br />

188 of the river embankments. Although cultivated and wild forms<br />

189 are often difficult to distinguish, the RIVER population was<br />

190 chosen to test for possible introgression because several individ-<br />

1<br />

2 Proc. R. Soc. Lond. B<br />

3<br />

ual beets displayed conspicuous morphological characters specific<br />

to cultivated lines.<br />

Approximately 20% of the individuals belonging to the<br />

COAST and FIELD populations were randomly sampled in<br />

order to achieve the best representation of the whole population.<br />

The sampling of the RIVER population was exhaustive (i.e. all<br />

individuals were collected). Altogether, 154 individuals were<br />

collected (see figure 1) and fresh leaves were desiccated using<br />

silica gels (Prolabo Inc.) prior to DNA extraction.<br />

(b) Genetic data collection<br />

Extraction and purification of total DNA was performed using<br />

a DNeasy 96 Plant Kit following the standard protocol for isolation<br />

of DNA from plant leaf tissue outlined in the DNeasy 96<br />

Plant protocol handbook (Qiagen Inc.).<br />

The maternal cultivated origin was assessed by means of a<br />

diagnostic chloroplastic PCR–RFLP marker. Weed beets carry<br />

the Owen cytoplasmic male sterility (CMS) (Boudry et al.<br />

1993), a trait maternally inherited and characterized by the<br />

inability of the plant to produce functional pollen. This CMS is<br />

typical of sugar beet cultivars because of its worldwide use in<br />

breeding. The Owen CMS cytoplasm is associated with one<br />

additional polymorphic HindIII site on chloroplast DNA (Ran &<br />

Michaelis 1995). Primers used, PCR conditions, and DNA<br />

digestion were applied as described by Ran & Michaelis (1995).<br />

HindIII-digested cpDNA products were separated using 0.8%<br />

agarose gel electrophoresis and visualized after ethidium bromide<br />

staining under UV light.<br />

Individuals were genotyped at six microsatellite loci (CT4,<br />

GTT1, GCC1, GAA1, BVM3 and CAA1) according to protocols<br />

previously described in Mörchen et al. (1996) and Viard et<br />

al. (2002). One additional unpublished microsatellite locus<br />

(CA2) was used with CCTTGCTAGTTGCTGCTGTG and<br />

GCATATGTACAAGAGAGCCGTTT as 5-3 primer<br />

sequences. CA2 is an imperfect and interrupted locus composed<br />

of TG repeats, allelic sizes range between 225 and 229 bp, and<br />

PCR conditions are identical to those described in Viard et al.<br />

(2002) except for the annealing temperature (55 °C). Electrophoresis<br />

and genotyping were performed on a LI-COR automated<br />

DNA sequencer model 4200s (LI-COR Inc., NB, USA).<br />

(c) Statistical treatments<br />

Allele frequencies, departures from Hardy–Weinberg equilibrium,<br />

linkage disequilibrium, allelic richness following the<br />

rarefaction procedure of El Mousadik & Petit (1996), gene<br />

diversity, and unbiased intrapopulation fixation index (Fis, a<br />

measure of departures from panmixia within populations) were<br />

calculated for each population using Fstat (Goudet 1995). Global<br />

differentiation among populations (Fst) was quantified following<br />

Weir & Cockerham (1984) and tested for significant<br />

departure from zero by randomly permuting multilocus genotypes<br />

among samples.<br />

The high number of alleles segregating at hypervariable<br />

microsatellite markers is likely to achieve a differentiation of all<br />

individuals that are uniquely defined, even with a small number<br />

of loci (e.g. Beaumont et al. 2001). To depict fine-scaled introgression<br />

from crop to wild relatives, and to test individual admixture<br />

proportions and the correspondence of genetic clusters with<br />

geographically labelled groups, we applied a model-based clustering<br />

algorithm introduced by Pritchard et al. (2000). Using the<br />

software Structure (Pritchard et al. 2000), this Bayesian<br />

method enables identification of clusters of genetically similar<br />

individuals from multilocus genotypes without prior knowledge<br />

191<br />

192<br />

193<br />

194<br />

195<br />

196<br />

197<br />

198<br />

199<br />

200<br />

201<br />

202<br />

203<br />

204<br />

205<br />

206<br />

207<br />

208<br />

209<br />

210<br />

211<br />

212<br />

213<br />

214<br />

215<br />

216<br />

217<br />

218<br />

219<br />

220<br />

221<br />

222<br />

223<br />

224<br />

225<br />

226<br />

227<br />

228<br />

229<br />

230<br />

231<br />

232<br />

233<br />

234<br />

235<br />

236<br />

237<br />

238<br />

239<br />

240<br />

241<br />

242<br />

243<br />

244<br />

245<br />

246<br />

247<br />

248<br />

249<br />

250<br />

251<br />

12 PROCB: proceedings of the royal society<br />

08-05-03 14:12:57 Rev16.04x PROCBD216P<br />

3


123<br />

1 Weed–wild interactions in <strong>Beta</strong> <strong>vulgaris</strong> J.-F. Arnaud and others 03PB0216.3<br />

2<br />

3580<br />

581<br />

sugar beet production area<br />

100 m<br />

500 m<br />

1500 m<br />

FIELD<br />

FIELD<br />

n = 40<br />

the Channel<br />

COAST<br />

RIVER<br />

n = 38 n = 76<br />

river of Wimereux<br />

River 31<br />

river of Wimereux<br />

bridge<br />

River 62<br />

582<br />

583<br />

584 Figure 1. Map of the sampled area. The three B. <strong>vulgaris</strong> populations are located within (COAST and RIVER) and in the<br />

585 vicinity (FIELD) of Wimereux city (northern France, 50°46 N; 1°37 W). n: sample size. The spatial distribution of the<br />

586 individuals sampled within the RIVER population is also given. Individuals characterized by the Owen cytoplasmic male<br />

587 sterility are indicated with a grey triangle, whereas individuals carrying a non-Owen cytoplasmic type are represented by a<br />

588 white circle. Scale bar, 50 m.<br />

252 of their population affiliation. In this approach, it is assumed<br />

253 that there are K populations contributing to the gene pool of<br />

254 the sampled populations. Individuals can have membership in<br />

255 multiple clusters, with membership coefficients summing to 1<br />

256 across clusters. Each run consisted of a burn-in period of<br />

257 200 000 steps followed by 106 MCMC (Monte Carlo Markov<br />

258 Chain) replicates, with the number K of specified clusters being<br />

259 from one to six and assuming that allele frequencies are uncorre-<br />

260 lated. Repeated runs of Structure produced identical results to<br />

261 those shown.<br />

262 3. RESULTS<br />

263 Chloroplastic DNA markers allowed us to distinguish<br />

264 the cytoplasm associated with Owen’s male sterility (found<br />

265 in cultivated sugar beets and their weedy relatives resulting<br />

266 from a crossing with a maternal cultivated plant) from<br />

267 other cytoplasms. All individuals from FIELD and<br />

268 COAST populations were characterized by the OwenCms<br />

269 and non-OwenCms haplotypes, respectively. This demon-<br />

270 strates that extensive seed dispersal does not occur from<br />

271 the fields into the coastal population of the area under<br />

272 study. However, the RIVER population was clearly par-<br />

273 titioned into two groups according to the cytoplasm<br />

1<br />

2 Proc. R. Soc. Lond. B<br />

3<br />

(figure 1). Twenty-five individuals out of 76, spatially<br />

clustered in the upstream part of the RIVER population,<br />

exhibited the OwenCms haplotype, indicating that one of<br />

their maternal ancestors was a cultivated individual.<br />

Therefore, in the further nuclear microsatellite analyses<br />

the RIVER population was partitioned into two subpopulations<br />

according to OwenCms and non-OwenCms lineages.<br />

No linkage disequilibrium occurred for microsatellite<br />

loci, except for pair CT4/CAA1 within the FIELD population<br />

( p 0.05, after Bonferroni adjustment). Multiple<br />

probability tests across all population samples yielded no<br />

significant p-adjusted values, showing the independence of<br />

each locus. Summary statistics about the genetic diversity<br />

within the four populations are given in table 1. Allele frequencies<br />

are available upon request. Allelic richness<br />

(Arich.) ranged from 2 to 16.95 across loci and populations,<br />

and amounts of gene diversity (He) were relatively<br />

high whatever the lineages considered (table 1). In contrast<br />

to OwenCms lineages (FIELD and OwenCms<br />

RIVER populations), no significant departures from<br />

Hardy–Weinberg equilibrium were detected for wild beet<br />

populations (COAST and non-OwenCms RIVER), as<br />

suggested by Fis values closed to zero (table 1). The<br />

274<br />

275<br />

276<br />

277<br />

278<br />

279<br />

280<br />

281<br />

282<br />

283<br />

284<br />

285<br />

286<br />

287<br />

288<br />

289<br />

290<br />

291<br />

292<br />

293<br />

294<br />

295<br />

296<br />

297<br />

12 PROCB: proceedings of the royal society<br />

08-05-03 14:12:57 Rev16.04x PROCBD216P<br />

3


123<br />

1 03PB0216.4 J.-F. Arnaud and others Weed–wild interactions in <strong>Beta</strong> <strong>vulgaris</strong><br />

2<br />

3618<br />

619 Table 1. Genetic diversity within the FIELD, COAST and RIVER populations.<br />

620 (The RIVER population was divided into two subpopulations according to the individual chloroplastic haplotypes (OwenCms<br />

621 versus non-OwenCms). A rich. : allelic richness estimated following the rarefaction method developed by El Mousadik & Petit<br />

622 (1996); He: expected heterozygosity (gene diversity); Fis: intrapopulation fixation index and its associated significance ( ∗ p<br />

623 624 0.05, ∗∗ p 0.01, ∗∗∗ p 0.001 after Bonferroni correction).)<br />

629 634<br />

639<br />

644<br />

648 652<br />

656<br />

666<br />

669 672<br />

675<br />

689<br />

703 717<br />

731<br />

745<br />

759<br />

773<br />

787<br />

801<br />

815<br />

829<br />

843<br />

857<br />

FIELD (n = 40) RIVER (n = 76) COAST (n = 38)<br />

OwenCms (n = 25) non-OwenCms (n = 51)<br />

locus A rich. He Fis A rich. He Fis A rich. He Fis A rich. He Fis<br />

CT4 6.62 0.82 0.33 ∗∗∗ 6.00 0.79 0.24 ∗ 5.34 0.60 0.07 8.49 0.62 0.05<br />

GTT1 3.00 0.66 0.28 ∗ 3.00 0.40 0.11 3.46 0.41 0.04 3.54 0.40 0.06<br />

GCC1 2.00 0.49 0.75 ∗∗∗ 2.00 0.37 0.07 2.00 0.49 0.04 2.00 0.50 0.00<br />

GAA1 2.62 0.33 0.09 2.00 0.39 0.085 2.97 0.38 0.13 3.00 0.45 0.04<br />

BVM3 9.52 0.67 0.52 ∗∗∗ 10.00 0.76 0.63 ∗∗∗ 15.39 0.91 0.20 ∗∗∗ 16.95 0.93 0.04<br />

CAA1 7.36 0.72 0.24 ∗ 9.00 0.74 0.46 ∗∗∗ 12.73 0.83 0.17 ∗∗ 12.21 0.85 0.08<br />

CA2 3.00 0.62 0.36 ∗ 4.00 0.59 0.46 ∗∗ 2.98 0.45 0.097 2.99 0.52 0.24<br />

mean 4.87 0.62 0.37 ∗∗∗ 5.14 0.57 0.33 ∗∗∗ 6.41 0.58 0.04 7.02 0.61 0.01<br />

873<br />

874 Table 2. Pairwise estimates of genetic differentiation (below diagonal) and their associate significance (above diagonal).<br />

875 876 ( ∗∗ p 0.01, n.s.: not significant, after Bonferroni correction.)<br />

882 888<br />

894<br />

900<br />

902 904<br />

906<br />

913<br />

920 927<br />

934<br />

941<br />

948<br />

955<br />

962<br />

969<br />

RIVER<br />

FIELD OwenCms non-OwenCms COAST<br />

FIELD —<br />

∗∗ ∗∗ ∗∗<br />

RIVER OwenCms 0.0536 —<br />

∗∗ ∗∗<br />

non-OwenCms 0.1726 0.2259 — n.s.<br />

COAST 0.1702 0.2257 0.0021 —<br />

298 observed heterozygote deficiencies in weeds (OwenCms<br />

299 lineages) could not be attributed to a particular locus, so<br />

300 that both selective effects and potential presence of null<br />

301 alleles can be neglected. Using permutation procedures<br />

302 implemented in Fstat, neither Arich., nor He showed sig-<br />

303 nificant differences between non-OwenCms and<br />

304 OwenCms lineages. Fis values were, however, significantly<br />

305 higher for OwenCms lineages ( p 0.001 after 10 000<br />

306 permutations).<br />

307 Although there were strong frequency oppositions, we<br />

308 did not find any diagnostic microsatellite alleles to dis-<br />

309 tinguish between cultivated–weed lineages and wild forms<br />

310 of B. <strong>vulgaris</strong> (data not shown). However, even at this<br />

311 microgeographical scale, noticeable significant genetic dif-<br />

312 ferentiation ( p 0.01) occurred between all pairwise<br />

313 population comparisons, with the clear exception of the<br />

314 COAST population and non-OwenCms individuals of the<br />

315 RIVER population (table 2). Such a lack of differentiation<br />

316 strongly suggests that all non-OwenCms individuals<br />

317 belong to the same breeding group. By assuming unin-<br />

318 formative priors on all the K inferred clusters, the Baye-<br />

319 sian clustering approach strongly strengthened this<br />

320 hypothesis. Genetic admixture analysis clearly indicated<br />

321 that the posterior probability for the proper number of<br />

322 clusters was maximum for K = 3 populations<br />

323 (LnP = 2794.7). By incorporating prior population<br />

324 information to assist the clustering, figure 2 revealed three<br />

325 clusters that: (i) discriminated well COAST/non-<br />

326 OwenCms RIVER beets (i.e. wild lineages) and<br />

1<br />

2 Proc. R. Soc. Lond. B<br />

3<br />

cluster 1<br />

River 31<br />

FIELD<br />

RIVER OwenCms<br />

RIVER non-OwenCms<br />

COAST<br />

River 62<br />

cluster 2 cluster 3<br />

Figure 2. Diagram of three inferred clusters of individuals<br />

assuming K = 3 populations using Structure (Pritchard et<br />

al. 2000). Each point shows the mean ancestry for an<br />

individual in the sample. The values of the three coefficients<br />

in the ancestry vector q(i) are given by the distances to each<br />

of the three sides of the equilateral triangle. After the<br />

clustering was performed, the points were labelled according<br />

to sampling location and cytoplasmic type, i.e. COAST<br />

(dark blue), FIELD (red), RIVER OwenCms (pink) and<br />

RIVER non-OwenCms (light blue).<br />

590<br />

591<br />

592<br />

593<br />

594<br />

595<br />

596<br />

597<br />

598<br />

599<br />

600<br />

601<br />

602<br />

603<br />

12 PROCB: proceedings of the royal society<br />

08-05-03 14:12:57 Rev16.04x PROCBD216P<br />

3


123<br />

1 Weed–wild interactions in <strong>Beta</strong> <strong>vulgaris</strong> J.-F. Arnaud and others 03PB0216.5<br />

2<br />

3606<br />

607<br />

180<br />

160<br />

140<br />

120<br />

rank<br />

100<br />

80<br />

River 31<br />

60<br />

40<br />

20<br />

River 62<br />

0 0.1<br />

0.2 0.3 0.4 0.5<br />

q<br />

0.6 0.7 0.8 0.9 1<br />

608<br />

609<br />

weeds<br />

610 Figure 3. Distribution of the mean individual admixture coefficients q estimated using Structure (Pritchard et al. 2000)<br />

611 without prior population information. In this analysis, K (the number of population contributing to the gene pool of all<br />

612 sampled individuals) is assumed to be 2. Individuals were ranked from lowest to highest q-values and ranks were plotted<br />

613 against q. Aq-value of 1 denotes a wild individual, whereas 0 denotes weedy individuals. Also displayed are lines giving the<br />

614 95% posterior probability intervals of q for each individual. The cytoplasmic status of individuals is represented by a grey<br />

615 diamond for a OwenCms cytoplasm and a white diamond for a non-OwenCms cytoplasm.<br />

wilds<br />

327 FIELD/OwenCms RIVER beets (i.e. weedy lineages); and<br />

328 (ii) clustered almost all wild individuals into the same<br />

329 cluster (i.e. cluster 1). Among the 89 individuals carrying<br />

330 the non-OwenCms haplotype, 80 fell into cluster 1, while<br />

331 no OwenCms individuals were assigned into this cluster.<br />

332 Instead, OwenCms individuals split into clusters 2 and 3<br />

333 in accordance with their spatial location, with, nonethe-<br />

334 less, some intermediate individuals (figure 2). Does the<br />

335 nuclear polymorphism mirror the cytoplasmic partitioning<br />

336 of individuals? By assuming K = 2 (LnP = 2837.5), we<br />

337 then reported the means for the individual admixture pro-<br />

338 portion qi and their 95% probability intervals in figure 3,<br />

339 as in Beaumont et al. (2001). It can be seen that there is<br />

340 strong evidence for two distinct groups. These two clusters<br />

341 were perfectly in accordance with the cytoplasmic classi-<br />

342 fication, with only six non-OwenCms (wild) individuals<br />

343 present in an intermediate position (0.2 q 0.8) with<br />

344 very wide probability limits for q-values. However, within<br />

345 the ‘wild part’ of the RIVER population (see figure 1),<br />

346 individuals ‘River 31’ and ‘River 62’ have a large pro-<br />

347 portion (0.912 and 0.976, respectively) of their nuclear<br />

348 genome that comes from the weedy lineage although they<br />

349 carry a non-OwenCms cytoplasm. Once again, no<br />

350 OwenCms individuals were assigned into the bulk of wild<br />

351 sea beets (figure 3).<br />

1<br />

2 Proc. R. Soc. Lond. B<br />

3<br />

4. DISCUSSION<br />

An increasing number of studies have documented gene<br />

exchanges from crops into their wild or weedy relatives,<br />

which can result in the long-term establishment of cultivar<br />

alleles in wild populations (Darmency 1994; Ellstrand et<br />

al. 1999). Lavigne et al. (2002) predicted the existence of<br />

substantial pollen-mediated gene flow from cultivated to<br />

wild beets, even in the case where wild beets are supposedly<br />

destroyed within 1000 m outside the fields. None the<br />

less, the use of nuclear microsatellite markers allowed us<br />

to depict a clear genetic cleavage between wild individuals<br />

and their weedy relatives. In addition to highlighting such<br />

a genetic distinctiveness, admixture analysis also suggested<br />

clustering in accordance with the maternal origin and very<br />

rare introgression events with only, at best, two individuals<br />

(1.29%) resulting from wild–weed hybridization events.<br />

Based on these results, this study surprisingly demonstrates<br />

that introgression via pollen dispersal into B. <strong>vulgaris</strong><br />

ssp. maritima populations occupying their natural<br />

habitat may not be the most likely route for transgene<br />

spread from agriculture (but see Gray & Raybould 1998).<br />

In fact, the occurrence of such gene exchanges may be<br />

limited because weedy beet populations usually suffer<br />

from a rapid turnover owing to farming practices (crop<br />

352<br />

353<br />

354<br />

355<br />

356<br />

357<br />

358<br />

359<br />

360<br />

361<br />

362<br />

363<br />

364<br />

365<br />

366<br />

367<br />

368<br />

369<br />

370<br />

371<br />

372<br />

373<br />

374<br />

375<br />

12 PROCB: proceedings of the royal society<br />

08-05-03 14:12:57 Rev16.04x PROCBD216P<br />

3


123<br />

1 03PB0216.6 J.-F. Arnaud and others Weed–wild interactions in <strong>Beta</strong> <strong>vulgaris</strong><br />

2<br />

3<br />

376 rotation) and have a limited flowering overlap with wild<br />

377 sea beets. The episodes of close parapatry between culti-<br />

378 vated beets and coastal populations are therefore neither<br />

379 continuous in time nor in space so that gene flow from<br />

380 the ‘weedy pollen pool’ may be spatially and temporally<br />

381 restricted. Weed plant species are indeed often distributed<br />

382 in ephemeral patches (Baker 1974; Hodkinson & Thomp-<br />

383 son 1997; McCauley 1997) and the location of the 25<br />

384 weed beets collected along the ‘Wimereux’ river matched<br />

385 well with recent dyke-building work and soil transport. In<br />

386 fact, 2 years after the first sampling, this population of<br />

387 weed beets was extinct, because of man-made disturb-<br />

388 ances (building and cleaning, personal observation2).<br />

389 In contrast to pollen dispersal, seeds remain in the vicin-<br />

390 ity of the seed source for the most part, and the long-<br />

391 distance component of seed dispersal distribution is often<br />

392 difficult to document (Levin 1981; Howe & Smallwood<br />

393 1982; Cain et al. 2000; and references therein). Many<br />

394 studies have reported that pollen dispersal is often the<br />

395 major contributor to gene flow in populations of wind-<br />

396 pollinated outcrossing plant species and that striking spa-<br />

397 tial genetic structuring is mainly the result of restricted<br />

398 seed dispersal (e.g. Hamrick & Nason 1996; McCauley<br />

399 1997; Latta et al. 1998; Laporte et al. 2001). None the<br />

400 less, we should bear in mind that dispersal occurs spatially<br />

401 by seed and pollen movement, but also occurs in time via<br />

402 the seed bank (Hodkinson & Thompson 1997; Raybould<br />

403 1999). In the present study, several lines of evidence show<br />

404 that seed migration can occur from fields to wild popu-<br />

405 lations, through weedy lineages via human activities such<br />

406 as soil transport. Individuals resulting from the emergence<br />

407 of a weed seed bank (generally found within sugar beet<br />

408 fields) were found only a few metres apart from typical<br />

409 sea beet individuals. Since most studies involving the<br />

410 assessment of transgene escape from crops to wild rela-<br />

411 tives generally focus only on pollen dispersal (reviewed in<br />

412 Lavigne et al. 2002) this last result was unexpected. It is<br />

413 interesting with regard to the processes associated with<br />

414 transgene escape in B. <strong>vulgaris</strong> (Gray & Raybould 1998),<br />

415 as well as other key aspects of the biology of plants, such as<br />

416 invasiveness or the evolution of metapopulation dynamics<br />

417 (Cain et al. 2000). Because beets have a long-lived seed<br />

418 bank and can shed thousands of seeds, this means that<br />

419 weeds are likely to reappear each spring despite stochastic<br />

420 disturbances (Desplanque et al. 2002). Together with<br />

421 stochastic founder effects, occasional long-distance seed<br />

422 dispersal, without regard to the extent of geographical sep-<br />

423 aration, also means that gene flow would not necessarily<br />

424 decline regularly as an exponential function of distance<br />

425 from the genetically engineered crop (Raybould & Gray<br />

426 1994; Van Raamsdonk & Schouten 1997).<br />

427 Evaluation of the relative contribution of pollen and<br />

428 seed movement using maternally (chloroplastic DNA) or<br />

429 biparentally (microsatellites) inherited markers is also<br />

430 especially interesting in weedy plant species (McCauley<br />

431 1997). In human-dispersed species, such as weedy forms<br />

432 of B. <strong>vulgaris</strong>, many colonization events may result from<br />

433 long-distance dispersal of a relatively few seeds into<br />

434 ephemeral vacant habitats, a situation in which extinction–<br />

435 recolonization models of genetic differentiation would be<br />

436 more adequate than classical models of gene flow<br />

437 (Hamrick & Nason 1996; McCauley 1997). Indeed, weed<br />

438 beets are adapted to man-disturbed habitats (e.g. very<br />

1<br />

2 Proc. R. Soc. Lond. B<br />

3<br />

high seed output in favourable environmental<br />

circumstances) and may display life histories equivalent to<br />

the demography assumed in the so-called metapopulation<br />

model (Hamrick & Nason 1996; Hanski 1999). Therefore,<br />

seed movement should be relatively common in the<br />

sugar beet production areas and may involve multiple<br />

colonization events through either the seed bank and/or<br />

human-mediated long-range dispersal. As (i) the weed<br />

population located along the river embankments is clearly<br />

differentiated from the neighbouring FIELD population;<br />

and (ii) substantial departures from Hardy–Weinberg<br />

equilibrium are found within both weed populations (<br />

Fis 0; see table 1), these findings highlight the occurrence<br />

of stochastic colonization events involving a mixture<br />

of genetically distinct individuals coming from several<br />

sources (Wahlund effect). Accordingly, an additional<br />

subclustering analysis using Structure showed that<br />

weedy individuals sampled within the sugar beet crop split<br />

into seven distinct clusters with the highest probability<br />

(results not shown). Hence, the evolution of B. <strong>vulgaris</strong><br />

weediness is likely to result from multiple exports of genetically<br />

differentiated crop–wild hybrids from the seed production<br />

areas, followed by secondary hybridization events<br />

(see also Viard et al. 2002).<br />

In conclusion, although pollen usually represents a significant<br />

vector for the spread of genetically modified traits,<br />

the present results suggest: (i) that seed flow may have<br />

a deeper and longer impact in connecting wild and crop<br />

relatives within the complex <strong>Beta</strong>; and (ii) point out the<br />

key role of a long-lived seed bank, a factor often neglected.<br />

The authors thank Jacqueline Bernard for technical advice and<br />

support. They are grateful to Henk Van Dijk and two anonymous<br />

referees for critical reading and helpful comments on<br />

earlier versions of the manuscript. This work was funded by a<br />

grant from the Bureau des Ressources Génétiques, the ACI<br />

Impact des OGM from the French Ministry of Research, the<br />

Contrat de Plan Etat/Région Nord-Pas-de-Calais. M.D. was<br />

supported by an INRA/Région Nord-Pas-de-Calais fellowship.<br />

REFERENCES<br />

Baker, H. G. 1974 The evolution of weeds. A. Rev. Ecol. Syst.<br />

5, 1–24.<br />

Bartsch, D., Lehnen, M., Clegg, J., Pohl-Orf, M., Schuphan,<br />

I. & Ellstrand, N. C. 1999 Impact of gene flow from cultivated<br />

beet on genetic diversity of wild sea beet populations.<br />

Mol. Ecol. 8, 1733–1741.<br />

Beaumont, M., Barratt, E. M., Gottelli, D., Kitchener, A. C.,<br />

Daniels, M. J., Pritchard, J. K. & Bruford, M. W. 2001 Genetic<br />

diversity and introgression in the Scottish wildcat. Mol.<br />

Ecol. 10, 319–336.<br />

Boudry, P., Mörchen, M., Saumitou-Laprade, P., Vernet, P. &<br />

Van Dijk, H. 1993 The origin and evolution of weed beets:<br />

consequences for the breeding and release of herbicide<br />

resistant transgenic sugar beets. Theor. Appl. Genet. 87,<br />

471–478.<br />

Cain, M. L., Milligan, B. G. & Strand, A. E. 2000 Long-distance<br />

seed dispersal in plant populations. Am. J. Bot. 87,<br />

1217–1227.<br />

Darmency, H. 1994 The impact of hybrids between genetically<br />

modified crop plants and their related species: introgression<br />

and weediness. Mol. Ecol. 3, 37–40.<br />

Desplanque, B., Hautekèete, N.-C. & Van Dijk, H. 2002<br />

Transgenic weed beets: possible, probable, avoidable? J.<br />

Appl. Ecol. 39, 561–571.<br />

439<br />

440<br />

441<br />

442<br />

443<br />

444<br />

445<br />

446<br />

447<br />

448<br />

449<br />

450<br />

451<br />

452<br />

453<br />

454<br />

455<br />

456<br />

457<br />

458<br />

459<br />

460<br />

461<br />

462<br />

463<br />

464<br />

465<br />

466<br />

467<br />

468<br />

469<br />

470<br />

471<br />

472<br />

473<br />

474<br />

475<br />

476<br />

477<br />

478<br />

479<br />

480<br />

481<br />

482<br />

483<br />

484<br />

485<br />

486<br />

487<br />

488<br />

489<br />

490<br />

491<br />

492<br />

493<br />

494<br />

495<br />

496<br />

497<br />

498<br />

499<br />

500<br />

501<br />

12 PROCB: proceedings of the royal society<br />

08-05-03 14:12:57 Rev16.04x PROCBD216P<br />

3


123<br />

1 Weed–wild interactions in <strong>Beta</strong> <strong>vulgaris</strong> J.-F. Arnaud and others 03PB0216.7<br />

2<br />

3<br />

502 El Mousadik, A. & Petit, R. J. 1996 High level of genetic differ-<br />

503 entiation for allelic richness among population of the argan<br />

504 tree [Argania spinosa (L.) Skeels] endemic to Morocco.<br />

505 Theor. Appl. Genet. 92, 832–839.<br />

506 Ellstrand, N. C., Prentice, H. C. & Hancock, J. F. 1999 Gene<br />

507 flow and introgression from domesticated plants into their<br />

508 wild relatives. A. Rev. Ecol. Syst. 30, 539–563.<br />

509 Goudet, J. 1995 Fstat (version 1.2). A computer program to<br />

510 calculate F-statistics. J. Hered. 86, 485–486.<br />

511 Gray, A. J. & Raybould, A. F. 1998 Reducing transgene escape<br />

512 routes. Nature 392, 653–654.<br />

513 Hails, R. S. 2000 Genetically modified plants—the debate con-<br />

514 tinues. Trends Ecol. Evol. 15, 14–18.<br />

515 Hamrick, J. L. & Nason, J. D. 1996 Consequences of dispersal<br />

516 in plants. In Population dynamics in ecological space and time<br />

517 (ed. O. E. Rhodes, R. K. Chesser & M. H. Smith), pp. 203–<br />

518 236. University of Chicago Press.<br />

519 Hanski, I. 1999 Metapopulation ecology. Oxford University<br />

520 Press.<br />

521 Hodkinson, D. J. & Thompson, K. 1997 Plant dispersal: the<br />

522 role of man. J. Appl. Ecol. 34, 1484–1496.<br />

523 Hornsey, K. G. & Arnold, M. M. 1979 The origin of weed<br />

524 beet. Ann. Appl. Biol. 92, 279.<br />

525 Howe, H. F. & Smallwood, J. 1982 Ecology of seed dispersal.<br />

526 A. Rev. Ecol. Syst. 13, 201–228.<br />

527 Laporte, V., Viard, F., Bena, G., Valero, M. & Cuguen, J.<br />

528 2001 The spatial structure of sexual and cytonuclear poly-<br />

529 morphism in the gynodioecious <strong>Beta</strong> <strong>vulgaris</strong> ssp. maritima:<br />

530 I/ at a local scale. Genetics 157, 1699–1710.<br />

531 Latta, R. G., Linhart, Y. B., Fleck, D. & Elliot, M. 1998 Direct<br />

532 and indirect estimates of seed versus pollen movement<br />

533 within a population of Ponderosa pine. Evolution 52, 61–67.<br />

534 Lavigne, C., Klein, E. K. & Couvet, D. 2002 Using seed purity<br />

535 data to estimate an average pollen mediated gene flow from<br />

536 crops to wild relatives. Theor. Appl. Genet. 104, 139–145.<br />

537 Levin, D. A. 1981 Dispersal versus gene flow in plants. Ann.<br />

538 Missouri Bot. Gard. 68, 233–253.<br />

McCauley, D. E. 1997 The relative contributions of seed and<br />

pollen movement to the local genetic structure of Silene alba.<br />

J. Hered. 88, 257–263.<br />

Mörchen, M., Cuguen, J., Michaelis, G., Hanni, C. & Saumitou-<br />

Laprade, P. 1996 Abundance and length polymorphism of<br />

microsatellite repeats in <strong>Beta</strong> <strong>vulgaris</strong> L. Theor. Appl. Genet.<br />

92, 326–333.<br />

Pritchard, J. K., Stephens, M. & Donnelly, P. 2000 Inference<br />

of population structure using multilocus genotype data. Genetics<br />

155, 945–959.<br />

Ran, Z. & Michaelis, G. 1995 Mapping of a chloroplast RFLP<br />

marker associated with the CMS cytoplasm of sugar beet<br />

(<strong>Beta</strong> <strong>vulgaris</strong>). Theor. Appl. Genet. 91, 836–840.<br />

Raybould, A. F. 1999 Transgenes and agriculture—going with<br />

the flow? Trends Plant Sci. 4, 247–248.<br />

Raybould, A. F. & Gray, A. J. 1994 Will hybrids of genetically<br />

modified crops invade natural communities? Trends Ecol.<br />

Evol. 9, 85–89.<br />

Saeglitz, C., Pohl, M. & Bartsch, D. 2000 Monitoring gene<br />

flow from transgenic sugar beet using cytoplasmic male-sterile<br />

bait plants. Mol. Ecol. 9, 2035–2040.<br />

Van Dijk, H. & Desplanque, B. 1999 European <strong>Beta</strong>: crops<br />

and their wild and weedy relatives. In Plant evolution in manmade<br />

habitats (ed. L. W. D. Van Raamsdonk & J. C. M. Den<br />

Nijs), pp. 257–270. Amsterdam: Hugo de Vries Laboratory.<br />

Van Raamsdonk, L. W. D. & Schouten, H. J. 1997 Gene flow<br />

and establishment of transgenes in natural plant populations.<br />

Acta Bot. Neerl. 46, 69–84.<br />

Viard, F., Bernard, J. & Desplanque, B. 2002 Crop–weed<br />

interactions in the <strong>Beta</strong> <strong>vulgaris</strong> complex at a local scale:<br />

allelic diversity and gene flow within sugar beet fields. Theor.<br />

Appl. Genet. 104, 688–697.<br />

Weir, B. S. & Cockerham, C. C. 1984 Estimating F-statistics<br />

for the analysis of population structure. Evolution 38,<br />

1358–1370. 573 574<br />

As this paper exceeds the maximum length normally permitted, the 575<br />

authors have agreed to contribute to production costs. 576 577<br />

539<br />

540<br />

541<br />

542<br />

543<br />

544<br />

545<br />

546<br />

547<br />

548<br />

549<br />

550<br />

551<br />

552<br />

553<br />

554<br />

555<br />

556<br />

557<br />

558<br />

559<br />

560<br />

561<br />

562<br />

563<br />

564<br />

565<br />

566<br />

567<br />

568<br />

569<br />

570<br />

571<br />

572<br />

1<br />

2 Proc. R. Soc. Lond. B<br />

3<br />

12 PROCB: proceedings of the royal society<br />

08-05-03 14:12:57 Rev16.04x PROCBD216P<br />

3

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!