Skip to main content

Advertisement

Log in

Using molecular markers for detecting domestication, improvement, and adaptation genes

  • Published:
Euphytica Aims and scope Submit manuscript

Abstract

Development of statistical tests to detect selection (strictly speaking, departures from the neutral equilibrium model) has been an active area of research in population genetics over the last 15 years. With the advent of dense genome sequencing of many domesticated crops, some of this machinery (which heretofore has been largely restricted to human genetics and evolutionary biology) is starting to be applied in the search for genes under recent selection in crop species. We review the population genetics of signatures of selection and formal tests of selection, with discussions as to how these apply in the search for domestication and improvement genes in crops and for adaptation genes in their wild relatives. Plant domestication has specific features, such as complex demography, selfing, and selection of alleles starting at intermediate frequencies, that compromise many of the standard tests, and hence the full power of tests for selection has yet to be realized.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  • Akey JM, Zhang G, Zhang K, Jin L, Shriver MD (2002) Interrogating a high-density SNP map for signatures of natural selection. Genome Res 12:1805–1814

    Article  PubMed  CAS  Google Scholar 

  • Biswas S, Akey JM (2006) Genomic insights into positive selection. Trends Genet 22:437–446

    Article  PubMed  CAS  Google Scholar 

  • Brinkman MA, Frey KJ (1977) Yield component analysis of oat isolines that produce different grain yields. Crop Sci 17:165–168

    Google Scholar 

  • Cavalli-Sforza LL (1966) Population structure and human evolution. Proc Royal Soc London Ser B 164:362–379

    Article  CAS  Google Scholar 

  • Clark RM, Linton E, Messing J, Doebley JF (2004) Pattern of diversity in the genomic region near the maize domestication gene tb1. PNAS 101:700–707

    Article  PubMed  CAS  Google Scholar 

  • Clark RM, Wagler TN, Quijada P, Doebley J (2006) A distant upstream enhancer at the maize domestication gene tb1 has pleotropic effects on plant and inflorescent architecture. Nat Genet 38:594–597

    Article  PubMed  CAS  Google Scholar 

  • Charlesworth B, Morgan MT, Charlesworth D (1993) The effect of deleterious mutations on neutral molecular variation. Genetics 134:1289–1303

    PubMed  CAS  Google Scholar 

  • Charlesworth D, Charlesworth B, Morgan MT (1995) The pattern of neutral molecular variation under the background selection model. Genetics 141:1619–1632

    PubMed  CAS  Google Scholar 

  • Doebley J, Stec A, Gustus C (1995) teosinte branched1 and the origin of maize: evidence for epistasis and the evolution of dominance. Genetics 141:333–346

    PubMed  CAS  Google Scholar 

  • Enard W, Przeworski M, Fisher SE, Lai CS, Wiebe V et al (2002) Molecular evolution of FOXP2, a gene involved in speech and language. Nature 418:869–872

    Article  PubMed  CAS  Google Scholar 

  • Ewens WJ (1972) The sampling theory of selectively neutral alleles. Theor Popul Biol 3:87–112

    Article  PubMed  CAS  Google Scholar 

  • Fay JC, Wu C-I (2000) Hitchhiking under positive Darwinian selection. Genetics 155:1405–1413

    PubMed  CAS  Google Scholar 

  • Ford MJ (2002) Applications of selective neutrality tests to molecular ecology. Mol Ecol 11:1245–1262

    Article  PubMed  CAS  Google Scholar 

  • Fu Y-X (1996) New statistical tests of neutrality for DNA samples from a population. Genetics 143:557–570

    PubMed  CAS  Google Scholar 

  • Fu Y-X (1997) Statistical tests of neutrality of neutrality against population growth, hitchhiking and background selection. Genetics 147:915–925

    PubMed  CAS  Google Scholar 

  • Fu Y-X, Li W-H (1993) Statistical tests of neutrality of mutations. Genetics 133:693–709

    PubMed  CAS  Google Scholar 

  • Hamblin MT, Casa AM, Su H, Murray SC, Paterson AH, Auadro CF, Kresovich S (2006) Challenges of detecting directional selection after a bottleneck: lessons from Sorhum bicolor. Genetics 173:953–964

    Article  PubMed  CAS  Google Scholar 

  • Hein J, Schierup MH, Wiuf C (2005) Gene genealogies, variation and evolution: a primer in coalescent theory. Oxford University Press, Oxford

  • Hudson RR, Kreitman M, Aguade M (1987) A test of neutral molecular evolution based on nucleotide data. Genetics 116:153–159

    PubMed  CAS  Google Scholar 

  • Innan H, Kim Y (2004) Pattern of polymorphism after strong artificial selection in a domestication event. PNAS 101:10667–10672

    Article  PubMed  CAS  Google Scholar 

  • Jensen JD, Kim Y, DuMont VB, Aquadro CF, Bustamante CD (2005) Distinguishing between selective sweeps and demography using DNA polymorphism data. Genetics 170:1401–1410

    Article  PubMed  CAS  Google Scholar 

  • Jensen MA, Charlesworth B, Kreitman M (2002) Patterns of genetic variation at a chromosome 4 locus of Drosophila melanogaster and D simulans. Genetics 160:493–507

    PubMed  CAS  Google Scholar 

  • Kaplan NL, Hudson RR, Langley CH (1989) The “hitchhiking effect” revisited. Genetics 123:887–899

    PubMed  CAS  Google Scholar 

  • Kayser M, Brauer S, Stoneking M (2003) A genome scan to detect candidate regions influenced by local natural selection in human populations. Mol Biol Evol 20:893–900

    Article  PubMed  CAS  Google Scholar 

  • Kim Y, Stephan W (2002) Detecting a local signature of genetic hitchhiking along a recombining chromosome. Genetics 160:765–777

    PubMed  CAS  Google Scholar 

  • Kimura M (1983) The neutral theory of molecular evolution. Cambridge Univ Press, UK

    Google Scholar 

  • Kreitman M (2000) Methods to detect selection in populations with applications to the human. Ann Rev Gemoics Hum Genet 1:539–559

    Article  CAS  Google Scholar 

  • Lewontin RC, Krakauer J (1973) Distribution of gene frequency as a test of the theory of the selective neutrality of polymorphisms. Genetics 74:175–195

    PubMed  CAS  Google Scholar 

  • Lu J, Tang T, Tang H, Huang J, Shi S, Wu C-I (2006) The accumulation of deleterious mutations in rice genomes: a hypothesis on the cost of domestication. Trends Genet 22:126–131

    Article  PubMed  CAS  Google Scholar 

  • Lynch M, Walsh B (1997) Genetics and analysis of quantitative traits. Sinauer Associates, Sunderland, MA

    Google Scholar 

  • Maynard Smith J, Haigh J (1974) The hitch-hiking effect of a favorable gene. Genet Res 23:23–35

    Article  Google Scholar 

  • McDonald JH, Kreitman M (1991) Adaptive protein evolution at the Adh locus in Drosophila. Nature 351:652–654

    Article  PubMed  CAS  Google Scholar 

  • Nielsen R (2001) Statistical tests of selective neutrality in the age of genomics. Heredity 86:641–647

    Article  PubMed  CAS  Google Scholar 

  • Nielsen R (2005) Molecular signatures of natural selection. Annu Rev Genet 39:197–218

    Article  PubMed  CAS  Google Scholar 

  • Nielsen R, Yang Z (1998) Likelihood models for detecting positively selected amino acid sites and applications to the HIV-1 envelope gene. Genetics 148:929–936

    PubMed  CAS  Google Scholar 

  • Olsen KM, Caicedo AL, Polato N, McClung A, McCouch S, Purugganan MD (2006) Selection under domestication: evidence for a sweep in the rice Waxy genomic region. Genetics 173:975–983

    Article  PubMed  CAS  Google Scholar 

  • Palaisa K, Morgante M, Tingey S, Rafalski A (2004) Long-range patterns of diversity and linkage disequilibrium surrounding the maize Y1 gene are indicative of an asymmetric selective sweep. PNAS 101:9885–9890

    Article  PubMed  CAS  Google Scholar 

  • Perlitz M, Stephan W (1997) The mean and variance of the number of segregating sites since the last hitchhiking event. J Math Biol 36:1–23

    Article  PubMed  CAS  Google Scholar 

  • Przeworski M (2002) The signature of positive selection at randomly chosen loci. Genetics 160:1179–1189

    PubMed  Google Scholar 

  • Przeworski M (2003) Estimating the time since the fixation of a beneficial allele. Genetics 164:1667–1676

    PubMed  Google Scholar 

  • Rosenberg NA, Nordborg M (2002) Genealogical trees, coalescent theory and the analysis of genetic polymorphisms. Nat Rev Genet 3:380 –390

    Article  PubMed  CAS  Google Scholar 

  • Sabeti PC, Reich DE, Higgins JM, Levine HZP, Richter DJ, Schaffner SF, Gabriel SB, Platko JV, Patterson NJ, McDonald GJ, Ackerman HC, Campbell SJ, Altshuler D, Cooper R, Kwiatkowski D, Ward R, Lander ES (2002) Detecting recent positive selection in the human genome from haplotype structure. Nature 419:832–837

    Article  PubMed  CAS  Google Scholar 

  • Sabeti PC, Schaffner SF, Fry B, Lohmueller J, Varilly P, Shamovsky O, Palma A, Mikkelsen TS, Altshuler D, Lander ES (2006) Positive natural selection in the human lineage. Science 312:1614–1620

    Article  PubMed  CAS  Google Scholar 

  • Schlštterer C (2003) Hitchhiking mapping: functional genomics from the population genetics perspective. Trends Genet 19:32–38

    Article  Google Scholar 

  • Simonsen KL, Churchill GA, Aquadro CF (1995) Properties of statistical tests of neutrality for DNA polymorphism data. Genetics 141:413–429

    PubMed  CAS  Google Scholar 

  • Stephan W, Wiehr THE, Lenz M (1992) The effect of strongly selected substitutions on neutral polymorphisms: analytic results based on diffusion theory. Theor Pop Biol 41:237–254

    Article  Google Scholar 

  • Storz JF (2005) Using genome scans of DNA polymorphism to infer adaptive population divergence. Mol Ecol 14:671–688

    Article  PubMed  CAS  Google Scholar 

  • Storz JG, Payseur BA, Nachman MW (2004) Genomic scans of DNA variability in humans reveal evidence for selective sweeps outside of Africa. Mol Biol Evol 21:1800–1811

    Article  PubMed  CAS  Google Scholar 

  • Tajima F (1989) Statistical methods for testing the neutral mutation hypothesis by DNA polymorphism. Genetics 123:585–595

    PubMed  CAS  Google Scholar 

  • Tenaillon MI, U’Ren J, Tenaillon O, Gaut BS (2004) Selection versus deomography: a multilocus investigation of the domestication process in maize. Mol Biol Evol 21:1214–1225

    Article  PubMed  CAS  Google Scholar 

  • Teshima KM, Coop G, Przeworski M (2006) How reliable are empirical genomic scans for selective sweeps? Genoem Res 16:702 –712

    Article  CAS  Google Scholar 

  • Vigouroux Y, McMullen M, Hittinger CT, Houchins K, Schulz L, Kresovich S, Matsuoka Y, Doebley J (2002) Identifying genes of agronomic importance in maize by screening microsatellites for evidence of selection during domestication. PNAS 99:9650–9655

    Article  PubMed  CAS  Google Scholar 

  • Voight BF, Kudaravlli S, Wen X, Pritchard JK (2006) A map of recent positive selection in the human genome. PLoS Biol 4:446–458

    Article  CAS  Google Scholar 

  • Wang R-L, Stec A, Hey J, Lukens L, Doebley J (1999) The limits of selection during maize domestication. Nature 398:236–239

    Article  PubMed  CAS  Google Scholar 

  • Wang ET, Kodama G, Baldi P, Moyzis RK (2006) Global landscape of recent inferred Darwinian selection for Homo sapiens. Proc Natl Acad Sci USA 103:135–140

    Article  PubMed  CAS  Google Scholar 

  • Waterson GA (1975) On the number of segregation sites. Theor Popul Biol 7:256–276

    Article  Google Scholar 

  • Whitt SR, Wilson LM, Tenaillon MI, Gaut BS, Buckler ES IV (2002) Genetic diversity and selection in the maize starch pathway. PNAS 99:12959–12962

    Article  PubMed  CAS  Google Scholar 

  • Wright SI, Gaut BS (2005) Molecular population genetics and the search for adaptive evolution in plants. Mol Biol Evol 22:506–519

    Article  PubMed  CAS  Google Scholar 

  • Wright SI, Bi IV, Schroeder SG, Yamasaki M, Doebley JF, McMullen MD, Gaut BS (2005) The effects of artificial selection on the maize genome. Science 308:1310–1314

    Article  PubMed  CAS  Google Scholar 

  • Yamasaki M, Tenaillon MI, Bi IV, Schroeder SG, Sanchez-Villeda H, Doebley JF, Gaut BS, McMullen MD (2005) A large-scale screen for artificial selection in maize identifies candidate agronomic loci for domestication and crop improvement. Plant Cell 17:2859–2872

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgments

Many thanks to the two careful reviewers, and Associate Editor H.-P. Piepho for their detailed comments that significantly improved the manuscript. This paper was initially presented at the 2006 Biometrics in Plant Breeding meeting in Zagreb, Croatia.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bruce Walsh.

Appendix: Details of frequency spectrum-based tests

Appendix: Details of frequency spectrum-based tests

Fu and Li’s D * and F * tests

Table 1 provides three different estimators of θ under the infinite-sites model. Tajima’s D is based on the contrast between two of these, but this leaves two other contrasts, which Fu and Li (1993) used as the basis for two new tests. Their D * test compares the segregating sites (S) versus singletons (η) estimator of θ,

$$ D^{\ast}=\frac{\widehat{\theta}_S-\widehat{\theta}_{\eta}} {\sqrt{\alpha_{\ast}S +\beta_{\ast} S^2}} $$
(A1a)
$$ \alpha_{\ast}=\frac{1}{a_n}\left({\frac{n+1}{n}-\frac{1}{a_n}}\right)-\beta_{\ast} $$
(A1b)
$$ \beta_{\ast} =\frac{1}{a_n^2+b_n}\left({\frac{b_n}{a_n^2}-\frac{2}{n}\left({1 + \frac{1}{a_n}-a_n+\frac{a_n}{n}}\right)- \frac{1}{n^2}}\right). $$
(A1c)

In contrast, their F * test compares the average pair-wise divergence (k) versus singletons (η) estimator of θ,

$$ F^{\ast}= \frac{\widehat{\theta}_k-\widehat{\theta}_{\eta}}{\sqrt{\alpha_FS +\beta_F S^2}} $$
(A2a)
$$ \alpha_F= \frac{1}{a_n}\left({\frac{4n^2+19n+3-12(n+1)a_{n+1}}{3n(n-1)}}\right) -\beta_F $$
(A2b)
$$ \beta_F =\frac{1}{a_n^2+b_n}\left({\frac{2n^4+110n^2-255n+153}{9n^2(n-1)} + \frac{2(n-1)a_n}{n^2} - \frac{8b_n}{n}} \right). $$
(A2c)

These expression are from Simonsen et al. (1995), with Eq. A2c correcting a typo in the original Fu and Li paper. Critical values for both tests are tabulated by Fu and Li (1993). While these tests are fairly widely used, Simonsen et al. (1995) found that they are not as powerful as Tajima’s test for detecting a selective sweep or population structure departures (bottlenecks or population subdivision). However, Fu (1997) found that both tests have more power than Tajima’s D for detecting signals of background selection.

Fu’s W and F S tests

Fu (1996, 1997) proposed several more refined tests for specific settings, such as too few alleles or too many alleles. These tests use the infinite alleles (as opposed to infinite sites) framework for sequence analysis (see Fig. 4). To develop these, we first need to introduce Ewen’s Sampling Formula (Evens 1972), which gives the probability (under the infinite alleles model) that we see K alleles (haplotypes) in a sample of size n as

$$ \Pr( K=k)=\frac{|\,S^k_n|\theta^k}{S_n(\theta)} $$
(A3a)

where

$$ S_n(\theta)=\theta(\theta+1)(\theta+2) \cdots (\theta+n-1) $$
(A3b)

and S k n is the coefficient on the θk term in the polynomial given by S n (θ). (S k n is called a Stirling number of the first kind). For example, the probability that only a single allele is seen in our sample is

$$ \Pr(K=1)=\frac{(n-1)!}{(\theta+1)(\theta+2) \cdots (\theta+n-1)} $$
(A4)

Using Eq. A3a, the mean and variance for the number of alleles can be found to be

$$ E(K) =1+ \theta\cdot \sum\limits_{j=2}^{n} \frac{1}{\theta+j-1} $$
(A5a)

and

$$ \sigma^2(K)=\theta\cdot \sum\limits_{j=1}^{n-1} \frac{j}{(\theta+j)^2} $$
(A5b)

Fu’s W test (1996) is based on Ewen’s sampling formula, and is as follows. Suppose we have an estimate \(\widehat{\theta}\) of θ and we observe k alleles in our sample. The probability of seeing k (or fewer) alleles in our sample under the null hypothesis is just

$$ W=\Pr( K \le k)=\sum\limits_{i=1}^k \Pr( K= k|\widehat{\theta})= \sum\limits_{i=1}^k \frac{|S^k_n|\widehat{\theta}^{k}} {S_n(\widehat{\theta})} $$
(A6)

The W test uses the Watterson estimator \(\widehat{\theta}=S/a_n\) so that

$$ S_n(\widehat{\theta})=S/a_n(S/a_n+1)(S/a_n+2) \cdots (S/a_n+n-1) $$

This is a test for a deficiency of rare alleles, and hence W is a one-sided test statistic. Fu (1996) showed that the W test is more powerful that Tajima’s D and Fu and Li’s D * and F * tests for detecting samples from a structured population (as also occurs with overdominant selection).

Fu’s F S test (1997) is the complement of his W statistic, being a test for excess rare alleles. It starts by computing the probability of seeing m or more alleles in our sample,

$$ S^{\prime}=\Pr( K \ge m)=\sum\limits_{i=m}^n \frac{|S^m_n|\widehat{\theta}^{m}} {S_n(\widehat{\theta})} $$
(A7a)

but now using \(\widehat{\theta}=k,\) the estimator of θ based on average number of pair-wise differences. Fu notes that S′ is not an optimal test statistic because its critical points are often too close to zero. Because of this, the test statistic S is the logit of S′,

$$ F_S=\ln \left({ \frac{S^{\prime}}{1-S^{\prime}}}\right) $$
(A7b)

F S is negative when there is an excess of rare alleles (as occurs with an excess of recent mutations as would occur with a selective sweep or population expansion), with a sufficiently large negative value being evidence for selection. Hence, F S is also a one-sided test statistic. Fu (1997) showed that F S is more powerful that Tajima’s and Fu and Li’s tests for detecting population growth/selective sweeps. Conversely, Fu and Li’s tests are more powerful for detecting background selection.

Fay and Wu’s H test

Fay and Wu (2000) and Kim and Stephan (2000) note that a distinct signal is left by a selective sweep that is not left by background selection. Specifically, it is common to see alleles that have newly arisen by mutation at high frequency following a sweep (as they hitched along for the ride). With background selection, this feature is not expected. This is the basis for Fay and Wu’s H test, which disproportionately weights derived alleles at high frequency. Their test requires an outgroup so that one can access whether an allele occurs in the outgroup or has recently been derived by mutation. Such derived alleles are expected to be at lower frequency (as under neutrality, the frequency of an allele is a rough indicator of its age, with older alleles being more frequent). The test proceeds as follows. Let S i denote the number of derived mutants foundi times in our sample of size n. For example, if there are five unique (derived) alleles, four alleles each appearing twice, and one allele appearing five times in our sample of size 18, then S 1 = 5, S 2 = 4, S 5 = 1. The estimate of θ from the average pair-wise difference expressed in terms of the S i is

$$ \widehat{\theta}_k=2\sum\limits_{i=1}^{n-1} \frac{S_i i(n-i)}{n(n-1)} $$
(A8a)

while an estimate of θ weighted by homozygosity is

$$ \widehat{\theta}_H=2\sum\limits_{i=1}^{n-1} \frac{S_i i^2}{n(n-1)} $$
(A8b)

Fay and Wu’s H test is given by the scaled difference of \(\widehat{\theta}_H- \widehat{\theta}_k.\)

Given that Fay and Wu’s test weights derived allele at high frequency, a significant H and D test is consistent with a selective sweep, while a significant D test, but not a significant H test suggests background selection or demographic features more likely account for the departure from neutrality. While widely used, the H test is not without problems. While it is largely insensitive to population bottlenecks, it is highly sensitive to population structure. Further, the power of H rapidly decreases over time following a sweep, while the D test retains substantial power over a much longer time after a sweep (Przeworski 2002).

Hudson–Kreitman–Aguade (HKA) test

Consider two species (or distant populations) A and B that are at mutation-drift equilibrium with population sizes N A  = N and N B  = α N, respectively. Further assume they separated T = τ/(2N) generations ago from a common population of size \(N^{\ast}=(N_A+N_B)/2=N(1+\alpha)/2,\) the average of the two current population sizes. Suppose \(i=1, \cdots, L\) unlinked loci are examined in both species. The amount of polymorphism for locus i in A is a function of \(\theta_i=4N_e\mu_i,\) while the amount of polymorphism for the same locus in B is a function of \(4N_B\mu_i=4(\alpha N_e)\mu_i=\alpha\theta_i.\) The resulting summary statistics used are LS A i values, for the number of segregating sites at locus i in A, another LS B i for the same loci in B, and L D i values, for the amounts of divergence (measured by the average number of differences between a random gamete from A and a random gamete from B). Given these 3L summary statistics, the HKA test statistic X 2 is given by

$$ X^2=\sum\limits_{i=1}^L \frac{\left({S_i^A - \widehat{E}(S_i^A)}\right)^2}{\widehat{Var}(S_i^A)} + \sum\limits_{i=1}^L \frac{\left({S_i^B - \widehat{E}(S_i^B)}\right)^2} {\widehat{Var}(S_i^B)} + \sum\limits_{i=1}^L \frac{\left({D_i - \widehat{E}(D_ii)}\right)^2}{\widehat{Var}(D_i)} $$
(A9)

where for n A samples from A and n B samples from B,

$$ \widehat{E}(S_i^A)= \widehat{\theta_i}a_{n_A},\quad \widehat{E}(S_i^A)=\widehat{\alpha}\widehat{\theta_i}a_{n_B} $$
(A10a)
$$ \widehat{Var}(S_i^A) =\widehat{\theta_i}a_{n_A} + \widehat{\theta_i}^2b_{n_A}, \widehat{Var}(S_i^B )= \widehat{\alpha} \widehat{\theta_i}a_{n_A} + \widehat{\alpha}^2\widehat{\theta_i}^2b_{n_B} $$
(A10b)
$$ \widehat{D_i}= \widehat{\theta_i}\left({\widehat{T} + \frac{1+\widehat{\alpha}}{2}}\right) $$
(A10c)
$$ \widehat{Var}(D_i)=\widehat{\theta_i}\left({\widehat{T} + \frac{1+\widehat{\alpha}}{2}}\right) +\left({\frac{ \widehat{\theta_i}(1+\widehat{\alpha})}{2}}\right)^2 $$
(A10d)

and a n and b n are given by Eq. 2. Equations A10a and (A10b) follow from Eq. 3, while Eq. A10c follows by re-writing

$$ \theta_i\left({T + \frac{1+\alpha}{2}}\right)=4N\mu_i\left( {\frac{\tau}{2N} + \frac{1+\alpha}{2}}\right)=2\mu_i\tau + 4\mu_i \frac{N (1+\alpha)}{2} $$

where the first term is the between-population divergence due to new mutations and the second term the divergence from partitioning of the polymorphism \(4N^{\ast}\mu_i\) in the ancestral population. Thus, the HKA test has L + 2 parameters to estimate, the \(L\theta_i^x\) values and two demographic parameters, T and α. The HKA test estimates these parameters and then (using Eq. A10) computes the goodness of fit X 2 statistic (Eq. A9), which is approximately χ2 distributed with 3L − (L + 2) = 2L − 2 degrees of freedom.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Walsh, B. Using molecular markers for detecting domestication, improvement, and adaptation genes. Euphytica 161, 1–17 (2008). https://doi.org/10.1007/s10681-007-9465-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10681-007-9465-8

Keywords

Navigation