You are on page 1of 12

REVIEW & INTERPRETATION

Genomic Selection for Crop Improvement


Elliot L. Heffner, Mark E. Sorrells, and Jean-Luc Jannink*

E.L. Heff ner and M.E. Sorrells, Dep. of Plant Breeding and Genetics,
ABSTRACT Cornell Univ., Bradfield Hall, Ithaca, NY 14853; J.-L. Jannink, USDA-
Despite important strides in marker technolo- ARS, R.W. Holley Center for Agriculture and Health, Cornell Univ.,
gies, the use of marker-assisted selection has Ithaca, NY 14853. Received 2 Sept. 2008. *Corresponding author
stagnated for the improvement of quantitative ( jeanluc.jannink@ars.usda.gov).
traits. Biparental mating designs for the detec-
Abbreviations: BLUP, best linear unbiased predictor; EBV, estimated
tion of loci affecting these traits (quantitative
breeding value; G×E, genotype × environment GEBV, genomic esti-
trait loci [QTL]) impede their application, and
mated breeding value; GS, genomic selection; LD, linkage disequilib-
the statistical methods used are ill-suited to the
rium; MAS, marker-assisted selection; QTL, quantitative trait locus; RR,
traits’ polygenic nature. Genomic selection (GS)
ridge regression; SR, stepwise regression; TBV, true breeding value.
has been proposed to address these deficien-
cies. Genomic selection predicts the breeding
values of lines in a population by analyzing their
phenotypes and high-density marker scores. A
key to the success of GS is that it incorporates
T he use of marker-assisted selection (MAS) in plant breeding
has continued to increase in the public and private sectors.
Most applications, however, have been constrained to simple,
all marker information in the prediction model, monogenic traits (reviewed by Xu and Crouch, 2008). While
thereby avoiding biased marker effect estimates
MAS has had significant impacts in backcrossing of major genes
and capturing more of the variation due to
into elite varieties (Holland, 2004), backcrossing is regarded as
small-effect QTL. In simulations, the correlation
between true breeding value and the genomic
the most conservative of breeding methods because improvement
estimated breeding value has reached levels of occurs through the pyramiding of only a few target genes (Lee,
0.85 even for polygenic low heritability traits. 1995). Gene pyramiding is inefficient for quantitative traits that
This level of accuracy is sufficient to consider are often controlled by many small-effect quantitative trait loci
selecting for agronomic performance using (QTL; Kearsey and Farquhar, 1998).
marker information alone. Such selection would Current MAS methods are better suited for manipulating a
substantially accelerate the breeding cycle, few major effect genes than many small-effect genes (Dekkers and
enhancing gains per unit time. It would dra- Hospital, 2002). Unfortunately, these small-effect genes underlie
matically change the role of phenotyping, which the complex polygenic traits that are crucial for the success of
would then serve to update prediction models new crop varieties (Crosbie et al., 2003). Two primary limitations
and no longer to select lines. While research to
to MAS are (i) the biparental mapping populations used in most
date shows the exceptional promise of GS, work
QTL studies do not readily translate to breeding applications and
remains to be done to validate it empirically and
to incorporate it into breeding schemes.
(ii) statistical methods used to identify target loci and implement

Published in Crop Sci. 49:1–12 (2009).


doi: 10.2135/cropsci2008.08.0512
© Crop Science Society of America
677 S. Segoe Rd., Madison, WI 53711 USA
All rights reserved. No part of this periodical may be reproduced or transmitted in any
form or by any means, electronic or mechanical, including photocopying, recording,
or any information storage and retrieval system, without permission in writing from
the publisher. Permission for printing and for reprinting the material contained herein
has been obtained by the publisher.

CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009 1


MAS have been inadequate for improving polygenic traits with significant markers in a selection index that would
controlled by many loci of small effect. The application explain a significant proportion of additive genetic vari-
of genomic selection (GS) proposed by Meuwissen et al. ance. In the first step, they were unable to estimate all
(2001) to breeding populations using high marker densi- marker effects simultaneously with simple regression due
ties is emerging as a solution to both of these deficiencies. to the lack of degrees of freedom. Therefore, they pro-
We review here current GS methods and their perfor- posed selecting the most significant markers from the pre-
mance. In addition, we present future directions for GS vious generation via multiple linear regressions and then
research and some exciting opportunities GS provides that reestimating effects of the selected markers in the current
could revolutionize crop improvement. generation with independent multiple regressions (Lande
and Thompson, 1990).
CURRENT MAS LIMITATIONS Lande and Thompson (1990) introduced this two-
The most common method of QTL detection is the use of step approach to handle large marker sets because they
a biparental mapping population. While these studies are estimated that hundreds of molecular markers would be
important to the understanding of genetic architecture, needed to capture a significant proportion of the additive
building mapping populations distinct from breeding genetic variance. In the early 1990s, genomewide marker
populations often strains the resources of a breeding pro- coverage was a limiting factor for MAS, but in recent
gram. Available resources limit the size of mapping popu- years, plant breeders have encountered a major shift in the
lations and, consequently, the accuracy of QTL position amount of genomic information available due to the rapid
and effect estimates (Dekkers and Hospital, 2002; Schön advances in marker technologies. Although genotyping
et al., 2004). Also, allelic diversity and genetic background is still a major expense, the declining costs per marker
effects that are present in a breeding program will not be data point have facilitated large-scale genotyping efforts
captured with a single biparental population. Therefore, in breeding programs. For example, the Monsanto Com-
multiple mapping populations are needed, QTL positions pany (St. Louis, MO) reported that from 2000 to 2006,
require validation, and QTL effects must be reestimated they experienced a sixfold decrease in cost per marker
by breeders in their specific germplasm. The validation data point and increased the volume of their marker data
in locally adapted germplasm is important because poor by 40-fold (Eathington et al., 2007). The availability of
estimates of the numerous small-effect QTL will lead to abundant markers and the reduction of genotyping costs
gains from MAS that are inferior to traditional pheno- will present new tools for plant breeders only if statistical
typic selection (Bernardo, 2001). Therefore, the resources methodologies for the utilization of genomewide marker
required for QTL detection coupled with validation and coverage are developed.
effect reestimation limit the effectiveness of biparental
population derived QTL for MAS in plant breeding pop- GENOMIC SELECTION
ulations (reviewed by Holland, 2004). FOR BREEDING VALUE ESTIMATION
To avoid this disconnect between biparental and breed- The two-step process of Lande and Thompson (1990) has
ing populations, linkage disequilibrium (LD)–based map- been criticized as an inefficient use of available data (Meu-
ping can be used for dissecting complex traits in breeding wissen et al., 2001): one would rather want to use all avail-
populations that already have extensive phenotypic data able data in a single step to achieve maximally accurate
across locations and years ( Jannink et al., 2001; Rafalski, estimates of marker effects. Genomic selection is a form of
2002). This strategy avoids the need to develop special MAS that simultaneously estimates all locus, haplotype,
mapping populations that impose an additional burden on or marker effects across the entire genome to calculate
breeding programs. Also, mapping within breeding pop- genomic estimated breeding values (GEBVs; Meuwissen et
ulations will allow for QTL identification and allelic value al., 2001). This approach contrasts greatly with traditional
estimates that can be directly utilized by MAS without MAS because there is not a defined subset of significant
the need for extensive validation (Breseghello and Sor- markers used for selection. Instead, GS analyzes jointly all
rells, 2006; Holland, 2004). However, low heritability, markers on a population attempting to explain the total
small population sizes, few large-effect QTL, confounding genetic variance with dense genomewide marker cover-
population structure, and arbitrary significance thresholds age through summing marker effects to predict breeding
found in current association mapping efforts allow iden- value of individuals (Meuwissen et al., 2001).
tification of only a few QTL with overestimated effects The central process of GS is the calculation GEBVs
(Beavis, 1998; Schön et al., 2004; Xu, 2003a). for individuals having only genotypic data using a model
To minimize the limitations for successful MAS, that was “trained” from individuals having both pheno-
Lande and Thompson (1990) proposed a visionary two- typic and genotypic data (Fig. 1; Meuwissen et al., 2001).
step approach: (i) select significant markers from large The population of individuals with both phenotypic and
marker sets, and (ii) combine phenotypic information genotypic data is known as the “training population”

2 WWW.CROPS.ORG CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009


as it is used to estimate model parameters
that will subsequently be used to calculate
GEBVs of selection candidates (e.g., breed-
ing lines) having only genotypic data (Fig.
1). These GEBVs are then used to select the
individuals for advancement in the breeding
cycle. Therefore, selection of an individual
without phenotypic data can be performed
by using a model to predict the individual’s
breeding value (Meuwissen et al., 2001).
To maximize GEBV accuracy, the training Figure 1. Diagram of genomic selection (GS) processes starting from the training
population must be representative of selec- population and selection candidates continuing through to genomic estimated
tion candidates in the breeding program to breeding value (GEBV)–based selection. Note that while we show here a single
occurrence of model training, training can be performed iteratively as new phenotype
which GS will be applied. and marker data accumulate.
Historically, estimated breeding values
(EBVs) for quantitative traits have been cal-
culated by best linear unbiased prediction (BLUP) based et al., 2001), and 15 to 20 kb in a diversity panel of sor-
only on phenotypic data of individuals and their relatives ghum (Sorghum bicolor; Hamblin et al., 2005). Examples
(Henderson, 1984). The use of EBVs via BLUP has been from diversity panels may give rough predictions of LD
popular in animal breeding and in recent years has been decay in a species, but because many factors affect LD,
used by plant breeders (reviewed by Piepho et al., 2007). individual breeding programs will need to determine LD
However, data on markers linked to known QTL can also decay rates on a case-by-case basis in their specific breed-
be used for calculation of EBVs (Fernando and Grossman, ing populations.
1989); this method was predicted to increase gains from Linkage disequilibrium estimates can be used to deter-
selection in animal breeding up to 38% (Meuwissen and mine target marker densities for GS. For example, Calus
Goddard, 1996). These results were encouraging, but they and Veerkamp (2007) used the average r 2 between adjacent
require extensive prior QTL discovery efforts in non- markers as a measure of their marker density relative to the
breeding populations. decay of LD. They found that for a high heritability trait,
average adjacent marker r 2 of 0.15 was sufficient, but for
MARKER DENSITY AND a low heritability trait, increasing the r 2 to 0.2 improved
LINKAGE DISEQUILIBRIUM the accuracy of GEBV predictions. These marker densities
Genomic selection differs from current MAS strategies may still be out of reach for some crops or populations.
because instead of only using markers that have a pre- Looking to the near future, however, high throughput
defined significant correlation with a trait, all markers are sequencing has made marker discovery affordable for most
used to estimate breeding values for each genotype. Conse- crop species, and the continued reduction of genotyping
quently, dense marker coverage is needed to maximize the costs will facilitate dense genomewide marker coverage
number of QTL in LD with at least one marker, thereby for all crop species (reviewed in Zhu et al., 2008). Note
also maximizing the number of QTL whose effects will that the conditions of complete genome saturation and of
be captured by markers. Target marker density will be at least one marker in LD with each QTL need not be
dictated by the rate of LD decay across the genome, as met to derive useful prediction models for GEBV. While
assessed by the relationship between intermarker coeffi- it is tempting to surmise a minimum number of markers
cient of determination, r 2, and genetic distance. needed to obtain useful GEBVs, the many factors affect-
Rate and pattern of LD decay are affected by popula- ing this number and the lack of empirical results currently
tion characteristics such as evolutionary history, mating available would make any guess meaningless. Clearly, this
system, population size, admixture, recombination rate, subject requires urgent attention.
and selection effects (Gaut and Long, 2003). Therefore,
LD decay rates are highly variable among species, popula- STATISTICAL MODELS
tions, and genomic regions. Examples of this variability AND PERFORMANCE
in LD decay rates include the following: 75 to 500 kb The challenge of QTL analysis is the selection of the
in a diversity panel of rice (Oryza sativa; Mather et al., appropriate statistical model to identify QTL and esti-
2007), 10 to 20 cM (roughly 50–100 Mb) in elite culti- mate their effects (Broman and Speed, 2002). In breeding
vars of wheat (Triticum aestivum; Chao, 2007; Maccaferri programs, statistical methods for GS will need to simul-
et al., 2005), 0.1 to 1.5 kb in diverse inbred lines of maize taneously estimate many marker effects from a limited
(Zea mays ssp. mays; Remington et al., 2001; Tenaillon number of phenotypes. A greater number of explanatory

CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009 WWW.CROPS.ORG 3


variables (markers) than observations (phenotyped lines) (Moreau et al., 2004). These complications were probably
leads to a lack of degrees of freedom that must be handled consequences of SR that detects only large effects and that
through the selection and use of the most appropriate sta- overestimates effects.
tistical model, that is, the model that results in the highest In a GS simulation by Meuwissen et al. (2001), SR
GEBV accuracy with consideration of model complex- resulted in low GEBV accuracy due to limited detec-
ity and computation requirements. In the assessment of tion of QTL. The simulated outcrossing population had
model performance, GEBV accuracy has a precise defi ni- an effective population size of 100 with a trait heritabil-
tion, namely, the Pearson correlation between the GEBV ity of 0.5. After 1000 generations of random mating to
and the true breeding value (TBV). Accuracy defined in establish mutation-drift equilibrium, generation 1001 had
this way is directly proportional to gain from selection a population size of 200 (100 males; 100 females). Two
when selecting on the GEBV, that is, R = irσA, where R generations (1002 and 1003) of size 2000 with 20 half-
is the response, i is the selection intensity, r is the accu- sib families of size 100 individuals were then simulated.
racy defined above, and σA is the square root of the addi- Generations 1001 and 1002 were used to train the model
tive genetic variance of TBV (Falconer and Mackay, 1996, while GEBV accuracy was calculated on generation 1003.
p. 189). We briefly describe here three models: stepwise Genotypic data consisted of 101 multi-allelic markers on
regression, ridge regression, and Bayesian estimation. each of 10 chromosomes of length 100 cM. Adjacent pairs
of markers were considered haplotypes such that 50,000
Stepwise Regression for MAS haplotype effects were estimated. The accuracy of GEBV
Traditional MAS considers marker effects as fi xed, requir- for SR (0.318) was less than that expected for strictly phe-
ing stepwise regression (SR) approaches that avoid the lack notype-based BLUP (about 0.4; Meuwissen et al., 2001).
of degrees of freedom problem by fitting markers singly In agreement with Lange and Whittaker (2001), Meuwis-
or in small groups. After the model selection process dur- sen et al. (2001) concluded that SR’s procedure to identify
ing which markers are added or removed from the model marker subsets is suboptimal for MAS in situations where
on the basis of arbitrary significance thresholds, nonsig- the majority of the additive genetic variance is generated
nificant markers are assigned an effect of zero and signifi- by many QTL. Note, however, that the GEBV accuracy of
cant marker effects are simultaneously tested to estimate SR depends on the details of the analysis: using the Meu-
their effects. This stepwise approach to set nonsignificant wissen et al. (2001) simulation design, Habier et al. (2007)
marker effects to zero is critical for maintaining model found that SR produced an accuracy of 0.61. Habier et al.
estimability (Lande and Thompson, 1990). Significance (2007) attributed this difference to the use of a less-strin-
thresholds that may maximize response to selection can- gent significance threshold than was used by Meuwissen
not be determined analytically, although guidelines have et al. (2001). This conclusion was supported by simula-
been established through simulation (Hospital et al., 1997; tions showing prediction accuracy changes with changes
Moreau et al., 1998). The general guideline is that liberal in significance thresholds (Piyasatian et al., 2007).
p-value thresholds improve selection gain (Hospital et al.,
1997; Moreau et al., 1998). Nevertheless, when only sig- Ridge Regression BLUP
nificant marker effects are estimated, only a portion of the for Genomic Selection
genetic variance will be captured (Goddard and Hayes, The ridge regression BLUP (RR-BLUP) method can
2007) and effects retained in the model can be greatly simultaneously estimate all marker effects for GS (Meu-
overestimated (Beavis, 1998; Hayes, 2007), particularly wissen et al., 2001; Whittaker et al., 2000). Rather than
when many effects are tested. categorizing markers as either significant or as having no
Limitations of SR for MAS in practice were reported effect, ridge regression shrinks all marker effects toward
by Moreau et al. (2004). In 300 test-crossed maize prog- zero (Breiman, 1995; Whittaker et al., 2000). The method
enies evaluated in 14 trials over 11 locations for dry grain makes the assumption that markers are random effects
yield and grain moisture, they discovered 16 QTL for with a common variance (Meuwissen et al., 2001; Table
dry grain yield and 12 QTL for grain moisture explain- 1). Equal variance does not assume that all markers have
ing 50% of the total phenotypic variance of both traits. the same effect (Bernardo and Yu, 2007) but that marker
When using an index combining phenotypic and marker effects are all equally shrunken toward zero. Nevertheless,
information for a single cycle followed by two cycles of the assumption that individual markers have the same vari-
marker-only selection, they observed no genetic gain ance is unrealistic, and therefore, RR-BLUP incorrectly
from the two cycles of marker selection (Moreau et al., treats all effects equally (Xu, 2003b). Despite the incorrect
2004). They suggested that this inefficiency of MAS could assumption of equal marker variance, RR-BLUP is supe-
be caused by fi xation of major effect loci in the first cycle rior to SR because it can simultaneously estimate effects
of selection and inaccurate estimation of remaining effects for all markers: by avoiding marker selection, it avoids the
resulting in no gain from the cycles of marker selection biases that go with that selection (Whittaker et al., 2000).

4 WWW.CROPS.ORG CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009


Also, a ridge regression approach is more appropriate than (Muir, 2007), the presumably incorrect assumption that
SR for instances in which there are few or no large effects underlies it can lead to overshrinking of large effects (Table
and many small effects (Breiman, 1995), as is the case with 1; Meuwissen et al., 2001; Xu, 2003b). Bayesian meth-
most quantitative traits. ods have been adopted to relax this assumption and bet-
In the simulation by Meuwissen et al. (2001), RR- ter model marker effects of differing sizes (Hayes, 2007).
BLUP had a GEBV accuracy of 0.732, which was 41 and Here, a separate variance is estimated for each marker, and
33% greater than SR and phenotype-based BLUP, respec- the variances are assumed to follow a specified prior dis-
tively. With higher SR significance thresholds, Habier et tribution (Meuwissen et al., 2001).
al. (2007) reported that RR-BLUP resulted in 4 and 11% Meuwissen et al. (2001) proposed two types of prior
increase in GEBV accuracy compared to SR and tradi- distribution for the marker variance. The first type of prior
tional BLUP, respectively. In addition to these studies, (BayesA) uses an inverted chi-square distribution with
Muir (2007) simulated 512 genotypes with a low herita- degrees of freedom and scale parameters chosen so that the
bility trait (h2 = 0.1) in each of four training generations. mean and variance of the distribution match the expected
These conditions resulted in an even higher RR-BLUP mean and variance of the marker variances. In the simula-
GEBV accuracy of 0.83 despite the lower heritability. tion design described above, BayesA outperformed both
This gain in GEBV accuracy was attributed to the four SR and RR-BLUP with a GEBV accuracy of 0.798. Dif-
training generations used by Muir (2007), as opposed to ferent parameter values for the BayesA inverted chi-square
two generations used in previous studies (Habier et al., prior distribution have also been proposed that place much
2007; Meuwissen et al., 2001). higher density on marker variances close to zero, thereby
In a GS simulation on a population derived from a forcing more marker effect estimates close to zero (ter
biparental cross of maize inbreds, Bernardo and Yu (2007) Braak et al., 2005; Xu, 2003b).
found that relative to phenotypic selection, the increase in The BayesA method of Xu (2003b) was applied to
selection gain from RR-BLUP was 18% greater than that data from a doubled haploid barley (Hordeum vulgare) pop-
from SR for a highly heritable trait (h2 = 0.8) controlled by ulation of 145 lines with 127 single nucleotide polymor-
20 QTL. For a trait with low heritability (h2 = 0.2) con- phism markers covering 1500 cM for yield, heading date,
trolled by 100 QTL, the increase in selection gain from maturity, test weight, lodging, and kernel weight. Xu
RR-BLUP was 43% greater than that from SR (Bernardo (2003b) reported that SR and BayesA both found large-
and Yu, 2007). Similar results were observed by Piyasatian effect QTL but that BayesA provided better QTL location
et al. (2007), who found that in the fi rst round of selec- and effect estimation. Also, in simulation of a population
tion in a simulated cross between two inbred parents, gain derived from a biparental inbred cross, ter Braak et al.
from selection from RR-BLUP was 109 and 32% greater (2005) found that BayesA prior parameters forcing more
than that of traditional BLUP and SR, respectively. marker effect shrinkage gave better estimates of QTL
effects than did the Meuwissen et al. (2001) parameters.
Bayesian Estimation A comparison of these different prior parameterizations in
The simplifying assumption of equal and fi xed marker an association genetics rather than linkage mapping con-
effect variances allows RR-BLUP parameters to be effi- text has not been done.
ciently computed using maximum likelihood methods The second type of prior distribution Meuwissen et al.
(Meuwissen et al., 2001). While RR-BLUP can provide a (2001) proposed (BayesB) contrasts with BayesA by having
conservative EBV by shrinking all marker effects equally a prior mass at zero, thereby allowing for markers with no
Table 1. General characteristics and trends of performance for traditional best linear unbiased predictor (BLUP) and genomic
selection methods. Note that these are general summaries based on current understanding of model performance.
Marker effect; Proportion Performance with increased Inbreeding
Large-effect Small-effect
Method variance of markers Marker QTL† depression; loss
QTL QTL
assumptions fitted in model density number of diversity
Traditional N/A N/A N/A N/A Captured only Captured only Yes
BLUP by phenotype by phenotype
Stepwise regression Fixed Subset Reduced Reduced Overestimated Excluded Marginally Reduced
RR-BLUP‡ Random; Equal All Reduced§ Increased Underestimated Captured Reduced
BayesA Random; All ? Reduced More accurately Captured Reduced
Unique All > 0 estimated
BayesB Random; All Insensitive§ Reduced More accurately Captured Reduced
Unique Some = 0 estimated

QTL, quantitative trait locus.

RR, ridge regression.
§
Source: Fernando (2007).

CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009 WWW.CROPS.ORG 5


effects. The inverted chi-square prior of BayesA may be set densities and colinearities; these will need to be resolved
to strongly regress variances toward zero, but it does not by improved statistical methods (ter Braak et al., 2005).
permit the value of zero itself. BayesB thus presents a more
realistic prior because we expect that some regions of the INCLUSION OF A POLYGENIC EFFECT
genome will carry no QTL, so that some markers should TERM ACCOUNTING FOR KINSHIP
have estimates of zero effect. The results from Meuwissen Phenotypic information from relatives contribute to an
et al. (2001) showed that BayesB had a GEBV accuracy individual’s EBV because EBVs vary according to the
of 0.848, greater than all other methods tested. Of the additive relationship (A) matrix, that is, a matrix that con-
Bayesian methods, BayesB not only was more accurate but tains for each pair of individuals the proportion of alleles
was also less computationally demanding. Meuwissen et for which they are identical by descent (van Arendonk et
al. (2001) concluded that Bayesian methods outperformed al., 1994; Lynch and Walsh, 1998, p. 751). When markers
RR-BLUP through better estimation of large-effect QTL are introduced into the analysis, some genetic effects will
by allowing for unequal variances. be captured by markers in LD with QTL, but residual
de Roos et al. (2007) used Bayesian modeling as genetic effects will still be assumed to vary according to
described by Meuwissen and Goddard (2004) in actual the A matrix. These residual effects can be captured by
dairy cattle data for a single chromosome containing 32 including a polygenic term in the model. In association
markers with one being a known causal mutation for fat mapping, the inclusion of this matrix has been popular-
percentage. They compared Bayesian GS that used all ized as a statistical control for population structure and
marker information to regression on the genotype at the familial relatedness (Yu et al., 2006; Zhao et al., 2007).
known causal mutation and to traditional BLUP with no The A matrix can be calculated on the basis of the ped-
markers. Using a cross validation population of 1135, they igree or the marker data, with pedigree information pro-
concluded that Bayesian GS and regression on the causal viding exact expected relationships and markers providing
mutation had similar accuracies (0.752 and 0.746, respec- estimated realized relationships. When marker number is
tively), with both being superior to traditional BLUP (EBV high enough that marker sampling plays a minor role (i.e.,
accuracy of 0.508). Interestingly, the GS analysis often did relationship estimates on the basis of markers are accurate),
not place the causal mutation in the correct marker bracket marker-estimated relationships will better reflect true
but was nevertheless able to calculate accurate GEBV. This relationships than will pedigree-expected relationships. In
robustness of GEBV accuracy provides evidence that GS particular, four mechanisms lead realized relationships to
can perform well for breeders in the absence of the discov- diverge from their expectation: random Mendelian seg-
ery of QTL (de Roos et al., 2007). regation, segregation distortion, selection, and pedigree
In the future, genotyping costs will decrease, but it recording errors. For example, parental contributions to
is unlikely that phenotyping costs will also decrease, thus inbreds vary from their expected 50% because of random
shifting goals toward reducing phenotyping and increas- Mendelian segregation during selfing. For the genomes
ing genotyping. Bernardo and Yu (2007) suggested this of maize and wheat, there is a 10% probability that single
shift would be feasible when the cost of a marker data seed decent–derived inbreds will have less than 38 and
point is 5000 times less than the cost of phenotyping a 43% genome contribution from one parent, respectively
single entry. Regardless of the threshold, it is desirable (Frisch and Melchinger, 2007).
to decrease the number of phenotypic records needed for The value of including a polygenic effect term in the
training models for accurate GEBVs. Simulations by Meu- model will depend strongly on marker density available in
wissen et al. (2001) showed that with 2200 phenotypic the study for two reasons. First, if density is such that all
records, RR-BLUP and BayesB had GEBV accuracies of QTL are in strong LD with a marker, all genetic effects
0.732 and 0.848, respectively. When the number of pheno- will be absorbed by markers and none will be left for the
typic records was reduced to 500, RR-BLUP and BayesB polygenic term to capture (Bernardo and Yu, 2007; Meu-
GEBV accuracies decreased to 0.579 and 0.708, respec- wissen et al., 2001; Zhong and Jannink, 2007). Second,
tively (Meuwissen et al., 2001). Thus, the effect of low even markers that are in linkage equilibrium with all QTL
numbers of phenotypic records was less severe for BayesB carry information about relationships among individuals,
than for RR-BLUP. In addition, Fernando (2007) found and this information contributes to the accuracy of GEBV
that in contrast to RR-BLUP, BayesB’s GEBV accuracy (Habier et al., 2007). Indeed, this contribution depends on
did not decline as the number of markers increased. These the number of markers included in the GS method, and
findings suggest that Bayesian methods may be better because SR uses only a subset of markers, it benefits least
suited to handling situations with increased colinearity from genetic relationship contributions to GEBV accu-
between markers caused by extremely large markers sets racy (Habier et al., 2007).
and limited phenotypic records (Table 1). Computational Research to look explicitly at the value of including a
issues may arise for Bayesian methods under high marker polygenic effect term used adjacent-marker r 2 as a measure

6 WWW.CROPS.ORG CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009


of marker density. For a high heritability trait (h2 = 0.5), selected, thus lowering the effective population size and
the polygenic effect term increased GEBV up to an adja- thereby increasing the loss of genetic variability. Tradi-
cent-marker r 2 of 0.14, while for a low heritability trait (h2 tional BLUP increases EBV accuracy by incorporating
= 0.1), the term made no difference already at an r 2 of 0.11 ancestor and collateral relative phenotypes in the calcula-
(Calus and Veerkamp, 2007). At lower adjacent-marker r 2, tion (Henderson, 1984). But including family information
the polygenic term fulfi lls its role of explaining genetic in EBV calculation increases the correlation between EBV
variance not absorbed by markers, and it therefore con- of family members, making it more likely that multiple
tributes to GEBV accuracy (Calus and Veerkamp, 2007; sibs will be selected (Wray and Thompson, 1990). Sibling
Villanueva et al., 2005). coselection, in turn also reduces effective population size.
Therefore, while increased selection intensity and a higher
SELECTION INDEX THEORY EBV accuracy lead to greater short-term gains from selec-
APPLIED TO GENOMIC SELECTION tion, they both may reduce long-term gains by decreas-
A selection index integrates and weights multiple traits ing genetic variation and increasing rates of inbreeding
to achieve greater gains than if traits with independent (Quinton et al., 1992).
thresholds are individually or collectively selected (Hazel Daetwyler et al. (2007) reviewed these issues and
and Lush, 1942; Hazel, 1943). Selection indices can incor- determined that GS differs from simple phenotypic
porate marker data as indirect selection traits (Lande and selection and traditional BLUP by using markers to more
Thompson, 1990; Neimann-Sorensen and Robertson, accurately estimate Mendelian sampling variation, that is,
1961; Smith, 1967). However, current MAS applied to deviations between siblings within families. Mendelian
loci selected by SR violates the selection index assump- sampling variation, generated by random segregation, is
tions of multivariate normality and small changes in created anew each generation. Selecting strictly on this
allele frequencies because selection is based on only few variation therefore enables sustained genetic progress by
large effect loci (Dekkers, 2007; Lande and Thompson, decreasing coselection of sibs and thus reducing inbreed-
1990). Because GS is based on many markers distributed ing and the loss of genetic variation (Woolliams et al.,
throughout the genome, index selection assumptions are 1999). Optimized selection schemes have been proposed
met, providing an opportunity to use index selection the- where parent combinations are restricted by their level
ory to predict response to GS (Dekkers, 2007). of coancestry to limit the loss of genetic variation and
Dekkers (2007) used selection index theory by add- the rate of inbreeding (Grundy et al., 1998; Meuwis-
ing marker-derived breeding values as a separate corre- sen, 1997). In these schemes, an individual’s selective
lated trait to the selection index (Lande and Thompson, advantage depends largely on the Mendelian sampling
1990). In a simulated pig breeding program, selection on term, that is, on its performance relative to its siblings
only marker data could outperform phenotypic selection (Avendaño et al., 2004). Unlike traditional BLUP based
for low heritability traits (0.1) even with moderate GEBV on pedigree data that account for average relationships,
accuracy (0.55). When marker and phenotypic data were tracking markers enables GS to also track the random
both used for a single trait, even greater accuracies were segregation that makes up the Mendelian sampling term.
observed. This increase was due to marker information The benefit is both more accurate EBVs and decreased
that allowed for within family selection (Dekkers, 2007). correlation between EBVs within families, countering
For two negatively correlated traits with heritabilities the mechanism whereby the use of family informa-
of 0.3 and 0.1, Dekkers (2007) found using only markers tion increases loss of genetic diversity (Daetwyler et
increased gains from selection over phenotypic selection al., 2007). Note that the greater emphasis placed by GS
by 8.5% for the index of the two traits and 66% for the low on the Mendelian sampling term does not completely
heritability trait alone. Using both markers and phenotype negate variable long-term genetic contributions among
increased gains from selection over phenotypic selection individuals and its consequent increase in inbreeding
by 21% for the index of the two traits and 80.5% for the rate. In particular, superior individuals carry superior
low heritability trait alone. These results show the poten- alleles, and selection of those alleles will, in turn, lead
tial of GS to increase gains for multiple traits especially their carriers to leave more off spring behind (Daetwyler
in cases where phenotypic data is available on selection et al., 2007). Thus, it still may be advisable to manage
candidates and traits have low heritability. rates of inbreeding (e.g., Avendaño et al., 2004) even in
the context of GS. Nevertheless, the advantages of GS
MAINTAINING GENETIC DIVERSITY AND in regard to inbreeding and the maintenance of genetic
REDUCING INBREEDING DEPRESSION diversity should prove valuable for crops such as alfalfa
Gains from selection can be increased by raising the selec- (Medicago sativa) that suffer from inbreeding depression
tion intensity or the accuracy of EBV of breeding lines. and for maintaining genetic variation in advanced cycle
Increased selection intensity reduces the number of lines breeding programs.

CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009 WWW.CROPS.ORG 7


GAINS FROM SELECTION and locations, and even possibly other breeding programs.
PER UNIT TIME This information sharing should provide GS with stability
Marker-assisted selection strategies increase gain mainly in the face of G×E.
through gain per unit time, rather than gain per cycle (Ber- Anticipating the effect of epistasis on the potential of
nardo and Yu, 2007; Edwards and Johnson, 1994; Hospital GS is difficult. Almost all GS prediction accuracy eval-
et al., 1997; Koebner and Summers, 2003; Meuwissen et uations derive from simulations that adopted additive
al., 2001; Muir, 2007). To ascertain GS’s impact on gains genetic models. There is current debate, at both theoreti-
per unit time, Schaeffer (2006) suggested a plan for imple- cal and empirical levels, of the likely importance of epista-
menting GS into a dairy breeding program. Through sis in the architecture of quantitative traits (Carlborg and
reduction in time and costs needed to prove the value of a Haley, 2004; Hill et al., 2008; Holland, 2007; Mackay,
bull, assuming a GEBV accuracy of 0.75, Schaeffer (2006) 2008). To examine this issue, it is essential to distinguish
determined that GS could provide a twofold increase in between the genotypic value versus the breeding value of
rate of genetic gain and save 92% of the costs of the cur- a line (Falconer and Mackay, 1996). The genotypic value
rent progeny test based breeding program. is the expected phenotype of the line given its genotype
In plants, the importance of generation time varies and includes additive and nonadditive genetic effects. The
between crops, but the goal of reducing cycle time remains. breeding value is the expected phenotype of line’s prog-
In maize, a crop that uses doubled haploids and off-season eny and includes only additive effects. The additive mod-
nurseries, test cross performance selection still requires at els used by GS should predict the breeding value rather
least 2 yr (Bernardo and Yu, 2007), providing an oppor- than genotypic value (Goddard and Hayes, 2007). Conse-
tunity for GS to reduce unit time per selection cycle by quently, correlations between GEBVs and line phenotypes
reducing the need for progeny test data in every cycle. In may well be lower than those obtained in additive effect
the more extreme case of oil palm (Elaeis guineensis Jacq.), simulations, but they should nevertheless reflect a line’s
which takes 19 yr to complete a cycle of selection, Wong value as a parent. For cases in which estimates of genotypic
and Bernardo (2008) reported that GS reduced the selection value are desired in the presence of epistasis, methods are
cycle to 6 yr. Even with small population sizes (N = 50) that currently being developed and tested (e.g., Gianola et al.,
adversely effected GEBV accuracy, their simulations indi- 2006; Gianola and van Kaam, 2008; Gonzalez-Recio et
cated that GS would outperform MARS and phenotypic al., 2008). Further empirical evaluation of the prediction
selection when considering gain per unit cost and time. accuracies of these methods should help address the ongo-
ing debate over the importance of epistasis in the mapping
GENOTYPE × ENVIRONMENT of genotype to phenotype. Because of the small contri-
INTERACTIONS AND EPISTASIS bution that epistasis makes to breeding value (Holland,
Genotype × environment (G×E) interaction is a challenge 2001), GS using simpler additive models should be effec-
in plant breeding because the large number of experimen- tive for maximizing gain from selection.
tal lines and environments, that is, locations and years,
make it impossible to test a line in all possible environ-
mental conditions of a breeding program’s target region FUTURE DIRECTIONS
(Allard and Bradshaw, 1964). Consider, however, that the Statistical Methods
genotype of any line is composed of alleles that, over time, A statistical model will more faithfully capture QTL infor-
will have been evaluated in a larger sample of target envi- mation as its assumptions about the underlying genetic
ronments than would be feasible for any particular line. architecture, made explicit in the prior distributions of
Thus, it may be possible to accurately predict GEBV even QTL effects or variance, are more correct (Meuwissen et
in the presence of high G×E. As an extreme example, for al., 2001). There are two obstacles to translating this fact
winter annual crops, a severe winter may only occur once into improved models. First, GS may gain in accuracy
a decade. Variety releases for the region need to be hardy to not just by capturing more QTL information but also by
such winters because crop failure even once per decade is better capturing relationship information (Habier et al.,
too frequent. With GS, a given generation of experimental 2007). There may be a tradeoff between the kinds of prior
lines need never experience a test winter if the alleles they distributions of effects that promote the use of these two
carry were characterized during a severe winter. Simi- information sources (Habier et al., 2007). Second, we sim-
lar cases include the infrequent but devastating conditions ply do not know, for any complex trait, what the under-
caused by severe drought, flooding, disease pressure, and lying genetic architecture is, and thus, we do not have
insect infestation. The broader insight that these examples adequate prior knowledge at our disposal. Therefore, sta-
illustrate is that with GS, lines are not evaluated solely on tistical models that are relatively insensitive to the under-
the basis of their own phenotypic performance, but on the lying architecture may be optimal for most populations,
basis of information shared across other lines, other years although identifying those models remains challenging.

8 WWW.CROPS.ORG CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009


Finally, the marker technologies on which GS meth- indexeng.html), and the Canadian COOL-DUDE (Yan
ods depend are constantly changing. Next-generation and Tinker, 2007). Adaptation of these tools to link with
sequencing technologies and improvement of genotyping GS and development of user-friendly GS analyses them-
platforms present breeders with powerful tools for char- selves are needed to take GS from theory to practice.
acterizing the genetic composition of their germplasm. As
these technologies continue to evolve, they will provide CHANGES TO BREEDING
quantitatively and qualitatively different information (e.g., PROGRAM STRUCTURE
copy number and epigenetic variation; Stranger et al., The accuracies of GEBV observed in research offer the
2007; Zhang et al., 2008), and statistical machinery will possibility that future elite and parental lines will be
also need to evolve to use this information efficiently to selected on their GEBV rather than on their phenotypic
increase prediction accuracy. records from extensive field testing. The most immediate
impact of this circumstance would be a great increase in
Software and Database Development the speed of the breeding cycle (Fig. 2; Wong and Ber-
While statistical methods of prediction must be continually nardo, 2008), thereby increasing selection gains per unit
advanced, an integral part of their performance will be the time. This shift would also fundamentally alter the role of
software packages used to implement them. In conjunc- phenotyping in plant breeding (Fig. 2). Note that Fig. 2
tion with this software, robust databases that can efficiently offers a somewhat futuristic view of the use of GS, contin-
link breeding lines, testing environments, genotypic data, gent on its validation in practice. We do not, at this point,
phenotypic data, and breeding programs will need to be advocate dispensing with phenotypic evaluation before
developed to simplify flow and use of information. While parent selection.
private breeding companies have invested heavily in data The purpose of phenotyping now is to select the best
management systems that will likely be efficient in execut- lines from a segregating population and to evaluate fewer
ing GS (e.g., Eathington et al., 2007), public sector breed- lines with greater replication in each cycle of selection.
ing programs also need database software that integrates But in a GS driven breeding cycle, the purpose of phe-
the wide variety of data they generate (Heckenberger et notyping is to estimate or reestimate marker effects. It is
al., 2008; Tinker and Yan, 2006). Recent developments in far from clear at this point whether it will be advanta-
the public sector are promising, such as the Barley Coor- geous to evaluate only the best lines or to evaluate few
dinated Agricultural Project Hordeum Toolbox (http:// lines with high replication. Figure 2 therefore separates
hordeumtoolbox.org/), the GDPDM database schema the germplasm improvement cycle from the prediction
that links with the association analysis software TASSEL model improvement cycle. Indeed, if we use the guide-
(http://www.maizegenetics.net); the German GABI- lines for optimal QTL linkage mapping, evaluation should
BRAIN project (http://brain.uni-hohenheim.de/eng/ include not just the best but the best and the worst lines

Figure 2. Flow diagram of a genomic selection breeding program. Breeding cycle time is shortened by removing phenotypic evaluation of
lines before selection as parents for the next cycle. Model training and line development cycle length will be crop and breeding program
specific. (GEBV = genomic estimated breeding value.)

CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009 WWW.CROPS.ORG 9


(Darvasi and Soller, 1992; Lander and Botstein, 1989) and Avendaño, S., J. Woolliams, and B. Villanueva. 2004. Mendelian
sampling terms as a selective advantage in optimum breeding
many unreplicated lines instead of few replicated lines schemes with restrictions on the rate of inbreeding. Genet.
(Knapp and Bridges, 1990). Figure 2 also emphasizes the Res. 83:55–64.
need for model updating and reevaluation. Marker effects Beavis, W.D. 1998. QTL analyses: Power, precision, and accuracy.
may change as a result of allele frequency changes (Muir, p. 145–162. In A.H. Patterson (ed.) Molecular dissection of
2007) or of epistatic gene action. Model updating with complex traits. CRC Press, Boca Raton, FL.
each breeding cycle should mitigate reduced gains from Bernardo, R. 2001. What if we knew all the genes for a quantita-
GS caused by these mechanisms. Thus, GS could radically tive trait in hybrid crops? Crop Sci. 41:1–4.
Bernardo, R., and J. Yu. 2007. Prospects for genomewide selection
change the practice of field evaluation for breeders. Of
for quantitative traits in maize. Crop Sci. 47:1082–1090.
course, regardless of the breeding method used, final field Breiman, L. 1995. Better subset regression using the nonnegative
evaluations of varieties across the target environments will garrote. Technometrics 37:373–384.
be needed before they are distributed to farmers. Breseghello, F., and M.E. Sorrells. 2006. Association mapping of
GS may also diminish the need for breeders to select kernel size and milling quality in wheat (Triticum aestivum L.)
parents strictly from the set of lines evaluated in their tar- cultivars. Genetics 172:1165–1177.
get environments (Goddard and Hayes, 2007). Once a Broman, K.W., and T.P. Speed. 2002. A model selection approach
predictive linear model is established for their target envi- for the identification of quantitative trait loci in experimental
crosses. J. R. Statist. Soc. Ser. B 64:641–656.
ronments, any genotype with high target environment
Calus, M., and R. Veerkamp. 2007. Accuracy of breeding val-
specific GEBV will become a candidate. Thus, GS should ues when using and ignoring the polygenic effect in genomic
facilitate germplasm exchange and increase the probabil- breeding value estimation with a marker density of one SNP
ity of selecting useful germplasm. per cM. J. Anim. Breed. Genet. 124:362–368.
Carlborg, O., and C.S. Haley. 2004. Epistasis: Too often neglected
CONCLUSIONS in complex trait studies. Nat. Rev. Genet. 5:618–625.
It has been predicted for more than two decades that Chao, S. 2007. Evaluation of genetic diversity and genome-wide
molecular marker technology would reshape breeding linkage disequilibrium among US wheat (Triticum aestivum L.)
germplasm representing different market classes. Crop Sci.
programs and facilitate rapid gains from selection (Stuber 47:1018–1030.
et al., 1982; Tanksley et al., 1989). The failure of cur- Crosbie, T.M., S.R. Eathington, G.R. Johnson, M. Edwards, R.
rent MAS to significantly improve polygenic traits has Reiter, S. Stark, et al. 2003. Plant breeding: Past, present,
thwarted this prediction. Genomic selection looks to and future. p. 1–50. In K.R. Lamkey and M. Lee (ed.) Plant
fulfi ll it by using genomewide marker coverage to accu- Breeding: The Arnel R. Hallauer Int. Symp., Mexico City.
rately estimate breeding values, accelerate the breeding 17–23 Aug. 2003. Blackwell, Oxford, UK.
cycle, and introduce greater flexibility in the relation- Daetwyler, H.D., B. Villanueva, P. Bijma, and J.A. Woolliams.
2007. Inbreeding in genome-wide selection. J. Anim. Breed.
ship between phenotypic evaluation and selection. To do
Genet. 124:369–376.
so, however, GS must shift from theory to practice. As Darvasi, A., and M. Soller. 1992. Selective genotyping for deter-
evident in this review and interpretation, GS has almost mination of linkage between a marker locus and a quantita-
exclusively been tested through simulation, and therefore, tive trait locus. Theor. Appl. Genet. 85:353–359.
its potential value should be assessed with cautious opti- de Roos, A.P., C. Schrooten, E. Mullaart, M.P. Calus, and R.F.
mism. The accuracy of GS and its cost effectiveness must Veerkamp. 2007. Breeding value estimation for fat percentage
now be evaluated in breeding programs to provide the using dense markers on Bos taurus autosome 14. J. Dairy Sci.
empirical evidence needed to warrant the addition of GS 90:4821–4829.
Dekkers, J.C.M. 2007. Prediction of response to marker-assisted
to the plant breeders’ toolbox.
and genomic selection using selection index theory. J. Anim.
Breed. Genet. 124:331–341.
Acknowledgments Dekkers, J.C.M., and F. Hospital. 2002. The use of molecular
The authors thank Adam Famoso, Michael Gore, and Jesse genetics in the improvement of agricultural populations. Nat.
Munkvold for their insights and critical reviews. Work of Elliot Rev. Genet. 3:22–32.
Heff ner was funded by the USDA National Needs Fellowship Eathington, S.R., T.M. Crosbie, M.D. Edwards, R.S. Reiter, and
Grant 2005-38420-15785. Work of Jean-Luc Jannink was par- J.K. Bull. 2007. Molecular markers in a commercial breeding
tially funded by the USDA-NRI Grant Numbers 2008-55301- program. Crop Sci. 47:S154–S163.
Edwards, M., and L. Johnson. 1994. RFLP for rapid recurrent
18746 and 2006-55606-16722. Additional funding for this
selection. p. 33–40. In Analysis of Molecular Marker Data:
research was provided by Hatch 149419.
Joint Plant Breeding Symp. Series. Am. Soc. of Horticultural
Sci. and CSSA, Corvallis, OR.
Falconer, D., and T. Mackay. 1996. Quantitative genetics. Long-
References
man, Harrow, UK.
Allard, R.W., and A.D. Bradshaw. 1964. Implications of geno-
Fernando, R.L. 2007. Genomic selection. Acta Agric. Scand. Ser.
type–environmental interactions in applied plant breeding.
Anim. Sci. 57:192–195.
Crop Sci. 4:503–508.

10 WWW.CROPS.ORG CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009


Fernando, R., and M. Grossman. 1989. Marker assisted selection Jannink, J.-L., M.C. Bink, and R.C. Jansen. 2001. Using complex
using best linear unbiased prediction. Genet. Select. Evol. plant pedigrees to map valuable genes. Trends Plant Sci. 6:337.
21:467–477. Kearsey, M., and A. Farquhar. 1998. QTL analysis in plants: Where
Frisch, M., and A.E. Melchinger. 2007. Variance of the parental are we now? Heredity 80:137–142.
genome contribution to inbred lines derived from biparental Knapp, S., and W. Bridges. 1990. Using molecular markers to estimate
crosses. Genetics 176:477–488. quantitative trait locus parameters: Power and genetic variances
Gaut, B.S., and A.D. Long. 2003. The lowdown on linkage dis- for unreplicated and replicated progeny. Genetics 126:769–777.
equilibrium. Plant Cell 15:1502–1506. Koebner, R.M.D., and R.W. Summers. 2003. 21st century wheat
Gianola, D., R.L. Fernando, and A. Stella. 2006. Genomic-assisted breeding: Plot selection or plate detection? Trends Biotech-
prediction of genetic value with semiparametric procedures. nol. 21:59–63.
Genetics 173:1761–1776. Lande, R., and R. Thompson. 1990. Efficiency of marker-assisted
Gianola, D., and J.B. van Kaam. 2008. Reproducing kernel hilbert selection in the improvement of quantitative traits. Genetics
spaces regression methods for genomic assisted prediction of 124:743–756.
quantitative traits. Genetics 178:2289–2303. Lander, E., and D. Botstein. 1989. Mapping Mendelian factors
Goddard, M., and B. Hayes. 2007. Genomic selection. J. Anim. underlying quantitative traits using RFLP linkage maps.
Breed. Genet. 124:323–330. Genetics 121:185–199.
Gonzalez-Recio, O., D. Gianola, N. Long, K.A. Weigel, G.J.M. Lange, C., and J.C. Whittaker. 2001. On prediction of genetic val-
Rosa, and S. Avendano. 2008. Nonparametric methods for ues in marker-assisted selection. Genetics 159:1375–1381.
incorporating genomic information into genetic evaluations: An Lee, M. 1995. DNA markers and plant breeding programs. Adv.
application to mortality in broilers. Genetics 178:2305–2313. Agron. 55:265–344.
Grundy, B., B. Villanueva, and J. Wooliams. 1998. Dynamic selec- Lynch, M., and B. Walsh. 1998. Genetics and analysis of quantita-
tion procedures for constrained inbreeding and their conse- tive traits. Sinauer, Sunderland, MA.
quences for pedigree development. Genet. Res. 72:159–168. Maccaferri, M., M.C. Sanguineti, E. Noli, and R. Tuberosa. 2005.
Habier, D., R. Fernando, and J. Dekkers. 2007. The impact of Population structure and long-range linkage disequilibrium
genetic relationship information on genome-assisted breeding in a durum wheat elite collection. Mol. Breed. 15:271–290.
values. Genetics 177:2389–2397. Mackay, T.F.C. 2008. The genetic architecture of complex behav-
Hamblin, M.T., M.G. Salas Fernandez, A.M. Casa, S.E. Mitchell, iors: Lessons from Drosophila. Genetica doi:10.1007/s10709-
A.H. Paterson, and S. Kresovich. 2005. Equilibrium processes 008-9310-6.
cannot explain high levels of short-and medium-range link- Mather, K.A., A.L. Caicedo, N.R. Polato, K.M. Olsen, S.
age disequilibrium in the domesticated grass Sorghum bicolor. McCouch, and M.D. Purugganan. 2007. The extent of
Genetics 171:1247–1256. linkage disequilibrium in rice (Oryza sativa L.). Genetics
Hayes, B. 2007. QTL mapping, MAS, and genomic selection. 177:2223–2232.
Available at http://www.ans.iastate.edu/section/abg/short- Meuwissen, T. 1997. Maximizing the response of selection with a
course/notes.pdf (verified 4 Nov. 2008). Animal Breeding & predefi ned rate of inbreeding. J. Anim. Sci. 75:934–940.
Genetics, Dep. of Animal Science, Iowa State Univ., Ames. Meuwissen, T.H.E., and M.E. Goddard. 1996. Marker-assisted
Hazel, L. 1943. The genetic basis for constructing selection indexes. selection in animal breeding schemes. p. 160. In Proc. Int.
Genetics 28:476–490. Soc. Anim. Genet., Tours, France. 21–25 July 1996.
Hazel, L., and J.L. Lush. 1942. The efficiency of three methods of Meuwissen, T.H.E., and M.E. Goddard. 2004. Mapping multiple
selection. J. Hered. 33:393–399. QTL using linkage disequilibrium and linkage analysis infor-
Heckenberger, M., H.P. Maurer, A.E. Melchinger, and M. Frisch. mation and multitrait data. Genet. Sel. Evol. 36:261–279.
2008. The plabsoft database: A comprehensive database man- Meuwissen, T.H.E., B.J. Hayes, and M.E. Goddard. 2001. Predic-
agement system for integrating phenotypic and genomic tion of total genetic value using genome-wide dense marker
data in academic and commercial plant breeding programs. maps. Genetics 157:1819–1829.
Euphytica 161:173–179. Moreau, L., A. Charcosset, and A. Gallais. 2004. Experimental
Henderson, C. 1984. Application of linear models in animal breed- evaluation of several cycles of marker-assisted selection in
ing. Univ. of Guelph, Ontario. maize. Euphytica 137:111–118.
Hill, W.G., M.E. Goddard, P.M. Visscher, and T.F.C. Mackay. Moreau, L., A. Charcosset, F. Hospital, and A. Gallais. 1998.
2008. Data and theory point to mainly additive genetic vari- Marker-assisted selection efficiency in populations of fi nite
ance for complex traits. PLoS Genet. 4:e1000008. size. Genetics 148:1353–1365.
Holland, J.B. 2001. Epistasis and plant breeding. Plant Breed. Rev. Muir, W.M. 2007. Comparison of genomic and traditional BLUP-
21:27–92. estimated breeding value accuracy and selection response
Holland, J.B. 2004. Implementation of molecular markers for under alternative trait and genomic parameters. J. Anim.
quantitative traits in breeding programs: Challenges and Breed. Genet. 124:342–355.
opportunities. p. 26. In T. Fischer et al. (ed.) New Directions Neimann-Sorensen, A., and A. Robertson. 1961. The association
for a Diverse Planet: Proc. for the 4th Int. Crop Science Con- between blood groups and several production characteristics
gress, Brisbane, Australia. 26 Sept.–1 Oct. 2004. Regional in three Danish cattle breeds. Acta Agric. Scand. 11:163–196.
Institute, Gosford, Australia. Piepho, H.P., J. Möhring, A.E. Melchinger, and A. Büchse. 2007.
Holland, J.B. 2007. Genetic architecture of complex traits in BLUP for phenotypic selection in plant breeding and variety
plants. Curr. Opin. Plant Biol. 10:156–161. testing. Euphytica 161:209–228.
Hospital, F., L. Moreau, F. Lacoudre, A. Charcosset, and A. Gal- Piyasatian, N., R.L. Fernando, and J.C. Dekkers. 2007. Genomic
lais. 1997. More on the efficiency of marker-assisted selection. selection for marker-assisted improvement in line crosses.
Theor. Appl. Genet. 95:1181–1189. Theor. Appl. Genet. 115:665–674.

CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009 WWW.CROPS.ORG 11


Quinton, M., C. Smith, and M. Goddard. 1992. Comparison of Villanueva, B., R. Pong-Wong, J. Fernandez, and M. Toro. 2005.
selection methods at the same level of inbreeding. J. Anim. Benefits from marker-assisted selection under an additive
Sci. 70:1060–1067. polygenic genetic model. J. Anim. Sci. 83:1747–1752.
Rafalski, A. 2002. Applications of single nucleotide polymor- Whittaker, J.C., R. Thompson, and M.C. Denham. 2000.
phisms in crop genetics. Curr. Opin. Plant Biol. 5:94–100. Marker-assisted selection using ridge regression. Genet. Res.
Remington, D.L., J.M. Thornsberry, Y. Matsuoka, L.M. Wilson, 75:249–252.
S.R. Whitt, J. Doebley, S. Kresovich, M.M. Goodman, and Wong, C.K., and R. Bernardo. 2008. Genomewide selection in
E.S. Buckler, IV. 2001. Structure of linkage disequilibrium oil palm: Increasing selection gain per unit time and cost with
and phenotypic associations in the maize genome. Proc. Natl. small populations. Theor. Appl. Genet. 116:815–824.
Acad. Sci. USA 98:11479–11484. Woolliams, J., P. Bijma, and B. Villanueva. 1999. Expected genetic
Schaeffer, L.R. 2006. Strategy for applying genome-wide selec- contributions and their impact on gene flow and genetic gain.
tion in dairy cattle. J. Anim. Breed. Genet. 123:218–223. Genetics 153:1009–1020.
Schön, C., S. Utz, B. Groh, S. Truberg, S. Openshaw, and A. Wray, N.R., and R. Thompson. 1990. Prediction of rates of
Melchinger. 2004. QTL mapping based on resampling in a inbreeding in selected populations. Genet. Res. 55:41–54.
vast maize testcross experiment confi rms the infi nitesimal Xu, S. 2003a. Theoretical basis of the Beavis effect. Genetics
model of quantitative genetics for complex traits. Genetics 165:2259–2268.
167:485–498. Xu, S. 2003b. Estimating polygenic effects using markers of the
Smith, C. 1967. Improvement of metric traits through specific entire genome. Genetics 163:789–801.
genetic loci. Anim. Prod. 9:349–358. Xu, Y., and J.H. Crouch. 2008. Marker-assisted selection in plant
Stranger, B.E., M.S. Forrest, M. Dunning, C.E. Ingle, C. Beazley, breeding: From publications to practice. Crop Sci. 48:391–407.
N. Thorne, R. Redon, C.P. Bird, A. de Grassi, and C. Lee. Yan, W., and N.A. Tinker. 2007. DUDE: A user-friendly crop
2007. Relative impact of nucleotide and copy number varia- information system. Agron. J. 99:1029–1033.
tion on gene expression phenotypes. Science 315:848–853. Yu, J., G. Pressoir, W.H. Briggs, I.V. Bi, M. Yamasaki, J.F. Doeb-
Stuber, C., M. Goodman, and R. Moll. 1982. Improvement of ley, M.D. McMullen, B.S. Gaut, D.M. Nielsen, and J.B. Hol-
yield and ear number resulting from selection at allozyme loci land. 2006. A unified mixed-model method for association
in a maize population. Crop Sci. 22:737–740. mapping that accounts for multiple levels of relatedness. Nat.
Tanksley, S., N. Young, A. Paterson, and M. Bonierbale. 1989. Genet. 38:203–208.
RFLP mapping in plant breeding: New tools for an old sci- Zhang, W., H.R. Lee, D.H. Koo, and J. Jiang. 2008. Epigenetic
ence. Biotechnology (N. Y.) 7:257–264. modification of centromeric chromatin: Hypomethylation of
Tenaillon, M.I., M.C. Sawkins, A.D. Long, R.L. Gaut, J.F. Doeb- DNA sequences in the CENH3-associated chromatin in ara-
ley, and B.S. Gaut. 2001. Patterns of DNA sequence polymor- bidopsis thaliana and maize. Plant Cell 20:25–34.
phism along chromosome 1 of maize (Zea mays ssp. mays L.). Zhao, K., M.J. Aranzana, S. Kim, C. Lister, C. Shindo, C. Tang, C.
Proc. Natl. Acad. Sci. USA 98:9161–9166. Toomajian, H. Zheng, C. Dean, and P. Marjoram. 2007. An
ter Braak, C.J.F., M.P. Boer, and M.C.A.M. Bink. 2005. Extend- Arabidopsis example of association mapping in structured sam-
ing Xu’s Bayesian model for estimating polygenic effects using ples. PLoS Genet. 3:e4, doi:10.1371/journal.pgen.0030004.
markers of the entire genome. Genetics 170:1435–1438. Zhong, S., and J.L. Jannink. 2007. Using quantitative trait loci
Tinker, N., and W. Yan. 2006. Information systems for crop per- results to discriminate among crosses on the basis of their
formance data. Can. J. Plant Sci. 86:647–662. progeny mean and variance. Genetics 177:567–576.
van Arendonk, J., B. Tier, and B. Kinghorn. 1994. Use of multiple Zhu, C., M. Gore, E.S. Buckler, and J. Yu. 2008. Status and pros-
genetic markers in prediction of breeding values. Genetics pects of association mapping in plants. Plant Genome 1:5–20.
137:319–329.

12 WWW.CROPS.ORG CROP SCIENCE, VOL. 49, JANUARY– FEBRUARY 2009

You might also like