Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Livestock Production Science 84 (2003) 63–73 www.elsevier.com / locate / livprodsci Model comparison for genetic evaluation of milk yield in Uruguayan Holsteins J.I. Urioste a , *, R. Rekaya b ,1 , D. Gianola b , W.F. Fikse c , K.A. Weigel b a ´ , Universidad de la Republica ´ ´ 780, 12900 Montevideo, Uruguay , Av. Garzon Facultad de Agronomıa b Department of Dairy Science, University of Wisconsin, Madison, WI 53706, USA c INTERBULL Centre, SLU, Box 7023, S750 07 Uppsala, Sweden Received 23 May 2002; received in revised form 31 December 2002; accepted 7 March 2003 Abstract Three models for genetic evaluation of milk yield of Uruguayan Holstein cattle were compared using 159 169 lactation records from 81 928 cows calving between 1989 and 1998. Model I included the effects of herd-year-season, parity by age group, additive genetic merit, permanent environment, and residual. Model II included all effects in Model I, as well as number of days open and length of the dry period. Model III included all factors in Model II, and it accommodated heterogeneity of variance within contemporary groups (CG) through a pre-adjustment of the data based on empirical Bayes estimates of the CG variance. Estimates of heritability for milk yield were 0.23, 0.24 and 0.25, and estimates of repeatability were 0.55, 0.56 and 0.57 for Models I, II and III, respectively. Models were contrasted by examining changes in sire ranking and by a cross-validation procedure, based on the ability of the models to predict first, second and later lactations. Data were divided into two subsets, and records from one subset were predicted using location parameters estimated from the other subset. A resampling procedure was used to minimise the dependency on the sample structure. Correspondence between observed and predicted values was assessed in terms of square root of empirical mean square errors of prediction, percentage squared bias and the coefficient of determination ‘R 2 ’. Adjustment for heterogeneous CG variance had a marked effect on rankings of animals, especially elite cows, where correlations between solutions from Models I and II versus Model III ranged from 0.53 to 0.80. The percentage of animals selected in common by each pair of models decreased when selection intensity increased. Cross-validation analyses suggested that an assumption of heterogeneity of CG variance is tenable, especially in later lactations, whereas some doubts arise in first lactations, most probably due to the data structure used in the analyses.  2003 Elsevier B.V. All rights reserved. Keywords: Dairy cattle; Milk yield; Heterogeneous variance; Genetic evaluation; Predictive ability *Corresponding author. Tel.: 1598-2-355-9636; fax: 1598-2359-3004. E-mail address: jorgeu@internet.com.uy (J.I. Urioste). 1 Present address: Department of Animal and Dairy Science, University of Georgia, Athens, GA 30603-2771, USA. 1. Introduction Best linear unbiased prediction (Henderson, 1973, 1984) has become the standard method for inferring breeding values, especially in dairy cattle. A large 0301-6226 / 03 / $ – see front matter  2003 Elsevier B.V. All rights reserved. doi:10.1016 / S0301-6226(03)00051-4 64 J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 number of countries employ an animal model (INTERBULL, 2000), where an additive genetic effect is fitted for each animal in the pedigree. Systematic environmental effects, such as age and parity of cow, duration of the dry period, number of days open and length of the previous and current calving intervals are often included as explanatory variables in the models for genetic evaluation of milk yield (INTERBULL, 2000). The first applications of mixed linear models for genetic evaluation of dairy cattle assumed constant genetic and residual components of variance across environments. However, considerable evidence has accumulated that there exists heterogeneity of variance for milk yield (Hill et al., 1983; Brother´˜ stone and Hill, 1986; Ibanez et al., 1996, 1999; Dodenhoff and Swalve, 1998). Ignoring such heterogeneity may lead to imprecise or biased predictions of genetic merit. Most countries now account for heterogeneity of variance in their national evaluation programs (e.g., Brotherstone and Hill, 1986; Jones ´˜ and Goddard, 1990; Meuwissen et al., 1996; Ibanez et al., 1996; Wiggans and Van Raden, 1991; RobertGranie´ et al., 1999). In several countries, a simple pre-adjustment of records is made, prior to fitting the animal model, in an attempt to reduce within-herd heterogeneous phenotypic variance. Across-herd genetic evaluation of dairy cattle in Uruguay started in 1992, and the system evolved rapidly towards a BLUP animal model for repeated lactations, assuming homogeneous variance. For computational reasons, and due to a lack of accurate estimates of variance parameters, the model has been kept as simple as possible. It is of interest to progress toward more biologically plausible models, such as those that consider heterogeneous variance and that include effects of reproduction on milk yield. Reproductive and management variables such as calving interval, length of dry period and number days open have been shown to influence milk yield (Schaeffer and Henderson, 1972; Funk et al., 1987; Sadek and Freeman, 1992; Lee et al., 1997; Berger and Lista, 1999). Studies on model performance based on goodness of fit or predictive ability are scarce in the literature. ´ Perez-Enciso et al. (1993) compared linear and Poisson mixed models for litter size in pigs. These authors split the data into two subsets, and predicted records in one of the subsets using location parame- ters estimated from the complementary subset. Estany and Sorensen (1995) in pigs, Olesen et al. ´˜ (1994) in sheep, and Ibanez et al. (1999) and Tempelman and Gianola (1999) in dairy cattle used similar procedures. Once a model with reasonable goodness of fit and predictive ability is found, it is desirable to obtain precise estimates of genetic parameters for this model. The objective of this study was to estimate genetic parameters for and compare three alternative models for genetic evaluation of milk yield in the Uruguayan Holstein population, based on their predictive ability of first, second and later lactations. The first model was the repeatability animal model that is currently used in practice. The second model included the effect of reproductive and management variables on milk yield, and the third model is as the second model but with an adjustment for heterogeneous phenotypic variance. 2. Material and methods 2.1. Data Milk yield records (305-day) from Holstein cows calving between 1989 and 1998 that had been included in the 1999 Uruguayan national genetic evaluation were used in the analysis. Incomplete records are routinely extended to 305-day lactation records by the last test-day method, with extension factors calculated by age, parity and month of calving. Production data and pedigree were provided ´ Rural del Uruguay (ARU) and by by the Asociacion Instituto Nacional para el Mejoramiento Lechero (INML). A 305-day lactation record was included in the analysis if it had at least four test-day yields and if the cow had at least one test day beyond 150 days in milk. Records for parities 3 through 5 were included, provided the cow’s records for the preceding lactations qualified for inclusion in the data set. Winter–Spring (June–November) and Summer–Autumn (December–May) calving seasons were defined. Herd-year-season contemporary groups (HYS) were formed by combining herd, year and season of calving classes, and only HYS-classes with more than five records were retained. Extended records received were given the same weight as completed J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 records. After edits, 159 169 lactations from 81 928 cows in 3600 HYS were available for analysis. Additional information is in Urioste et al. (2001). The pedigree file included 99 192 animals, of which 82.6% had milk records and 64.3% had at least one known parent. 2.2. Models Three alternative models were compared. Model I is the model currently used in the Uruguayan genetic evaluation, and Model II included additional information on explanatory variables related to reproduction. Model III contemplated an adjustment for heterogeneous variance in addition to the reproductive variables. The specification for Model I was: y ijkl 5 HYS i 1 L j 1 a k 1 pe k 1 e ijkl (1) where y ijkl 5305-day milk yield of cow k in lactation l, lactation-age class j and HYS i; HYS i fixed effect of herd-year-season i (i51,2, . . . 3600); L j 5fixed effect of the jth combination of lactation and age class ( j51,2, . . . ,26); a k 5random additive genetic effect of animal k (k51,2, . . . ,99 192); pe k 5random permanent environmental effect of cow k (k5 1,2, . . . ,81 928); e ijkl 5random residual term. The vector of additive genetic effects was assumed to have the distribution a | N(0,As 2a ), where 0 is a vector of population means, A is a known matrix of additive genetic relationships between animals and s 2a is the additive genetic variance. The permanent environmental effects were taken to be independent and identically distributed, with mean 0 and variance s 2pe . The random residuals were independently distributed, each with mean 0 and variance s 2e . Additive genetic, permanent and residual effects were mutually independent. A description of the age-parity classes is given in Table 1. Dry-off and calving dates are reported routinely in the Uruguayan recording schemes. To identify information regarding the impact of reproductive measures, preliminary analyses were performed using a set of fixed models. These included all fixed terms in Model I plus the effects of length of previous calving interval (CI), number of days open (DO) and length of previous dry period (DP) or previous days in milk (DIM). There was a part-whole relationship between 65 Table 1 Parity-age classes and age ranges within parity Parity No. of levels No. of lactations Range of ages (months) 1 2 3 4 5 5 7 6 6 2 57 111 46 760 29 859 16 694 8745 22–53 30–71 48–93 54–180 78–180 CI, DP and DIM, and these variables were consequently measuring the same biological trait. It was decided to keep DP only. Further, because DO in the current lactation is still not recorded routinely in Uruguay, it was approximated as the difference between current CI (computed from current and next calving date) and a fixed gestation length (280 days), to account for effects of pregnancy on current lactation yield. Final classes for DO and DP, and the number of observations per class are in Table 2. Initially, a class consisted of a 5-day interval, and adjacent classes were combined when differences between levels were not significant. Missing observations for DO or DP were grouped into a specific class. A model including both DO and DP accounted for little additional variation (R 2 of 0.51) as compared with a model including DP only (R 2 of 0.50). However, DO constitutes a different source of biological variability, so it was decided to include both variables. Model III had the same specification as Model II, but heterogeneity of variance within contemporary groups (CG) was accommodated through a pre-adjustment of records, using an empirical Bayes estimator of the CG variance (Urioste et al., 2001). Briefly, herd-year-season levels were subdivided into herd-year-season-parity-lactation length CG classes, and the usual estimate of variance within a CG ( sˆ 2i , where i 5 1,2, . . . ,8955) was obtained. A fixed linear model including effects of herd, calving period, season, production level, contemporary group size, milk recording system, parity number and length of lactation was fitted to the logarithm of these estimated variances. Predicted values of vari0 ances using this model are represented as e k 9 ig ; here, k9 i is a suitable incidence vector and g 0 is a solution for the parameters of the model. Combining J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 66 Table 2 Classes for length of dry period and number of days open Class Dry period (days) No. of lactations Days open (days) No. of lactations 1 2 3 4 5 6 7 8 9 10 11 12 13 0–20 21–30 31–40 41–45 46–55 56–70 71–80 81–90 91–110 111–130 131–150 151–360 Missing data 2050 1914 3662 2696 7659 15 433 19 429 17 461 10 750 16 373 13 853 17 757 80 132 0–40 41–65 66–120 121–150 151–360 361–600 Missing data 2633 9703 31 629 10 383 24 253 23 496 77 072 0 ŝ 2i with e k 9 ig , an empirical Bayes estimator of the posterior mean of the variance within the ith CG is: ni 0 0 s˜ i2 (g 0 ,n ) 5 e k9 ig 1 ]ssˆ i2 2 e k 9 ig d n *i where n 58955-rank (model for log-variances) is a degree of belief parameter, ni 5 n i 2 1 (n i is the number of records for CG i ), and n * i 5 n 1 ni . Records were then adjusted as: y ij 2 mˆ i y ijC 5 mˆ i 1 ]]] s s̃(g 0 ,n ) base where sbase 5744.3 kg, based on preliminary analyses of the data. In each of the three models, additive genetic, permanent environmental and residual variances, heritability and repeatability were estimated via restricted maximum likelihood using the VCE software (Neumaier and Groeneveldt, 1998). Conditionally on these estimates, single trait BLUP breeding values and estimates of fixed effects were computed using the program JAA20 developed by Dr. Ignacy Misztal. 2.3. Comparison between models Models were contrasted by examining changes in the predicted breeding values and by using a crossvalidation procedure, using re-sampling. Rank correlations and regressions between predictions of breeding values were computed for bulls and cows at different hypothetical percentages of animals selected. The cross-validation procedure focused on the ability of the models to predict first, second and later lactation phenotypic records. The objective of the re-sampling was to reduce the dependency of several end-points examined on the specific structure of the sample to be predicted. To illustrate, consider prediction of first lactation records. These were divided into two randomly created sets, each containing approximately 50% of the records, and such that all fixed effects levels were present in both sets. One of the two sets (which we will refer to as ‘training set’) was chosen at random and merged with the data on second and later lactations to obtain BLUP of breeding values and BLUE of fixed effects. The breeding values and fixed effects were calculated only once, and these were used to predict the appropriate lactation records in other set, the ‘prediction set’, which initially included all first lactation records. This process was repeated 1000 times, such that the ‘prediction set’ varied at random (a new sample of records was taken each time) and contained, on average, 50% of records coming from the ‘training set’. The described procedure allowed computing two measures of model quality for each ‘replication’. One of the measures was goodness of fit, that is, the ability of the model to reproduce an observation that belongs to the ‘training set’ but that appears, with 50% chance, in the ‘predictive set’. The second measure centered on predictive performance, or ability of a model to predict an observation in the J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 ‘predictive set’ that was not included in the ‘training set’. Within each ‘replication’, for each lactation and model, agreement between observed and fitted or predicted values was assessed using the square root of the empirical mean squared error statistic: ]]]]] n 1 s SME 5 ] ( y i 2 yˆ i )2 n S i [S œ O where yˆ i is the fitted or predicted value for y i , y i is the observed record, and S is the appropriate set of observations. An additional end-point calculated was the ‘percentage squared bias’ (PSB), proposed by Ali and Schaeffer (1987), computed as: PSB 5 100( y 2 yˆ )9( y 2 yˆ ) /y9y where y and yˆ are observed and predicted values, respectively. Finally, a simple measure of agreement between predicted values and observations that had been omitted in the predictive set, the determination coefficient ‘R 2 ’ (by analogy with the usual regression statistic), was calculated as: ‘R 2 ’ 5 O( y 2 yˆ ) /(O y 2 ny¯ ), 2 2 2 computed over members of the predictive set. For each end-point, the mean and the standard deviations were calculated over the 1000 replicates. 3. Results and discussion Average milk yield for the Uruguayan Holstein in the sample was 4888 kg for a 305-day lactation, with a standard deviation of 1144 kg, a minimum of 1544 kg and a maximum yield of 11 650 kg. Table 3 presents estimates of variance components, Table 3 REML estimates of genetic and phenotypic parameters, by model (s 2a 5 additive genetic variance; s 2pe 5 permanent environment variance; s 2e 5 residual variance; s 2P 5 phenotypic variance) 2 a 2 pe 2 e 2 P 2 s (kg ) s (kg 2 ) s (kg 2 ) s (kg 2 ) Heritability Repeatability Model I Model II Model III 135 543 182 576 265 355 583 474 0.23 0.55 136 505 181 754 253 633 571 892 0.24 0.56 155 334 192 135 264 242 611 711 0.25 0.57 67 heritability and repeatability for milk yield. Models II and III gave slightly higher estimates of heritability and repeatability than Model I. Heritabilities were in the lower range of estimates from literature, probably influenced by the weak structure of the pedigree information: about 75% of the cows were grade and reliable pedigree information is unavailable for a large percentage of animals. However, the results are consistent with values obtained in populations with similar average levels of production. ˜ et al. (1989) found a heritability of 0.16 Carabano and an additive genetic variance of 108 608 kg 2 for Spanish Holsteins producing 4982 kg milk, on average. Stanton et al. (1991) estimated the variances between sires and heritabilities of milk yield from the United States and three Latin American countries. They found evidence of a lower additive genetic variance in the latter. Heritability in USA was 0.25, whereas for the Latin American countries it was 0.21. When milk records were adjusted to a common standard deviation of 1000 kg, these authors found that heritability was 0.26 for the USA population and 0.23 for the combined Latin American data. Hill et al. (1983) reported a heritability of 0.24 for British Friesian herds with low production level. Estimates of genetic variance found by these authors, however, were higher than those obtained here. All estimates of components of variance were larger in Model III. For example, the estimate of additive genetic variance was about 15% larger in Model III than in Models I or II. Increased heritability for Model III suggests that a failure to accounting for heteroscedasticity between contemporary groups may mask some of the existing genetic variation. Conditional on the variance components estimated (Table 3), solutions pertaining to the effects of age (within lactation), number of days open, and length of dry period on milk yield are presented in Fig. 1. Lactation-age effects indicated that cows reached mature production at the third lactation. Within lactation, production increased with age at calving, particularly in young cows. DO and DP both influenced yields, but their effects were smaller than those reported in other studies (Schaeffer and Henderson, 1972; Funk et al., 1987; Sadek and Freeman, 1992; Lee et al., 1997). Milk production increased with DO. Cows with DP less than 40 days produced markedly less milk. Yield increased with DP up to 68 J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 Fig. 1. Systematic environmental effects affecting milk yield; (A) parity and age of calving; (B) number of days open (Model II: filled bars; Model III: dotted bars); (C) length of previous dry period (Model II: filled bars; Model III: dotted bars). J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 70–100 days and decreased thereafter. Schaeffer and Henderson (1972) found an ‘optimum’ dry period of 50–59 days, whereas in the study of Funk et al. (1987) maximum production was found at 60–69 days. In practice, differences are small within a range of 615 days. The estimates of the effects of DO and DP on yield were similar for the two models (II and III) where the reproductive information was included in the explanatory structure. Rank correlations between models for predicted breeding values obtained with the three models are in Table 4. For the entire data set, correlations were larger than 0.98 and were similar for bulls and cows. For high selection intensity (0.1, 1 and 5% animals hypothetically selected), correlations between BLUPs from Models I and II were high. However, rank correlations between evaluations from Models I and II with those from Model III were much lower, ranging from 0.63 to 0.88 for sires and 0.53 to 0.80 for cows. Clearly, the adjustment for heterogeneous variance had a marked effect on rankings. A similar picture was observed by Robert-Granie´ et al. (1999): the overall correlation between evaluations from a homogeneous and heterogeneous variance model was high (0.98–0.99), but for elite cows the correlation was 0.54–0.78, reflecting an important re-ranking in this fraction of the population. This is possibly due to the fact that a cow’s own records and those of her dam and maternal relatives are usually made within a single herd, a situation where the effects of heterogeneity of variance on genetic evaluation can be especially important. Table 5 summarizes changes in absolute value of the BLUP EBV for milk yield. Again, changes were 69 larger when the comparison involved Model III, and especially when we analysed top fractions of animals selected. In particular cases, absolute changes in the BLUPs were around 450 kg, which is more than one genetic standard deviation in any of the models. Relationships between results from all models were also assessed by means of regressions on animals’ rankings. Departures from 0 for the intercept and from 1 for the slope of the regression line indicate differing ranking positions for animals selected by two different models. In this sense, regression lines are more informative than simple rank correlation measures, which cannot be significantly tested for values different from 1. Table 6 presents regressions on rankings of sires and cows between the three models, both for the overall data and for the top 5% of the animals. Overall regressions showed a clear departure of expected values. This was more marked in cows than in sires. Regressions involving top selected animals showed the same trend. Fig. 2 relates the percentage of animals (both sires and cows) selected in common by two different models, at different percentages of selected animals. The proportion of animals in common decreases when selection intensity increases. Percentage animals in common was lower for cows than for sires. For example, approximately 10% of the sires were not present in the top 5% of animals, when selection was based on models I or II instead of model III. For cows, the percentage of animals not present increased to 15–16%. Most visible changes should occur with the top animals (sires and elite cows), whereas selection at herd level, where 70–75% of cows are retained, would not change very much. In Table 4 Rank correlations between predicted breeding values from different models, by type of animal and percentage selected No. of animals Model I vs. Model II Model I vs. Model III Model II vs. Model III 2321 96 871 0.99 0.99 0.98 0.98 0.99 0.99 Sires selected with Model III 1% 23 5% 116 0.95 0.97 0.63 0.85 0.72 0.88 Cows selected with Model III 0.1% 97 1% 969 5% 4844 0.98 0.97 0.98 0.53 0.69 0.78 0.57 0.69 0.80 Overall Sires Cows J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 70 Table 5 Absolute values of the changes in predicted breeding value, by type of animal and percentage selected (M-I: Model I; M-II: Model II; M-III: Model III) Average (kg) Minimum (kg) Maximum (kg) M-I vs. M-II M-I vs. M-III M-II vs. M-III 16.1 34.6 28.4 0 0 0 165.5 449.5 467.6 Top 1% M-I vs. M-II M-I vs. M-III M-II vs. M-III Sires 16.4 61.2 56.2 Cows 20.2 71.9 65.5 Sires 0.4 0.9 3.4 Cows 0 0.1 0.3 Sires 46.3 187.8 208.9 Cows 115.1 449.5 467.1 Top 5% M-I vs. M-II M-I vs. M-III M-II vs. M-III 21.7 57.2 45.6 18.9 60.5 53.3 0.1 0.2 0 0 0 0.1 73.2 288.4 234.5 149.8 449.5 467.6 Overall Table 6 Regressions on rankings of animals, by type of animal and percentage selected (M-I: Model I; M-II: Model II; M-III: Model III) Intercept Sires Slope Cows Sires Overall M-I vs. M-II M-I vs. M-III M-II vs. M-III M-II vs. M-I M-III vs. M-I M-III vs. M-II 7.563.1 19.365.0 12.364.0 7.563.1 19.365.0 12.364.0 289.63619.62 762.24631.75 478.63625.20 289.63619.62 762.24631.75 478.63625.20 5% M-I vs. M-II M-I vs. M-III M-II vs. M-III M-II vs. M-I M-III vs. M-I M-III vs. M-II 21.361.9 5.565.3 3.564.3 5.261.7 21.463.5 15.963.2 4.83614.51 262.81654.58 244.90649.93 185.95613.36 1351.87624.33 1275.32624.17 the study of Robert-Granie´ et al. (1999), change from homogeneous to heterogeneous models resulted in 30–40% new cows in the top Holstein cow population. Sire ranking may also be affected when daughters are non-randomly distributed among high and low variance environments (Hill, 1984). RobertGranie´ et al. (1999) reported the case of foreign bulls as being used in herds with high production level and more variable environments, and therefore being 0.9960.00 0.9860.00 0.9960.00 0.9960.00 0.9860.00 0.9960.00 21.361.9 5.565.3 3.564.3 5.261.7 21.463.5 15.963.2 Cows 0.9960.00 0.9860.00 0.9960.00 0.9960.00 0.9860.00 0.9960.00 1.0260.00 1.1160.02 1.0860.02 0.9060.00 0.3660.01 0.3960.01 more affected when a correction for heterogeneity is performed. In Table 7, the predictive ability of the models is illustrated. In general terms, MSE was reduced with the complexity of the model, and increased with parity number. The coefficient of determination R 2 increased with model complexity and decreased with lactation number. The notorious exception was Lactation 1, for which there were no big differences J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 71 Fig. 2. Percentage of sires (s) and cows (d) selected in common between Models I and III (Model I in the legend), and Models II and III (Model II in the legend), at different percentages of selected animals. Table 7 Square root of empirical Mean Square Error (SME), Percentage Squared Bias (PSB) and coefficient of determination ‘R 2 ’ for prediction of different lactations with Models I, II and III Lactations Model I Model II Model III 1 SME (lack of fit) (kg) SME (pred. error) (kg) Average SME (kg) PSB (%) R 2 (%) 396.362.1 581.463.3 489.162.1 1.1860.01 82.360.2 389.762.0 578.263.2 484.462.0 1.1560.01 82.660.2 422.162.0 657.863.3 541.962.2 1.4460.01 80.660.2 2 SME (lack of fit) (kg) SME (pred. error) (kg) Average SME (kg) PSB (%) R 2 (%) 434.862.5 635.163.3 555.862.4 1.2060.01 77.360.2 421.062.4 620.663.2 542.062.3 1.1360.01 78.460.2 408.462.2 606.362.8 528.362.0 1.0860.01 79.460.2 3 and later SME (lack of fit) (kg) SME (pred. error) (kg) Average SME (kg) PSB (%) R 2 (%) 470.762.6 677.262.9 599.662.2 1.2360.01 73.460.2 459.062.5 664.462.8 587.362.2 1.1860.01 74.560.2 443.462.1 657.062.7 577.462.0 1.1460.01 74.560.2 between Models I and II (recall that the effect of DP is non-existent in lactation 1), and where Model III seems to be the worst model. This result was ´ ˜ et al. (1999) used unexpected. The study of Ibanez first lactations only and compared three models: homogeneous genetic and residual variance, homogeneous genetic and residual variance after data standardisation (like ours) and a heterogeneous genetic and residual variance. They found that a phenotypic preadjustment for heterogeneity corrected for the phenotypic dispersion and was better than the purely homogeneous model, but dispersion of predicted genetic values of animals in environments with large heritability was underestimated with respect to the heterogeneous variance model. For our data, the larger SME values for Model III could be interpreted as a consequence of the data structure: new herds entering the evaluation typically reported a few first lactation records, producing poor estimates of the Bayesian estimator and consequently (under) overestimating the adjustment for heterogeneity of variances. In later parities, however, production stress causes important variation in milk yield, and a correction for heterogeneity of variance seems to be advisable. Prediction bias, measured as PSB, was very low in all models and lactations. 72 J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 ´ ˜ et al. (1999) also found small values of PSB Ibanez for milk yield, 3.02% for a homogeneous model and slightly smaller values, 2.84 or 2.87%, for two heterogeneous models. Some primary effects of accounting for heterogeneous variance in national genetic evaluations have been discussed by Weigel and Lawlor (1994), and apply to the Uruguayan situation: (1) increased fairness of genetic evaluations; (2) increased overall accuracy of evaluation for superior cows, thereby increasing potential for genetic progress through maternal selection pathways; (3) improved accuracy of sire evaluations for (a) sires imported from foreign countries and used non-randomly with respect to herd variance and (b) breeder proven bulls with many progeny in a single herd with high or low variance 4. Conclusions The general results obtained in this study show that the current Uruguayan genetic evaluation model could be improved in various aspects. Accurate estimations of parameters are now available (Table 3), and the effects of DO and DP proved to be significant (see Fig. 1). The problem of heterogeneous variances in genetic evaluations of dairy cattle is that above average animals in the more variable herds may be over-evaluated. A greater proportion of animals would then be selected from the more variable herds. Cow evaluation was found to be much more sensitive to violations of the assumed homogeneous variance. Sire ranking may also be affected when daughters are non-randomly distributed among high and low variance environments (Hill, 1984). This could be the case of Uruguay, since there is no organised progeny testing of young bulls, and a majority of sires have daughters in just one herd. Also, foreign bulls (more expensive semen) tend to be used in herds with a high production level. Accuracy of evaluations depends on how well the assumptions of the model match the data. Our results, presented in Table 7, suggest that the assumption of heterogeneous variance is tenable, especially in later lactations, whereas some doubts arise in first lactations, most probably due to the structure of the database and pedigree information used in the analyses. Given the relative simplicity of the results found here, and with the necessary care in data selection to avoid poor estimates of the Bayesian estimator proposed by Urioste et al. (2001) to deal with heterogeneity of variances, incorporation of an adjustment for heterogeneity of variances to the Uruguayan Holstein genetic evaluation is strongly recommended. Acknowledgements ARU and INML are thanked for supplying the data, and the Department of Dairy Science, University of Wisconsin-Madison, for hosting J.I. Urioste’s visit. Dr. Ignacy Misztal is acknowledged for his ´ JAA programs, and Facultad de Agronomıa, Uruguay, and World-Wide Sires, Inc., Visalia, CA, for economic support. References Ali, T.E., Schaeffer, L.R., 1987. Accounting for covariances among test day milk yield in dairy cows. Can. J. Anim. Sci. 67, 637–644. ´ entre comportamiento Berger, A., Lista, O., 1999. Relacion productivo y reproductivo de vacas Holando. Tesis Ingeniero ´ ´ Agronomo, Universidad de la Republica Montevideo, Uruguay, 100 pp. Brotherstone, S., Hill, W.G., 1986. Heterogeneity of variance amongst herds for milk production. Anim. Prod. 42, 297–303. ˜ M.J., Van Vleck, L.D., Wiggans, G.R., Alenda, R., 1989. Carabano, Estimation of genetic parameters for milk and fat yields of dairy cattle in Spain and the United States. J. Dairy Sci. 72, 3013–3022. Dodenhoff, J., Swalve, H.H., 1998. Heterogeneity of variances across regions of northern Germany and adjustment in genetic evaluation. Livest. Prod. Sci. 53, 225–236. Estany, J., Sorensen, D., 1995. Estimation of genetic parameters for litter size in Danish Landrace and Yorkshire pigs. Anim. Sci. 60, 315–324. Funk, D.A., Freeman, A.E., Berger, P.J., 1987. Effects of previous days open, previous days dry and present days open on lactation yield. J. Dairy Sci. 70, 2366–2373. Henderson, C.R., 1973. Sire evaluation and genetic trends. In: Proc. Animal Breeding and Genetics symposium in honor of Dr, J.L. Lush. American Society of Animal Science and American Dairy Science Association. Champaign, IL, USA, pp. 10–41. Henderson, C.R., 1984. In: Applications of Linear Models in Animal Breeding. University of Guelph, Guelph, Ontario, Canada. J.I. Urioste et al. / Livestock Production Science 84 (2003) 63–73 Hill, W.G., 1984. On selection among groups with heterogeneous variance. Anim. Prod. 39, 473–477. Hill, W.G., Edwards, M.R., Ahmed, M.-K.A., Thompson, R., 1983. Heritability of milk yield and composition at different levels and variability of production. Anim. Prod. 36, 59–68. ˜ M.J., Alenda, R., 1999. Identification of ´˜ Ibanez, M.A., Carabano, sources of heterogeneous residual and genetic variances in milk yield data from the Spanish Holstein-Friesian population and impact on genetic evaluation. Livest. Prod. Sci. 59, 33–49. ˜ M.J., Foulley, J.L., Alenda, R., 1996. ´˜ Ibanez, M.A., Carabano, Heterogeneity of herd-period phenotypic variances in the Spanish Holstein-Friesian cattle: sources of heterogeneity and genetic evaluation. Livest. Prod. Sci. 45, 137–147. INTERBULL 2000. National Genetic Evaluation Programmes for Dairy Production Traits Practiced in Interbull Member Countries 1999–2000. Bulletin no. 24, Department of Animal Breeding and Genetics, SLU, Uppsala, Sweden, 111 pp. Jones, L.P., Goddard, M.E., 1990. Five years experience with the animal model for dairy evaluations in Australia. In: Proc. 4th WCGALP, Edinburgh, Scotland, XIII, pp. 382–385. Lee, J.K., Van Raden, P.M., Norman, H.D., Wiggans, G.R., Meinert, T.R., 1997. Relationship of yield during early lactation and days open during current lactation with 305-day yield. J. Dairy Sci. 80, 771–776. Meuwissen, T.H.E., De Jong, G., Engel, B., 1996. Joint estimation of breeding values and heterogeneous variances of large data files. J. Dairy Sci. 79, 310–316. Neumaier, A., Groeneveldt, E., 1998. Restricted Maximum Likelihood estimation of covariances in sparse linear models. Genet. Sel. Evol. 30, 3–26. ´ Olesen, I., Perez-Enciso, M., Gianola, D., Thomas, D.L., 1994. A comparison of normal and nonnormal mixed models for 73 number of lambs born in Norwegian sheep. J. Anim. Sci. 72, 1166–1173. ´ Perez-Enciso, M., Tempelman, R.J., Gianola, D., 1993. A comparison between linear and Poisson mixed models for litter size in Iberian pigs. Livest. Prod. Sci. 35, 303–316. ´ C., Bonaıti, ¨ B., Boichard, D., Barbat, A., 1999. Robert-Granie, Accounting for variance heterogeneity in French dairy cattle genetic evaluation. Livest. Prod. Sci. 60, 343–357. Sadek, M.H., Freeman, A.E., 1992. Adjustment factors for previous and present days open considering all lactations. J. Dairy Sci. 75, 279–287. Schaeffer, L.R., Henderson, C.R., 1972. Effects of days dry and days open on Holstein milk production. J. Dairy Sci. 55, 107–112. Stanton, T.L., Blake, R.W., Quaas, R.L., Van Vleck, L.D., ˜ M.J., 1991. Genotype by environment interaction for Carabano, Holstein milk yield in Colombia, Mexico and Puerto Rico. J. Dairy Sci. 74, 1700–1714. Tempelman, R.J., Gianola, D., 1999. Genetic analysis of fertility in dairy cattle using negative binomial mixed models. J. Dairy Sci. 82, 1834–1847. Urioste, J.I., Gianola, D., Rekaya, R., Fikse, W.F., Weigel, K.A., 2001. Evaluation of extent and amount of heterogeneous variance for milk yield in Uruguayan Holsteins. Animal Sci. 72, 259–268. Weigel, K.A., Lawlor, T.J., 1994. Adjustment for heterogeneous variance in genetic evaluations for conformation of United States Holsteins. J. Dairy Sci. 77, 1691–1701. Wiggans, G.R., Van Raden, P.M., 1991. Method and effect of adjustment for heterogeneous variance. J. Dairy Sci. 74, 4350– 4357.