18
This article was downloaded by: [Northeastern University] On: 30 October 2014, At: 09:46 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Acta Agriculturae Scandinavica, Section A — Animal Science Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/saga20 A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle H. M. Nielsen a , A. F. Groen b , S. Østergaard c & P. Berg a a Department of Genetics and Biotechnology , Danish Institute of Agricultural Sciences, Research Centre Foulum , Tjele, Denmark b Animal Breeding and Genetics group , Wageningen Institute of Animal Sciences, Wageningen University , Wageningen, The Netherlands c Department of Animal Health , Welfare and Nutrition, Danish Institute of Agricultural Sciences, Research Centre Foulum , Tjele, Denmark Published online: 01 Feb 2007. To cite this article: H. M. Nielsen , A. F. Groen , S. Østergaard & P. Berg (2006) A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle, Acta Agriculturae Scandinavica, Section A — Animal Science, 56:1, 16-32, DOI: 10.1080/09064700600836786 To link to this article: http://dx.doi.org/10.1080/09064700600836786 PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content. This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http:// www.tandfonline.com/page/terms-and-conditions

A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

  • Upload
    p

  • View
    212

  • Download
    0

Embed Size (px)

Citation preview

Page 1: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

This article was downloaded by: [Northeastern University]On: 30 October 2014, At: 09:46Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registered office: MortimerHouse, 37-41 Mortimer Street, London W1T 3JH, UK

Acta Agriculturae Scandinavica, Section A — AnimalSciencePublication details, including instructions for authors and subscription information:http://www.tandfonline.com/loi/saga20

A stochastic model for the derivation of economicvalues and their standard deviations for productionand functional traits in dairy cattleH. M. Nielsen a , A. F. Groen b , S. Østergaard c & P. Berg aa Department of Genetics and Biotechnology , Danish Institute of Agricultural Sciences,Research Centre Foulum , Tjele, Denmarkb Animal Breeding and Genetics group , Wageningen Institute of Animal Sciences,Wageningen University , Wageningen, The Netherlandsc Department of Animal Health , Welfare and Nutrition, Danish Institute of AgriculturalSciences, Research Centre Foulum , Tjele, DenmarkPublished online: 01 Feb 2007.

To cite this article: H. M. Nielsen , A. F. Groen , S. Østergaard & P. Berg (2006) A stochastic model for the derivationof economic values and their standard deviations for production and functional traits in dairy cattle, Acta AgriculturaeScandinavica, Section A — Animal Science, 56:1, 16-32, DOI: 10.1080/09064700600836786

To link to this article: http://dx.doi.org/10.1080/09064700600836786

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) containedin the publications on our platform. However, Taylor & Francis, our agents, and our licensors make norepresentations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose ofthe Content. Any opinions and views expressed in this publication are the opinions and views of the authors,and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be reliedupon and should be independently verified with primary sources of information. Taylor and Francis shallnot be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and otherliabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to orarising out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematicreproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in anyform to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://www.tandfonline.com/page/terms-and-conditions

Page 2: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

ORIGINAL ARTICLE

A stochastic model for the derivation of economic values and theirstandard deviations for production and functional traits in dairy cattle

H. M. NIELSEN1, A. F. GROEN2, S. ØSTERGAARD3 & P. BERG1

1Department of Genetics and Biotechnology, Danish Institute of Agricultural Sciences, Research Centre Foulum, Tjele,

Denmark, 2Animal Breeding and Genetics group, Wageningen Institute of Animal Sciences, Wageningen University,

Wageningen, The Netherlands, and 3Department of Animal Health, Welfare and Nutrition, Danish Institute of Agricultural

Sciences, Research Centre Foulum, Tjele, Denmark

AbstractThe objective of this paper was to present a model of a dairy cattle production system for the derivation of economic valuesand their standard deviations for both production and functional traits under Danish production circumstances. Thestochastic model used is dynamic, and simulates production and health in a dairy herd. Because of indirect effects betweentraits, the phenotypic levels of (related) traits can change as a result of genetic changes. Economic values for milk productionand body weight were 0.28 and �/0.76 t/kg per cow-year respectively. For incidence of milk fever, mastitis, retainedplacenta and laminitis economic values were �/402.1, �/162.5, �/79.0 and �/210.2 t/incidence per cow-year. The eco-nomic values for involuntary culling rate, stillbirth and conception rate were �6.66, �/1.63, and 1.98 t/% per cow-year,respectively and the economic value for days from calving to first heat was �/0.94 t/day per cow-year. Standard deviations ofeconomic values expressing variation in realised profit of a farm before and after a genetic change were computed using alinear Taylor series expansion. Expressed as coefficient of variation, standard deviations of economic values based on 1000replicates ranged between 0.07 (milk production) to 16 (retained placenta).

Keywords: Milk production, health traits, breeding objective.

Introduction

Modern dairy cattle breeding has increased produc-

tion levels successfully. However, the upward trend

in milk production per cow has been associated with

undesirable side-effects, such as an increase in

production diseases and reproductive problems

(Rauw et al., 1998). In order to avoid this deteriora-

tion of functional traits, a balanced improvement of

production and functional traits is required. In

Denmark, selection for functional traits has been

practised for several years. A total merit index is

used, and includes milk yield and several functional

traits, such as calving ease, fertility, diseases, and

longevity (Principles of Danish Cattle Breeding,

2001). However, economic values applied in the

breeding goal are not well documented. In addition,

reports of economic values for disease traits are

scarce and breeding goal definition is still of ongoing

interest (Groen et al., 1997).

Derivation of economic values for production

traits in dairy cattle has been thoroughly discussed

in the literature (e.g., Groen, 1989a). Including

functional traits in breeding goals complicates the

derivation of economic values, as many of these traits

are difficult to define and model in an appropriate

way. Also, there is growing public concern about

welfare and ethical issues related to farm animals

(Olesen et al., 2000). Objective methods are being

considered as the most appropriate method to derive

economic values when the production system can be

accurately described. Objective methods are based

on modelling, where equations along with defini-

tions, descriptions and assumptions represent the

behaviour of a production system (Groen et al.,

1997). Using a bio-economic model to derive

Correspondence: H. M. Nielsen, Department of Animal and Aquacultural Sciences, Norwegian University of Life Sciences, Box 5003, NO-1432 As, Norway.

E-mail: [email protected]

Acta Agriculturae Scand Section A, 2006; 56: 16�32

ISSN 0906-4702 print/ISSN 1651-1972 online # 2006 Taylor & Francis

DOI: 10.1080/09064700600836786

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 3: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

economic values it is possible to examine the effect

of different prices, levels and sizes of the prod-

uction system (Groen et al., 1997). In addition, a

bio-economic model is an efficient tool to model

relationships between functional traits (e.g., repro-

duction and culling). Traditionally, deterministic

bio-economic models have been used to derive

economic values (e.g., Groen, 1989a).

Economic values are derived from the change in

profit of the farm brought about by a genetic change

in a given trait without changing the genetic level of

other traits (Groen et al., 1997). When economic

values are derived as the partial derivative of the

profit equation, genetic levels of other traits in the

breeding objective will not change. However, when

economic values are derived using a bio-economic

model as in this study, interactions between genetic

levels of traits must be zero in order to avoid changes

in genetic levels of other breeding goal traits.

According to Hazel (1943) the economic value of

a trait depends upon the amount by which profit

may be expected to increase for a unit of genetic

improvement in that trait. Profit is a function of

genetic and environmental factors. In earlier studies

concerning derivation of economic values, little

differentiation was made between genetic and phe-

notypic potential of individual animals. This as-

sumes that there is a one-to-one relationship

between genetic change and phenotypic expression,

which may not be the case (Vargas et al., 2002;

Quinton, 2005). Economic values should be derived

under optimal management (Goddard, 1983). En-

vironmental factors can be classified into permanent

(e.g., parity) or temporary effects (e.g., stage of

lactation and health status). Both permanent and

temporary effects interact with the genotype of the

animal and affect feed intake and feed apportioning.

Consequently, phenotypic expressions of traits are

influenced by environmental factors (Bryant et al.,

2005). For example, milk production capacity of a

cow can only be fully expressed at the phenotypic

level as long as the level of feed intake capacity is

high enough to secure sufficient energy to meet

potential milk yield (Vargas et al., 2002). Recently,

Quinton et al. (2006) found that the economic value

for survival depended on the phenotypic level of

litter size in pigs due to a non-linear interaction

between litter size and perinatal survival. Therefore,

the phenotypic expression of survival after a genetic

change was dependent on the phenotypic level of

litter size and a genetic change in survival did not

necessarily result in the same magnitude of change at

the phenotypic level. Therefore, it is important to

consider the level of related traits when deriving

economic values.

The objective of this paper is to present a

stochastic model for the derivation of economic

values and their standard deviations under Danish

production circumstances. This study builds on

a model (Sørensen et al., 1992; Østergaard et al.,

2000, 2003), which previously has been used

for decision support to study management effects

in a herd. However, this dynamic model simulates a

dairy cattle production system and health traits in

detail. In addition, by using a stochastic simulation

model, uncertainty of economic values expressing

variation in realized profit of a farm before and after

a genetic change due to variation in performance of

cows can be obtained.

Material and methods

The derivation of economic values was based mainly

on the existing SimHerd model (Sørensen et al.,

1992; Østergaard et al., 2000, 2003). A module

(SimProfit) was added to derive economic values

and to model breeding goal traits, which could not

be modelled in SimHerd. A description of the overall

model is given in this paper. However, results

presented are limited to traits modelled in SimHerd.

First, an overall description of the model and the

assumed farm characteristics are presented and the

applied profit equations are given subsequently.

Secondly, details on different traits in the profit

equations are given and the method used to derive

economic values is described.

Model

The normative approach was chosen because of the

possibilities of detailed modelling of the production

circumstances and the flexibility to investigate effects

of price changes, for example. The overall model to

derive economic values is given in Figure 1. The

SimHerd model is a dynamic (discrete, weekly time-

stepping) model, simulating production and state

changes of a dairy herd with additional young stock.

The SimHerd model was developed and validated

for Danish production circumstances (Sørensen

et al., 1992; Østergaard et al., 2000, 2003). An

animal at any given time is characterized by its age,

lactation stage, lactation number, actual body

weight, milk yield, pregnancy status, and disease

occurrence. All discrete events at the animal level

(e.g., heat detection, conception, diseases and in-

voluntary culling) are stochastic variables randomly

sampled from relevant distributions (see Sørensen

et al., 1992; Østergard et al., 2000), which introduce

variation among phenotypic performance of cows. In

a dynamic model, production of the herd is the

accumulation of state changes and production of

Stochastic model for traits in dairy cattle 17

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 4: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

individual cows over time (Sørensen et al., 1992).

Genetic effects at the animal level are expressed at

the herd level through the realised phenotypic

expressions given at the herd level.

The number of heifer and bull calves at the farm is

controlled by a probability of birth of a heifer calf

and by the stillbirth rate in the herd. Heifer calves

are raised at the farm to be used as replacements.

The heifer enters the milking herd if the actual

number of cows in the herd is less than the

maximum capacity or if at least one cow has been

selected for culling. Surplus heifers for replacements

are sold at the age of calving. A replacement heifer

will be purchased if no heifers born within the herd

are available at the time of culling of a cow and the

herd size has reached the specified minimum. Herd

dynamics arise from state changes of the animals and

from variation in number of animals due to culling in

the herd. The culling strategy is controlled by herd

size (maximum number of cows in the herd), a

maximum number of days open (specified separately

for low and high yielding cows) and low milk yield

(specified separately for first, second and third parity

cows, respectively).

Genetic levels for milk yield, body weight, repro-

duction, culling, stillbirth and diseases are the basis

for phenotypic performance of cows in the SimHerd

model. Phenotypic performance of a cow is a

combination of genetic effects and environmental

components from the stochastic element of the

model. In addition, traits are related via non-linear

relationships (indirect effects). Change in genetic

level of traits is performed by independent changes

in level of each trait. However, due to the indirect

effects the final phenotypic outcome of (related)

traits can change when genetic changes are per-

formed. For example, genetic level for milk yield

partly determines potential feed intake, which also

influences phenotypic cow performance of body

weight. The phenotypic performance of cows and

input and output of animals from SimHerd are

entered into SimProfit. In SimProfit, calving perfor-

mance, temperament, milking speed and beef pro-

duction of bull calves are simulated deterministically

and these traits are uncorrelated. Input and output

from both SimHerd and SimProfit and economic

parameters are combined to give revenue and costs

per farm per year. Details on modelling of the

different traits are given in later paragraphs.

Farm characteristics and production system

An average Danish farm with regard to production

level, production system and management strategy

was chosen as the basic situation. The simulated

dairy farm represented a loose-house production

system with cubicles and a 6�/2 herringbone milking

parlour. Beef production from bull calves was

evaluated as an integrated part of the dairy cattle

production system, which is common practice in

Denmark. In 2000, the average farm size in Den-

mark was 70 milking cows with additional young

stock. In this study, maximum number of cows in the

herd was 72. Cows were inseminated until day 210

or day 168 after calving for high and low yielding

cows, respectively. After designation for culling, both

high and low yielding cows were latest culled at day

392 after calving. Voluntary culling thresholds of

cows were 25, 32, and 33 kg energy corrected milk

(ECM) per day as an average over weeks 1 to 24 after

TIFORPMISDREHMIS

stiart fo slevel citeneG

esae gnivlaC

tnemarepmeT

deeps gnikliM

)sevlac llub( noitacifissalc ssacraC

)sevlac llub( niag yliaD

scimonocE stsoc dna seuneveR

seulav cimonocE

setats woC

egA

rebmun noitatcaL

egats noitatcaL

sciteneg woC ecnerrucer esaesiD

yticapac dleiy kliM

,revef klim ,sititsam( ksir esaesid cisaB

)atnecalp deniater ,sitinimal

taeh tsrif ot gnivlac morf syaD

noitpecnoc fo ecnahC

thgiew erutaM

gnilluc yratnulovni fo ksiR

etar htribllitS

yticapac ekatni deeF

erocs noitidnoC ygetarts dreH

gnilluc yratnuloV

ygetarts gnideeF

eR p etarts noitcudor yg

cipytonehP

woc

ecnamrofrep

slaitnetop woC

+

tnemnorivnE

)scitsahcotS(

Figure 1. Overview of the model used to derive economic values.

18 H.M. Nielsen et al.

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 5: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

calving for first, second, and later parities, respec-

tively.

Until 24 weeks after calving all milking cows and

culled cows were fed ad libitum the same total mixed

ration with a fill unit of 0.37 (volume per Scandina-

vian feeding unit). Later, cows were assigned to two

feeding groups according to their current milk yield.

High yielding cows after 24 weeks after calving

(ECM above 23 and 28 for first parity and later

parity cows) were fed a ration with a fill unit of 0.46.

This ration was also used for the dry cows, but these

were fed restrictively. Low yielding cows (ECM

above 18 and 22 for first and later parity cows) after

24 weeks since calving were fed a ration with a fill unit

of 0.55. Feeding rations and feeding groups for cows

might possibly change with increased yield levels of

cows. However, preliminary analysis showed that

the feeding rations and groups described above were

also optimal for cows at higher yield levels. Therefore

the same feeding plan and feeding groups before

and after a genetic change in milk yield were used.

Profit equations

Profit of the farm was calculated as:

P�R�C

where,

R�/revenues of the farm (t/year)

C�/costs of the farm (t/year)

Milk, beef production and sales of surplus heifers

provided the revenues of the farm. Revenue from

beef production came both from fattening of bull

calves and culling of cows. Beef revenue of culled

cows was determined by the live weight of culled

cows and price per kg live weight. Variation in

carcass classification was not considered. However,

differentiation was made for culled cows, which had

a slaughter value, and dead cows, for which the

farmer had to pay a disposal price (Daka, 2001).

Subsidy denotes price compensation paid to the

farmer due to the low market price on beef. This

price compensation was a specific amount per

slaughtered bull calf (SUBSBC) or per slaughtered

cow (SUBS) (Direktoratet for FødevareErhverv,

2000). The number of milking cows on the farm is

given as number of cow-years defined as all feeding

days divided by days in the year:

R�N�[(ECM�PECM)

�NSCOW�(SUBS�(BCOW�PBCOW))

�(NDCOW�PDCOW)

�NBC�(SUBSBC�RBCALV)

�RHEIF�PRHEIF] (1)

where,

N�/number of milking cows on the farm (cow-year)

ECM�/energy corrected milk (kg/cow-year)

PECM�/price per kg of energy corrected milk (t)

NSCOW�/number of slaughtered cows per cow-

year

BCOW�/average live weight per slaughter cow (kg)

PBCOW�/price of slaughter cow (t/kg)

SUBS�/subsidy per slaughtered cow (t)

NDCOW�/number of dead cows per cow-year

PDCOW�/cost per dead cow (t)

NBC�/number of bull calves finished per cow-year

SUBSBC�/subsidy per bull calf (t)

RBCALV�/revenue per bull calf per cow-year (t/

year)

RHEIF�/number of heifers sold per cow-year

PRHEIF�/price of mature heifer (t/heifer)

Farm costs were defined using the approach by

Groen et al. (1997):

C�CVC�CFC�CFF

where

CVC�/variable cow costs

CFC�/fixed cow costs

CFF�/fixed farm costs

CVC�N�(CONC�PCONC�ROUG�PROUG

�Ccalving�Cmastitis�Cmilkfever�Crplacenta

�Claminitis�Ctemper�Cmilking�OTHER

�AICOW�NRHEIF�PRHEIF

�NBC�CBC) (2)

where further,

CONC�/concentrate (Scandinavian feeding unit

(SFU)/cow-year, 1 SFU�/7.89 MJ NE)

PCONC�/price of concentrate (t/SFU)

ROUG�/roughage (SFU/cow-year)

PROUG�/price of roughage (t/SFU)

OTHER�/other costs (t/cow-year)

Cmastitis�/cost of mastitis (t/cow-year)

Cmilkfever�/cost of milk fever (t/cow-year)

Crplacenta�/cost of retained placenta (t/cow-year)

Claminitis�/cost of laminitis (t/cow-year)

Ccalving�/cost of calving ease (t/cow-year)

Cmilking�/cost of milking (t/cow-year)

Ctemper�/cost of temperament (t/cow-year)

AICOW�/insemination costs (t/cow-year)

NRHEIF�/number of replacement heifers per cow-

year

PRHEIF�/cost of replacement heifer (t/heifer)

CBC�/cost per bull calf (t/bull calf)

Variable costs (CVC) included costs of feed, dis-

eases, calving, milking, temperament, insemination,

replacement and costs of producing bull calves.

Stochastic model for traits in dairy cattle 19

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 6: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

Costs of mastitis, milk fever, retained placenta and

laminitis per cow value were calculated as the cost in

t per incidence (Table I) multiplied by the number

of incidences per cow-year. Costs per incidence of

mastitis, milk fever and laminitis included labour,

veterinary treatment (fee and medicine) and dis-

carded milk (Danish Agricultural Advisory Centre,

2002), whereas costs for retained placenta only

included labour and veterinary treatment. Perma-

nent reduced milk yield and reduced feed intake was

not accounted for. Replacement costs included costs

of rearing all heifers born at the farm and cost of

purchased heifers. Bull calves from the milking cows

in the herd were kept and raised at the farm.

Therefore, revenues of selling newborn bull calves

and costs of purchasing newborn bull calves were not

included in the profit equations. Other costs

(OTHER) were mainly veterinary service, which

was not specified in the cost equation and cost of

milking control association (SJFI, 2000). Fixed cow

costs (CFC) included costs of labour, milking

parlour, electricity and housing and were based on

statistics from Economics of Agricultural enterprises

(SJFI, 2000). Prices in EURO (current value No-

vember 10, 2005: 1 t�/1.18 US$) used to derive

economic values are given in Table I. Average market

prices (2001) of milk (Danish Dairy Board, 2001),

beef (Danish Meat Board, 2002) and feed (LK,

2001) were used.

Traits

Data for cow performance were chosen to represent

an average Holstein cow in Denmark and were

supplied from the national cattle database (Bund-

gaard & Hoej, 2000). Economic values were derived

for 10 different traits. The traits were milk yield,

conception rate, days from calving until first heat,

stillbirth, incidence of mastitis, incidence of retained

placenta, incidence of milk fever, incidence of

laminitis, body weight, and involuntary culling rate.

Conformation traits were not included since they are

predictors for functional traits (e.g., leg conforma-

tion as predictor for laminitis). Therefore, most

conformation traits do not have a direct economic

value and should be included in the selection index

instead of in the breeding goal. Genetic levels of milk

yield, stillbirth, days from calving until first heat,

conception rate, involuntary culling rate, and body

weight are in Table I, whereas genetic levels for the

disease traits are in Table II (given as basic disease

risks). However, more traits were modelled to give

Table I. Applied prices (t) and genetic levels of traits before and after genetic change of 20%.

Abbreviation Value

Prices/costs

Price of energy corrected milk (t/kg) PECM 0.34

Price of culled cow (live weight) (t/kg) PBCOW 0.94

Subsidy (t/slaughter cow) SUBS 53

Price per dead cow paid by the farmer PDCOW 100

Price of concentrate (t/SFU) PCONC 0.15

Price of roughage (t/SFU) PROUG 0.13

Price of replacement heifer (t/heifer) PRHEIF 942

Cost of mastitis (t/incidence) COSTmas 131

Cost of milk fever (t/incidence) COSTmf 167

Cost of retained placenta (t/incidence) COSTrpl 48.4

Cost of laminitis (t/incidence) COSTlam 112

Labour price (t/hour) Plabour 15.5

Insemination cost (t/insemination) AI 14.8

Other costs (t/cow-year) OTHER 202

Fixed cow costs (t/cow-year) CFC 471

Fixed farm costs (t/farm/year) CFF 33647

Trait Level in basic situation Level after genetic change

Energy corrected milk (kg/cow-year)1 8340 10008

Stillbirth (%), first lactation 12 10

Stillbirth (%), later lactations 8 6

Days from calving to first heat2 35 28

Conception rate (%)2 60 72

Involuntary culling rate3 24 19

Body weight (kg/cow)4 630 756

1Milk yield per year for a cow in third lactation.2Average of all lactations.3% of involuntary culling in the herd per year.4Mature body weight for a cow in third lactation adjusted for foetus weight.

20 H.M. Nielsen et al.

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 7: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

performance of cows, heifers and bull calves to be

included in the profit equations (Equations 1 and 2).

A detailed description of all traits is given in the

following paragraphs.

Milk yield, body weight and feed intake capacity. Milk

yield of the cow was described by milk yield capacity,

milk yield potential and actual phenotypic milk yield

as described by Sørensen et al. (1992). Milk yield

potential at a certain stage of lactation was modelled

using an incomplete Gamma function (Wood,

1967). Milk yield capacity described the aver-

age predicted yield in energy corrected milk 1 to

24 weeks after calving for a third parity cow not

diseased. Average daily level of milk yield capacity

for a third parity cow was set to 33.4 kg, correspond-

ing to a lactation yield of 8340 kg ECM per cow-year

(see also Table I). For each individual cow, milk yield

capacity assigned at the time of birth was sampled

from a normal distribution with a mean correspond-

ing to the average genetic milk yield capacity in the

herd and a standard deviation of 2. Additionally, at

each calving a random deviation was sampled from a

normal distribution and added to the milk yield

capacity of the cow to create variation in milk yield

between lactations of the same cow.

Change in genetic level of body weight was

modelled by increasing mature weight (defined as

the weight of a mature cow with a body condition

score of 2). Mature weight of the cow was 630 kg.

Variation in actual expressed (phenotypic) body

weight for the cow was created through variation in

milk yield capacity and disease occurrence. Growth

potential was modelled using a Gompertz growth

function (Sørensen et al., 1992). Potential body

condition score (BCS) was simulated from a Gom-

pertz weight curve with body condition score of

3 (scored from 1 to 5) (Østergaard et al., 2003).

Actual BCS of the cow was calculated from simu-

lated actual body weight corrected for foetus weight

and from potential BCS. Feed intake capacity of the

cow was modelled using equations corrected for

parity, stage of lactation and weight at calving.

Potential feed intake was then calculated from feed

intake capacity and milk yield capacity as described

by Sørensen et al. (1992). From potential feed

intake, actual energy intake was calculated depen-

dent on feed composition. Next, energy require-

ments based on milk yield potential and growth

potential were calculated. Finally, actual energy

intake was compared with requirements based on

growth potential and potential milk yield to give

actual (phenotypic) milk yield and change in body

weight. In situations where available energy was

insufficient to meet milk yield potential and potential

growth, a rule for distribution between milk produc-

tion and growth was used to calculate actual

phenotypic milk yield and body weight. Details

regarding equations for distribution rules between

milk production and growth are described in Sør-

ensen et al. (1992).

Table II. Risk factors and effects of diseases.

Mastitis Milk fever Retained placenta Laminitis

Risk factors

Base risk1 (level in basic situation) 0.259 0.075 0.112 0.067

Base risk (level after genetic change�/�/20% from basic situation) 0.207 0.060 0.090 0.054

Parity 1 vs. parity 3, OR2 0.8 0.01 0.60 0.50

Parity 2 vs. parity 3, OR 0.9 0.25 0.85 0.75

Parity 4 vs. parity 3, OR 1.0 1.8 1.15 1.25

Same disease in previous lactation, OR 1.5 4.0 2.0 1.2

Threshold for high or low BCS3 4.0

High risk BCS vs. normal risk BCS, OR4 4.3

Milk yield potential, OR per kg above herd average 1.04 1.04

Milk fever in current lactation, OR 1.1 2.0

Lactation stage, a-parameter in gamma-distribution. 0.60

Lactation stage, b-parameter in gamma-distribution 86

Effects of diseases

Death, risk (%) 1 8 1

Risk of removal (%)5 4 2

1Base risk is defined as the risk for a cow in third lactation with average yield capacity and without any previous cases of diseases.2OR�/odds ratio, an OR of 0.8 means that the relationship between diseased and not diseased first parity cows is 0.8 times the relationship

between diseased and not diseased third parity cows.3BCS (body condition score) threshold used for grouping cows into high or low risk BCS.4The risk of a cow with BCS above the BCS threshold vs. the risk of a cow below the BCS threshold.5Risk of removal in the week where the cow fell ill given that the cow did not die from the disease.

Stochastic model for traits in dairy cattle 21

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 8: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

Health, fertility, calving and survival traits. Health

traits were modelled using disease risks represented

by a logistic regression model as described by

Østergaard et al. (2000). Starting point is the

definition of a standard cow defined as a cow in

third lactation with average yield capacity and with-

out any previous case of diseases. Disease risk for the

standard cow is called the basic disease risk, the

basic risk being the intercept in the regression model.

For cows deviating from the standard cow (e.g.,

cows in second lactation and cows with milk yield

above average herd level), other risks were added to

the basic disease risk to give a total risk of each

disease. Other risks of diseases were dependent on

the trait under consideration (Table II) and were

modelled by odds ratios (OR). The total disease risk

is calculated as 1/[1�/((1/basic risk�/1)/PORixi)],

the ORi being the OR for each risk factor and xi a

value for each risk factor. Total disease risk for

mastitis for a first parity cow with milk yield capacity

2 kg above herd average is thus 1/[1�/((1/0.259�/1)/

(0.8*1.042))]�/0.232. If the total disease risk is

higher than a random number from a uniform

distribution, the disease is triggered. Parity as a risk

factor of laminitis was based on literature estimates

(Groehn et al., 1992). Milk fever was a risk factor for

both mastitis and retained placenta. Therefore, a

cow suffering from milk fever would have an

increased risk of developing both mastitis and

retained placenta. Indirect effects of diseases were

also increased risk of dying and risk of culling in the

week where the cow fell ill, given that the cow did

not die from the disease. Disease traits in the

breeding goal were defined as disease incidence,

but genetic changes were obtained by changing the

relevant basic disease risk (Table II).

Fertility traits included conception rate and days

from calving to first heat. Chance of conception and

showing heat were triggered stochastically from

average herd rate and occurrence for each cow

sampled from a distribution. Conception rate was

60% and number of days from calving until first heat

was 35. In addition to the farmers’ culling strategy

and the culling and death due to diseases, culling of

cows was controlled by a risk of involuntary culling

expressed as percentage of involuntary culling at the

farm per year. Involuntary culling in the herd did not

depend on voluntary culling in the herd or on

maximum herd size allowed. The trait involuntary

culling rate was modelled by varying the risk of

involuntary culling in the herd with a starting value

of 24%.

Different frequencies of stillbirth for first and later

parities were used. Stillbirth was modelled through

an average herd rate and occurrence for each cow

was sampled from a distribution. Stillbirths of cows

in first and later parities were 12% and 8%,

respectively. When performing the genetic change,

stillbirth rate for both first parity cows and later

parity cows was changed. Cost of calving ease

(Ccalving) was calculated treating calving ease as a

categorical trait with four classes (see Meijering,

1986) as given in Equation 3.

Ccalving�(F(t2�m)�F(t1�m))�cewh

�(F(t3�m)�F(t2�m))�cdifficult

�(1�F(t3�m))�cvethelp) (3)

where,

F(ti)�/cumulative distribution function of a stan-

dard normal distribution N(0,1) with thresholds ti,

i�/1,2,3

m�/mean

cewh�/cost of class ‘easy with help’

cdifficult�/cost of class ‘difficult’

cvethelp�/cost of category ‘difficult with veterinary

assistance’

Costs of calving differed according to degree of

farmer labour and veterinary assistance. Percentage

of cows was 75, 21.4, 2.3, and 1.3 (Bundgaard &

Hoej, 2000) and costs were 0, 4, 15.5 and 178t for

classes easy, easy with help, difficult and difficult

with veterinary assistance, respectively.

Workability traits. Two different workability traits

were considered, milking speed and temperament.

Costs were mainly derived from assumed labour

requirements associated with handling the cow.

Assuming an optimized system the milking costs

were derived based on the number of cows, the

handling time per cow and the milking time of the

cows. The milking time of the cow was dependent on

the milk production and the milk flow rate. Milking

speed as an average over lactation and parities was

2.3 l/min and the handling time per cow per milking

was assumed to be 0.5 min.

MILKCOST�PLABOUR

��

HTIME��

1�KGMILK

FLOWRATE

��

where,

MILKCOST�/milking costs (t/cow-year)

PLABOUR�/labour price (t/min)

HTIME�/handling time (min/cow)

KGMILK�/milk yield (l/cow)

FLOWRATE�/milk flow rate (l/min)

The cost of temperament (Ctemper) was modelled

using a threshold model with two classes of tempera-

ment (1�/nervous, 2�/normal). Handling costs of a

22 H.M. Nielsen et al.

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 9: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

cow in class ‘nervous’ were assumed to correspond

to 94 t/cow-year, whereas handling of cow in class

‘normal’ was assumed not to include costs. Percen-

tages of cows in temperament class 1 and 2 were 4.7

and 95.3, respectively (Bundgaard & Hoej, 2000).

Ctemper�(F(t1�m))�cnervous�(1�F(t1�m))

�cnormal

where,

F(t1)�/cumulative distribution function of a stan-

dard normal distribution with threshold t1cnervous, cnormal�/labour costs (t/cow-year) of tem-

perament class ‘‘nervous’’ and ‘‘normal’’ respectively

derived as labour time (hour/cow-year) times labour

price (t/hour)

Beef production from bull calves

Gain and classification represented beef production

characteristics for bull calves. Cost per produced

bull calf (CBC) was calculated as:

CBC�(1�MORTBC)

�((CFEED�CNONFEED)�FEEDDAYS)

CFEED�FEEDgain�PFEED

where,

MORTBC�/mortality of bull calves (%)

CFEED�/total feed costs (t/day)

PFEED�/price of feed (t/SFU)

CNONFEED�/non-feed costs (t/day)

FEEDgain�/feed requirements for gain (SFU/kg

gain)

Non-feed costs included veterinary treatment, la-

bour and miscellaneous (SJFI, 2000). Parameters

used to calculate cost of beef production are given in

Table III.

The number of days (FEEDDAYS) to feed the

calf was calculated as

FEEDDAYS�WSL � BWEIGHT

DAYGAIN

where,

WSL�/live weight at slaughter (kg)

BWEIGHT�/birth weight (kg)

DAYGAIN�/average daily gain (kg/day)

In Denmark, carcass classification of bulls is based

on the European classification system (EUROP)

(EEC, 1981). Carcasses are scored for form using

a 5-point scale (E,U,R,O,P) and 3 subclasses

(�/,0,�/) giving 15 different classes. However, owing

to very low frequency of animals in some classes, the

15 classes were merged into five classes. Average

revenue from beef production (RBCALV) depen-

dent on classification was calculated with a threshold

model with five classes of classification.

RBCALV� (F(t1�m))�pO�

�(F(t2�m)�F(t1�m))�pO

�(F(t3�m)�F(t2�m))�pO�

�(F(t4�m)�F(t3�m))�pP�

�(1�F(t4�m))�pP;

where,

F(ti)�/cumulative distribution function of a stan-

dard normal distribution with thresholds ti (I�/

1, . . . ,4)

PO�, PO, PO�, PP�, PP�/price per kg slaughter

weight of calf with classification class O�/, O, O�/,

P�/, P, respectively (Danish Meat Board, 2002)

Frequencies of bull calves in classes O�/, O, O�/,

P�/, P were 2.3, 17.9, 56.8, 20.7, and 2.3% and

prices were 2.22, 2.15, 2.09, 1.9, and 1.76 t/kg

slaughter weight. Slaughter weight was calculated by

multiplying live weight at slaughter (WSL) with the

dressing percentage (DPCT).

Economic values

Economic values were derived with a fixed number

of cows as basis of evaluation (Groen et al., 1997).

Profit was chosen as the interest of selection, because

usually the farmer decides which animals to use.

When deriving economic values, the model was first

run with all 10 traits at their basic levels (Tables I

and II) using 1000 replicates (1 basic situation).

Secondly, the model was run after changing the level

of each of the 10 traits one at a time by 20% from the

level in the basic situation. This gave 10 alternative

situations, which were each run using 1000 repli-

cates. Economic values for the 10 traits were derived

by comparing farm profit before (basic situation)

and after changing the genetic level of each trait

Table III. Biological parameters, prices and cost applied in beef

production from bull calves.

Parameter Abbreviation Value

Biological parameters

Average gain (g/day) (bull calves) DAYGAIN 1236

Live weight at slaughter (kg) WSL 450

Birth weight (kg) BWEIGHT 40

Mortality (%) MORTBC 7

Dressing percentage (%) DPCT 51

Feed requirements for gain (SFU/kg) FEEDgain 5.9

Prices and costs

Feed price (t/SFU) PFEED 0.16

Non-feed costs (t/day) CNONFEED 0.47

Subsidy (t/bull calf) 259

Stochastic model for traits in dairy cattle 23

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 10: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

(alternative situation) from the basic situation. Each

simulation was run over a 10-year period. However,

to overcome the influence of the initial herd and to

allow the effect of management and production

system to be expressed, results for each simulation

were based on an average of the last five years of 10

years simulated.

Using a fixed number of cows as the basis of

evaluation, change in revenues and costs are gen-

erally due to the difference between marginal rev-

enue and marginal cost of production only, i.e., fixed

cow costs and fixed farm costs do not change (Groen

et al., 1997). The number of cow-years in each

alternative could differ from the number of cow-

years in the basic situation due to the stochastic

element of the model. Consequently, fixed costs per

cow varied as well. To account for the change in the

number of cow-years in the alternative situation, the

level of each trait was adjusted for number of cow-

years in the alternative situation (see Equation 4).

When deriving economic values, change in profit of

the farm was divided by the actual realized pheno-

typic change in the respective traits and not the

genetic change. For the disease traits this means that

the trait level was disease incidence and not basic

disease risk, which was used when performing the

genetic change.

EV�PROFITalt � PROFITbasic

LEVELalt � COWSalt � LEVELbasic � COWSbasic

(4)

where,

EV�/economic value (t/trait unit per cow-year)

PROFITalt�/farm profit (t/year) for the alternative

situation as an average of 1000 replicates

PROFITbasic�/farm profit (t/year) for basic situa-

tion as an average of 1000 replicates

LEVELalt�/level of trait for the alternative situation

as an average of 1000 replicates

LEVELbasic�/level of trait for the basic situation as

an average of 1000 replicates

COWSalt�/number of cow-years for the alternative

situation as an average of 1000 replicates

COWSbasic�/number of cow-years for the basic

situation as an average of 1000 replicates

Standard deviations of economic values were

computed using a linear Taylor series expansion,

where covariances and variances between parameters

in Equation 4 were accounted for. These standard

deviations represent variation in realized profit of the

farm due to variation in expression of cow perfor-

mance traits before and after a genetic change.

Unless specified otherwise, standard deviations are

the variation in economic values between two

randomly chosen replicates and not the inaccuracy

on the estimated economic values (standard errors),

which can be calculated by dividing by the square

root of the number of replicates.

Simulated scenarios

To describe the farm in terms of production para-

meters the model was first run with all traits at their

basic levels (Tables I and II). Secondly, using

incidence of mastitis as an example, economic values

were derived for different marginal changes (change

in basic risk of mastitis (Table II) of 5% to 50%).

Next, economic values were derived for 10 different

traits (milk yield, conception rate, days from calving

until first heat, stillbirth, incidence of mastitis,

incidence of retained placenta, incidence of milk

fever, incidence of laminitis, body weight, and

involuntary culling rate).

Results

Production parameters in the basic situation

After running the model with all traits at their basic

level, average milk yield of cows in the herd was

8519 kg ECM per cow-year, feed intake was 5560

SFU per cow-year and average body weight of

slaughter cows was 566 kg (Table IV). Number of

first incidence of mastitis, retained placenta, milk

fever and laminitis per cow-year were 0.32, 0.17,

0.10 and 0.07, respectively. The average productive

herd life was 961 days.

Marginal changes

With a marginal change of less than 20%, eco-

nomic values for mastitis varied from �/141.1 to

�/375.5 t/incidence per cow-year (see Table V).

However, from 20% change to 50% change the

economic values ranged from �/148.9 to �/168.4

t/incidence per cow-year. Standard deviations de-

creased with increasing marginal changes. With 5%

change and 20% change, standard deviations of the

economic values were 2860 and 682, respectively.

Because economic values for the disease traits were

unstable (the standard errors of the estimates were

high) with a change of less than 20% and because

the standard deviations of the economic values

depended on the size of the change, economic

values for all traits are represented with a 20%

change.

Production parameters after genetic change

After changing the genetic level of diseases by 20%

the simulation model gave a decrease in incidences

of mastitis and retained placenta with 0.07 and 0.03

24 H.M. Nielsen et al.

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 11: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

incidences per cow-year (see Table VI). For milk

fever and laminitis corresponding numbers were

0.02 and 0.01 incidences per cow-year.

Changing the genetic level for milk yield increased

milk yield in the herd by 1238 kg per cow-year.

Correspondingly, feed intake increased by 516 SFU

per cow-year (Table VI). The result of increased milk

yield was also a weight loss of 16 kg per cow-year.

The weight loss was due to the rule of energy

distribution between milk yield and growth, which

was used in situations where available energy was

insufficient to meet potential milk yield and growth.

By increasing the yield capacity of the cow, relatively

more energy was used for milk production and less

was available for gain. Likewise, increasing the

mature weight of the cow resulted in decreased

milk yield (Figure 2). This effect was seen when

available energy was insufficient to meet potential

milk yield and growth because relatively more energy

was used for gain.

Economic values and standard deviations

Changes in revenues and cost of the farm and

correspondingly economic values are given in Table

VII. The economic value for energy corrected milk is

from increased revenue from milk and increased feed

cost due to increased feed requirements. The in-

crease in genetic level of milk production resulted in

an increase in farm profit corresponding to 24,859 t

per year. The economic value per kg energy cor-

rected milk was 0.28 t per cow-year. Estimated

economic value of body weight was �/0.76 t/kg per

cow-year, and was mainly derived from increased

revenue from beef production of culled cows and

increased feed requirements. Due to indirect effects

Table IV. Simulated cow performance (mean, standard deviations and coefficient of variation) for the basic situation based on 1000

replicates.

Trait Mean Standard deviation Coefficient of variation

Number of cow-years 70.8 0.2 (0.003)

Energy corrected milk (kg/cow-year) 8519 63.9 (0.007)

Mastitis (incidence/cow-year) 0.32 0.03 (0.09)

Retained placenta (incidence/cow-year) 0.17 0.02 (0.12)

Milk fever (incidence/cow-year) 0.10 0.02 (0.2)

Laminitis (incidence/cow-year) 0.07 0.01 (0.14)

Feed intake (SFU/cow-year) 5560 29.4 (0.005)

Calvings per cow-year 1.2 0.02 (0.02)

Number of slaughter cows (farm/year) 25.1 2.0 (0.08)

Weight of slaughter cows (kg/cow) 566 7.4 (0.01)

Replacement rate (%/farm/year) 39.1 3.1 (0.08)

Dead cows (%/farm/year) 3.3 1.0 (0.3)

Herd life (days) 961 82.8 (0.09)

Stillborn calves (%/farm/year) 9.2 1.4 (0.15)

Number of heifers sold (farm/year) 11.3 2.5 (0.22)

Number of bull calves produced per year 44.1 2.1 (0.05)

Inseminations per cow-year 1.9 0.05 (0.03)

Table V. Economic values and standard deviations (S.D.) for incidence of mastitis with different magnitude of change from the level of

mastitis in the basic situation. Standard errors of estimated economic values (S.E.) based on 1000 replicates are given in parentheses.

Change in level of trait from

basic situation Mastitisalt1 Change in profit

Economic value

(t/incidence per cow-year)

Mean9/S.D. (S.E.)

�/5% 0.309/0.03 4119/2161 �/375.59/2860 (90.4)

�/10% 0.299/0.03 3209/2036 �/141.19/1327 (42.0)

�/15% 0.279/0.02 6029/2133 �/183.09/915 (28.9)

�/20% 0.259/0.02 7349/2125 �/162.59/682 (21.6)

�/25% 0.249/0.02 8879/2074 �/150.89/505 (16.0)

�/30% 0.229/0.02 10409/2081 �/148.99/442 (14.0)

�/35% 0.209/0.02 13649/2088 �/168.49/368 (11.6)

�/40% 0.199/0.02 14319/2102 �/154.69/322 (10.2)

�/45% 0.179/0.02 16299/2080 �/155.79/282 (8.9)

�/50% 0.159/0.02 17719/2098 �/153.79/259 (8.2)

1Level of trait (incidence/cow-year) and standard deviation in the alternative situation; basic situation 0.329/0.03 (0.09).

Stochastic model for traits in dairy cattle 25

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 12: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

between milk yield, feed intake and body weight,

milk yield and consequently milk revenues were

reduced, when mature body weight was increased

(Figure 2).

Estimated economic value for involuntary culling

was �/6.66 t/% involuntary culling per cow-year.

Increased average herd life means that fewer replace-

ment heifers are needed, i.e., more heifers can be

sold and relatively more cows are in the older age

groups. Therefore, an increase in milk production

was seen. In addition, a small increase in incidence

of milk fever due to the risk factor of parity was seen.

For mastitis, retained placenta, milk fever and

laminitis economic values were �/162.5, �/79.0,

�/402.1 and �/210.2 t per incidence per cow-year,

respectively. The economic values for diseases arise

mainly from decreased costs associated with dis-

eases. However, for milk fever, a small increase in

average herd life was observed.

The economic value of conception rate of 1.98 (t/

% per cow-year) originates from decreased costs of

inseminations per cow and from increased milk

revenues due to increased average herd life (fewer

cows were culled because of failure to get in calf).

The economic value for number of days from

insemination until first heat was �/0.94 t/day per

cow-year. This value was from shorter calving

interval and decreased costs related to culling due

to more cows below the culling threshold of days

open. However, a small increase in insemination

costs was observed, because of lower chance of

conceiving for cows inseminated at an earlier stage

after calving. Reducing the stillbirth rate among

calves means that more heifer calves and bull calves

can be reared. In this model, heifers were reared at

the farm, and at calving they were either sold or

transferred to the milking herd. The price difference

between the revenue of a sold heifer and rearing the

heifer was limited. Hence, the economic value of

stillbirth was mainly determined from the profit in

producing bull calves.

Standard deviations of the economic values ran-

ged between 0.02 (milk production) to 2936 (lami-

nitis). The economic value for production traits had

the lowest standard deviation, whereas the highest

standard deviations were found among the func-

tional traits. The standard deviations of economic

values were from replicates of a farm with about 70

cows and express variation in realized profit due to

variation in expression of traits before and after a

genetic change. These standard deviations are the

variation in economic values between two randomly

chosen replicates and not the inaccuracy in the

estimate of the economic value (standard error).

On average over the 1000 replicates, the genetic

change was realized (e.g., mastitis incidence in theTab

leV

I.C

ow

per

form

an

cele

vel

sfo

rth

ebasi

csi

tuati

on

an

dm

arg

inal

chan

ge

aft

erch

an

gin

ggen

etic

leve

lof

20%

.

Ch

an

ge

inle

vel

of

Tra

itB

asi

cE

CM

Con

cep

tion

rate

D.

f.c.

tofi

rst

hea

tS

tillb

irth

Mast

itis

Ret

ain

ed

pla

cen

taM

ilk

feve

rL

am

init

is

Bod

y

wei

gh

t

Involu

nta

ry

cullin

g

Nu

mb

erof

cow

-yea

rs70.7

70.0

40.0

70.0

30.0

40.0

10.0

10.0

20

00.2

4

En

ergy

corr

ecte

dm

ilk

(kg/c

ow

-yea

r)8519.3

41238.2

512.0

621.2

71.5

42.4

22.0

8.2

82.1

�/2

03.0

744.2

5

Mast

itis

(in

cid

ence

/cow

-yea

r)0.3

2�

/0.0

10

0.0

1�

/0.0

2�

/0.0

70

00

00

Ret

ain

edpla

cen

ta(i

nci

den

ce/c

ow

-yea

r)0.1

70

0.0

10

00

�/0

.03

00

00

Milk

feve

r(i

nci

den

ce/c

ow

-yea

r)0.1

0�

/0.0

10.0

10

00

0�

/0.0

20

00.0

1

Lam

init

is(i

nci

den

ce/c

ow

-yea

r)0.0

70

0.0

10

00

00

�/0

.01

00

Fee

din

take

(SF

U/c

ow

-yea

r)5559.0

2516.2

83.0

39.9

�/0

.66

0.7

70.8

33.3

80.2

225.7

21.9

9

Wei

ght

of

slau

ghte

rco

w(k

g/c

ow

-yea

r)566.1

8�

/16.1

40.3

7�

/0.6

2�

/0.7

1�

/0.5

3�

/0.2

9�

/0.0

7�

/0.4

191.9

26.3

Dea

dco

ws

(%/f

arm

/yea

r)3.3

2�

/0.1

20.1

0.1

1�

/0.0

6�

/0.0

7�

/0.0

4�

/0.1

9�

/0.1

10

0.0

4

Rep

lace

men

tra

te(%

/farm

/yea

r)39.0

7�

/3.5

1�

/3.3

7�

/0.6

50.0

2�

/0.1

�/0

.06

�/0

.3�

/0.0

40.0

8�

/4.4

5

Her

dlife

(day

s)961.0

399.1

294.3

915.2

8�

/1.6

51.7

60.1

96.5

9�

/0.3

8�

/3.1

6127.8

4

Sti

llb

irth

calv

es(%

/farm

/yea

r)9.3

�/0

.11

�/0

.22

�/0

.1�

/1.9

60.0

3�

/0.0

6�

/0.0

6�

/0.0

90.0

4�

/0.2

Inse

min

ati

on

sper

cow

-yea

r1.8

80

�/0

.28

0.0

50

00

00

0�

/0.0

1

26 H.M. Nielsen et al.

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 13: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

herd lower after the genetic change than in the basic

situation as illustrated in Figure 3a,b). Conse-

quently, average farm profit increased. Genetic

change in some replicates was not realized in the

herd and incidence of mastitis was higher than

average incidence of mastitis in the basic situation.

In some of the replicates, profit was therefore lower

after a genetic change than in the basic situation.

Discussion

The aim of this study was to present a model of a

dairy cattle production system for the derivation of

economic values and their standard deviations. The

model was suitable for the derivation of economic

values and standard deviations for both production

and functional traits in dairy cattle assuming Danish

production circumstances. However, more sensitiv-

ity analyses need to be performed for further

validation of the model. Additionally, this study

only considered one set of production parameters.

Because of uncertainty among future production

circumstances, scenario studies should show sensi-

tivity to change in production circumstances.

Model assumptions

Economic values were derived using a marginal

change of 20%. A change of 20% in trait level was

chosen because of unstable economic values (high

standard errors of the economic values) for disease

traits due to the stochastic element of the model (see

Table V). For other traits such as milk production, a

smaller change (e.g., 1%) was possible. Because

standard deviations and standard errors of the

estimated economic values depended on the size of

the change and to be able to compare standard

deviations between traits, the same change was

performed for all traits. Preferably, smaller changes

would have been performed. This would correspond

more closely to the actual genetic change obtained

after one generation of selection. A change of 20% in

milk yield can be obtained after 10 to 20 years of

selection (Danish Cattle Federation, 2003). Eco-

nomic values obtained in this study correspond to

economic values in the long run. However, for other

traits such as disease traits, genetic improvement is

at a lower rate (Danish Cattle Federation, 2003).

Therefore, economic values might be biased when

derived using a large change; the actual realized

change is small and economic values change with

magnitude of the change. In addition, for some traits

profit is a non-linear function of the trait level

(Dekkers et al., 1995). Performing a large change

for such traits might influence economic values.

Moreover, in practice some traits are restricted to

zero change by using a restricted index. However,

restricting changes in traits to zero change by using a

restricted index reduces selection response com-

pared to when the trait is given an appropriate

economic value derived from using a profit equation

(Gibson & Kennedy, 1990).

In principle, economic values should be derived by

dividing change in profit by genetic change. In earlier

studies, profit was divided by genetic change and the

number of cows before a genetic change (e.g.,

Groen, 1989b). In this study however, profit was

divided by the phenotypic realized changed (see

Equation 4). One of the reasons for dividing by

phenotypic change rather than genetic change was

because the equation used to derive economic values

differed from earlier studies. In this study, correction

500

550

600

650

700

750

800

850

900

950

1000

662 693 725 756 788 819 851 882 914 945

Body weight (kg/cow)

,)woc/g

k( thgie

w ydoB

01/)raey-woc/g

k( kli

M

Pweight Gweight Pmilk

Figure 2. Level of phenotypic milk yield (kg/cow-year), phenotypic weight (kg/cow-year) and genetic body weight (kg per mature third

parity cow) of cows with different changes of the genetic level of body weight.

Stochastic model for traits in dairy cattle 27

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 14: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

Table VII. Revenues and costs (t/farm per year) for the basic situation and marginal costs and revenues after a change in level of trait by 20% and economic values (EV), standard deviation (SD)

and coefficient of variation (CV) for production and functional traits.

Trait Basic ECM

Conception

rate

D. f. c. to first

heat Stillbirth Mastitis

Retained

placenta Milk fever Laminitis Body weight

Involuntary

culling

Revenues

Milk (ECM) 202858 29622 496 594 146 85 71 273 68 �/4820 1739

Culled cows 14461 �/1720 �/1416 �/326 34 �/5 �/4 �/6 30 2213 �/1628

Bull calves 29369 �/743 112 395 620 58 13 �/31 33 �/69 �/426

Heifers 11056 1849 2610 921 778 �/18 27 163 95 �/52 2013

Total revenues 257744 28999 1809 1591 1574 115 104 385 218 �/2729 1702

Costs

Feed 55502 5381 83 134 23 15 15 55 8 2192 399

Mastitis 2961 �/115 61 70 30 �/593 26 23 21 31 23

Retained placenta 578 �/8 23 6 �/1 �/5 �/116 �/13 �/13 0 16

Milk fever 1195 �/94 114 36 �/16 �/9 �/13 �/252 �/13 30 123

Laminitis 581 �/2 23 6 0 3 �/3 �/1 �/115 �/5 1

Temp. 313 0 0 0 0 0 0 0 0 0 1

Milking 7646 4 7 3 4 1 1 3 1 0 22

Insemination 1967 7 �/286 53 6 2 2 4 3 0 �/5

Calving 393 �/9 1 5 0 0 0 �/1 0 0 �/6

Replacement 37271 �/484 358 493 803 �/82 �/13 �/35 80 �/2 �/978

Other 14287 8 15 6 8 2 2 5 1 1 48

Bull production 22460 �/568 86 302 474 44 10 �/24 25 �/53 �/326

Total var. costs 145153 4120 485 1114 1330 �/624 �/89 �/236 �/1 2194 �/682

Fixed cow costs 33336 20 34 14 18 4 4 12 3 2 112

Fixed farm costs 33647 0 0 0 0 0 0 0 0 0 0

Total farm costs 212136 4139 519 1129 1348 �/619 �/85 �/224 2 2196 �/570

Profit 45608 24859 1290 463 226 734 190 609 216 �/4925 2271

Change in level of trait 1238.25 9.1 �/7 �/1.96 �/0.07 �/0.03 �/0.02 �/0.01 91.92 �/4.90

Change in cow-years 0.04 0.07 0.03 0.04 0.01 0.01 0.02 0 0 0.24

EV1 0.28 1.98 �/0.94 �/1.63 �/162.5 �/79.0 �/402.1 �/210.2 �/0.76 �/6.66

SD 0.02 4.5 6.0 21 682 1249 2012 2936 0.55 6.4

CV 0.07 2.3 6.4 12.9 4.2 15.8 5.0 14.0 0.72 1.0

1t/trait unit per cow-year.

28

H.M

.N

ielsenet

al.

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 15: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

number of cow-years before and after a genetic

change could differ due to the stochastic element

of the model (number of cow-years was multiplied

by trait level). Phenotypic levels of traits were

defined per cow-year, whereas definition of genetic

trait levels differed between traits but was for most

traits defined for an average third parity cow.

Secondly, genetic changes for the disease traits

were performed at the underlying level (basic disease

risk). By using phenotypic outcome of the model in

the equation, the economic values were expressed on

the preferred scale (per disease incidence). In

principle, phenotypic change should correspond

closely to genetic change when using 1000 replicates.

Dividing by phenotypic change rather than genetic

change would then not influence economic values.

However, dividing by phenotypic change rather than

genetic change might have influenced the economic

value for milk yield because the increase in the

phenotypic level of milk yield per cow-year did not

correspond to the increase at the genetic level. This

was due to indirect effects where feed intake capacity

of the cow limited expression of genetic milk yield

capacity of the cow, and to herd effects such as age

structure of cows because genetic level was expressed

as milk yield per third parity cow per year, whereas

phenotypic level of milk in the herd was expressed

per cow-year.

Indirect effects influenced the economic values for

both body weight and milk production where the

economic weight for body weight was more negative

than expected due to feed intake as an indirect effect.

Increasing milk production decreased the average

weight of culled cows, because feed intake capacity

was limiting the phenotypic expression of both

weight and milk yield potential of the cow. This

effect was, however, counterbalanced with the ex-

tended herd life of the cows, which decreased

replacement costs. Prior studies estimating eco-

nomic values for dairy cattle production traits used

models where feed intake was determined only by

nutrient requirements for maintenance and produc-

tion. However, in the model presented by Vargas et

al. (2002), performance of the cow was dependent

on the genetic potential, feed availability and feed

intake capacity. This was also the case for the model

presented in the current study, which had an

influence especially on the economic value for body

weight.

The indirect effects applied in this study were the

main reason for the high economic value for milk

fever. The model included underlying risks for

getting a disease, where some of those risks were

the risk of having another disease. For example, milk

fever was specified as a major risk factor for mastitis

and retained placenta. Indirect effects of diseases

were also increased risk of dead cows. Therefore,

reducing the incidence of milk fever at the same time

increased average herd life through reduced invo-

luntary culling and death of cows.

In the model, the farmer’s replacement policy was

determined by a threshold for low yield and a

threshold for maximum number of days open. These

voluntary cullings were not changed after changing

the genetic level of milk yield and conception rate in

the herd. A change in genetic level of those traits

might result in a change in the optimum voluntary

replacement rate. Theoretically, the production

system must be optimal in the basic situation and

should be re-optimized after a genetic change.

However, Dekkers (1991) found less than 1% bias

in economic values for milk yield, involuntary culling

and conception rate when the policy optimum for

the level of trait before a genetic change was also

used after a genetic change. Failing to change culling

thresholds might be increasingly important in this

study due to the indirect effects and the large change

performed when deriving the economic values. For

example, increased genetic level of milk yield influ-

enced herd life of cows because voluntary culling of

0.20 0.24 0.28 0.32 0.36 0.40 0.44

(a)

0.15 0.19 0.23 0.27 0.31 0.35

(b)

Figure 3. Distribution of incidence of mastitis in the herd expressed as incidence per cow-year for the basic situation (a) and after a genetic

change (b). The mean level of incidence of mastitis per cow-year was 0.329/0.03 in the basic situation and 0.259/0.02 after the genetic

change.

Stochastic model for traits in dairy cattle 29

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 16: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

cows was kept at the same level before and after a

genetic change.

Economic values

Among the disease traits milk fever had the highest

absolute economic value. This was as expected due

to the high cost of an incidence of milk fever. The

economic value for body weight estimated in this

study was �/0.76 t/kg per cow-year. In the review by

Koenen et al. (2000), economic values for body

weight were found in the range of �/1.28 to 0.02 t/

kg per cow per year. However, studies differ with

respect to methods used, prices and restrictions on

the production system applied. In most studies the

economic values were based only on marginal feed

costs and marginal return from beef production. In

this study an increase in the mature weight of the

cow was associated with decreased milk production,

because feed intake of the cow was limiting the

expression of milk yield capacity and growth. This

was the main reason why the economic value for

body weight found in this study was at the upper end

of the range of economic values of live weight

reviewed by Koenen et al. (2000).

The level of economic value of conception rate

found in this study (1.98 t/% per cow-year) is in the

range of economic values reported by others. Boi-

chard (1990) found economic values of conception

rate from 1.14 to 2.14 $/cow per year, whereas the

economic value of conception rate estimated by

Vargas et al. (2002) was 2.42 $/cow per year. The

economic value for milk was 0.28 t/kg per cow-year.

This result is in agreement with other studies using a

fixed number of cows as the base of evaluation

(Groen, 1989b; Bekman & van Arendonk, 1993).

The economic value of involuntary culling esti-

mated in this study was high (�/6.66 t/% involun-

tary culling per cow-year) compared to other studies

(Dekkers, 1991; Rogers et al., 1998) and higher than

expected, because some involuntary culling was

accounted for via the diseases. The economic value

was mainly due to reduced replacement costs and

increased milk yield at the farm due to more cows in

later parities, but might also be due to a more

efficient use of the cow barn. With a low rate of

involuntary culling, the farmer will rarely experience

the fact that a cow is involuntarily culled without

having a replacement heifer available. This was

confirmed by the relatively large change in the

number of cows when the rate of involuntary culling

was reduced. The relatively high economic value of

involuntary culling might also partly be explained by

the way that involuntary culling was modelled.

Voluntary culling only covers non-satisfactory repro-

duction relative to the milk production; all other

culling (except culling due to diseases) is performed

via the percentage of involuntary culling in the herd.

Standard deviations of economic values

Estimated standard deviations of economic values

were high, especially for functional traits. No other

studies were found where standard deviations of

economic values were estimated. However, due to

higher standard deviations among cow performance

of functional traits and especially disease traits

(Table IV), standard deviations of their economic

values were expected to be higher than for produc-

tion traits. There was no variation among mature

body weight of the cows in the herd. The variation in

actual expressed body weight for the cow was only

from the indirect effects of milk yield capacity and

disease occurrence, which might have underesti-

mated the standard deviation of the economic value

for body weight.

The standard deviations of the economic values

were from replicates of a farm and express variation

in realized profit due to expression of cow traits

before and after a genetic change. Variances of profit,

number of cow-years at the farm, and expression of

cow traits and their covariances before and after a

genetic change were used to derive the standard

deviations of the economic values and consequently

influenced the size of the standard deviations (Equa-

tion 4). Variation in realized profit of the farm was

from variation in expression of traits due to the

stochastic component of performance of the cows in

the model. On average over the 1000 replicates,

genetic change was realized at the farm (e.g., mastitis

incidence in the herd decreased) and consequently

profit increased, as reflected by the low standard

errors (see Table V). In some replicates genetic

change was not realized and profit decreased.

The variation in profit was from both variation in

expression of the traits for which the economic

values were derived and variation in other traits

due to the indirect effects between traits. In addition,

profit of the farm was a function of all traits in the

stochastic model. Even in replicates with realization

of a genetic change of a given trait, variation in non-

related traits contributed to variation in profit (e.g.,

mastitis incidence decreased after a genetic change

in mastitis but milk yield per cow was lower than in

the basic situation). In a stochastic model as used in

this study, random variation among the traits implies

that a genetic change is not always realized at

phenotypic level and reflected in farm profit. There-

fore, realized farm profit after genetic improvement

in a given trait is uncertain, which correspond to

what farmers are exposed to in the real world.

Results from this paper show that this type of

30 H.M. Nielsen et al.

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 17: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

uncertainty differs between traits. From a farmer

perspective, variation in profit due to expression of

traits is highly relevant since farm profit is a result of

the realized genetic improvement in the herd. There-

fore, variation in economic values due to expression

of traits could impact relative selection influence,

because the farmer would tend to put less selection

emphasis on traits, which are highly uncertain, such

as the functional traits. As discussed by Dekkers and

Gibson (1998), farmer acceptance of the breeding

goal is important for optimum impact of selection. If

economic values of traits differ with respect to

uncertainty, farmers may be reluctant to adopt the

breeding goal, which influences the relative selection

influence on traits.

Derived economic values were based on para-

meters from an average Danish dairy farm and

consequently also average herd size. This approach

ignores differences between farm sizes when estimat-

ing standard deviations of the economic values.

However, differences in farm size could be ac-

counted for by estimating standard deviations of

the economic values for different groups of farm

sizes. Ideally, uncertainty of economic values should

include both variation between farms, variation in

prices of inputs and outputs and variation in

expression of traits, which all contribute to variation

in profit and economic values. Both uncertainty of

economic values due to variation in prices and

uncertainty due to variation in expression of traits

are expected to differ between traits. Consequently

both types of uncertainties could possibly impact

farmers’ selection emphasis on a given trait. Amer

and Hofer (1994) suggested including the sampling

distribution of the economic values from variation in

prices when deriving selection index weights. They

showed a reduction in selection response when

economic values were uncertain. However, the

potential use of the standard deviations from this

study to account for reduction in selection response

due to uncertain economic values as suggested by

Amer and Hofer (1994) is beyond the scope of this

paper.

Acknowledgements

The authors thank Lars Gjøl Christensen, Jørn

Pedersen and Jan Tind Sørensen for valuable

discussions about the simulation model and com-

ments on the manuscript. H.M. Nielsen acknowl-

edges the EU for the Marie Curie Fellowship while

staying at Wageningen University, The Netherlands.

References

Amer, P.R., & Hofer, A. (1994). Optimum bias in selection index

parameters estimated with uncertainty. Journal of Animal

Breeding and Genetics , 111 , 89�101.

Bekman, H., & van Arendonk, J.A.M. (1993). Derivation of

economic values for veal, beef and milk production traits

using profit equations. Livestock Production Science , 34 , 35�56.

Bundgaard, E., & Hoej, S. (2000). Direct Access to the Cattle

Database with Livestock Registration. Annual Report 1999,

National Committee on Danish Cattle Husbandry, Aarhus,

Denmark.

Boichard, D. (1990). Estimation of the economic value of

conception rate in dairy cattle. Livestock Production Science ,

24 , 187�204.

Bryant, J., Lopez-Villabos, N., Holmes, C., & Pryce, J. (2005).

Simulation modelling of dairy cattle performance based on

knowledge of genotype, environment and genotype by

environment interactions: current status. Agricultural Sys-

tems , 86 , 121�143.

Daka. (2001). Afgifter for indsamling og forarbejdning af selvdøde

dyr: Available: http://www.daka.dk/dk_index.html. (In Dan-

ish.)

Danish Agricultural Advisory Centre. (2002). Handbog i kvæg-

hold. Landbrugsforlaget. 199 pp. (In Danish.)

Danish Cattle Federation (2003). Arsstatistik-Avl 2002�2003.

Report no. 106. 334 pp. (In Danish.)

Danish Dairy Board. (2001). Afregningsmodel for mælk. Avail-

able: http://www.lr.dk. (In Danish.)

Danish Meat Board. (2002). Statistisk tabelmateriale for 2001.

Okse- og kalvekød. Available: http://www.meatboard.dk.

Dekkers, J.C.M. (1991). Estimation of economic values for dairy

cattle breeding goals: bias due to suboptimal management

policies. Livestock Production Science , 29 , 131�149.

Dekkers, J.C.M., Birke, P.V., & Gibson, J.P. (1995). Optimum

linear selection indexes for multiple generation objectives

with non-linear profit functions. Animal Science , 61 , 165�175.

Direktoratet for FødevareErhverv. (2000). Vejledning om slagte-

præmie for kvæg og handyrpræmie. Report. Ministry of

Food, Agriculture and Fisheries. 16 pp. (In Danish.)

EEC. (1981). Community scale for the classification of carcasses

of adult bovine animals. Reg. ECC, No 1208/81, Reg. EEC,

No 2930/81.

Gibson, J.P., & Kennedy, B.W. (1990). The use of constrained

selection indexes in breeding for economic merit. Theoretical

and Applied Genetics , 80 , 801�805.

Goddard, M.E. (1983). Selection indices for non-linear profit

functions. Theoretical and Applied Genetics , 64 , 339�344.

Groehn, J.A., Kaneene, J.B., & Foster, D. (1992). Risk factors

associated with lameness in lactating dairy cattle in Michi-

gan. Preventive Veterinary Medicine , 14 , 77�85.

Groen, A.F. (1989a). Cattle breeding goals and production

circumstances. Ph.D. thesis, Animal Breeding and Genetics

Group, Wageningen University. 167 pp.

Groen, A.F. (1989b). Economic values in cattle breeding. I.

Influences of production circumstances in situations without

output limitations. Livestock Production Science , 22 , 1�16.

Groen, A.F., Steine, T., Colleau, J.-J., Pedersen, J., Pribyl, J., &

Reinsch, N. (1997). Economic values in dairy cattle breed-

ing, with special reference to functional traits. Report of an

EAAP-working group. Livestock Production Science , 49 , 1�21.

Groen, A.F., & Korver, S. (1989). The economic value of feed

intake capacity of dairy cows. Livestock Production Science ,

22 , 269�281.

Stochastic model for traits in dairy cattle 31

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014

Page 18: A stochastic model for the derivation of economic values and their standard deviations for production and functional traits in dairy cattle

Hazel, L.N. (1943). The genetic basis for constructing selection

indexes. Genetics , 28 , 476�490.

Koenen, E.P.C., Berentsen, P.B.N., & Groen, A.F. (2000).

Economic values of live weight and feed-intake capacity of

dairy cattle under Dutch production circumstances. Livestock

Production Science , 66 , 235�250.

LK. (2001). BudgetKalkuler 2001. The Danish Agricultural

Advisory Centre. 63 pp. (In Danish.)

Meijering, A. (1986). Dystocia in dairy cattle breeding. Ph.D.

thesis, Animal Breeding and Genetics Group, Wageningen

University, Wageningen. 179 pp.

Olesen, I., Groen, A.F., & Gjerde, B. (2000). Definition of animal

breeding goals for sustainable production systems. Journal of

Animal Science , 78 , 570�582.

Østergaard, S., Sørensen, J.T., & Kristensen, A.R. (2000). A

stochastic model simulating the feeding-health-production

complex in a dairy herd. Journal of Dairy Science , 83 , 721�733.

Østergaard, S., Sørensen, J.T., & Houe, H. (2003). A stochastic

model simulating milk fever in a dairy herd. Preventive

Veterinary Medicine , 58 , 125�143.

Principles of Danish Cattle Breeding. (2001). Recording of

production data, recording of breeding data, calculation

methods and breeding values. The Danish Agricultural

Advisory Centre. Fourth Edition 53 pp.

Quinton, V.M., Wilton, J.W., Robinson, J.A., & Mathur, P.M.

(2006). Economic weights for sow productivity traits in

nucleus pig populations. Livestock Science , 99 , 69�77.

Rauw, W.M., Kanis, E., Noordhuizen, E.N., & Grommers, F.J.

(1998). Undesirable side-effects of selection for high produc-

tion efficiency in farm animals: a review. Livestock Production

Science , 56 , 13�33.

Rogers, G.W., van Arendonk, J.A.M., & McDaniel, B.T. (1988).

Influence of involuntary culling on optimum culling rates

and annualized net revenue. Journal of Dairy Science , 71 ,

3463�3469.

SJFI. (2000). Economics of Agricultural enterprises 1998/99.

Report 83, serie B from the Danish Institute of Agri-

cultural and Fisheries Economics, Copenhagen. 148 pp.

(In Danish.)

Sørensen, J.T., Kristensen, E.S., & Thysen, I. (1992). A stochastic

model simulating the dairy herd on a PC. Agricultural

Systems , 39 , 177�200.

Vargas, B., Groen, A.F., Herrero, M., & van Arendonk, J.A.M.

(2002). Economic values for production and functional traits

in Holstein cattle of Costa Rica. Livestock Production Science ,

75 , 101�116.

Wood, P.H.P. (1967). Algebraic model of the lactation curve in

cattle. Nature , 216 , 164�165.

32 H.M. Nielsen et al.

Dow

nloa

ded

by [

Nor

thea

ster

n U

nive

rsity

] at

09:

46 3

0 O

ctob

er 2

014