You are on page 1of 157

ENERGY NUTRITION IN RUMINANTS

ENERGY NUTRITION IN
RUMINANTS

E. R.0RSKOV
Applied Research Department, The Rowett Research Institute, Aberdeen, UK

and

M.RYLE
39, Hunshelf Park, Stocksbridge, Sheffield, UK

ELSEVIER APPLIED SCIENCE


LONDON and NEW YORK
ELSEVIER SCIENCE PUBLISHERS LTD
Crown House, Linton Road, Barking, Essex IGIl 8JU, England

Sole Distributor in the USA and Canada


ELSEVIER SCIENCE PUBLISHING CO., INC.
655 Avenue of the Americas, New York, NY 10010, USA

WITH 43 TABLES AND 9 ILLUSTRATIONS


ISBN·13:978·94·010·6823·9 e·ISBN·13:978·94·009·0751·5
DOl: 10.1007/978·94·009·0751·5
© 1990 ELSEVIER SCIENCE PUBLISHERS LTD
Softcover reprint of the hardcover 1st edition 1990
British Library Cataloguing in Publication Data

0rskov, E. R. (Egil Robert), 1934-


Energy nutrition in ruminants.
1. Livestock: Ruminants. Feeding & nutrition
I.. Title II. Ryle, M. (Margaret)
636.2084
ISBN 97&-94-010-6823-9

Library of Congress Cataloging in Publication Data


0rskov, E. R.
Energy nutrition in ruminants/E. R. 0rskov and M. Ryle.
p. em.
Includes bibliographical references.
ISBN 97&-94-010-6823-9
1. Ruminants-Nutrition. 2. Ruminants-Feeding and feeds.
3. Bioenergetics. I. Ryle, Margaret. II. Title.
SF95.077 1990
636.2'08~c20
89·71460
CIP

No responsibility is assumed by the Publisher for any injury and/or damage to persons or
property as a matter of products liability, negligence or otherwise, or from any use or
operation of any methods, products, instructions or ideas contained in the material herein.

Special regulations for readers in the USA


This publication has been registered With the Copyright Clearance Center Inc. (Ccq, Salem,
Massachusetts. Information can be obtained from the CCC about conditions under which
photocopies of parts of this publication may be made in the USA. All other copyright
questions, including photocopying outside the USA, should be referred to the publisher.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval
system, or transmitted in any form or by any means, electronic, mechanical, photocopying,
recording, or otherwise, without the prior written permission of the publisher.
PREFACE

This book is intended to be a companion volume to 'Protein


Nutrition in Ruminants' (1982, Academic Press), which emphasized
both the role of proteins and new systems for their evaluation. Here
the focus is on energy-yielding nutrients and problems involved in
evaluating them. Nonetheless in both volumes there is explicit
recognition of the interdependence of energy and protein nutrition.
I have not attempted to review comprehensively all the literature
relating to ruminant energy nutrition and must apologize to
colleagues whose work is not fully reported. Where possible tables
and figures are taken from the studies of our group at the Rowett
Research Institute since, if for no other reason, I am most familiar
with these data.
I have first considered the nutrition of the newborn and have
stressed the role of behaviour 'in determining whether nutrients
enter or bypass the rumen. The development of the rumen, the
principles of anaerobic fermentation and the roles of various
. species of rumen bacteria, protozoa and fungi in relation to
different substrates, are summarized. This is followed by accounts
of the factors affecting the utilization of different substrates and the
v
vi Preface

absorption and metabolism of the end-products of fermentation


and digestion, together with estimates of digestive capacity in
various segments of the gut. The ruminant's requirements for
energy-yielding nutrients is considered in relation to the per-
formance of various activities and to environmental conditions,
particular attention being paid to the requirement for glucose
precursors.
The final chapters are largely concerned with the control of feed
intake, especially of roughages. It is argued that current feed
evaluation systems are inadequate, because they do not take into
account how much of each feed an animal actually consumes. This
deficiency is particularly important where the diet consists mainly
of roughages, ·as is typical in developing countries. A provisional
new system is proposed, based on information regarding (i) the
fraction of soluble material in the feed, (ii) the proportion of the
insoluble material which is potentially fermentable, and (iii) the
rate at which insoluble but fermentable material is degraded. Data
are presented which show that this system can predict both feed
intake by test animals and their consequent growth rate.
I hope this book will be of value to undergraduate and
postgraduate students specializing in nutrition, to nutritionists in
commercial and government-supported organizations, and to those
who teach in the fascinating field of ruminant nutrition. While the
volume is no definitive treatise, I trust that it will prove both
stimulating and challenging.
Dr. Margaret Ryle wrote Chapter 2 and edited the whole book.
We are both most grateful to Dr. R.N.B. Kay for suggesting many
constructive amendments, and to Mrs. R. Milton, who helped
greatly in the preparation of the manuscript.

E.R.0rskov
CONTENTS

Preface . v

1. Nutrient supply to the newborn ruminant 1

2. Energy nutrition of rumen micro-organisms 10

3. Manipulation of rumen fermentation and associative


effects 28

4. Host animal control of microbial fermentation and host


animal digestion 43

5. Absorption of nutrients 52

6. Energy metabolism of the host animal 63

vii
viii Contents

7. Utilization of the energy of absorbed nutrients 84

8. Feed quality and feed intake .102

9. Feed evaluation, past and present · 122

10. Towards future feed evaluation systems · 133

Index · 145
CHAPTER 1

NUTRIENT SUPPLY TO THE


NEWBORN RUMINANT

I. Introduction
II. Physiology of stomach development
A. The oesophageal groove reflex
B. Development of enzyme competence
i. Proteolytic enzymes
ii. Lipase
iii. Carbohydrases
III. Milk replacers
IV. Utilization of energy from milk
V. Transition to fermentable feeds

I. INTRODUCTION

The moment the navel cord is broken, during passage from the
uterus to the external environment, major changes occur in the
energy nutrition of the young. In fact, considering the complexity
Gfthe changes, it is remarkable that mortality is so low. One major
change is in the route by which nutrients are supplied. Before birth,
2 Energy Nutrition in Ruminants

glucose, amino acids etc. are delivered directly via the placenta.
Afterwards they must first be consumed as milk and then absorbed
from the small intestine. The high rate of success with which this
major change occurs is due to the close matching of the composition
of the colostrum, the newborn's first energy source, with intestinal
cell permeability to large molecules. At the same time, large
amounts of globulins obtained from the colostrum help to protect
against prevalent pathogens while the immune system is developing.
Another very important change requiring immediate adaptation
concerns the control of body temperature which, in itself, may
require considerable energy. This adaptation occurs rapidly after
birth, more efficiently than in humans and much more rapidly than
in newly hatched chicks, which cannot control their body
temperature at all for some hours and are better regarded as
walking eggs! For ruminants that are likely to be born in a cold
environment, other provisions also help to ensure survival. These
include a subcutaneous layer of brown fat which in effect acts like
an electric blanket. Brown fat accumulates only towards the end of
gestation so, for example, lambs that are born even one week
premature are much more sensitive to adverse temperatures than
those carried to term.

II. PHYSIOLOGY OF STOMACH DEVELOPMENT

There are many excellent books describing the development of the


ruminant stomach. Moreover, many relevant articles have been
published in the regular symposia on ruminant physiology which
are held every five years, as well as elsewhere. Therefore only a
summary will be given here.

IIA. The Oesophageal Groove Reflex

The rumen is poorly developed at birth, in contrast to the


abomasum. Indeed, the volumes of the two are similar, although in
the mature animal that of the rumen is at least ten times that of the
abomasum. Milk and milk replacers are channelled directly to the
Nutrient supply to the newborn ruminant 3

abomasum via the so-called oesophageal groove. This is essentially


a gutter-shaped structure that, when closed, prevents the fluid from
entering the rumen. There have been many studies on the
oesophageal groove's reflex closure. Initially it was thought that
some substances in the milk stimulated closure when present in the
mouth or pharynx (Wester, 1926). However, Watson (1944) pointed
out that the behaviour pattern played a crucial role, together with
tactile stimulation from the teat which provides the milk. 0rskov et
al. (1970) carried out several studies on this aspect and produced
convincing evidence regarding the nature of the reflex closure. They
showed that it occurred in a manner similar to the conditioned
reflexes described by Pavlov (1927). Even when the receptors in the
mouth and pharynx were bypassed by injecting liquid directly into
the oesophagus, provided that the lamb was shown the bottle or
bucket from which it was normally fed, the fluid entered the
abomasum. In addition the lamb displayed the classic juvenile
excitement associated with suckling, including head-butting, tail-
shaking etc. On the other hand if the lamb was not shown the
feeding bottle or bucket, the liquid passed into the rumen and there
was no excited behaviour. If regular feeding from bottle or bucket
was continued, the reflex closure could be maintained well into
adult life, regardless of the feeding regime otherwise practised. The
use of this mechanism to bypass the rumen with suspensions of
protein etc. has been discussed in more detail by 0rskov (1982).
Although the conditioned nature of the oesophageal groove
reflex has many advantages for the animal, it also has disadvantages
when young ruminants are weaned early and must receive milk or
milk substitutes from unfamiliar containers. In order to imprint the
new feeding method it is generally necessary to wean the young
from its dam within 2~8 h of birth. Later weaning, especially to
bucket feeding, is difficult. With some breeds of cattle it is virtually
impossible to train the calves to drink from buckets and, in general,
training to suck from artificial teats is easier. Unsuccessful
imprinting is revealed by the absence of juvenile excitement. In such
animals a large proportion of the milk enters the premature rumen
and undergoes lactic fermentation there, leading to problems of
acidosis and partial destruction of the protein.
4 Energy Nutrition in Ruminants

lIB. Development of Enzyme Competence

i. Proteolytic Enzymes
The action of the rennin and acid secreted into the abomasum
results in clotting of the milk and delays the passage of protein and
fat to the small intestine. Pepsin production is generally relatively
low in young ruminants. It increases with age and apparently also
with the presence of proteins other than casein (Garnot et al.,
1974). Replacement of milk protein with others of vegetable or
microbial origin results in lower digestibility. Thus estimates of the
apparent digestibility of soya protein range from 70 to 90%
(Walker & Kirk, 1975; Nistan et al., 1971; Raven & Robinson,
1959), while those for casein are normally around 95%. The most
successful substitute for casein in young ruminants appears to be
fish protein hydrolysate, with an apparent digestibility of about
92% (Toullec, 1974; Soliman et al., 1977). Rennin, of course, delays
only the passage of casein from the abomasum, not that of proteins
from other sources. It would therefore appear that a milk substitute
in which all or a large part of the proteins are of vegetable or fish
origin may have to be given at more frequent intervals than one in
which casein is the protein source.

ii. Lipase
Milk fat presents no digestive problem. Lipase is present both in the
saliva and in the pancreatic juice (Ternouth et al., 1971). Although
the fat content varies widely between sheep and cows, as well as
between different breeds of cows, these differences have never to the
authors' knowledge led to problems of digestion of the butterfat by
the offspring. It may be relevant that the casein clot slows down the
passage of lipids to the small intestine, so reducing the risk of
exceeding the lipolytic capacity of the pancreatic lipase or the
capacity for absorption.
Milk fat can generally be replaced by fats from many sources, of
both animal (lard, tallow) and vegetable (coconut, palm kernel)
origin, provided they are homogenized and emulsified. Nevertheless
when Walker & Kirk (1975) compared a range of vegetable fats
Nutrient supply to the newborn ruminant 5

with butterfat, they found that the digestibility of the latter was
highest.

iii. Carbohydrases
The ability of the young to digest lactose is not surprising. In fact,
the lactose content of the food can be varied widely. Penning (1975)
tested lactose to fat ratios of 35:29, 45:20 and 55: 11 and found little
difference in the performance of lambs, although they tended to
grow faster on the high-fat diet. However, when the capacity for
digesting lactose is exceeded, or when the digestion is impaired, it
will be fermented in the large intestine and thi~ will lead to scouring.
Both Glimp (1972) and Molenat & Theriez (1972) found that the
maximum level of lactose efficiently utilized by lambs was 42% of
the dry matter.
While one would expect early development oflactase activity, the
early capacity to digest maltose is more surprising. Maltase activity
increases rapidly after birth and in lambs is already very high by the
second week (Walker, 1959).
The ability to digest small amounts of starch also develops early
and also increases rapidly soon after birth. Ternouth et al. (1971)
found a six-fold increase in calves between the first and third week
of life. Thivend et al. (1979), using artificially reared lambs fitted
with ileal cannulae, found that their ability to digest starch
was quite substantial. When the proportion of starch in the dry
matter of an artificial milk replacer was increased progressively
from 19.6 to 35.7%, only at the highest level did substantial
quantities reach the large intestine (Table 1.1). The post-ruminal
digestion of raw starch remains slow and limited even in mature
ruminants, although the capacity to digest gelled and partially
hydrolysed starch appears to be much greater (Mayes & 0rskov,
1974).
There is general agreement that sucrase activity is not present in
the small intestine of either immature or mature ruminants
(Siddons, 1968; 0rskov et al., 1972). Addition of sucrose to milk
almost invariably leads to scouring since it provides a readily
available substrate for fermentation in the large intestine.
6 Energy Nutrition in Ruminants

TABLE 1.1
Composition of Milk Replacer and Digestibility of Partially Hydrolysed Maize
Starch (Protamyl) in the Small and Large Intestines of Lambs

Diet 1 2 3

Protamyl (g/kg) 196 279 357


Casein (g/kg) 373 355 339
Butterfat (g/kg) 328 272 218
Intake of starch (g/d) 49 70 100
Digestion of starch before ileum (%) 95.2 91.0 83.6
Digestion of starch in large intestine (%) 4.5 8.6 15.7

From Thivend et at. (1979).

III. MILK REPLACERS

Soliman et at. (1977) prepared a milk replacer consisting of fish


hydrolysate, emulsified lard plus coconut oil and partially
hydrolysed starch. They fed this to lambs 8 times in every 24 h, to
see whether frequent feeding of this milk substitute would permit
normal growth. Table 1.2 summarizes the results, which show that
it is indeed possible to replace all milk components, providing that
the technical facilities for processing substitute materials are
available. However, the management of feeding based on whole
milk or, better still, on milk that is suckled from the cow or ewe, is
much easier. Moreover, with natural suckling, growth rates are
generally higher and mortality lower than when similar quantities
of milk are supplied from buckets (Paredes et at., 1981).

IV. UTILIZATION OF ENERGY FROM MILK

In a large-scale comparative slaughter trial with lambs, Walker &


Jagusch (1967) showed that the efficiency with which the metabo-
lizable energy of milk was utilized was 71 %, when the level of
feeding was above maintenance. As will be seen in Chapter 7, the
maximum efficiency for volatile fatty acids (VF As) is generally
about 60%. It is of course to be expected that the energy consumed
in milk will be metabolized more efficiently than the energy in the
Nutrient supply to the newborn ruminant 7

TABLE 1.2
Effect of Replacing all Milk Constituents by Non-milk Derivatives

Fat source: Butter Butter Lard+ Lard+


coconut coconut
Carbohydrate source: Lactose Protamyla Protamyl Protamyl
Protein source: Casein Casein Casein Fish protein
Live weight gain (gjd) 177 183 154 170
Food conversion ratio b 1.02 1.06 1.08 1.00
Dry matter digestibility (%) 98.8 96.4 96.4 95.4
Digestibility of protein (%) 96.4 94.9 94.4 93.6
Digestibility of lipids (%) 98.9 98.3 98.2 97.7
Digestibility of starch (%) 99.8 99.8 99.9

From Soliman et al. (1977).


aprotamyl is hydrolysed maize starch.
bkg dry matter jkg live weight gain.

older, functioning ruminant's diet, since with milk there are no


losses of methane or fermentation heat and the constituents can be
metabolized directly.

V. TRANSITION TO FERMENTABLE FEEDS

Young ruminants usually begin to eat solid food 2-3 weeks after
birth. At the same time they begin to acquire the typical rumen flora
and fauna. Although most rumen bacteria function as strict
anaerobes, they nevertheless occur in the external environment.
Many rumen organisms, in particular the protozoa, are acquired
from other animals via the saliva. Others are introduced in feed
contaminated with faeces. Even animals reared in total isolation
develop a relatively normal rumen flora, though they remain free of
protozoa.
Rumen development depends mainly on the stimulus of volatile
fatty acid formed during the fermentation of ingested carbohydrate.
Stimulation by bulky, fibrous feeds is not important (Warner &
Flatt, 1965). In other words, the more that readily fermentable
carbohydrate is consumed, the more rapidly does the rumen
develop, until it can supply sufficient nutrients to meet the animal's
need. Normally the rumen reaches its mature proportions relative
8 Energy Nutrition in Ruminants

to the abomasum by 2-6 months. In general, this development


occurs more rapidly in sheep than in cattle (see Van Soest, 1982).
However, the age at which these proportions are attained can be
affected by the level of consumption of liquid milk, either from the
dam or from artificial rearing systems. The greater the quantity of
milk consumed, the less is the young ruminant inclined to eat solid
food and, consequently, the slower is rumen development. In order
to encourage rapid development the amount of milk offered should
be reduced after 3-4 weeks of age. Weaning is possible before the
rumen has reached its mature proportions, provided that the solid
feed offered is rapidly degradable and can therefore yield sufficient
VF A. Lambs weaned at 25, 35 or 45 days, and given concentrate,
recovered their weaning weights after 23, 15 and 5 days respectively.
Moreover, nipid growth was subsequently achieved despite
rumen:abomasum proportions that were still immature (0rskov et
al., 1973).

REFERENCES

Garnot, P., Valles, E., Thapon, J-L., Toullec, R. & Tomassone, R-D. (1974)
Influence of dietary proteins on rennin and pepsin content of pre-ruminant calf
veal J. Dairy Res. 41, 19-23
Glimp, H.A. (1972) Effect of diet composition on performance of lambs reared
from birth on milk replacer J. Anim. Sci. 34, 1085-1088
Mayes, R.W. & 0rskov, E.R. (1974) The utilization of gelled maize starch in the
small intestine of sheep Brit. J. Nutr. 32, 143-153
Molenat, G. & Theriez, C-M. (1972) Artificial milk feeding of lambs. 2. Effect of
fat content of milk replacers Ann. Zootech. 21, 385-399
Nistan, Z., Volcani, R., Gordin, S. & Hasdai, A. (1971) Growth and nutrient
utilization by calves fed milk replacers containing milk or soybean protein
concentrate heated to various degrees J. Dairy Sci. 54, 1294-1299
0rskov, E.R. (1982) Protein Nutrition in Ruminants Academic Press, London
0rskov, E.R., Benzie, D. & Kay, RN.B. (1970) The effect of feeding procedure
on closure of the oesophageal groove in sheep Brit. J. Nutr. 24, 785-795
0rskov, E.R., Mayes, R.W. & Mann, S.D. (1972) Postruminal digestion of
sucrose in sheep Brit. J. Nutr. 28, 425--432
0rskov, E.R., Fraser, e. & Gill, J.e. (1973) A note on the effect of time of weaning
and weight at slaughter on feed utilization of intensively fed lambs Anim. Prod.
16,311-314
Paredes, L., Capriles, M., Parra, R. & Marguer, N. (1981) The performance of
calves reared by restricted suckling with matter of high milk production
potential Trop. Anim. Prod. 6, 368-372
Nutrient supply to the newborn ruminant 9
Pavlov, I.P. (1927) Conditioned Reflexes Trans!. by O.V. Aarep. Oxford University
Press, Oxford
Penning, I.M. (1975) Nutrition of the Liquid-fed Lamb Ph.D. Thesis, University of
Reading
Raven, A.M. & Robinson, K.L. (1959) Studies on the nutrition of the young calf.
2. The nutritive value of unhydrogenated palm oil, unhydrogenated palm-
kernel oil and butter fat, as additions to a milk diet Brit. J. Nutr. 13, 178-190
Siddons, R.C. (1968) Carbohydrase activities in the bovine digestive tract
Biochem. J. 108, 839-844
Soliman, H.S., 0rskov, E.R. & Smart, R.1. (1977) Milk replacers based on non-
milk constituents for lambs Proc. Nutr. Soc. 36, 52A
Ternouth, J.H., Siddons, R.C. & Toothill, J. (1971) Pancreatic secretion in the
milk fed calf Proc. Nutr. Soc. 30, 89A
Thivend, P., Clark, C.F.S., 0rskov, E.R. & Kay, R.N.B. (1979) Digestion of
partially hydrolyzed starch in milk replacers by the young lamb Ann. Rech. Vet.
10, 422--424
Toullec, R. (1974) The·use of soluble fish protein concentrates in milk replacers
Proc. 3rd European Symposium on the Use of Fish Meal in Animal Feeding
pp 68-72 Internal. Assocn. of Fish Meal Manufacturers, Peterborough
Van Soest, P.J. (1982) Nutritional Ecology of the Ruminant 0 & B Books,
Corvallis, OR
Walker, D.M. (1959) The development of the digestive system of the young
anima!. 3. Carbohydrase enzyme development in the young lamb J. agric. Sci.,
Camb. 53, 374-380
Walker, D.M. & Jagusch, K. T. (1967) Influence of ambient temperature on energy
utilization for milk production in the cow. In Blaxter, L.L., Kielanowski, J. &
Thorbek, G. (Eds) Proc. 4th Symp. on Energy Metabolism of Farm Animals
pp 187-193 E.A.A.P. Publication No. 12 Oriel Press, Newcastle upon Tyne
Walker, D.M. & Kirk, R.D. (1975) The utilization by pre-ruminant lambs of milk
replacers containing isolated soya bean protein Austr. J. agric. Res. 26,
1025-1035
Warner, R.G. & Flatt, W.P. (1965) Anatomical development of the ruminant
stomach. In Dougherty, R.W., Allen, R.S., Burroughs, W., Jacobson, N.L. &
McGiliiard, A.D. (Eds) Physiology of Digestion in the Ruminant pp 24-38
Butterworth, London
Watson R.H. (1944) Studies on Deglutition in Sheep Bulletin No. 180 Council for
Scientific & Industrial Research, Melbourne
Wester, I. (1926) Die Physiologie und Pathologie der Vormagen beim Rinde Berlin
CHAPTER 2

ENERGY NUTRITION OF RUMEN


MICRO-ORGANISMS

I. Introduction
II. Rumen bacteria
A. Cellulolytic bacteria
B. Amylolytic bacteria
C. Soluble carbohydrates
D. Other energy sources
III. Rumen fungi
IV. Rumen protozoa
A. General characteristics
B. Cellulolytic ciliates
C. Amylolytic ciliates
D. Soluble carbohydrates
E. Other energy sources
F. Products of ciliate fermentation
V. Some interactions between bacteria and ciliates
VI. pH and feed-related effects
VII. Implications of anaerobiosis for the energy nutrition
of rumen micro-organisms
VIII. Conclusions
10
Energy nutrition of rumen micro-organisms 11

I. INTRODUCTION

The rumen has often been compared to a fermentation vat.


However, although its temperature and anaerobic state are stable
and its pH fluctuations are limited, nothing else is reminiscent of
industrial conditions. Both the chemical composition and the
physical structure of the ingested substrates can vary widely and
rapidly. Instead of one or two types of micro-organism providing
the enzymes for fermentation, there may be significant numbers of
many species belonging to each of three main groups - bacteria,
fungi and protozoa. The relative numbers of the different species
vary with the composition and structure of the feed and their
mutual interactions' are highly complex. Yet although the propor-
tions of the different end-products also vary, they nevertheless
normally consist principally of carbon dioxide, methane and the
three volatile fatty acids (VFAs) acetic acid, propionic acid and
butyric acid, together with ammonia, traces of some other VFAs
and sometimes lactic acid.
This chapter outlines the characteristics of each main group of
micro-organisms in relation to energy nutrition, and considers
some aspects of their interactions with one another and with the
host animal and its diet. Further general information is available in
the following recent reviews and research papers: Coleman (1979,
1980, 1985a, 1986a, 1986b), Wolin (1979), Demeyer (1981),
Williams (1982, 1986), Orpin (1984), Williams et al. (1984),
Williams & Strachan (1984), Williams & Coleman (1985),
Mountfort (1987), Ryle & 0rskov (1987), Veira (1986).

II. RUMEN BACTERIA

The rumen contents usually include about 1010-1011 bacteria/ml,


up to 75% being associated with food particles. The overall density
does not vary greatly with the host's diet but the relative numbers
of the different species present may be greatly affected by the
substrates available for fermentation. Probably several hundred
12 Energy Nutrition in Ruminants

species of bacteria occur in the rumen but only about 30 are present
at densities of at least 10 7 /ml in one or more species of ruminant.
The most important are those which ferment cellulose.

IIA. Cellulolytic Bacteria

Vertebrates lack enzymes capable of breaking down cellulose and


hemicellulose, which are complex fi-linked polymers of glucose and
pentoses. Cellulolytic micro-organisms confer a special advantage
upon ruminants (and a few other vertebrate groups) by making
these abundant sources of energy available to them. There are
several species of cellulolytic rumen bacteria, the most important
being Ruminococcusalbus, Ruminococcusflavefasciens and Bactero-
ides succinogenes. These attach themselves firmly to plant frag-
ments and secrete enzymes which disrupt the backbone of the
cellulose molecule, remove the side-chains from it and further
hydrolyse the resulting oligosaccharide fragments. Cellobiose, a
disaccharide, is one of the soluble products. It is used both by the
bacteria which produce it and by other micro-organisms which
cannot themselves break down cellulose. If cellobiose is not further
hydrolysed it can inhibit the attachment of B. succinogenes to
cellulose and suppress the cellulase activity of R. albus. Glucose,
another product of cellulolysis, can also inhibit the activity of some
enzymes. In general, one finds many other species in association
with cellulolytic bacteria, probably depending for energy on the
soluble products which they release. The three species listed above
also break down hemicellulose - another important structural
polysaccharide of plant tissues - but do not appear to be able to
use the products. These are degraded further by different species.
Pure cultures of various bacteria produce many substances not
ordinarily detectable in tl).e rumen. For example, B. succinogenes
produces succinate from cellulose, but in vivo this is converted to
propionate by another species. Since propionate is one of the major
products of fermentation from which the host ruminant can derive
glucose, the linked action of the two species is important. In
contrast, much of the acetate produced by R. flavefasciens may
contribute directly to the VFA pool.
Energy nutrition of rumen micro-organisms 13

Cellulolytic bacteria colonize the epidermal surfaces of plant


fragments within 5 min of their entering the rumen. Providing that
sufficient ammonia is present they multiply rapidly. Dietary urea
can provide ammonia and so promote efficient utilization of fibrous
roughage. If, however, the rumen pH falls below about 6.0, both
the multiplication of cellulolytic bacteria and cellulolysis slow
down. Below 5.6 they cease altogether; so the presence 9f dietary
components which promote acidity will inhibit the digestion of
roughage (see Chapter 3).

lIB. Amylolytic Bacteria

Some species of bacteria are equipped with a wide range of enzymes


and can ferment a number of substrates, though they may prefer to
use one rather than another. Others, with a smaller repertoire, can
ferment only a few substrates. Most of the bacteria which hydrolyse
starch are unable to use cellulose. However, more species possess
amylolytic enzymes than cellulolytic ones. Thus, although rumi-
nants seldom have access to large quantities of starch either in the
wild or in many rural areas of the world, they can nevertheless
readily adapt to this source of energy. Thanks to these rumen
bacteria, which begin the process of converting starch to VFAs,
cattle can be fattened on cereal grains.
Provided sufficient ammonia is available, the bacteria multiply
rapidly. The rate of growth may affect the end-products. For
instance, Streptococcus bovis produces acetate and ethanol when
multiplying slowly, but when multiplying rapidly it produces
lactate. This species is more acid-tolerant than most rumen micro-
organisms. It is not normally very abundant but tends to become
dominant if the accumulation of lactic acid outruns the buffering
system. Severe acidosis (PH 5.0) can occur when the diet is changed
abruptly from cellulose to starch or sucrose. Presumably when this
happens the populations of micro-organisms required to convert
lactate to acetate and propionate are either not present at a
sufficient density, or multiply less rapidly than the population of
S. bovis and are then increasingly inhibited as the pH falls.
14 Energy Nutrition in Ruminants

Meanwhile, the population of S. bovis itself may double every 20


min.
Some species of bacteria can ferment dextrins and maltose as well
as starch but cannot use most mono- and disaccharides. Others,
unable to initiate the fermentation of starch itself, may associate
with those primarily responsible and perhaps contribute indirectly
by removing soluble end-products of fermentation.

nc. Soluble Carbohydrates


Soluble oligosaccharides and sugars are the end-products of the
extra-cellular, degradation of cellulose, starch and other carbo-
hydrate polymers. After absorption their intra-cellular hydrolysis
provides bacteria (and other ruminant micro-organisms) with the
energy required for generating adenosine triphosphate (ATP). ATP
provides energy for the synthesis of most biologically significant
substances, and hence is essential for the growth and multiplication
of most organisms. It is used up as rapidly as it is formed. Since
sugars are readily absorbed from the surrounding fluid, optimal
concentrations permit very high bacterial growth rates.
During the generation of ATP both the reduced nucleotide,
NADH, and reduced ferredoxin are produced. These are recycled
by an oxidation step which involves the release of molecular
hydrogen. However, except at very low partial pressures, hydrogen
inhibits the oxidizing enzyme and brings the ATP-generating
system to a halt. The energy from saccharide hydrolysis is then
diverted into other synthetic pathways, including the production of
two VF As, propionate and butyrate. This diversion can be limited
in the rumen by species of methanogenic bacteria. These gain
energy for biosynthesis by combining hydrogen with carbon dioxide
to form methane. The methane accumulates in the rumen in
gaseous form until expelled during eructation, together with excess
carbon dioxide. The activity of methanogenic bacteria can produce
more than 200 litres of methane per day in a 500 kg cow.
Conditions which restrict methanogenesis favour the production of
propionate and increase the propionate/acetate ratio in the end-
Energy nutrition of rumen micro-organisms 15

products offermentation. Conditions which favour methanogenesis


promote A TP synthesis and microbial growth, as well as acetate
production, at the cost of losing up to 15% of the feed energy as
methane.

lID. Other Energy Sources

Lipids
Only low levels of lipids (up to about 7%) are acceptable in the
ruminant diet. At higher levels the free fatty acids released when the
lipids are hydrolysted inhibit fibre digestion, possibly by coating
food particles and preventing bacterial attachment. If higher levels
of fat are to be used they must be protected from hydrolysis in the
rumen. The glycerol derived from hydrolysed fats is converted to
VF As by some rumen bacteria. Others hydrogenate unsaturated
fatty acids to saturated ones.

Proteins
Most soluble dietary proteins are rapidly and completely hydro-
lysed in the rumen. Part of the amino acid yield can be incorporated
directly into the protein of a few bacterial species as well as some
protozoa. The remainder serves as an energy source, being further
degraded to VF As and ammonia. Several species of bacteria are
involved in the production of ammonia, much of which is recycled
during bacterial growth and multiplication. The excess is absorbed
into the host's portal blood and converted to urea in the liver.

Volatile Fatty Acids


Bacteria which utilize VF As are present in the rumen, but they
multiply so slowly that the rate of the ruminal outflow ensures a
very low population density. Thus, of the final products of rumen
fermentation (a) the VFAs are metabolized only by micro-
organisms which do not thrive in the rumen environment; (b) the
carbon dioxide is produced faster than it is used and constitutes
16 Energy Nutrition in Ruminants

(depending to some extent on the type offermentation) about 65%


of the eructated gas, and (c) the methane - which constitutes the
remaining 35% - cannot supply any micro-organisms with energy
in the anaerobic conditions of the rumen.

III. RUMEN FUNGI

The considerable contribution of phycomycete fungi to ruminal


fermentation has been recognized only quite recently, yet they may
constitute up to 8% of the intra-ruminal biomass. In the flagellated
stage of the life cycle they colonize damaged regions of plant tissue
within 2 h of ingestion, moving towards them in response to
diffusing soluble material. The vegetative stage then develops
rapidly. By 22 h up to 30% of the larger food particles may be
infected and invasion already considerable, with fungal rhizoids
penetrating the plant cell walls. Pure cultures of some rumen fungi
ferment cellulose in vitro, yielding mainly acetate, lactate, carbon
dioxide and hydrogen. Cellulase released into the medium accounts
for much of the degradation. Hemicellulose and xylan are also
fermented, as well as starch and sugars. In studies so far undertaken
lignin does not appear to be susceptible to attack by rumen fungi.
Nevertheless, this group of micro-organisms may be particularly
important for the degradation of those plant structural materials
which predominate in coarse roughage.

IV. RUMEN PROTOZOA

IV A. General Characteristics

A few flagellates can be found in the rumen, but the vast majority
of the protozoa there belong to the class Ciliata. These ciliates
include some of the most complex unicellular organisms that are
known. Their total biomass in the rumen is generally similar to that
Energy nutrition of rumen micro-organisms 17

of the bacteria but, on suitable feeds, may be more than three times
as great or almost zero. However, since they are all much larger
than bacteria, their density in the rumen contents is normally only
about 10 5-106 /m!.
The different species range from approximately 25 to 250 11m in
length. They are grouped into seventeen genera within the sub-class
of Entodiniomorphida and two genera within the sub-class of
Holotricha. These two main groups differ markedly both in their
morphology and in their metabolism. The particular combination
of ciliate species present can vary with the host species, the
geographical location, the diet and even the individual animal
within the flock or herd.
The intervals between consecutive ciliate cell divisions are longer
than those of bacteria, being of the order of 0.5-2 days. This slower
rate of mUltiplication might theoretically lead to their elimination
in the outflow of fluid from the rumen. However, many remain
enmeshed among - and often attached to - the larger food
fragments which are excluded from the outflow. In addition,
between meals the Holotrichs congregate in large numbers on the
walls of the reticulum (Abe et ai., 1981). Partly in consequence of
this sequestration, the fraction of the ciliate population which
passes from the rumen per day is substantially less than the
equivalent fraction of the bacterial population. However, a large
proportion (possibly up to two-thirds) appear to be lysed within the
rumen (Ffoulkes & Leng, 1984).
The ciliates differ from the bacteria in several other important
aspects: (1) They are highly motile, so, in spite of much smaller
numbers, they invade newly ingested food almost as rapidly as the
bacteria. Orpin (1985) showed that in vitro they move towards food
particles and quickly attach to them, leaving few ciliates swimming
freely in the fluid. (2) They are able to store surplus carbohydrate
in the form of a characteristic insoluble polymer, amylopectin. (3)
They are more easily destroyed by acid conditions than are many
bacteria, the Holotrichs being most sensitive and some small
Entodiniomorphs least so. (4) Like other animals they cannot
synthesize amino acids from simple compounds of nitrogen. They
depend mainly on bacteria, which they engulf and digest, using the
18 Energy Nutrition in Ruminants

released amino acids to build their own proteins. Depending on the


species of both prey and predator, individual ciliates have been
observed to engulf anything from 200 to 105 bacteria/h. Overall,
about I % of the bacteria present in the rumen may be engulfed by
ciliates every minute. Some large ciliates also obtain amino acids by
engulfing their smaller relations. (5) Unlike the bacteria, ciliates are
not essential to rumen fermentation, though they can contribute to
its efficiency.

IVB. Cellulolytic Ciliates

A few species. of the genus Epidinium are-involved in the physiCal


disruption of plant tissues. They secrete enzymes which promote
the separation of cells, the breakdown of cell walls and the
fragmentation of plant material. Together with other large
Entodiniomorphs they rapidly engulf quite large fragments and
digest them intracellularly.
Although the Holotrichs have only a low capacity for fermenting
cellulose, considerable cellulase activity has been demonstrated in a
number of Entodiniomorph species. In suitable conditions more
than half the cellulase activity in the rumen is associated with the
ciliate population, together with enzymes that depolymerize
hemicellulose, pectin and other structural carbohydrates. More-
over, free protozoal cellulases are absorbed onto cellulose. Thus a
substantial proportion of ruminal fibre degradation may be because
of enzymes released when ciliates die, perhaps as a result of
exposure to oxygen during rumination or to hypotonic rumen fluid
following drinking (Coleman, 1985b). It is therefore not surprising
that the rate of fibre digestion in the rumen is often reduced when
ciliates are eliminated, though subsequent bacterial fermentation in
the caecum and colon may largely compensate the host animal.
Different species of ciliates probably collaborate in the degra-
dation of cellulose and other substrates. Eadie (1962) described two
distinct stable associations of Entodiniomorph species. The Type A
association is characterized in particular by the presence of one
species and the Type B by another, various additional species
Energy nutrition of rumen micro-organisms 19

occurring in both associations. Recently Coleman (1986b) reported


that the ratio of cellulase to amylase activity was substantially
greater in a Type B than in a Type A ciliate population.

IVe. Amylolytic Ciliates

Total ciliate numbers increase rapidly when starch is plentiful. All


Entodiniomorphs use it, engulfing the grains avidly, fermenting
them intracellularly and storing much of the surplus carbohydrate
as granules of amylopectin. Some also release sugars into the
surrounding fluid. Certain species, which cannot utilize cellulose,
largely depend on starch as their source of energy. In contrast, one
of the two Holotrich genera cannot use stardi at all and the other
ferments it to only a limited extent.

IVD. Soluble Carbohydrates

Most Entodiniomorphs probably absorb and use soluble carbo-


hydrates present in the surrounding fluid, but the Holotrichs
depend on them. These ciliates move quickly towards sources of
diffusing soluble material and rapidly absorb sugars from the
rumen fluid, storing the surplus as amylopectin. Coleman (1979)
estimated that up to one-third of the sugars ingested by the host
could be converted to amylopectin. The soluble carbohydrates
fermented by Holotrichs include glucose, fructose, sucrose and
cellobiose. When these sugars are plentiful, total ciliate numbers
can be very high, due to the multiplication of Holotrichs rather
than of Entodiniomorphs.

IVE. Other Energy Sources

Lipids
Ciliates are normally responsible for about 30-40% of the lipolysis
which occurs in the rumen. Entodiniomorphs ingest oil droplets
and both they and Holotrichs take up long-chain fatty acids. They
hydrogenate unsaturated fatty acids and so increase the rumen
content of saturated ones. In sheep, the plasma level of the latter is
20 Energy Nutrition in Ruminants

reduced by defaunation (i.e. by eliminating the protozoa). About


75% of microbial lipid is normally associated with the ciliates, so
their contribution to the host's supply may at times be significant.

Proteins
Ciliates are probably not very important in relation to the
hydrolysis of dietary protein. Nugent & Mangan (1981) estimated
that they were responsible for only 10% ofleafprotein degradation.
Their specific proteolytic activity can be as low as one-tenth that of
the bacteria (Brock et at., 1982). On the other hand, part of the
large bacterial amino acid pool which they consume is degraded,
rather than reassembled into protozoal protein. In consequence of
this, plus the slow outflow of intact ciliates from the rumen, the
total amount of microbial protein available to the host can in
certain circumstances be appreciably increased by defaunation.

IVF. Products of Ciliate Fermentation

Rumen ciliates cultured in vitro produce substantial amounts of


acetate, butyrate, lactate, hydrogen and carbon dioxide, but little
propionate. Their overall contribution to rumen fermentation has
been examined by comparisons between normal host animals and
defaunated ones. In the absence of ciliates both the proportion of
propionate and the total amount ofVFA produced are often - but
not always - greater, while the proportions of both butyrate and
acetate may decline. More lactate may be produced, but less
methane. However, the effects of defaunation are inconsistent and
likely to vary with the dominant species of protozoa originally
present. Thus Holotrichs produce much lactate whereas Entodinio-
morphs produce mainly butyrate and acetate. Moreover, defauna-
tion leads not only to a rapid increase in the total biomass of
bacteria and fungi, but also to changes in the relative importance of
different species. It is therefore difficult to establish whether
changes in the end-products of fermentation are due to the absence
of ciliates or to secondary shifts in the microbial population.
Defaunated animals on high-roughage diets tend to be pot-
bellied, the reduced rumen fermentation of cellulose and other
Energy nutrition of rumen micro-organisms 21

structural carbohydrates being associated with a greater volume of


rumen contents and with more prolonged caecal fermentation.
Faecal excretion of dry matter is also sometimes increased.
Although defaunation may increase the proportion of microbial
protein available to the host, on high-roughage diets it may reduce
the overall rate of fermentation and, with it, both the total protein
and the total VFA leaving the rumen.

V. SOME INTERACTIONS BETWEEN BACTERIA AND


CILIATES

Several interactions between different rumen micro-organisms,


both competitive and collaborative, have already been mentioned.
The most striking is the dependence of the ciliates on bacteria for
amino acids. Defaunation can lead to a threefold increase in
bacterial numbers. On the other hand, the bacteria themselves
extensively use both the protein and the amylopectin of lysed
ciliates (Cottle et al., 1984).
Some engulfed bacteria can function symbiotically within their
ciliate predators. Others, remaining independent, also interact
metabolically with ciliates. For example, methanogenic bacteria
attach themselves to several Entodiniomorph species when bacterial
hydrogen is in short supply. Indeed, ciliate-generated hydrogen is
probably a major source of rumen methane. The end-products of
fungal cellulolysis are similarly processed by methanogenic bac-
teria. Some ciliates absorb traces of oxygen from the rumen fluid,
so helping to maintain an environment tolerable to stricter
anaerobes, including many bacteria.
Attention has been drawn to the interactions of various bacterial
species where cellulose is being degraded. Ciliates, in particular
Holotrichs, which congregate at points from which soluble
carbohydrates are diffusing, also probably contribute to bacterial
polysaccharide degradation by removing some of the end-products.
Both in vitro and in vivo the digestibility of cellulose is greatest if
both cellulolytic bacteria and ciliates are present. Yoder et al.
(1966) found that in vitro a mixed population of ciliates, free from
22 Energy Nutrition in Ruminants

bacteria, digested 7% of a standard quantity of cellulose during a


given period. Within the same period a comparable sample of
mixed rumen bacteria digested 40%. When both ciliates and
bacteria were present in optimal proportions they acted syner-
gistically and digested more than 60%. If, however, sugars are
plentiful in the fluid, Holotrich ciliates do not congregate on and
attach themselves to fibrous material. So, for example, over-
abundant molasses added to a high-roughage diet might reduce
cellulose digestion.
With high-concentrate diets the most significant way in which the
ciliates affect bacterial function is by stabilizing the environment.
Their rapid sequestration of much of the starch and soluble
carbohydrate ,that enter the rumen restricts the rate of lactic acid
formation and limits the fluctuations of rumen pH. This is
important because the survival of any population of micro-
organisms probably depends on the minimum pH to which it is
subjected rather than on the average. (There are, nevertheless,
limits to the amount of carbohydrate that ciliates can take up, so a
large excess can result in acid accumulation in spite of their initial
presence.) The continuing production ofVFAs between feeds, from
protozoal amylopectin, is a further advantage gained by the host
from the ciliates living within its rumen.

VI. pH AND FEED-RELATED EFFECTS

As noted earlier, different micro-organisms tolerate different pH


ranges. Streptococcus bovis withstands far more acid conditions
than the cellulolytic bacteria. Holotrichs are eliminated at pH levels
which are not lethal to small Entodiniomorphs. Thus two factors
jointly determine which micro-organisms predominate in the rumen
ecosystem: the substrates available and the pH of the rumen fluid.
Bicarbonate secreted in the saliva is the most important buffering
agent in the system. More copious secretion provides more effective
pH stabilization. Since salivation is stimulated by chewing and
rumination, the physical structure of the feed has a marked
influence on the survival of many rumen micro-organisms. Thus the
Energy nutrition of rumen micro-organisms 23

rumen fluid of cattle fed on whole sugar-cane has a high pH and


contains many Holotrichs, but that of cattle fed only on molasses
may contain none, despite the abundance of soluble carbohydrates.
In general, the physical as well as the chemical nature of the feed
influences both the efficiency of fermentation and the ratios
between its end-products via effects on the various micro-organisms
inhabiting the rumen.

VII. IMPLICATIONS OF ANAEROBIOSIS FOR THE


ENERGY NUTRITION OF RUMEN MICRO-ORGANISMS

Because of the strict anaerobic conditions in the rumen, only that


energy which can' be extracted without the use of oxygen is
available to the micro-organisms. The VFAs are, so to speak, their
excretion products. These are absorbed directly from the rumen
into the host animal's portal system (see Chapter 5) and serve as a
major source of energy while undergoing further, oxidative,
degradation. The fermentation of carbohydrate is by far the most
important source of energy for the rumen anaerobes. Although
they can also ferment protein, the amount of energy so obtained is
very small, probably less than 1% of the total protein energy. The
proportion of the energy in hydrolysed fat obtained through the
fermentation of glycerol is even smaller.
The stoichiometry of fermentation of hexose to the three main
VFAs is shown below:
Acetic acid
CaH120a + 2H 20 -+ 2CH aCOOH + 2C0 2+ 4H2
Propionic acid
CaH120a + 2H2 -+ 2CH a-CH 2-COOH + 2H 20
Butyric acid
C 6 H 120 a -+ CHa -CH 2-CH 2-COOH + 2C0 2+ 2H2
4H2 + CO 2 -+ CH 4 + 2H 20
The capture of carbohydrate energy into metabolites useful to the
host animal is much greater when propionic acid is produced than
when acetic acid is. The generation of hydrogen and hence of
methane - a waste product - is greatest when acetic acid is
24 Energy Nutrition in Ruminants

formed. While energy captured in the multiplying microbial cells is


not immediately available to the host, the biggest loss is in released
hydrogen and in the methane produced from it. As can be seen
above, the production of 2 moles of propionic acid from 1 mole of
hexose requires additional hydrogen. Propionic acid thus acts as a
hydrogen sink and reduces the loss of energy through methane
formation. These relationships were outlined by Hungate (1966).
The capture of hexose energy into VFAs, assuming complete
fermentation, can be calculated from the following expression:
E = 0.622pa+ 1.092pp+ 1.560pb x 100 (2.1)
pa+pp+2pb
where E is the efficiency of conversion -of hexose energy to VFA,
and pa, pp and pb are the proportions of acetic, propionic and
butyric acids respectively. The relationships are illustrated in Fig.
2.i. These relationships, described by 0rskov et al. (1968), also
showed that the heat of fermentation, i.e. the energy not accounted
for in end-products, was 6.4% regardless of the type of fer-
mentation. This implies that the overall energy available for micro-
organisms is the same regardless of the proportions of the different
VFAs that they produce, though the amount of A TP generated
could vary between species. If the quantities of the VF As produced
are known, methane production can be derived from the expression
M = 0.5a-0.25p+0.5b (2.2)
where a, p and b are moles of acetic, propionic and butyric acids
respectively. The above stoichiometric relationships indicate that
the capture of fermentation energy into energy that can be utilized
by the host animal can vary considerably. This variation will be
indicated by differences in methane production, not by differences
in the heat of fermentation.

VIII. CONCLUSIONS

This brief description of the rumen ecosystem can only suggest its
complexity. Many aspects which have been studied extensively are
not mentioned here. Many other aspects have not yet been
Energy nutrition of rumen micro-organisms 25

.....
.....
84 ......
..... .....
W .....
.....
.....
>. .....
.,
c: 80
u .....
:Q
....
.....,
76
c:
0

.,>
·iii
L
72
c:
0
U

68

64

50 70 75
Molar proportion of acetic acid % (pa)

Fig. 2.i. Effect of type of fermentation on efficiency of conversion of hexose.


%pb: Molar proportion of butyric acid; 100 - (pa + pb): molar proportion of pro-
pionic acid. The dotted line encloses the molar proportions encountered in prac-
tice. (From 0rskov et al. (1968).)

investigated. The interactions between the numerous different


micro-organisms are probably best compared to an immensely
elaborate mobile: move one component and all the rest will shift at
least a little. Yet it is the flexibility of this complex system which
makes possible the conversion of a wide range of substrates to the
VFAs on which the ruminant depends for energy.

REFERENCES

Abe, M., Iriki, T., Tobe, N. & Shibui, H. (1981) Sequestration of Holotrich
protozoa in the reticulo-rumen of cattle Appl. Env. Microbiol. 41, 758-765
26 Energy Nutrition in Ruminants
Brock, F.M., Forsberg, CW. & Buchanan-Smith, J.G. (1982) Proteolytic
activity of rumen microorganisms and effects of proteinase inhibitors Appl. Env.
Microbiol. 44, 561-569
Coleman, G.S. (1979) The role of rumen protozoa in the metabolism of ruminants
given tropical feeds Trop. Anim. Prod. 4, 199-213
Coleman, G.S. (1980) Rumen ciliate protozoa Adv. Parasitol. 18, 121-173
Coleman, G.S. (1985a) The cellulase content of 15 species of entodiniomorphid
protozoa, mixed bacteria and plant debris isolated from the ovine rumen J.
agric. Sci., Camb. 104, 349-360
Coleman, G.S. (l985b) Possible causes of the high death rate of ciliate protozoa
in the rumen J. agric. Sci., Camb. 105, 39-43
Coleman, G.S. (l986a) The distribution of carboxymethyl-cellulase between
fractions taken from the rumens of sheep containing no protozoa or one of five
different protozoal populations J. agric. Sci., Camb. 106, 121-127
Coleman, G.S. (l986b) The amylase activity of 14 species of entodiniomorphid
protozoa and the distribution of amylase in rumen digesta fractions of sheep
containing no protozoa or one of seven differel1t protozoal populations J. agric.
Sci., Camb. 107, 709-721
Cottle, D.J., Nolan, J.V. & Leng, R.A. (1984) Turnover of protozoa and bacteria
in the rumen of sheep Proc. Austr. Soc. Anim. Prod. 12, 138
Demeyer, D.1. (1981) Rumen microbes and digestion of plant cell walls Agric. &
Env. 6, 295-337
Eadie, J.M. (1962) Inter-relationships between certain rumen ciliate protozoa J.
gen. Microbiol. 29, 579-588
Ffoulkes, D. & Leng, R.A. (1984) Dynamics of protozoa in the rumen of cattle
Anim. Prod. in Austr. 15, 679
Hungate, R.E. (1966) The Rumen and its Microbes Academic Press, New York
Mountfort, D.O. (1987) The rumen anaerobic fungi FEMS Microbiol. Rev. 46,
401-408
Nugent, J.H. & Mangan, J.L. (1981) Characteristics of the rumen proteolysis of
Fraction I (l8S) leaf protein from lucerne (Medicago sativa L.) Brit. J. Nutr. 46,
39-58
0rskov, E.R., Flatt, W.P. & Moe, P.W. (1968) Fermentation balance approach to
estimate extent of fermentation and efficiency of volatile fatty acid formation in
ruminants J. Dairy Sci. 51, 1429-1435
Orpin, CG. (1984) The role of ciliate protozoa and fungi in the rumen digestion
of plant cell walls Anim. Feed Sci. Technol. 10, 121-144
Orpin, CG. (1985) Association of rumen ciliate populations with plant particles
in vitro. Microb. Ecol. II, 59-70
Ryle, M. & 0rskov, E.R. (1987) Rumen ciliates and tropical feeds World Anim.
Rev. 64, 21-30
Veira, D.M. (1986) The role of ciliate protozoa in nutrition of the ruminant J.
Anim. Sci. 63, 1547-1560
Williams, A.G. (1982) The metabolism and significance of ciliate protozoa in the
rumen ecosystem Rep. Hannah Res. [nst. pp. 93-110
Williams, A.G. (1986) Rumen holotrich ciliate protozoa Microbiol. Rev. 50,25-49
Williams, A.G. & Coleman, G.S. (1985) Hemicellulose-degrading enzymes in
ciliate protozoa Curro Microbiol. 12, 85-90
Energy nutrition oj rumen micro-organisms 27
Williams, A.G. & Strachan, N.H. (1984) The distribution of polysaccharide-
degrading enzymes in the bovine rumen digesta ecosystem Curro Microbio!. 10,
215-220
Williams, A.G., Withers, S.E. & Coleman, G.S. (1984) Glycoside hydrolases of
rumen bacteria and protozoa Curro Microbio!. 10,287-294
Wolin, M.J. (1979) The rumen fermentation: a model for microbial interactions in
anaerobic ecosystems Adv. Microb. Eco!. 3,49-77
Yoder, R.D., Trenkle, A. & Burroughs, W. (1966) Influence of rumen protozoa
and bacteria upon cellulose digestion in vitro. J. Anim. Sci. 25, 609-612
CHAPTER 3

MANIPULATION OF RUMEN
FERMENTATION AND ASSOCIATIVE
EFFECTS

I. Effect of substrate
A. Fibre
B. Starch
C. Sugars
D. Lipids
II. Effects of rumen environment
A. pH
B. Additives
C. Outflow rate
III. Associative effects
A. Negative associative effects
1. Rumen pH causing depression of the rate of
fermentation of fibre
11. Substrate competition
111. Depression of starch digestibility
B. Positive associative effects
IV. Conclusions
28
Rumen fermentation manipulation 29
The reasons for manipulating rumen fermentation will be discussed
in more detail in Chapters 6 and 7. As outlined in Chapter 2, the
type of rumen fermentation determines the extent to which
hydrogen, made available by anaerobic fermentation, is incor-
porated into compounds of use to the animal- in particular
propionic acid - or is lost in the reduction of carbon dioxide to
methane, with subsequent eructation. The type of fermentation can
also directly influence the host animal's metabolism by affecting its
endocrine status (see Chapter 7). Thus if dairy cows absorb too
much propionic acid from their rumens their blood insulin levels
rise, which can seriously affect both the production and the
composition of their milk. Similarly, in lambs, too high a
proportion of propionic acid, relative to the other fatty acids,
results in the deposition of undesirable branched-chain soft fat
(Duncan et ai., 1974). Therefore methods of manipulating rumen
fermentation, to optimize the ratios of different end-products of
fermentation, are of practical importance, particularly for lactating
animals.

I. EFFECT OF SUBSTRATE

lA. Fibre

It was established many years ago that the end-products of fibre


fermentation, i.e. of the activity of cellulolytic organisms, include a
high proportion of acetic acid (see reviews by Blaxter, 1962, and
0rskov, 1975). The poorer the quality of the roughage, the greater
- in general - is the proportion of acetic acid within the total
fatty acids. High acetic acid Pfoduction is associated with high
levels of available hydrogen and hence with high methane
production. However, soluble sugars modify the process. Better-
quality roughages contain more soluble sugars and these tend to
favour fermentation processes which yield greater proportions of
propionic and butyric acids (Hungate, 1966). Table 3.1, from
Armstrong (1964), summarizes the effects of some different
30 Energy Nutrition in Ruminants

TABLE 3.1
Effect of Maturity of Hay on Type of Fermentation, in Terms of Molar
Proportions of Volatile Fatty Acids

Type Proportions of individual VFAs


of hay (molar %)

Acetic Propionic Butyric


acid acid acid

Rye grass Young 62.7 22.7 14.6


Mature 68.6 20.8 10.6
Timothy Young 67.2 20.7 12.1
Mature 71.7 18.0 10.3

From Armstrong (1964).

roughage diets on rumen fermentation, mature hay giving a higher


proportion of acetic acid than hay made from young grass.

m. Starch
Unlike the cellulolytic microbes, which produce mainly acetic acid,
starch-fermenting organisms normally generate relatively more
propionic acid. However, the proportion produced is affected by
rumen pH. Table 3.2 illustrates the effects on VF A production of
various pH values, due to differences in the processing of barley,
wheat and maize (0rskov et al., 1974). The various pH values were
caused by differences in the secretion of saliva, which arose from
differences in the time required for chewing the feed and for
rumination. Similar effects were obtained by supplementing ground
cereal diets with bicarbop.ate (Mould & 0rskov, 1984). Ground
cereals fed to lambs can result in very high levels of propionic acid
and, as mentioned above, these give rise to abnormal fat
metabolism that is characterized by the production of large
quantities of odd-numbered and branched-chain fatty acids (see
Chapter 7). It should also be noted that, under exceptional
circumstances, where a. large population of rumen protozoa is
Rumen fermentation manipulation 31

TABLE 3.2
Effect of Processing of Different Cereals on Rumen pH and Volatile Fatty Acid
Concentrations in Rumen Fluid

Cereal Rumen Total VFA Proportions of individual VFAs


and pH (mequiv / litre) (molar %)
treatment
Acetic Propionic Butyric Higher
acid acid acid acids
Barley
LW 6.4 86 52.5 30.1 12.0 5.4
PO 5.4 102 45.0 45.3 7.0 2.7
Maize
LW 6.1 84 47.2 38.7 8.8 5.3
PO 5.2 90 41.3 4:n 10.0 5.6
Oats
LW 6.7 65 65.0 18.6 11.7 4.8
PO 6.1 73 53.2 37.5 5.4 3.9
Wheat
LW 5.9 78 52.3 32.2 8.6 6.9
PO 5.0 100 34.2 42.6 16.0 7.2
SE of 0.14 11 2.4 3.2 2.0
means
From 0rskov et al. (1974).
LW: Loose, whole.
PO: Pelleted, ground.
8 observations/mean.

maintained on a high-starch diet, butyric rather than propionic acid


predominates.

Ie. Sugars

The fermentation pattern induced by soluble sugars, including


molasses, is less predictable than that induced by fibre or even than
that induced by starch-based diets. Undoubtedly this is at least
partly due to the fact that many types of organism can utilize these
compounds. Even cellulolytic organisms can utilize glucose directly
(Hungate, 1966). Table 3.3, from 0rskov & Oltjen (1967), compares
32 Energy Nutrition in Ruminants

the types of fermentation which occur with different carbohydrates.


Soluble sugars generally give rise to high overall concentrations of
VFAs in the rumen, due to the rapid rate of fermentation.

ID. Lipids

Although the glycerol of lipids can provide an energy source for


microbes, the free fatty acids cannot be used in the anaerobic
conditions prevailing in the rumen. High concentrations of these
acids inhibit fibre digestion and consequently result in a smaller
proportion of acetic acid while, at the same time, the total amount
of substrate fermented is reduced. Therefore lipids should be ad<ied
to high-fibre diets with caution.

II. EFFECTS OF RUMEN ENVIRONMENT

IIA. pH

As illustrated above, the pH can alter the type of fermentation,


particularly of high-starch and high-soluble-sugar diets. Conse-
quently, the variability between animals in saliva production can
influence the type offermentation (see Chapter 4), as can conditions
which selectively inhibit the multiplication of some organisms. For
example, few types of protozoa are present below pH 6.0. Mould &
0rskov (1984) reduced rumen pH by infusing mineral acids (Table
3.4). When the pH was reduced below 6, intake decreased rapidly
because cellulose fermentation was abolished, although the pro-
portion of acetic acid present in the rumen actually increased
despite the low pH.

lIB. Additives

As with rumen pH, additives which selectively inhibit the


multiplication of some types of bacteria or protozoa will alter the
fermentation pattern. During the last decade much work has been
carried out with the aim of producing an additive that will increase
Rumen fermentation manipulation 33

TABLE 3.3
Effect of High Percentages of Different Carbohydrates on the Ruminal
Fermentation Patterns of Steers Fed on Purified Diets

Carbohydrate Rumen Total VFA Proportions of individual VFAs


pH (mequiv/litre) (molar %)

Acetic Propionic Butyric Higher


acid acid acid acids
Celluloses 6.9 87.1 73.7 18.3 4.8 3.2
Starch 6.7 76.3 60.4 24.7 10.4 4.5
Starch + 6.0 114.4 57.1 28.9 9.9 4.1
glucose
Sucrose 5.8 121.7 49.6 23.2 20.2 7.0
Glucose 5.7 102.6 38.0 22.3 25.8 13.9
SE of 0.2 9.3 2.8 2.9 1.0
means

From 0rskov & Oltjen (1967).

TABLE 3.4
Effect of Acidity, Controlled by Infusion of Mineral Acid (H 2S0 4 , H 3 P0 4 and
HCI), on the Type of Fermentation of a Grass-Hay Diet

Concentration of Rumen Proportions of individual VFAs


infusion pH (molar %)
(mM)
Acetic Propionic Butyric
acid acid acid

o 6.6 62.4 25.7 9.0


0.20 6.4 64.5 22.4 9.2
0.25 6.2 70.7 18.8 8.4
From Mould & 0rskov (1984).

the proportion of propionic acid relative to total volatile fatty


acids. The so-called ionophores; such as Monensin, slightly reduce
the proportions of acetic and butyric acids in the rumen, as well as
the production of methane, while increasing the proportion of
propionic acid (see Table 3.5). However, such compounds also act
l:!$ antibiotics, so it is uncertain whether their effects are due to
changes in the rumen fermentation pattern, in the microflora of the
post-ruminal tract, or indeed in the metabolism of the host animal.
34 Energy Nutrition in Ruminants

TABLE 3.5
Effect of Monensin on Molar Proportions of Volatile Fatty Acids In
Roughage-fed Cattle

Monensin Proportions of individual VFAs


(ppm) (molar %)

Acetic Propionic Butyric


acid acid acid
o 67.0 22.1 10.9
1 63.5 26.7 9.8
5 62.0 29.3 8.7

Adapted from Richardson et al. (1976).

IIC. Outflow Rate

There is little information regarding the effects on the rumen


fermentation pattern either of the rate of dilution of the rumen
contents or of the outflow rates of liquids and solids. The type of
nutrient absorbed by the host can, however, be affected by
promoting the escape of starch particles from rumen fermentation,
since their subsequent digestion yields glucose. Post-ruminal
digestion of starch can be enhanced by decreasing rumen
fermentation via decreased grain processing, particularly of maize.
In Table 3.6 rolled barley, flaked maize, ground maize, and cracked
maize are compared. When barley is rolled its starch ferments much
more rapidly than similarly processed maize starch and very little
escapes rumen fermentation. Cooked maize starch also ferments
rapidly, but a large proportion of raw starch from highly processed
maize escapes rumen fermentation.

III. ASSOCIATIVE EFFECTS

This term is generally used when a mixture or combination of feeds


has a nutritional value different from the sum of the values of its
components. For example, the apparent digestibility of a mixture
may be different from the sum of the apparent digestibilities of its
Rumen fermentation manipulation 35
TABLE 3.6
Effects of Different Cereals and Different Processing Methods on Percentage of
Starch Escaping Rumen Fermentation in Sheep

Treatment % escaping % of intake


rumen fermentation in faeces
Rolled barley 6.2 0.8
Flaked maize 5.4 0.7
Ground maize 12.1 0.5
Cracked maize 14.2 0.8
SE of means 3.2 2.0

From 0rskov et al. (1969).

constituents. (It must of course be assumed that neither the mixture


nor any individual constituent lacks nitrogen, sulphur or necessary
minerals). 'Non-additivity' or 'interaction' would be a more
appropriate term, but 'associative effect' is the generally accepted
one. The phenomenon was first described in some detail by Forbes
et al. (1931), since when there has been some controversy about its
importance. No feed evaluation system takes it into account: each
assumes that the feed value of a mixed diet is represented by the
sum of its individual components' values. However, an example of
the possible magnitude of an associative effect is shown in Table
3.7, from Mould et al. (1983a). When rolled barley was added to
ground hay the observed digestibility of dry matter was sub-
stantially less than what would be expected on the basis of simple
additivity. In this experiment, assuming that the depression related
to the hay only, its calculated digestibility was reduced by 37%.
Such a substantial change cannot be ignored. Other important
examples of associative effects are discussed below.

IlIA. Negative Associative Effects

i. Rumen pH Causing Depression of the Rate of Fermentation of


Fibre
The greatest single cause of negative associative effects is un-
doubtedly low rumen pH. Mould et al. (1983b) showed conclusively
that when rumen pH was less than about 6.2 fibre digestion was
36 Energy Nutrition in Ruminants
TABLE 3.7
Effect on Digestibility of Feeding Rolled Barley Together with Ground Hay

Barley Dry matter digestibility (%) Depression in


(% of diet) digestibility (%)
Observed Expected
Dry matter Hay
0 52.2 52.2
33 56.2 60.9 -7.7 -13.4
46 57.7 64.3 -10.2 -23.4
57 59.0 67.2 -12.2 -37.2
100 78.4 78.4
From Mould et al. (l983a).

TABLE 3.8
Effect of Type of Concentrate and Added Bicarbonate on rumen pH and 24 h
Degradability of the Insoluble Fraction of Hay

Rumen pH Degradability

-NaCO a + NaCO a -NaCO a + NaCO a


Control 6.6 46.7
Whole barley 6.0 6.7 26.2 32.0
Pelleted barley 6.2 6.9 27.6 37.6
Whole maize 5.9 6.8 13.8 24.2
Pelleted maize 6.1 6.6 19.3 27.1
SE of means 0.08 0.06 1.67 1.67
From Mould et al. (1983b).
Concentrate was 65 % of diet.

depressed (Table 3.8). The subsequent effect on digestibility


depended largely on particle size. With ground roughage the
depression was greater than with long roughage, presumably due to
greater saliva secretion with the latter. However, the intake of long
roughage was also reduced. Istasse et al. (1986) showed that the
extent to which fibre digestion was inhibited could be predicted
accurately from the extent of rumen pH depression below 6.0 and
the duration of this depression (cf. Chapter 2). In many feed
evaluation systems the digestibility of the diet is assessed at
maintenance levels of feeding, where rumen pH is rarely depressed
Rumen fermentation manipulation 37
below 6.0 and where, therefore, negative associative effects are not
recognized. This is probably one of the reasons for the fact that
most feed evaluation systems do not take them into account.
Both feeding regimes and feed processing methods can influence
the magnitude of associative effects. 0rskov & Fraser (1975)
observed that lambs ate more hay and digested it better when given
a supplement of whole barley rather than one of rolled barley, the
rumen pH being maintained at a higher level. Similarly, Table 3.9
shows how the degree of processing of a cereal supplement affected
the intake of long hay by cattle (0rskov et al., 1978). The more
extensive the processing, the greater was the depression of hay
intake. In addition to the degree of proQessing, the type of
concentrate can affect the extent to which associative effects occur,
again through rumen pH changes. Thus Fahmy et al. (1984) found
that when the diet included 70% of rolled barley, the digestibility
of ammonia-treated straw fell from 53 to 22%, yet with 70% of
molassed sugar-beet pulp it fell to only 40% (Table 3.10). Thus the
extent of negative associative effects is determined mainly by rumen
pH, and therefore by the type and level of concentrate and by the
degree of its processing.

ii. Substrate Competition


Mould et al. (1983b) noted in regard to digestibility that, while
most of the deviation from additivity could be attributed to rumen
pH, some of it could not. This they termed 'substrate competition'
or 'carbohydrate effects'. Such effects are particularly apparent
where the diet contains a large proportion of soluble sugars, such
as molasses. In this situation fibre digestion can be reduced in the
absence of any depression of rumen pH. 0rskov & Hovell (1978)
found that fibre digestion was less rapid when steers received a diet
based on whole-crop sugar-cane than when they received a hay diet
based on napier grass, despite similar rumen pH values. It is
possible that rapid utilization of micro-elements by fast-growing
..bacteria may deprive the more slowly growing cellulolytic organisms
of these nutrients. Alternatively, species of ciliate protozoa which
remove the sugar end-products of cellulolysis may satisfy their
38 Energy Nutrition in Ruminants
TABLE 3.9
Effect of Method of Processing Barley Grain on Hay Intake by Steers (Grain
Offered at Approx. 40 g/kgwo.7s/d)

Processing method Intake of hay Total intake


(g dry matter/kgWO· 75 ) (g dry matter/kgWO· 75 )

Whole 42.1 83.5


Alkali treated 43.0 80.4
Torrifieda 38.1 79.6
Crimped 35.1 76.6
Rolled 34.9 76.8
Ground 34.4 76.7
Ground and pelleted 30.5 70.9
SE of means 1.4 1.4
From 0rskov et al. (1978).
aHeated under infra-red source to induce cracking of seed coat.

TABLE 3.10
Effects of Sugar-beet Pulp and Rolled Barley on Dry Matter Digestibility of Diets
Based on Ammonia-treated Straw: Comparison of Observed Digestibility with
that Predicted on the Assumption of Additivity

Diet % dry matter digestibility Calculated


digestibility
Observed Predicted of straw (%)
Treated straw only 53.3 53.3 53.3
Treated straw + 70 % 70.5 74.5 40.3
sugar-beet pulp
Treated straw + 70 % 65.4 74.9 21.8
rolled barley
From Fahmy et al. (1984).

requirements elsewhere and so no longer associate with the


cellulolytic bacteria, thus retarding fibre degradation. It may be
possible to dilute the sugar. with high-quality fibre to eliminate the
inhibitory effect, or even to give rise to a positive associative effect.

iii. Depression of Starch Digestibility


While negative associative effects are usually due to a depression in
fibre degradation, they can - with some diets - be related to a
depression in the digestion of starch. For instance, whole or
Rumen fermentation manipulation 39
TABLE 3.11
Effects of Different Diets Fed to Sheep on Percent Degradation of Untreated
Straw, Following Incubation for Various Periods in Nylon Bags within the Rumen

Diet Rumen Rumen % degradation after


ammonia pH
(mg/litre) 8h 16h 24h 48 h
Untreated straw 268 6.9 19.3 29.8 16.5 45.3
Ammonia-treated straw 244 6.8 19.5 30.6 38.7 53.1
Hay 212 6.5 17.3 31.5 40.0 50.2
From Silva & 0rskov (1988).

cracked maize, fed alone, has a high digestibility in both cattle and
sheep, due to a low rate of outflow, although the starch of maize is
fermented much more slowly than that of barley (0rskov et al.,
1969). However, when fibrous roughage is added to such diets, the
passage rate of small particles increases and substantial quantities
of starch can be excreted in the faeces. Similar observations by
Nordin & Campling (1976) and by Joanning & Johnson (1979)
confirmed that the additivity may be negative when such feeds are
mixed or combined.

llIB. Positive Associative Effects

Although negative associative effects are more common, positive


associative effects also occur. So far, these have been observed
where the feed consists largely of poor-quality roughage. Silva &
0rskov (1988) provided an excellent illustration. They fed sheep on
straw, allliTIonia-treated straw or hay. Urea (and sulphur) was
added to the untreated straw and to the hay, to ensure that the
degradation rate was not limited by the rumen ammonia con-
centration. Subsequently, samples of the three feeds were incubated
(in nylon bags) in the rumen environment developed on each diet,
to see if their degradation rates were similar. The most important
observation was that untreated straw was fermented more rapidly
in the rumens of animals that had been fed on ammonia-treated
straw than in the rumens of animals fed on untreated straw (Table
40 Energy Nutrition in Ruminants

3.11). The authors' interpretation, that lack of readily available


substrates limited colonization in the rumens of animals fed
untreated straw, led to a search for supplements capable of
increasing the degradation rate of such diets. Table 3.12 shows how
sugar-beet pulp proved to be such a supplement, increasing both
digestibility and intake of untreated straw (Silva et aI., 1989).
Many Asian farmers have observed the same pheI}omenon: cattle
fed on rice straw are habitually given small quantities of greens to
increase straw intake.
It is clear from the above discussion that both negative and
positive associative effects occur, but they are difficult to quantify
in most systems of feed evaluation. Nylon bag incubations, in
fistulated animals provide a powerful tool for studying these effects.
These bags can be employed to determine optimum combinations
of feed, optimum feeding regimes and optimum degrees of
processing concentrates to give maximum utilization of mixed
feeds.

IV. CONCLUSIONS

Despite the importance of fermentation patterns, only generaliza-


tions can be made regarding the effects of either their deliberate or
their accidental manipulation. Due to the many contributory
factors it is extremely difficult to predict the effects of different diets.
In general, greater fermentation of cellulosic substrates results in
greater proportions of acetic acid. Conversely, greater fermentation
of starches results in greater proportions of propionic acid. The
type of fermentation resulting from diets high in sugars, including
molasses, is particularly difficult to predict. It is therefore extremely
unlikely that feeds can be described in terms of the amount and
types of VF As produced from them during rumen fermentation.
Negative and positive associative effects can be very important
when feeds are combined. Many of the former can be minimized by
attention to the feeding regime, to the extent to which cereal grains
are processed and to the types of concentrate suitable for different
combinations of feed.
Rumen fermentation manipulation 41
TABLE 3.12
Effects of Fish Meal (FM; 50 gjkg straw) and Sugar-beet Pulp (SBP; 150 gjkg
straw) on Intake and Digestibility of Straw Supplemented with Urea or
Treated with Ammonia

Treatment Supplement Straw Straw


intake digestibility
(gjd) (%)
Untreated + urea 414 49
Untreated + urea FM 480 56
Untreated + urea SBP 505 57
Untreated + urea FM and SBP 480 59
Ammonia treated 729 57
Ammonia treated FM 690 59
Ammonia treated SBP 717 59
Ammonia treated FM and SBP 658 64
SE of means 42.0 1.8
From Silva et al. (1989).

REFERENCES

Armstrong, D.G. (1964) Evaluation of artificially dried grass as a source of energy


for sheep J. agric. Sci., Camb. 62, 399-406
Blaxter, K.L. (1962) The Energy Metabolism of Ruminants Hutchinson, London
Duncan, W.R.H., 0rskov, E. R., Frazer, C. & Garton, G.A. (1974) Effect of
processing of dietary barley and of supplementary cobalt and cyanocobalamine
on the fatty acid composition of lamb triglyceride with special reference to
branched chain fatty acids Brit. J. Nutr. 32, 71-75
Fahmy, S.T.M., Lee, N. & 0rskov, E.R. (1984) Digestion and utilization of straw.
2. Effect of different supplements on the digestion of ammonia-treated straw
Anim. Prod. 38, 75-81
Forbes, E.B., Braman, W.W., Kriss, M. & Swift, RW. (1931) The metabolizable
energy and net energy values of corn meal when fed exclusively and in
combination with alfalfa hay J. agric. Res. 43, 1015-1026
Hungate, RE. (1966) The Rumen and its Microbes Academic Press, New York
Istasse, L., Smart, RI. & 0rskov, E.R (1986) Comparison between two methods
of feeding concentrate to sheep given a diet high or low in concentrate, with or
without buffering substances Anim. Feed Sci. Technol. 16, 37-49
Joanning, S.W. & Johnson, D.E. (1979) Nutrient digestibility depressions in corn
silage--{;orn grain mixtures fed to steers J. Anim. Sci. 49, Suppl.l, p 379
Mould, F.L. & 0rskov, E.R. (1984) Manipulation of rumen fluid pH and its
influence on cellulolysis in sacco, dry matter degradation and the rumen
microflora of sheep offered either hay or concentrate Anim. Feed Sci. Technol.
16, 1-14
Mould, F.L., 0rskov, E.R. & Gould, S.A. (l983a) Associative effects of mixed
42 Energy Nutrition in Ruminants
feeds. II. The effect of dietary addition of bicarbonate salts on the voluntary
intake and digestibility of diets containing various proportions of hay and
barley Anim. Feed Sci. Techno!. 10, 31-47
Mould, F.L., 0rskov, E.R. & Mann, S.O. (I983b) Associative effects of mixed
feeds. I. Effects of type and level of supplementation and the influence of rumen
fluid pH on cellulolysis in vivo and dry matter digestion of various roughages
Anim. Feed Sci. Technol. 10, 15-30
Nordin, M. & Campling, R.C. (1976) Effect of the amount and form of roughage
in the diet on digestibility of whole maize grain in cows and steers J. agric. Sci.,
Camb. 87, 213-219
0rskov, E.R. (I975) Manipulation of rumen fermentation for maximum food
utilization World Rev. Nutr. Diet 22, 152-182
0rskov, E.R. & Fraser, C. (1975) The effect of processing of barley based
supplements on rumen pH, rate of digestion and voluntary intake in sheep Brit.
J. Nutr. 34,493-500
0rskov, E.R. & Hovell, F.D. de B. (I978) Rumen digestion of hay by cattle given
sugar cane diets or pangola hay Trap. Anim. Prod. 3, 9-11
0rskov, E.R. & Oltjen, R.R. (1967) Influence of carbohydrate and nitrogen
sources on the rumen volatile fatty acids and ethanol of cattle fed purified diets
J. Nutr. 93, 222-228
0rskov, E.R., Fraser, C. & Kay, R.N.B. (1969) Dietary factors influencing the
digestion of starch in the rumen and small and large intestine of early weaned
lambs Brit. J. Nutr. 23, 217-226
0rskov, E.R., Fraser, C. & Gordon, J.G. (1974) Effect of processing of cereals on
rumen fermentation, digestibility, rumination time and firmness of sub-
cutaneous fat in lambs Brit. J. Nutr. 32, 59-69
0rskov, E.R., Soliman, H.S. & Macdearmid, A. (I978) Intake of hay by cattle
given supplements of barley subjected to various forms of physical treatment or
treated with alkali J. agric. Sci., Camb. 90, 611-616
Richardson, L.F., Raun, A.P., Potter, E.L., Cooley, C.O. & Rathmacher, R.P.
(1976) Effect of Monensin on rumen fermentation in vitro and in vivo J. Anim.
Sci. 43, 657-664
Silva, A.T. & 0rskov, E.R. (1988) Fibre degradation in the rumens of animals
receiving hay and untreated or ammonia treated straw Anim. Feed Sci. Technol.
19,277-287
Silva, A.T., Greenhalgh, J.F.D. & 0rskov, E.R. (1989) Influence of ammonia
treatment and supplementation on the intake, digestibility and weight gain of
sheep and cattle on barley straw diet Anim. Prod. 48, 99-108
CHAPTER 4

HOST ANIMAL CONTROL OF


MICROBIAL FERMENTATION AND
HOST ANIMAL DIGESTION

I. Control of rumen pH
A. Roughage feeding
B. Concentrate feeding
C. Rumen pH and feeding behaviour
II. Host animal control of urea recycling
III. Control of outflow rate
IV. Control of post-ruminal digestion

In some respects it can be said that when the ruminant animal has
consumed its food the rest of the process is left essentially to the
rumen microbes. However, the individual host animal can, directly
or indirectly, affect and cause differences in the environment
provided for fermentation. This is clearly illustrated by the fact
that, when feed degradation rates were measured in a group of
aIJ,imals, the greatest source of variation was between individuals
rather than between measurements made at different times (Mehrez
& 0rskov, 1977). Moreover, the variation in ruminal outflow rate
43
44 Energy Nutrition in Ruminants

also has a large between-animal component (0rskov et al., 1988).


In this chapter the most important sources of these differences
between animals are discussed.

I. CONTROL OF RUMEN pH

IA. Roughage Feeding

Cellulolytic bacteria need a rumen pH between about 6.2 and 7.0 in


order to multiply rapidly (see Chapter 2). On grass and forage diets
the pH is normally within that range ~ecause the animals secrete
copious amounts of saliva, which contains mainly sodium,
potassium, bicarbonate and phosphate. Kay (1966) reviewed the
literature on salivation and reported a range of 6-12 litresjd in
sheep and 110-170 litresjd in cattle. Thus the volume secreted daily
is usually greater than that of the filled rumen. It appears that the
amount of saliva secreted during the consumption of roughage and
its subsequent rumination, together with the carbon dioxide
produced during fermentation, is essentially geared to maintaining
a pH which is optimal for cellulose fermentation. The output
depends mainly on the time spent eating and ruminating, but saliva
is also secreted plentifully between periods of chewing. Even fasting
animals produce a small steady flow. Indeed, if the animals are
given small amounts of concentrate, although they eat and ruminate
for only brief periods a rumen pH of 6-7 is maintained, in contrast
to pH 5-6 induced by high levels of concentrate feeding.
The saliva serves to provide a buffered medium for the microbes
rather than to neutralize all the VFA produced. In fact the quantity
of saliva secreted is enough to neutralize only about one-third of
the VF A. Most VF A is .absorbed in its acid form through the
rumen epithelium, which is highly permeable to undissociated
VF A, and is neutralized by the blood buffering system until
metabolized. The bicarbonate and phosphate of the saliva have pK
values (i.e. pH values at which the salt is half dissociated from
hydrogen ions) of 6.2 and 7.4 respectively and they maintain the pH
of the rumen contents above 6, unless VFA is produced at an
Host control of fermentation and digestion 45
TABLE 4.1
Effects of Feeding Lambs with Whole and Pelleted Barley on Rumen pH,
Rumination time, Number of Swallowed Boluses and Digestibility

Barley Rumen Rumination Swallowed Digestibility


pH time boluses (%)
(min/24 h) (no./24 h)
Whole 6.2 402 521 78.0
Pelleted 5.2 216 263 76.7
SE 0.1 41 62 0.5
From 0rskov et af. (1974).

overwhelming rate. Thus although most of the VFA - which has


a pK value of about 4.8, well below the normal rumen pH - is
present in its neutralized form, most of that which is absorbed
comes· from the small fraction present in the acid form. Conse-
quently, when experimental animals are being sustained by
intragastric infusion of nutrients, it is neither necessary nor
desirable to neutralize all the VFA. This became very clear during
the development of the technique of intragastric nutrition (0rskov
et al., 1979). Attempts to neutralize VFA with sodium carbonate
before infusion led immediately to electrolyte imbalance, as well as
to problems of osmotic pressure. Since even infused animals secrete
some saliva of their own, artificial saliva had to be delivered at a
rate which in fact amounted to only about one-sixth of that
required to neutralize all the infused VFA. The situation thus
reflected the natural state, most VFA being absorbed in the
undissociated acid form, the form in which it is produced by the
rumen micro-organisms.

m. Concentrate Feeding
Whereas the optimum pH range for cellulose digestion is
maintained on forage-based diets, with more intensive concentrate
feeding it is not. There are several reasons for this. Concentrate-
based diets are usually more digestible, i.e. fermentable. Therefore
more VFA is produced per unit weight than with forage. At the
same time both the higher density and the smaller particle size
46 Energy Nutrition in Ruminants

TABLE 4.2
Effect of Eating Pattern on Rumen pH and Degradation of Washed Hay
Incubated in the Rumen of Sheep Maintained on a Mixed Diet of 65 %
Concentrate and 35 % Hay

Meal Sheep Rumen pH Duration of Degradability of


size pH < 6.0 hay (%) after
Mean Range (hj24 h)
24 h 48 h 72 h

Small A 6.1 5.9-6.3 7 22.1 36.9 44.7


Large B&C 5.8 5.3-6.6 19 10.3 21.9 27.6
From Istasse et al. (1986).

result in much more rapid consumption ana less rumination time.


Consequently, saliva production is relatively low. As Table 4.1
shows, lambs fed pelleted grain spent less time ruminating than
lambs fed whole grain. Although the digestibility of the two diets
was similar, the rumen pH in the lambs fed pelleted grain, which
salivated less, was 5.2 compared with 6.2 in the lambs fed whole
grain (0rskov et al., 1974).

Ie. Rumen pH and Feeding Behaviour

If rumen pH is a function of salivation, it is clear that feeding


behaviour may also affect it. Animals which eat slowly are likely to
produce more saliva than those which eat rapidly. Concentrate
given in restricted amounts will usually be consumed quickly. Table
4.2 contrasts the daily variation in rumen pH of one sheep which
ate small meals of a mixed diet at frequent intervals, with those of
two sheep which ate larger meals less frequently but consumed
similar total amounts (lstasse et al., 1986). The rate of degradation
of cellulosic feeds was much greater in the one animal which
showed a high and constant rumen pH.
It may be argued that if the proportion of cellulosic feeds in the
diet is small, rumen pH values below 6.2 will not have very serious
consequences for the digestion of the diet. However, a high-
concentrate diet may greatly extend the rumen retention time of
long roughage particles added to it, due to their low degradation
Host control of fermentation and digestion 47
rate, and so impede further intake. If the rumen pH approaches 5.0
another problem arises, that of rumenitis which, in cattle, often
leads to liver abscesses (Fell et ai., 1968).

ll. HOST ANIMAL CONTROL OF UREA RECYCLING

When the fermentation of roughage is limited by the N content of


the feed, endogenous urea entering the rumen by way of the saliva
and the rumen epithelium provides additional N for microbial
protein synthesis, so improving forage digestibility and intake.
Cereal diets also may contain an excess of fermentable energy
relative to N and th~refore similarly promote the use of endogenous
N by the micro-organisms. Utilization of blood urea in this way
decreases urinary excretion of urea, for the kidney is well able to
reabsorb most of the urea filtered at the glomerulus, especially
when water is scarce and the urine is concentrated. However, when
fermentation and microbial protein synthesis is limited by lack of
fermentable carbohydrate, the urea recycled to the rumen will
simply be reabsorbed as ammonia and ultimately excreted as
urinary urea. The rates at which urea is supplied both in the saliva
and by diffusion through the rumen epithelium depend on its
concentration in the blood. Animals on sub-maintenance rations of
poor N-limited roughages utilize tissue protein for energy, and the
additional urea released by the liver adds to the amount of blood
urea available for cycling to the rumen. This, in turn, may help to
remedy the situation by enhancing digestibility and intake.

IiI. CONTROL OF OUTFLOW RATE

The previous sections were concerned with the host animal's ability
to influence rumen pH and rumen ammonia concentration, by
varying saliva secretion and the recycling of endogenous urea
respectively. Evidently a large inflow of saliva must - of necessity
---;-lead to a large outflow of liquid, which could occur via the
rumen wall as well as by passage to the post-ruminal gut. Table 4.3,
based on an intragastric nutrition trial (0rskov et ai., 1986) shows
48 Energy Nutrition in Ruminants
TABLE 4.3
Effect of Ruminal and Abomasal Osmotic Pressure on Liquid Outflow Rate in
Three Sheep Maintained by Intragastric Nutrition

Mean Mean SEM

Liquid infused (litres/d) 9.9 9.9


{mOSmOlfkg water 288 373 25
Rumen pH 6.38 6.92 0.09
Outflow rate (litres/d) 5.9 9.6 0.9
{ mOsmol/kg water 264 299 12
Abomasum pH 3.28 3.25 0.49
Outflow rate (litres/d) 7.7 11.4 0.7
From 0rskov et al. (1986).

that increased osmotic pressure of the rumen fluid greatly


accelerates its outflow to the post-ruminal gut. However, the
precise control of liquid and particle outflow is still only poorly
understood. For example, as will be discussed further in Chapter 8,
environmental temperature can easily affect it. Also, since the
structure of the feed largely determines the rate of salivation, this
too modifies fluid outflow rate.
Individual animals also differ in the outflow rate of solids from
the rumen. Table 4.4, from 0rskov et al. (1971), shows how one of
a pair of lambs fed on maize fermented a far smaller proportion of
the intake within the rumen than the other. This, in tum, resulted
in very large differences both in the extent of starch digestion in the
small and large intestine and in ruminal microbial protein
production, indicated by the difference in the concentration of
diaminopimelic acid in the abomasum. Animals also differ with
respect to the concentration of dry matter in the rumen. This may
in part be due to differences in water consumption, and could be
related to the need to maintain a suitable osmotic pressure.
Animals may also differ in their control of the concentration of
solids in the digesta flowing from the reticulum to the omasum,
relative to the rather higher concentration in the reticulo-rumen.
Differences between the abilities of individual animals to
influence the rumen environment or between the levels at which the
various parameters tend to be stabilized can all affect the
Host control of fermentation and digestion 49

TABLE 4.4
Site of Digestion of a Maize Diet; Differences Between Two Lambs and
Consequences for Nitrogen Metabolism

Lamb

A B
Starch intake (gjd) 942 778
Fermented in rumen (gjd) 540 757
Digested in small intestine (gjd) 324 21
Fermented in large intestine (gjd) 57 o
Excreted in faeces (gjd) 21 o
Apparent digestibility of protein (%) 47 72
DAPAa: gj16 g abomasal N 3.56 4.95
Crude protein in faeces (gjd) 59 27

From 0rskov et al. (1971).


aDiamirtopimelic acid.

digestibility of the diet. A recent experiment with dairy cows


(0rskov et at., 1988) led to several important conclusions. Despite
similar diets, differences were demonstrated between animals with
respect to the fractional outflow rates of both liquid and particles
(Table 4.5). There was no significant correlation between the
fractional flow of liquid and that of particles. Differences between
animals in the particle outflow rate had consequences for the
apparent organic matter digestibility. Particle outflow was highly
and significantly negatively correlated with apparent digestibility
but there was no correlation between the latter and liquid outflow.
(The quantitative aspects of outflow rate and digestibility will be
discussed further in Chapter 8.) Interestingly, there was no
correlation between intake and digestibility in cows fed ad libitum,
although these parameters were significantly correlated within cows
when intake was restricted at various levels.

IV. CONTROL OF POST-RUMINAL DIGESTION

Little is known regarding the extent to which host animals exert any
control at the post-ruminallevel. The entry into the small intestine
of potentially digestible substrates is largely determined by the diet
50 Energy Nutrition in Ruminants
TABLE 4.5
Effect on Apparent Digestibility of Selecting Cows on the Basis of Low or High
Fractional Outflow of Small Particles from the Rumen

Organic matter intake (g/kgW O• 75 ) 101 99


Outflow rate of small particles (fraction/h) 0.0257 0.0309
Apparent digestibility (%) 75.1 71.4
From 0rskov et al. (1988).
Three cows in each group.

and by events at the level of the rumen. Table 4.4 shows that in one
lamb so much starch could pass into the abomasum that the
capacity for post-ruminal starch digestion was exceeded, while in
the other lamb this was clearly not the case.
The amount of fermentable substrate which enters the large
intestine is also largely determined by the diet and by events within
the rumen. Diets consisting of small cellulosic particles are more
likely to escape rumen fermentation than diets consisting of long
particles of straw etc. Thus, once again, between-animal differences
with respect to fermentation in the large intestine are largely
governed by between-animal differences in rumen function. In pigs,
forage-type diets induce a rapid hypertrophy of the large intestine
(Stephen, 1976). This is analogous to the way in which different
species of ruminants vary in caecum--colon capacity relative to
reticulo-rumen capacity, apparently reflecting the amount of
potentially digestible fibre that passes through the rumen without
degradation (Hofmann, 1989).

REFERENCES

Fell, B.F., Kay, M., Whitelaw, F:.G. & Boyne, R. (1968) The relationship between
the acidity of the rumen contents and rumenitis in calves fed on barley Res. Vet.
Sci. 10, 181-187
Hofmann, R.R. (1989) Evolutionary steps of ecophysiological adaptation and
diversification of ruminants: a comparative view of their digestive system
Oecologia 78, 443-457
Istasse, L., Smart, R.I. & 0rskov, E.R. (1986) Comparison between two methods
of feeding concentrate to sheep given a diet high or low in concentrate with or
without buffering substances Anim. Feed Sci. Techno!. 16, 37-49
Host control of fermentation and digestion 51

Kay, R.N.B. (1966) The influence of saliva on digestion in ruminants World Rev.
Nutr. Diet 6, 292-325
Mehrez, A.Z. & 0rskov, E.R. (1977) The use of a Dacron bag technique to
determine rate of degradation of protein and energy in the rumen J. agric. Sci.,
Camb. 88, 645-650
0rskov, E.R., Fraser, C. & McDonald, 1. (1971) Digestion of concentrate in sheep.
3. Effect of rumen fermentation of barley and maize diets on protein digestion
Brit. J. Nutr. 26, 477-486
0rskov, E.R., Fraser, C. & Gordon, J.G. (1974) Effect of processing of cereals on
rumen fermentation, digestibility, rumination time and firmness of subcu-
taneous fat Brit. J. Nutr. 32, 59-69
0rskov, E.R., Grubb, D.A., Wenham, G. & Corrigall, W. (1979) The sustenance
of growing and fattening ruminants by intragastric infusion of volatile fatty
acids and protein Brit. J. Nutr. 41, 553-558
0rskov, E.R., MacLeod, N.A. & Kyle, D.J. (1986) Flow of nitrogen from the
rumen and abomasum in cattle and sheep given"protein free nutrients by
intragastric infusion Brit. J. Nutr. 56, 241-248
0rskov, E.R., Ojwang, I. & Reid, G.W. (1988) A study on consistency of
difference between cows in rumen outflow rate of fibrous particles and other
substrates and consequences for digestibility and intake of roughages Anim.
Prod. 47, 45-51
Stephen, T.G. (1976) A Preliminary Evaluation of the Value of Grass and Grass
Pulp for Non-lactating Sows M.Sc. Thesis, University of Aberdeen
CHAPTER 5

ABSORPTION OF NUTRIENTS

I. Absorption of volatile fatty acids from the rumen


A. Effect of rumen pH
B. Effect of blood flow
II. Post-ruminal digestion and absorption
A. Small intestine
B. Large intestine

I. ABSORPTION OF VOLATILE FATTY ACIDS FROM


THE RUMEN

Volatile fatty acids are the ruminant animal's main source of


energy. As described in Chapter 2, they are produced when the feed
is fermented in the rumen. They are similarly produced during the
fermentation of feed residues in the large intestine (although this
generally accounts for only a small proportion of the total VF A
produced).
About 80% of the VFAs produced in the rumen are absorbed
through its wall, the surface area of which is greatly enlarged by
52
Absorption of nutrients 53

numerous papillae. These papillae grow and regress in response to


changes in food consumption and VF A concentration (Hofmann,
1989). The rest of the rumina1 VFA is absorbed from the omasum
and abomasum. At normal rumen pH values (between 6 and 7)
95% or more of the VFAs are in their dissociated, ionized form (see
Chapter 4), yet the rumen wall is so much less permeable to
dissociated than to undissociated VF A that most is absorbed in the
latter, acid form, in which it is also produced. The absorption of
VFA by the papillae creates an alkaline environment near their
surfaces. B.F. Fell & E.R. 0rskov (unpublished) found that the pH
between the papillae of concentrate-fed lambs was about 1 unit
higher than that of the main rumen contents.

IA. Effect of Rumen pH

Evidently, if the VFAs are absorbed mainly in the undissociated


form, absorption must be greater at lower rumen pH values. A
controversial issue, that has generated many lively debates, relates
to the relative proportions of the different VF As found in the
rumen. Do they reflect the molar proportions in which they are
produced? Early work by Danielli et al. (1945) with epithelial slices,
showed clearly that the absorption rates of different VF As could
differ. At rumen pH values at or above 7 the order was acetic
acid> propionic acid> butyric acid, while at lower pH values the
sequence was reversed. Because this work was carried out in vitro
it has largely been ignored. Most workers still assume that the VF A
proportions in the rumen represent the proportions in which they
are produced, disregarding the possibility of differential absorption.
Indeed, some Australian workers (Leng, 1966) concluded from
isotope measurements that it was valid to do so. However,
Thorlacius & Lodge (1973), using animals with rumen pouches,
confirmed the importance of taking rumen pH into account when
determining the relative absorption rates. It is significant also that
Sutton & Morant (1978) found a closer relationship between the
molar ratios produced and those observed in the rumen when cows
received a hay diet than when they received a concentrate diet,
which would promote acid conditions.
54 Energy Nutrition in Ruminants

Since VFA production rates are not steady, it has proved difficult
to obtain valid measurements of absorption using radioactive
isotopes. However, the development of the intragastric nutrition
technique (0rskov et al., 1979) - which allows nutrient intake to
be limited to VFA infused into the rumen and proteins into the
abomasum - opened new perspectives for this type of study. A
constant rate of infusion can be maintained and the proportions of
the various VFAs in the infusate can be compared with those in the
rumen. In a recent trial (MacLeod et al., 1984), the rumen pH was
varied by altering the amount of buffer infused and it was clearly
shown that the difference between the VFA molar ratio of the
infusate and that of the rumen contents increased as the pH
decreased (Fig. 5.i).Although the infusatt3contained 63 molar % of
acetic acid, the rumen contents contained 73% at pH 5.5 and 66%
at pH 6.6. Thus differential absorption rates must be taken into
account, particularly at low rumen pH values, in any study ofVFA
absorption.
N.A. MacLeod & E.R. 0rskov (unpublished) have also looked
for possible effects of osmotic pressure on VFA absorption rate.
The three main VFAs were not affected differentially but the
overall rate declined as osmotic pressure increased, particularly
above 330-350 mOsmol/kg water. At these osmotic pressures a
high proportion ofVFA enters the abomasum in the liquid outflow.
Here it causes disturbances by acting as a buffer, so tending to raise
the abomasal pH above its normal range of 2.3-2.8 and
consequently stimulating additional secretion of gastric acid and
also of abomasal nitrogen (0rskov et aI., 1986).

lB. Effect of Blood Flow

The effect of blood flow ,on the absorption of VFAs has recently
been excellently reviewed by Dobson (1984) and by Barnes et al.
(1986). Dilation of the blood vessels in the rumen epithelium and
increased blood flow, which occur within 2-3 h after feeding,
appear to be responses to the CO 2 and VF A produced in the rumen.
Blood flow to the rumino-reticular muscle also increases as a result
of a faster contraction rate. The enhanced ruminal blood flow
Absorption of nutrients 55

760

740 •• • • •
• • • • •
'0 0 720 • ••
0 0 0 0
'i:j E
!'II- 700
•• 0
0
0

• •
0 0
vO 0
:,:; E 0
~ E 680 0
i
<{~ 000
660 0 • • 0

640
-- -- -- -- -- -- --
-- --
270
~~
-- -- -- --
v- 250
• •
0 e
!'II 0 000 0 0
vE
-
.-c - 230 • ••• 0
0
0
o 0 0 0
0

•• •••
o 0 0 0 0
.- E
g-E 210 •• • • •
L.

Q.
~

190 • •
110 -- -- -- -- -- -- -- -- -
• ••
'O~

'u 0 90
!'II E 0 000 0
v-
·C "0 70 • ••• 0
0
i •
0
0

~E
:> E
0
•• ••
0
i e ee i i
1Xl~ 50

7·0 0
0
J:
Q.
6 '5
0 0
0
• •
• •
0 0
...c 6'0
000
0
0
E
• 0 0 0 0 0 0
0

• • • •

:>
5·5
••• •
•• • • •
a:: ~
5 '0
0 2 4 6 8 10 12 14 16 18 20
Time (h)

Fig. S.i. The effect of fluctuating rumen pH on the molar proportions of VFAs
(mJlloljmol) in the rumens of two lambs (0 and.) wholly maintained by VF A
infusion. The broken horizontal lines indicate the molar proportions in the infusate.
(From MacLeod et al. (1984).)
56 Energy Nutrition in Ruminants

promotes an increased rate of absorption in response to increasing


VFA production. There is insufficient information regarding other
possible influences on the flow of blood to the alimentary tract,
such as the rate at which nutrients are removed from the circulation.
If the latter had an effect the intake of feed might be reduced even
in situations where there were no physical limitations to intake.
Assuming that absorption rate depends on blood flow, evidence
from ruminants maintained exclusively on continuous VFA
infusion (0rskov et ai., 1979) may be valuable in this context. Since
VFAs are absorbed in the acid, undissociated form, the buffer used
for intragastric infusions is administered in quantities much smaller
than that needed to neutralize the acids. It follows that any decrease
in absorption rate will decrease rumen pH. Indeed, rumen pH is
very sensitive to changes in absorption rate. When animals are
given protein-free infusates for a long period rumen pH decreases,
signifying reduction in VFA absorption. If animals receiving
intragastric infusion are stressed by infection or environmental
change the rumen pH declines similarly. On the other hand, rumen
pH increases in animals during rapid growth and also in those
which are lactating.

II. POST-RUMINAL DIGESTION AND ABSORPTION

IIA. Small Intestine

When ruminants are normally fed on roughage-based diets, the


digesta that undergo changes and absorption in the small intestine
consist, to a large extent, of microbial matter, along with any
dietary protein and lipids which have escaped ruminal degradation.
It is of interest to knO\y whether concentrate feeds and other
supplements can be digested and absorbed there too. Investigation
of this question is possible because such nutrients can be made to
bypass the rumen by means of the oesophageal groove reflex
(0rskov et ai., 1970a) or can be protected from rumen degradation,
thus avoiding losses of heat and gases during fermentation (cf.
Chapter 1).
Absorption of nutrients 57
140

120

100

-
~

'0
en
80
"
11\
0
60
v
:>
(5
40

20

0 40 80 120 160 200 240 280 320 360 400 440 480 520
Glucose infu sed (g Id)

Fig. S.ii. Effects of glucose infusion via the abomasum on glucose passing the
terminal ileum (-) and excreted in the faeces (---) of two sheep (0 and .).
(From 0rskov et al. (1971) .)

The complete absence of endogenous sucrose-hydrolysing


enzymes in ruminants was discussed in Chapter 1. Sucrose entering
the small intestine will appear almost quantitatively in the terminal
ileum, where local fermentation probably accounts for the small
amounts which disappear (0rskov et al., 1972). The limited
capacity for starch digestion has also been referred to. This relates
to raw starch in particular, the low amylase activity and brief
retention time severely limiting the amount digested in the small
intestine to about 200 gjd in sheep. Using gelled starch, Mayes &
0rskov (1974) showed that maltose accumulated, indicating that
the relatively low maltase activity of the intestinal brush border
limited its digestion. The amount of maltose digested in the small
intestine ofthe sheep was 200-300 gjd.
Even the capacity to absorb glucose is limited. Under ideal
circumstances, with continuous abomasal infusion, 0rskov et al.
(1971) showed that a maximum of about 300 g could be absorbed
in 24 h by a 40 kg sheep (Fig. 5.ii). When a similar quantity was fed
intermittently from a bottle, large amounts escaped absorption in
th"e small intestine and reached the caecum-colon, where they
altered the pH, the fermentation products and the number of
bacteria in the caecal fluid (Mann & 0rskov, 1973). Table 5.1, from
58 Energy Nutrition in Ruminants

TABLE 5.1
Effects on Caecal pH and Flora of 250 g/d of Different Carbohydrates
Entering the Rumen or the Abomasum (by Means of the Oesophageal Groove
Technique)

Carbohydrate Point of entry Caecal pH Viable bacterial


count x 10-8

Rumen 6.6 1
Sucrose
Abomasum 5.6 19
Rumen 5.0 2
Maltose
Abomasum 6.2 37
Rumen 6.6 10
Glucose
Abomasum 6.0 68
Rumen 6.6 200"
Cellulose
Abomasum 6.4 176"

From Mann & 0rskov (1973).


liTotal count.

the above paper, shows the effects of feeding glucose, maltose,


sucrose and cellulose, both into the rumen and directly to the
abomasum, by bottle-feeding. It can be inferred that some of the
glucose, sucrose and maltose given intermittently by bottle, at the
rate of 200 gj d, passed into the large intestine, since the pH
decreased and the numbers of viable bacteria in the caecum
increased. In contrast, cellulose caused no marked caecal changes.
It is clear from Fig. 5.ii that the capacity for glucose absorption is
rather limited in the normally functioning ruminant. This, in tum,
limits the extent to which post-ruminal digestion of oc-linked
carbohydrates, such as starch, should be encouraged, even if it were
to reduce the ruminal fermentation losses of heat and methane.

DB. Large Intestine

The mechanisms of production and absorption of VFAs in the


large intestine are probably similar to those in the rumen. The
buffering is obviously supplied not by saliva but by bicarbonate-
rich intestinal secretions. The pH of ileal fluid is usually between 7
and 8. The mucous epithelium of the caecum can absorb VFA
rapidly (Sakata, 1987). Some experiments have been carried out to
Absorption of nutrients 59

'0
~ 75
~"
"
E
»
L
'0
Cl

.... 50
:;,
Q.
....:;,
o
.c.
u
L

...."III o
I.

.-.-.-. 4f
ii 25
u

tJ."
/0
0-0-0- -0
0-0- ......
o
I I , I I I I I I I I I I I I
o 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Day of infusion
I , , , , , I , , , I I I , I I

o 20406080100 140 180 220 260 300


Starch infused (g/d)

Fig. S.iii. Influence of caecal starch infusion on daily faecal output of starch by
two sheep (0 and .) which received a basal diet of 900 gld of dried grass
pellets. (From 0rskov et at. (l970b).)

measure the capacity of the large intestine for fermentation. 0rskov


et al. (1970b) infused varying amounts of starch into the caecum of
sheep (Fig. 5.iii) and noted the level at which starch first appeared
in the faeces. This occurred just above 100 gjd. The infused starch
altered the type of fermentation to a pattern comparable to that
induced by starch in the rumen. This change was reflected in the
colon and in the faeces (see Table 5.2). When the capacity for starch
fermentation is exceeded, the faeces become increasingly soft and
watery, similar in consistency and VFA content to rumen fluid. If
the animals are given a bottle of sucrose solution scouring occurs
within 3-4 h. In such circumstances caecal and even faecal pH can
be below 5. The caecal contents of lambs given large amounts of
60 Energy Nutrition in Ruminants
TABLE 5.2
Proportions of Different VFAs (molar %) in the Colon, Caecum and Faeces of
Two Sheep when no Starch was Infused and when 300 g of Starch were Infused
Daily into the Caecum

Sheep Starch Site Acetic Propionic Butyric Higher Total VFA


infused acid acid acid acids concentration
(gld) (mmoillitre
faecal H 2 O)
A 0 Colon 66.0 15.4 6.6 12.1
A 300 Colon 61.4 10.6 20.6 7.4
B 0 Caecum 77.1 7.8 4.7 10.5
B 300 Caecum 65.1 9.8 14.8 10.3
A 0 Faeces 68.5 20.5 6.7 4.4 64.8
A 300 Faeces 70.2 5.1 19.5 5.2 227.6
B 0 Faeces 69.0 14.6 5.5 10.9 64.8
B 300 Faeces 62.9 8.8 18.3 10.0 118.8

From 0rskov et al. (l970b).

milk by bottle also become quite acid and may contain high
concentrations of both VF A and lactate, without untoward effect
(Robson & Kay, 1972).
Hofmann (1989) summarized his and his colleagues' excellent
and extensive observations on the morphology of the ruminant gut.
Those species that habitually select rich browse diets ferment the
forage rapidly but briefly in the rumen, under rather acid
conditions. Much potentially degradable fibre leaves the rumen but
is fermented further in the caecum-colon, which is considerably
enlarged, having a capacity 1/6-1/10 of that of the rumen,
compared with 1/15-1/30 in grazing species such as cattle and
sheep. Adaptation of large intestine capacity and mucosal archi-
tecture may occur in animals that must face major seasonal
variations in the quantity and quality of food available to them.
Such adaptations to changes of diet occur in domesticated species.
However, caecal fermentation does not permit utilization of the
microbial protein. Unlike that formed in the rumen, it cannot be
absorbed and is lost in the faeces (0rskov et al., 1970b; 0rskov,
1982; Kay, 1983). This disadvantage must be borne in mind when
formulating diets which may encourage caecal fermentation.
Absorption of nutrients 61

REFERENCES

Barnes, R.J., Comline, R.S. & Dobson, A (1986) The control of splanchnic blood
flow In Milligan, L.P., Grovenor, W.L. & Dobson. A (Eds) Control of
Digestion and Metabolism in Ruminants pp 41-59 Proc. 6th Int. Symp. on
Ruminant Physiology, Banff, Canada. Prentice-Hall, Englewood Cliffs, NJ
Danielli, J.F., Hitchcock, M.W.S., Marshall, R.A. & Phillipson, AT. (1945) The
mechanism of absorption from the rumen as exemplified by the behaviour of
acetic, propionic and butyric acids J. expo BioI. 22, 75-84
Dobson, A. (1984) Blood flow and absorption from the rumen Quart. J. expo
Physiol. 69, 599-606
Hofmann, R.R. (1989) Evolutionary steps of ecophysiological adaptation and
diversification of ruminants: a comparative view of their digestive system
Oecologia 78, 443-457
Kay, R.N.B. (1983) Rumen function and physiology Vet. Rec. 113, 6-9
Leng, R.A. (1966) Volatile fatty acid production in the rumen of sheep Proc.
Austr. Soc. Anim. Prod. 4, 389-394
MacLeod, N.A., 0rskov, E.R. & Atkinson, T. (1984) The effect of pH on the
relative proportions of ruminal volatile fatty acids in sheep sustained by
intragastric infusions J. agric. Sci., Camb. 103, 459-462
Mann, S.O. & 0rskov, E.R. (1973) The effect of rumen and post-rumen feeding
of carbohydrates on the caecal micro flora of sheep J. appl. Bact. 36, 475-484
Mayes, R.W. & 0rskov, E.R. (1974) The utilization of gelled maize in the small
intestine of sheep Brit. J. Nutr. 32, 143-153
0rskov, E.R. (1982) Protein Nutrition in Ruminants Academic Press, London
0rskov, E.R., Benzie, D. & Kay, R.N.B. (l970a) The effect of feeding procedure
on closure of the oesophageal groove in sheep Brit. J. Nutr. 24, 785-795
0rskov, E.R., Fraser, c., Mason, V.c. & Mann, S.O. (l970b) The influence of
starch digestion in the large intestine of sheep on caecal fermentation, caecal
microflora and faecal nitrogen excretion Brit. J. Nutr. 24, 671-682
0rskov, E.R., Mayes, R.W. & Penn, A (1971) The capacity for removal of glucose
from the small intestine in mature sheep Proc. Nutr. Soc. 30, 43A-44A
0rskov, E.R., Mayes, R.W. & Mann, S.O. (1972) Post ruminal digestion of
sucrose in sheep Brit. J. Nutr. 28, 425-432
0rskov, E.R., Grubb, D.A., Wenham, G. & Corregall, W. (1979) The sustenance
of growing and fattening ruminants by intragastric infusion of volatile fatty
acids and protein Brit. J. Nutl". 41, 553-558
0rskov, E.R., MacLeod, N.A. & Kyle, D.J. (1986) Flow of nitrogen from the
rumen and abomasum in cattle and sheep given protein-free nutrients by
intragastric infusion Brit. J. Nutr. 56, 241-248
Robson, M.G. & Kay, R.N.B. (1972) Changing patterns of fermentation and
mineral absorption in the large intestine of lambs weaned from milk to
concentrates Proc. Nutr. Soc. 3 I, 62A
Sakata, T. (1987) Stimulating effect of short chain fatty acids on epithelial cell
proliferation in the rat intestine: A possible explanation for trophic effects of
fermentable fibre, gut microbes and luminal trophic factors Brit. J. Nutr. 58,
95-103
62 Energy Nutrition in Ruminants
Sutton, J.D. & Morant, S.V. (1978) Measurement of the rate of volatile fatty acid
production in the rumen In Osbourn, D.F, Beever, D.E. & Thomson, D.J. (Eds)
Ruminant Digestion and Feed Evaluation pp 71-79 Agricultural Research
Council, London
Thorlacius, S.O. & Lodge, G.A. (1973) Absorption of steam volatile fatty acids
from the rumen of the cow as influenced by diet, buffers and pH J. Anim. Sci.
53,279-288
CHAPTER 6

ENERGY METABOLISM OF THE HOST


ANIMAL

I. Basal and fasting metabolism


II. Estimation of energy maintenance
III. Cost of different activities
A. Standing
B. Eating
C. Rumination
D. Walking
E. Draught power
F. Fat and protein deposition
G. Pregnancy
H. Lactation
I. Wool and fibre production
J. Keeping warm and keeping cool
i. Low environmental temperatures
ii. High environmental temperatures

In this chapter the cost of the different activities performed by


ruminants will be discussed briefly, although for some of these it is
as yet poorly documented.
63
64 Energy Nutrition in Ruminants

I. BASAL AND FASTING METABOLISM

Basal metabolism is generally defined as the heat production of a


completely quiescent animal in a post-absorptive state, within a
thermoneutral environment. However, although this state can be
achieved with humans, it is extremely difficult to achieve with other
animals. Consequently the term 'fasting metabolism' has been
adopted for them. To measure fasting metabolism the animals are
normally kept in a respiration chamber, where activities other than
standing or lying are minimal. One problem in ruminants is that it
takes a long time to reach a truly fasting or post-absorptive state.
Usually 5 days without food is advocated before measurements are
made, but even at the end of that period fermentation may still be
continuing in the rumen or large intestine.
In many feed evaluation systems the value for fasting metabolism
serves as the base-line against which the utilization of different
feeds is measured, particularly utilization for energy maintenance.
However, in recent years this practice has been challenged (Webster
et al., 1974), mainly on the grounds that values obtained while the
animal is fasting will vary with its nutritional state prior to fasting.
This is because the sizes of the metabolically active organs, such as
the liver, kidney etc., are greatly affected by the level of nutrition
(Koong et al., 1985). As 5 days of fasting allows little adjustment
of organ size, fasting metabolism measured in well-nourished
animals is greater than that measured in animals that have recently
been fed at maintenance level or have been under-nourished.
Another argument against the use of fasting metabolism as the
base-line stems from the oxidation of fat, which occurs in the
fasting animal and is associated with greatly increased levels of p-
hydroxybutyrate in the blood. Moreover, some of the blood amino
acids - presumably generated during protein turnover - are
oxidized to provide a source of glucose precursors, while the
associated N is excreted. The difference in N excretion during
fasting and during energy maintenance by propionic acid or glucose
infusion can be seen in Table 6.1 and Fig. 6.i, from KuVera et al.
(1988). Figure 6.i is based on their own data and those of Asplund
et al. (1985), obtained by the technique of intragastric nutrition. In
Energy metabolism of the host animal 65

100
f:::"
f:::"
f:::"f:::"

.. •
80 f:::"
~
• f:::"
0 f:::"

c: f:::"f:::" o
0 f:::" f:::"
....Q.I
l-
V
x
60 .f:::,,&

f:::"
0
0
f:::"
0 o
Q.I
0 f:::"
z

40

o 40 80 120 160 200 240 280 320 360


Energy supply (kJ/kgWO'75 /d)

Fig. 6.i. Reduction in excretion of N by fasting cattle and sheep which resulted
from the infusion of glucose or propionic acid (l00 % = N excretion in the ab-
sence of infusion): 6" cattle receiving glucose; ,A., cattle receiving propionic acid;
0, sheep receiving glucose; . , sheep receiving propionic acid. (Adapted from
KuVera et al. (1988).)

general, fasting N excretion is some 40% greater than basal N


excretion or N excretion during energy balance. These character-
istics of fasting metabolism show that fasting is a state of glucose
deficiency and thus comparable to a nutritionally unbalanced diet.
It could therefore be argued that it should not be used as the basis
for measuring utilization of nutritionally balanced diets. Ku Vera et
al. (1988) showed that, by infusing small amounts of glucose or
glucose precursors into the abomasum, fasting N excretion could
be reduced to basal levels at an energy input of about one-third of
the maintenance energy need, while blood levels of p-hydroxy-
butyrate also fell to normal levels. Heat production did not increase
as a result of this glucose infusion. Indeed it sometimes decreased,
0\
0\

TABLE 6.1
Effects of Abomasal Infusions of Glucose or VFA on the Heat Production, Urinary N Excretion and Plasma Metabolites of
Otherwise Fasting Friesian Steers ~
~

~
Substance Heat Urinary N Plasma concentration of ~

infused (MJ/d) .production excretion ~


(kJ/kgWO· 75 /d) (mg N/kgWO· 75 /d) p-hydroxybutyrate Insulin Urea Glucose ~
:::;..
(mmol/ litre) (p./mlitre) (mmol/ litre) (mmol/ litre) §.
None 365 616 1.08 8.9 2.6 3.8 s·
~
Glucose 1.5 360 533 1.03 8.0 2.3 3.7 §
Glucose 3.1 368 482 0.79 8.9 1.9 3.8 s·
Glucose 4.6 360 396 0.59 8.4 1.6 3.8 §
Glucose 7.7 363 389 0.44 8.6 1.2 3.9 t;
Glucose 10.8 365 354 0.46 9.0 1.2 4.0
VFA 10.8 400 365 0.47 9.9 1.2 3.8
From KuVera et al. (1988).
Each value is mean of three steers.
Energy metabolism of the host animal 67

while the energy content of the urea and other nitrogenous


compounds excreted in the urine declined (Table 6.1).
In view of these results assessment of the utilization of feed
energy for maintenance would be better based on an energy intake
level which ensures no increase in N excretion, rather than on
fasting heat production. Indeed the changes which result from
fasting may well be the source of the distinction between metabolic
efficiency below maintenance level (Km) and above it (Kf), adopted
by the Agricultural Research Council (1980). Figure 6.ii (from
0rskov, 1982a) shows the slope for energy balance plotted against
energy input above maintenance level, corresponding to Kf.
Extension below this level towards fasting metabolism demonstrates
a change of slope. This has given rise to a concept of a difference
between the efficiency of energy utilization below energy main-
tenance (Km) and that above it (Kf). It is argued here that fasting
metabolism is equivalent to metabolism associated with a very
unbalanced diet, as indicated by Fig. 6.i. Its characteristics lead to
overestimation of the slope and consequently give rise to apparently
greater efficiency of feed utilization below maintenance level. If the
first measurement of a feed utilization trial were made at about 0.3
times energy maintenance, it is unlikely that differences between Km
and Kf would be detected. Although this question requires further
investigation, the consequence of the hypothesis suggested by Fig.
6.ii is that feed energy utilization is similar above and below the
maintenance level. Since it is known that both above and below this
level the simultaneous synthesis and degradation of both fat and
protein tissue accelerate in response to increased feed intake, this is
not an unreasonable assumption.

II. ESTIMATION OF ENERGY MAINTENANCE

While methods of estimating the energy requirement for main-


tenance are open to dispute, it is a useful term for expressing the
level of the exogenous nutrient supply. It is defined as the
metabolizable energy (ME) input per day at which the animals are
in energy balance. Many values are given in the literature, which
has been summarized by the Agricultural Research Council (1980),
68 Energy Nutrition in Ruminants

GI
u
c
ra
ra
.0
>.
en
L
GI
c
W

Fasting -+

Metabolizable energy input


Fig. 6.ii. Relationship between energy balance and energy input above and below
maintenance level: implications for concepts of Km and Kf. (From 0rskov
(1982a).)

with an average between 420 and 460 kJjkgWO. 75 for both sheep
and cattle. 0rskov & McDonald (1970), using comparative
slaughter data, obtained a value of 420 kJ jkgWO. 75 for lambs.
The maintenance energy need can be determined in a respiration
chamber, usually with one feeding level just below and one just
above that need. Nevertheless, average values must be applied to
animals in the field with some care. Energy maintenance values are
generally expressed relative to live weight, which assumes - apart
from the effects of previous nutrition on organ weights mentioned
earlier - that neither the proportions of fat and protein in the
body nor, in particular, the weight of the gut contents is important.
There is relatively little evidence indicating that the proportions of
fat and protein affect the maintenance requirement per unit either
of total body weight or of metabolic body weight. Blaxter (1962),
on the basis of Schiemann's (1958) results, concluded that there was
probably no difference between fat and thin animals in their
maintenance requirement per unit body weight.
Energy metabolism of the host animal 69
Concerning gut volume and weight, it is highly unlikely that the
energy per kilogram required for maintaining the gut and for work
associated with the gut contents equals that required to maintain
body tissue. Mould et al. (1982) showed that the gut of some
Bangladeshi cattle accounted for 33% of the live weight, whereas in
European cattle it accounts for only 20%. It is improbable that the
gut maintenance requirement, relative to body weight, i!) the same
for both. This question of gut content is well illustrated by some
results obtained at the Rowett Research Institute (0rskov et aI.,
1975). Lambs were given either whole barley or whole oats ad
libitum and were slaughtered at 35 kg live weight. In order to
calculate the efficiency of utilization of digestible energy for
fattening, the estimated intake required for maintenance was, in
each case, subtracted from the total consumption. The results
indicated that oats were far better utilized than barley (Table 6.2).
However, when the lambs were slaughtered, the gut contents of
those fed on oats weighed 9.4 kg while the gut contents of those fed
on barley weighed 4.7 kg. Calculation of the efficiencies of
utilization based on estimated empty body weights showed no
difference between the feeds in utilization of digestible energy,
which is what would be expected.

III. COST OF DIFFERENT ACTIVITIES

IlIA. Standing

One of the most common activities of mammals is standing up.


This incurs a cost and statements about fasting metabolism ought
to include the proportion of time that the animal was standing. The
extra cost of standing, compared with lying down, may be expressed
in terms of energy/kg live wt. The Agricultural Research Council
(1980) estimate, based on the findings of Blaxter (1962), is 10 kJ /kg
live wt./d, which amounts to a total of 5 MJ /d for a 500 kg animal.
KuVera et al. (1988) obtained a value for the difference between
continuously standing up and continuously lying down, of 12 J /kg
live wt./min, which amounts to 8.6 MJ/d for a 500 kg animal.
-.]
o

TABLE 6.2
Effects of Oats and Barley and of Duration of Diet on Weight of Lambs' Gut Contents and on Intake above Maintenance ~
Level of Digestible Organic Matter (DOM) and the Efficiency of its Utilization (Lambs Slaughtered at 35 kg live weight) ""
~
~

Oats Barley Days to Gut Total DOM DOM intake Empty body ~
(%) (%) slaughter contents intake above wt gain S.
(kg) (kg) estimated (kg/kg DOM ~.
;:,:
maintenance above s·
(kg) maintenance) :;.;,
o 100 74 4.7 42.1 19.5 0.92
§

45 55 91 6.7 49.4 19.9 0.95 §
100 o 120 9.4 51.7 12.7 1.36 t::;'

From 0rskov et al. (1975).


Energy metabolism of the host animal 71

Thus, relative to a maintenance energy requirement of about 45


MJ /d, the energy required to keep upright all the time is substantial
- up to about 20% of the total expenditure. Summers et al. (1988)
recently reported that the resting metabolic rate was 26% higher in
standing ewes than in those lying down. These results emphasize
the importance of recording the animal's position when measure-
ments are made of fasting or resting metabolism.

nm. Eating

Depending on the type of diet -large/long particles or ground


and pelleted, concentrate or roughage, solid or liquid - ruminants
can spend from a few minutes to more than 8 hours per day eating.
While eating they are normally standing up. Consequently, it is
extremely important to recognize that differences in the utilization
of dietary energy can sometimes be due to differences in the time
spent eating.
While it is relatively easy to measure the apparent cost of eating,
other factors associated with this activity, such as excitement and
standing, complicate the process. Measurement is complicated
further by the absorption of nutrients. Most feeds contain a soluble
fraction which ferments rapidly to VF As. These in turn induce
increased blood flow to the rumen and intestine, resulting in the
classical increase in heat production at the beginning of the meal
described many years ago by Rubner (1902). The true measurement
of the cost of eating can thus be confused with or confounded by
the so-called heat increment of feeding. These confusions could
theoretically be overcome by using animals fitted with oesophageal
cannulae to prevent the feed from entering the rumen, but this has
not as yet apparently been attempted.
The true energy cost of eating appears to be due almost entirely
to the cost of chewing. The rate of ingestion, i.e. dry matter
consumed per unit time, is unimportant (Holmes et al., 1978). Data
from Holmes et al. (1976, 1978) and Adam et al. (1984) give an
average value of 32 J /kg live wt. Imino However, KuVera et al.
(1988) gave indigestible feed to steers (otherwise maintained by
intra gastric nutrition) to avoid any effects of nutrient absorption,
72 Energy Nutrition in Ruminants

and obtained an average value of 11.6 Jjkg live wt.jmin. It is


evidently important to determine which is the correct value since,
for 500 kg animals which spend 8 hjd eating, the cost would be
either 2.8 or 7.8 MJ jd, i.e. either 6% or 17% of the total
maintenance energy requirement. Nevertheless, whichever value is
correct, the cost is quite sizeable and can make a substantial
difference to productivity, particularly at low levels of production.
For example Wainman et al. (1972) found that the Kf oflong, dried
grass was 40% above maintenance level, while that of the same
grass when pelleted was 52%. This could well be attributable to
time spent eating.

nIe. Rumination
Depending on the type of feed, a ruminant may spend up to 8 hjd
ruminating. The duration depends largely on the physical form of
the diet. Since rumination is associated with increased saliva flow,
the duration is important with some feeds for maintaining a stable
rumen environment. In general, the cost of rumination appears to
be considerably less than the cost of eating. Graham (1964)
reported a value of 16 Jjkg live wt.jmin for sheep. However,
KuVera et al. (1988), using the intragastric nutrition technique,
obtained a value of 9.3 Jjkg live wt.jmin, which is close to their
estimate for the cost of eating when confounding effects due to
excitement, increased blood flow etc. are avoided (see above). It is
to be expected that the cost of the physical work involved in the two
activities will be similar.

nID. Walking

Most measurements of energy expenditure have been made on


stationary animals kept indoors. Nevertheless, measurement of the
energy cost of walking, both on the level and on a gradient, has
been attempted. Blaxter (1962) summarized the available data and
arrived at a value of about 2 Jjkg live wt.jhorizontal metre. Thus
the additional energy expenditure of a 500 kg animal walking 6 km
daily, on the level, amounts to 6 MJ jd. Blaxter also estimated that,
Energy metabolism of the host animal 73

under normal conditions in a thermoneutral environment, heat


production out of doors was 10-15% greater than indoors. If the
outdoor activities are the same for animals fed at the maintenance
level and for animals consuming three times as much, the cost of
these activities, expressed as a percentage of heat production, will
evidently vary with the level of nutrition. The proportional increase
in energy expenditure will be greater at lower levels of feed intake.

HIE. Draught Power

There is little doubt that the greatest contribution of cattle and


buffaloes to human needs is that of draught power, yet it is the
contribution which has received least attention. All physical work
by ruminants involves walking, the cost of which has been
discussed. However, Lawrence (1986) noted that walking accounts
for a different proportion of the energy consumption, depending on
the type of work. He developed a simple formula with which he
examined the energy utilization of large oxen (about 650 kg) when
ploughing and when pulling a cart. Table 6.3 shows that even when
ploughing a very large proportion of the expended energy was
needed simply for walking. Assuming a maintenance energy
requirement of 400 kJ/kgWO,75, i.e. 50 MJ, the total energy
requirement for work relative to simple maintenance was about
1.75. However, the results of several workers suggest that the cost
per day of physical work in fact seldom exceeds 1.7 times
maintenance (Pearson, In press) and is generally about 1.3-1.4
times maintenance. Moreover, it is rare for cattle or buffalo to work
every day, so maintenance costs relative to work output are usually
even greater.
Nevertheless, female cattle, which are often used for draught
power, may simultaneously be pregnant and/or lactating, which
dramatically increases their energy requirement. Their draught
power is mostly used in areas where crop residues and poor-quality
roughages provide much or all of their diet. Such feeds may so limit
consumption that an extra 1.7 times maintenance requirement
cannot be eaten. The animals must then rely on stored body fat if
they are to perform sustained physical work.
74 Energy Nutrition in Ruminants
TABLE 6.3
Effects of Type of Work on the Proportion of the Total Energy Expenditure that
is Used for Walking

Ploughing Pulling 500 kg cart


medium soil on tarmac road
Time spent working (h) 5.5 5.5
Distance travelled (km) 11.6 19.5
Total energy used (MJ) 36.5 36.2
Energy used for doing work (MJ) 21.3 6.5
Energy used for doing work (%) 58.4 18.0
Energy used for walking (MJ) 14.4 24.2
Energy used for walking (%) 39.3 66.9
Adapted from Lawrence (1986).

llIF. Fat and Protein Deposition

Kellner's original feed evaluation system (see Chapter 9) expressed


the utilization of dietary energy in terms of its ability to produce
fat. However, the fat was' not actually measured and some of the
retained energy, even in mature animals, must have been protein.
Kotarbinska & Kielanowski (1967), using comparative slaughter
data, attempted to establish values for the cost of protein and fat
deposition in pigs. They assumed a constant maintenance need and
used a system of regression analysis. The results indicated that fat
deposition cost 13.45 kcal/g or 56.35 kJ /g. Assuming that fat
contains 39 kJ / g, this implies an efficiency for fat deposition of
69%. The same authors were also the first to obtain some
measurement of the cost of protein deposition, namely 11.03 kcaljg
or 46.2 kJ / g. Assuming that protein contains about 23 kJ / g, the
efficiency of deposition was about 50%. Other authors have since
confirmed that the synthesis of protein is generally less efficient
than that of fat.
The problem when using multiple regression to calculate the cost
of fat and protein deposition is that these two processes are usually
positively correlated. This invalidates the procedure. Moreover,
usually much more energy is stored as fat than as protein, so the
error associated with estimating the cost of the latter is relatively
much greater. 0rskov & McDonald (1970) carried out a very large
Energy metabolism of the host animal 75

slaughter experiment which avoided some of these difficulties.


Lambs were fed at different levels of intake and with different
proportions of protein in the diet. They were slaughtered at total
body weights of either 27 or 40 kg. There were large variations both
in number of days to slaughter and in the fat protein ratios of the
carcasses. It was possible to use mUltiple regression analysis,
entering into the equation the total metabolic weight-days (MWD,
i.e. mean body weight during the experiment, in kgWO. 75 , times no.
of days), and the energy retained respectively in fat and in protein.
The coefficients which gave the lowest residual standard error to
account for the intake of metabolic energy (MEl), expressed in kJ,
could then be defined. In this way the following equation was
derived:

MEl = 420MWD+3.03 x kJprotein + 1.26 X kJ rat


In other words, the average maintenance requirement was 420
kJ jkgWO. 75. As the energy content of protein is 23 kJ j g and the
efficiency of its deposition was 33%, its total cost was about 70
kJjg. Similarly, since the energy content of fat is 39 kJjg and the
efficiency of its deposition was 79%, its total cost was about 49
kJjg.
It is unlikely that the costs of protein and fat deposition remain
constant. More recent data on protein synthesis certainly suggest
that the turnover rate varies both with the stage of maturity and
with other factors (Reeds et ai., 1985). There is also some turnover
of fat, though the cost of this is likely to be much less important
than that of protein. Although the values obtained by different
workers vary, there is nevertheless general agreement that protein
deposition is relatively more expensive in terms of energy than fat
deposition.

IIIG. Pregnancy

The term 'conceptus' includes both the foetus and its associated
tissues, i.e. the placenta with its cotyledons, the amnion, amniotic
fluid, etc. The conceptus grows dramatically towards the end of
76 Energy Nutrition in Ruminants

pregnancy and its demand for energy is large. Robinson et al.


(1980) estimated that the efficiency of ME utilization for the
synthesis of conceptus tissues was 14.5% when the dam's diet
contained 10.5 MJ of ME/kg dry matter. This estimate was based
both on their own studies and on those of Graham (1964), Sykes &
Field (1972) and Rattray et al. (1973). Using this value they
calculated that the ME required per kg of lamb at term was 1.13
MJ/d. A ewe carrying twin lambs that will weigh 3.5 kg each at
birth needs 7.9 MJ/d for their growth. The maintenance re-
quirement of a 70 kg ewe is about 10 MJ /d (assuming a need of 420
kJ /kgW O' 75), so her total requirement at the end of her twin
pregnancy is about 1.8 times her maintenance level. The need. is
generally maximal shortly before parturition, but by that time the
foetuses restrict her gut volume which, in tum, limits her potential
intake of poor-quality roughages. Increased outflow partly com-
pensates for restricted gut volume (Ngongoni et al., 1987) but this
results in reduced digestibility. The practical consequence is that
either the animal must receive a diet which includes some
concentrate to permit an increased degradation rate, or she will
support the pregnancy by using her own body fat and protein. Since
the main energy source for the foetus is glucose, it is important that
some of the feeds consumed can generate both protein and glucose.
Pregnancy toxaemia is the classic symptom of glucose deficiency in
underfed ewes that are carrying a heavy foetal burden.

IIIH. Lactation

Just as the measurement of protein utilization in dairy cows is


difficult (0rskov, 1982b), so also is that of energy utilization. This
is because the animals are often either in negative or in positive
energy balance, i.e. losing or gaining body tissue and simultaneously
yielding energy in milk. The work from the Energy Metabolism
Laboratory at Beltsville, U.S.A., is outstanding in this area (Moe et
al., 1970). These workers submitted data from over 500 balance
trials to a regression analysis similar to that used by 0rskov &
McDonald (1970). They arrived at the following very important
Energy metabolism of the host animal 77

values: (i) a maintenance requirement of 457 kJ /kgW O. 75; (ii) an


efficiency with which ME was used for milk production of 61.6%;
and (iii) an efficiency of tissue loss to support lactation of 81.0%.
The efficiency of energy deposition in dry cows was 59.2%, which
is higher than generally reported elsewere. When 0rskov et al.
(1969) infused VFA into the rumens of lactating cows kept in
respiration chambers, they obtained a value for the efficiency of
ME utilization for milk production of 78%. The proportions of
acetic and propionic acids did not affect this value. Whitelaw et al.
(1986), having adjusted their data to zero change in body tissue,
estimated an efficiency of 84%. There is as yet no information
regarding the relationship between milk -composition and the
efficiency of ME use for lactation. In general, metabolizable energy
is used more efficiently for lactation than for fattening and tissue
gain. The Agricultural Research Council (1980) estimated the
energy cost of energy retention to be about 5% more than for
lactation. It is likely that the greater efficiency of milk production
results from the absence of turnover, since there is none in the milk
secreted into the alveoli, while turnover continues in deposited
protein and fat.

III I. Wool and Fibre Production

Apart from associated waxy oils, wool, hair, hom etc. are virtually
100% protein. However, even in specially selected animals, wool
production constitutes only a very small amount of retained energy
compared with the maintenance energy requirement. As a result
there are no good estimates of the energy cost of production of
wool etc. It is likely to be more efficient than protein deposition in
active tissues since this dead n1aterial undergoes no turnover. If a
70 kg ewe, selected for rapid wool growth, produces 30 g ofwool/d,
this amounts to a net energy increment of 0.71 MJ/d. Assuming
30% efficiency for the process, there is an additional ME
requirement of 2.4 MJ/d, resulting in a total energy need 1.25 times
the maintenance requirement of 9.7 MJ/d. This is probably an
extreme estimate since it was based on an assumed efficiency similar
78 Energy Nutrition in Ruminants

to tissue protein deposition. For animals not selected for fibre yield
the energy required is likely to be substantially less, the total
amounting to no more than 1.1 times the maintenance energy need.
It should be noted, however, that since wool and fibre growth
proceed during the determination of the maintenance energy need,
the cost is normally included in that value.

llIJ. Keeping Warm and Keeping Cool

i. Low Environmental Temperatures


So far, when discussing fasting metabolism and resting metabolic
rate, it has been assumed that the animals are kept at an
environmental temperature above a certain critical level. This
critical temperature is defined as that below which the animal must
increase heat production in order to maintain normal body
temperature. Blaxter (1962) published what is probably the most
detailed account of this topic. Here it will be discussed only briefly.
Jhere is very little quantitative evidence on the increased feed
requirements which arise when the environmental temperature falls
below the critical level. There are many reasons for this. In
particular, the animals may be able to adapt in some way to the new
environmental stress. It will be recalled that the efficiency of
ruminants' utilization of pure nutrients, such as VF As, seldom
exceeds about 60%. Metabolizable energy from normal feeds is
generally used less efficiently than that, some being lost via the heat
increment of feeding and the energy costs of eating and ruminating.
Thus considerable amounts of heat are generated and the critical
temperature at which heat production must be increased is low.
Blaxter's (1962) results provide a good illustration. The critical
temperature of steers with normal hair coats was 18 D e when
fasting, 7D e on a maintenance level of feeding and - 1De when they
were fed to gain 500 gjd. It is of practical importance to recognize
this effect of feeding level. For example Webster et al. (1970)
showed that grain-fed calves, which grew at a rate of about 1 kgjd
at 20 D e, grew almost equally well in outside temperatures of
- 20 D e if they were sheltered, though wind exposure slightly
reduced their performance. These authors also showed that for
Energy metabolism of the host animal 79
well-fed steers adapted to a cold environment the critical tem-
perature was -lS°C. The adaptation included less shedding of hair
during the cold season.
Real problems can, however, occur when limited feed is combined
with high wind speeds and rain, a combination which is not
unusual. Some breeds of animals adapt to a cold environment
better than others, probably as a result of selection. The indigenous
cattle in the Scottish Highlands have much longer winter coats than
do the breeds developed in the south of Britain. Blaxter (1962)
demonstrated the effect of coat length on the critical temperature of
sheep at maintenance level offeeding. For wool lengths of 1, lO, 50
and 100 mm it was 2S, 22, 9 and - 3°C respectively.
When the environmental temperature falls below the critical level
the animal's first reaction is generally a shivering thermogenesis.
The muscles near the skin generate heat by their rhythmic
contraction. Later so-called non-shivering thermogenesis occurs, in
which heat is produced unaccompanied by this muscular activity.
The energy source utilized during cold exposure generally
consists offat, not protein (Blaxter, 1962). The efficiency of this fat
utilization must of course be very close to lOO%. Nevertheless,
there are problems in quantitatively defining, in the field,
temperature, humidity and wind speed in the animal's microclimate,
type of hair coat and, not least, feed intake. Consequently it is very
difficult to establish the amount of additional feed required to
prevent excessive fat loss in adverse temperatures. Suffice it to say
that crises can arise rapidly if undernutrition, low environmental
temperatures, high winds and rain or sleet coincide, whereas well-
fed ruminants can withstand severe climatic conditions with little
loss of performance. During pregnancy the effects of combined
undernutrition and environmental stress are further aggravated
because of the glucose requirement for foetal growth: pregnancy
toxaemia will often occur.

ii. High Environmental Temperatures


In many parts of the world there are more problems of high than
of low temperatures and there are breeds of ruminants which have
adapted accordingly. The difficulties associated with introducing
80 Energy Nutrition in Ruminants

exotic cattle from temperate regions to the tropics are well known.
When the temperature of the environment equals or exceeds that of
the body, surplus heat must be eliminated by vaporization of water
from the skin and the respiratory tract. The animals usually
respond to high environmental temperatures by reducing activity
and feed intake, so reducing the obligatory loss of heat. This was
well demonstrated in a series of trials carried out in Missouri
(Kibler & Brody, 1956; Johnson et ai., 1958; Johnson & Yeck,
1964). The invariable consequence when the external temperature
rose towards that of the body was a reduction in feed intake. As a
result, productivity, including milk yield and growth rate, declined.
If high temperatures are combined with high humidity, so redUGing
evaporation, the problem obviously worsens and a crisis can occur.
Usually the body temperature rises, facilitating heat dispersal, but
ifit exceeds about 40°C the metabolic rate also increases. Tolerance
of raised body temperatures is limited for most domestic animals.
Again, there are adaptations to hot environments. Thus
Australian Merino sheep withstand high external temperatures
better than Hampshires or Southdowns, and show smaller increases
in body temperature (Miller & Monge, 1946). Camels, on the other
hand, tolerate wide fluctuations of the body temperature - from
34 to 41°C - without adverse physiological effects. The tendencies
for ruminants in hot, arid zones to have pale-coloured hair and skin
and to store fat in humps or tails rather than subcutaneously, may
be related to temperature control. Other behavioural and physio-
logical adaptations, such as nocturnal feeding and the production
of concentrated urine and dry faeces, reduce heat uptake and
conserve water needed for evaporative cooling (Schmidt-Nielsen,
1964; Yousef, 1987).

REFERENCES

Adam, I., Young, B.A., Nicol, A.M. & Degen, A.A. (1984) Energy cost of eating
in cattle given diets of different form Anim. Prod. 38, 53-56
Agricultural Research Council (1980) The Nutrient Requirements of Farm
Livestock No.2. Ruminants 2nd Edition Commonwealth Agricultural Bureaux,
Slough
Energy metabolism of the host animal 81
Asplund, J.M., 0rskov, E.R., Hovell, F.D. de B. & MacLeod, N.A. (1985) The
effect of intragastric infusion of glucose, lipids or acetate on fasting N
excretion and blood metabolites in sheep Brit. J. Nutr. 54, 189-195
Blaxter, K.L. (1962) The Energy Metabolism of Ruminants Hutchinson Scientific
& Technical, London
Graham, N.McC. (1964) Energy cost offeeding activities and energy expenditure
of grazing sheep Austr. J. agric. Res. 15,969-973
Holmes, C.W., Stephens, D.B. & Toner, J.N. (1976) Heart rate as a possible
indicator of the energy metabolism of calves kept out-of-doors Livestock Prod.
Sci. 3, 333-341
Holmes, C.W., McLean, N.A. & Lockyer, K.J. (1978) Changes in the rate of heat
production of calves during grazing and eating N. Z. J. agric. Res. 21, 107-112
Johnson, H.D. & Yeck, R.G. (1964) Environmental physiology and shelter
engineering. LXVIII. Age and temperature effects on TDN, water consumption
and balance of dairy cows and heifers exposed to temperatures of 35 to 90°F
Research Bulletin No. 865 University of Missouri
Johnson, H.D., Ragdale, A.C. & Yeck, R.G. (1958) Environmental physiology
and shelter engineering. LIX. Effect of constant environmental temperatures of
50° and 80°F on water and feed consumption of Brahman, Santa Gertrudis and
Shorthorn cattle Research Bulletin No. 683, University of Missouri
Kibler, H.H. & Brody, S. (1956) Environmental physiology and shelter
engineering. VIII. Influence of diurnal temperature cycles on heat production
and cardiorespiratory activities in Holstein and Jersey cows Research Bulletin
No. 601 University of Missouri
Koong, L.J., Ferrell, c.L. & Nienaber, J.A. (1985) Assessment of inter-
relationships among levels of intake and production, organ size and fasting heat
production in growing animals J. Nutr. 115, 1383-1390
Kotarbinska, M. & Kielanowski, J. (1967) Energy balance studies with growing
pigs by the comparative slaughter technique In Blaxter, K.L., Kielanowski, J.
& Thorbek, G. (Eds) Energy Metabolism of Farm Animals pp 299-310 E.A.A.P.
Publication No. 12 Oriel Press, Newcastle upon Tyne
KuVera, J.C., 0rskov, E.R. & MacLeod, N.A. (1988) Energy exchanges in cattle
nourished by intragastric nutrition In Van der Honig, Y. (Ed.) Proc. E.A.A.P
Energy Metabolism Symposium pp 271-274, PUdok, Wageningen
Lawrence, P.R. (1986) A review of the nutrient requirement of draught oxen In
Copland, J. (Ed.) Draught Animal Power for Production pp 58-63 A.C.I.A.R.
Proceedings Series No. 10, A.C.I.A.R., Canberra
Miller, J.C. & Monge, L. (1946) Body temperature and respiration rate, and their
relation to adaptability in sheep J. Anim. Sci. 5, 147-153
Moe, P.W., Tyrrell, H.F. & Flatt, W.P. (1970) Partial efficiency of energy use for
maintenance, lactation, body gain and gestation in the dairy cow In Schurch, A.
& Wenk, G. (Eds) Energy Metabolism in Farm Animals pp 65-68 E.A.A.P.
Publication No. 13 Juris Druck & Verlag, Zurich
Mould, F.L., Saadullah, M., Haque, M., Davis, D., Dolberg, F. & 0rskov, E.R.
(1982) Investigation of some of the physiological factors influencing intake and
digestion of rice straw by native cattle in Bangladesh Trop. Anim. Prod. 7,
174-181
Ngongoni, N.T., Robinson, J.J., Kay, R.N.B., Stephenson, R.G.A., Atkinson, T.
82 Energy Nutrition in Ruminants
& Grant, I. (1987) The effect of altering the hormone status of ewes on the
outflow rate of protein supplements from the rumen and so on protein
degradability Anim. Prod. 44, 395-404
0rskov, E.R. (1982a) Maintenance and growth in ruminants. Introductory
comments In Ekern, A. & Sundst01, F. (Eds) Energy Metabolism of Farm
Animals pp 141-146 The Agricultural University of Norway, Aas
0rskov, E.R. (1982b) Protein Nutrition in Ruminants Academic Press, London
0rskov, E.R. & McDonald, I. (1970) The utilization of dietary energy for
maintenance and for protein and fat deposition in young growing sheep In
Schiirch, A. & Wenk, G. (Eds) Energy Metabolism in Farm Animals pp 121-125
E.A.A.P. Publication No. l3 Juris Druck & Verlag, Zurich
0rskov, E.R., Flatt, W.P., Moe, P.W., Munson, P.W., Henken, R.W. & Katy, I.
(1969) The influence of ruminal infusion of volatile fatty acids on milk yield and
composition and energy utilization by lactating cows. Brit. J. Nutr. 23, 443-463
0rskov, E.R., Duncan, W.R.H. & Carnie, C.A. (1975) Cereal processing and food
utilization in sheep. 3. The effect of replacing whole barley by whole oats on
food utilization and firmness of subcutaneous fat in sheep Anim. Prod. 21, 51-58
Pearson, A.R. (In press) Reduced output of well-fed buffaloes carting loads on the
Terai in East Nepal Trop. Anim. Prod. & Health
Rattray, P.V., Garrett, W.N., East, N.E. & Henman, N. (1973) Net energy
requirements of ewe lambs for maintenance, gain and pregnancy and net energy
values of feedstuffs for lambs J. Anim. Sci. 37, 853-857
Reeds, P.J., Nicholson, B.A. & Fuller, M.F. (1985) Contribution of protein
synthesis to energy expenditure in vivo and in vitro In Moe, P.W., Tyrell, H.F.
& Reynolds, P.J. (Eds) Energy Metabolism of Farm Animals pp 6-9 E.A.A.P.
Publication No. 32 Rowman & Littlefield, NJ
Robinson, J.J., McDonald, I., Fraser, C. & Gordon, J.G. (1980) Studies on
reproduction in prolific ewes. 6. The efficiency of energy utilization for
conceptus growth J. agric. Sci., Camb. 94, 333-338
Rubner, M. (1902) Die Gesetz des Energieverbrauchs bei die Erhniirung (Quoted by
Blaxter, 1962)
Schiemann, R. (1958) Kritische Betragungen iiber den Entwicklung der starke-
wertlehre Oscar Kellner Dtsch. Akad. Landwirt. Wiss. Ab. No. 31
Schmidt-Nielsen, K. (1964) Desert Animals Oxford University Press, Oxford
Summers, M., McBride, B.W. & Milligan, L.P. (1988) Components of basal
energy expenditure In Dobson, A. & Dobson, M.J. (Eds) Aspects of Digestive
Physiology in Ruminants pp 257-285 Cornell University Press, Ithaca, NY
Sykes, A.R. & Field, A.c. (1972) Effects of dietary deficiencies of energy, protein
and calcium on the pregnant ewe III. Some observations on the use of
biochemical parameters in controlling energy undernutrition during pregnancy
and on the efficiency of utilization of energy and protein for foetal growth J.
agric. Sci., Camb. 87, 127-l33
Wainman, F.W., Blaxter, K.L. & Smith, J.S. (1972) The utilization of the energy
of artificially dried grass prepared in different ways J. agric. Sci., Camb. 78,
441-447
Webster, A.J.F., Chlumecky, J. & Young, B.A. (1970) Effect of cold environments
on the energy exchanges of young beef cattle Canad. J. Anim. Sci. 50, 89-100
Webster, A.J.F., Brockway, J.M. & Smith, J.S. (1974) Prediction of the energy
Energy metabolism of the host animal 83
requirements for growth in beef cattle 1. The irrelevance of fasting metabolism
Anim. Prod. 19, 127-139
Whitelaw, F.G., Milne, J.S., 0rskov, E.R. & Smith, J.S. (1986) The nitrogen and
energy metabolism of lactating cows given abomasal infusions of casein Brit. J.
Nutr. 55, 537-556
Yousef, M.K. (1987) Principles of bioclimatology and adaptation In Johnson,
H.D. (Ed.) Bioclimatology and the Adaptation of Livestock World Animal
Science Series B5 pp 17-31 Elsevier, Amsterdam
CHAPTER 7

UTILIZATION OF THE ENERGY OF


ABSORBED NUTRIENTS

I. Volatile fatty acids


II. Glucose
A. Sources of glucose
B. Glucose requirement for growth
C. Glucose requirement for lactation
D. Glucose requirement during pregnancy
III. Body reserves
A. Use of body reserves during fasting
B. Use of body reserves to fuel protein deposition
C. Use of body fat to support lactation
IV. Composition of nutrients and endocrine changes

I. VOLATILE FATTY ACIDS

In view of the importance of volatile fatty acids (VFAs) as a source


of energy for ruminants, this section will emphasize their
significance. Since the Cambridge team, led by Sir John Barcroft,
84
Use of energy of absorbed nutrients 85
established that VF As were in fact the main source of energy for
these animals, much effort has been devoted to investigating their
role in detail. This has led to many interesting debates. One of these
was related to the finding that metabolizable energy (ME) from
long, poor-quality roughages was utilized less efficiently than that
from concentrate feeds. Since roughages tend to yield relatively
more acetic acid and concentrates relatively more propionic acid
(see Chapter 3), one hypothesis was that the amount of acetic acid
produced was causally related to poor ME utilization.
This hypothesis was tested some years ago by Armstrong and his
colleagues at the Hannah Research Institute (Armstrong & Blaxter,
1957 a,b; Armstrong et al., 1957, 1958). They observed that
different VF As infused into the rumens of fasting sheep were
utilized with equal efficiency. On the other hand, when the sheep
were in positive energy balance and the infused VF As were added
to a basal diet of hay, there were differences in their utilization. The
observed values were 33, 57 and 62% for acetic, propionic and
butyric acid respectively. They also attempted to use VF A mixtures
with high and low proportions of acetic acid. Here, however, the
results were much less conclusive, a mixture high in acetic acid
being utilized less efficiently and a mixture low in acetic acid being
utilized more efficiently than expected (see 0rskov et al., 1979).
Using a different approach, 0rskov & Allen (1966 a, b, c) and
0rskov et al. (1966) were unable to confirm Armstrong et al.'s
(1957, 1958) observations when they fed VFA salts to groups of
growing lambs. Moreover, they found no differential effects of
acetic and propionic acid on heat production by dairy cows,
although milk production was affected (0rskov et al., 1969). Later,
Tyrrell et al. (1979) observed that acetic acid was used more
efficiently when infused into the rumen of cows receiving a high-
concentrate diet than when infused similarly into cows receiving a
high-roughage diet. The VF A contributed 15% of the ME of each
diet. Although acetic acid generally constitutes the largest molar
proportion of the total VF A (see Chapter 3), it contributes a
<;onsiderably smaller proportion of the total energy derived from
nutrients, due to its relatively low calorific value. Indeed, the energy
contribution of acetic acid is far less than is often expected. Table
86 Energy Nutrition in Ruminants

7.1 lists a series of mixtures with different molar proportions of the


VFAs together with their contributions to the energy value of each
mixture. It will be appreciated that the addition of 15% ME as
acetic acid to a basal diet already yielding much of this acid through
fermentation will result in a molar proportion well above
physiological levels. Interpretation of experimental results must
take this into account. Any effect on the utilization of a basal diet
which follows the addition of supplementary VF A will tend to be
interpreted as a result of the direct effect of the supplement but this
is not necessarily the case.
The development of the technique of total intragastric nutrition
(0rskov et al., 1979) made definitive work in this field possible.
Ruminants could be maintained exclusively on intra-ruminal
infusions of VF A and a buffer solution which stabilized rumen pH
a.nd osmotic pressure within normal limits. At the same time,
different levels of proteins could be infused into the abomasum. A
wide range of ratios of the different VF As were compared directly
in terms of heat production. In general, this was measured both at
the estimated level of VFA intake required for energy maintenance
and at twice that level. However, the mixture which contained 85%
acetic acid was given at 1 and 1.5 times maintenance level, because
with larger amounts it was difficult to maintain normal rumen pH.
The results presented in Table 7.2 show clearly that the large
differences in utilization between roughages and concentrates
cannot be explained by differences in the molar proportions of the
VF As. There were no consistent differences in the utilization of the
different mixtures. Table 7.2 also includes estimates of utilization
based on the coefficients derived by Armstrong et al. (1957, 1958)
from their studies. As acetic acid contributes a relatively small
proportion of the energy (see Table 7.1), the differences, even
considering the finding of a relatively poor utilization of acetic acid,
were not very large. They were certainly not large enough to explain
the differences sometimes reported between fibrous roughages and
concentrates. The explanation for these must be sought elsewhere,
such as in activities associated with standing, eating and rumination
(see Chapter 6). Recent results (0rskov & MacLeod, 1990) showed
that when the capacity for acetic acid utilization was exceeded-
Use of energy of absorbed nutrients 87
TABLE 7.1
Volatile Fatty Acids in Mixtures Expressed as Molar % and as Percent of Total
Energy

Molar %
Acetic acid 35 45 55 65 75 85
Propionic acid 55 45 35 25 15 5
Butyric acid 10 10 10 10 10 10
% of energy
Acetic acid 22 30 39 48 59 72
Propionic acid 62 53 43 33 21 7
Butyric acid 16 17 18 19 20 21
Adapted from 0rskov et al. (1979).

TABLE 7.2
Efficiency of Utilization of Different Mixtures of Volatile Fatty Acids for
Fattening (Kt), Measured during Intragastric Nutrition and as Predicted from
Armstrong & Blaxter's (1957b) results

Molar % Kf Kf

Acetic Propionic Butyric Intragastric Armstrong


acid acid acid nutrition & Blaxter

45 45 10 0.64 0.50
55 35 10 0.57 0.48
65 25 10 0.61 0.46
75 15 10 0.61 0.44
From 0rskov et al. (1979).

well above physiological proportions - heat production decreased


and acetic acid was excreted in urine. Thus, inefficient oxidation of
acetic acid could not be induced.
As mentioned earlier, infusion of acetic acid into the rumens of
dairy cows affects the composition of the milk differently from
infusions of propionic acid, whereas heat production is unaffected
by the type of VFA (0rskov et al., 1969). This has now been
confirmed in dairy cows maintained solely by intragastric infusion
(0rskov & MacLeod, 1982). If the absorption of propionic acid
exceeds the capacity of the liver to handle it, some enters the
peripheral circulation and stimulates the pancreas to secrete insulin,
88 Energy Nutrition in Ruminants

TABLE 7.3
Effect of Molar Proportions of Volatile Fatty Acids on Glucogenic Energy,
Expressed as Percent of Total Energy in the Mixture

Molar % Glucogenic
energy
Acetic Propionic Butyric (%)
acid acid acid
45 45 10 53
55 35 10 48
65 25 10 36
75 15 10 21
From 0rskov (1980).

so leading to" the utilization of energy for tissue synthesis and to


depression of the cow's milk yield and milk fat percentage.
It is well known that acetic acid cannot give rise to a net synthesis
of glucose. Acetate utilization depends on a source of three-carbon
units, such as propionic acid or glucose, which can generate the
oxaloacetic acid needed to sustain oxidation via the Krebs citric
acid pathway. However, only a small amount of oxaloacetic acid is
degraded in this process, most being regenerated, so the re-
quirement for glucose precursors for this purpose is relatively low.
With regard to fat synthesis, the enzyme NADPH is necessary
for the production of long-chain fatty acids from two-carbon units.
These are provided by VFA and are also generated from glucose by
oxidation via the pentose phosphate pathway. In the lactating
animal the need for glucose precursors is great, since they are
required both to provide NADPH for the synthesis of fat and for
the synthesis of lactose.

D.GLUCOSE

DA. Sources of Glucose

There is essentially no disagreement about the pathways of


oxidation of the different VF As. In Table 7.3 (from 0rskov, 1980)
it was assumed that only propionic acid served as a direct source of
Use of energy of absorbed nutrients 89

glucose. The glucogenic energy in the mixtures of VFAs varied


from about 21 to 53% of the total energy. However, there are also
other sources of glucose, in particular those microbial amino acids
- about 20% of the total - that are sub-optimal for ruminant
protein synthesis (Storm et aI., 1983). Some of these amino acids
are glucogenic and available as a source of three-carbon units in
place of glucose. In addition, amino acid lost in tissue turnover
amounts to about 350 mg N/kgWO. 75 per day, which is equivalent
to approximately 15% of the metabolic energy required for
maintenance. The efficiency with which amino acids yield glucose
precursors is not fully understood, but some authors estimate that
it can amount to 50% of their energy.
Early infusion work (Armstrong & Blaxter, 1957a) showed no
differences in heat production between mixtures of VF As in which
from 100% down to 9.9% of the energy was derived from
propionic acid, which probably implies that even less is needed.
Indeed 0rskov et al. (1979) used a mixture containing only 7.7% of
the glucogenic energy as propionic acid. More recently (E.R.
0rskov and N.A. MacLeod, unpublished), almost 100% acetic acid
was used, and no metabolic problems were found, at the
maintenance level of energy feeding. Table 7.3 shows clearly that
there is little or no possibility of glucogenic precursors limiting
metabolism at a low level of intake, because of the low requirement.
When animals are fed well below maintenance, fat is oxidized.
Consequently it is of interest to calculate the glucogenic energy in
triglycerides. The glycerol represents only 4-5% of the total
triglyceride energy. This implies that in the fasting animal, were it
to use only fat, the supply of glucose precursors would possibly be
limiting. In consequence, more amino acids are oxidized rather
than re-incorporated into protein in the process of tissue turnover,
so providing additional glucose precursor. It is significant that
intragastric nutrition studies with sheep and cattle showed that
nitrogen excretion in fasting animals is always about 40% greater
than in animals receiving sufficient VFA to meet their estimated
~aintenance energy requirement (see Fig. 6.i). It might be
concluded that additional protein is being oxidized in the fasting
animal, in order to supply glucose precursors; yet energy from VF A
90 Energy Nutrition in Ruminants

TABLE 7.4
Effect of Level of Production on the Theoretical Requirement for Glucogenic
Energy for Lambs and Steers

Species Live Level of ME Requirement for


wt. production requirement glucogenic energy
(kg) (g live wt. (MIld) (% ME)
gainld)
Lambs 20 0 3.9 9.9
20 300 9.4 21.6
20 600 14.9 24.6
40 300 14.9 21.2
40 600 23.2 24.2
Steers 200 negative a 11.1 9.9
200 0 22.2 9.9
200 1000 42.7 19.6
200 2000 70.8 23.6
400 0 37.4 9.9
400 1000 68.6 18.8
400 2000 110.9 23.2

From 0rskov (1980).


aRations 0.5 x maintenance level.

sufficient for only about 20-30% of the estimated maintenance


requirement reduced nitrogen excretion to near basal levels
(KuVera et al., 1988). Calculation of the amount of glucose
precursors in the VF A mixtures used indicates that the additional
quantity needed by fasting sheep, for example, amounts to only
20-40 g/ d. It should be noted that, as expected, glucose deficiency
is greatest when body fat supplies a major part of the energy, i.e.
when the intake is one-third of that needed for maintenance, or less.

lIB. Glucose Requirement for Growth

An estimate can be made of the glucose requirement during growth,


in particular for fat synthesis. Hovell (1972) calculated that 8 mol
of three-carbon compounds were required for the production of 1
mol of triglyceride, using a total of 32.9 mol of two-carbon units.
This amounts to a requirement of 29.9% of glucogenic energy when
fat is synthesized. A similar or probably lower amount is needed for
Use of energy of absorbed nutrients 91

protein synthesis. On the basis of these assumptions, 0rskov (1980)


calculated the glucose requirements of growing Suffolk lambs and
Hereford steers at different theoretical growth rates and at different
stages of maturity (Table 7.4). It is significant that the estimated
glucose or three-carbon requirement, even at growth rates of 600
gj d in lambs and 2 kgj d in steers, is less than that yielded by any
of the VFA mixtures given in Table 7.2, except the mixture
containing 75 molar % of acetic acid. This suggests that adequately
fed growing animals are unlikely ever to encounter the problem of
deficiency of glucose precursors. Nevertheless, this conclusion has
been challenged by Leng & Preston (1976) and Preston & Leng
(1986), who argued that glucose deficiency Gan be a constraint in
animals consuming low-quality roughages.

lIe. Glucose Requirement for Lactation

An attempt can be made to calculate, as for maintenance and


growth, the glucogenic energy required for lactation. Here there are
additional demands for glucogenic precursors, namely for the
lactose secreted in the milk and for the NADPH needed to fuel milk
fat synthesis. Assuming that the glucose required for maintenance
is the same as in growing animals, Table 7.5 illustrates the
theoretical requirement for glucogenic energy at different milk
yields. It can be seen, by comparison with Table 7.2, that when the
animals obtain sufficient energy from their diet to meet the energy
needs of lactation, the glucogenic energy generated is likely to be
sufficient for milk production. In contrast, if it is assumed, as in the
last two lines of Table 7.5, that the animals' requirements are not
met and that the production of 10-20 kg milkjd is being fuelled by
body tissue breakdown, then the glucogenic energy available from
the feed may well be insufficient. In that case, the increased need for
it in early lactation can lead to a real possibility of glucose
deficiency, with resultant acetonaemia.
, It has recently been observed (0rskov et ai., 1987) that milk
production in underfed cows can be stimulated by feeding various
amounts of fishmeal (Table 7.6). At the highest levels of fishmeal
92 Energy Nutrition in Ruminants

TABLE 7.5
Effect of Level of Milk Production and the Extent of Negative Energy Balance
on the Need for Glucogenic Energy in Cows Weighing 600 kg

Level of ME ME Glucogenic energy


production requirement intake requirement
(kg FCMa/d) (MJ/d) (MJ/d) (% ME)

0 61.9 61.9 9.9


10 110.5 110.5 22.0
20 159.1 159.1 26.6
40 256.3 256.3 30.7
40 256.3 207.8 34.3
60 353.5 353.5 32.6
60 353.5 256.3 39.0

From 0rskov (1980).


aFat corrected milk.

intake the cows were consuming only the dietary energy required to
produce about 15 kg milk/d, yet they produced 35 kg/d. Most of
the cows on that treatment developed acetonaemia. However, cows
which received less fishmeal, and were consequently in slightly less
negative energy balance, showed no signs of acetonaemia although
they supported the production of at least 10 kg milk/d by utilizing
body tissue for periods of 3 months. This experiment provided
some confirmation that, even for milk production, available glucose
precursors are seldom a constraint. Only in the exceptional
circumstances illustrated in Table 7.5, where high-yielding animals
stop eating or are prevented from eating, is glucose deficiency likely
to occur.

fiD. Glucose Requirement During Pregnancy

The main source of energy for the foetus is glueose, or its


precursors. Consequently, animals carrying large foetal loads are
susceptible to glucose deficiency. The greater the load, the greater
is the susceptibility. In sheep the term 'twin lamb disease', to denote
pregnancy toxaemia, is highly explicit. As in lactation, a deficiency
is likely to occur only when energy needs are not supplied by the
Use of energy of absorbed nutrients 93

TABLE 7.6
Fish Meal and Dry Matter Intake, Milk Yield and Composition and
Calculated Energy Deficit During Week 2 of Test Period

Fish Dry Milk FCMa Milk composition Calculated


meal matter yield yield ME deficit
(g/kg) intake (kg/d) (kg/d) Fat N (MJ/d)
feed (kg/d) (g/kg) (g/kg)
0 12.2 23.6 28.5 54.0 5.90 -77
40 13.8 26.1 29.2 48.1 5.93 -64
80 13.8 27.1 32.0 52.0 5.97 -80
120 13.6 28.5 34.9 54.8 5.84 -97
SE of 0.6 1.8 1.8 3.0 1.36
means
From 0rskov et al. (1987).
Each value is the mean of eight observations.
aFat corrected milk.

feed and stored body fat is used. This condition can also develop
when the feed is of such poor quality that the animal's needs are not
met, and it is exacerbated in later pregnancy, when rumen volume
is restricted by the volume of the uterus. Pregnancy toxaemia can
also occur as a result of a stressful environment, for example when
sheep are reluctant or unable to seek food during and after
snowstorms. The condition can be cured either by injecting glucose
intravenously or by drenching with a suitable precursor, such as
propionic acid, which is not susceptible to bacterial degradation.

III. BODY RESERVES

Ruminant body reserves are extensively utilized and subsequently


replenished, especially by those animals which live in arid or cool
temperate zones with highly seasonal nutrient supplies. Yet there is
relatively little information on the efficiency with which these
reserves are used. They consist mainly of triglycerides stored in
various parts of the body, including perineal fat, the tails of sheep
and the humps of zebu cattle and camels.
94 Energy Nutrition in Ruminants

llIA. Use of Body Reserves During Fasting

It is very difficult to measure the efficiency of utilization of body


reserves. However, fasting metabolism can be measured and it can
be argued that the relative difference in heat production between
the states of fasting and of energy equilibrium provides a measure
of the efficiency of utilization of body reserves relative to dietary
ME, i.e. fasting metabolism is measured against the nutrients used
to supply maintenance energy. Nevertheless Armstrong & Blaxter
(1957a) found no increase in heat production when sufficient
glucose to supply maintenance energy was infused into the
abomasum of fasting animals. Recently KuVera et al. (1988) m,ade
similar observations (see Chapter 6). They also found that while
glucose supplied at energy maintenance levels did not augment heat
production, a mixture ofVFAs did. This suggests that the efficiency
with which exogenous, absorbed VF A energy is utilized is lower
than that for body fat. On the other hand the efficiency with which
glucose is used for oxidation is similar to that of body fat. It is
relevant that the Agricultural Research Council (1980) defines the
maintenance need as 1.35 times the fasting metabolism, indicating
that 1 kJ of body fat is approximately equal to 1.35 kJ of dietary
ME. It would be of interest to determine whether absorbed;
exogenous fat is utilized with similar efficiency. However, as
discussed in Chapter 6, body reserve utilization during fasting may
not be maximally efficient. It is well known that in blood from
fasting animals the concentrations of p-hydroxybutyrate and free
fatty acids are high, suggesting a sluggish metabolism of the body
reserves. This can, however, be corrected by a small supply of
glucose.
It is considered that the increase in nitrogen excretion during
fasting indicates a lack ·of glucose precursors. These are being
supplied from the breakdown of proteins while they are sim-
ultaneously required for amino acid salvage during protein
turnover. Nitrogen excretion responds very clearly to small
amounts of glucose. One might speculate whether fasting heat
production could be reduced by supplying small amounts of
glucose precursors to assist in adequate utilization of body fat.
Use of energy of absorbed nutrients 95

Such a hypothesis is difficult to confirm but it is important in


relation to the concept that metabolizable energy is used with
different efficiencies in different nutritional states, as discussed in
Chapter 6.

IIIB. Use of Body Reserves to Fuel Protein Deposition

The extent to which accumulated body fat can serve to fuel growth
has only recently been recognized. Intragastric nutrition studies
showed that fasting sheep which received only the required protein,
by abomasal infusion, attained positive protein balance. Moreover,
the efficiency with which the dietary protein- was utilized equalled
that in well-nourished sheep and cattle (0rskov et al., 1983; Hovell
et al., 1983). Fattet et al. (1984) carried out a comparative slaughter
experiment and measured the extent to which lambs could use body
fat to support lean growth. One group received restricted amounts
of a straw diet, sufficient for only half the maintenance energy
requirement, plus rumen-undegradable protein (fishmeal). Another
group received straw almost sufficient for energy maintenance, plus
fishmeal. The extent to which the lambs were able to utilize body fat
to fuel protein deposition is shown in Table 7.7. The efficiency of
this process is not known, but the results open up new avenues of
research, since more information is needed to compare utilization
of body fat with that of energy sources stored by other means. Thus
it may be more efficient to make maximal grazing use of range land
for storage of body fat while the feed quality is good - as achieved
by freely grazing wild and domestic animals - rather than to
conserve the grass crop as medium-quality hay or silage. Optimal
management will probably involve both policies but the strategic
use and manipulation of body reserves to support growth may well
add flexibility to ruminant production systems.

IIIC. Use of Body Fat to Support Lactation

While the use of body fat to support lactation has been generally
recognized, its potential importance was not appreciated before the
studies of Flatt and his co-workers at Beltsville (Flatt et al., 1965).
96 Energy Nutrition in Ruminants
TABLE 7.7
Effect of Feeding Straw and Fish Meal on Changes in Body Composition of Lambs
during a 92 day Feeding Period

Treatments: Straw: Low Low High High


Fish meal: + +
Energy intake (kJ /kgWO' 75 / d) 235 307 456 488
Change in live wt. (kg) -4.32 0.29 0.08 6.22
Change in empty body wt. (kg) -5.05 0.64 -0.80 4.18
Change in fat content (kg) -3.53 -1.53 -1.40 -0.93
Change in protein content (kg) -0.47 0.48 -0.14 0.89
From Fattet et al. (1984).

TABLE 7.8
Effects of Abomasal Infusion of Glucose and Casein on Milk Yield, Milk
Composition and Energy Deficit of Cows in Early Lactation

Casein Glucose Milk Milk FCMa Milk Net energy


(g/d) (g/d) yield fat (kg/d) protein deficit
(kg/d) (g/kg) (g/kg) (MJ/d)
0 750 16.8 48.2 18.9 25.2 20.5
250 500 19.8 49.8 22.7 28.4 30.5
500 250 21.6 51.0 25.2 29.6 38.6
750 0 21.4 54.0 26.1 31.5 41.0
SE of means 0.8 1.2 0.6
From 0rskov et al. (1977).
aFat corrected milk.

They described work with a cow, Lorna, which lost 70-105 MJjd,
equivalent to a daily loss of 2-3 kg of body fat. The efficiency with
which body reserves are used to support lactation is not well
understood. It is difficult to measure, due to the confounding effects
of the dietary energy supply and of the maintenance need. Moe et
al. (1970) calculated an efficiency of 85% for the use of body
reserves, compared with an efficiency of 60% when dietary
metabolizable energy was used to support lactation. The former
probably arises in part from the extent to which lipids from body
fat are incorporated directly into milk fat. Little information is
available on this point but 0rskov et al. (1969) showed that the
Use of energy of absorbed nutrients 97

milk fat composition in fasting cows was similar to that of body fat.
Generally speaking, milk fat is elevated when body reserves are
used extensively to support lactation. In one study (0rskov et al.,
1977), body fat utilization was increased by infusing casein into the
abomasum of cows which otherwise received feed sufficient only for
the production of about 10 kg milk/d. Table 7.8 shows that both
milk fat and protein concentrations, and also milk yield, were
increased by the casein infusions, although the calculated energy
deficit simultaneously increased.

IV. COMPOSITION OF NUTRIENTS AND ENDOCRINE


CHANGES

So long as ruminants consume roughage-based diets there is


relatively little fluctuation during the day either in VF A production
and uptake or in protein flow to the small intestine. There is thus
less need for endocrine systems to regulate blood concentrations of
absorbed nutrients than in monogastric animals, such as the pig or
human, which may consume starchy meals at widely spaced
intervals. However, when ruminants receive diets that result in the
rapid generation and absorption of propionic acid, insulin
production rises. Increasing concentrations of insulin stimulate
most types of cell to take up nutrients from the blood. In lactating
animals this can deprive the mammary glands of sufficient nutrients,
leading to low milk fat and eventually to reduced milk yields.
0rskov et al. (1969) showed that when milk fat was reduced by
infusing a VF A mixture high in propionic acid, heat production
was unchanged, but the distribution of energy between milk and
body tissue was altered. Sutton (1980) demonstrated that this effect
was particularly marked in dairy cows when concentrate was given
twice rather than six times a day. Istasse et al. (1987), using the
intragastric nutrition technique, simulated continuous feeding by
continuous infusion of VF As and casein, and intermittent feeding
by delivering a part of the propionic acid in two 3 h pulses within
each 24 h. They also compared the propionic acid dose with doses
of glucose or casein of equal energy value, delivered continuously
or in pulses to the abomasum. Some of the results are illustrated in
98 Energy Nutrition in Ruminants

60

40

20

0
tI


<-

I
+' 120
::>
E

/-
100
c
:J

--I
oil
c
80

60

. ./1 /
40

20
~
:~
4~~!:\' 4
\~~::--:
f /

o
08:00 12:00 16:00 20:00
T i me of day (h)

Fig. 7.i. Changes in plasma concentrations of insulin in cows in response to the


following intragastric infusions : continuous propionic acid and casein (0) ; in-
termittent propionic acid (.) ; intermittent casein (.); casein-free (0); con-
tinuous glucose (6); intermittent glucose (.. ). _ : Duration of intermittent
infusions. (From Istasse et al. (1987) .)
Use of energy of absorbed nutrients 99
Fig. 7.i. They showed that each treatment affected blood insulin
and blood glucose differently. Pulses of propionic acid caused
increases of blood propionate and insulin (but no change in blood
glucose), both returning to normal levels within 2-3 h after the end
of the pulse. Glucose infusion led to increases in blood glucose and
insulin levels, which again returned to normal within 2-3 h. Casein,
on the other hand, caused an increase in insulin but a decrease in
blood glucose, the return to normal levels after the end of the
infusion being generally slower than after propionic acid or glucose
infusion. Casein also led to increased blood free amino acids. To
summarize, increased blood levels of propionic acid, glucose and
also free amino acids were associated with increased insulin
production, which' probably explains why amino acids caused a
reducti.on in blood glucose concentration (see also Bassett, 1972).
Although there are no reports of blood acetic acid affecting
insulin levels, blood propionic acid certainly does. In most
circumstances propionic acid does not reach the peripheral
circulation, being converted in the liver to succinic acid, glucose,
etc. However, when large amounts of concentrates are fed
intermittently a rise in plasma propionate occurs. Thus the feeding
regime can have major implications for ruminant metabolism. In
sheep and goats - but not in cattle - when the capacity of the
liver to metabolize propionic acid is exceeded, both propionic acid
itself and its metabolite, methylmalonic acid, appear in the blood.
These interfere with normal fat synthesis, resulting in a large
proportion (sometimes more than 20%) of odd-numbered and
branched-chain fatty acids. The latter have very low melting points
and give rise to the 'soft fat' syndrome described by Duncan et al.
(1974).

REFERENCES

Agricultural Research Council (1980) The Nutrient Requirements of Ruminant


Livestock Commonwealth Agricultural Bureaux, Slough
Armstrong, D.G. & Blaxter, K.L. (1957a) The heat increment of steam-volatile
fatty acids in fasting sheep Brit. J. Nutr. 11, 247-272
Armstrong, D.G. & Blaxter, K.L. (I957b) The utilization of acetic, propionic and
butyric acids by fattening sheep Brit. J. Nutr. 11,413-425
100 Energy Nutrition in Ruminants
Armstrong, D.G., Blaxter, K.L. & Graham, N.McC. (1957) The heat increment of
mixtures of steam-volatile fatty acids in fasting sheep Brit. J. Nutr. II, 392-408
Armstrong, D.G., Blaxter, K.L., Graham, N.McC. & Wainman, F.W. (1958) The
utilization of the energy of two mixtures of steam-volatile fatty acids by
fattening sheep Brit. J. Nutr. 12, 177-188
Bassett, J.M. (1972) Plasma glucagon concentrations in sheep: their regulation and
relation to concentrations of insulin and growth hormone Austr. J. Bioi. Sci. 25,
1277-1287
Duncan, W.R.H., 0rskov, E.R., Fraser, C. & Garton, G.A. (1974) The effect of
processing of dietary barley and of supplementary cobalt and cyanocobalamine
on the fatty acid composition of lamb triglyceride with special reference to
branched chain fatty acids Brit. J. Nutr. 32, 71-75
Fattet, I., Hovell, F.D. de B., 0rskov, E.R., Kyle, D.J. & Smart, R.I. (1984)
Undernutrition in sheep. The effect of supplementation with protein on protein
accretion Brit. J. Nutr. 52, 561-574
Flatt, W.P., Coppock, C.E. & Moore, L.A. (1965) Energy balance studies With
lactating, non-pregnant dairy cows consuming rations with a varying hay to
grain ratio In Blaxter, K.L. (Ed.) Energy Metabolism pp 121-130 E.A.A.P.
Publication No. 11, Academic Press, London
Hovell, F.D. de B. (1972) The Utilization of Salts of Volatile Fatty Acids by
Growing Lambs Ph.D. Thesis, University of Aberdeen
Hovell; F.D. de B., 0rskov, E.R., Grubb, D.A. & MacLeod, N.A. (1983) Basal
urinary nitrogen excretion and growth response to supplemental protein by
lambs close to energy equilibrium Brit. J. Nutr. 50, 173-187
Istasse, L., MacLeod, N.A., Goodall, E. & 0rskov, E.R. (1987) Effect on plasma
insulin of intermittent infusion of propionic acid, glucose or casein into the
alimentary tract of non-lactating cows maintained on a liquid diet Brit. J. Nutr.
58, 139-148
KuVera, J.c., 0rskov, E.R. & MacLeod, N.A. (1988) Energy transactions in cattle
nourished by intragastric nutrition In Van Der Honig, Y. (Ed.) Proc. E.A.A.P.
Energy Metabolism Symposium, pp 271-274, Pudok, Wageningen
Leng, R.A. & Preston, T.R. (1976) Sugar cane for cattle production. Present
constraints, perspectives and research priorities Trop. Anim. Prod. 1, 1-12
Moe, P.W., Tyrrell, H.F. & Flatt, W.P. (1970) Partial efficiency of energy use for
maintenance, lactation, body gain and gestation in the dairy cow In Schiirch, A.
& Wenk, O. (Eds) Energy Metabolism of Farm Animals pp 65-67 E.A.A.P.
Publication No. 13 Juris Druck & Verlag, Zurich
0rskov, E.R. (1980) Possible nutritional constraints in meeting the energy and
protein requirements of highly productive ruminants In Ruckebusch, Y. &
Thivend, P. (Eds) Digestive Physiology and Metabolism in Ruminants pp
309-327 MTP Press, Lancaster
0rskov, E.R. & Allen, D.M. (1966a) Utilization of salts of volatile fatty acids by
growing sheep. 1. Acetate, propionate and butyrate as sources of energy for
young growing lambs Brit. J. Nutr. 20, 295-305
0rskov, E.R. & Allen, D.M. (1966b) Utilization of salts of volatile fatty acids by
growing sheep. 3. Effect of frequency of feeding on the utilization of acetate and
propionate by young growing lambs Brit. J. Nutr. 20, 509-517
0rskov, E.R. & Allen, D.M. (1966c) Utilization of salts of volatile fatty acids by
Use of energy of absorbed nutrients 101

growing sheep. 4. Effects of type of rumen fermentation of the basal diet on the
utilization of salts of volatile fatty acids for nitrogen retention and body gains
Brit. J. Nutr. 20, 519-532
0rskov, E.R. & MacLeod, N.A. (1982) Effect of volatile fatty acid composition
and protein on energy utilization and milk composition in cows sustained by
intragastric nutrition In Ehern, A. & Sundstol, F. (Eds) Energy Metabolism of
Farm Animals pp 22-25 The Agricultural University of Norway, Aas
0rskov, E.R. & MacLeod, N.A. (1990) Dietary induced thermogenesis in
ruminants and feed evaluation Proc. Nutr. Soc. (In press)
0rskov, E.R., Hovell, F.D. de B. & Allen, D.M. (1966) Utilization of salts of
volatile fatty acids by growing sheep. 2. Effect of stage of maturity and hormone
implantation on the utilization of volatile fatty acid salts as sources of energy for
growth and fattening Brit. J. Nutr. 20, 307-315
0rskov, E.R., Flatt, W.P., Moe, P.W., Munson, A.W., Henken, R.W. & Katz, I.
(1969) The influence of rumina I infusion of volatile fatty acids on milk yield and
composition and energy utilization by lactating cows Brit. J. Nutr. 23,443-453
0rskov, E.R., Grubb,D.A. & Kay, R.N.B. (1977) Effect of postruminal glucose
or protein supplementation on milk yield and composition in Friesian cows in
early lactation and negative energy balance Brit. J. Nutr. 39, 397-405
0rskov, E.R., Grubb, D.A., Wenham, W. & Corregall, W. (1979) The sustenance
of growing and fattening ruminants by intragastric infusion of volatile fatty
acids and protein Brit. J. Nutr. 41, 553-558
0rskov, E.R., MacLeod, N.A., Fahmy, S.T.M., Istasse, L. & Hovell, F.D. de B.
(1983) Investigation of nitrogen balance in dairy cows and steers nourished by
intragastric infusion. Effect of submaintenance energy input with or without
protein Brit. J. Nutr. 50, 99-107
0rskov, E.R., Reid, G.W. & Tait, A.G. (1987) Effect of fish meal on the
mobilization of body energy in dairy cows Anim. Prod. 45, 345-348
Preston, T.R. & Leng, R.A. (1986) Matching Livestock Production Systems to
Available Resources International Livestock Centre for Africa, Addis Ababa
Storm, E., Brown, D.S. & 0rskov, E.R. (1983) The nutritive value of rumen
micro-organisms in ruminants. 3. The digestion of microbial amino and nucleic
acids in, and losses of endogenous nitrogen from, the small intestine of the sheep
Brit. J. Nutr. 50, 479-485
Sutton, 1.0. (1980) Digestion and end product formation in the rumen from
production rations In Ruckebusch, Y. & Thivend, P. (Eds) Digestive Physiology
and Metabolism in the Ruminant pp 271-290 MTP Press, Lancaster
Tyrrell, H.F., Reynolds, P.l. & Moe, P.N. (1979) Effect of diet on partial efficiency
of acetate use for body tissue synthesis by mature cattle J. Anim. Sci. 48,
598-606
CHAPTER 8

FEED QUALITY AND FEED INTAKE

I. Introduction
II. Feed-related Factors
A. Extent of digestion
B. Rate of digestion
C. Rate of reduction of large to small particles
D. Prediction of intake from feed characteristics
III. Animal-related Factors
A. Control of intake of concentrate feeds
B. Rumen volume
C. Effect of lactation
D. Effect of temperature
E. Recovery from low-level nutrition
F. Effect of physical work
IV. Conclusions
102
Feed quality and feed intake 103

I. INTRODUCTION

The control of food intake in mammals is highly complex. Some of


the factors known to be involved in the initiation of hunger and in
the inhibition of eating are listed below:
Initiation of hunger Inhibition of eating
Metabolic demand unsatisfied Metabolic demand satisfied
Palatable food Unpalatable food
Social stimuli Social stimuli
Endocrine stimuli Gut distension
Pharmacological stimuli Disturbance; fear; pain
Nausea
High temperatures

In ruminants gut distension is particularly important, but in


general actual intake is the product both of factors promoting
hunger and of factors promoting satiety. Some of the factors are
poorly understood, such as those involved in the extreme
seasonality of intake in some species. For example, red deer stags
consume twice as much in spring and early summer as in winter,
even when given the same diet (Kay, 1985).
The prediction of feed intake, in particular of fibrous roughage,
is one important aspect of ruminant nutrition and will be discussed
at length in this chapter. This subject has been the focus of many
lively debates at numerous meetings. There are good reasons for
this, since knowledge of feed values is often of little interest without
knowledge of how much the animals will consume. It would be of
great advantage to be able to predict consumption from the
characteristics of the feed itself. The more the animal relies on
slowly degraded roughages, the more relevant are these character-
istics. Simple estimates of digestibility or metabolizability are often
of little value if there is no information regarding the amounts that
will be consumed. Both the feed evaluation systems based on net
energy, e.g. starch equivalents and feed units, and those based on
metabolizable energy (ME) and total digestible nutrients (TDN)
suffer from this disadvantage. In recent years some characteristics
104 Energy Nutrition in Ruminants

60

50
.~.-
g 40
..-
I'll
"0
I'll
~ 30
/ b a+b

,/'
QJ
o
20

10 L-------------
,.
a I I ,

a 24 48 72 96
Time (h)

Fig. S.i. Description of the degradation of fibrous feeds, p = a + b(l- e- ct ),


derived from the results of intra-ruminal incubations.

of roughage which affect voluntary intake have been identified.


Many authors - notably Mertens & Ely (1979), Van Soest (1982)
and Ellis et al. (1987) - have contributed to these findings, and to
a more dynamic description of digestion than that based only on
data for digestibility in vitro or in vivo. Nevertheless, although it is
recognized that feed characteristics and animal factors are jointly
involved in the control of intake, it is necessary to discuss their
respective contributions separately. (Certain aspects of palatability
which modify intake, such as the tannin content of the feed, will not
be discussed.)

II. FEED-RELATED FACTORS

When dealing with a feed which may impose physical limitations on


intake, various characteristics should be considered. The potential
extent of digestion determines the minimal proportion which
Feed quality and feed intake 105

persists as indigestible residue, a residue that takes up space while


it is retained in the stomach. At the same time the rate of digestion
determines how long potentially digestible material must occupy
space. The rate at which long particles are reduced to small particles
presents yet another possible constraint. In contrast, it is clear that
the water-soluble fraction of roughage occupies little or no space in
the stomach. Therefore, in order to describe those characteristics of
roughages which relate to their consumption, an attempt will be
made to analyse each of these aspects separately.

IIA. Extent of Digestion

The potential extent of digestion can be determined by measuring


the loss of dry matter from samples of feed, contained in nylon
bags, which are incubated within the rumen for various periods of
time. It is depicted in Fig. S.i as the asymptote of the curve for
percent degradation against time. The curve may be expressed
mathematically, using the formula of0rskov & McDonald (1979):

(S.l)

where p is the percent degradation at time t, and a, band care


constants. It follows that (a + b) is the potential degradability of
the material. However, it should be emphasized that (a + b) is not
synonymous with digestibility since the latter will vary with the
time spent in the rumen, i.e. the rumen retention time. As discussed
elsewhere, rumen retention time varies with rumen volume and the
level of feeding, and possibly also with the rate at which large
particles are reduced in size (see below). Although (a + b) should
be constant for a feed when determined under rumen conditions
that promote cellulolysis, digestibility in vivo is not in fact constant.
It varies both with the level of feeding and with the nature and
quantity of any supplements which either enhance or depress the
.degradation rate. Some standardization is, however, achievable by
measuring digestibility in vivo at maintenance energy intake.
Ideally (a + b) should be separated into the two components.
106 Energy Nutrition in Ruminants

Thus a represents the intercept of the curve, i.e. the material which
dissolves immediately in the rumen fluid, but which occupies little
or no space there. In contrast, b is the insoluble but potentially
degradable fraction. A lag phase sometimes occurs before degra-
dation is detectable. This can result in an apparently negative value
for a. However, the true value can be estimated directly, from the
fraction of the contents lost from a nylon bag when washed.
Alternatively, McDonald's (1981) modified equation may be used.
(See Chapter 10 for further discussion.)
It follows from the formula that 100 - (a + b) is a measure of
the absolutely indigestible material in the feed, expressed as a
percentage, i.e. the minimal proportion Jhat will be passed into.the
faeces and that always occupies space. The difference between this
value and the observed digested proportion can sometimes be used
t() identify problems of depressed or inhibited degradation.
However, it should also be pointed out that when conditions in the
rumen are very unsuitable for cellulolytic activity some com-
pensatory degradation can occur in the caecum and large intestine.
This is usually accompanied by increased faecal N because, unlike
in the rumen, the proteins of the microbes involved in the
degradation of cellulose in the hind gut are not subsequently
digested.

liB. Rate of Digestion

The value of c in the equation indicates the rate of degradation of


insoluble materials. In combination with rumen retention time it
determines the quantity of material which is degraded, and hence
the fraction of potentially degradable but insoluble material which
is actually utilized. Unlike a and b, c is not a constant value of a feed
unless specifically determined under optimal conditions for cellu-
lolysis. On the other hand, the differences in degradation rate which
are associated with different rumen environments can be used to
assess the extent to which cellulolysis is inhibited. For instance if
digestibility in vivo is far less than predicted, it is worth examining
in detail the conditions for cellulolysis in the rumen, where it may
Feed quality and feed intake 107

be inhibited by acidity, by inadequate ammonia concentration, etc.


The value c as a feed characteristic should therefore be determined
under the most ideal conditions for cellulolysis.

nc. Rate of Reduction of Large to Small Particles


The establishment of rate constants defining the reduction of large
particles to particles small enough to permit outflow has proved
difficult and is likely to continue to do so. The process is related
both to animal behaviour (chewing and rumination) and to the feed
characteristics. Thus the level of susceptibility of cell walls to
microbial erosion and the degree of friability of the fibres play an
important part (see McLeod & Minson, 1988).
For a long time it was considered that particle size in the rumen
effluent was controlled at the reticulo-omasal orifice. This hy-
pothesis is now largely discarded, because the size ranges of
particles in the faeces of sheep, goats and cattle are rather similar
although the size of the reticulo-omasal orifice differs several-fold
between these species. It now seems (Ullyat et ai., 1986) that the
particles are filtered through the meshwork of solid digesta. This
results from the biphasic movements of the reticulum, which
alternately promote the outflow to the omasum of the liquid phase
- and with it a suspension of small, dense particles from the depths
of the reticulum - and project it back onto the solids.
It is nevertheless important to know whether the rate of particle
reduction is a real constraint to intake, i.e. whether it is slower or
faster than the rate of outflow of small particles. If particle size
reduction is more rapid than the outflow of small particles, feed
intake will mostly be controlled by the feed characteristics defined
above, namely the rate and the extent of digestion. Mira et al.
(1983) observed no differences in the intake of straw by cattle
whether it was ground or only chopped, suggesting that the rate of
particle size reduction did not limit intake. Similarly, 0rskov et al.
(1988a), using chromium-mordanted straw as a marker (prepared
by the technique of Uden et aI., 1980), found that the mean rumen
retention time was reduced by only about 10% (from 53 to 48 h) as
a result of grinding the straw. While results so far indicate that for
108 Energy Nutrition in Ruminants

most roughages the time taken to reduce large to small particles is


not a serious constraint on intake, it is also clear that no
generalizations should be made. Some roughages, such as sisal
pulp, contain extremely strong fibrous particles which are un-
doubtedly reduced very slowly and can therefore seriously constrain
feed intake.

lID. Prediction of Intake from Feed Characteristics

The prediction of levels of intake from degradation characteristics


was first undertaken some years ago. Thus Chenost et al. (1970)
found that, for 82 forages, the results of 24 h incubations with
nylon bags eorrelated better with voluntary intake than did the
results of digestibility trials in vivo. Similarly, Hovell et al. (1986)
showed that the relationship between the intake of four forages and
(a + b) was better than that between intake and digestibility in vivo.
In a recent experiment designed to investigate the voluntary
intake of straw by steers, 0rskov et al. (l988b) identified five types
of straw which differed in all three of the degradation character-
istics, a, band c, referred to in Section IIA of this chapter. To
increase the range of variability, samples of the same straws were
treated with ammonia, giving a total of 10 types. The untreated
samples were sprayed with a solution of urea to provide 2% by
weight of this substance, so ensuring that their rates of degradation
were not limited by a deficiency of N. The values of the three
characteristics for each straw were obtained from its degradation
curve. This was determined by incubating samples in nylon bags
within the rumens of sheep, during a series of time intervals. The
values of a, band c which described each curve with least residual
error were then computed. Table 8.1 shows the fitted values for the
disappearance of dry matter. It also shows how the values of a, b
and c varied between the different straws and that treatment with
ammonia on the whole increased the value of b but did not
consistently alter c.
Each type of straw was subsequently fed to a group of 8 steers for
60 days. Total dry matter intake, digestible dry matter intake and
growth rate were recorded. The possibilities of predicting the level
TABLE 8.1
The Effects of Straw Type, Variety and Ammonia Treatment on Degradation Characteristics in Nylon Bags Incubated for
7-72 h in the Rumen, with Values of the Constants in the Formula p = a+b(l-e- C' )

Type Variety Ammonia Disappearance of dry matter (DM) a b c Residual


~
II>
Treatment from nylon bags (g/100 g DM) SD s::...
~

7h 24h 48h 72 h
~
~
12.9 24.3 33.4 36.0 6.0 32.9 0.0337 0.25 §
Winter barley Gerbel
+ 16.9 33.0 46.5 53.8 7.9 54.4 0.0258 1.21 s::...
14.2 28.3 37.4 41.0 5.1 38.2 0.0391 0.97 ~
II>
Winter barley Igri s::...
+ 17.8 33.7 44.8 49.6 7.9 45.2 0.0351 0.63

17.4 36.8 47.3 50.6 3.4 48.7 0.0483 0.66 i:)
Spring barley Corgi ;>;-
+ 22.9 46.6 60.0 64,5 6.4 60.4 0.0457 2.04 II>

16.6 32.3 44.3 50.1 7.5 48.0 0.0303 1.09


Spring barley Golden Promise
+ 21.4 40.3 52.8 57.9 9.3 52.1 0.0376 0.34
16.5 30.7 40.8 45.2 7.7 40.9 0.0345 1.76
Winter wheat Norman
+ 20.7 39.3 51.9 57.2 9.0 51.9 0.0364 0.81
From 0rskov et al. (l988b).

0
-
1.0
110 Energy Nutrition in Ruminants

of each of these parameters from the asymptote of the degradation


curve (a + b), from the asymptote plus rate constant c, and from
the three separate factors, a, band c, were then tested. This was
achieved by submitting all the data to multiple regression analysis.
The values of the correlation coefficient, r, indicating the predictive
efficiencies of the factors used, are presented in Table 8.2, together
with the regression coefficients for the equations. This table shows
that dry matter intake, digestible dry matter intake and growth rate
could be predicted with great accuracy from multiple regression
equations based on the values for a, band c. Since a, band c were
not correlated this was a statistically valid exercise. It can be seen
in Table 8.2 that, for each parameter pr.~dicted, the addition of the
rate constant c to (a + b) in the muitiple regression always
decreased the residual error and increased the value of the
correlation coefficient.
In this experiment, predictive accuracy was achieved despite
neglect of the elusive rate constant expressing the rate of reduction
of large to small particles. This, however, may simply imply either
that its value was similar for all these particular cereal straws, or
that in each case it was greater than the value for outflow rate from
the rumen. It by no means implies that the rate of particle reduction
can always be ignored.

DI. ANIMAL-RELATED FACTORS

Although, as shown above, feed characteristics can apparently


determine the relative levels of intake of different straws by groups
of similar animals, it is well known that individual animals vary in
their consumption of the same materials. It has already been noted
that variations in the duration of chewing and/or rumination could
result in different levels of intake if the rate of reducing particle size
were a limiting factor. However, the fact that outflow rate increases
with feeding level suggests again that the rate of particle reduction
is generally of minor importance. Animals vary considerably in
their capacity to consume roughages, a large intake being the
consequence either of a rapid outflow or of increased rumen
TABLE 8.2
Prediction ofIntake of Total and Digestible Dry Matter and of the Growth Rate of Steers from Degradation Characteristics
Generated from the Equation p = a+b(! _e-ct ), with Regression Coefficients Used to Derive Predicted Values of the Y
Variable from the Factors of this Equation.

Y Factors Coefficients applied to Intercept r Residual ~


~
variable used SD 1::1...
~
(a+b) c a b ;:::
I:)

Total dry (a+b) 0.0766 0.572 0.83 0.452 ~


matter intake (a+b)+c 0.0748 40.7 0.822 0.89 0.375
--.
I:)
;:s
1::1...
(kg/d) a+b+c 56.4 0.159 0.0658 -1.56 0.88 0.383
~
~
Digestible dry (a+b) 0.0642 1.258 0.86 0.33 1::1...
matter intake (a+b)+c 0.0624 39.0 -2.595 0.96 0.195 s·
(kg/d) a+b+c 37.7 0.0554 0.0640 -2.576 0.95 0.204 ~
~
Growth (a+b) 0.0175 -0.595 0.84 99
rate (a+b)+c 0.0170 9.55 -0.922 0.91 77
(g/d) a+b+c 17.02 0.0571 0.0126 -1.267 0.95 54
From 0rskov et al. (1988b).
(a+b), Asymptote of degradation curve.
tv
-
TABLE 8.3
Examples of Different Feedstuffs Expressed as a, band c from the Equation p = a + b(l- e- ct )

Feedstuff Variety Ammonia a .b c Residual


treated SD

Spring barley Doublet -7.20 78.91 0.0419 0.43


straw Klaxon -11.93 62.60 0.0736 5.35 ti'
Winter barley Pipkin 7.66 46.51 0.0518 1.43 "'
~
~
straw Opera 4.14 44.61 0.0507 0.83
Norman -3.70 83.78 0.0136 3.29 ~
~
Winter wheat Renard -3.23 61.70 0.0220 1.05 ::;:
straw Norman + -2.50 83.97 0.0295 2.11 c·
::s
Renard + 3.00 81.97 0.0234 2.58 s·
Trafalgar 5.21 65.61 0.0209 1.40 ~
Oat Dula 1.90 61.96 0.0265 0.57 §
straw Trafalgar + 2.14 8];42 0.0296 3.30 s·
§
Dula + 2.72 78.13 0.0271 3.27 0::-
Hay (barn dried) 4.96 69.72 0.0772 6.83
Barley grain 18.5 81.5 0.0340
(caustic treated)
Barley grain (rolled) -6.04 90.41 0.2500 1.28
Cotton seed 32.2 60.2 0.0820
Soya bean 9.9 80.1 0.0820
Ground nut 12.1 87.9 0.0640
Sunflower 19.8 77.8 0.0610

From E. R. 0rskov (unoublished).


Feed quality and feed intake 113

volume (discussed below), the latter permitting long retention time


(low fractional clearance) combined with a normal outflow rate.

InA. Control of Intake of Concentrate Feeds

There is much disagreement concerning what constitutes con-


centrate and what constitutes roughage. When are we dealing with
so-called metabolic regulation of intake and when are we dealing
with physical limitations? Farmers often refer to bought-in,
processed feeds as concentrates and to home-grown materials as
roughage. Thus roots, such as swedes, are considered to be
roughage, although in terms of their degradation rate they resemble
concentrates. Perhaps only fibrous feeds should be referred to as
roughages. Alternatively it might be more appropriate to define all
feeds in terms of the constants a, band c used above, rather than
to attempt more precise definition of the traditional categories.
Table 8.3 lists a series of different feeds together with their
properties, expressed in terms of their degradation characteristics
(see also Chapter 10).
Description in these terms may help to define groups of feeds
which can be expected to share physical or metabolic limitations to
intake, including characteristics relevant to particular animal
species or ages. For example Andrews et al. (1969) demonstrated
that the consumption of oats (but not barley) by lambs weighing 20
kg, which presumably had rumens of immature proportions, was
limited by the physical properties of the feed. Yet in older and
heavier lambs (40 kg) intake was regulated primarily by the
metabolic capacity of the animals, the intake of oats being about
equal to that of barley in terms of digestible dry matter, but greater
in terms of total weight (Table 8.4).
It can be readily understood that the consumption of high-
quality diets, such as those based on cereals, will increase when
requirements change due to lactation, increased physical work
load, low environmental temperatures, seasonal variation in rate of
growth and fat deposition, etc. Animals receiving poor-quality
diets, which impose physical limitations to intake, also often
consume more under such circumstances, as a result of metabolic
114 Energy Nutrition in Ruminants
TABLE 8.4
Intake of Dry Matter and Digestible Dry Matter when Two Concentrate Diets
were Offered to Lambs Weighing 20, 30 and 40 kg

Live weight Mean intake (gjd)


(kg)
Dry matter Digestible dry matter

Barley Oats Barley Oats


20 927 780 730 493
30 1038 1094 818 691
40 1150 1410 906 891
From Andrews et al. (1969).

factors that stimulate hunger and suppress satiety. It is assumed


here that the rumen environment for cellulolysis remains constant.
This implies that, if rumen volume continues unchanged, rumen
retention time and hence feed digestibility has to be reduced.
Nevertheless, whatever the diet, animals will in general vary their
intake when their energy requirements alter.

I11B. Rumen Volume

Rumen volume can be of critical importance for determining the


intake of poor-quality roughages. This is vividly illustrated in early-
weaned calves and lambs, where the rumen is insufficiently
developed for the consumption of sufficient roughage for survival.
Older animals, of similar weight but with better-developed rumens,
survive successfully on the same diet. As quoted above, Andrews et
al. (1969) showed that even the intake of oats can be physically
restricted in early-weaned lambs weighing less than 40 kg.
Differences in rumen v:olume between individuals and between
breeds have largely been ignored, although they deserve attention.
In many developing countries poor-quality roughages, such as
straw, may provide up to 90% of the diet. The voluntary intake of
straw by the indigenous breeds of cattle is often consistently greater
than that of breeds in developed countries presented with similar
material. For instance, the voluntary intake relative to body weight
Feed quality and feed intake 115

of urea-supplemented barley straw by Friesian heifers was 42


g/kgWo. 75 , whereas that of similar quality urea-supplemented rice
straw by small zebu cows in Bangladesh was 75 g/kgWo. 75 (Mould
et al., 1982). These authors subsequently found that the filled gut of
Bangladeshi zebu cattle represented 33% of live body weight,
compared with about 18% for 'western' breeds. Regarding the
latter, it is likely that pot-bellied animals, with large rumen volumes
permitted by large abdominal cavities, have been consistently
culled in the process of selecting beef cattle with a high killing-out
percentage. In the process, the capacity to consume and utilize
poor-quality roughage may have declined. Weyreter et al. (1987)
recently reported the interesting observation that native heathland
sheep had a bigger rumen volume and longer retention time than
so-called improved sheep. Moreover, they extracted more nutrients
from their feed and lost less weight when maintained on poor-
quality roughage (wheat straw).
There is little or no direct information either on the inheritance
of gut volume or on whether it is possible to select ruminants for a
greater capacity to utilize poor-quality roughages. In a recent trial
(0rskov et al., 1988a) three Friesian cows exhibiting a high
fractional outflow of fibrous particles (i.e. fraction of rumen
particle content leaving the rumen per hour) were compared with
three exhibiting a low fractional outflow. Since the voluntary intake
of the two groups was similar, the latter animals were expected to
show higher apparent digestibility than the former. Their rumen
retention time, which is inversely related to fractional outflow, was
of course the greater and it is likely that this was due to larger
rumens. Using these two groups of cows, the effects on fractional
particle outflow of different types of feed, with high or low
proportions of roughage, and of different levels of feeding, were
examined. Table 8.5 shows the results. The differences between the
two groups persisted, regardless of diet and levels of feeding.
Moreover, as the table clearly indicates, this was reflected in the
digestibility of the diet. This consistency of differences suggests that
characteristics determining roughage utilization may be heritable
and presents a challenge to breeders to produce animals capable of
high production performance on poor-quality roughages.
-0\

TABLE 8.5
Mean Food Intake, Outflow Rate and Apparent Digestibility in Cows Selected for Low (LO) and High (HI) Fractional
Outflow from the Rumen, when Offered Low- and High-Roughage Diets ad libitum and in Restricted Amounts

Parameter Fractional Low roughage High roughage ~


outflow '"
~
'.,:
Ad lib Restricted Ad lib Restricted
Live weight LO 599 599 598 596 ~
of cows (kg) HI 590 583 587 583
S.
5'
;::
Dry matter intake LO 148.7 105.8 107.4 76.0
(g/kgwo. 75 /d) HI 145.6 103.5 106.1 74.8 s·
~
Fractional LO 0.0298 0.0257 0.0275 0.0254 §
outflow/h HI 0.0323 0.0309 0.0302 0.0316 s·
0.731 0.657 0.655 §
Dry matter LO 0.688 ~
digestibility HI 0.674 0.696 0.636 0.640
Organic matter LO 0.710 0.751 0.673 0.670
digestibility HI 0.696 0.714 0.651 0.656
From 0rskov et al. (1988a).
Three cows in each group.
Feed quality and feed intake 117

IIIC. Effect of Lactation

Lactation has an effect of special interest. The greater requirement


for energy, and consequent more rapid removal of metabolites
from the blood, is associated with hypertrophy of the rumen
(discussed in detail by Forbes, 1986). Feed intake is greatly
increased, showing that the constraint imposed by rumen volume
on the maximum intake of non-lactating, non-pregnant cows is
much reduced. Tulloh (1966) compared lactating and non-lactating
twin cows. The weights of their filled rumens were 116 and 90 kg
respectively and those of the digesta 55 and 39 kg. Most of the
changes in feed intake occurring during lactation can be related to
increased rumen volume. There is little or no change in digestibility,
indicating only minor alterations of fractional outflow (Fell et at.,
1972). It is interesting that the greatest feed intake occurs in mid-
lactation, after peak milk production. Forbes (1982) suggested that
the size of the liver may be implicated, since this enlarges relatively
slowly in response to the increased metabolic demands of lactation.
However, the control of rumen hypertrophy is not well understood.

IIID. Effect of Temperature

Feed intake increases with a fall in ambient temperature, as was


originally highlighted by Graham et at. (1982) and later by various
Canadian workers (see the excellent review by Kennedy et at.,
1986). Unlike the effects of lactation, lowered temperatures
probably do not lead to an increase in rumen volume. On the other
hand, the fractional outflow of fibrous particles is definitely
augmented and, in consequence, the digestibility of roughages is
diminished (Westra & Christopherson, 1976; Kennedy et at., 1976).
This effect has been reported both in cattle and in sheep maintained
at low temperatures, and also in sheep after shearing in cool
conditions. The digestibility of concentrates is not reduced, since
their high rate of degradation ensures complete utilization
regardless of temperature and retention time (Williams & Innes,
1982). The extent to which the digestibility of roughage is reduced
when the fractional outflow increases will depend on the rate of
degradation relative to the rate of outflow of small particles, i.e. on
1I8 Energy Nutrition in Ruminants

the quality of the feed. Kennedy (1985) showed that when the
ambient temperature was reduced from 20-25°C to 0-5°C, the
digestibility of cell wall constituents of rapidly degradable alfalfa
fibre fell from 45 to 42%, but that of more slowly degraded brome
grass fibre fell from 53 to 43%. Degradation rates measured in
nylon bags are unaffected by external temperature, as one would
expect if rumen fermentation rate is unchanged. Thus intake
responses following a reduction in temperature seem to be mediated
wholly through increased fractional outflow, i.e. decreased re-
tention time.

IIIE. Recovery from Low-level Nutrition

It is well known that ruminants at a low level of nutrition during


winter or dry seasons will increase consumption when this
constraint is removed, and eat more feed relative to body weight
than will previously well-nourished animals. Metabolic factors
influence appetite to a considerable extent but, apart from these, it
is not known whether increased intake is related to greater rumen
volume, to a higher fractional outflow or to both. Animals
recovering from a low-protein status also increase their feed
consumption, as a result of high anabolic uptake and rapid removal
of nutrients from the circulation. Egan (1970) observed that sheep
infused post-ruminally with protein increased their intake of
roughage-based diets, apparently as a result of increased rumen
volume. Similar effects on intake were obtained in protein-deficient
lambs which received concentrate, supplemented with urea to meet
microbial needs. Table 8.6 summarizes the comparable results of
0rskov et af. (1973), who gave lambs different levels of protein by
bottle in order to bypass the rumen via the oesophageal groove, and
so avoid microbial degradation. Again, feed intake was enhanced.
Thus corrections of metabolic deficiencies or constraints can lead to
increased appetite, even for diets which impose physical restrictions
on intake.
Responses to the relief of metabolic constraints, like responses to
the generally increased metabolic activity of lactation, must not be
confused with the effects of altered microbial nutrition. It is
Feed quality and feed intake 119
TABLE 8.6
Daily Intake, in g dry matter/d, of a Barley/Urea Diet by Lambs of Various
Live Weights which Received either no Supplement or Increments of Fish-
protein Concentrate by Bottle

Supplement Live weight (kg)


(g/d)
25 35 45
0 851 1078 1265
17 994 1190 1415
34 927 1196 1561
51 1003 1241 1416
SEM of means 48 44 43
From 0rskov et al. (1973).
Each value is the mean of four animals.

assumed above that microbial nutrition suffered no constraint.


However, if the supply of nitrogen, or some other essential factor
such as phosphorus, restricts microbial nutrition, the potential
maximal degradation rate will not be achieved and both feed intake
and digestibility will be reduced.

llIF. Effect of Physical Work

Very little information is available on the effects of physical work


on the consumption of roughage diets and on the kinetics of their
digestion, yet, compared with other ruminants, the vast numbers of
animals carrying out the physical work of cultivation and transport
have less time available in which to feed and to ruminate, although
they do apparently ruminate while working. This is an area in
which information is urgently needed.

IV. CONCLUSIONS

Ruminant intake is affected by feed texture and flavour, by


~ntimicrobial substances in forage (such as tannins and mimosine),
by disturbance, pain and sickness, by social facilitation and
inhibition, and by high temperatures. Nevertheless, as shown
120 Energy Nutrition in Ruminants

above, under normal management, dietary and animal factors


relating to ruminant digestion have an overwhelming and increas-
ingly predictable influence on feed intake, and hence on productive
performance.

REFERENCES

Andrews, R.P., Kay, M. & 0rskov, E.R. (1969) The effect of different dietary
energy concentrations on the voluntary intake and growth of intensively fed
lambs Anim. Prod. 11, 173-185
Chenost, M., Grenet, E., Demarquilly, e. & larrige, R. (1970) The use of the nylon
bag technique for the study of forage digestion in the rumen and for predicting
feed values Proc. II th Int. Grassland Congr., Surfers Paradise pp 697-701 Univ.
of Queensland Press, St. Lucia
Egan, A.R. (1970) Utilization by sheep of casein administered per duodenum at
different levels of roughage intake Austr. J. agric. Res. 21, 85-94.
Ellis, W.e., Wylie, M.l. & Matis, 1.H. (1987) Dietary-digestive interactions
determining the feeding value of forages and roughages In 0rskov, E.R. (Ed.)
Feed Science World Animal Science B4 pp 177-225 Elsevier, Amsterdam
Fell, RF., Campbell, R.M., Mackie, W.S. & Weekes, T.E.e. (1972) Changes
associated with pregnancy and lactation in some extra-reproductive organs of
the ewe J. agric. Sci., Camb. 79, 397-407
Forbes, 1.M. (1982) The role of the liver in the control of food intake Proc. Nutr.
Soc. 41, 123-126
Forbes, 1.M. (1986) The effect of sex hormones, pregnancy and lactation on
digestion, metabolism and voluntary food intake In Milligan, L.P., Grovum,
W.L. & Dobson, A. (Eds) Control of Digestion and Metabolism in Ruminants
pp 420-435 Reston Books, Nl
Graham, A.D., Nicol, A.M. & Christopherson, R.I. (1982) Rumen motility
responses to adrenaline and noradrenaline and organ weights of warm- and
cold-acclimatized sheep Canad. J. Anim. Sci. 62, 777-786
Hovell, F.D.de R, Ngambi, 1.W., Barber, W.P. & Kyle, D.l. (1986) The voluntary
intake of hay by sheep in relation to its degradability in the rumen as
measured in nylon bags Anim. Prod. 42, 111-118
Kay, R.N.R (1985) Seasonal variation in appetite in ruminants In Haresign, W.
(Ed.) Recent Developments in Ruminant Nutrition pp 195-215 Butterworth,
London
Kennedy, P.M. (1985) Effect of rumination on reduction of particle size of rumen
digesta by cattle Austr. J. agric. Res. 36, 819-828
Kennedy, P.M., Christopherson, R.I. & Milligan, L.P. (1976) The effect of cold
exposure of sheep on digestion, rumen turnover time and efficiency of microbial
synthesis Brit. J. Nutr. 36, 231-242
Kennedy, P.M., Christopherson, R.I. & Milligan, L.P. (\986) Digestive responses
to cold In Milligan, L.P., Grovum, W.L. & Dobson, A. (Eds) Control of
Digestion and Metabolism in Ruminants pp 285-306 Reston Books, Nl
McDonald, I. (\981) A revised model for the estimation of protein degradability
Feed quality and feed intake 121

in the rumen J. agric. Sci., Camb. 96, 251-252


McLeod, M.N. & Minson, D.J. (1988) Large particle breakdown by cattle eating
ryegrass and alfalfa J. Anim. Sci, 66, 992-999
Mertens, D.R. & Ely, L.O. (1979) A dynamic model of fibre digestion and passage
in the ruminant for evaluating forage quality J. Anim. Sci. 49, 1085-1095
Mira, J.J.F., Kay, M. & Hunter, E.A. (1983) A comparison of long and shredded
cereal straw for beef cattle Anim. Prod. 36, 87-92
Mould, F.L., Saadullah, M., Haque, M., Davis, c., Dolberg, F. & 0rskov, E.R.
(1982) Investigation of some of the physiological factors influencing intake and
digestion of rice straw by native cattle of Bangladesh Trop. Anim. Prod. 7,
174-181
0rskov, E.R. & McDonald, I. (1979) The estimation of protein degradability in
the rumen from incubation measurements weighted according to rate of passage
J. agric. Sci., Camb. 92, 499-503
0rskov, E.R., Fraser, C. & Pirie, R. (1973) The effect ofgy-passing the rumen with
supplements of protein and energy on intake of concenfrates by sheep Brit. J.
Nutr. 30, 361-367
0rskov, E.R., Ojwang, I. & Reid, G.W. (1988a) A study on consistency of
differences between cows in rumen outflow rate of fibrous particles and other
substrates and consequences for digestibility and intake of roughages Anim.
Prod. 47, 45-51
0rskov, E.R., Reid, G.W. & Kay, M. (1988b) Prediction of intake by cattle from
degradation characteristics of roughages Anim. Prod. 46, 29-34
Tulloh, N.M. (1966) Physical studies of the alimentary tract of grazing cattle. III.
Seasonal changes in capacity of the reticulo-rumen of dairy cattle N. Z. J. agric.
Res. 9, 252-260
Uden, P., Colucci, P.E. & Van Soest, P.J. (1980) Investigation of chromium,
cerium and cobalt as markers in digesta. Rate of passage studies J. Sci. Food &
Agric. 31, 625-632
Ullyatt, M.J., Dellow, D.W., John, A., Reid, C.S.W. & Waghorn, G.c. (1986)
Contribution of chewing during eating and rumination to the clearance of
digesta from the ruminoreticulum In Milligan, L.P., Grovum, W.L. & Dobson,
A. (Eds) Control of Digestion and Metabolism in Ruminants pp 498-515 Reston
Books, NJ
Van So est, P.J. (1982) Nutritional Ecology of the Ruminant 0 & B Books,
Corvallis, OR
Westra, R. & Christopherson, R.J. (1976) Effects of cold on digestibility, retention
time of digesta, reticulum motility and thyroid hormones in sheep Canad. J.
Anim. Sci. 56, 699-708
Weyreter, H., Heller, R., Dellow, D.; Lechner-Doll, M. & Engelhardt, W.V.
(1987) Rumen fluid volume and retention time of digesta in an indigenous and
a conventional breed of sheep fed a low quality fibrous diet J. Anim. Physiol. &
Anim. Nutr. 58, 89-100
Williams, P.E.V. & Innes, G.M. (1982) Effects of short term cold exposure on the
digestion of milk replacer by young preruminant calves Res. Vet. Sci. 32,
383-386
CHAPTER 9

FEED EVALUATION, PAST AND


PRESENT

I. Historical aspects
A. Introduction
B. The total digestible nutrient system
C. Net energy or starch equivalent systems
II. Current methods
A. Current methods of calorimetry
1. Indirect closed-circuit calorimetry
ii. Indirect open-circuit calorimetry
111. Open and shut chambers
iv. Open-circuit respiration hoods
B. Current feed evaluation systems
i. Net energy systems
ii. Metabolizable energy systems
III. Limitations of current feed evaluation systems
122
Feed evaluation, past and present 123

I. mSTORICAL ASPECTS

IA. Introduction

The desire to establish an 'exchange rate' between different animal


feeds is not new. It has existed since animal production became a
commercial occupation, indeed since animals first served man's
need for cultivation and transport. Tyler (1975) gave a brilliant
account of its history, with numerous references extending back to
observations on the value of different feeds as early as 2500 BC
Flatt (1988) also summarized many historical aspects of feed
evaluation. There is therefore no need for an elaborate account
here.
Although differences in the values of alternative feeds were no
doubt widely recognized from early times, tables of relative values
were probably first compiled in Germany. Tyler (1975) noted that
in 1725 feeds such as hay were evaluated in Bavaria in terms of
Straw Units, straw being the most abundant staple fodder for cattle
in that region. The famous Captain Middleton, of Kent (quoted by
Tyler, 1975), used Hay Units, other feeds being compared with
meadow hay. Thus 1 tonne of hay was estimated to be equal in
value to 8 tonnes of turnips, 3 tonnes of carrots or 500 kg of linseed
oil cake. Although Middleton was apparently the first to publish a
hay-based evaluation system, Thaer (1809-1812) is generally given
the credit for this innovation. For about 50 years the hay equivalent
was widely used. Morton's (1855) encyclopaedia listed nine authors,
including Thaer, who evaluated various feeds in this way. Some of
the values which they obtained are presented in Table 9.1. These
were derived from feeding trials, not from digestion trials, weight
gain being the yardstick.

lB. The Total Digestible Nutrient System

As frequently occurs, critical scientists eventually realized that the


standard, hay itself, was too variable to be the basis of satisfactory
comparisons. Henneberg & Stohmann (1860), at Weende Gottingen,
found that the daily maintenance requirement of an ox could vary
124 Energy Nutrition in Ruminants

TABLE 9.1
Hay Equivalents as Summarized by Morton (1855)

Author % Water

Block Thaer Middleton Schweizer

Meadow hay 100 100 100 100 11


Rye straw 200 666 267 19
Barley straw 193 150 200 11
Carrots 366 300 338 300 88
Barley grain 33 76 35 13
Linseed cake 43 13

from 4 to 7 kg, depending on the quality of the hay. To replace hay


as a standard they developed a procedure based on the digestibility
of protein, carbohydrate and fat. This total digestible nutrient
(TDN) system provided the basis for methods of feed evaluation. It
was adopted in the U.S.A. and, with fairly minor modifications, is
still used there, despite the fact that TDN procedures do not take
into account energy lost in methane and urine.

Ie. Net Energy or Starch Equivalent Systems


Energy lost in methane and urine is generally higher relative to
digestible energy with roughage feeds than with concentrates.
Consequently, to compare the entire range of feeds it is essential to
take these losses into account. The net energy or starch equivalent
system was developed in order to make such adjustments. During
the first half of the twentieth century it became the basis for feed
evaluation in Europe. Kellner (1905) published a comprehensive
book on this system and generally receives the credit for its
development, although other workers were also involved. He had
access to respiration chambers and was able to study changes in the
rate of fat deposition in mature cattle when various pure nutrients
were added to a basal diet. The amount of fat deposited per kg of
test nutrient was expressed in terms of the amount deposited per kg
of starch fed. The results are usually presented in kcal, but have
been recalculated in kJ for Table 9.2. It is of interest that digestible
Feed evaluation, past and present 125

TABLE 9.2
Starch Equivalents of Pure Nutrients; after Kellner and the Scandinavian Feed
Unit System

Nutrient Fat deposited/kg nutrient (kJ) Starch equivalent


factor
Starch 9448 1.00
Fibre 10057 1.00
Protein 9280 0.94

fibre yielded the same value as starch. It is also interesting that the
value for protein was lower. However, Kellner soon found that
when normal diets were examined in terms of-their content of pure
nutrients, the obserVed values were often lower than those expected,
especially for roughage. He therefore instituted a fibre correction
factor to be applied to the crude fibre content of the diet.
A similar net energy procedure, the Scandinavian feed unit
system, was developed by Johannes Fjord. It was based on the
amount of fat deposited as a result of feeding 1 kg of barley to
ruminants, estimated as equal to 1650 kcal net energy. When this
and the starch equivalent systems were eventually integrated, the
factor for protein was increased from 0.94 (see Table 9.2) to 1.43 on
the basis of the ratio of the calorific values for protein and
carbohydrate. However, neither the old nor the new protein factor
appear to have been supported by any experimental evidence.

II. CURRENT METHODS

IIA. Current Methods of Calorimetry

With the advent of equipment designed for continuously recording


and handling large amounts of data, and for accurately measuring
the composition of air, many respiration chambers were established
in many parts of the world and, with them, almost as many feed
evaluation systems! Most of these respiration chambers depend on
indirect calorimetry, in which the heat produced is estimated from
the carbon dioxide exhaled and the oxygen consumed, together
126 Energy Nutrition in Ruminants

with methane loss. (Direct calorimetry - as the name implies-


measures the heat produced directly.) There are essentially three
types of indirect calorimetry.

i. Indirect Closed-circuit Calorimetry


In this system the same air is circulated through the chamber for
periods of 24 h. It is pumped through a solution of KOH and the
absorbed CO 2 is measured gravimetrically. Inputs of oxygen
compensate for the absorbed CO 2 and maintain the air pressure. At
the end of each run the composition of the circulating air is
determined. The system is extremely accurate and robust but lends
itself to small animal studies only. The: quantity of KOH solution
required to remove the CO 2 produced by cattle, for example, is
prohibitively expensive for practical purposes.

ii. Indirect Open-circuit Calorimetry


Here the composition of the air entering and leaving the animal
chamber is analysed. Oxygen consumed and CO 2 produced are
estimated by difference. This system requires extremely sensitive
and accurate equipment to determine the small differences in air
composition. In addition, it requires very accurate measurement of
the flow rate of the air as it is pumped out ofthe chamber. Methane
production is usually measured and therefore - because some
methane is produced in the hindgut -- a completely closed chamber
must be used. The system can be connected directly to data-logging
equipment, the results being stored on computer tape. It is at
present the technique most commonly used for feed evaluation.

iii. Open and Shut Chambers


In order to retain some advantages of a closed circuit, without
having to use a CO 2 absorbent, Sir Kenneth Blaxter and his
colleagues developed the so-called open and shut chamber (Blaxter
et al., 1972). In principle, air is circulated in a large, closed chamber
for about 30 min, its composition being measured at the beginning
and end of this period. It is then replaced with fresh air. The
procedure is repeated as many times as required. Accuracy is
enhanced because the volume of circulating air is known, and the
Feed evaluation, past and present 127

differences in composition which develop over 30 min are much


larger than those measured by open-circuit calorimetry. Essentially
the same principle was used by Feng Yanglian et al. (1985) at the
Rowett Research Institute. His animals expired air into a very large
balloon fitted with non-return valves. Differences between the
composition of the inhaled air and the contents of the balloon were
large. The drawback of this method is its failure to measure
methane generated in the hindgut.

iv. Open-circuit Respiration Hoods


There are, in addition to the three main systems of indirect
calorimetry, simple systems which involve pumping air through a
hood placed over the animal's face. Provided backflow of expired
air is prevented, measurement of the flow rate and of the
composition of the intake and exhaust air permit calculation of
heat production. Such methods are usually employed only for short
periods of measurement. However, equipment derived from the
principles of both the respiration hood and the open-circuit
chamber was developed for additional study of intragastric
nutrition (KuVera et al., 1988) and can be used for this purpose
during 24 h periods. This nutritional procedure eliminates hindgut
methane production, so there is no need to enclose the animal
completely, as in the open-circuit chamber.

lIB. Current Feed Evaluation Systems

In principle, only three systems of feed evaluation are currently


used, based respectively on the measurement of net energy, the
measurement of metabolizable energy and the measurement of
total digestible nutrients (TDN), discussed earlier.

i. Net Energy Systems


The system developed by Kellner has been further elaborated at the
Oscar Kellner Institute, in Rostock, G.D.R. As before, different
factors are defined for digestible crude protein, digestible ether
extract, digestible crude fibre and digestible N-free extract. A
series of correction factors are added for low-quality roughage
128 Energy Nutrition in Ruminants

(Schiemann et al., 1971). The authors claimed great accuracy for


the method, but the inclusion of such factors as 'digestible crude
protein' is surprising, since this term has very little biological
meaning in ruminants (see 0rskov, 1982).
A measure developed in the Netherlands is derived by first
computing the metabolizable energy in digestible nutrients, then
converting it to net energy. This value is further transformed to a
unit (VEM) based on the net energy obtained from 1 g of barley,
i.e. 1.65 kcal (Van Es, 1970). In France, Vermorel (1978) developed
a similar system based on 1 kg of barley, one unit being equivalent
to 1.73 Meal. A comparable Swiss system expresses the results in
joules instead of calories.
A net energy system, based on the work ofW.P. Flatt, P.W. Moe
& H.F. Tyrrell at Beltsville, was developed in the United States for
use with dairy cows (see Flatt, 1988). However, the older system,
based on TON, is still very much in evidence there. Lofgreen &
Garrett (1968) developed another procedure, based on comparative
slaughter experiments involving both sheep and cattle, for use with
fattened animals. This is widely applied, but it is based on the body
composition of breeds favoured in the United States and is not
necessarily applicable to breeds markedly different from them in
this respect.

ii. Metabolizable Energy Systems


In the U.K. the net energy system based on starch equivalents has
recently been replaced by one based on metabolizable energy (ME),
expressed in megajoules (Agricultural Research Council, 1980).
ME is the gross energy less energy in faeces, urine and methane.
This was adopted because of observed differences between the
utilization of energy for maintenance, fattening, growth and
lactation. Measurements 'On animals differing in their physiological
function resulted in different estimates of the net energy values for
various feeds. By basing the system on the animal's requirement for
ME, corrections can be applied appropriate to the level and type of
production. The ME values of the diet can be calculated from the
metabolizability of the gross energy, estimated at the maintenance
level of feeding. Differences in efficiency due to diet digestibility are
Feed evaluation, past and present 129

computed both as differences between its efficiency for maintenance


(Km) and its efficiency for production above maintenance (Kf). The
particular advantage of the system is that the ME values remain
constant for each type of feed. In practice, however, complications
arise when calculating the animal's requirement for a specific level
of production. Differences in the utilization of ME for particular
functions have to be taken into account when formulating rations.
In contrast, net energy systems simply allocate correction factors
for fattening, lactation etc. The major departure of the Agricultural
Research Council's ME system from earlier ones is, therefore, the
assumption that the efficiency of utilization is itself greater below
the maintenance level of feeding (Km) than- above maintenance
(Kf). This assumption was discussed in Chapter 6, where its validity
was questioned. The validity of allocating urinary energy to
attributes of the feed is also open to question, since the largest part
of urinary energy has been metabolized by the animal. It could thus
be argued that only faecal and methane energy should be subtracted
from the gross energy.

III. LIMITATIONS OF CURRENT FEED EVALUATION


SYSTEMS

During the 19th century the Hay Equivalent proved to be


insufficiently accurate, especially when applied outside the area in
which it was developed. Although barley is a more standard
product, a system based on it is generally applicable only where
barley or other such grains are normally fed to ruminants, i.e. in
very few countries of the world. The new net energy and
metabolizable energy systems attempt to describe the type of feed
resource independently from the need of the animal, using calories
or joules as the reference units. However, the more varied the
resources, the more inadequate is the description of feed values in
terms of a single static property. This criticism can be levelled at all
the feed evaluation systems so far developed.
There are two problems. First, the metabolizability of a feed is
not constant. It can be altered extensively by changes in both the
130 Energy Nutrition in Ruminants

retention time and the degradation rate (see Chapter 8). Assessment
at the maintenance level of feeding, as in the ME system, goes some
way towards reducing these sources of variation. Thus the
degradation rate would probably be near optimal, with high rumen
pH, and variations in outflow rate would probably be of the same
order as normal between-animal variations. The difficulty lies in
extrapolating to levels differing from maintenance, as do most
feeding routines. The feeding level correction of the ME system
applied to a high concentrate diet nearly always yields a constant
ME value, regardless of feeding circumstances, since degradation is
always rapid. However, ME values for roughage feeds and mixed
diets are far from constant as the degradation rate varies greatly,
depending ort rumen pH, substrate competition, the physical
structure of the feed and the outflow rate, which increases with the
level of feeding (see Chapter 8).
The second and most serious shortcoming of current feed
evaluation systems relates to their failure to include information on
the amount that the animals will voluntarily consume. This aspect
was also discussed in Chapter 8. It is of relatively little value to
know the metabolizability of a feed, or its potential or net energy
value, if one knows neither how much will be eaten nor the
maximum proportion that can be tolerated in the total diet while
achieving the desired feed consumption.
The basis of a further criticism, summarized in Chapter 6, relates
to the evaluation of feeds relative to fasting metabolism as in, for
example, the ME system. Fasting metabolism, involving the
utilization of body fat, the degradation of protein and increased N
excretion, appears to be influenced by protein and glucose
deficiency. Sometimes heat production may even be reduced as a
result of alleviating the glucose deficiency during fasting by infusing
small amounts into the abomasum. Certainly, the excretion ofN in
the urine is reduced by such infusions. The problem of fasting
metabolism has given rise to the confusing concepts of so-called
Km and Kf, which assume differences in the efficiency of feed
utilization below and above the maintenance level of feeding.
Despite these limitations it is evident that the established systems
of feed evaluation have served and will continue to serve a very
Feed evaluation, past and present 131

useful purpose in the countries where they have been developed.


They are particularly useful there for allocating feed to animals
kept indoors, especially when it is of high-quality and is rationed.
However, when feeding routines (for dairy cows in particular) are
based on the unrestricted supply of completely mixed diets, these
evaluation systems do not adequately predict consumption, and
hence the cost of sustained performance. Moreover, there is a trend
towards indoor beef production with unrestricted supplies of either
high-quality fattening diets or low-quality roughage. In either case
it is essential for the future of these production methods that feed
evaluation systems should provide information on the animals'
probable feed consumption.

REFERENCES

Agricultural Research Council (1980) The Nutrient Requirements of Ruminant


Livestock Commonwealth Agricultural Bureaux, Slough
Blaxter, K.L., Brockway, J.M. & Boyne, A.W. (1972) A new method for
estimating the heat production of animals Quart. J. expo Physiol. 57, 60-72
Feng Yanglian, Mollison, G.S., Smith, J.S. & Brockway, J.M. (1985) A new
closed circuit mask system for estimating heat production in animals In Moe,
P.W., Tyrrell, H.F. & Reynolds, P.J. (Eds) Energy Metabolism of Farm Animals
pp 140-144 E.A.A.P. Publication No.32 Rowman & Littlefield, NJ
Flatt, W.P. (1988) Feed evaluation systems: historical background In 0rskov,
E.R. (Ed.) World Animal Science; B4 Feed Science pp 1-22 Elsevier,
Amsterdam
Henneberg, W. & Stohmann, F. (1860) Beitrage fur Begrundung einer nationellen
Futterung der Wiedenkauer Vol.1 F.U. Schmetschtke & M. Bruhn, Braun-
schweig
Kellner, O. (1905) Erniihrung der landwertschaftlichen Nutztiere 1st Edition Paul
Parey, Berlin
KuVera, J.e., MacLeod, N.A. & 0rskov, E.R. (1988) Energy exchanges in cattle
nourished by intragastric infusions of nutrients. In Van der Honig, Y. & Close,
W.H. (Eds) Energy Metabolism of Farm Animals pp 271-274 Proc. 11th
E.A.A.P. Symp. on Energy Metabolism in Farm Animals, Pudok, Wageningen
Lofgreen, G.P. & Garrett, W.N. (1968) A system for expressing net energy
requirements and feed values for growing and finishing beef cattle J. Anim. Sci.
27,793-806
Morton, J.e. (1855) A Cyclopaedia of Agriculture Blackie & Son, London
0rskov, E.R. (1982) Protein Nutrition in Ruminants Academic Press, London
Schiemann, R., Nekring, K., Hoffmann, L., Jentsch, W. & Chudy, A. (1971)
Energetische Futterbestung und Energienormen VEB Deutscher Landwert-
schaftsverlag, Berlin
132 Energy Nutrition in Ruminants
Thaer, A. (1809-1812) Grundsatze der rationellen Landwirtschaft Vol. 1-4 Berlin.
Quoted by Flatt (1988)
Tyler, C. (1975) Albrecht Thaer's hay equivalents: fact or fiction? Nutr. Abstr. &
Rev. 45, 1-11
Van Es, A.J.H. (1970) Stimulation of the growth of veal calves fed liquid milk
replacers In Schurch, A.S. & Wenk, C. (Eds) Energy Metabolism in Farm
Animals pp. 97-100 E.A.A.P. Publication No. 13 Juris Druck & Verlag, Zurich
Vermorel, M. (1978) Feed evaluation for ruminants. II. The new energy systems
proposed in France Livestock Prod. Sci. 5, 347-365
CHAPTER 10

TOWARDS FUTURE FEED


EVALUATION SYSTEMS

I. A promising approach
A. The problem of the lag phase
B. Use of index values
C. Problems of negative associative effects
D. Problems of outflow rate
II. Methods of measurement
A. Soluble fraction
B. Insoluble but degradable fraction
C. Degradation rate constant
III. Animal req uiremen ts

I. A PROMISING APPROACH

It was shown in Chapter 8 that the intake of straw by steers could


be predicted quite accurately from information on the degradation
characteristics of the feed, generated from the formula p = a + b
(1 - e- ct ). In the same way, M. Gill & Y. Nakashima (personal
communication) predicted with similar accuracy the intake of
silage-based diets by sheep. This formula has generated useful
133
134 Energy Nutrition in Ruminants

information relating to the restrictions of gut capacity. The soluble


material a contained in roughage contributes little to rumen fill. In
contrast, the insoluble but fermentable material b contributes to the
volume occupied initially, and the rate constant e determines the
duration of its diminishing occupation. The insoluble and in-
digestible material, 100 - (a + b), occupies a constant volume
until voided.

IA. The Problem of the Lag Phase

Certain problems relating to this model must, however, be


considered. Application of the original equation to experimental
data often results in a negative value for a (see Table 8.3), although
this is in reality impossible. With protein supplements and with
grains, insoluble material disappears rapidly. However, with
cellulosic materials there is often an initial lag phase during which
the rumen micro-organisms colonize the feed particles, before the
insoluble material corresponding to b begins to disappear (see
Chapter 2). Indeed if the dry matter in nylon bags is weighed at
short intervals after the beginning of incubation in the rumen, it
appears to get heavier, due to accumulating adhering microbes.
This lag phase gives rise to negative values of a when the
degradation curve is extrapolated back to zero, as clearly illustrated
in Fig lOj. It is probably important in practice, since during this
period all the insoluble but fermentable material is retained in the
rumen and thus contributes to possible restrictions on further
intake. Negative values for a are often combined with very high
values for b, although (a + b) obviously always indicates the
asymptote of the degradation curve, i.e. the total proportion of
fermentable material. Figure 8j illustrates the curve assuming no
lag phase.
The problem of negative values can be tackled by measuring the
actual amount of soluble material. If this is denoted by a', the lag
phase L can be calculated from the expression: L = lie (bla + b
- a'), as described by McDonald (1981). Figure 10j illustrates the
problem and its solution. If b' = (a + b) - a', the prediction
equations presented in Chapter 8 (Table 8.2) can be reanalysed in
Towards future feed evaluation systems 135
60

50

c 40
o
....1\1
~ 30
L
01
~
o 20

I I ,

8 16 24 _ 48
_ _ _ _ _ _72
_____ J

Ti me (h)

Fig. to.i. Description of degradation of fibrous residues, p = a' + b'(l- e- ct ). The


soluble fraction is indicated by the broken line. L, Lag phase.

terms of these substituted factors, the values of c remaining


unchanged. These amended results are presented in Table 10.1,
which also includes values obtained when the lag phase, L, was
included in the equation.
The correlation coefficients between the measured and predicted
values for digestibility, dry matter intake, etc. are presented. The
values of r indicate the accuracy with which the various parameters
can be predicted. Table 10.1 shows how each degradation
characteristic contributed to this accuracy. When only the
asymptote (a + b) was used the prediction was not very reliable. It
was significantly improved for all parameters by including the rate
constant, c. Substitution of a' and b' for a and b resulted in a further
significant improvement in the accuracy with which digestibility,
dry matter intake and growth rate were predicted (r = 0.90, 0.93
and 0.95 respectively). However, inclusion of the lag phase, L, led
to no further significant improvement.
This approach seems to offer great scope. It also indicates that no
adequate description of a roughage feed can be obtained from a
single property of it. The soluble fraction, the potential digestibility
of the insoluble fraction and the rate constant are all important.
136 Energy Nutrition in Ruminants
TABLE 10.1
Accuracy of Estimating Digestibility, Dry Matter Intake, Digestible Dry Matter
Intake and Growth Rate of Steers from Feed Degradation Characteristics, as
Indicated by the Multiple Correlation Coefficients (r) between Factors of the
Degradation Equation and these Parameters; Comparison with Multiple Corre-
lation Coefficients between the Index Value and the same Parameters

Factors used in Digestibility Dry matter Digestible Growth rate


multiple intake dry matter
regression intake
analysis

(a+b) 0.77 0.83 0.86 0.84


(a+b)+c 0.85 0.89 0.96 0.91
a'+b'+c 0.90 0.93 0.96 0.95
a' +b' +c+L 0.91 0.93 0.95 0.95
Index value 0.74 0.95 0.94 0.96

E. R. 0rskov (unpublished).
(a+b), Asymptote of degradation curve.
a', Soluble fraction of feed.
b' = (a+b)-a'.
L, Lag phase.

lB. The Use of Index Values

For practical purposes it may sometimes be useful to derive a single


value for each feed. In this way, different feeds could be ranked in
order of their ability to provide a diet adequate for maintenance or
for any other specified intake level. Such a definition has been
attempted (0rskov, 1989), on the basis of the regression equation
used to predict the dry matter intake of steers (Y) from a', b' and
c and to generate the values of r, shown in Table 10.1, from the
equation:
(10.1)
Since this regression was calculated to achieve the lowest residual
error, the regression coefficients allocated to each of the factors
provide measures of their importance as contributors to the
prediction of intake. In order to simplify this equation, the
coefficient for a' was transformed to 1 by dividing each component
of the equation by Xl' This procedure generated coefficients for b'
Towards future feed evaluation systems 137
TABLE 10.2
Effects of Type and Treatment of Straws Fed to Steers on Degradation
Characteristics (using Measured Solubility as the a' Value and the Lag Phase
(L) Calculated According to McDonald (1981», Ranked According to Index
Values Based on Weighted a', b' and c

Straw Ammonia a' b' L Index Dry mailer


treated (h) value intake (kg/d)
Type Variety

Winter barley Gerbel 12.5 26.3 0.0359 6.7 30.2 3.43


Winter barley Igri 13.6 29.7 0.0389 6.5 33.3 3.56
Spring barley Golden Promise 15.0 40.5 0.0304 5.7 37.3 4.43
Winter barley Igri + 15.9 37.3 0.0350 5.5 37.8 4.82
Winter wheat Norman 19.3 29.4 0.0343 9.7 37.9 4.57
Winter barley Gerbel + 16.0 46.3 0.0257 6.2 39.6 4.70
Spring barley Corgi 16.0 36.1 0.0481 6.2 40.1 5.16
Spring barley Golden Promise + 20.1 41.2 0.0l77 6.1 44.1 4.93
Winter wheat Norman· + 24.5 37.5 0.0364 9.7 46.4 5.81
Spring barley Corgi + 19.0 47.7 0.0457 5.1 47.2 5.86

E. R. 0i-skov (unpublished).
Symbols as in Table 10.1.

and e of 0.4 and 200 respectively. Using these coefficients, the sum
of a' + O.4b + 200e was defined as the index value. For example,
the index value for Gerbel straw is
12.5 + (0.4 x 26.3) + (200 x 0.0359) = 30.2 (10.2)
When the straw was treated with ammonia the index value changed
to
16.0 + (0.4 x 46.3) + (200 x 0.0257) = 39.6 (10.3)
The predictive accuracy of the index value for the four parameters
of Table 10.1 is indicated by the figures in the bottom line of that
table. It is of course to be expected that dry matter intake would be
highly correlated with the index value, since the latter was derived
from the regression equation for this parameter. Nevertheless, the
correlation with growth rate (r·= 0.96) is encouraging. Table 10.2
shows the values of the degradation characteristics, a', b' , e and L
for 10 different types of straw ranked in order of their index values,
together with their daily intake by steers. When ranked by index
value the sequence is very similar to that when ranked by intake.
It may be possible to use such index values to predict the
minimum feed quality required to enable an animal to consume
138 Energy Nutrition in Ruminants

sufficient to meet its maintenance energy need. It was calculated


from the present data that an index value of about 35.5 was
required if the steers were to consume sufficient for their need
(assumed to be 450 kJ of ME/kgWo. 75 /d). It must be emphasized
here that the proposed index values do not refer to a single
biological characteristic and have no biological meaning, yet they
may prove useful for ranking feeds in the order of their potential
for consumption in sufficient quantity to meet the requirements for
various levels of production. Table 10.3 presents a range of feeds
analysed in the authors' laboratory, with their values of a', b', c and
L and their index values. More data are needed to elaborate this
approach and to confirm that the sameJactors apply to the whole
range of those ruminant feeds that impose physical limitations on
consumption.
At the same time, values generated for high-quality feeds may
well also prove to be important for formulation of complete diets,
in which different qualities of feed are intimately mixed. In recent
unpublished work from the authors' laboratory, complete diets
were formulated from hay or ammoniated wheat straw combined
with different amounts of concentrate. The absence of associative
effects was assumed. The ranking of the combined index values
corresponded to intake better than did that of the digestibility of
the complete feeds.
Finally, it should be pointed out that further improvements are
possible in the prediction of performance on diets with abnormally
high lipid content or ash content (such as animal excreta). This
could be achieved by describing the disappearance specifically of
organic matter or of fermentable energy. Care will also be needed
where soluble materi;:tl may include indigestible phenolics, as in
alkali-treated straws, or antimicrobial substances.

Ie. Problems of Negative Associative Effects

Negative associative effects, discussed in Chapter 3, result mainly


from less than optimal fermentation conditions, which reduce the
degradation rate constant. Consequently, the index value of a feed,
established experimentally, may not be realized in practice due to
Towards future feed evaluation systems 139
TABLE 10.3
Description of Feeds in Terms of the Factors of the Exponential Equation and
the Index Value

Type offeed a' b' c L Index


value
Spring barley straw (Colt) 10.3 33.8 0.0466 4.8 33.1
Spring barley straw (Corgi) 12.8 37.1 0.0580 6.7 39.2
Spring barley straw (Doublet) 10.9 39.9 0.0495 5.8 36.8
Winter barley straw (Gerbel) 6.6 39.1 0.0247 3.3 27.2
Oat straw (Ballad) 11.4 38.2 0.0240 2.7 31.5
Rice straw (Sasanisiki) 17.1 36.0 0.0399 4.2 39.5
Maize stover 15.6 46.7 0.0356 12.8 41.4
Barley leaf blade 15.6 70.2 0.0672 5.0 57.1
Barley stems 13.5 36.4 0.0406 7.3 26.2
Oat leaf 11.3 49.4 0.0352 3.9 38.1
Oat stems 12.4 29.8 0.0152 1.5 27.1
Rice leaf 15.1 37.2 0.0340 5.2 36.8
Rice stems 30.0 33.5 0.0484 4.7 53.1
Maize cob 12.5 41.5 0.024 16.1 33.9
Maize leaf 19.7 38.0 0.041 14.2 41.5
Maize stem 14.1 36.9 0.032 11.2 35.5
Hay 21.5 49.6 0.037 3.2 59.0
E. R. 0rskov & W. Shand (unpublished).
Symbols as in Table 10.1.

constraints within the specific fermentation conditions. Such


constraints need to be studied when new feeding systems are
developed. The realization that negative associative effects are
generally due to a reduced rate constant should help to ensure that
optimal fermentation conditions are sought and therefore to ensure
that feeds are utilized with maximum efficiency.

ID. Problems of Outflow Rate

The predictive accuracy shown in Table 10.1 is impressive,


accounting for almost 90% of the variation in feed intake, as
represented by the high correlation coefficients; yet it may well be
possible in some instances to improve it further, by measuring the
rate at which long particles of feed are degraded to small particles
and by measuring the fractional outflow of the latter. The
140 Energy Nutrition in Ruminants

complexity of measuring the outflow rate of particles has been


discussed recently in an excellent paper by Kennedy & Murphy
(1988). Simultaneous complete information on the fractional
outflow and on the rate of degradation should, at least theoretically,
provide a still better solution for evaluating both roughages and the
ability of an animal to extract the ma#mum amount of fermentable
substrate during their passage through the gut. (Fermentation in
the hindgut of undegraded but potentially digestible material will
tend, of course, to increase the actual overall efficiency of feed
utilization above this estimate.) The conclusions are, however,
likely to be influenced by animal production systems, feeding levels,
type of animal (in particular, differ~nces in gut volume),etc.
Nevertheless, as demonstrated in Table 10.1, description of the feed
itself can come quite close to predicting intake and performance.

IT. METHODS OF MEASUREMENT

Assuming that the three most important characteristics which


determine the energy yield of a feed are its content of soluble
material, its content of potentially degradable material and the
degradation rate constant, we must consider methods for measuring
these properties.

ITA. Soluble Fraction

This can be determined by a variety of methods, including solubility


in a buffer, loss by solution from nylon bags when washed and
adjusted for small particle loss, and - probably the most ap-
propriate - measurement of those materials in cell contents and
cell walls by neutral detergent analysis (cf. Van Soest, 1963).

ITB. Insoluble but Degradable Fraction

So far, only biological methods are capable of measuring this


parameter adequately. Together with the soluble fraction it
comprises the asymptote in the exponential equation given at the
Towards future feed evaluation systems 141

beginning of this chapter. It can be determined by the prolonged


incubation in the rumen of nylon bags containing samples of the
test feed. This characteristic could perhaps also be determined by
measuring gas production during prolonged incubations in an in
vitro system (Menke et al., 1979). It is unlikely that gross chemical
methods could generate the required information. Ramanzin et al.
(1986) and Reid et al. (1988) showed that neither the fihre content,
determined by various methods, nor even the lignin content was
closely correlated with the potential extent of digestion. Both the
architecture of the plant cell walls and the lignin content appear to
be important. Preliminary studies suggest that a way forward may
be found by calibrating near-infra-red analysis data against data
obtained by incubating nylon bags in the rumen (I. Murray,
unpublished).

lIe. Degradation Rate Constant

This parameter is without doubt the most difficult to determine


accurately, yet its importance was clearly demonstrated in Chapter
8 and in Tables 8.2 and 10.1. It can be determined by incubating
nylon bags of the test material in the rumen for a series of time
intervals, and could perhaps also be measured by means of in vitro
incubations. An approximate rate constant might perhaps even be
generated from a series of incubations using cellulase enzymes
combined with measurements of gas production (Menke et al.,
1979).
In conclusion, it should be pointed out that the measurements
indicated above are far less expensive to carry out than those based
on indirect calorimetry, yet their ability to predict intake and
associated growth rate is far superior.

III. ANIMAL REQUIREMENTS

Recently developed procedures for determining the protein require-


ments of ruminants permit more precise allocation of the supply
than formerly (0rskov, 1982). More precise measurement of the
142 Energy Nutrition in Ruminants

energy need is similarly required. Information is lacking about the


animals' metabolic capacity, their capacity to produce milk of
varying composition and their capacity for weight gain with
varying fat lean ratios. Moreover, it is necessary to know what type
of resource can meet either overall need or a desired proportion of
the maximal potential. It may not prove cost-effective to provide
for maximum potential metabolic activity, owing to restraints that
depend on the quality and price of available feeds. For example,
lambs marketed at closely similar weights and composition can
range from 3 to 12 months of age. High-quality feed, with a high
index value, is essential if the lambs are to reach slaughter weight
by 3 months. A different resource strategy serves for fattening
lambs by 1 year. The latter will obviously be much less efficient
from the point of view of energy utilization, since four times as
much will be used simply for maintenance, yet it may nevertheless
be more cost-effective since it can be achieved with cheaper feed.
In Chapter 8, reference was made to large differences between
animals in rumen retention time and therefore in the efficiency of
roughage utilization. The work of Mould et al. (1982), in
Bangladesh, and of 0rskov et al. (1988) showed clearly that cattle
can vary greatly in their capacity to use poor-quality feeds, and that
such feeds also vary greatly. It should therefore be possible to select
animals with a view to matching them with the quality which is
available locally for ruminant consumption. Equally, it may be
possible to develop feed resources that match more precisely the
animals' requirement and capacity (see Tuah et al., 1986).
With the advent of the effective manipulation of fat stores and
the demonstration that stored fat can fuel protein deposition
(Fattet et al., 1984), workers involved in the energy nutrition of
ruminants face a new challenge. The energetics of the use of fat for
protein deposition needs to be investigated. This question is
particularly relevant where the nutritional supply fluctuates, since
dietary protein can be so manipulated that animals gain weight in
the form of lean tissue, despite an energy balance which is negative
in terms of its exogenous supply. The ability of the animals to store
fat and the cost of fat storage compared with other methods of feed
conservation also warrant investigation. In some parts of the world
Towards future feed evaluation systems 143

the most successful ruminants may be those which consume large


amounts when feed is plentiful and its quality good, and which
deposit fat that can subsequently be used for growth, draught
power, pregnancy and lactation.
Thus the possibility of manipulating fat stores, like other sources
of energy, and of selecting animals to match resources, provide new
challenges to animal breeders as well as to those concerned with
animal nutrition. Moreover, these questions are relevant to anyone
who is concerned about the maximal use of renewable resources.
Ruminant animals are uniquely equipped to exploit the renewable
low-quality fibrous plant resources of the world and, from them, to
provide many and varied services for mankind.

REFERENCES

Fattet, I., Hovell, F.D. de 8., 0rskov, E.R., Kyle, D.J. & Smart, R.I. (1984)
Undernutrition in sheep. The effect of supplementation on protein accretion
Brit. J. Nutr. 52, 561-574
Kennedy, P.M. & Murphy, M.R. (1988) The nutritional implications of differential
passage of particles through the ruminant alimentary tract Nutr. Res. Rev. 1,
189-208
McDonald, I. (1981) A revised model for the estimation of protein degradability
in the rumen J. agric. Sci., Camb. 96, 251-252
Menke, K.H., Raab, L., Salenski, A., Steingass, H., Fritz, D. & Schneider, W.
(1979) The estimation of the digestibility and metabolizable energy content of
ruminant feedingstuffs from the gas production when they are incubated with
rumen liquor in vitro J. agric. Sci., Camb. 93, 217-222
Mould, F.L., Saadullah, M., Haque, M., Davis, C., Dolberg, F. & 0rskov, E.R.
(1982) Investigation of some of the physiological factors influencing intake and
digestion of rice straw by native cattle of Bangladesh Trap. Anim. Prod. 7,
174-181
0rskov, E.R. (1982) Protein Nutrition in Ruminants Academic Press, London
0rskov, E.R. (1989) Recent advances in evaluation of roughages as feeds for
ruminants In Farrell, D.J. (Ed.) A'dvances in Animal Nutrition pp 102-108
University of New England Printery, Armidale
0rskov, E.R., OJ wang, I. & Reid, G.W. (1988) A study on consistency of
differences between cows in rumen outflow rate of fibrous particles and other
substrates and consequences for digestibility and intake of roughages Anim.
Prod. 47, 45-51
Ramanzin, M., 0rskov, E.R. & Tuah, A.K. (1986) Rumen degradation of straw.
2. Botanical fractions of straw from two barley cultivars Anim. Prod. 43,
271-278
144 Energy Nutrition in Ruminants
Reid, G.W., 0rskov, E.R. & Kay, M. (1988) A note on the effect of variety, type
of straw and ammonia treatment on digestibility and on growth rate in steers
Anim. Prod. 47, 157-160
Tuah, A.K., Lufadeju, E. & 0rskov, E.R. (1986) Rumen degradation of straw.
1. Untreated and ammonia-treated barley, oat and wheat straw varieties and
triticale straw Anim. Prod. 43, 261-269
Van Soest, P.J. (1963) The use of detergents in the analysis of fibrous feeds. 11. A
rapid method for the determination of fibre and lignin J. Assoc. Off. agric.
Chem. 46, 829-835
INDEX

Absorption of nutrients, 23, 44, 52-60 dextrins, 14


Acetonaemia, 91-2 fructose, 19
Anaerobiosis, 21, 23 hemicellulose, 12, 16, 18
lactose, 5, 88, 91
maltose, 5, 14, 57-8
Body reserves pectin, 18
fasting, use during, 64, 67, 89, 94-5 starch, 5, 13, 19, 30-1, 34, 38-9, 57,
fat synthesis and deposition, 88, 59
90-1,93, 142 sucrose, 5, 13, 19, 57-9
lactation, use during, 76-7, 95-7 xylan, 16
late pregnancy, use during, 76, 93
low temperatures, use at, 79
physical work, use for, 73 Digestibility
protein deposition, use for, 95, 142 associative effects, feed mixtures,
Breed and species differences 34-6, 37--40, 138-9
breeds, 69, 79-80, 114-15, 128, 142 cellulose, 21-2, 45-6
species, 8, 50, 60, 80, 99, 103, 107 depression of, 106, 118
late pregnancy, in, 76
milk substitutes, of, 4
Carbohydrates particle outflow rate from rumen
cellobiose, 12, 19 and, 39,48-9,105,114-15,117
cellulose and cellulolysis, 12, 16, prediction of, 135
18-19, 21-2, 29, 32, 37-8,44-7, rumen pH and, 35-6
106-7 starch, of, 38-9
145
146 Index

Environmental temperature prediction of feed intake and


critical temperature, 78-9 growth rate, 108, 110, 133-6,
feed intake and, 78-80 137-9, 140
rumen outflow rate and, 48 soluble fraction of feed, 105-6,
134, 140
history, 123-5
Fatty acids hay and straw units, 123
branched chain, 29, 30, 99 net energy or starch equivalent
free fatty acids, 32, 94 systems, 124--5
hydrogenation of, 19 total digestible nutrient systems,
inhibition of fibre digestion by, 15, 123-4
32 Feeding
lactic acid/lactate, 3, 13, 16, 20, 22, behaviour and rumen pH, 46-7
60 heat increment of, 71, 78
Feed evaluation level, 105, 115
animal requirements and, 141-3 regimes, 37, 97, 99, 131
calorimetry, methods Feed intake, 130
indirect closed-circuit, 126 degradation characteristics and, 113
indirect open-circuit, 126 depression of, 32, 36, 80, 117
open and shut chambers, digestibility and, 48-9, 103-4
126-7 energy requirement and, 113-14,
open-circuit respiration hoods, 117
127 environmental temperature and,
current systems of feed evaluation 78-80, 117-18
limitations, 35, 40, 64-5, 67, 103, extent of digestion and, 105-6
129-31 feed particle size and, 107
metabolizable energy systems, lactation and, 117
128-9 late pregnancy, 76
net energy systems, 127-8 nitrogen supply and, 47, 118
degradation characteristics method outflow rate-from rumen and, 117
degradable fraction of feed, 106, particle size reduction and, 107-8
110, 134, 140-1 physical work and, 119
degradation curve, 104-5, 108, prediction of, 103, 108-10, 133, 135,
134-5 137-9
degradation rate, 105-6, 110, rate of digestion and, 106-7
113, 134, 138, 140-1 recovery from low-level nutrition,
formulae, 105, 133, 135 118-19
index values, 136-7, 142 restriction of, 134
inhibition of degradation, 106 rumen volume and, 113-15, 118
lag phase, 106, 134-5 stimulation of, 39
measurement of characteristics, Feeds
108, 134, 140-1 concentrates, 22, 37, 44-6, 56, 76,
negative associative effects, 138-9 85-6,97,99,113,117-19,124
non-degradable fraction of feed, cereal processing, 30, 34, 37
104-6, 110, 134 cereals, 30, 34-9, 45-6, 69, 78,
outflow rate from rumen, 106-7, 113, 118-19
139-40 digestibility of, 45-6, 103-6, 114-15,
particle size reduction, 105, 107, 117-18
110,139 metabolizability of, 129-30
Index 147

milk replacers, 6 Growth rate prediction, 108, 110, 135,


molasses, 22-3, 31-2, 37-8 137
quality of, 113, 118, 137-8, 142 Gut
roughages, 16,20-2,29-30,32,36-9, distension and feed intake, 103
44-7, 56, 73, 76, 85--6, 91, 97, enzyme development of, 4-5
103-5,108,110,113-15,117, post-ruminal absorption, 56-8, 60
124, 135, 140, 142 post-ruminal digestion, 34, 48-50,
structure of, 22-3, 48, 107, 141 56-8
sugar-cane, 23, 37 post-ruminal fermentation, 48, 50,
supplements, 40, 91-2, 105 58--60
Fermentation size of, 69, 115, 134
hind-gut, in, 5,20-1,49-50, 52,
57--60, 106
rumen, in, Insulin, 29, 87, 97
additives, effects of, 32-3 Intragastric nutrition
associative effects" negative, 34-7 applications of technique, 45, 47,
associative effects, positive, 34-5, 54, 56, 71-2, 77, 86-7, 89,95,
39-40 97
carbon dioxide production, 14, methanogenesis, absence of during,
20, 44 127
cellulose, 12, 18,29, 32, 44, 106-7
end-products of, II, 20
heat of, 24 Metabolizable energy
lipids, 32 efficiency of utilization, 67, 69,
outflow rate from rumen and, 34, 75-8, 85
39, 107 body reserves, of, 76, 79, 94, 96
particle size and, 34, 36 glucose, of, 94
pH of rumen contents and, 30-2, milk, of, 6
35--6, 37, 60 volatile fatty acids, of, 6, 94
rate of, 20-1, 34, 38-9,47 requirement for
soluble carbohydrates, 14, 19, 29, body maintenance, 67-8
31-2,37 draught power, 73
starch, 13, 19, 30-1, 38-9 eating, 71-2
substrate composition, effects of, fat and protein deposition, 74-5
37-8 lactation, 76-7
pregnancy, 75--6
rumination, 72
Glucose standing, 69, 71
absorption, 57 temperature control, high
deficiency, 65, 90, 91-3 temperatures, 79-80
fermentation in hind-gut, 58 temperature control, low
growth, requirement for, 90-1 temperatures, 78-9
lactation, requirement for, 91-2 walking, 72-3
precursors, 12, 34, 64-5, 76, 88-90, wool and fibre production, 77-8
91-4 Micro-organisms
pregnancy, requirement for, 76, 79, acquisition of, by new-born
92-3 ruminants, 7
volatile fatty acid mixtures, yields ATP production by, 14,24
from, 91 bacteria, II, 20, 37-8
148 Index

biomass of, 16-17 Nitrogen


carbon dioxide production by, 14, ammonia as source of, 39, 47
20 excretion in fasting animals, 64-5,
cellulolysis by, 12, 16, 18, 21-2, 23, 67, 89-90, 94
29, 37-8 faecal, 60, 106
ciliate protozoa, 16-19, 30-2, 37 intake of, and roughage
amylopectin and carbohydrate consumption, 118
sequestration by, 17, 19, 21-2 re-cycled urea as source of, 47
defaunation, effects of, 18-21 tissue amino acids as source of, 47,
Entodiniomorphs, 17, 18-21 64
Holotrichs, 17-23 Nylon bag technique: applications,
lipid content of, 19-20 39-40, 105-6, 108, 118, 134,
motility of, 17 140-1
colonization of feed particles, 134
density, 11, 17
fungi, 16, 20-1 Oesophageal groove reflex, 2
hydrogen production by, 14, 20-1, experimental use, 3, 56, 118
23-4
interactions between, 14, 18-19,
21-2,37-8 Pregnancy
lipid degradation by, 15, 19-20, 23 body reserves, use of in, 76, 93
methanogenesis by, 14,20-1, 23-4, feed intake in late, 76
29,33 glucose requirement in, 76, 79, 92-3
multiplication rates of, 13-14, 17 metabolizable energy requirement
nitrogen sources for, 13, 15, 17-18, in, 76
20-1,47 protein requirement in late, 76
oxygen absorption by, 21 toxaemia, 76, 79, 92-3
pH sensitivity of, 13, 17, 22-3, 32
pH stabilization by, 22
protein degradation by, 15,20, Rumen
23 carbon dioxide production, 15-16,
soluble carbohydrate utilization by, 54
14, 16, 19, 21-2, 23 development, 7-8
species and genera, 12-13, 16-17, 18, dry matter concentration in, 48
22 hypertrophy during lactation, 117
starch fermentation by, 13, 16, 19, osmotic pressure of content, 48
23 outflow rate/retention time, 34, 39,
symbiosis of, 21-2 46-9, 105-7, 113-15, 117, 142
volatile fatty acid production by, papillae, 53
14,20,22,23 particle filtration in, 107
volatile fatty acid utilization by, 15 pH, 22-3, 32, 35-7, 44-7, 53-4, 56
Milk volume, 2,7,113-15,117-18
composition, 29, 87-8, 96-7 Rumenitis, 47
replacers, 6 Ruminants
undernourished animals, basal metabolism, 64-5
production by, 91-2 behaviour, 107
volatile fatty acid ratios, effects on energy maintenance, 64, 67-8,
production, 85 89-90, 138
yield, 29, 97 fasting metabolism, 64-5, 67, 130
Index 149

fat utilization, 64 acetic acid/acetate, 12-16, 20,23,


individual differences, 32, 48-50, 29, 32, 53-4, 85-9
110, 114-15 butyric acid/butyrate, 14,20,29, 31,
tissue protein utilization, 47, 64 33,53
volatile fatty acid utilization, 84-8 concentration, 31
Rumination time, 30, 44-6, 72 efficiency of utilization, 85
energy ratios, 85--6, 89-90
heat production and, 86--7
Saliva, 30
milk composition and, 87~8
composition, 22, 44
molar ratios, 29, 53-4, 85-6
secretion rate, 32
pK of, 45
urea in, 47
production, 14, 15-16,45, 52, 58, 60,
volume, 22, 30, 36, 44-6
97
Soft-fat syndrome, 99
propionic acid/propionate, 12, 14,
Stoichiometry of rumen fermentation, 20,23-4,29-31,33,53,85
23-4 glucose precursor, as, 12, 64,
88-90,93
Volatile fatty acid insulin, and, 29, 87, 97, 99
absorption, 44-5, 52-3, 56, 58, 71 rumen development and, 7-8
blood flow, effect of, 54, 56 utilization by rumen bacteria, 15
rumen pH, effect of, 53-4 utilization by ruminants, 84-8

You might also like