Você está na página 1de 123

A

G U I D E

TO

D I S T I L L E R S

D R I E D

G R A I N S

W I T H

S O L U B L E S

User Handbook

(D D G S)

DDGS Handbook Table of Contents

Introduction

Tab 1

Use of DDGS in Beef Diets

Tab 2

Use of DDGS in Dairy Diets

Tab 3

Use of DDGS in Poultry Diets

Tab 4

Use of DDGS in Swine Diets

Tab 5

Use of DDGS in Aquaculture Diets

Tab 6

Use of DDGS in Companion Animal Diets

Tab 7

Physical & Chemical Characteristics of DDGS

Tab 8

Nutrient Composition of DDGS: Variability and Measurement

Tab 9

Factors that Impact DDGS Pricing & Transportation

Tab 10

Ethanol Production and its Co-Products

Tab 11

Frequently Asked Questions about DDGS

Tab 12

U.S. DDGS Suppliers List

Tab 13

Glossary of Terms

Tab 14

Website Links

Tab 15

An Introduction to U.S. DDGS


Distillers dried grains with solubles (DDGS) is a valuable feed ingredient which is a coproduct of drymill ethanol production from grains. In ethanol production, the starch is fermented
to obtain ethyl alcohol, but the remaining components of the grain kernel (endosperm, germ),
preserve much of the original nutritional value of the grain, including energy, protein and
phosphorous. Drymill plants recover and recombine these components into a variety of animal
feed ingredients. DDGS is a popular dried form of these combined components, available to
domestic and international customers as an ingredient for livestock and poultry rations. As the
U.S. ethanol industry continues to grow, a greater quantity of DDGS will be available for feeds
in the domestic and export market and a wider diversity of distillers co-products with different
nutritional characteristics will become available for specific animal feeding applications.
Corn is the primary feedstock for drymill ethanol production in the United States. In certain
locations sorghum and other grains may also be used. Every bushel of grain (25.4 kg of corn and
sorghum, slightly different weight for other grains) in the process produces 11.8 liters (2.7
gallons) of ethanol and 7.7 kg (18 pounds) of DDGS. The ethanol industry in the United States is
expanding rapidly, resulting in a fast-growing supply of DDGS in the marketplace. In January
2007, the Renewable Fuels Association reported that 112 operating drymill ethanol plants have a
combined capacity of 5.53 billion gallons of ethanol annually, and that 83 more plants are either
under construction or expanding, which could add another 6.0 billion gallons of production
capacity within the next two years. DDGS production from these ethanol plants reached 8.5
million metric tons in calendar year 2006, and is expected to climb to 36 million tons by 2010.
DDGS offers an opportunity for cost savings in animal feed rations, and will be available in
abundant quantities in coming years.
This DDGS User Handbook is intended as a guide to feed manufacturers and animal producers,
enabling them to understand how DDGS may fit into feed rations for livestock, poultry and fish,
and how to purchase and handle DDGS. The handbook includes information on current research
regarding DDGS use in cattle, swine, poultry, fish and companion animals. Other chapters
describe the variability and measurement of nutritional characteristics of DDGS, and provide
information on buying DDGS from the United States.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be formulated by a
qualified nutritionist. The USGC has no control over the nutritional content of any specific product which may be
selected for feeding. Potential buyers should consult an appropriate nutritionist for specific recommendations.
USGC makes no warranties that these recommendations are suitable for any particular herd or for any particular
animal. The USGC disclaims any liability for itself or its members for any problems encountered in the use of these
recommendations. By reviewing this material, buyers agree to these limitations and waive any claims against USGC
for liability arising out of this material.

For more information, please contact the U.S. Grains Council at 202-789-0789 or email
grains@grains.org. You may also visit our website at www.grains.org.

01 - Introduction

Use of DDGS in
B e e f D i e ts

User Handbook

Use of U.S. Distillers Co-Products in


Beef Cattle Diets
The U.S. beef cattle industry has been a major consumer of wet and dried corn distillers coproducts for decades. As a result, there has been a considerable amount of research conducted to
evaluate the feeding value of corn distillers co-products to cattle. Most of the research that has
been conducted on feeding distillers grains to finishing beef cattle. Several excellent research
summaries and feeding recommendations have been published (Erickson et al., 2005; Tjardes
and Wright, 2002; Loy et al., 2005a; Loy et al., 2005b; Schingoethe, 2004).

Nutrient Composition of Distillers Co-Products for Beef Cattle


There are several different forms of distillers co-products produced in dry grind ethanol
plants. The liquid that is removed from the mash is called thin stillage, which can be returned to
the cooking and distillation processes, sold directly as high moisture cattle feed or dehydrated to
produce condensed distillers solubles (CDS). The residual solids or coarse grains fraction is
referred to as wet distillers grains, and can be used as cattle feed or dried to produce dried
distillers grains (DDG). Condensed distillers solubles can be used as cattle feed or blended with
distillers grains to produce distillers grains with solubles. Distillers grains with solubles are
sold wet (WDGS; 30% dry matter), modified (MDGS; 50% dry matter), or dried (DDGS; 90%
dry matter). Because there are several wet and dry forms of distillers co-products available, it is
important to obtain an actual nutrient analysis of the co-products intended to be used because the
nutrient content can vary widely among. Some of the reasons for variation in nutrient content of
distillers co-products include variation in fermentation and distillation efficiencies, different
drying processes and temperatures and the amount of condensed distillers solubles blended into
various co-products. Commonly reported nutrient values for several distillers co-products are
shown in Table 1 (Tjardes and Wright, 2002).
Table 1: Concentrations of Selected Nutrients in Various Corn Distillers Co-products
(100% Dry Matter Basis).
Nutrient
CDS1
WDG2
MDGS3
DDG4
DDGS5
30-50
25-35
50
88-90
88-90
Dry Matter, %
20-30
30-35
30-35
25-35
25-32
Crude protein, %
50
45-53
45-53
40-50
43-53
Degradable intake protein,
% of CP
9-15
8-12
8-12
8-10
8-10
Fat, %
10-23
30-50
30-50
40-44
39-45
NDF, %
75-120
70-110
70-110
77-88
85-90
TDN, %
2.21-2.54
1.98-2.43
1.98-2.43 1.96-2.21 2.16-2.21
NEm, Mcal/kg
1.76-2.05
1.54-1.76
1.54-1.76 1.48-1.54 1.50-1.54
NEg, Mcal/kg
0.03-0.17
0.02-0.03
0.02-0.03 0.11-0.20 0.17-0.26
Calcium, %
1.30-1.45
0.50-0.80
0.50-0.80 0.41-0.80 0.78-1.08
Phosphorus, %
1
4
Condensed distillers solubles.
Dried distillers grains.
2
5
Wet distillers grains.
Distillers dried grains with solubles.
3
Adapted from Tjardes and Wright (2002).
Modified distillers grains with solubles.

02 - Use of DDGS in Beef Diets

Thin stillage contains only 5-10% dry matter and can be successfully fed to beef cattle as a
water replacement. CDS provide significant amounts of protein and energy to the diet, and is
often used to add moisture to condition diets. However, its use should be limited to no more than
20% of diet dry matter because its high fat content can depress fiber intake and digestion
(Tjardes and Wright, 2002).
Distillers grains with or without solubles is an excellent energy source for cattle. In the United
States, finishing beef cattle have successfully been fed as much as 40% DDGS of ration dry
matter as a replacement for corn grain. When adding DDGS to the diet at this level, it is used
primarily as an energy source, and supplies more protein and phosphorus than required for
finishing feedlot cattle. In one research study (Ham et al., 1994), the net energy gain (NEgain) of
DDGS for beef cattle was 21% higher than the value of dry-rolled corn. Conservatively, most
nutritionists consider DDGS to have an apparent energy value equal to corn grain when fed to
finishing cattle at levels ranging from 10-20% of total ration dry matter. In many studies, feeding
DDGS at levels of 15-20% of the diet dry matter improved growth rate and feed conversion of
finishing beef cattle compared to when diets containing corn grain were fed. This performance
improvement is often a result of reduced sub-acute acidosis and fewer problems with cattle going
off-feed. The starch in corn grain is more likely to cause acidosis, laminitis and fatty liver
when fed at high levels to finishing beef cattle. However, these potential problems are greatly
reduced when feeding DDGS because of the low residual starch content (less than 2%) and the
high amount of highly digestible fiber.
Distillers grains with or without solubles are a very good protein source and are high in
ruminally undegradable protein (RUP), or bypass protein. Since DDGS goes through a drying
process, there is potential for burning which can cause a chemical reaction called the Maillard or
browning reaction. When this reaction occurs, it causes some of the carbohydrate and protein to
be bound in a chemical form that makes it unavailable to the animal. Therefore, light colored
DDGS that has a sweet and fermented smell should be used to achieve the best feeding value and
growth performance for beef cattle. Marketers of DDGS often discount the price of dark and heat
damaged DDGS to account for the reduction in feeding value. Acid detergent insoluble nitrogen
(ADIN) can be used to determine the extent of protein damage in DDGS. Once the ADIN value
is determined in the laboratory, this value is multiplied by a factor of 6.25 to calculate the
appropriate protein value for DDGS. This calculated protein value represents the amount of
crude protein in DDGS that is unavailable and can be compared to the actual crude protein value
to determine the extent of protein damage. The proportion of RUP in DDGS is approximately
60-70% compared to 30% for soybean meal. However, Erickson et al. (2005) indicated that the
high RUP value of DDGS is due to the innate characteristics of the protein rather than drying or
moisture content, and does not appear to be influenced by ADIN since protein efficiency (kg
gain/kg supplemental protein) appears to stay the same or increase as the amount of ADIN in
DDGS increased.
Distillers grains, with or without solubles, are low in calcium but high in phosphorus and
sulfur. Depending upon the feeding level, adding distillers grains to the diet may allow complete
removal of other supplemental phosphorus sources from the mineral mixture previously fed. Due
to the high levels of wet or dried DGS fed, beef cattle feedlot diets contain excess phosphorus
relative to their requirement. This results in excess phosphorus being excreted in manure and
must be considered when developing manure management plans. Due to the low calcium level of
DDGS, supplemental calcium sources (e.g. ground limestone or alfalfa) must be added to the diet
to maintain a calcium to phosphorus ratio between 1.2:1 to no more than 7:1 to avoid reductions
02 - Use of DDGS in Beef Diets

in animal performance and urinary calculi (Tjardes and Wright, 2002). Distillers grains with and
without solubles can sometimes be high in sulfur and contribute significant amounts of sulfur to
the diet. If more than 0.4% sulfur from feed (dry matter basis) and water is consumed,
polioencephalomalacia in cattle can occur. Furthermore, sulfur interferes with copper absorption
and metabolism, which is worsened in the presence of molybdenum. Therefore, in geographic
regions where high sulfur levels are found in forages and water, the level of DDGS that can be
added may need to be reduced (Tjardes and Wright, 2002).

Finishing Cattle
Most of the DDGS research has involved using it primarily as an energy source in diets for
finishing cattle. DDGS is very palatable and readily consumed by beef cattle. Furthermore,
feeding DDGS does not change the quality or yield of beef carcasses, and it has no effect on the
sensory or eating characteristics of beef. Feeding WDGS results in better performance than
feeding DDGS to finishing cattle (Erickson et al., 2005). Replacement of corn with wet distillers
grains has consistently resulted in a 15-25% improvement in feed conversion when 30-40% of
corn is replaced with WDGS in the diet (DeHaan et al., 1982; Farlin, 1981; Firkins et al., 1985;
Fanning et al., 1999; Larson et al., 1993; Trenkle, 1997a; Trenkle 1997b; Vander Pol et al.,
2005a). This improvement in feed conversion is primarily due to WDGS having 120-150% of
the energy value of corn (Erickson et al., 2005). Drying appears to reduce the energy value to
102-127% of the energy value of dry rolled corn in high forage diets. It appears that the high
energy values of WDGS and DDGS are a result of acidosis control (Erickson et al., 2005).
Vander Pol et al. (2005c) showed that when finishing cattle are fed diets containing 10-20%
DDGS of diet dry matter, there is no benefit for supplementing diets with urea, suggesting that
nitrogen recycling was occurring. However, Erickson et al. (2005) suggested that to be
conservative, it may be best to follow National Research Council (1996) guidelines for
degradable intake protein supplementation when formulating diets containing less than 20%
DDGS.
A few studies have evaluated the quality and sensory characteristics of beef from cattle fed
distillers grains. Roeber et al. (2005) evaluated beef color, tenderness and sensory
characteristics of beef strip loins from two experiments where wet or dry distillers grains were
fed to Holstein steers at levels up to 50% of the ration. There were no differences in tenderness,
flavor or juiciness. Similarly, Jenschke et al. (2006) showed that finishing beef cattle fed diets
containing up to 50% wet distillers grains (dry matter basis) produced steaks that did not differ
in tenderness, amount of connective tissue, juiciness or off-flavor intensity. In fact steaks from
cattle fed 0-10% wet distillers grains diets were most likely to have an off-flavor compared to
steaks from cattle fed 30-50% wet distillers diets. Finally, Gordon et al. (2002) fed diets
containing 0, 15, 30, 45, 60, or 75% DDGS to finishing heifers during a 153 day finishing trial
and observed that there was a small linear improvement in tenderness of steaks from cattle fed
increasing amounts of DDGS.
Less research has been conducted related to feeding DDGS to other ages of cattle. However,
DDGS is an excellent feed ingredient to supplement energy and protein when cattle are fed low
quality forages. When added to diets containing forages low in phosphorus, the phosphorus in
DDGS will be of significant value. Other potential uses of DDGS include using it as a creep feed

02 - Use of DDGS in Beef Diets

for nursing calves, a supplement for grazing cattle and a supplement for low quality forages and
crop residues that might be fed to growing calves, gestating beef cows or developing beef
heifers.

Beef Cows
Unlike for finishing beef cattle, less research has been conducted on feeding DDGS to beef
cows. Loy et al. (2005) published an excellent summary of results on DDGS in beef cow diets.
The best applications for DDGS in beef cow diets are in situations where 1) supplemental protein
is needed (especially when feeding low quality forages) to replace corn gluten feed or soybean
meal, 2) a low starch, high fiber energy source is needed to replace corn gluten feed or soy hulls
and 3) when a source of supplemental fat is needed.
DDGS as a Supplemental Protein Source
Researchers have shown that when DDGS was supplemented to provide 0.18 kg of protein/day
to beef cows grazing on a native winter range in Colorado, it compared favorably to alfalfa hay
or cull navy beans (Smith et al., 1999). Shike et al. (2004) compared performance effects of
feeding corn gluten feed or DDGS as a supplement to ground alfalfa hay to lactating Simmental
cows and observed that cows fed DDGS gained more weight, but produced less milk compared
to cows fed corn gluten feed. However, there were no differences between cows fed DDGS and
those fed corn gluten feed on calf weights and rebreeding performance. Loy et al. (2005)
reported that in a subsequent study conducted at the University of Illinois, researchers compared
supplementing diets for lactating Angus and Simmental cows consisting of ground corn stalks
with either DDGS or corn gluten feed. Cows nursing calves were limit-fed total mixed rations
and there were no differences in milk production and calf weight gains between cows
supplemented with DDGS or corn gluten feed.
DDGS as an Energy Source
DDGS is an effective energy supplement when fed with low quality forages. Summer and
Trenkle (1998) showed that DDGS and corn gluten feed were superior supplements to corn in
corn stover diets, but not in the higher quality alfalfa diets. Corn stover (stalks) are low in
protein, energy and minerals, but are low in cost and readily available in major corn producing
states in the United States. When low quality forages (e.g. corn stover) are fed to gestating beef
cows in good condition, feeding 1.4 to 2.3 kg of DDGS per day, during the last 1/3 of gestation
will meet their protein and energy requirements (Loy et al., 2002). For beef cows fed low quality
forage (e.g. corn stalks) in early lactation, supplementing with 2.7 to 3.6 kg of DDGS will meet
their protein and energy requirements (Loy et al., 2002).
DDGS as a Supplemental Fat Source
Supplemental fat may improve reproduction in cow herds experiencing suboptimal pregnancy
rates (less than 90%). Loy et al. (2002) indicated that feeding supplements with similar fatty acid
profiles to corn oil (found in DDGS), improved pregnancy rates. They also showed that fat
supplementation works best in feeding situations where protein and/or energy supplementation is
already necessary.

02 - Use of DDGS in Beef Diets

Replacement Heifers
Very little research has been conducted on feeding DDGS to replacement heifers. However,
based upon numerous studies for finishing cattle, DDGS would be an excellent source of RUP
and energy for developing replacement heifers. In a study by MacDonald and Klopfenstein
(2004), replacement heifers grazing brome grass were supplemented with 0, 0.45, 0.90, 1.36, or
1.81 kg DDGS per day. These researchers observed that for each 0.45 kg of DDGS
supplemented, forage consumption decreased by 0.78 kg per day and average daily gain
increased by 27 g per day.
Loy et al. (2003) evaluated the value of supplementing the ration, daily or three times per
week, with DDGS in high forage diets for growing crossbred heifers. These heifers were
provided ad libitum access to grass hay (8.7% crude protein) and were supplemented with DDGS
or dry rolled corn. The supplements were fed at two levels and offered either daily or three times
per week in equal proportions. For heifers that were supplemented daily, they ate more hay and
gained faster, but were not more efficient than heifers supplemented three times per week. At
both the low and high supplementation levels, heifers fed DDGS had better average daily gain
ADG and feed conversion than heifers fed the dry rolled corn (Table 2). These authors calculated
that the net energy value of DDGS was 27% higher than for corn grain.
Table 2: Growth performance of growing heifers fed native grass hay and supplemented
with either corn or DDGS for at two supplementation levels.
Lowa
Highb
0.37
0.71
ADG, kg/d
Corn
0.45
0.86
DDGS
15.9
9.8
DM Intake/ADG
Corn
12.8
8.0
DDGS
a
Low = supplement fed at 0.21% of body weight
b
High = supplement fed at 0.81% of body weight
Source: Loy et al. (2003a).
In a subsequent study, Loy et al. (2004) fed cannulated heifers either no supplement, DDGS
supplemented daily, DDGS supplemented alternating days, dry rolled corn daily, or dry rolled
corn on alternating days. As expected, hay intake was higher for heifers that received no
supplementation compared to those that did, but there were no differences in feed intake between
heifers supplemented with DDGS or corn. Heifers that were supplemented with DDGS had
higher rates of rumen fiber disappearance than heifers supplemented with corn.
Stalker et al. (2004) conducted two experiments to evaluate the effects of supplemental
degradable protein requirements when DDGS was fed as an energy source in forage based diets.
Diets were formulated to be deficient (less than 100 g/day) in degradable protein but contained
excess metabolizable protein. Their results showed that adding urea to meet the degradable
protein intake requirement is not necessary when DDGS is used as an energy source in forage
based diets.
Morris et al. (2005) showed that when individually fed heifers were provided high or low
quality forage diets that supplementation with either 0, 0.68, 1.36, 2.04, or 2.72 kg DDGS per
day that forage intake decreased and average daily gain increased. These results suggest that
DDGS can be an effective forage supplement to increase growth at times when availability of
forage may be limited.
02 - Use of DDGS in Beef Diets

Summary
Corn DDGS is an excellent energy and protein source for beef cattle in all phases of
production. It can effectively be used as an energy source and be fed up to 40% of ration dry
matter intake for finishing cattle with excellent growth performance and carcass and meat
quality. However, at this high feeding rate excess protein and phosphorus will be fed.
The best applications for using DDGS in beef cow diets are in situations where 1)
supplemental protein is needed (especially when feeding low quality forages) to replace corn
gluten feed or soybean meal, 2) a low starch, high fiber energy source is needed to replace corn
gluten feed or soy hulls and 3) when a source of supplemental fat is needed.
For growing heifers, adding urea to meet the degradable protein intake requirement is not
necessary when DDGS is used as an energy source in forage based diets. DDGS can be an
effective forage supplement to increase growth at times when availability of forage may be
limited.

Literature Cited
Bremer, V.B., G.E. Erickson, T.J. Klopfenstein, M.L. Gibson, K.J. Vander Pol, M.A. Greenquist. 2005. Feedlot
performance of a new distillers byproduct (Dakota Bran) for finishing cattle. J. Anim. Sci. 83:(Suppl. 1).
Cooper, R.J., C.T. Milton, T.J. Klopfenstein, T.L. Scott, C.B. Wilson, and R.A. Mass. 2002. Effect of corn
processing on starch digestion and bacterial crude protein flow in finishing cattle. J. Anim. Sci. 80:797-804.
DeHaan, K., T. Klopfenstein, R. Stock, S. Abrams, and R. Britton. 1982. Wet distillers co-products for growing
ruminants. Nebraska Beef Rep. MP 43:33.
Erickson, G.E., T.J. Klopfenstein, D.C. Adams, and R.J. Rasby. 2006. Utilization of Corn Co-Products in the Beef
Industry. Nebraska Corn Board and the University of Nebraska. www.nebraskacorn.org. 17 pp.
Fanning, K., T. Milton, T. Klopfenstein, and M. Klemesrud. 1999. Corn and sorghum distillers grains for
finishing cattle. Nebraska Beef Rep. MP 71 A:32.
Farlin, S.D. 1981. Wet distillers grains for finishing cattle. Anim. Nutr. Health 36:35.
Firkins, J.L., L.L. Berger, and G.C. Fahey, Jr. 1985. Evaluation of wet and dry distillers grains and wet and dry
corn gluten feeds for ruminants. J. Anim. Sci. 60:847.
Gordon, C.M., J.S. Drouillard, R.K Phebus, K.A. Hachmeister, M.E. Dikeman, J.J. Higgins, and A.L. Reicks.
2002. The effect of Dakota Gold Brand dried distillers grains with solubles of varying levels on sensory and color
characteristics of ribeye steaks. Cattlemans Day 2002, Report of Progress 890. Kansas State University. pp. 72-74.
Gustad, K., T.J. Klopfenstein, G. Erickson, J. MacDonald, K. Vander Pol, and M. Greenquist. 2006. Dried
distillers grains supplementation to calves grazing corn residue.
Ham, G.A., R.A. Stock, T.J. Klopfenstein, E.M. Larson, D.H. Shain, and R.P. Huffman. 1994. Wet corn distillers
co-products compared with dried distillers grains with solubles as a source of protein and energy for ruminants. J.
Anim. Sci. 72:3246.
Holt, S.M., and R.H. Pritchard. 2004. Composition and nutritive value of corn co-products from dry milling
ethanol plants. South Dakota State Beef Report.
Jenschke, B.E., J.M. James, K.J. Vander Pol, C.R. Calkins, and T.J. Klopfenstein. 2006. Wet distillers grains plus
solubles do not increase liver-like off-flavors in cooked beef. Nebraska Beef Report, University of NebraskaLincoln, pp. 115-117.
Larson, E.M., R.A. Stock, T.J. Klopfenstein, M.H. Sindt, and R.P. Huffman. 1993. Feeding value of wet distillers
co-products from finishing ruminants. J. Anim. Sci. 71:2228.
Loy, T.W., T.J. Klopfenstein, G.E. Erickson, and C.N. Macken. 2003. Value of dry distillers grains in high fiber
diets and effect on supplementation frequency. Nebraska Beef Cattle Report MP 80-A:8.

02 - Use of DDGS in Beef Diets

Loy, T.W., J.C. MacDonald, T.J. Klopfenstein, and G.E. Erickson. 2004. Effect of distillers grains or corn
supplementation frequency on forage intake and digestibility. Nebraska Beef Cattle Report MP 80-A:22-24.
Loza, P.L., K.J. Vander Pol, G.E. Erickson, R.A. Stock, and T.J. Klopfenstein. 2004. Corn milling co-products
and alfalfa levels in cattle finishing diets. J. Anim. Sci. 82 (Suppl. 1):158.
MacDonald, J.C. and T.J. Klopfenstein. 2004. Dried distillers grains as a grazed forage supplement. Nebraska
Beef Cattle Report MP 80-A:22-24.
Morris, S.E., T.J. Klopfenstein, D.C. Adams, G.E. Erickson, and K.J. Vander Pol. 2005. The effects of dried
distillers grains on heifers consuming low or high quality forages. Nebraska Beef Report MP 83-A:18-20.
NRC. 1996. Nutrient Requirements of Beef Cattle (7th ed.). National Academy Press, Washington, DC.
Owens, F.N., D.S. Secrist, W.J. Hill, and D.R. Gill. 1997. The effect of grain source and grain processing on
performance of feedlot cattle: a review. J. Anim. Sci. 75:868-879.
Roeber, D.L., R.K. Gill, and A DiCostanzo. 2005. Meat quality responses to feeding distillers grains to finishing
Holstein steers. J. Anim. Sci. 83:2455-2460.
Shike, D.W., D.B. Faulkner, and J.M. Dahlquist. 2004. Influence of limit-fed dry corn gluten feed and distillers
dried grains with solubles on performance, lactation, and reproduction of beef cows. J. Anim. Sci. 82 (Suppl. 2):96.
Smith, C.D., J.C. Whitlier, D.N. Schutz, and D. Conch. 1999. Comparison of alfalfa hay and distillers dried
grains with solubles alone and in combination with cull beans as protein sources for beef cows grazing native winter
range. Beef Program Report. Colorado St. Clin.
Stalker, L.A., T.J. Klopfenstein, D.C. Adams, and G.E. Erickson. 2004. Urea inclusion in forage-based diets
containing dried distillers grains. Nebraska Beef Cattle Report MP 80-A:20-21.
Stalker, L.A., D.C. Adams, and T.J. Klopfenstein. 2006. A system for wintering beef heifers using dried distillers
grain. Nebraska Beef Report MP 88-A:13.
Stock, R.A., J. M. Lewis, T.J. Klopfenstein, and C.T. Milton. 1999. Review of new information on the use of wet
and dry milling feed by-products in feedlot diets. Proc. Am. Soc. Anim. Sci. Available at:
http://www.asas.org/jas/symposia/proceedings/0924.pdf.
Summer, P., ans A. Trenkle. 1998. Effects of supplementing high or low quality forages with corn or corn
processing co-products upon digestibility of dry matter and energy by steers. Iowa State University Beef Research
Report ASL-R1540.
Tjardes, J. and C. Wright. 2002. Feeding corn distiller's co-products to beef cattle. SDSU Extension Extra. Ex
2036, August 2002. Dept. of Animal and Range Sciences. pp. 1-5.
Trenkle, A. 1997a. Evaluation of wet distillers grains in finishing diets for yearling steers. Beef research Report
Iowa State University ASRI 450.
Trenkle, A. 1997b. Substituting wet distillers grains or condensed solubles for corn grain in finishing diets for
yearling heifers. Beef Research report Iowa State University ASRI 451.
Vander pol, K.J., G. Erickson, T. Klopfenstein, and M. Greenquist. 2005a. Effect of level of wet distillers grains
on feed lot performance of finishing cattle and energy value relative to corn. J. Anim. Sci. 83(Suppl. 2):25.
Vander Pol, K.J., G.E. Erickson, and T. Klopfenstin. 2005b. Economics of wet distillers grains use in feedlot
diets. J. Anim. Sci. 83(Suppl. 2):67.
Vander Pol, K.J., G.E. Erickson, and T.J. Klopfenstein. 2005c. Degradable intake protein in finishing diets
containing dried distillers grains. J. Anim. Sci. 83(Suppl. 2):62.
Vander Pol, K.J., G.E. Erickson, M.A. Greenquist, and T.J. Klopfenstein. 2006. Effect of Corn Processing in
Finishing Diets Containing Wet Distillers Grains on Feedlot Performance and Carcass Characteristics of Finishing
Steers. 2006 Nebraska Beef Report.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in understanding
generally-accepted feeding levels. However, all rations for specific herds should be formulated by a qualified nutritionist.
The USGC has no control over the nutritional content of any specific product which may be selected for feeding. Potential
buyers should consult an appropriate nutritionist for specific recommendations. USGC makes no warranties that these
recommendations are suitable for any particular herd or for any particular animal. The USGC disclaims any liability for
itself or its members for any problems encountered in the use of these recommendations. By reviewing this material,
buyers agree to these limitations and waive any claims against USGC for liability arising out of this material.

02 - Use of DDGS in Beef Diets

Use of DDGS in
Da i ry D i e ts

User Handbook

Use of U.S. DDGS in Dairy Cattle Rations


Distillers dried grains with solubles (DDGS) is a very good protein source for dairy cows. The
protein content in high quality DDGS is typically more than 30% on a dry matter. DDGS is a
good source of ruminally undegradable protein (RUP), or by-pass protein, for cattle (Table 1).
Most of the readily degradable protein in corn is degraded during the fermentation process,
resulting in a proportionately higher level of RUP than found in corn. The quality of protein in
DDGS is fairly good, but as for most corn co-products, lysine is the first limiting amino acid. As
a result, milk production can sometimes be increased when dairy cows are fed rations containing
supplemental ruminally protected lysine and methionine, or when DDGS is blended with other
high protein ingredients that contain more lysine. However, in most situations feeding rations
containing DDGS results in milk production being as high, or higher, than when dairy cows are
fed rations containing soybean meal as the protein source. It also is important to recognize that
dark colored corn DDGS usually indicates heat damage of the protein, which may lead to
reduced milk production. In a study by Powers et al. (1995), dairy cows fed diets containing dark
colored DDGS had lower milk production than cows fed diets containing light colored DDGS.
Therefore, it is important to use high quality sources of light colored DDGS in dairy cows diets
to achieve maximum milk production.
DDGS is also a very good energy source for dairy cattle. Energy values for high quality DDGS
are 10-15% higher than values previously reported by the National Research Council (NRC,
2001). DDGS contains more energy than corn. Furthermore, because almost all of the starch in
corn is converted to ethanol during the fermentation process, the fat and fiber concentrations in
DDGS are increased by a factor of three compared to corn. DDGS contains high amounts of
neutral detergent fiber (NDF) but low amounts of lignin. This makes DDGS a highly digestible
fiber source for cattle, and reduces digestive upsets compared to corn. The highly digestible fiber
in DDGS also allows it to serve as a partial replacement for forages and concentrates in diets for
dairy and beef cattle.
Table 1. Nutrient Composition of High Quality U.S. Corn DDGS for Ruminants.
Nutrient
Corn DDGS (% of Dry Matter)
Crude protein
30.1
a
RUP % of crude protein
55.0
NEmaintenance, Mcal/kg
2.07
NEgain, Mcal/kg
1.41
NElactation, Mcal/kg
2.26
Neutral detergent fiber (NDF)
41.5
Acid detergent fiber (ADF)
16.1
Ether extract
10.7
Ash
5.2
Calcium
0.22
Phosphorus
0.83
Magnesium
0.33
Potassium
1.10
Sodium
0.30
Sulfur
0.44
a

RUP = ruminally undegradable protein

03 - Use of DDGS in Dairy Diets

Source: Schingoethe (2004)

Many questions have been asked regarding the inclusion of distillers grains in dairy cow diets.
Can distillers grains-based diets support the same level of milk production as traditional dairy
cow diets? Does the high concentration of polyunsaturated fat in distillers grains cause milk fat
depression? Does the low concentration of lysine result in lower milk protein production? Does it
matter whether distillers grains are fed as wet distillers grains with solubles (WDGS) or as
distillers dried grains with solubles (DDGS)? How much distillers grains can be included into
dairy cow diets and what effect does it have on milk fat and protein composition and overall milk
production?
To answer these questions, a meta-analysis of previous experiments that involved feeding
distillers grains to lactating dairy cows was conducted (Kalscheur, 2005). Twenty-three studies
investigating the inclusion of distillers grains in dairy cow diets were compiled into a database
with 96 treatment comparisons. These studies were published between 1982 and 2005, and it is
recognized that distillers grains quality has changed over this time period. All studies were
included in the analysis to determine the overall effect of feeding distillers grains to dairy cows.
To evaluate the level of inclusion on lactation performance, treatments were divided into five
distillers grains dietary inclusion level ranges, 0, 4-10%, 10-20%, 20-30%, and more than 30%
on a dry matter basis. The form of the distillers grains wet or dried was also identified. The
impact of dietary inclusion level and form of distillers grains was evaluated on dry matter
intake, milk production and milk fat and protein percentage.

Effect of feeding distillers grains on dry matter intake


Dry matter intake (DMI) was affected by both dietary inclusion level and by the form of the
distillers grains (Table 2). Intake was increased by the addition of distillers grains in dairy cow
diets. For cows fed DDGS, intake increased as the dietary DDGS inclusion level increased, and
was greatest for cows fed 20-30% DDGS. These cows consumed 0.7 kg more feed (dry matter,
or DM, basis) than cows fed the control diets containing no DDGS. Cows fed more than 30%
DDGS consumed about the same amount of feed as cows which consumed control diets.
While diets with DDGS up to the 20-30% inclusion rate stimulated DMI, DMI of cows fed
DWGS diets was greatest at lower inclusion levels, 4-10% and 10-20% rate. When WDGS was
included at concentrations greater than 20%, DMI decreased. In addition, cows fed more than
30% WDGS ate 2.3 kg/d less than the control group, and 5.1 kg/d less than those fed the 4-10%
inclusion rate.
In general, distillers grains are considered to be highly palatable, and research supports this
because DMI is stimulated when distillers grains are included up to 20% of the DM in dairy
cow diets. Decreased intake at higher inclusion levels may be caused by higher dietary fat
concentrations, or in the case of WDGS, high dietary moisture concentrations.

03 - Use of DDGS in Dairy Diets

Table 2: Dry matter intake and milk yield of dairy cows fed increasing levels of distillers
grains as either dried or wet.
Inclusion level
(DM basis)
0%
4 10%
10 20%
20 30%
> 30%
SEM

Dried
23.5c
23.6bc
23.9ab
24.2a
23.3bc
0.8

DMI, kg/d
Wet
20.9b
23.7a
22.9ab
21.3ab
18.6c
1.3

All
22.2b
23.7a
23.4ab
22.8ab
20.9c
0.8

Dried
33.2
33.5
33.3
33.6
32.2
1.5

Milk, kg/d
Wet
31.4
34.0
34.1
31.6
31.6
2.6

All
33.0
33.4
33.2
33.5
32.2
1.4

a,b,c

Values within a column followed by a different superscript letter differ (P < 0.05). No superscript
within a column indicates that there was no significant difference between distillers grains dietary
inclusion level.

Effect of feeding distillers grains on milk production


Milk production was not impacted by the form of distillers grains fed, but there was a
curvilinear response to increasing distillers grains in dairy cow diets (Table 2). Cows fed diets
containing 4-30% distillers grains produced the same amount of milk, approximately 0.4 kg/d
more, than cows fed diets containing no distillers grains. When cows were fed the highest
inclusion rate (more than 30%) of distillers grains, milk yield tended to decrease. These cows
produced 0.8 kg/d less milk than cows fed no distillers grains. Cows fed more than 20% WDGS
decreased in milk production. This was most likely related to the decreased DMI.

Effect of feeding distillers grains on milk composition


Milk fat percentage varied among inclusion levels and was not significantly affected by
inclusion level or form (Table 3). With the current data set, the inclusion of distillers grains does
not support the theory that feeding distillers grains results in milk fat depression. Many factors
play an important role in causing milk fat depression. When formulating diets, it is important to
include sufficient fiber from forages in order to maintain rumen function. Distillers grains
provide 28-44% NDF, but this fiber is finely processed and rapidly digested in the rumen. As
such, fiber from distillers grains is not considered ruminally effective fiber and should not be
considered equal to forage fiber. High levels of fat provided from distillers grain may also
impact rumen function leading to milk fat depression, but it is often a combination of dietary
factors which lead to significant reduction in milk fat percentage.
Table 3. Milk fat and protein percentage from
dairy cows fed increasing levels of distillers grains.
Inclusion level
(DM basis)
Fat, %
Protein, %
0%
3.39
2.95a
4 10%
3.43
2.96a
10.1 20%
3.41
2.94a
20.1 30%
3.33
2.97a
> 30%
3.47
2.82b
SEM
0.08
0.07
03 - Use of DDGS in Dairy Diets

a,b

Values within a column followed by a different superscript letter differ (P < 0.05). No superscript
within a column indicates that there was no significant difference between distillers grain inclusion level.

Milk protein percentage was not different for cows fed diets containing 0-30% distillers
grains, and the form of the distillers grains did not alter composition (Table 2). However, milk
protein percentage decreased 0.13 percentage units when distillers grains was included at
concentrations greater than 30% of the diet compared to cows fed control diets. At the higher
inclusion levels, distillers grains most likely replaced all other sources of protein
supplementation. At these high levels of inclusion, lower intestinal protein digestibility, lower
lysine concentrations and an unbalanced amino acid profile may all contribute to lower milk
protein percentage. It should be noted that the lower milk protein percentages were most evident
in studies conducted in the 1980s and 1990s. Newer studies are not as consistent in showing
this effect. Lysine is very heat sensitive and can be negatively affected by processing and drying.
Improved processing and drying procedures in the fuel-ethanol plants built in recent years may
have improved amino acid quality of the product.

Other factors to consider


Dietary inclusion level of distillers grains is not the only factor that needs to be considered
when formulating lactating dairy cow diets with distillers grains. Other dietary factors that may
affect milk production and milk composition when distillers grains are added to the diet include
type of forage, ratio of forage to concentrate, high oil content of distillers grains and formulating
diets on an amino acid basis. In addition, the form of the distillers grains, wet or dried, may also
affect cow performance. The impact of these dietary factors on milk production and milk
composition was evaluated using the same 23 published reports as described previously. There
were 96 treatment comparisons included in this database.
Type of forage
To evaluate whether type of forage had an impact on animal performance, each diet was
identified by the ratio of corn silage to alfalfa. Twenty-three diets contained 100% corn silage,
38 diets contained 55-75% corn silage, 19 diets contained 45-54% corn silage and 16 diets
contained only alfalfa silage or hay (0% corn silage) as the forage source. In general, a
combination of forages is preferred to balance nutrient requirements and provide effective fiber
for normal rumen fermentation. However, the type of forages included in dairy cow diets is
mostly dictated by local supply. In some areas, alfalfa can be grown effectively, and therefore, it
may be the predominant forage included in dairy cow diets whereas in other regions of the
United States, corn silage predominates.
This review found that forage type had no impact on dry matter intake, milk production, or
milk fat composition. Forage type did, however, affect milk protein composition. Cows fed diets
containing 55-75% corn silage produced milk with the highest concentration of protein at 3.04%.
Cows fed 100% alfalfa/grass with 0% corn silage resulted in the lowest concentration of protein
at 2.72%. Cows fed 45-54% corn silage and 100% corn silage produced milk with intermediate
levels of protein at 2.98 and 2.82%, respectively. Cows fed diets with a blend of corn silage and
alfalfa produced milk with greater milk protein percentage suggesting that diets formulated with
one forage source are more likely to be insufficient in the amino acids needed to maximize milk
protein percentage.

03 - Use of DDGS in Dairy Diets

Forage to concentrate ratio


Forage to concentrate ratio is a second dietary factor that may affect lactation performance of
the dairy cow when distillers grains are included in the diet. To evaluate the effect of forage to
concentrate ratio, treatments were classified into one of three categories: diets containing less
than 50% forage, diets containing 50% forage and 50% concentrate and diets containing more
than 50% forage. Dry matter intake, milk production and milk protein percentage were not
affected by forage to concentrate ratio. The percentage of milk fat, however, was reduced by
0.36% in diets containing less than 50% forage.
This supports the hypothesis that lack of forage in the diet, resulting in insufficient effective
fiber, is a major contributing factor for causing reduced milk fat percentage rather than simply
the inclusion of distillers grains in the diet. Upon initial consideration, neutral detergent fiber
levels appear adequate because of the fiber provided by distillers grains. However, this fiber has
a small particle size and does not provide effective fiber needed for normal rumen function. A
recent experiment conducted at South Dakota State University tested this hypothesis directly
(Cyriac et al., 2005). As forage decreased in the diet from 55 to 34%, milk fat % decreased
linearly from 3.34 to 2.85% even though NDF % remained similar across diets. Therefore, when
formulating diets containing high levels of distillers grains, it is important to be certain that they
contain adequate levels of effective fiber from forage. The remaining fiber from distillers grains
will be quickly digested to volatile fatty acids (VFAs) in the rumen.
High oil content of distillers grains
The high oil content of distillers grains is a potential concern when it is included in dairy cow
diets. Corn oil in distillers grains is relatively high in linoleic acid, which is an unsaturated fatty
acid. High levels of vegetable oil can potentially cause incomplete biohydrogentation in the
rumen resulting in milk fat depression. This review of previously published studies did not find a
strong relationship between dietary distillers grain inclusion and milk fat depression. However,
it is possible that there could be interactions between the concentration of oil and the lack of
effective fiber which can result in milk fat depression.
Formulating diets on an amino acid basis
Finally, the effect of formulating diets on an amino acid basis was evaluated. This analysis
included experiments where rumen-protected lysine and methionine, or a source of lysine, such
as blood meal, was added to the diets. Lysine may be deficient in diets where corn feedstuffs are
the predominant ingredients in dairy cow diets. Milk protein percentage tended to increase when
diets included a source of lysine. Additional research is needed to determine if supplemental
lysine would allow for additional amounts of distillers grains to be included in dairy cow diets.
Feeding DDGS to Lactating Dairy Cows in Hot, Humid Sub-Tropical Climates
Most of the DDGS research involving dairy cattle has been conducted in temperate climates.
The U.S. Grains Council sponsored a feeding trial on a commercial dairy farm in central Taiwan
from September to November 2003 (Chen and Shurson, 2004). The objectives of this feeding
trail were to compare the feeding value of DDGS with corn, soybean meal and roasted soybeans
in lactating dairy cow rations and test the feasibility of DDGS in dairy rations in a hot and humid
sub-tropical environment.

03 - Use of DDGS in Dairy Diets

The trial was conducted on a commercial dairy farm located in Tainan County, Taiwan. The
location of the farm was about 20 km south of the Tropic of Cancer. The dairy herd consisted of
a total 600 cattle, including 290 milking cows. The main barn of this dairy was a typical freestall facility with an exercise lot for each pen. The barn was equipped with a sprinkler and
misting system for evaporative cooling during the hot season. A double 12 stall milking parlor
with automatic take-offs milking machines was operated by 4 milkers.
Fifty primparous Holstein cows were randomly assigned to the control and DDGS treatment
groups based on their days in milk (DIM), pre-treatment milk production and body condition
score (BCS). The average DIM of two groups was the same (149 56 days). The average milk
production of the control and DDGS group at grouping was 22.3 2.8 kg and 22.4 3.7 kg,
respectively. The average BCS of the control and DDGS group at grouping was 3.0 0.3 kg and
3.1 0.3 kg, respectively. The feeding trial consisted of a two-week adjustment period to allow
the cows to adapt to the pen, followed by an eight-week experimental period for data collection.
Cows were fed a total mixed ration (TMR) containing either 0% (control) or 10% DM from
DDGS. DDGS partially replaced some of the soybean meal, corn, steam-flaked corn and roasted
soybeans in the TMR ration. The rations were formulated using Cornell Net Carbohydrate and
Protein System (Barry, et al., 1994) to meet the requirement of metabolizable protein (MP),
metabolizable energy (ME), calcium and phosphorus.
The average daily dry matter intake (DMI) of the control and DDGS groups were 17.8 1.2
and 17.6 1.0 kg, respectively. The addition of DDGS did not influence the DMI of the
experimental animals and there was no pen effect on DMI (Table 4), but the actual DMI was
lower than the DMI prediction by Cornell Net Carbohydrate and Protein System (version
4.26;Barry, et al., 1994). This DMI discrepancy might result from the heat-stressed conditions
experienced during the trial. Although the trial was conducted from September to November, the
cows were still under a heat-stressed environment (temperature/humidity index greater than
72).(Figure 1).
90.0
85.0
80.0

THI

75.0
70.0
65.0
60.0
55.0
11/18/2003 14:59

11/15/2003 17:59

11/12/2003 20:59

11/7/2003 2:59

11/9/2003 23:59

11/4/2003 5:59

11/1/2003 8:59

10/29/2003 11:59

10/26/2003 14:59

10/23/2003 17:59

10/20/2003 20:59

10/15/2003 2:59

10/17/2003 23:59

10/9/2003 8:59

10/12/2003 5:59

10/6/2003 11:59

10/3/2003 14:59

9/30/2003 17:59

9/27/2003 20:59

9/24/2003 23:59

9/22/2003 2:59

9/19/2003 5:59

9/16/2003 8:59

Date

Time

50.0

Figure 1. Temperature-Humidity-Index (THI) during the


commercial feeding trial in Taiwan.

03 - Use of DDGS in Dairy Diets

The average milk production of all cows in the control and DDGS groups on each dairy herd
improvement (DHI) day is shown in Figure 3. Cows in the DDGS group tended to have a higher
average milk production than cows in the control group. There was no difference in milk
production before ration treatment (9/6/2003 and 9/21/2003 DHI). After the feeding the
experimental rations, the cows in the DDGS group produced more milk than the cows in the
control group on each DHI test day. The increase in milk production of cows fed the DDGS
ration may have been due to the high feeding value of DDGS or lower days in milk (DIM) of the
DDGS group. It is unlikely that this difference was due to a pen effect because there was no
difference in milk production between the two groups during the adapting (pre-treatment) period.
The removal of mastitis cows from the trial resulted in a difference of DIM between two
groups, but this difference was small (6 days). Therefore, DDGS may have a real advantage for
supporting higher milk production of mid-lactating cows under heat-stressed conditions. Both
groups showed a significant drop in milk production in the last DHI test. The THI increased
during this period of time (Figure 2) and feeding poor quality corn silage obtained from a new
silage bag were two possible reasons to explain this phenomenon.
25

Milk Production (KG)

20

15
Control
DDGS
10

0
2003/9/6

2003/9/21

2003/10/6

2003/10/21

2003/11/5

2003/11/20

DHI Date

Figure 2. Average Milk Production of Cows fed the Control and DDGS TMR.
The DHI data from the animals that completed the trial were used for statistical analysis
(Table 4). Cows in the DDGS group produced significantly more (0.9 kg/d/h) milk than the cows
in the control group. The ration containing DDGS provided more fat to the DDGS group and
could be a primary factor for supporting higher milk production. However, DDGS is highly
digestible and may contain some unidentified compounds that enhance rumen function and
animal performance. Although milk fat percentage was not different between treatments or pens,
cows in the DDGS group tended to produce more milk fat per day than cows in the control
group. The higher milk fat production can be attributed to the higher level of milk production of

03 - Use of DDGS in Dairy Diets

cows in the DDGS group. Although the addition of 10% DDGS to the ration significantly
decreased the milk protein percentage, the amount of milk protein produced per day was not
affected. One of the concerns regarding the use of DDGS in the lactating dairy cow rations is its
high fat content, which may interfere with ruminal fermentation and may decrease microbial
protein production and milk protein. However, the higher level of milk production of cows in the
DDGS group compensated for the negative effects of feeding DDGS on milk protein percentage.
Both dietary treatment and pen effects were observed for percentage of lactose in milk, but it is
not clear why these responses were observed. The body condition score was not significantly
different between dietary treatments during the trial.
Table 4. Effects of Feeding TMR1 with and without 10% DDGS on
Milk Production, Milk Composition and Body Condition Score of
Mid-Lactating Cows under Heat-stressed Conditions.
Response
Treatment (T)
Pen (P)
SE
P- value
variable
Control DDGS
1
2
T
P
TP
DMI, kg/d2
17.8
17.6
17.8
17.6 0.20 0.32
0.29 0.012
Milk, kg/d
19.5
20.4
19.8
20.1 0.44 0.04
0.46 0.003
Fat, %
4.51
4.45
4.43
4.53 0.13 0.61
0.41 0.69
Fat, kg/d
0.86
0.91
0.87
0.91 0.03 0.10
0.22 0.07
Protein, %
3.45
3.32
3.41
3.37 0.04 0.001 0.17 0.73
Protein, kg/d
0.66
0.68
0.67
0.67 0.02 0.40
0.97 0.02
Lactose, %
4.85
4.90
4.92
4.83 0.03 0.07
0.004 0.84
Total Solids, % 13.5
13.4
13.5
13.4 0.16 0.36
0.77 0.63
3
MUN, mg/dL
11.2
11.8
12.3
12.8 0.50 0.23
0.80 0.04
4
4
SCC, 10 /ml
26.9
35.4
35.9
26.4 13.8 0.54
0.49 0.76
BCS5
2.96
3.01
0.21
1

TMR = total mixed ration


DMI = dry matter intake
3
MUN = milk urea nitrogen

SCC = somatic cell count


BCS = body condition score

Feeding DDGS to Growing Dairy Heifers


Although DDGS is considered to be an excellent energy and protein source for ruminants,
there is very little information on feeding DDGS to growing dairy heifers. Kalscheur and Garcia
(2004) suggested that data from experiments on feeding DDGS to growing beef cattle could be
extrapolated, with caution, to expected responses for growing dairy cattle. When wet or dried
distillers grains were fed to growing beef calves, there were no differences in growth rate or
protein accretion (Kalscheur and Garcia, 2004). However, when dried rolled corn was replaced
with wet distillers grains or DDGS, to provide 40% of dry matter intake, growth rate and feed
conversion were improved (Kalscheur and Garcia, 2004). Growing cattle fed wet distillers
grains generally have higher feed conversion than cattle fed DDGS. At high DDGS feeding
levels, variable amounts of heat-damaged protein among DDGS sources are less of a concern for
growing cattle because they consume protein in excess of their requirements (Kalscheur and
Garcia, 2004). Therefore, DDGS can be added to growing heifer rations at levels up to 40% of
dry matter intake to achieve excellent growth rate and feed conversion.

03 - Use of DDGS in Dairy Diets

Summary
DDGS is a good source of protein, fat, phosphorus and energy for lactating dairy cows.
Distillers grains can be included in dairy cow diets up to 20% of the ration without decreasing
dry matter intake, milk production and milk fat and protein percentage. Inclusion of DDGS 2030% also supports milk production equal to or greater than diets with no DDGS; however, milk
production from cows fed diets containing wet distillers grains decreases when wet distillers
grains are included at more than 20% of the diet. Milk fat percentage varies, but was not
significantly changed by the inclusion of distillers grains in the diet. Milk protein percentage
decreased at the highest distillers grains dietary inclusion levels. More research on using
distillers grains from newer ethanol plants is needed to determine if improved quality
corresponds to improved performance. Consequently, distillers grains from todays ethanol
plants may not affect milk protein percentage as did distillers grains from the 1980s and
1990s. In addition, studies investigating rumen function are needed to determine the impact of
distillers grains on milk fat concentration.
Distillers grains can replace more expensive sources of protein, energy and minerals in dairy
cow diets. However, when balancing diets containing DDGS, nutritionists must follow
acceptable nutritional guidelines to prevent an imbalance of nutrients. DDGS can be effectively
used in a total mixed ration by mid-lactating dairy cows under heat-stressed climatic conditions,
and is a potential high quality co-product for the dairy industry in sub-tropical and tropical
regions of the world. Although there has been limited research to evaluate feeding DDGS to
growing dairy heifers, DDGS has been added to growing beef cattle rations at levels up to 40%
of dry matter intake to achieve excellent growth rate and feed conversion.

References
Barry, M. C., D. G. Fox, T. P. Tylutki, A. N. Pell, J. D. O'Connor, C. J. Sniffen, and W. Chalupa. 1994. The
Cornell net carbohydrate and protein system for evaluating cattle diets. 3rd ed. Cornell University, Ithaca, NY.
Chen, Yuan-Kuo and J. Shurson. 2004. Evaluation of distillers dried grains with solubles for lactating cows in
Taiwan. http://www.ddgs.umn.edu/international-translations/Taiwanese%20(Yuan-Kuo%20Chen%202004).pdf
Cyriac, J., M. M. Abdelqader, K. F. Kalscheur, A. R. Hippen, and D. J. Schingoethe. 2005. Effect of replacing
forage fiber with non-forage fiber in lactating dairy cow diets. 88(Suppl. 1):252
Kalscheur, K. F. Impact of feeding distillers grains on milk fat, protein, and yield. Distillers Grains Technology
Council. 9th Annual Symposium. Louisville, KY. May 18, 2005.
Kalscheur, K.F. and A.D. Garcia. 2004. Use of by-products in growing dairy heifer diets. Extension Extra, South
Dakota State University. ExEx 4030, 3 pp.
National Research Council. 2001. Nutrient Requirements of Dairy Cattle. 7th Rev. Ed. National Academy of Sci.,
Washington, DC.
Powers, W.J., H.H. Van Horn, B. Harris, Jr., and C.J. Wilcox. 1995. Effects of variable sources of distillers grains
plus solubles on milk yield and composition. J. Dairy Sci. 78:388-396.
Schingoethe, D.J. 2004. Corn Co products for Cattle. Proceedings from 40th Eastern Nutrition Conference, May
11-12, Ottawa, ON, Canada. pp 30-47.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in understanding generallyaccepted feeding levels. However, all rations for specific herds should be formulated by a qualified nutritionist. The USGC has
no control over the nutritional content of any specific product which may be selected for feeding. Potential buyers should consult
an appropriate nutritionist for specific recommendations. USGC makes no warranties that these recommendations are suitable
for any particular herd or for any particular animal. The USGC disclaims any liability for itself or its members for any problems
encountered in the use of these recommendations. By reviewing this material, buyers agree to these limitations and waive any
claims against USGC for liability arising out of this material.

03 - Use of DDGS in Dairy Diets

Use of DDGS in
P o u lt ry D i e ts

User Handbook

Use of U.S. DDGS in Poultry Diets


Historical Use of DDGS in Poultry Diets
Use of distillers dried grains with solubles in poultry diets has historically been about a 5%
inclusion rate due to limitations such as supply and pricing (Waldroup et al., 1981) and
variability in nutrient content and digestibility (Noll et al., 2001). In past decades, DDGS was
used in poultry diets primarily as a source of unidentified factors that promoted growth and egg
hatchability. Distillers dried solubles (DDS) or DDGS were added to poultry diets at levels of
less than 10% of the diet. Couch et al. (1957) showed that adding 5% DDGS improved turkey
growth rates by 17-32%. Day et al. (1972) reported improvements in broiler body weights when
2.5% DDS or 5% DDGS were added to the diet in one of three trials. Couch et al. (1957) also
observed improvements in turkey breeder hatchability during the second half of lay with dietary
inclusion of dried alfalfa meal, condensed fish solubles and DDS. Manley et al. (1978) observed
an improvement in egg production when 3% DDGS was added to diets of hens experiencing a
low rate of egg production in late lay. Some researchers have hypothesized that responses to
unidentified factors may partially be attributed to improvements in feed palatability. Alenier and
Combs (1981) noted that chicken layer hens preferred diets containing 10% DDGS or 15% DDS
over a corn-soybean meal diet without DDGS. However, Cantor and Johnson (1983) were unable
to document a feed preference effect for diets containing DDGS compared to corn-soybean meal
diets. Most of the improved responses from these early studies have been attributed to DDS and
DDGS providing vitamins, and perhaps trace minerals that were lacking in poultry diets. Now
that the requirements for essential nutrients have been established and the availability of a variety
of commercial nutrient supplements, these responses are less likely to occur when distillers coproducts are added to poultry diets.

Nutrient Value of DDGS for Poultry


DDGS can supply a significant amount of energy, amino acids and phosphorus to poultry diets.
However, Spiehs et al. (2002) showed that the nutrient content of DDGS can vary among and
within ethanol plants, but nutrient levels are generally higher than those published by the
National Research Council (NRC, 1994). The only nutrient with a coefficient of variation less
than 5% among ethanol plants was dry matter, whereas crude protein, fat, fiber and some amino
acids had coefficients of variations less than 10%. The first two limiting amino acids in poultry
diets are lysine and methionine and, unfortunately, the coefficients of variation for these amino
acids are high (17.3 and 13.6, respectively). Furthermore, Spiehs et al. (2002) showed that the
coefficient of variation for phosphorus was also high (11.7%). In a subsequent study, Noll et al.
(2003) obtained 22 DDGS samples from four different ethanol plants and observed lower
average levels of protein, ash, fiber, methionine, lysine and phosphorus compared to levels
observed by Spiehs et al. (2002). However, this may have been due to fewer numbers of sources
and samples analyzed compared to the 118 samples from 10 ethanol plants in the study reported
by Spiehs et al. (2002). Noll et al. (2003) showed that the coefficients of variation were lower
within plants than among plants.

04 - Use of DDGS in Poultry Diets

Energy
In recent studies, researchers have used metabolizable energy values of 2,865 kcal apparent
metabolizable energy (AME)/kg, 2,905 kcal true metabolizable energy (TME)/kg and 2,805 kcal
TME/kg for DDGS in feeding trials with turkeys (Noll et al., 2004), broilers (Lumpkins et al.,
2004) and layers (Lumpkins et al., 2005), respectively, without negative effects on feed
conversion and with dietary inclusion levels of 10%. Batal and Dale (2004) obtained an average
TME value for DDGS of 2,831 kcal/kg with roosters, whereas Roberson (K. D. Roberson,
Michigan State University, personal communication) determined AME values of 2,760 and
2,750 kcal/kg for DDGS in turkey poults and laying hens, respectively. The experimentally
derived AME value of 2,750 kcal/kg is considered to be a more adequate estimate of the energy
value of DDGS in market turkey toms compared to the value of 2,480 kcal/kg from NRC (1994),
or an experimentally derived TME value of 2,980 kcal/kg (Noll et al., 2005). Conservatively, a
value of 2,755 kcal ME/kg can be used to avoid overestimating the energy content of DDGS.
Regardless, it is important to note that these recent estimates of energy are substantially higher
than the value of 2,480 kcal ME/kg reported in NRC (1994).
Amino acids
Recent research results have also shown that the amino acid content and digestibility of light
colored DDGS sources is higher than values reported in NRC (1994). For example, lysine
digestibility of DDGS can be as high as 83% compared to 65% the value reported in the
poultry NRC (1994) reported by Ergul et al. (2003). Cromwell et al. (1993) first demonstrated
that lightness (L*) and yellowness (b*) of color were highly correlated with chick weight gain
(0.74 and 0.72, respectively) and feed conversion (0.69 and 0.74, respectively). Ergul et al.
(2003) also confirmed that lightness and yellowness of color of DDGS appear to be reasonable
predictors of digestible lysine content among light colored DDGS sources for poultry (Figure 1).

04 - Use of DDGS in Poultry Diets

Minerals
DDGS is also high in phosphorus (0.73%; Noll et al., 2003). Unlike phosphorus availability in
corn, the availability of phosphorus in DDGS is higher for poultry. Lumpkins and Batal (2005)
obtained phosphorus availability estimates of 54 and 68%, whereas Martinez et al. (2004)
obtained bioavailability estimates for phosphorous of 69, 75, 82, and 102% for different DDGS
samples. The sodium content of DDGS can range 0.01-0.48% averaging 0.11%. Therefore,
dietary adjustments for sodium content may be necessary if the source of DDGS being used
contains high levels of sodium, in order to avoid potential problems with wet litter and dirty
eggs.
Xanthophyll
DDGS can contain as much as 40 ppm of xanthophyll. The xanthophyll content of DDGS has
been shown in commercial field and university research trials to significantly increase egg yolk
color when fed to laying hens (Shurson et al., 2003 and Roberson et al., 2005, respectively), and
increase skin color of broilers when included at levels of 10% of the diet.

Feeding U.S. DDGS to Chicken Layers


There has been a limited amount of research conducted on the use of high quality corn DDGS
in layer diets. Matterson et al. (1966) showed that DDGS could be added to laying hen diets at
levels of 10-20%, which accounted for about 30% of the total dietary protein, without synthetic
lysine supplementation, with no effect on egg production. Harms et al. (1969) reported that
adding 10% DDGS to a layer diet to replace a portion of the dietary protein did not affect egg
production or egg weight. Jensen et al. (1974) reported that feeding diets containing DDGS
resulted in an improvement in interior egg quality (Haugh units), but it was not a consistent
response.
More recently, Lumpkins et al. (2005) fed Hy-line W-36 laying hens high energy (2,871 kcal
TMEn/kg) and low energy ((2,805 kcal TMEn/kg) diets, with and without 15% DDGS from 22 to
42 weeks of age. The DDGS used in this study had color values of L* = 58.52, a* = 6.38, and b*
= 20.48. There were no significant differences in egg production for layers fed the 0 and 15%
DDGS high energy diets during the entire 22-week experiment. However, adding 15% DDGS to
the low energy diet slightly reduced egg production from 26 to 34 weeks of age, but there was no
difference after 34-weeks of age. There were no differences in egg weights, specific gravity,
shell breaking strength, feed conversion, body weight or mortality between the four dietary
treatments throughout the entire experiment. There was no difference in Haugh units between
dietary treatments from 25 to 31 weeks of age. At 43 weeks of age, layers fed the low energy,
15% DDGS diet had lower Haugh units compared to hens fed the high energy, 15% DDGS diet.
Furthermore, feeding the 15% DDGS diets had no appreciable effect on egg yolk color. Based
upon these results, the researchers concluded that DDGS is a very acceptable feed ingredient in
layer diets and the maximal dietary inclusion level of DDGS should be 10-12% in high energy
commercial diets, but lower dietary inclusion rates may be necessary in lower energy diets.
Roberson et al. (2005) conducted two experiments where diets containing 0, 5, 10, or 15%
DDGS were fed to laying hens to determine if egg production parameters or yolk color would be
affected. In the first experiment, a source of light colored DDGS was added to diets fed from 48
to 56 weeks of age and then a brown colored DDGS source was added to diets from 58 to 67
weeks of age. Egg production measurements were not different at most ages. However, as

04 - Use of DDGS in Poultry Diets

dietary level of DDGS increased, there was a linear decrease in egg production (52-53 weeks of
age), egg weight (63 weeks of age), egg mass (51 and 53 weeks of age) and specific gravity (51
weeks of age). Egg yolk color increased linearly as dietary level of DDGS increased throughout
the experiment. In experiment 2, egg yolk redness (a*) increased linearly as dietary DDGS level
increased. These results showed that egg yolk color becomes more red within one month of
feeding diets containing 10% DDGS or more of a light colored DDGS, and that egg yolk color
becomes more red by two months of feeding diets containing 5% DDGS. The researchers
concluded that feeding layer diets containing up to 15% DDGS did not affect egg production but
the variable results in experiment 1 suggest that a level less than 15% DDGS should be used.
Shurson et al. (2003) conducted a commercial layer feeding trial in Jalisco, Mexico, to evaluate
egg production, egg quality and egg yolk color under practical feeding conditions in Mexico.
There were no differences in dry matter, crude protein, crude fat, ash, calcium and phosphorus
content between the control and 10% DDGS diets. However, the addition of 10% DDGS
provided significantly more xanthophyll to the DDGS diet (11.8 parts per million, or ppm) than
the control diet (10.2 ppm) and the difference in xanthophyll content of the experimental diets
tended to be the greatest during the first four weeks of the trial (Figure 1). Xanthophyll content
of the DDGS diets appeared to decline during the trial which reflects the expected loss of
xanthophyll content of DDGS during the 16 week storage period (4 weeks prior to starting the
trial plus the 12 week trial).
There were no differences in average hen body weight during the first two weeks of the trial,
but hens fed the DDGS diet were heavier than hens fed the control diet for weeks 3 through 12
(Figure 2). This suggests that the energy content of the DDGS diet was higher than the control
diet because average weekly feed consumption was not different between hens fed the control
and DDGS diets (Figure 3).
As shown in Figure 4, average percentage of production was not different between layers fed
the control and DDGS diets during weeks 1, 2, 3, 4, and 9. However, hens fed the DDGS diet
had a higher percentage of production during weeks 5, 6, 7, 8, 10, 11 and 12. These results
suggest that feeding layer diets containing 10% DDGS may result in an increase in egg
production compared to feeding a common control diet used in Jalisco. The decrease in
percentage of production that occurred during week 9 was a result of a subclinical outbreak of
infectious bronchitis, along with feeding mycotoxin contaminated sorghum during this time
period. Layers fed the DDGS diet appeared to return to a high percentage of production more
quickly than hens fed the control diet.

04 - Use of DDGS in Poultry Diets

Figure 1: Xanthophyll Content of Control and DDGS Diets


During the 12-Week Layer Trial.
14.00
12.00
10.00
8.00

DDGS
Control

6.00
4.00
2.00
0.00
Week 1

Week 2

Week 4

Week 5

Week 6

Week 7

Week 8

Week 9

Week 10

Figure 2: Average Hen Body Weight (kg) During the 12-week DDGS Trial.
2.000
1.800
1.600
1.400
1.200

Control
DDGS

1.000
0.800
0.600
0.400
0.200
0.000
Week Week Week Week Week Week Week Week Week Week Week Week
1
2
3
4
5
6
7
8
9
10
11
12

04 - Use of DDGS in Poultry Diets

Figure 3: Average Weekly Feed Consumption (kg) per


Replicate of Hens Fed Control and DDGS Diets.
3500

3000

2500

2000

Control
DDGS

1500

1000

500

0
Week 1

Week 2

Week 3

Week 4

Week 5

Week 6

Week 7

Week 8

Week 9 Week 10 Week 11 Week 12

Figure 4: Average Percentage of Production by Week for


Layers Fed Control and DDGS Diets
90.0
80.0
70.0
60.0
50.0

Control
DDGS

40.0
30.0
20.0
10.0
0.0
Week 1 Week 2 Week 3 Week 4 Week 5 Week 6 Week 7 Week 8 Week 9 Week
10

Week
11

Week
12

As shown in Table 2, there were no overall differences in % mortality and % prolapsed hens
between layers fed the control and DDGS diets. During the 12-week feeding period, the
percentage production of first class eggs tended to be higher for hens fed the DDGS diet

04 - Use of DDGS in Poultry Diets

compared to hens fed the control diet. Hens fed the DDGS diet produced an average of 3.7 more
eggs during the 12-week feeding period compared to hens fed the control diet. Furthermore, hens
fed the DDGS diets tended to produce heavier eggs than hens fed the control diet. However, the
percentage of first class eggs of the total eggs produced was lower for layers fed the DDGS diet.
The lower percentage of first class eggs of total eggs produced for hens fed the DDGS diet was
due to the higher percentage of broken eggs (1.22 vs. 0.75%), no shell eggs (0.02 vs. 0.01%),
dirty eggs (2.18 vs. 1.37%), and double yolk eggs (0.12 vs. 0.08%). Although there were
significant dietary treatment differences for no shell eggs and double yolk eggs, the percentage
of the total eggs produced was extremely low and is not of great importance. The higher number
of broken eggs for hens fed the DDGS diets is likely due to the production of slightly larger eggs
that often did not fit through the opening in cages where the birds were housed. It is unclear why
feeding the DDGS diet in this experiment resulted in an increase in the percentage of dirty eggs
compared to eggs from hens fed the control diet.
Table 1. Effect of Feeding a Layer Diet Containing DDGS on
Hen Mortality and Prolapses and Egg Production and Quality
Response variable
Average number hens/wk/pyramid
% hen mortality
% prolapsed hens
% production of first class eggs
Total number of eggs produced
Average % production
Egg weight produced/pyramid, kg
Average egg weight produced/hen/day, kg
Total number of first class eggs
% first class eggs
Total number of broken eggs
% broken eggs
Total number of no shell eggs
% no shell eggs
Total number of dirty eggs
% dirty eggs
Total number of double yolk eggs
% double yolk eggs
Average egg Haugh units
Average egg specific gravity
Average yolk color

Control
3,948
1.99
0.49
66.2
224,533
68.7
14,576
0.308
219,565
97.8
1,683
0.75
26.3
0.01
3,073
1.37
185
0.08
92.6
7.41
10.63

DDGS
3,828
1.80
0.52
68.9
229,294
72.4
14,659
0.320
221,156
96.5
2,806
1.22
48.4
0.02
4,999
2.18
284
0.12
93.2
7.34
10.81

SE
51.2
0.13
0.07
1.09
2324
1.01
158.2
0.005
2338
0.20
116
0.05
4.45
0.002
341
0.15
16.9
0.008
0.46
0.06
0.02

P value
0.12
0.30
0.76
0.10
0.17
0.02
0.72
0.11
0.64
0.003
0.0001
0.0001
0.003
0.006
0.001
0.002
0.001
0.003
0.45
0.51
0.02

There were no overall differences in egg albumin quality (measured as Haugh units) and egg
shell quality (measured as specific gravity) between dietary treatment groups (Table 1).
However, hens fed the DDGS diet produced heavier eggs during week 6, week 10 and week 11
than hens fed the control diet (Figure 6). Furthermore, hens fed the DDGS diet produced eggs
with a darker colored egg yolk, which is very desirable to the Mexican consumer, compared to
feeding the control diet (Table 1). However, as shown in Figure 7, these differences in egg yolk

04 - Use of DDGS in Poultry Diets

color were greater during the early weeks of the production cycle compared to the later portion
of the feeding trial and this pattern corresponds with the declining level of xanthophyll content of
DDGS shown in Figure 1.
Figure 6. Average Egg Weight (g) by Week for Layers Fed Control and DDGS Diets.
72.00
70.00
68.00
66.00

Control
DDGS

64.00
62.00
60.00

W
ee
k
W 1
ee
k
W 2
ee
k
W 3
ee
k
W 4
ee
k
W 5
ee
k
W 6
ee
k
W 7
ee
k
W 8
ee
W k9
ee
k
W 10
ee
k
11

58.00

Figure 7. Differences in Yolk Color (Roche Units) in Eggs


Produced by Layers Fed Control and DDGS Diets.
11.40

11.20

11.00

10.80

Control
DDGS

10.60

10.40

10.20

10.00

9.80
Week 1 Week 2 Week 3 Week 4 Week 5 Week 6 Week 7 Week 8 Week 9 Week 10 Week 11

Results from this study show that adding 10% DDGS to practical chicken layer diets used in
Jalisco can provide a significant improvement in % production and egg yolk color compared to
typical control diets routinely used. However, because eggs produced by layers tended to be
slightly larger than eggs produced by hens fed the control diets, the percentage of broken eggs

04 - Use of DDGS in Poultry Diets

may increase depending upon the types of cages used in commercial layer facilities. There were
no differences between hens fed the DDGS diets compared to the control diets for mortality,
prolapses, egg albumin quality and egg shell quality. However, feeding diets containing 10%
DDGS appeared to increase the number of dirty eggs. Potential reasons for an increase in the
number of dirty eggs for hens fed the DDGS diet are unknown but may have been due to slightly
different management conditions among the test barns used in this study.

Feeding DDGS to Broilers


Researchers have observed positive results when DDGS is added to broiler diets. In an early
study by Day et al. (1972), weight gain of broilers was increased when low levels of DDGS (2.5
and 5%) were added to the diet compared to broilers fed the control diet. Waldroup et al. (1981)
concluded that DDGS can be added to broiler diets at levels up to 25% to achieve good
performance if dietary energy level is held constant.
In more recent studies, Lumpkins et al. (2004) conducted two experiments to evaluate dietary
energy and protein density and DDGS inclusion rate in broiler diets. In the first experiment, two
dietary nutrient densities (high = 22% protein, 3,050 kcal MEn/kg and low = 20% protein, 3,000
kcal MEn/kg) contained either 0 or 15% DDGS. Chicks were fed experimental diets from 0 to 18
days of age. Weight gain and feed conversion were the highest for chicks fed the high density
diet compared to the low density diet, but performance was not different between chicks fed the
0 or 15% DDGS diets within diet nutrient density level. In the second experiment, they fed
chicks isocaloric and isonitrogenous diets containing 0, 6, 12, or 18% DDGS for a 42-day
feeding period. There were no differences in weight gain, feed conversion or carcass yield
throughout the experiment as dietary DDGS level increased, except for a depression in gain and
feed conversion of chicks fed the level of 18% DDGS in the starter period. These researchers
concluded that DDGS from modern ethanol plants is an acceptable ingredient in broiler diets and
recommended a 6% dietary inclusion rate in the starter period and 12-15% DDGS in grower and
finisher phases.
The U.S. Grains Council has been involved in several broiler trials in Taiwan. In a
comprehensive study conducted in 2005 (Jin-Jenn Lu and Yuan-Kuo Chen, 2005), researchers
wanted to determine the effect of different dietary inclusion rates of DDGS on growth
performance, skin color and carcass quality of domestic colored chickens. Results from this
study showed that adding 20% corn DDGS to domestic colored chicken diets had no negative
effects on weight gain, feed efficiency, meat quality, protein metabolism and fat metabolism.
The xanthophylls in DDGS can be effectively absorbed and deposited in the abdominal fat pad
and skin of broilers. DDGS can be stored effectively for up to 12 weeks without losing
xanthophyll concentration. Although the xanthophylls in DDGS can not completely replace
artificial pigments to meet the color requirement for the Taiwan market, 20% DDGS plus onehalf of the amount of artificial pigments can achieve the desired carcass quality and color of the
abdominal fat pad and skin. When the cost of diets without supplementation of artificial
pigments are the same between treatments as in this trial, adding 20% DDGS can decrease the
supplementation of artificial pigments by 50% thereby saving a significant amount in feed cost.
These results show that DDGS is a good alternative feedstuff for efficient domestic colored
chicken production and its use in diets for domestic colored chickens is encouraged.
Additional broiler trials were conducted in the commercial feed industry in Taiwan in 2004.
Growth performance results from two trials are shown in Tables 2 and 3. These results suggest

04 - Use of DDGS in Poultry Diets

that excellent growth performance can be obtained when adding 10% DDGS to starter, grower
and finisher broiler diets, equal to typical commercial broiler diets used in the Taiwan broiler
industry. These results are consistent with previously published results from the University of
Georgia (Lumpkins et al., 2003) showing that high quality DDGS can be added up at levels to
12% of starter, grower and finisher broiler diets without having any negative effects on growth
performance.
Table 2: Growth Performance of Broiler Chickens
Fed Diets Containing 0 or 10% DDGS in Taiwan (Trial 1).
Control 10% DDGS
Standard
P value
Deviation
160
160
Number of birds, d 0
152
157
Number of birds, d 38
95.0
98.1
Livability, %
Avg. body weight, g
Day 0
Day 14
Day 29
Day 38
Avg. feed intake, g
Day 0-14
Day 14-29
Day 29-38
Day 0-38
Avg. gain, g/d
Day 0-14
Day 14-29
Day 29-38
Day 0-38
Feed/Gain
Day 0-14
Day 14-29
Day 29-38
Day 0-38

42
434
1336
2028

42
441
1346
2001

0.76
12.82
51.50
46.24

0.34
0.22
0.69
0.21

466
1368
1417
3251

471
1401
1432
3305

20.42
82.31
59.51
131.09

0.62
0.39
0.58
0.39

392
902
1521
1986

399
904
1487
1959

12.74
45.74
53.78
46.19

0.24
0.91
0.18
0.20

1.19
1.52
0.93
1.60

1.18
1.55
0.96
1.65

0.03
0.05
0.07
0.06

0.57
0.16
0.33
0.08

04 - Use of DDGS in Poultry Diets

10

Table 3: Growth Performance of Broiler Chickens Fed Diets


Containing 0 or 10% DDGS on a Commercial Farm in Taiwan (Trial 2).
Control
10% DDGS
30,000
30,000
Initial number of birds
28,950
28,584
Final number of birds
% Livability
96.5
95.3
Avg. body weight/bird, kg
Day 32
Day 36
Avg. feed intake/bird, kg
Day 0-36
Feed/Gain
Day 0-36
Avg. feed cost NT/kg
Cost per kg gain, NT

1.76
1.96

1.72
1.90

3.51

3.21

1.79
10.05
17.99

1.69
9.87
16.68

Turkeys
Noll (2004) summarized results from three trials where diets containing up to 12% DDGS were
fed to market toms during the grower-finisher period and found no difference in body weight
gain and feed conversion compared to the control corn-soybean meal-meat meal diets. Roberson
(2003) conducted two experiments using Large White female turkeys to evaluate the effects of
increasing dietary DDGS level on growth performance. In the first experiment, corn-soybean
meal diets containing 0, 9, 18 or 27% DDGS were fed to growing turkeys from 56 to 105 days of
age. Body weight linearly decreased with increasing level of DDGS in the diet at 105 days of
age. However, feed conversion improved from 77 to 105 days of age as dietary DDGS level
increased. Roberson (2003) noted that the incidence of pendulous crops increased for birds fed
diets with high levels of DDGS. In the second experiment, diets containing 0, 7 or 10% DDGS
were fed in the grower period, with half of the birds fed the 10% DDGS in the grower period fed
7% DDGS in the finisher period. There were no differences among dietary treatments for body
weight gain or feed conversion in this experiment. He concluded that DDGS can be effectively
included at 10% of growing-finishing diets for turkey hens if the proper nutrient values for
DDGS are used.

Ducks
The U.S. Grains Council sponsored a recent study conducted at the I-lan Branch of the
Livestock Research Institute in Taiwan, where researchers evaluated the effects of feeding diets
containing dried distillers grains with solubles on the production performance and egg quality of
brown Tsaiya duck layers (Huang et al., 2006). ducks from 14 weeks of age up to 50 weeks of
age were randomly assigned to one of four dietary treatments containing 0, 6, 12 or 18% DDGS.
Diets were isocaloric and isonitrogenous diet and contained 2,750 kcal/kg ME and 19% crude
protein (CP). Results from this study suggested that adding DDGS at levels up to 18% of the diet
for laying ducks had no significant effect on feed intake, feed conversion or quality of the egg
shell. When laying ducks were fed the 18% DDGS diet egg production rate increased in the cold
season. Egg weight tended to be higher when 12% or 18% of DDGS was included in the diets.
04 - Use of DDGS in Poultry Diets

11

Yolk color was linearly improved with increasing amounts of DDGS in the laying duck diets.
The xanthophylls in DDGS can be well utilized by the laying ducks. When DDGS was used in
duck laying diets, fat percentage of yolk and linoleic acid content of yolk was increased. DDGS
can be efficiently used in the diets of duck layers to improve the yolk characteristics without
influencing the productive performance.

Summary
Current recommended maximum dietary inclusion levels for DDGS are 10% for meat birds
and 15% for chicken layers. Higher levels of DDGS can be used successfully with appropriate
diet formulation adjustments for energy and amino acids (Noll et al., 2004; Waldroup et al.,
1981). When formulating diets containing DDGS, digestible amino acid values should be used
especially for lysine, methionine, cystine and threonine. Diets should also be formulated by
setting minimum acceptable levels for tryptophan and arginine due to the second limiting nature
of these amino acids in DDGS protein.

Literature Cited
Abe, C., N. J. Nagle, C. Parsons, J.Brannon, and S. L. Noll, 2004. High protein corn distiller dried grains as a feed
ingredient. Poultry Sci. 83 (Suppl. 1):264.
Alenier, J.C. and G.F. Combs, Jr. 1981. Effects on feed palatability of ingredients believed to contain unidentified
growth factors for poultry. Poultry Sci. 60:215-224.
Batal, A. B. and N. M. Dale, 2004. True metabolizable energy and amino acid digestibility of distillers dried
grains with solubles. Poultry Sci. 83 (Suppl 1):317.
Cantor, A.H. and T.H. Johnson. 1983. Effects of unidentified growth factor sources on feed preference of chicks.
Poultry Sci. 62:1281-1286.
Combs, G.F. and E.H. Bossard. 1969. Further studies on available amino acid content of corn distillers dried
grains with solubles. In Proceedings Distillers Feed Research Council Conference. Pp. 53-58.
Couch, J.R., A.A. Kurnick, R.L. Svacha, and B.L. Reid. 1957. Corn distillers dried solubles in turkey feeds
summary and new developments. In Proceedings Distillers Feed Research Council Conference. Pp. 71-78.
Cromwell, G.L., K.L. Herkleman, and T.S. Stahly. 1993. Physical, chemical, and nutritional characteristics of
distillers dried grains with solubles for chicks and pigs. J. Anim. Sci. 71:679-686.
Day, E.J., B.C. Dilworth, and J. McNaughton. 1972. Unidentified growth factor sources in poultry diets. In
Proceedings Distillers Feed Research Council Conference. Pp. 40-45.
Ergul, T., C. Martinez Amezcus, C. M. Parsons, B. Walters, J. Brannon and S. L. Noll, 2003. Amino acid
digestibility in corn distillers dried grains with solubles. Poultry Sci. 82 (Suppl. 1): 70.
Harms, R.H., R.S. Moreno, and B.L. Damron. 1969. Evaluation of distillers dried grains with solubles in diets of
laying hens. Poultry Sci. 48:1652-1655.
Huang, J.F., M.Y. Chen, H.F. Lee, S.H. Wang, Y.H. Hu, and Y.K. Chen. 2006. Effects of Corn Distillers Dried
Grains with Soluble on the Productive Performance and Egg Quality of Brown Tsaiya Duck Layers. Personal
communication with Y.K Chen agape118@so-net.net.tw.
Jensen, L.S., L. Falen, and C.H. Chang. 1974. Effect of distillers grains with solubles on reproduction and liver fat
accumulation in laying hens. Poultry Sci. 53:586-592.
Lumpkins, B., A. Batal and N. Dale, 2004. Evaluation of distillers dried grains with solubles as a feed ingredient
for broilers. Poultry Sci. 83:1891-1896.
Lumpkins, B.S. and A.B. Batal. 2005. The bioavailability of lysine and phosphorus in distillers dried grains with
solubles. Poultry Science 84:581-586.
Lumpkins, B., A. Batal and N. Dale, 2005. Use of distillers dried grains plus solubles in laying hen diets. J. Appl.
Poultry Sci. 14:25-31.

04 - Use of DDGS in Poultry Diets

12

Manley, J.M., R.A. Voitle, and R.H. Harms. 1978. The influence of distillers dried grains with solubles (DDGS)
in the diet of turkey breeder hens. Poultry Sci. 57:726-728.
Martinez Amezcua, C., C. M. Parsons, and S.L. Noll. 2004. Content and relative bioavailability of phosphorus in
distillers dried grains with solubles in chicks. Poultry Sci. 83:971-976.
Matterson, L.D., J. Tlustohowicz, and E.P. Singsen. 1966. Corn distillers dried grains with solubles in rations for
high-producing hens. Poultry Sci. 45:147-151.
National Research Council. 1994. Nutrient Requirements of Poultry, 9th Revised Edition, National Academy
Press, Washington, DC.
Noll, S., V. Stangeland, G. Speers, and J. Brannon. 2001. Distillers grains in poultry diets. 62nd Minnesota
Nutrition Conference and Minnesota Corn Growers Association Technical Symposium, Bloomington, MN.
September 11-12, 2001.
Noll, S., C. Abe, and J. Brannon. 2003. Nutrient composition of corn distillers dried grains with solubles. Poultry
Science 82(Supplement):71.
Noll, S. L., V. Stangeland, G. Speers, C. M. Parsons, and J. Brannon, 2003. Market tom turkey response to protein
and threonine. Poultry Sci. 82 (Suppl. 1): 73.
Noll, S. L., J. Brannon, and V. Stangeland, 2004. Market turkey performance and inclusion level of corn distillers
dried grains with solubles. Poultry Sci. 83 (Suppl. I): 321.
Noll, S. 2004. DDGS in poultry diets: Does it make sense. Midwest Poultry Federation Pre-Show Nutrition
Conference, River Centre, St. Paul, MN. March 16, 2004.
Noll, S. L., J. Brannon, J. L. Kalbfleisch, and K. D.Roberson, 2005. Metabolizable energy value for corn distillers
dried grains with solubles in turkey diets. Poultry Sci. 84 (Suppl. 1):
Roberson, K. D., J. L. Kalbfleisch, W. Pan and R. A. Charbeneau, 2005. Effect of corn distillers dried grains with
solubles at various levels on performance of laying hens and yolk color. Intl J. Poultry Sci. 4(2):44-51.
Shurson, G.C., C. Santos, J. Aguirre, and S. Hernndez. 2003. Effects of Feeding Babcock B300 Laying Hens
Conventional Sanfandila Layer Diets Compared to Diets Containing 10% Norgold DDGS on Performance and Egg
Quality. A commercial field trial sponsored by the Minnesota Corn Research and Promotion Council and the
Minnesota Department of Agriculture.
Spiehs, M.J., M.H. Whitney, and G.C. Shurson. 2002. Nutrient database for distillers dried grains with solubles
produced from new ethanol plants in Minnesota and South Dakota. J. Anim. Sci. 80:2639.
Waldroup, P. W., J.A. Owen, B.E. Ramsey, and D.L. Whelchel, 1981. The use of high levels of distillers dried
grains plus solubles in broiler diets. Poultry Sci. 60:1479-1484.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be formulated by a
qualified nutritionist. The USGC has no control over the nutritional content of any specific product which may be
selected for feeding. Potential buyers should consult an appropriate nutritionist for specific recommendations.
USGC makes no warranties that these recommendations are suitable for any particular herd or for any particular
animal. The USGC disclaims any liability for itself or its members for any problems encountered in the use of these
recommendations. By reviewing this material, buyers agree to these limitations and waive any claims against USGC
for liability arising out of this material.

04 - Use of DDGS in Poultry Diets

13

Use of DDGS in
S w i n e D i e ts

User Handbook

Use of U.S. DDGS in Swine Diets


Historical Use of DDGS in Swine Diets
Historically, limited amounts (less than 3% of total production) of distillers co-products were
used in swine diets until about the year 2000. During the past 60 years, research has been
conducted to evaluate three types of distillers co-products in swine diets distillers dried
solubles (DDS), distillers dried grains (DDG) and distillers dried grains with solubles (DDGS).
In the 1940s and 1950s, most of the research on feeding distillers co-products to swine focused
on evaluating DDS. Performance trials were conducted to measure growth rate and feed
conversion of pigs when DDS was added to starter (Krider et al., 1944; Catron et al., 1954) and
grower-finisher diets (Fairbanks et al., 1944; Beeson et al, 1959). Several studies were also
conducted to determine if DDS could replace common protein (Fairbanks et al., 1945; Hanson,
1948; Winford et al., 1951) and vitamin (Krider and Terrill, 1949) supplements in corn-based
diets during various phases of production.
Beginning in the late 1950s, researchers continued to evaluate growth performance of pigs fed
distillers co-products (Livingstone and Livingston, 1966; Combs and Wallace, 1969; and
Combs and Wallace, 1970), but interest in identifying unidentified growth factor(s) in
distillers co-products and their effects on swine growth performance became a research focus
(Beeson et al., 1959; Couch et al., 1960; Conrad, 1961; Wallace and Combs, 1968).
In the 1970s and 1980s, construction of large scale ethanol plants occurred and researchers
began to focus on evaluating DDGS. A series of titration experiments were conducted to
determine maximal inclusion rates of DDGS that could be added to starter (Wahlstrom and
Libal, 1980; Orr et al., 1981; Cromwell et al., 1985) and grower-finisher diets (Wahlstrom et al.,
1970; Smelski and Stothers, 1972; Cromwell et al., 1983). Additional studies focused on amino
acid content of DDGS and the effect of lysine supplementation on performance of pigs fed diets
containing DDGS (Wahlstrom and Libal, 1980; Cromwell et al. 1983; Cromwell and Stahly,
1986).
From 1986 until 1998, very little research was conducted to evaluate the use of distillers coproducts in swine feeds, even though several new dry-grind fuel ethanol plants were being built.
These relatively new,dry-mill ethanol plants use state-of-the-art engineering designs,
fermentation technologies and drying processes compared to older plants that were built and
operating decades before. Consequently, the nutrient content and digestibility of DDGS
produced by these modern ethanol plants are higher than published in National Research Council
(NRC) in 1998.

Nutritional Value of DDGS for Swine


High quality DDGS has a digestible and metabolizable energy value equal to or greater than
corn. Spiehs et al. (1999) was the first to report that the digestible energy (DE) and metabolizable
energy (ME) values were similar to energy values for corn (3.49 Mcal/kg and 3.37 Mcal/kg,
respectively). Fu et al. (2004) reported that the ME and net energy (NE) values for DDGS were
3.25 Mcal/kg and 2.61 Mcal/kg, respectively, whereas Hastad et al. (2004) reported much higher
values for DE, ME and NE (3.87 Mcal/kg, 3.60 Mcal/kg, and 2.61 Mcal/kg, respectively). Stein

05 - Use of DDGS in Swine Diets

et al. (2006) confirmed that the DE and ME value of DDGS for swine is equal to, or greater than
corn (3,639 kcal DE/kg and 3,378 kcal ME/kg).
Like the low protein quality (low lysine and poor amino acid balance) of corn, DDGS is also
low in lysine relative to its crude protein content. Threonine is the second limiting amino acid
after lysine, and should be monitored during diet formulation when using more than 10% DDGS
in swine diets. Amino acid digestibility can also vary among DDGS sources. Stein et al. (2006)
showed that the range in true lysine digestibility coefficients for swine ranges from 43.9-63.0%.
Fastinger and Mahan (2006) reported a similar range in standardized ileal lysine digestibility
values (38.2-61.5%) when five sources of DDGS were evaluated. Lightness and yellowness of
color of DDGS appear to be reasonable predictors of digestible lysine content among DDGS
sources for swine (Pederson et al., 2005). In order to ensure excellent pig performance when
adding DDGS to swine diets, only light colored sources should be used and diets should be
formulated on a digestible amino acid basis if more than 10% DDGS is included in the diet.
DDGS is an excellent source of available phosphorus for swine. Whitney et al. (2001) showed
that relative phosphorus availability in DDGS was 90%, using dicalcium phosphate as the
inorganic phosphorus reference source.

Use of DDGS in starter diets


Whitney and Shurson (2004) conducted two experiments to determine the effects of increasing
dietary levels (0-25%) of DDGS on growth performance of early-weaned pigs. A total of 96
crossbred pigs (BW = 6.18 0.14 kg) were blocked by gender and ancestry, and pigs within each
block were randomly assigned to one of six dietary treatments (4 pigs/pen, 4 pens/treatment) in
each of two growth performance experiments. Dietary treatments consisted of providing 0, 5, 10,
15, 20, or 25% DDGS during Phases 2 and 3 of a 3-phase nursery feeding program. Pigs in
Experiment 1 were slightly older (19.0 vs. 16.9 days of age) and heavier (7.10 vs. 5.26 kg) at the
beginning of the experiment compared to pigs in Experiment 2. All pigs were provided a
commercial pelleted diet for the first 4 days post-weaning, and were then switched to their
respective experimental Phase 2 diets (fed for a subsequent 14 days), followed by Phase 3
experimental diets (fed for an additional 21 days). Experimental diets were formulated to contain
equivalent apparent ileal digestible lysine (1.35 and 1.15%) and methionine + cystine (0.80 and
0.65%), ME (3340 and 3390 kcal/kg), calcium (0.95 and 0.80%) and total phosphorus (0.80 and
0.70%) within Phases 2 and 3, respectively.
Overall growth rate, ending body weight, and feed conversion of pigs were similar among
dietary treatments regardless of dietary DDGS level fed for both experiments. In Experiment 1,
feed intake was unaffected by dietary treatment. In Experiment 2, however, increasing dietary
DDGS level linearly decreased feed intake during Phase 2, and tended to decrease voluntary feed
intake over the length of the experiment. These results suggest that high quality DDGS can be
included in Phase 3 diets for nursery pigs at dietary levels up to 25%, without negatively
affecting growth performance after a two-week acclimation period. Satisfactory growth
performance can also be achieved when adding up to 25% DDGS in Phase 2 diets for pigs
weighing at least 7 kg in body weight. Including these high levels immediately post-weaning,
however, may negatively influence feed intake, resulting in poorer initial growth performance.
More recently, Gaines et al. (2006) conducted two trials to evaluate the effect of dietary levels
of DDGS and choice white grease on growth performance in the late nursery phase of growth
(more than 11 kg BW). The first trial was conducted to evaluate dietary DDGS inclusion rates of
05 - Use of DDGS in Swine Diets

0, 15 and 30% without supplemental fat. The second trial used the same dietary levels of DDGS
as in the first trial, but also evaluated the effect of adding 0 or 15% choice white grease to the
diet on growth performance. There was no effect of dietary DDGS inclusion level or fat source
on average daily gain. In the second trial, both feeding diets containing DDGS and the addition
of 5% choice white grease improved the gain:feed ratio, which was attributed to lower feed
intake.

Use of DDGS in Grower-Finisher Diets


Whitney et al. (2006c) conducted a study to determine the effects of feeding diets containing 0,
10, 20 or 30% DDGS on growth performance and carcass characteristics of grower-finisher pigs.
They used a total of 240 crossbred pigs with an initial body weight of about 28.6 kg, and
assigned them to one of four diet sequences in a five-phase grower-finisher feeding program.
Corn-soybean meal diets were formulated on total lysine basis, and also contained up to 4%
soybean oil as a supplemental fat source. Soybean oil was chosen as the supplemental fat source
for this study because we did not have the ability to use animal fats at the location where this
study was conducted. Therefore, these experimental diets contained unusually high levels of
unsaturated fatty acids compared what is currently being fed to grower-finisher pigs in the U.S.
pork industry.
As shown in Table 1, pigs fed the diets containing 10% DDGS grew at the same rate,
consumed the same amount of feed and had the same feed conversion as pigs fed the control
corn-soybean meal diets. Feeding diets containing 20% DDGS resulted in reduced growth rate
but feed conversion was not significantly affected. However, feeding diets containing 30%
DDGS reduced growth rate and feed conversion compared to pigs fed the corn-soybean meal
control diets or the diets containing 10% DDGS. This reduction in performance at higher DDGS
inclusion rates was likely due to formulating diets on a total amino acid basis and not accounting
for the digestibility of amino acids in DDGS, which likely resulted in not meeting the pigs amino
acid requirements at the 20 and 30% dietary inclusion rates for DDGS.
Table 1: Effect of Dietary DDGS Level on Overall Growth
Performance of Grower-Finisher Pigs.
0% DDGS
10% DDGS 20% DDGS 30% DDGS
a
0.86
0.86a
0.83bc
0.81bd
Average Daily Gain
(ADG), kg
2.38
2.37
2.31
2.35
Average Daily Feed
Intake (ADFI), kg
2.76a
2.80a
2.92b
2.76a
Feed/Gain (F/G)
117a
117a
114b
112b
Final Wt., kg
a, b

Means within row with unlike superscripts are different (P < .05).
Means within row with unlike superscripts are different (P < .10).

c, d

At the end of the feeding portion of this study, pigs were slaughtered to obtain carcass (Table
2), muscle (Table 3) and fat (Table 4) quality measurements. Carcass weight and dressing
percentage of pigs fed the 0 and 10% DDGS diets were the same and greater than those from
pigs fed the 20 and 30% DDGS diets. The lighter carcass weights of pigs fed the 20% and 30%
DDGS diets were a result of reduced growth rate and lighter live weights compared to pigs fed
the control (0%) and 10% DDGS diets. However, there was no difference in backfat thickness or

05 - Use of DDGS in Swine Diets

percentage of carcass lean among the different DDGS feeding levels. Pigs fed the 0% DDGS
diets had greater loin depths compared to pigs fed the 30% DDGS diets, with intermediate loin
depths from pigs fed either 10 or 20% DDGS. The differences in loin depth were influenced by
the differences in slaughter weight of pigs among the four dietary treatments. These results
indicate that, although growth performance was negatively affected by feeding diets containing
20 or 30% DDGS, carcass composition was largely unaffected as indicated by the similar fat
depths and percent carcass lean across dietary treatments.
Furthermore, none of the muscle quality measurements except 11-day purge loss were affected
by dietary DDGS level (Table 3). It is unclear why muscle from pigs fed the 20% DDGS had a
higher 11-day purge loss compared to muscle from pigs fed the control diet, but 11-day purge
loss was not different between the 0, 10 and 30% DDGS treatments. These data indicate adding
DDGS at levels up to 30% in swine finishing diets did not have meaningful effects on pork
muscle quality.
Iodine number increased linearly, and thus, belly fat became more unsaturated, as the dietary
concentration of DDGS increased (Table 4). Researchers have clearly established that feeding
diets containing an unsaturated fat source can alter the degree of saturation in pork fat. Lea et al.
(1970) indicated that adequately firm pork fat has an iodine number below 70. Boyd (1997)
suggested that the iodine value threshold for pork fat in the United States should be set at 74. In
our study, iodine values were greater than 70, but less than 74, for the diets containing 30%
DDGS and about 70 for the pigs fed the 20% DDGS diets. A significant amount of unsaturated
fatty acids was supplied to experimental diets from supplemental soybean oil in addition to the
corn oil present in DDGS in this study. We estimate, based on NRC (1998), that a typical swine
finishing diet without supplemental fat (85% corn, 11% soybean meal) would contain about 3%
unsaturated fatty acids. By comparison, we estimated our phase 5 control diet contained 4.33%
unsaturated fatty acids and the Phase 5 diet with 30% DDGS contained 4.96% unsaturated fatty
acids. We expect that if an animal fat source, which is lower in unsaturated fatty acid
concentration, were added to these diets, or if no supplemental fat was added, the iodine values
of carcass fat from pigs fed high concentrations of DDGS would be lower and the negative
effects of adding high levels of DDGS to the diets on pork fat quality would be less. The effect
of DDGS feeding on iodine number was reflected in the analysis of belly firmness score. Lower
belly firmness scores indicated that bellies from pigs that were fed 30% DDGS were softer than
bellies from pigs fed 0 or 20% DDGS. Softer bellies were most likely a consequence of elevated
concentrations of dietary unsaturated lipids supplied by soybean oil and DDGS.
Table 2: Effects of Dietary DDGS Level on
Carcass Characteristics of Grower-Finisher Pigs.
0% DDGS 10% DDGS 20% DDGS 30% DDGS
117
119
113
112
Slaughter weight, kg
c
c
d
86.6
81.6
80.7d
85.7
Carcass weight, lbs
c
c
d
73.4
72.8
72.1
71.9d
Dressing %
21.3
21.8
21.1
20.6
Fat depth, mm
ac
b
c
53.9
54.8
51.6d
56.5
Loin depth, mm
52.6
52.0
52.6
52.5
% Carcass lean
a, b
c, d

Means within row with unlike superscripts are different (P < .05).
Means within row with unlike superscripts are different (P < .10).

05 - Use of DDGS in Swine Diets

Table 3: Muscle Quality Characteristics from


Grower-Finisher Pigs Fed Diets Containing 0, 10, 20, and 30% DDGS.
0%
10%
20%
30%
c
54.3
55.1
55.8
55.5
L*
3.2
3.2
3.1
3.1
Color scored
e
2.2
2.0
2.1
2.1
Firmness score
f
1.9
1.9
1.7
1.9
Marbling score
5.6
5.6
5.6
5.6
Ultimate pH
a
b
2.4
2.8
2.5
2.1
11-d purge loss, %
0.7
0.7
0.7
0.7
24-hr drip loss, %
18.7
18.5
18.3
18.8
Cooking loss, %
21.4
21.5
21.8
22.1
Total moisture loss, %g
h
3.4
3.4
3.3
3.3
Warner-Bratzler sheer force, kg
a, b

Means within row with unlike superscripts are different (P < .05).
0 = black, 100 = white
d
1 = pale pinkish gray/white; 2 = grayish pink; 3 = reddish pink; 4 = dark reddish pink; 5 = purplish
red; 6 = dark purplish red
e
1 = soft, 2 = firm, 3 = very firm
f
Visual scale approximates % intramuscular fat content (NPPC, 1999)
g
Total moisture loss = 11-d purge loss + 24-h drip loss + cooking loss
h
Measure of tenderness
c

Table 4: Fat Quality Characteristics of Market Hogs Fed


Corn-Soybean Meal Diets Containing 0, 10, 20 and 30% DDGS.
0%
10%
20%
30%
DDGS
DDGS
DDGS
DDGS
3.15a
3.00ab
2.84bc
2.71c
Belly thickness, cm
24.4a
25.1a
21.3b
27.3a
Belly firmness score, degrees
d
de
d
25.9
23.8
25.4
22.4e
Adjusted belly firmness score,
degrees
66.8d
68.6e
70.6f
72.0f
Iodine number
a, b, c
Means within row with unlike superscripts are different (P < .10).
d, e, f
Means within row with unlike superscripts are different (P < .05).
Based upon these results, including 10% DDGS in conventional swine grower-finisher diets
has no detrimental effects on pig performance, carcass quality or pork quality. When diets are
formulated on a total amino acid basis, it appears that inclusion rates of 20% or higher result in
depressed growth performance. Including DDGS at concentrations of 20 to 30% of the diet, and
using soybean oil as a supplemental fat source for grower-finisher pigs does not affect muscle
composition or quality, but decreases the saturation of fatty acids, resulting in softer bellies and
may negatively affect further processing traits.
A recent commercial field trial conducted by the University of Minnesota and Land O
Lakes/Purina Feed was conducted in the summer of 2006 to further evaluate the impact of
feeding conventional corn-soybean meal grower-finisher diets with or without 10% DDGS
on pork fat quality. Two cooperating pork producers were selected for this study. Each producer
had typical commercial 1,000 head finishing barns and were located in southern Minnesota. Each

05 - Use of DDGS in Swine Diets

40-pen barn was a double curtain sided building with 8 foot pits, utilized pit fans for ventilation
and weighted baffle ceiling air inlets. Both farms had common genetics consisting of Monsanto
Genepacker sows mated with Monsanto EB terminal line boar semen. Overall health status of
both groups of pigs was very good. Feed for both farms was formulated and provided by Land
O Lakes/Purina Feed. Producer A fed typical corn-soybean meal diets, whereas Producer B fed
corn-soybean meal diets containing 10% DDGS. An eight-phase mixed sex feeding program was
used and the last finisher diet contained 4.5g Paylean. Diets within each phase contained similar
nutrient levels with and without 10% DDGS. All diets within each phase contained the same
level of choice white grease as the supplemental fat source (supplemental levels ranged from
1.25-3.75% depending on the diet phase).
One hundred twenty eight pigs were randomly selected from each group for evaluation of carcass
traits. At 24 hours postmortem, a total of 48 mid-belly samples were collected from each dietary
treatment group, with equal numbers of barrows (n=12) and gilts (n=12) from each farm. From
the 48 mid-belly samples, a visual color score (on a scale from 1-4 with 1 = pale and 4 = dark)
was determined by a group of six panelists using a visual system for Japanese pork fat color
scores. All belly fat samples were then analyzed to determine complete fatty acid profiles. Iodine
value and mean melting point were calculated using fatty acid data from each sample.
As shown in Table 5, pigs fed the 10% DDGS grew equally well, consumed less feed, had
better feed conversion and lower feed cost per pound of gain compared to pigs fed the cornsoybean diets without DDGS. At slaughter, there were no differences in carcass weight, backfat
thickness or percentage of ham, loin and belly relative to total carcass weight (Table 6). In
addition, there were no differences in loin depth or percentage of lean muscle in the carcasses
between the two groups. These results are in agreement with the growth performance and carcass
composition results obtained in the study conducted by Whitney et al. (2006c) and clearly show
that feeding corn-soybean meal diets containing 10% DDGS have no negative on growth
performance and carcass characteristics of grower-finisher pigs. In fact, the producer who fed the
DDGS diets in this study obtained the same carcass quality at a lower feed cost per pound of gain
compared to the producer who fed diets without DDGS.
When the composition and quality characteristics of belly fat from these pigs were evaluated,
there were no differences in color score based upon Japanese pork fat quality standards (Table
7), nor were there any differences in mean melting point of the belly fat. However, bellies from
pigs fed the 10% DDGS diets had a higher iodine value than pigs fed the diets without DDGS.
This is also in agreement with the results obtained in the study reported by Whitney et al. (2006)
shown in Table 4. The iodine values are similar and below the suggested maximum threshold of
70. These results clearly show that feeding diets containing 10% DDGS to grower-finisher pigs
have negative effects on pork fat quality. As expected, the levels of linoleic acid, polyunsaturated
fatty acids and omega 6 fatty acids increase in belly fat when pigs are fed diets containing 10%
DDGS, but are well within accepted standards of acceptable pork fat quality.

05 - Use of DDGS in Swine Diets

Table 5: Growth Performance, Feed Usage and Feed Cost of


Grower-Finisher Pigs Fed Diets Containing 0 or 10% DDGS.
0% DDGS 10% DDGS
0.82
0.83
ADG, kg
2.24
2.10
ADFI, kg
2.54
2.73
F/G
258.5
251.2
Kg Feed/Head
0.077
0.073
Feed Cost/Lb Gain, $
Table 6: Carcass Characteristics of Grower-Finisher
Pigs Fed Diets Containing 0 or 10% DDGS.
0% DDGS 10% DDGS
96.1
95.2
Carcass weight, kg
27.3
27.8
Last rib backfat, mm
25.3
24.8
Tenth rib backfat, in.
11.74
11.74
Ham, %
7.93
7.91
Loin, %
10.51
10.41
Belly, %
68.0
68.0
Loin depth, mm
56.36
56.47
% Carcass lean
Table 7: Mid-Belly Fat Quality Characteristics of Carcasses
from Grower-Finisher Pigs Fed Diets Containing 0 or 10% DDGS.
Measurement
0% DDGS
10% DDGS
Japanese fat color score
1.76
1.81
Mean melting point, C
29.3
28.7
a
Iodine value
66.7
68.3b
Oleic acid (18:1), %
47.39c
45.12d
c
Linoleic acid (18:2), %
11.94
13.98d
Saturated fatty acids, %
33.99
34.26
c
Monounsaturated fatty acids, %
51.78
49.47d
Polyunsaturated fatty acids, %
14.02c
16.11d
Total omega 3 fatty acids, %
0.98
0.96
c
Total omega 6 fatty acids, %
13.02
15.14d
Omega 6:omega 3 ratio
13.28c
15.78d
a, b
Means within row with unlike superscripts are different (P < .05).
c, d
Means within row with unlike superscripts are different (P < .0001).
Based upon these research results, there is no reason for concern when feeding grower-finisher
diets containing 10% DDGS on carcass or pork quality. The composition of some fatty acids
(e.g. linoleic acid, polyunsaturated fatty acids and omega 6 fatty acids) in pork fat increase with
the addition of DDGS to corn-soybean meal diets, but do not alter the acceptability based upon
current industry standards. Furthermore, there is no evidence suggesting that feeding growerfinisher pigs diets containing 10% DDGS will decrease the quality and acceptability of U.S. pork
in the Japanese export market.

05 - Use of DDGS in Swine Diets

Gralapp et al. (2002) conducted two experiments to evaluate the impact of the level of DDGS
added to grower-finisher diets on manure characteristics, odor emissions and growth
performance. Three diets containing 0, 5 or 10% DDGS were fed to 72 finishing pigs during six
four-week periods. They found that ADG and feed efficiency were reduced at higher DDGS
dietary inclusion rates, but there was a tendency for higher feed intake in pigs fed the 10%
DDGS diet, 2.91 kg/day vs. 2.73 and 2.75 kg/day for pigs fed the 5 and 0% DDGS diets,
respectively. These results confirm that pig growth performance is not affected when fed diets
containing 10% DDGS compared to feeding typical corn-soybean meal diets. DeDecker et al.
(2005) showed that feeding grower-finisher diets containing 30% DDGS could be achieved
without any negative effects on growth performance, but carcass yield decreased linearly as
dietary DDGS level increased.

Use of DDGS in gestation and lactation diets


Three studies have been conducted to determine the optimum inclusion rate of DDGS in diets
for sows during gestation and lactation (Thong et al., 1978; Monegue and Cromwell, 1995;
Wilson et al., 2003), and recommendations for maximum dietary inclusion rates have been
published based upon results obtained by Thong et al., 1964 and Monegue and Cromwell, 1995
(Weigel et al., 1997; Pork Industry Handbook, 1998). As a result of limited information of
feeding DDGS to sows, current recommendations for DDGS inclusion for use of DDGS in sow
diets are somewhat different. The Feed Co-Products Handbook (Weigel et al., 1997) lists the
maximum inclusion rate for DDGS to be up to 50% in gestation diets and up to 20% in lactation
diets. The Pork Industry Handbook, however, recommends slightly lower levels of DDGS usage,
suggesting up to 40% in gestation diets and a maximum inclusion rate of 10% in lactation diets
(PIH Factsheet #112).
Thong et al. (1978) conducted an experiment using 64 gilts to evaluate the use of DDGS as a
replacement for soybean meal in a corn-soybean meal diet fed during gestation. To conduct this
experiment, sows were fed diets containing either 0, 17.7 or 44.2% DDGS during gestation. All
diets were formulated to contain 0.42% total dietary lysine. Number of pigs farrowed per litter
and average pig birth weight were not significantly affected by dietary treatment. The authors
concluded that DDGS could replace soybean meal on a lysine-equivalent basis as a source of
supplemental amino acids at levels up to 44.2% of the diet for gestating sows.
Monegue and Cromwell (1995) compared reproductive performance of sows fed a fortified
corn-soybean meal diet to sows fed diets containing 40 or 80% corn gluten feed (CGF) and sows
fed diets containing 40 or 80% DDGS during gestation. A total of 90 parity 4 crossbred sows (18
sows/dietary treatment) were used in this study. Diets contained similar levels of total lysine and
were fed at different levels to equalize ME intake at 6.2 Mcal/sow/day. Sows were allowed to
consume a fortified corn-soybean meal diet ad libitum during the subsequent 28-day lactation
period. Farrowing rates averaged 91% and were not affected by dietary treatment. Gestation
weight gains tended to be greater in sows fed the CGF and DDGS diets indicating that the energy
in these co-products was well utilized. Lactation feed intake and sow weight loss during lactation
were similar among dietary treatments. Litter size at birth and pig birth weights were not affected
by dietary treatment, although numerically, sows fed the 80% DDGS had slightly smaller litters.
Litter size weaned and litter weaning weights were not different among dietary treatments,
although feeding the 80% CGF diet and the DDGS diets during lactation numerically reduced
litter size weaned and increased individual pig weight at weaning. There were no differences in

05 - Use of DDGS in Swine Diets

litter weaning weight and pig survival percentage to weaning among dietary treatments. Days for
sows to return to estrus following weaning were similar among dietary treatment groups and
averaged 4.7 days. The authors concluded that diets containing high levels of CGF and DDGS,
up to 80% of the gestation diet, are well utilized, and do not appear to impair reproductive or
lactation performance.
More recently, Wilson et al. (2003) conducted a two-parity study utilizing 93 multiparous sows
to determine the effects of feeding diets containing 50% DDGS in gestation and 20% DDGS in
lactation on sow reproductive performance. Nutrient balance was also determined from day 100
to day 105 of pregnancy using 14 gestating sows. Sows were allotted based on parity and initial
body weight to one of two gestation diets (0 or 50% DDGS, corn-soybean meal based diets), and
one of two lactation diets (0 or 20% DDGS, corn-soybean meal based diets). Sows were fed a
daily amount of feed based on 1% of sow body weight plus 100 g, 300 g and 500 g per day on
days 0 to 30, 31 to 60 and 61 to 90 days of gestation, respectively. Sows were provided ad
libitum access to feed during lactation. Sows remained on their respective dietary treatment
combinations through two reproductive cycles. No differences in sow gestation weight gain, pigs
born alive per litter, litter birth weight, or average pig birth weight were observed between sows
fed 0 and 50% DDGS diets during gestation for both reproductive cycles. Dietary treatment
combination had no effect on litter size, litter birth weight or litter weaning weight during the
first reproductive cycle, but sows fed 0% DDGS gestation and lactation diets weaned fewer pigs
per litter during the second reproductive cycle. Pre-weaning mortality was higher for sows fed
the 50% DDGS gestation diet and 20% DDGS lactation diet compared to other treatment
combinations during the first reproductive cycle, but dietary treatment combinations had no
effect on pre-weaning mortality during the second reproductive cycle. Sows fed the 0% DDGS
gestation diet and the 20% DDGS lactation diet had lower lactation feed intake, which primarily
occurred within the first seven days of lactation, but this effect was not observed during the
second reproductive cycle. Wean-to-estrus interval was higher for sows fed the 0% DDGS
gestation and lactation diet treatment combination compared to sows fed the 50% DDGS
gestation, 20% DDGS lactation diet combination and the 50% DDGS gestation, 0% DDGS
lactation diet combination during the first reproductive cycle. No wean-to-estrus interval
differences were observed during the second reproductive cycle. Sows fed the 50% DDGS diet
in late gestation consumed more energy, nitrogen, sulfur and potassium, and had greater
nitrogen, sulfur and phosphorus retention than sows fed the 0% DDGS gestation diet. These
results indicate that feeding a gestation diet containing 50% DDGS will support good
reproductive performance. However, feeding a 20% DDGS lactation diet may reduce feed intake
during the first week post-partum if sows were fed a corn-soybean meal diet during gestation and
not provided an adjustment period to adapt to a high DDGS diet during lactation.
Hill et al. (2005) conducted a study to determine if lactating sows could utilize diets containing
15% DDGS to maintain body weight and lactation performance while decreasing manure
phosphorus excretion. Their results showed that the inclusion of 15% DDGS in a lactation diet
supports good sow performance while maintaining and perhaps reducing manure phosphorus
excretion.

DDGS and manure management


Spiehs et al. (2000) conducted a 10-week trial to measure odor and gas characteristics of swine
manure and energy, nitrogen, and phosphorus balance of grow-finish pigs fed corn-soybean meal

05 - Use of DDGS in Swine Diets

based diets containing 0 or 20% DDGS. Sixteen PIC barrows weighing 57.6 3.8 kg were
randomly assigned to one of two dietary treatments (eight pigs/treatment): control (0% DDGS)
and 20% DDGS. A three-phase diet sequence was used. Calculated total lysine and phosphorus
levels were identical for both diets within each phase. Manure from each pig, housed in
collection cages, was collected daily except during the last three days of weeks 2, 6 and 10, when
total fecal and urinary excretion was collected for nutrient balance measurements. Urine and
feces were mixed and emptied into simulated anaerobic manure pits according to respective
dietary treatments. Air samples were collected weekly from the headspace above each simulated
pit and analyzed for hydrogen sulfide (H2S) and ammonia (NH3). Air samples collected during
weeks 0, 2, 5 and 8 were evaluated for odor detection level utilizing a human odor panel and
olfactometer.
Dietary treatment had no effect on H2S, NH3 or odor detection levels over the 10-week trial.
Pigs fed the DDGS diets had greater nitrogen (N)and gross energy (GE) intake in all three of the
growth phases, but average daily feed intake was not different among treatments. Dietary DE and
ME (kcal/kg) were not different between the two experimental diets. Percentage of nitrogen
retention was not different between dietary treatments, but feeding DDGS tended to increase N
intake and excretion during all three phases. Percentage of phosphorus retention was not
different between dietary treatments. These results suggest that feeding 20% DDGS has no effect
on H2S, NH3 and odor levels over a 10-week manure storage period compared to feeding cornsoybean meal diets. Feeding DDGS increases GE intake and improves phosphorus utilization
during late finishing phases, but also increases N excretion. When diets containing DDGS are
formulated on an available phosphorus basis using the available phosphorus value obtained by
Whitney et al. (2001), one would expect the phosphorus excretion in swine manure to be
reduced.

Effect of Feeding DDGS on Gut Health of Growing Pigs


Whitney et al. (2006a, b) conducted two experiments to determine if including DDGS in the
diet of young growing pigs reduces the incidence or severity of clinical signs, fecal shedding,
intestinal lesions and/or cellular infection indicating porcine proliferative enteropathy (ileitis)
after challenge with Lawsonia intracellularis. In the first experiment, 80 pigs were weaned at 17
days of age and were randomly allotted (blocked by sex and weight) to one of four treatment
groups. A negative control group was unchallenged and fed a control corn-soybean meal diet.
The remaining 3 groups were inoculated orally with 1.5 x 109 L. intracellularis per pig after a
four-week dietary adaptation period, and were fed either a control corn-soybean meal diet or a
similar diet containing 10 or 20% DDGS. On day 21 post-challenge, all pigs were euthanized
and intestinal mucosa was examined for the presence of lesions. Ileal tissue samples were
analyzed to determine presence and proliferation of L. intracellularis. Challenging pigs reduced
ADFI, ADG and G/F by 25, 55 and 40%, respectively, during the three-week post-challenge
period. Dietary treatment did not affect growth performance. Gross lesions were observed in
63% of challenged pigs compared to 0% in the negative control group. Including DDGS in the
diet did not positively affect lesion prevalence and length, proliferation of L. intracellularis, or
severity of lesions. In the second experiment, 100 pigs were managed similar to the first
experiment, except the L. intracellularis dosage rate for challenging pigs was reduced by 50%.
Treatment groups consisted of a negative control group and four treatments in a 2x2 factorial
arrangement testing the effect of 10% dietary DDGS inclusion and/or antimicrobial regimen.

05 - Use of DDGS in Swine Diets

10

Antimicrobial regimen consisted of providing 30 mg BMD/kg diet (supplied continuously in


the diet), with chlortetracycline (Aureomycin) pulsed at 500 mg/kg from 3 days prior to 11 days
post-challenge. Feeding diets containing 10% DDGS reduced ileum and colon lesion length and
prevalence and reduced the severity of lesions in the ileum and colon compared to other
challenged pigs. Pigs fed the antimicrobial regimen reduced prevalence and severity of lesions in
the jejunum and tended to have reduced total tract lesion length. The combination of DDGS and
antimicrobial resulted in no differences in length, severity or prevalence of lesions, but fecal
shedding of L. intracellularis was reduced on day 14 post-challenge. The proportion of intestinal
cells infected with L. intracellulariswas reduced when DDGS or antimicrobials were fed. In
conclusion, it appears that dietary inclusion of DDGS may aid the young growing pig in resisting
a moderate ileitis challenge similar to a U.S.-approved antimicrobial regimen, but under more
severe challenges, DDGS may not be effective.

Recommended Maximum Inclusion Rates of DDGS in Swine Diets


Based upon current research results, the maximum usage rate of DDGS in swine diets is as
follows:
Production Phase
Nursery pigs (>7 kg)
Grow-finish pigs
Developing gilts
Gestating sows
Lactating sows
Boars

Maximum % of Diet
30
20
20
50
20
50

These recommendations assume that high quality DDGS is free of mycotoxins. Nursery diets
containing up to 30% DDGS will support growth performance equivalent to feeding pigs cornsoybean meal based diets provided that diets are formulated on a digestible amino acid and
available phosphorus basis. Similarly, grower-finisher and gilt development diets containing
levels up to 30% DDGS should provide equivalent growth performance compared to pigs fed
corn-soybean meal diets if they are formulated on a digestible amino acid and available
phosphorus basis. However, due to concerns of reduced belly firmness and soft pork fat at high
levels of DDGS inclusion, we recommend no more than 20% DDGS be added to grower-finisher
diets. If the DDGS supplier has a quality control program that includes screening corn and/or
DDGS for mycotoxins, developing gilt diets can contain up to 20% DDGS.
For sows, up to 50% DDGS can be successfully added to gestation diets, and 20% DDGS can
be added to the lactation diet if DDGS is free of mycotoxins. If there are no assurances that
DDGS is mycotoxin free, no more than 20% should be added to gestation diets and no more than
10% DDGS should be added to lactation diets to minimize the risk of mycotoxicosis. However,
when switching sows from a corn-soybean meal diet to diets containing DDGS, gestation diets
should be initially formulated to contain 20% DDGS and then the level of DDGS can be
increased when each new batch of feed is made to allow the sows to adapt to the DDGS diets and
avoid reduced feed intake. Similarly, when switching from a corn-soybean meal diet to a DDGS
diet for lactating sows, begin feeding a 10% DDGS diet to allow the sows to adapt

05 - Use of DDGS in Swine Diets

11

(approximately 5 to 7 days) before feeding the maximum recommended level to avoid potential
reductions in feed intake.

Literature Cited
Beeson, W.M., D.L. Jeter, and J.H. Conrad. 1959. Effect or organic and inorganic sources of unidentified growth
factors on the growing pig. Proc. Distillers Feed Conf. 14:62-69.
Boyd, R.D. 1997. Relationship between dietary fatty acid profile and body fat composition in growing pigs. PIC
USA T & D Technical Memo 153. Pig Improvement Company,USA, Franklin, Kentucky.
Catron, D.V., F. Diaz, V.C. Speer and G.C. Ashton. 1954. Distillers dried solubles in pig starters. Proc. Distillers
Feed Conf. 9:49-51.
Combs, G.E. and H.D. Wallace. 1969. Dried distillers grains with solubles in pig starter diets. Florida Agric.
Expt. Station, Gainesville. Mimeograph Series No. AN69-14.
Combs, G.E. and H.D. Wallace. 1970. Dried distillers grains with solubles for growing finishing pigs. Florida
Agric. Expt. Station, Gainesville. Mimeograph Series No. AN70-13.
Conrad, J.H. 1961. Recent research and the role of unidentified growth factors in 1961 swine rations. Proc.
Distillers Feed Conf. 16:41-51.
Couch, J.R., H.D. Stelzner, R.E. Davies, and C.W. Deyoe. 1960. Isolation of an unidentified factor from corn
distillers dried solubles. Proc. Distillers Feed Conf. 15:11-19.
Cromwell, G.L., T.S. Stahly, H.J. Monegue, and J.R. Overfield. 1983. Distillers dried grains with solubles for
growing-finishing swine. Kentucky Agric. Expt. Station, Lexington. Progress Report 274. p. 30-32.
Cromwell, G.L., T.S. Stahly, and H.J. Monegue. 1985. Distillers dried grains with solubles and antibiotics for
weanling swine. Kentucky Agric. Expt. Station, Lexington. Progress Report 292. p. 10-11.
Cromwell, G.L. and T.S. Stahly. 1986. Distillers dried grains with solubles for growing finishing swine. Proc.
Distillers Feed Conf. 41:77-87.
Cromwell, G.L., K.L. Herkelman, and T.S. Stahly. 1993. Physical, chemica, and nutritional characteristics of
distillers dried grains with solubles for chicks and pigs. J. Anim. Sci. 71:679-686.
DeDecker, J.M., M. Ellis, B.F. Wolter, J. Spencer, D.M. Webel, C.R. Bertelsen and B.A. Peterson. 2005. Effects
of dietary level of distiller dried grains with solubles and fat on the growth performance of growing pigs. J. Anim.
Sci. 83(Suppl. 2):49.
Fairbanks, B.W., J.L. Krider, and W.E. Carroll. 1944. Distillers by-products in swine rations. I. Creep-feeding and
growing-fattening rations. J. Anim. Sci. 3:29-40.
Fairbanks, B.W., J.L. Krider, and W.E. Carroll. 1945. Distillers by-products in swine rations. III. Dried corn
distillers solubles, alfalfa meal, and crystalline B-vitamins compared for growing-fattening pigs in drylot. J. Anim.
Sci. 4:420-429.
Fastinger, N.D. and D.C. Mahan. 2006. Determination of the ideal amino acid and energy digestibilities of corn
distillers dried grains with solubles using grower-finisher pigs. J. Anim. Sci. 84:1722-1728.
Fu, S.X., M. Johnston, R.W. Fent, D.C. Kendall, J.L. Usry, R.D. Boyd, and G.L. Allee. 2004. Effect of corn
distillers dried grains with solubles (DDGS) on growth, carcass characteristics and fecal volume in growing
finishing pigs. J. Anim. Sci. 82 (Suppl. 2):50.
Gaines, A, B. Ratliff, P. Srichana, and G. Allee. 2006. Use of corn distillers dried grains and solubles in late
nursery pig diets. J. Anim. Sci. 84(Suppl.2):89.
Gralapp, A.K., W.J. Powers, M.A. Faust, and D.S. Bundy. 2002. Effects of dietary ingredients on manure
characteristics and odorous emissions from swine. J. Anim. Sci. 80:1512-1519.
Hanson, L.E. 1948. Swine feeding trials with distillers solubles. Proc. Distillers Feed Research Conf. 3:47-56.
Hastad, C.W., M.D. Tokach, J.L. Nelssen, R.D. Goodband, S.S. Dritz, J.M. DeRouchey, C.N. Groesbeck, K.R.
Lawrence, N.A. Lenehan, and T.P Keegan. 2004. Energy value of dried distillers grains with solubles in swine diets.
J. Anim. Sci. 82 (Suppl. 2):50.
Hill, G.M., J.E. Link, M.J. Rincker, K.D. Roberson, D.L. Kirkpatrick, and M.L. Gibson. 2005. Corn distillers
grains with solubles in sow lactation diets. J. Anim. Sci. 83 (Suppl. 2):82.

05 - Use of DDGS in Swine Diets

12

Krider, J.L, B.W. Fairbanks and W.E. Carroll. 1944. Distillers by-products in swine rations. II. Lactation and
growing-fattening rations. J. Anim. Sci. 3:107-119.
Krider, J.L. and S.W. Terrill. 1949. Recent work at the Illinois station on distillers grain solubles in swine rations.
Proc. Distillers Feed Research Conf. 4:21-33.
Lea, C. H., P. A. T. Swoboda, and D. P. Gatherum. 1970. A chemical study of soft fat in crossbred pigs. J. Agric.
Sci. Camb. 74:1-11.
Livingstone, R.M. and D.M.S. Livingston. 1966. A note on the use of distillers by-products in diets for growing
pigs. Anim. Prod. 11:259-261.
Monegue, H.J. and G.L. Cromwell. 1995. High dietary levels of corn byproducts for gestating sows. J. Anim. Sci.
73 (Suppl. 1):86.
National Research Council. 1998. Nutrient Requirements of Swine. 10th ed. National Academy Press,
Washington, D.C.
Orr, D.E., W.F. Owsley and L.F. Tribble. 1981.Use of corn distillers dried grains, dextrose, and fish meal. Proc.
29th Swine Short Course, Texas Agric. Expt. Station, Lubbock, pp. 48-50.
Pederson, C., A. Pahm, and H.H. Stein. 2005. Effectiveness of in vitro procedures to estimate CP and amino acid
digestibility coefficients in dried distillers grain with solubles by growing pigs. J. Anim. Sci. (Suppl. 2) 83:39.
Pork Industry Handbook. 1998. Relative value of feedstuffs for swine. Purdue University, Factsheet 112.
Smelski, R.B. and S.C. Stothers. 1972. Evaluation of corn distillers dried grains with solubles for finishing pigs.
Proc. Western Sec. Am. Soc. Anim. Sci. 23:122-127.
Spiehs, M.J., G.C. Shurson, and M.H. Whitney. 1999. Energy, nitrogen, and phosphorus digestibility of growing
and finishing swine diets containing distillers dried grains with solubles. J. Anim. Sci. 77:188 (Suppl. 1).
Spiehs, M.J., M.H. Whitney, G.C. Shurson, and R.E. Nicolai. 2000. Odor characteristics of swine manure and
nutrient balance of grow-finish pigs fed diets with and without distillers dried grains with solubles. J. Anim. Sci.
78:69 (Suppl. 2).
Spiehs, M.J., M.H. Whitney, and G.C. Shurson. 2002. Nutrient database for distillers dried grains with solubles
produced from new ethanol plants in Minnesota and South Dakota. J. Anim. Sci. 80:2639.
Stein H. H., M. L. Gibson, C. Pedersen, and M. G. Boersma. 2006. Amino acid and energy digestibility in ten
samples of distillers dried grain with solubles fed to growing pigs. J. Anim. Sci. 84: 853-860.
Thong, L.A., A.H. Jensen, B.G. Harmon, and S.G. Cornelius. 1978. Distillers dried grains with solubles as a
supplemental protein source in diets for gestating swine. J. Anim. Sci. 46:674-677.
Wahlstrom, R.C., C.S. German, and G.W. Libal. 1970. Corn distillers dried grains with solubles in growingfinishing swine rations. J. Anim. Sci. 30:532-535.
Wahlstrom, R.C. and G.W. Libal. 1980. Effect of distillers dried grains with solubles in pig starter diets. SDSU
Swine Day. Bull. No. 80-6. p. 14-16. Brookings, SD.
Wallace, H.D. and G.E. Combs. 1968. Distillers dried corn solubles as a source of unidentified nutritional
factor(s) for the gestating-lactating sow. Florida Agric. Expt. Station, Gainesville. Mimeograph Series No. AN69-3.
Weigel, J.C., D. Loy, and L. Kilmer. 1997. Feed Co-Products of the Dry Corn Milling Process. Renewable Fuels
Association and National Corn Growers Association. Washington, D.C. and St. Louis, MO.
Whitney, M.H., M.J. Spiehs, and G.C. Shurson. 2001. Availability of phosphorus availability of distillers dried
grains with solubles for growing swine. J. Anim. Sci. 79 (Suppl. 1):108.
Whitney, M.H. and G.C. Shurson. 2004. Growth performance of nursery pigs fed diets containing increasing
levels of corn distillers dried grains with solubles originating from a modern Midwestern ethanol plant. J. Anim.
Sci. 82:122-128.
Whitney, M.H., G.C. Shurson, and R.C. Guedes. 2006a. Effect of dietary inclusion of distillers dried grains with
solubles on the ability of growing pigs to resist a Lawsonia intracellularis challenge. J. Anim. Sci. 2006. 84:1860
1869.
Whitney, M.H., G.C. Shurson, and R.C. Guedes. 2006b. Effect of including distillers dried grains with solubles in
the diet, with or without antimicrobial regimen, on the ability of growing pigs to resist a Lawsonia intracellularis
challenge. J. Anim. Sci. 2006. 84:18701879.

05 - Use of DDGS in Swine Diets

13

Whitney, M.H, G.C. Shurson, L.J. Johnston, D. Wulf, and B. Shanks. 2006c. Growth performance and carcass
characteristics of pigs fed increasing levels of distillers dried grains with solubles. J. Anim. Sci. 84:(in press).
Wilson, J.A., M.H. Whitney, G.C. Shurson, and S.K. Baidoo. 2003. Effects of adding distillers dried grain with
solubles (DDGS) to gestation and lactation diets on reproductive performance and nutrient balance. J. Anim. Sci. 81:
(Suppl. 1).
Winford, E.J., W.P. Garrigus, and C.E. Barnhart. 1951. Distillers dried solubles as a protein supplement for
growing and fattening hogs in drylot. Kentucky Ag. Expt. Station. Bull. No. 577. p.3-16. Lexington, KY.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be formulated by a
qualified nutritionist. The USGC has no control over the nutritional content of any specific product which may be
selected for feeding. Potential buyers should consult an appropriate nutritionist for specific recommendations.
USGC makes no warranties that these recommendations are suitable for any particular herd or for any particular
animal. The USGC disclaims any liability for itself or its members for any problems encountered in the use of these
recommendations. By reviewing this material, buyers agree to these limitations and waive any claims against USGC
for liability arising out of this material.

05 - Use of DDGS in Swine Diets

14

Use of DDGS in
A q u a c u lt u r e D i e ts

User Handbook

Use of U.S. DDGS in Aquaculture Diets


Aquaculture is one of the fastest growing food producing industries in the world. Fish meal has
traditionally been used in commercial fish feed as a major source of dietary protein for many
years. However, when global fish meal production declines and fish meal prices increase, fish
nutritionists begin considering less expensive plant protein sources. Plant protein sources have
traditionally been considered to be inferior to fish meal in fish diets. However, when two or more
complimentary plant protein sources, such as distillers dried grains with solubles (DDGS) and
soybean meal, are added to the diet, the potential exists to replace all of the fish meal in the diet.
Therefore, fish nutritionists are continually evaluating alternative plant protein sources as a
means to reduce or replace expensive fish meal sources in order to reduce diet cost. Due to its
moderately high protein content, relatively low phosphorus content and low cost compared to
fish meal, there is increasing global interest in using DDGS in aquaculture diets. Furthermore,
DDGS does not contain anti-nutritional factors found in other protein sources such as soybean
meal (trypsin inhibitors) or cottonseed meal (gossypol).
Aquaculture, like livestock and poultry production around the world, is also subject to
increasing environmental regulations. The two nutrients of greatest concern in fish farm effluent
water are nitrogen and phosphorus. Soybean meal and distillers dried grains with
solubles(DDGS) are relatively high in protein but much lower in phosphorus than fish meal. As a
result, substituting DDGS and soybean meal for fish meal in aquaculture diets reduces the total
phosphorus level in the diet and lowers the level of phosphorus in fish farm discharge water.
This paper describes the results of recent studies involving the feeding of diets containing
DDGS and to provide revised maximum dietary inclusion rates for DDGS based upon these
results.

Channel Catfish (Ictalurus punctatus)


One of the attractive features of using DDGS in channel catfish diets is that it does not contain
anti-nutritional factors found in other protein sources such as soybean meal (trypsin inhibitors Wilson and Poe, 1985; Shiau et al., 1987) or cottonseed meal (gossypol Jauncey and Ross,
1982; Robinson, 1991). Tidwell et al. (1990) conducted an experiment over an 11-week period
where channel catfish fingerlings were fed diets containing 0, 10, 20 and 40% distillers grains
with solubles, replacing some of the corn and soybean meal. After the 11-week feeding period,
there were no significant differences in individual fish weight, percentage survival, feed
conversion or protein efficiency ratio among dietary treatments (Table 1). However, fish fed the
20% DDGS diet were slightly shorter in length compared to fish fed the other dietary treatments.

06 - Use of DDGS in Aquaculture Diets

Table 1: Length, survival, final body weight, feed conversion


and protein efficiency ratio (PER) in channel catfish fingerlings
fed diets containing four levels of distillers grains with solubles.
0% DDGS 10% DDGS 20% DDGS 40% DDGS
Length, mm
115.2
114.1
107.4
117.8
Survival, %
67.5
70.0
80.0
90.0
Final body weight, g
17.3
15.2
13.2
16.5
Feed/gain (F/G)
2.85
3.23
3.20
2.60
PER
0.99
0.87
0.88
1.05
In a study conducted by Webster et al. (1993), cage reared juvenile catfish were fed diets
containing 0, 10, 20 or 30% DDGS to partially replace corn and soybean meal in the diet. There
were no differences in individual fish weights, survival, feed conversion, carcass composition,
carcass waste (head, skin, viscera) and organoleptic properties of the filets among dietary
treatments. Results from this study suggest that up to 30% DDGS can be added to channel
catfish diets with no negative effects on growth performance, carcass composition or flavor
qualities of the filets. Therefore, DDGS is considered an acceptable ingredient in diets for
channel catfish (Tidwell et al., 1990; Webster et al., 1991).

Rainbow Trout (Oncorhynchus mykiss)


Feed for carnivorous fish like rainbow trout (Oncorhynchus mykiss) requires large amounts of
fish meal (300 to 500 g/kg diet). As a result, when fish meal prices are high, nutritionists begin
evaluating alternative protein sources such as DDGS to use as partial replacements for fishmeal.
Cheng and Hardy (2004a) reported that they had unpublished data indicating that apparent
digestibility coefficients of nutrients in DDGS were high for rainbow trout. Apparent
digestibility coefficients for crude protein, essential amino acids and non-essential amino acids
were 90.4%, more than 90% (except for threonine, 87.9%), and more than 86% (except for
cystine, 75.9%), respectively. However, they pointed out that one of the limitations of using
DDGS in rainbow trout diets is the relatively low levels of the most limiting amino acids (lysine
and methionine), which are much lower than found in fish meal. Therefore, supplemental
synthetic lysine and methionine is necessary in order to achieve satisfactory growth performance.
Cheng and Hardy (2004a) conducted a six-week feeding trial to determine the effects of
feeding diets containing 0, 7.5, 15 and 22.5% DDGS with or without synthetic lysine and
methionine supplementation to assess the nutritional value of DDGS in diets for 50 g rainbow
trout. Survival rate of all fish used in the study was 100%. Fish fed diets containing 15% DDGS,
or replacing 50% of fish meal on an isonitrogenous and isocaloric basis had similar weight gain
and feed conversion compared to fish fed the fish meal-based diet. This suggests that DDGS can
be added up to 15% of the diet, or can replace up to 50% of the fish meal in a diet without the
need for supplemental lysine and methionine to achieve satisfactory growth performance. In
addition, DDGS could be used at the 22.5% dietary inclusion level, or replace up to 75% of the
fish meal in rainbow trout diets with lysine and methionine supplementation. Furthermore,
Cheng et al. (2003) showed that when soybean meal, DDGS and 1.65 g/kg of methionine
hydroxyl analogue (MHA) were added to rainbow trout (50 g in initial body weight) diets to
replace 50% of the fish meal, weight gain, feed conversion, crude protein and phosphorus

06 - Use of DDGS in Aquaculture Diets

retention were significantly improved compared to fish fed the an equivalent diet without MHA
supplementation.
Cheng and Hardy (2004b) also evaluated the effects of phytase supplementation on apparent
digestibility coefficients of nutrients in DDGS, as well as growth performance and apparent
nutrient retention of rainbow trout fed diets containing DDGS, phytase and varying levels of a
trace mineral premix. Apparent digestibility coefficients in DDGS diets (30% of total diet)
containing different levels of phytase (0, 300, 600, 900 and 1200 FTU/kg of diet) ranged from
49-59% for dry matter, 79-89% for crude fat, 80-92% for crude protein, 51-67% for gross
energy, 74-97% for amino acids- and 7-99 % for minerals. When DDGS was included at a rate
of 15% of the diet and supplemented with lysine, methionine, and phytase but different levels of
trace mineral premix supplementation, there were no differences in weight gain, feed conversion,
survival, body composition and apparent nutrient retention among fish fed all diets except for
fish fed a diet without trace mineral supplementation. These results suggest that phytase was
effective in releasing most of the minerals and that trace mineral supplementation could be
reduced when phytase is added to rainbow trout diets.
Stone et al. (2005) evaluated the effects of extrusion on nutritional value of diets containing
corn gluten meal and DDGS for rainbow trout and observed that the extent of fish meal
replacement in the diet depends upon the ratio of DDGS to corn gluten meal used. Their results
suggest that up to 18% dietary inclusion of these co-products could replace about 25% of the fish
meal in practical diets without negatively affecting growth performance. They also found that
extrusion of diets containing DDGS and corn gluten meal was of no benefit compared to feeding
cold-pelleted diets.

Freshwater Prawns (Macrobrachium rosenbergii)


A few studies have been conducted on feeding diets containing DDGS to freshwater prawns. In
an initial study, Tidwell et al. (1993a) fed juvenile freshwater prawns (0.66 g) one of three
isonitrogenous diets (29% crude protein) containing 0, 20 or 40% DDGS. There were no
differences among dietary treatments for average yield (833 kg/ha), survival (75%), individual
weight (57 g) and feed conversion (3.1). These results show that levels of up to 40% DDGS can
be included in practical diets for prawns stocked at a density of 19,760/hectare to achieve good
performance.
In a subsequent study, Tidwell et al. (1993b) evaluated the effects of partially replacing fish
meal with soybean meal and DDGS in diets for pond-raised freshwater juvenile prawns (0.51 g).
Three diets were formulated to contain 32% crude protein and contained 15, 7.5 or 0% fish meal.
Fish meal was replaced with a variable percentage of soybean meal and a fixed percentage of
DDGS (40%). There were no differences among dietary treatments for average yield, survival,
individual weight and feed conversion. They noted that replacement of fish meal with soybean
meal and DDGS increased dietary levels of glutamine, praline, alanine, leucine and
phenylalanine and decreases in aspartic acid, glycine, arginine and lysine levels in the diets.
Fatty acid profiles of the diets also changed when soybean meal and DDGS replaced fish meal.
Concentrations of 16:0, 18:2n-6, and 20:1n-9 increased and concentrations of 14:0, 16:1n-7,
18:1n9, 18:3n-3, 20:5n-3, 22:5n-3 and 22:6n-3 decreased. These results suggest that fish meal
can be partially or totally replaced with soybean meal and DDGS in diets for freshwater prawns
raised in ponds in temperate areas. Coyle et al. (1996) showed that juvenile prawn (more than 2

06 - Use of DDGS in Aquaculture Diets

g) showed that DDGS can be consumed directly by prawn, and that DDGS may serve a dual role
as both feed and a pond fertilizer.

Tilapia (Oreochromis niloticus)


Tilapia (Oreochromis niloticus) is a popular warm water fish grown throughout the world. Wu
et al. (1994) reported that diets containing either corn gluten meal (18%) or DDGS (29%) and
32% or 36% crude protein, resulted in higher weight gains for tilapia than fish fed a commercial
fish feed containing 36% crude protein and fish meal for tilapia with initial weight of 30 g.
In a subsequent study, Wu et al., (1996) evaluated the growth responses over an eight week
feeding period of smaller tilapia (0.4 g initial weight) by feeding diets containing up to 49%
DDGS, up to 42% corn gluten feed or up to 22% corn gluten meal, at dietary crude protein levels
of 32%, 36% and 40%. Of the eight diets fed, the highest weight gain was achieved by feeding
the 36% protein commercial control diet (5,320% increase) and the 40% protein diet containing
35% DDGS (5,100% increase). The highest feed conversion was achieved by feeding the control
diet (1.05) and two 40% protein diets containing either 35% DDGS (1.13) or 30% corn gluten
feed (1.12). The highest protein efficiency ratio (weight gain/protein fed) was obtained by
feeding the control diet (3.79) and two 36% protein diets containing 49% DDGS (3.71) or 42%
corn gluten feed (3.55). From these results, these researchers concluded that feeding diets
containing 32%, 36% and 40% protein and 16- 49% protein-rich ethanol co-products will result
in good weight gain, feed conversion and protein efficiency ratio for tilapia fry.
When using DDGS in aquaculture diets, it is also important to know if lower protein diets
containing higher amounts of ethanol co-products (DDGS, corn gluten feed, corn gluten meal)
and synthetic amino acids can support satisfactory growth performance. Wu et al., (1997)
evaluated growth performance of tilapia fry (0.5 g initial weight) over an eight week feeding
period by feeding diets containing 28 or 32% protein, synthetic lysine and tryptophan and 5492% ethanol co-products. There were no statistically significant differences in feed conversion
(1.76 vs. 1.43) and protein efficiency ratio (1.82 vs. 2.21) among fish fed the 28% protein diet
containing 82% DDGS and synthetic lysine and tryptophan or the 67% corn gluten feed and 26%
soy flour, and fish fed the control 32% protein diet (FCR = 1.25, PER = 2.05). Based upon these
results, there is a potential to use DDGS and other ethanol co-products, along with synthetic
amino acid supplementation, to formulate all-plant diets and replace all of the fish meal in
juvenile tilapia.
Tidwell et al. (2000) evaluated growth, survival and body composition of cage-cultured Nile
tilapia fed pelleted and unpelleted DDGS in polyculture with freshwater prawn. Growth rate was
higher for fish fed the pelleted DDGS than for fish fed the unpelleted DDGS. However, fish fed
a commercial catfish diet had greater increases in individual weight, individual length, specific
growth rate and feed conversion compared to fish fed the diets including DDGS either pelleted
or unpelleted. Although growth was higher for fish fed the commercial diet, the cost of
production was significantly higher ($0.66/kg gain) compared to fish fed the unpelleted and
pelleted DDGS diets ($0.26/kg gain and $0.37/kg gain, respectively). Production of prawn was
1,449 kg/ha and adding tilapia in polyculture increased total pond productivity by 81%. These
researchers concluded that feeding DDGS provides economical growth of tilapia and that
polyculture of tilapia may improve overall pond efficiency in freshwater prawn production ponds
in temperate climates.

06 - Use of DDGS in Aquaculture Diets

Conclusions
Based upon recent research studies, maximum dietary inclusion rates of DDGS are shown in
Table 2.
Table 2: Current, Revised Recommendations
for Maximum Dietary Inclusion Rates of DDGS for Various Species of Fish.
Species
% DDGS
Comments
Catfish
Up to 30%
Trout
Up to 15%
Without synthetic lysine and methionine supplementation
Trout
Up to 22.5% With synthetic lysine and methionine supplementation
Salmon
Up to 10%
Freshwater Up to 40%
Can replace some or all of the fish meal in the diet
Prawns
Shrimp
Up to 10%
No studies are available but based upon research results
with freshwater prawns, a minimum of 10% DDGS in
shrimp should be acceptable.
Tilapia
Up to 35%
Without synthetic lysine and supplementation in high
protein diets (40% crude protein)
Tilapia
Up to 82%
With synthetic lysine and tryptophan supplementation in
low protein diets (28% crude protein)

References
Cheng, Z.J., R.W. Hardy, and M. Blair. 2003. Effects of supplementing methionine hydroxyl analogue in soybean
meal and distillers dried grain-based diets on the performance and nutrient retention of rainbow trout
[Oncorhynchus mykiss (Walbaum)]. 2003. Aquaculture Research 34:1303-1310.
Cheng, Z.J. and R.H. Hardy. 2004a. Effects of microbial phytase supplementation in corn distillers dried grains
with solubles on nutrient digestibility and growth performance of rainbow trout (Oncorhynchus mykiss). Journal of
Applied Aquaculture 15:83-100.
Cheng, Z.J. and R.W. Hardy. 2004b. Nutritional value of diets containing distillers dried grain with solubles for
rainbow trout (Oncorhynchus mykiss). Journal of Applied Aquaculture 15:101-113.
Coyle, S., T. Najeeullah, and J. Tidwell. 1996. A preliminary evaluation of naturally occurring organisms, distiller
by-products, and prepared diets as food for juvenile freshwater prawn (Macrobrachium rosenbergii). Journal of
Appled Aquaculture 6:57-66.
Jauncey, K., and B. Ross. 1982. A guide to tilapia feeds and feeding. University of Stirling, Institute for
Aquaculture, Stirling, UK.
Robinson, E.H. 1991. Improvement of cottonseed meal protein with supplemental lysine in feeds for channel
catfish. Journal of Applied Aquaculture 1 (2):1-14.
Shiau, S.Y., J. L. Chuang, and G.L. Sun. 1987. Inclusion of soybean meal in tilapia (Oreochromis niloticus x O.
aureus) diets at two protein levels. Aquaculture 65:251-261.
Stone, D.A.J., R.W. Hardy, F.T. Barrows, and Z.J. Cheng. 2005. Effects of extrusion on nutritional value of diets
containing corn gluten meal and corn distillers dried grain for rainbow trout, Oncorhynchus mykiss. Journal of
Applied Aquaculture 17:1-20.
Tidwell, J.H., C.D. Webster, and D.H. Yancey. 1990. Evaluation of distillers grains with solubles in prepared
channel catfish diets. Transactions of the Kentucky Academy of Science 51:135-138.
Tidwell, J.H., C.D. Webster, J.A. Clark, and L.R. DAbramo. 1993a. Evaluation of distillers dried grains with
solubles as an ingredient in diets for pond culture of the freshwater prawn Macrobrachium rosenbergii. Journal of
the World Aquaculture Society 24:66-70.

06 - Use of DDGS in Aquaculture Diets

Tidwell, J.H., C.D. Webster, D.H. Yancey, and L.R. DAbramo. 1993b. Partial and total replacement of fish meal
with soybean meal and distillers by-products in diets for pond culture of the freshwater prawn (Macrobrachium
rosenbergii). Aquaculture 118:119-130.
Tidwell, J.H., S.D. Coyle, A. VanArnum, C. Weibel, and S. Harkins. 2000. Growth, survival, and body
composition of cage cultured Nile tilapia Oreochromis niloticus fed pelleted and unpelleted distillers grains with
solubles in polyculture with freshwater prawn Macrobrachium rosenbergii. Journal of the World Aquaculture
Society 31:627-631.
Webster, C.D., J.H. Tidwell, and D.H. Yancey. 1991. Evaluation of distillers grains with solubles as a protein
source in diets for channel catfish. Aquaculture 96:179-190.
Webster, C.D., J.H. Tidwell, L.S. Goodgame, and P.B. Johnsen. 1993. Growth, body composition, and
organoleptic evaluation of channel catfish fed diets containing different percentages of distillers grains with
solubles. The Progressive Fish-Culturist 55:95-100.
Weigel, J.C., D. Loy, and L. Kilmer. 1997. Feeding co-products of the dry corn milling process. Renewable Fuels
Association and National Corn Growers Association. Washington, D.C. and St. Louis, MO p. 8.
Wilson, R.P., and W.E. Poe. 1985. Effects of feeding soybean meal with varying trypsin inhibitor activities on
growth of fingerling channel catfish. Aquaculture 46:19-25.
Wu, Y.V., R.R. Rosati, D.J. Sessa, and P.B. Brown. 1994. Utilization of protein-rich ethanol co-products from
corn in tilapia feed. Journal of American Oil Chemists Society 71:1041-1043.
Wu, Y.V., R.R. Rosati, and P.B. Brown. 1996. Effect of diets containing various levels of protein and ethanol
coproducts from corn on growth of tilapia fry. Journal of Agricultural Food Chemistry 44:1491-1493.
Wu, Y.V., R.R. Rosati, and P.B. Brown. 1997. Use of corn-derived ethanol coproducts and synthetic lysine and
tryptophan for growth of tilapia (Oreochromis niloticus) fry. Journal of Agricultural Food Chemistry 45:2174-2177.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be formulated by a
qualified nutritionist. The USGC has no control over the nutritional content of any specific product which may be
selected for feeding. Potential buyers should consult an appropriate nutritionist for specific recommendations.
USGC makes no warranties that these recommendations are suitable for any particular herd or for any particular
animal. The USGC disclaims any liability for itself or its members for any problems encountered in the use of these
recommendations. By reviewing this material, buyers agree to these limitations and waive any claims against USGC
for liability arising out of this material.

06 - Use of DDGS in Aquaculture Diets

Use of DDGS in
C o m pa n i o n A n i m a l D i e ts

User Handbook

Effects of Feeding U.S. DDGS to


Companion Animals
Horses
Very little research has been conducted related to feeding diets containing distillers
dried grains with solubles (DDGS) to horses and other companion animals. Estimations
of digestible energy of distillery co-products range from 11.5 to 14.2 megajoules (MJ)/kg
of dry matter (DLG, 1995) and the relatively high concentration of crude protein in
DDGS allow it to partially replace soybean meal or dried skimmed milk powder in horse
feeds (Frape, 1998). Furthermore, the relatively high oil content in DDGS may be an
important energy source for performance horses (DLG, 1995; Orme et al., 1997).
Leonard et al. (1975) evaluated in vitro cellulose digestibility of dried distillers solubles
(DDS) and DDGS and showed that both ingredients stimulated cellulose digestion. They
also evaluated the effect of adding DDGS to the diet at levels of 0%, 5% and 10%, or
administering half of the DDGS (10% of total diet) directly into the cecum via a fistula,
four hours after each feeding. The cecally-administered DDGS resulted in significantly
higher cellulose digestibility (32.4%) compared to feeding an equivalent amount of
DDGS in the diet (27.2%). Similar trends were found for dry matter and gross energy
digestibility, but there were no significant differences among treatments. In a subsequent
trial, Leonard et al. (1975) fed diets containing corn, bluegrass hay and DDGS at levels of
0%, 9.1% and 18.2%, and found a linear increase in protein digestibility with increasing
dietary levels of DDGS. There were no differences among dietary treatments for dry
matter, cellulose or gross energy digestibility. These authors concluded that DDGS
contains some unidentified factors that stimulate cellulose digestion by the organisms
present in the cecum of the horse and that DDGS can be used effectively in horse diets.
Although horses can utilize the nutrients in DDGS quite well, one of the potential
issues that could limit the use of DDGS in horse feeds is palatability. Equines are very
sensitive to dietary inclusion of novel feed ingredients. Pagan (1991) conducted a series
of feed preference and digestibility trials to determine the suitability of using DDGS as a
feed ingredient for horses. In the feed preference trials, he fed horses pelleted diets
containing 0%, 5%, 10% or 20% DDGS in two tests over six consecutive days. Horses
showed no preference differences between diets containing 0%, 5% or 10% DDGS, and
feeding the 20% DDGS diet was preferred more frequently than pellets containing lower
levels of DDGS. These results suggest that DDGS can be used effectively in pelleted
horse feeds at levels up to 10% of the diet, without any negative effects on palatability,
while increasing the DDGS dietary inclusion level to 20% may actually increase feed
preference. In the digestibility trials, protein digestibility tended to decrease as the level
of DDGS increased in the diet and dry matter digestibility was slightly depressed when
horses were fed the 5% and 10% DDGS diets. However, fat and total digestible nutrients
(TDN) digestibility was no different among diets containing varying levels of DDGS.
Pagan (1991) concluded that DDGS appears to be a suitable ingredient for horses at
inclusion rates up to 20% of the diet.

07 - Use of DDGS in Companion Animal Diets

Hill (2002) evaluated eating behavior and feed intake responses of horses fed various
proportions of wheat distillers grains and concentrate at ratios of 1:0, 0.75:0.25,
0.50:0.50 and 0:1. When wheat distillers grains were offered at a rate of 0.75 of dietary
dry matter and not soaked prior to feeding, there was a significant reduction in the rate of
feed ingestion and the number of chews per kg of dry matter. If the concentrate was
soaked before feeding, there was an increase in the number of feeding bouts when 0.25 of
the concentrate was replaced with wheat distillers grains. However, feed consumption
processes were not affected until 0.5 of the concentrate dry matter was replaced with
wheat distillers grains. Based upon these results, Hill (2002) concluded that wheat
distillers grains can be used as a substitute for other energy and protein ingredients in
horse rations, but the dietary inclusion rate depends on the method of presentation of the
feed to the horse. Soaking of the concentrate before feeding reduced the level of the
distillery co-product that can be incorporated into the ration in order to meet the desired
amount of dry matter intake.

Rabbits
Very little research has been conducted to evaluate the feeding value of DDGS for
rabbits. One study was conducted in Spain where researchers compared the nutrient
digestibility of wheat bran, corn gluten feed and DDGS in New Zealand White x
Californian crossbred rabbits (Villamide et al., 1989). The basal diet contained a low
amount of energy (2,200 kcal/kg dry matter) and a high energy to protein ratio (25 kcal
DE/g digestible protein). Although the dietary fiber content was similar, energy and acid
detergent fiber (ADF) digestibility was highest for rabbits fed the DDGS diet (74.0% and
58.3%, respectively) compared to rabbits fed diets containing wheat bran (59.4% and
9.6%, respectively) and corn gluten feed (65.0% and 27.7%, respectively). Furthermore,
rabbits fed the DDGS diet had the highest level of protein digestibility (70.1%) compared
to rabbits fed the wheat bran (66.6%) and corn gluten feed (61.4%) diets. These results
suggest that DDGS is a suitable ingredient for rabbit diets and provides more digestible
energy, ADF and protein than wheat bran and corn gluten feed.

Dogs and Cats


While there are no published scientific reports on incorporating DDGS into cat foods,
there have been a few studies conducted that DDGS can be effectively used in dry,
extruded dog foods. Studies were conducted at the University of Illinois (Allen et al.,
1981) to evaluate nutrient digestibility of diets containing DDGS for both adult and
immature Pointer dogs. Supplementation of diets with low levels (4 to 8%) of DDGS had
no effect on the apparent digestibility of dry matter or starch by adult dogs. Adding
moderate levels (16.1%) of DDGS to the diet decreased dry matter digestibility, but had
no effect on starch or energy digestibility. Feeding diets containing high levels (26.1%)
of DDGS decreased dry matter and energy digestibility, but had no effect on crude
protein digestibility in adult dogs. Growing puppies that were fed diets containing a
moderate amount (14.1%) of DDGS had lower dry matter and energy digestibility, but
digested more ADF compared to puppies fed diets that contained no DDGS. Nitrogen (N)
intake and fecal N were reduced by dietary DDGS supplementation, but there was no
effect on urinary N, total N excretion, absorbed N or N retention.

07 - Use of DDGS in Companion Animal Diets

Additional research conducted by Corbin (1984) has shown that up to 10% DDGS can
be included in diets for growing puppies to achieve good food intake and body growth
(Table 1). Including DDGS in diets for older, more mature dogs can be very
advantageous for controlling weight gains due to its high fiber content. Weigel et al.
(1997) suggested that diets for mature dogs could include up to 25% DDGS depending on
age and activity level to achieve good intestinal health.
Table 1: Effects of Feeding a Diet Containing 10% DDGS to
Growing Puppies on Food Intake, Weight Gain and Body Length.
0% DDGS
10% DDGS
Number of puppies/treatment
12
12
Initial body weight, kg
3.34
3.42
Final body weight, kg
10.15
10.28
Food intake/10 weeks, kg
21.34
27.96
Weight gain, kg
6.80
6.86
Feed/Gain
3.13
4.07
Increase in body length, cm
22.25
21.97

Conclusions
Based upon the limited research information available, it appears that DDGS is a very
suitable ingredient for use in horse, rabbit and dog diets. Current feeding
recommendations are shown in Table 2.
Table 2: Recommended Maximum Dietary Inclusion Rates
for DDGS in Diets for Horses, Rabbits and Dogs.
Species
Maximum DDGS Inclusion Rate
Horses
Up to 20% of the diet
Rabbits
Up to 20% of the diet
Growing Puppies
Up to 10% of the diet
Adult Dogs
Up to 25% of the diet depending on age and activity level

Literature Cited
Allen, S.E., G.C. Fahey, Jr., J.E. Corbin, J.L. Pugh, and R.A. Franklin. 1981. Evaluation of byproduct
feedstuffs as dietary ingredients for dogs. J. Anim. Sci. 53:1538-1544.
Corbin, J. 1984. Distillers dried grains with solubles for growing puppies. Distillers Feed Conference.
39:28-33.
Deutsche Landwirtschafts Gesellschaft, DLG. 1995. Futtewettabellen Ppferde. 3. Ausgabe DLG,
Frankfurt am Main, Germany.
Frape, D. 1998. Equine Nutrition and Feeding. Blackwell Science, London.
Hill, J. 2001. Effect of level of inclusion and method of presentation of a single distillery by-product on
the processes of ingestion of concentrate feeds by horses. Livestock Production Science 75:209-218.
Leonard, T.M., J.P. Baker, and J. Willard. 1975. Influence of distillers feeds on digestion in the equine. J.
Anim. Sci. 40:1086-1092.
Orme, C.E., R.C. Harris, D. Marlin, and J. Hurley. 1997. Metabolic adaptation to a fat supplemented diet
by the thoroughbred horse. British Journal of Nutrition 78:443-455.

07 - Use of DDGS in Companion Animal Diets

Pagan, J.D. 1991. Distillers dried grains as an ingredient for horse feeds: Palatability and digestibility
study. Distillers Feed Conference. 46:83-86.
Villamide, M.J., J.C. de Blas, and R. Carabano. 1989. Nutritive value of cereal by-products for rabbits. 2.
Wheat bran, corn gluten feed and dried distillers grains and solubles. Journal of Applied Rabbit Research
12:152-155.
Weigel, J.C., D. Loy, and L. Kilmer. 1997. Feeding co-products of the dry corn milling process.
Renewable Fuels Association and National Corn Growers Association. Washington, D.C. and St. Louis,
MO p. 8.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be
formulated by a qualified nutritionist. The USGC has no control over the nutritional content of any specific
product which may be selected for feeding. Potential buyers should consult an appropriate nutritionist for
specific recommendations. USGC makes no warranties that these recommendations are suitable for any
particular herd or for any particular animal. The USGC disclaims any liability for itself or its members for
any problems encountered in the use of these recommendations. By reviewing this material, buyers agree to
these limitations and waive any claims against USGC for liability arising out of this material.

07 - Use of DDGS in Companion Animal Diets

P h ys i c a l & C h e m i c a l
C h a ra c t e r i st i c s o f D D G S

User Handbook

Physical & Chemical Characteristics


of U.S. DDGS
Physical and chemical characteristics of distillers dried grains with solubles (DDGS) vary
among sources and can influence its feeding value and handling characteristics. These
characteristics include color, smell, particle size, bulk density, pH, flowability, shelf life stability
and hygroscopicity.

Color
Color of DDGS can vary from being very light yellow in color to being very dark brown in
color. Differences in color among DDGS sources are influenced by;
the natural color of the feedstock grain being used,
the amount of solubles added to grains before drying,
drying time, and drying temperature.
The color of corn kernels can vary among varieties and has some influence on final DDGS
color. Corn-sorghum blends of DDGS are also somewhat darker in color than corn DDGS
because of the bronze color of many sorghum varieties.
When a relatively high proportion of solubles are added to the mash (grains fraction) to make
DDGS, the color becomes darker. Noll et al. (2006) conducted a study where they evaluated
color in batches of DDGS where approximately 0, 30, 60 and 100% of the maximum possible of
syrup was added to the mash before drying. Actual rates of solubles addition to the mash were 0,
12, 25 and 42 gallons/minute. As shown in Table 1, increasing solubles addition rate to the mash
resulted in a decrease in L* (lightness of color) and b* (yellowness of color), with an increase in
a* (redness of color). Similar results were also reported by Ganesan et al. (2005).
Table 1. The Effect of the Rate of Solubles
Addition to Mash on Color Characteristics of DDGS
0
12
25
42
Pearson
gal/min gal/min gal/min gal/min Correlation
Color (CIE Scale)
59.4
56.8
52.5
46.1
- 0.98
L*
8.0
8.4
9.3
8.8
0.62
a*
43.3
42.1
40.4
35.6
- 0.92
b*

P
Value
0.0001
0.03
0.0001

Adapted from Noll et al. (2006).

Dryer temperatures in dry-grind ethanol plants can range from 127 to 621 C. The amount of
time DDGS spends in the dryer also influences the color. In general, the higher the dryer
temperature and the longer DDGS remains in the dryer, the darker the resulting DDGS will be.

Smell
High quality DDGS has a sweet, fermented smell. DDGS that has a burned or smoky smell has
been overheated.
08 - Physical & Chemical Characteristics of DDGS

Particle Size, Bulk Density and pH


Particle size and particle size uniformity of feed ingredients are important considerations of
livestock and poultry nutritionists when selecting sources and determining the need for further
processing when manufacturing complete feeds or feed supplements. Particle size affects:
1. Nutrient digestibility as particle size is reduced, nutrient digestibility and feed
conversion is improved. This is due to the increasing amount of surface area of an
ingredient that is exposed and available for digestive enzymes to act upon.
2. Mixing efficiency a more uniform particle size in a mixture of ingredients will reduce
mixing time in order to achieve a uniformly distributed mix of ingredients in a complete
feed.
3. Amount of ingredient segregation during transport and handling particle and ingredient
segregation (separation) occurs when particles of different sizes and bulk densities are
blended together and transported or handled.
4. Pellet quality is often defined as the hardness of the pellet and percentage of fines in the
complete feed after pelleting. For corn-soybean meal based diets, a low mean particle
size (400 microns) generally results in a higher quality pellet (less % fines).
5. Bulk density is a measure describing the weight of an ingredient per unit volume. In
general, bulk density can be increased by reducing particle size to increase the weight of
a feed ingredient or complete feed per unit of volume.
6. Palatability and sorting of meal or mash diets depending on the animal, a finely ground,
powdery feed will reduce feed intake and cause bridging in feeders and storage bins.
Extremely coarsely ground feeds can also reduce palatability.
7. Incidence of gastric ulcers in swine, the incidence of gastric ulcers increases as the
mean particle size of the diet is reduced.
Bulk density is an important factor to consider when determining the storage volume of
transport vehicles, vessels, containers, totes and bags. Bulk density affects transport and storage
costs. Low bulk density ingredients have higher cost per unit of weight. It also affects the amount
of ingredient segregation that may occur during handling of complete feeds. High bulk density
particles settle to the bottom of a load during transport, whereas low bulk density particles rise to
the top of a load.
Several unpublished studies have been conducted at the University of Minnesota to compare
particle size and bulk density among DDGS sources. The first study was conducted during the
summer of 2001 where representative samples of DDGS (4.5 kg) were obtained from 16 ethanol
plants in Minnesota, South Dakota and Missouri. From these samples, a 200 gram subsample of
DDGS from each plant was screened through five U.S. sieves and the weight of the DDGS not
filtering through each screen was determined and recorded. The fine particles that filtered
through all screens were collected in the pan and weighed. The U.S. sieve numbers and their
corresponding size of screen openings (microns) were #10, #16, #18, #20 and #30 representing
2000, 1180, 1000, 850 and 600, respectively. The size of DDGS particles collected in the pan
was less than 600 microns. The weights of DDGS collected on each screen were then used to
calculate the percentage of weight of each fraction of the total separated. In addition to
determining the average particle size (geometric mean), the coefficient of variation (CV) and

08 - Physical & Chemical Characteristics of DDGS

standard deviation (SD) of particle size within and among ethanol plants were calculated. These
results are shown in Table 3.
Bulk density (lbs/cubic foot) was determined by filling a one quart container and weighing the
amount of DDGS to fill the container (results shown in Table 3). Samples were also visually
evaluated for color and the presence of syrup balls.
The average particle size among the 16 ethanol plants was 1,282 microns (SD = 305, CV=
24%), and ranged from 612 microns in plant 6 to 2,125 microns in plant 15. Thus, there is
considerable variation in average particle size of DDGS originating from these modern ethanol
plants. As a point of reference, the target mean particle size for meal or mash diets for swine and
poultry is 600-800 microns. Only plants 6 and 7 were close to this target range. All other plants
produced coarser DDGS particles suggesting that further grinding of DDGS may be warranted to
reduce the mean particle size, improve particle size uniformity and optimize nutrient digestibility
of DDGS in a complete mixed feed. Plant 15 had the highest mean particle size (2125 microns).
Ethanol plants that produced DDGS with high amounts of syrup balls tended to have a higher
mean particle size.
Bulk density averaged 35.7 lbs/cubic foot (SD = 2.79, CV = 7.8%), but ranged from 30.8 to
39.3 lbs/cubic foot. However, the correlation between mean particle size and bulk density was
surprisingly very low (r = 0.05) which may be due to the variable amounts of syrup balls among
the samples collected.
Table 3: Mean and Variation of Particle Size
Among Ethanol Plants and Bulk Density of DDGS in 2001
Plant
1
2
3
4
5
6
7
8a
8b
9
10
12
13
14
15
16
Average

Particle Size Mean


1398
1322
1425
1370
1255
612
974
1258
1142
1337
1488
1235
1198
1229
2125
1148
1282.25

Standard Deviation
2.32
2.00
1.62
1.84
1.68
2.75
2.15
1.70
1.84
1.78
1.62
1.75
1.87
2.09
1.56
2.25
1.93

Bulk Density
36.3
39.2
36.8
36.3
33.5
39.3
36.1
33.7
30.8
31.8
38.2
31.4
35.9
39.2
37.6
35.1
35.7

CV %
0.17
0.15
0.11
0.13
0.13
0.45
0.22
0.14
0.16
0.13
0.11
0.14
0.16
0.17
0.07
0.20
0.15

68% of the particles


will fall between:
603
3243
661
2644
880
2309
745
2521
747
2108
223
1683
453
2094
740
2139
621
2101
751
2380
919
2411
706
2161
641
2240
588
2569
1362
3315
510
2583
697
2406

Two additional DDGS nutrient analysis and physical characteristics surveys were conducted by
researchers at the University of Minnesota in 2004 (34 samples from ethanol plants in 11
different states) and 2005 (35 samples). As shown in Tables 4 and 5, average particle ranges

08 - Physical & Chemical Characteristics of DDGS

were 665-737 m, but the range in particle size is extremely large 73 to 1217 m. The pH of
DDGS sources averages 4.1 but can range from 3.6-5.0.
Table 4: Particle Size, Bulk Density, and pH of 34 DDGS Sources Analyzed in 2004.
Average
Range
SD
CV, %
665
256 - 1087
257.48
38.7
Particle size, m
31.2
24.9 35.0
2.43
7.78
Bulk density, lbs/ft3
4.14
3.7 4.6
0.28
6.81
pH
Table 5: Particle Size, Bulk Density, and pH of 35 DDGS sources analyzed in 2005.
Average
Range
SD
CV, %
737
73 1217
283
38.0
Particle size, m
3
25.2
22.8 31.5
8.6
34.2
Bulk density, lbs/ft
4.13
3.6 5.0
0.33
7.91
pH

Flowability
Flowability is defined as the ability of granular solids and powders to flow during discharge
from transportation or storage containments. Flowability is not an inherent natural material
property, but rather a consequence of several interacting properties that simultaneously influence
material flow (Rosentrater, 2006). Flowability may be affected by a number of synergistically
interacting factors including product moisture, particle size distribution, storage temperature,
relative humidity, time, compaction pressure distribution within the product mass, vibrations
during transport and/or variations in the levels of these factors throughout the storage process
(Rosentrater, 2006).Other factors that may affect flowability include chemical constituents,
protein, fat, starch and carbohydrate levels as well as the addition of flow agents.
Under certain conditions, DDGS can exhibit poor flowability (AURI and MN Corn Growers
Assoc., 2005). (National Corn-to-Ethanol Research Center, 2005). Clumping or caking can
occur as a result of loading DDGS into trucks, rail cars or containers if it has not been cooled and
cured properly before loading. This often causes flowability problems and difficulty unloading
DDGS. Reduced flowability and bridging of DDGS in bulk storage containers and transport
vehicles has limited the acceptability of some DDGS sources for some customers and rail car
owners.
Research studies are being conducted to determine the factors that cause flowability problems
and potential solutions to reduce these problems.. The Agricultural Utilization Research Institute
and the Minnesota Corn Growers Association (2005) studied a limited number of DDGS samples
under laboratory conditions. They reported that relative humidity greater than 60% seemed to
reduce flowability of a DDGS sample, which is likely due to the products ability to absorb
moisture. While moisture from both the environment and the DDGS itself likely influence
flowability, many other factors have been suggested as possible controllers of flowability such as
particle size, content of solubles, dryer temperature, moisture content at dryer exit, and others.
Interventions to improve flowability of DDGS have been limited to trial and error approaches
within ethanol plants. Some of these interventions involve regulating the completeness of
fermentation, adjusting moisture content and changing particle size. However, results from these
studies have not been published in the scientific literature. Iowa Limestone Company Resources
(2003) investigated the effectiveness of including 2% calcium carbonate in DDGS as a
08 - Physical & Chemical Characteristics of DDGS

flowability agent. They reported a 6-12% reduction in the angle of repose determined in a
laboratory setting when calcium carbonate was added to DDGS after drying. Determination of
flowability under practical industry conditions was not attempted in their study. Because
moisture and relative humidity seem to play an important role in flowability of DDGS, some
have suggested that the use of zeolites and/or grain conditioners may control moisture migration
in DDGS. However, no controlled studies of this concept have been reported.
Since flow behavior of a feed material is multidimensional, there is no single test that
completely measures the ability of a material to flow (Rosentrater, 2006). Shear testing
equipment is the primary equipment used to measure the strength and flow properties of bulk
materials. They also measure the amount of compaction as well as the bulk strength of materials
(Rosentrater, 2006). Another approach for measuring the flowability of granular materials
involves measuring four main physical properties: angle of repose, compressibility, angle of
spatula, and coefficient of uniformity (e.g. cohesion) (Rosentrater, 2006).

Shelf Life Stability


Since the moisture content of DDGS is usually between 10-12%, there is minimal risk of
spoilage during transit and storage unless water leaks into transit vessels or storage facilities. No
research studies have been conducted to demonstrate that preservatives and mold inhibitors are
necessary to prevent spoilage and extend shelf life of DDGS.
Unless the moisture content of DDGS exceeds 12-13%, the shelf life of DDGS appears to be
many months. In a U.S. Grains Council field trial, DDGS was shipped from an ethanol plant in
South Dakota in a 40 ft. container to Taiwan. Upon arrival in Taiwan, DDGS was put into 50 kg
bags and stored in a covered steel pole barn for 10 weeks during the course of the dairy feeding
trial on a commercial dairy farm located about 20 km south of the Tropic of Cancer.
Environmental temperatures averaged more than 32 C and humidity was in excess of 90%
during this storage period. Samples of DDGS were collected upon arrival at the farm and again
after the 10 week storage period. There was no change in peroxide value (measure of oxidative
rancidity of oil) during this trial. Presumably, this may be due to the high amount of natural
antioxidants present in corn which are further increased during a heating process.

Hygroscopicity
Limited information exists regarding the hygroscopicity (ability to attract moisture) of DDGS.
However, it appears that under humid climatic conditions, DDGS will increase in moisture
content during long-term storage. The U.S. Grains Council sponsored a broiler field trial in
Taiwan, where moisture content of DDGS was monitored during storage at a commercial feed
mill from March 16 to June 10, 2004. A random sample of DDGS was obtained weekly from
storage at the feed mill and analyzed for moisture over a 13-week storage period. Moisture
content of DDGS increased from 9.05% at the beginning of the storage period to 12.26% at the
end of the 13-week storage period (Table 6). As expected, crude protein concentration did not
change in DDGS and no aflatoxin was present initially or at the end of the storage period.

08 - Physical & Chemical Characteristics of DDGS

Table 6: Laboratory analysis results for moisture, crude protein


and aflatoxin of DDGS during storage at the commercial feed mill in Taiwan.
Sample Date Sample Number Moisture, % Crude protein, % Aflatoxin, ppb
16-Mar-04
17-Mar-04
24-Mar-04
31-Mar-04
7-Apr-04
14-Apr-04
21-Apr-04
28-Apr-04
5-May-04
12-May-04
19-May-04
27-May-04
3-Jun-04
10-Jun-04

1
2
3
4
5
6
7
8
9
10
11
12

9.05
10.17
10.65
10.70
10.71
10.76
10.93
11.02
11.28
11.16
11.70
11.88
12.13
12.26

27.60
27.61
27.59
27.63
27.62
27.73
27.71
27.62
27.54
27.61
27.63
27.61
27.50
27.53

0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00

References
Agricultural Utilization Research Institute (AURI), and Minnesota Corn Growers Association. 2005. Distillers
Dried Grains Flowability Report. Waseca, MN.
Cromwell, G.L., K.L. Herkleman, and T.S. Stahly. 1993. Physical, chemical, and nutritional characteristics of
distillers dried grains with solubles for chicks and pigs. J. Anim. Sci. 71:679-686.
Ferrer, E., A. Algria, Farre, G. Clemente, and C. Calvo. 2005. Fluorescence, browning index, and color in infant
formulas during storage. J. Agric. Food Chem. 53:4911-4917.
Noll, S., C. Parsons, and B. Walters. 2006. Whats new since September 2005 in feeding distillers co-products to
poultry. Proceedings from the 67th Minnesota Nutrition Conference & University of Minnesota Research Update
Session: Livestock Production in the New Millenium, St. Paul, MN. pp. 149-154.
Ganesan, V. K.A. Rosentrater, and K. Muthukumarappan. 2005. Effect of moisture content and soluble levels on
the physical and chemical properties of DDGS. ASAE paper No. 056110. St. Joseph, MI.
ILC Resources. 2003. CaCO3 treatment of DDGS. In house study provided by R.H. Bristol.
National Corn to Ethanol Research Center (NCERC). 2005. Website at: www.ethanolresearch.com/. Accessed
June 13, 2006.
Pederson, C., A. Pahm, and H.H. Stein. 2005. Effectiveness of in vitro procedures to estimate CP and amino acid
digestibility coefficients in dried distillers grain with solubles by growing pigs. J. Anim. Sci. (Suppl. 2) 83:39.
Rosentrater, K.A. 2006. Understanding Distillers grain Storage, Handling, and Flowability Challenges. Distillers
Grains Quarterly. First Quarter 2006. pp. 18-21.
Shurson, J., S. Noll, and J. Goihl. 2005. Corn by-product diversity and feeding value to non-ruminants. Proc. MN
Nutr. Conf. pg. 50 68.
University of Minnesota. 2005. DDGS website at: www.ddgs.umn.edu. Accessed October 3, 2006.

08 - Physical & Chemical Characteristics of DDGS

The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be formulated by a
qualified nutritionist. The USGC has no control over the nutritional content of any specific product which may be
selected for feeding. Potential buyers should consult an appropriate nutritionist for specific recommendations.
USGC makes no warranties that these recommendations are suitable for any particular herd or for any particular
animal. The USGC disclaims any liability for itself or its members for any problems encountered in the use of these
recommendations. By reviewing this material, buyers agree to these limitations and waive any claims against USGC
for liability arising out of this material.

08 - Physical & Chemical Characteristics of DDGS

Nutrient Composition
of DDGS

User Handbook

Nutrient Content of DDGS:


Variability and Measurement
Comparison of nutrient composition of distillers co-products
Fuel ethanol is produced by dry grind ethanol plants as well as wet mills in the United States.
Corn is the primary grain used in both wet mills and dry-grind ethanol plants because of its high
fermentable starch content compared to other feedstocks. Some ethanol plants use sorghum or
blend corn with barley, wheat or sorghum to make ethanol, depending on geographical location,
cost and availability of these grains relative to corn.
While dry-mill ethanol plants produce a variety of co-products (see section 11), distillers dried
grains with solubles (DDGS) is the most important co-product being marketed internationally for
use in dairy, beef, swine, poultry and aquaculture feeds. Nutrient values of high quality DDGS
produced by modern ethanol plants in the United States are generally higher than those published
in the swine NRC (1998), poultry NRC (1994), dairy NRC (2001) and beef NRC (1996) Nutrient
Requirement publications.
Wet mills produce corn gluten feed, corn gluten meal and corn germ meal.
The beverage alcohol industry also produces DDGS (less than 1 % of total DDGS production),
but it is often dark in color, tends to be more variable in nutrient content and has lower levels of
digestible nutrients than DDGS from modern fuel ethanol plants. Brewers dried grains is a coproduct of the beer manufacturing industry and consist of the dried residue of barley malt and
other grains that have been used to provide maltose and dextrins for fermenting. The high fiber
level (18-19%) of brewers dried grains limits its use in some animal diets.
The co-products of dry mill, wet mill and the beverage alcohol industries are nutritionally
different from each other and have different economic value in various types of animal and
poultry feeds. Table 1 shows a comparison of the nutrient content of high quality DDGS,
compared to published values in the Swine National Research Council (NRC, 1998) for DDGS,
corn gluten feed, corn gluten meal and brewers dried grains.
The primary nutritional advantages of high quality DDGS compared to corn gluten feed, corn
gluten meal and brewers dried grains are the high oil and available phosphorus content. The
digestible energy (DE) and metabolizable (ME) values of high quality DDGS are significantly
higher than corn gluten feed and brewers dried grains, but less than corn gluten meal. Amino
acid levels of DDGS are lower than corn gluten meal, but comparable to corn gluten feed and
brewers dried grains.

09 Nutrient Composition of DDGS: Variability & Measurement

Table 1: Comparison of the nutrient composition (as fed basis)


between high quality DDGS and DDGS, corn gluten feed,
corn gluten meal and brewers dried grains (NRC, 1998).
Brewers
Corn
High
DDGS
Corn
Dried Grains
Gluten
Quality
NRC
Gluten
NRC (1998)
Meal
(1998)
Feed
Corn
NRC (1998) NRC (1998)
DDGS
Dry Matter, %
89
93
90
90
92
Crude Protein, %
27.2
27.7
21.5
60.2
26.5
Crude Fat, %
9.5
8.4
3.0
2.9
7.3
Acid Detergent Fiber
14.0
16.3
10.7
4.6
21.9
(ADF), %
Neutral Detergent
38.8
34.6
33.3
8.7
48.7
Fiber (NDF), %
DE, kcal/kg
3953
3200
2990
4225
2100
ME, kcal/kg
3580
2820
2605
3830
1960
Arginine, %
1.06
1.13
1.04
1.93
1.53
Histidine, %
0.68
0.69
0.67
1.28
0.53
Isoleucine, %
1.01
1.03
0.66
2.48
1.02
Leucine, %
3.18
2.57
1.96
10.19
2.08
Lysine, %
0.74
0.62
0.63
1.02
1.08
Methionine, %
0.49
0.50
0.35
1.43
0.45
Cystine, %
0.52
0.52
0.46
1.09
0.49
Phenylalanine, %
1.32
1.34
0.76
3.84
1.22
Threonine, %
1.01
0.94
0.74
2.08
0.95
Tryptophan, %
0.21
0.25
0.07
0.31
0.26
Valine, %
1.34
1.30
1.01
2.79
1.26
Calcium, %
0.05
0.20
0.22
0.05
0.32
Chlorine, %
no data
0.08
0.22
0.06
0.15
Magnesium, %
0.13
0.19
0.33
0.08
0.16
Phosphorus, %
0.79
0.77
0.83
0.44
0.56
Avail. Phosphorus, %
0.71
0.59
0.49
0.07
0.19
Potassium, %
0.84
0.84
0.98
0.18
0.08
Sodium, %
0.22
0.20
0.15
0.02
0.26
Sulfur, %
0.44
0.30
0.22
0.43
0.31
Copper, mg/kg
6
57
48
26
21
Iron, mg/kg
121
257
460
282
250
Manganese, mg/kg
13
24
24
4
38
Selenium, mg/kg
no data
0.39
0.27
1.00
0.70
Zinc, mg/kg
75
80
70
33
62
Beta carotene, mg/kg
no data
3.5
1.0
-0.2
Vitamin E, mg/kg
no data
no data
8.5
6.7
-Niacin, mg/kg*
no data
75
66
55
43
Pantothenic acid,
no data
14.0
17.0
3.5
8.0
mg/kg

09 Nutrient Composition of DDGS: Variability & Measurement

Riboflavin, mg/kg
Vitamin B12, mg/kg
Biotin, mg/kg
Choline, mg/kg
Folacin, mg/kg
Thiamin, mg/kg
Vitamin B6, mg/kg

no data
no data
no data
no data
no data
no data
no data

8.6
0.0
0.78
2637
0.90
2.9
8.0

2.4
0.0
0.14
1518
0.28
2.0
13.0

2.2
0.0
0.15
330
0.13
0.3
6.9

1.4
0.0
0.24
1723
7.10
0.6
0.7

High Protein DDGS and Other New Types of Ethanol Co-Products


Several ethanol production companies and other research groups have developed a variety of
modified processes to enhance ethanol yield and change the resulting co-products from dry grind
ethanol plants. The most widely discussed processes involve using new enzyme technology to
increase the crude protein content of DDGS, removing the germ and/or bran from corn prior to
fermentation and removing the phosphorus prior to producing DDGS. We may expect to see a
diversity of new, improved DDGS and other dry-mill co-products as the ethanol industry
continues to expand. Each new product will have its specific nutritional profile, and must be
evaluated carefully to establish its appropriate place in animal rations.

Factors Affecting Variability in Nutrient Content and Digestibility of


DDGS Sources
Nutritionists want consistency and predictability in the feed ingredients they purchase and use.
The nutrient content of DDGS can vary among DDGS sources (Table 2), and has been shown to
vary over time within plants (Spiehs et al., 2002).
Table 2: Averages and Ranges in Composition of Selected Nutrients
(100% Dry Matter Basis) Among 32 U.S. DDGS Sources (www.ddgs.umn.edu)
Nutrient
Average (CV)
Range
Crude protein, %
30.9 (4.7)
28.7 - 32.9
Crude fat, %
10.7 (16.4)
8.8 - 12.4
Crude fiber, %
7.2 (18.0)
5.4 - 10.4
Ash, %
6.0 (26.6)
3.0 - 9.8
Calculated ME (swine), kcal/kg
3810 (3.5)
3504 4048
Lysine, %
0.90 (11.4)
0.61 - 1.06
Arginine, %
1.31 (7.4)
1.01 - 1.48
Tryptophan, %
0.24 (13.7)
0.18 - 0.28
Methionine, %
0.65 (8.4)
0.54 - 0.76
Phosphorus, %
0.75 (19.4)
0.42 - 0.99
Olentine (1986) listed a number of variables in the raw materials and processing factors that
contribute to variation in nutrient composition of distillers co-products (Table 3).

09 Nutrient Composition of DDGS: Variability & Measurement

Table 3: Factors influencing nutrient composition


of distillers co-products. (Olentine, 1986)
Raw Materials
Fermentation
Types of grains
Yeast quality and quantity
Grain variety
Temperature
Grain quality
Time
Soil conditions
Cooling
Fertilizer
Agitation
Weather
Acidity and production control
Production and harvesting methods
Distillation
Grain formula
Type: vacuum or atmospheric, continuous
or batch
Processing Factors
Direct or indirect heating
Grind Procedure
Change in volume during distillation
Fineness
Processing
Duration
Type of screen: stationary, rotating, or
Cooking
vibratory
Amount of water
Use of centrifuges
Amount of pre-malt
Type of presses
Temperature and time
Evaporators
Continuous or batch fermentation
Temperature
Cooling time
Number
Conversion
Dryers
Type, quantity and quality of malt
Time
Fungal amylase
Temperature
Time and temperature
Type
Dilution of converted grains
Amount of syrup mixed with grain
Volume and gallon per bushel or
grain bill
Quality and quantity of grain
products

To manage the diversity among DDGS sources, some commercial feed manufacturers require
identity preservation of selected DDGS sources and use a preferred suppliers list when
purchasing DDGS.
The three most important factors affecting variability of nutrient content in DDGS are:
variation in the nutrient content of the corn delivered to the ethanol plant,
variations in the mixture ratio of the two components of DDGS in the plant, and
differences in drying time and temperatures .

09 Nutrient Composition of DDGS: Variability & Measurement

Variation in nutrient content of corn


Much of the variation in nutrient content of DDGS is likely due to the normal variation among
corn varieties and geographic location where corn is grown. Reese and Lewis (1989) showed that
corn produced in Nebraska in 1987 ranged between 7.8-10.0% crude protein, 0.22-0.32 % lysine,
and 0.24-0.34% phosphorus (Table 4).
Table 4: Overall Average, Minimum, and Maximum Values for Nutrients in Corna
Nutrient
Average Minimum Maximum
Crude protein, %
8.6
7.8
10.0
Lysine, %
0.26
0.22
0.32
Calcium, %
0.01
0.01
0.01
Phosphorus, %
0.28
0.24
0.34
Selenium, parts per million (ppm)
0.12
0.10
0.16
Vitamin E, IU/lb
3.9
1.9
5.8
a

88% dry-matter basis

In the dry mill plants as the starch is removed from corn kernels to produce ethanol, the nutrients
in the DDCS co-product become concentrated, and the variability of nutrients in the corn will be
more pronounced in the DDGS.

Variation in the rate of solubles added to grains


DDGS is produced in a dry mill ethanol plant by blending two output streams from the ethanol
production process: condensed distillers solubles and the grains fraction (see a complete
explanation of this process in Section 11). The official definition of DDGS published by the
Association of American Feed Control Officials (AAFCO) requires that at least 75% of the
solids in the whole stillage be mixed with the wet cake from an ethanol plant. Ethanol plants
may vary the amount of solubles in the blend above the 75% minimum. Since the typical nutrient
content of each output stream is different (see Table 5), variation from plant to plant in the ratio
of blending the two components of DDGS will affect nutrient composition of the DDGS.
Table 5: Nutrient Content and Variability of Distillers Grains
and Distillers Solubles on (100% Dry Matter Basis).
Grains Fraction
Average
Minimum
Maximum
Dry Matter, %
34.3
33.7
34.9
Crude Protein, %
33.8
31.3
36.0
Crude Fat, %
7.7
2.1
10.1
Crude Fiber, %
9.1
8.2
9.9
Ash, %
3.0
2.6
3.3
Calcium, %
0.04
0.03
0.05
Phosphorus, %
0.56
0.44
0.69

09 Nutrient Composition of DDGS: Variability & Measurement

Solubles Fraction
Dry Matter, %
Crude Protein, %
Crude Fat, %
Crude Fiber, %
Ash, %
Calcium, %
Phosphorus, %

27.7
19.5
17.4
1.4
8.4
0.09
1.3

23.7
17.9
14.4
1.1
7.8
0.06
1.2

30.5
20.8
20.1
1.8
9.1
0.12
1.4

Source: Knott et al. (2004)

Noll et al. (2006) evaluated the nutrient composition and digestibility of batches of DDGS
produced with varying levels of solubles added to the wet grains. The DDGS samples produced
contained solubles added at approximately 0, 30, 60 and 100% of the maximum possible addition
of solubles to the grains. This corresponds to adding 0, 12, 25 and 42 gallons of syrup to the
grains fraction per minute. Dryer temperatures decreased as the rate of solubles addition to the
grains decreased. DDGS samples were analyzed for color, particle size, moisture, crude fat,
crude protein, crude fiber, ash, phosphorus, lysine, methionine, cystine and threonine. Digestible
amino acids were determined using cecectomized roosters and true metabolizable energy
(TMEn) was determined using intact young turkeys.
Particle size increased and was more variable as increasing additions of solubles were added to
the grains fraction. As shown in Table 6, adding increasing amounts of solubles resulted in
darker colored DDGS (reduced L*) and less yellow color (reduced b*). Increased addition of
solubles resulted in increased crude fat, ash, TMEn (poultry), magnesium, sodium, phosphorus,
potassium, chloride and sulfur, but had minimal effects on crude protein and amino acid content
and digestibility.
Table 6: Effects of increasing solubles addition to the grains fraction
during DDGS production on color, nutrient content, TMEn (poultry)
and amino acid digestibility (100% dry matter basis).
P value
Measurement
0
12
25
42
Pearson
gal/min gal/min gal/min gal/min correlation
coefficient
Color L*
59.4
56.8
52.5
46.1
- 0.98
0.0001
Color a*
8.0
8.4
9.3
8.8
0.62
0.03
Color b*
43.3
42.1
40.4
35.6
- 0.92
0.0001
Moisture, %
9.52
9.75
10.74
13.83
0.93
0.06
Crude fat, %
7.97
9.14
9.22
10.53
0.96
0.04
Crude protein, %
31.96
32.65
32.46
31.98
0.03
NS
Crude fiber, %
9.17
7.76
10.08
6.50
- 0.51
NS
Ash, %
2.58
3.58
3.72
4.62
0.97
0.03
Lysine, %
1.04
1.05
1.09
1.04
0.02
NS
Methionine, %
0.63
0.64
0.59
0.62
- 0.13
NS
Cystine, %
0.61
0.61
0.53
0.62
0.16
NS
Threonine, %
1.20
1.22
1.20
1.20
- 0.18
NS
Phosphorus, %
0.53
0.66
0.77
0.91
0.99
0.002
09 Nutrient Composition of DDGS: Variability & Measurement

TMEn, kcal/kg
Lys digestibility, %
Met digestibility, %
Cys digestibility, %
Thr digestibility, %
Arg digestibility, %

2712
78.2
90.9
87.2
85.9
92.1

2897
76.0
88.6
87.6
83.2
90.7

3002
69.7
86.3
80.7
80.5
86.7

3743
75.0
87.3
80.3
77.3
88.5

0.94
- 0.90
- 0.92
- 0.95
- 0.99
- 0.99

0.06
NS
NS
NS
0.02
0.07

Lysine Digestibility: Drying Time and Temperatures


Among sources of light colored DDGS, Ergul et al., (2003) showed that true lysine
digestibility coefficients ranged between 59-83 % for poultry and Stein et al. (2005) showed a
similar range in true lysine digestibility coefficients for swine (44 to 63%).
Results from a recent, unpublished, collaborative study involving swine nutrition researchers at
South Dakota State University, University of Minnesota and Degussa have shown that the lysine
content ranged from 0.52-1.13% and the Standardized True Ileal lysine digestibility values
ranged from 17.7%-74.4% among 34 different DDGS sources. Because of the variation in amino
acid digestibility among DDGS sources, and the need to know amino acid digestibility for swine
among DDGS sources, research is underway at the University of Minnesota to evaluate the
accuracy of using several in vitro laboratory procedures to predict amino acid digestibility of
DDGS sources before they are used to formulate and manufacture swine diets.
It is likely that much of the difference in lysine digestibility among light colored DDGS
sources is due to drying time and temperature used to produce DDGS. Dryer temperatures can
range from 260 to 1150 F, depending on the ethanol plant (Table 7).
Table 7: Dryer temperatures used by selected dry-grind
ethanol plants in the United States to produce DDGS
Plant
Dryer (1) Temp (F)
Dryer (2) Temp (F)
A
700-800
750-800
B
1050-1100
C
590
D
1150
E
445
F
960
497
G
791
595
H
837
770
I
850
260
J
550-700
K
875
640
L
1100
M
1000
N
900
930
O
950
P
940
860-880

09 Nutrient Composition of DDGS: Variability & Measurement

Since amount and length of heating is highly correlated to lysine digestibility (Figure 1), it is
not surprising that a fairly wide range in lysine digestibility exists among light colored DDGS
sources.
Some dry-mill ethanol plants use process modifications to produce ethanol and DDGS. For
example, some plants use cookers to add heat for fermentation and use less enzymes, while other
plants will use more enzymes and do not rely on the use of cookers to facilitate fermentation.
Theoretically, use of less heat could improve amino acid digestibility of DDGS, but no studies
have been conducted to determine how these processes impact final nutrient composition and
digestibility.

To reduce the risk of under or over-valuing DDGS in animal feeds, DDGS customers must
select their sources based on availability of information on current and complete nutrient
profiles, and, ideally, have a rigorous DDGS quality assurance program. In addition, ethanol
plants should develop a database of sample analyses to show how consistent the co-products
being produced are over time. Accurate, fast, and inexpensive in vitro procedures need to be
identified and/or developed to estimate amino acid digestibility among DDGS sources for swine
and poultry.

Assessment of Nutrient Content and Digestibility in DDGS


Perhaps the biggest challenge of using distillers dried grains with solubles (DDGS) in animal
feeds is to know the nutrient content and digestibility of the source being used.
Until more accurate in vitro procedures can be developed, color measurement with
Minolta or Hunter lab spectrophotometers appears to be the most consistent predictor of
lysine digestibility among DDGS sources.
Use of Acid Detergent Insoluble Nitrogen (ADIN) as a predictor of protein and amino
acid digestibility in DDGS is not as accurate as it is for heat damaged forages.

09 Nutrient Composition of DDGS: Variability & Measurement

Although the use of enzyme assays such as IDEA and pepsin/pancreatin and reactive
lysine procedures are promising in vitro procedures for predicting digestible crude
protein and amino acid content, more refinements are need to improve their accuracy.
Calibrations for amino acids and energy in DDGS can be developed using NIRS, but the
quality of these calibrations is dependent on the calibration method used. Overall, the
quality of these calibrations is reasonable, but lower than that obtained from other feed
ingredients.

Color
The color of distillers dried grains with solubles (DDGS) can range from being very light
golden in color to being dark brown in color (Figure 1). Color differences are due to initial grain
color, the amount of solubles added to the grains to make DDGS and the drying time and
temperature used.

Figure 1: Differences in color among DDGS sources.


Noll et al. (2006) showed that the amount of solubles added to the grains affects color as
shown in Table 1. Ganesan et al. (2005) also showed that lightness (L*) of DDGS color
decreased with increased soluble percentage addition (r2 = 0.76), while redness (a*) and
yellowness (b*) values increased with increased soluble percentage addition (r2 = 0.63 and 0.72,
respectively).

09 Nutrient Composition of DDGS: Variability & Measurement

Table 1: Effects of increasing solubles addition to the grains fraction during DDGS
production on color, nutrient content, total mechanical energy (TME)n (poultry) and
amino acid digestibility (100% dry matter basis).
Measurement
0
12
25
42
Pearson
P
gal/min gal/min gal/min gal/min
correlation
value
coefficient
Color L*
59.4
56.8
52.5
46.1
- 0.98
0.0001
Color a*
8.0
8.4
9.3
8.8
0.62
0.03
Color b*
43.3
42.1
40.4
35.6
- 0.92
0.0001
The lightness of color of DDGS samples appears to be moderately correlated with total lysine
content, where lighter colored samples tended to have more total lysine and higher lysine
digestibility. The amount and length of heating is highly correlated to color and lysine
digestibility and due to the wide range in dryer temperatures, there is a wide range in lysine
digestibility among DDGS sources. Lower amino acid digestibility and can lead to reduced
growth performance when fed to swine and poultry.
When heat is applied to feed ingredients, a browning or Maillard reaction occurs resulting in
the formation of high molecular weight polymeric compounds known as melanoidins. The
degree of browning (measured via absorbance at 420 nm) is used to assess the extent the
Maillard reaction has taken place in foods. Digestibility of lysine is the most affected by the
extent of the Maillard reaction. Lightness and yellowness of DDGS color appear to be reasonable
predictors of digestible lysine content among light colored DDGS sources for poultry (Figure 1;
Ergul et al., 2003) and swine (Cromwell et al., 1993; Pederson et al., 2005). However, among
sources of light colored DDGS, Ergul et al., (2003) showed that true lysine digestibility
coefficients ranged from 59-83 % for poultry and Stein et al. (2005) showed a similar range in
true lysine digestibility coefficients for swine (44-63%). Cromwell et al. (1993) evaluated the
relationship between Hunter Lab color scores of various sources of DDGS and this relationship
to acid detergent insoluble nitrogen (ADIN) and growth performance of pigs (Table 2).
Hunter and Minolta colorimeters have been used in human food industry to measure color.
Color is defined by Commission Internationale dEclairage, in Vienna, Austria, as lightness or
L* (0 dark, 100 light), a* for yellowness-redness and b* for blueness-greenness. These
colorimeters have not been widely used in the animal feed industry as methods to predict amino
acid concentration and digestibility in feed ingredients. However, with the variation in color
among DDGS sources and the relatively high correlation between lightness of color (L*) and
yellowness of color (b*) and high lysine digestibility, Hunter and Minolta colorimeters are now
being used to assess DDGS quality.
Cromwell et al. (1993) were the first to demonstrate that DDGS color is correlated to pig and
chick growth performance. In that study, six DDGS sources were combined to mix three
different growing pig diets containing DDGS of light, medium and dark color. As shown in
Table 2, pigs fed the diet made from the combination of dark DDGS sources (A and E) had
slower growth, reduced feed intake and reduced feed efficiency compared to pigs fed the diet
containing the light colored DDGS (B and D). The researchers concluded that the darker the
DDGS, the poorer the growth performance of the pigs.

09 Nutrient Composition of DDGS: Variability & Measurement

10

Table 2: Effect of acid detergent insoluble nitrogen and color score on growth performance
of pigs fed three blended sources of DDGS
Hunter Lab Color3
Source
ADIN %
ADG, g2 ADFI, g2
F/G2
L*
a*
b*
A
29.0
6.5
12.7
27.1
218
1,103
5.05
E
31.1
6.1
13.1
36.9
G
38.8
6.8
16.5
16.0
291
1,312
4.52
I
41.8
6.5
18.8
26.4
B
53.2
4.7
21.8
8.8
390
1,416
3.61
D
51.7
7.1
24.1
12.0
1

Modifed form Cromwell et al., 1993.


Difference among diets P < 0.01
3
L = lightness (0 = black; 100 = white). The higher the a* and b* values, the greater the degree of
redness and yellowness, respectively.
2

Figure 2 was developed (Urriola, 2006) using data from Cromwell et al. (1993), which includes
amino acid composition of nine DDGS sources and their color readings. There was a moderate
correlation (r2 = 0.47) between lightness of DDGS color and lysine content, where lighter colored
samples tended to have more total lysine. Other researchers have shown similar relationships
between extreme heat damage of feed ingredients (dark color) and reduced total lysine
concentration (Finot 2005; Friedman 1992, van Barneveld et al., 1994a).
Figure 2: Correlation between lightness of DDGS color (L*) and amino acid composition of
9 DDGS sources (adapted from Cromwell et al. (1993).

1
0.9

Lysine = 0.0092x + 0.359


2
R = 0.47

0.8
0.7
0.6
0.5
0.4

Methionine = 0.0003x + 0.502


2
R = 0.01

0.3
0.2

Tryptophan = 0.0005x + 0.178


2
R = 0.06

0.1
0
20

25

30

35
40
45
Hunter Lab lightness (L*)

50

09 Nutrient Composition of DDGS: Variability & Measurement

55

11

Other researchers have also shown that DDGS color and digestible amino acid content are
correlated (Ergul et al., 2003; Fastinger et al., 2006; Batal and Dale, 2006). Ergul et al. (2003)
collected 22 DDGS samples from four different ethanol plants that utilized modern ethanol
production processes to determine digestible amino acids using cecectomized roosters. Color
(L*, a* and b*) was measured in each sample using a Minolta colorimeter and found that
lightness and yellowness of color of DDGS samples were correlated to digestible lysine content
(r2 = 0.67 and 0.77 for L* and b*, respectively). Fastinger et al. (2006) also used cecectomized
roosters to determine the correlation of lightness of DDGS color (L*) of five DDGS sources with
lysine digestibility and reported a correlation of r2 = 0.86. Similarly, Batal and Dale (2006) found
a high correlation (r2 = 0.74) between lightness of color and digestible lysine content for poultry
among DDGS sources.

Acid Detergent Insoluble Nitrogen (ADIN)


The Cornell Net Carbohydrate Protein Model used in ruminant nutrition separates dietary
nitrogen into four fractions (Licitra et al., 1996). Fraction A represents non-protein nitrogen;
fraction B represents true protein nitrogen and is divided into sub-fractions B1, B2 and B3 based
on the degree of solubility; and fraction C is acid detergent insoluble nitrogen (ADIN). ADIN is
the nitrogen residue leftover in acid detergent fiber (ADF). Ruminant nutritionists have been
interested in using ADIN as an indicator of digestibility and heat damage in forages and nonforage proteins for many years, with high correlations between ADIN and nitrogen digestibility
in heat damaged forages.
However, the accuracy of using of ADIN to predict heat damage in non-forage proteins has
appears to be less than for forages. Klopfenstein and Britton (1987) showed that ADIN is poorly
correlated to nitrogen digestibility (r = -0.27) in DDGS. Akayezu et al. (1998), showed that the
relationship between ADIN and protein digestibility is not constant across ADIN values and that
the best correlations between ADIN and intestinal availability of rumen undegradable protein
(RUP) are obtained when ADIN values are higher than 13% of total nitrogen. Van Soest and
Mason (1991) suggested that low correlations between ADIN and nitrogen digestibility observed
are due to the confounding effects of endogenous nitrogen losses. When data are mathematically
corrected for endogenous nitrogen losses, the correlations increase (r = 0.84).
The ADIN relationship to amino acid digestibility has been less studied in monogastric animals
compared to ruminants. Cromwell et al., (1993) analyzed ADIN in nine DDGS sources and
found that it ranged between 8.8-36.9% and darkness of color of DDGS samples increased as
ADIN values increased (r2 = 0.62). They also found that chickens fed diets low in ADIN content
had better growth performance than chickens fed diets with higher amounts of ADIN. Chicken
growth performance and ADIN were highly correlated (r2 = 0.86 and 0.72 for weight gain and
feed/gain, respectively).

Near Infrared Spectrophotomtery (NIRS)


No studies have been published regarding the feasibility of using near infrared
spectrophotometry (NIRS) on predicting amino acid and energy content of DDGS. However,
there is tremendous interest in using this technology in the ethanol and feed industry if reliable
calibrations can be achieved. In an unpublished study, researchers at the University of Minnesota
and North Carolina State University collaborated to evaluate the feasibility of making NIRS

09 Nutrient Composition of DDGS: Variability & Measurement

12

calibrations for amino acids and energy for DDGS. Samples of DDGS (n = 103) that were
chemically analyzed by Spiehs et al. (2002) were used for NIRS calibrations at North Carolina
State University. Samples were ground using a Retsch grinder through a 0.5 mm screen. Aliquots
were analyzed (in duplicate) for gross energy using an IKA model C5000 bomb calorimeter.
Ground samples were analyzed using a NIR Systems model 6500 spectrophotometer using a
half-size rectangular cup. Scans were obtained from 400 to 2,500 nm. Spectral data were
derivatized to the second order and smoothed, and calibrations were developed using partial least
squares (PLS) regression with cross-validation (20 segments) using The Unscrambler after
removal of outliers.
Lysine, methionine, threonine and energy calibrations were developed using PLS1 (Table 3).
With this method, calibrations were developed for individual parameters. The calibrations
obtained were reasonable, with over 75% of the variation explained for lysine and energy, and 53
and 66% of the variation for threonine and methionine, respectively. The disappointing
calibrations for threonine was likely a result of the low variation (CV = 6.2%) within the DDGS
samples. A better calibration was expected for methionine because of the large variation among
DDGS samples and because the calibrations for methionine in feedstuffs such as meat and bone
meal are very successful. A reasonable calibration for energy was developed even though there
was low variation in energy content between the DDGS samples (CV = 1.9%).
Table 3: Calibration statistics obtained using PLS1.
Ra
Rmsepb, % R2
CV, %
Lysine
0.89
0.064
0.79
16.2
Methionine 0.81
0.044
0.66
14.2
Threonine
0.73
0.046
0.53
6.2
Energy

0.87

37

0.76

1.9

R is the correlation between actual and predicted values.


b
Rmsep is the prediction error.

The data were also used to make a PLS2-type calibration (Table 4). With this type of
calibration, multiple dependent variables (e.g. all amino acids) were calibrated at the same time.
The PLS2 method of calibration is a faster method of calibrating and the calibration developed
can take the interdependence of the variables into consideration, therefore resulting in
calibrations that are more biologically relevant and robust. A disadvantage of PLS2 is that the
calibration is not optimized for individual parameters because outliers for a specific parameter
are not removed. For amino acids, this may be especially important because three different
assays are used. As shown in Table 4, the calibration results obtained with PLS1 are better than
the calibrations obtained with PLS2. This difference was much greater for lysine and threonine
and minimal for methionine.
Table 4: Calibration statistics obtained using PLS2.
Rmsepb, %
R2
CV, %
Ra
Thr
0.61
0.050
0.37
6.2
Cys
0.74
0.035
0.55
9.4
Val
0.65
0.078
0.43
7.3
Met
0.80
0.046
0.64
14.2
09 Nutrient Composition of DDGS: Variability & Measurement

13

Ile
Leu
Phe
His
Lys
Arg
Trp
TAAc

0.71
0.84
0.82
0.76
0.73
0.75
0.60
0.79

0.065
0.125
0.052
0.036
0.089
0.065
0.017
1.008

0.50
0.70
0.68
0.58
0.53
0.56
0.36
0.63

8.5
6.5
6.5
7.8
16.2
8.7
9.1
6.6

R is the correlation between actual and predicted values.


Rmsep is the prediction error.
c
TAA = total amino acids
b

These results suggest that calibrations for amino acids and energy in DDGS can be developed
using NIRS. The quality of these calibrations is dependent on the calibration method used, with
PLS1 calibrations preferred over PLS2 calibrations. Overall, the quality of these calibrations was
reasonable, but lower than that obtained from other feed ingredients such as meat and bone meal,
which is also a very heterogeneous and heat processed material. This may be due to DDGS assay
variation over time, lower amino acid concentrations in DDGS compared to meat and bone meal,
or the possibility that the small sample size analyzed by NIRS was not representative of those
assayed for amino acids.

Enzyme-based in vitro procedures


Novus International, Inc. has developed a patented enzyme-based assay (IDEA) designed to
rapidly predict amino acid digestibility of feed ingredients used in poultry diets including
soybean meal, meat and bone meal, poultry by-product meal and feather meal with assay times
ranging from two hours to less than one day for animal and plant proteins, respectively.
Schasteen et al. (2005) conducted a study to evaluate the applicability of IDEA for predicting the
amino acid digestibility of DDGS. They collected 28 DDGS samples and analyzed them using
the IDEA procedure along with determining true amino acid digestibility using a precision-fed
cecectomized rooster assay. Their results showed that true lysine digestibility varied among
DDGS samples and ranged between 59.1%-83.6% with an average of 70.3%, and the IDEA
analysis of these samples resulted in a strong correlation of IDEA values with the true lysine
digestibility values (r2 = 0.88). Crude protein concentrations in these samples ranged from
24.5%-30.2%. The other true amino acid digestibility values determined for these samples did
not vary to the same extent as lysine (25%), with cysteine having the next greatest variability
(20%) and methionine and alanine showing the least variation (8%). The IDEAanalysis showed
poor correlations (r2 = less than 0.5) for all amino acids other than lysine. These results suggest
that variations in poultry in vivo lysine digestibility values exist among DDGS sources and that
IDEAcan provide a good prediction of in vivo poultry digestibility of lysine for DDGS but not
other amino acids.
Pedersen et al. (2005) evaluated the effectiveness of two enzyme-based in vitro procedures (i.e.
a pepsin-based procedure and a pepsin/pancreatin-based procedure) to estimate crude protein and
amino acid digestibility coefficients in DDGS by growing pigs. Standardized ileal digestibility
coefficients of crude protein and amino acids were measured in 14 samples of DDGS using
growing pigs fitted with a T-cannula in the distal ileum of the small intestine. The pepsin
procedure estimated crude protein digestibility after 16 hours of incubating DDGS samples with
09 Nutrient Composition of DDGS: Variability & Measurement

14

pepsin at pH 1. For the pepsin/pancreatin procedure, DDGS samples were incubated with pepsin
for six hours at pH 2 followed by a 16 hour incubation with pancreatin at pH 6.8. Samples were
filtered after the incubation period and the filtrate was analyzed for crude protein concentration.
The correlation between in vivo crude protein digestibility and the digestibility obtained using
the pepsin procedure was low (r2 = 0.29), but was improved when the pepsin/pancreatin
procedure was used (r2 = 0.55). These researchers concluded that the pepsin/pancreatin
procedure may potentially be used to predict the in vivo digestibility of crude protein and amino
acids in DDGS, but additional work is needed to improve the correlations.

Reactive Lysine
Pahm et al. (2006) evaluated the use of a reactive lysine (homoarginine) procedure as a potential
in vitro procedure to predict lysine digestibility of DDGS for growing pigs. Results from this
study showed that DDGS samples and ileal digesta should be guanidinated for three days in a 0.6
molar methylisourea solution to achieve good recovery of amino acids and that the reactive
lysine procedure may be used as an in vitro method to predict the Standardized Ileal Digestibility
of lysine content in DDGS.

Literature Cited
Akayezu, J.M., J.G. Linn, S.R. Harty, and J.M. Cassady. 1998. Use of distillers grains and co-products in
ruminant diets. Proceedings of the 59th Minnesota Nutrition Conference, Bloomington, MN. September 1998.
Available: http://www.ddgs.umn.edu/info-dairy.htm. Accessed: August, 2006.
Batal, A.B., and N.M. Dale. 2006. True metabolizable energy and amino acid digestibility of distillers dried grains
with solubles. J. Appl. Poult. Res. 15:89-93.
Batal, A.B. and N.M. Dale, 2004. True metabolizable energy and amino acid digestibility of distillers dried grains
with solubles. Poultry Sci. 83 (Suppl 1):317.
Cromwell, G.L., K.L. Herkelman, and T.S. Stahly. 1993. Physical, chemical, and nutritional characteristics of
distillers dried grains with solubles for chicks and pigs. J. Anim. Sci. 71:679-686.
Davies, C.G.A, and T.P. Labuza. 2000. The Maillard reaction applications to confectionery products. Available:
http://courses.che.umn.edu/05fscn4111-1f/Readings%20pdf/maillard-confectionary.pdf. Accessed: August 2006.
Ergul, T., C. Martinez Amezcua, C. Parsons, B. Walters, J. Brannon and S.L. Noll. 2003. Amino acid digestibility
in corn distillers dried grains with solubles. Proceedings at the Poultry Science Association Meeting, Madison, WI.
July 2003. Available: http://www.ddgs.umn.edu/info-poultry.htm. Accessed: August 2006.
Ergul, T., C. Martinez Amezcus, C.M. Parsons, B. Walters, J. Brannon and S. L. Noll, 2003. Amino acid
digestibility in corn distillers dried grains with solubles. Poultry Sci. 82 (Suppl. 1): 70.
Erickson, G.E., T.J. Klopfenstein, D.C. Adams, and R.J. Rasby. 2006. Utilization of Corn Co-Products in the Beef
Industry. Nebraska Corn Board and the University of Nebraska. www.nebraskacorn.org. 17 pp.
Fastinger, N.D. and D.C. Mahan. 2006. Determination of the ileal amino acid and energy digestibilities of corn
distillers dried grains with solubles using grower-finisher pigs. J. Anim. Sci. 84:1722-1728.
Ferrer E., A. Alegra Farr, G. Clemente, and C. Calvo. 2005. Fluorescence, browning index, and color in infant
formulas during storage. J. Agric. Food Chem. 53:4911-4917.
Finot, P.A. 2005. The absorption and metabolism of modified amino acids in processed foods. J. of AOAC Int.
88:894:903.
Friedman, M. 1992. Dietary impact of food processing. Annual Reviews in Nutrition. 12:119-137.
Fu, S.X., M. Johnston, R.W. Fent, D.C. Kendall, J.L. Usry, R.D. Boyd, and G.L. Allee. 2004. Effect of corn
distillers dried grains with solubles (DDGS) on growth, carcass characteristics, and fecal volume in growing
finishing pigs. J. Anim. Sci. 82 (Suppl. 2):50.

09 Nutrient Composition of DDGS: Variability & Measurement

15

Ganesan, V., K.A. Rosentrater, and K. Muthukumarappan. 2006. Effect of moisture content and soluble levels on
the physical and chemical properties of DDGS. ASAE Annual International Meeting, Tampa, Fl, July 17-20.
Ham, G.A., R.A. Stock, T.J. Klopfenstein, E.M. Larson, D.H. Shain, and R.P. Huffman. 1994. Wet corn distillers
co-products compared with dried distillers grains with solubles as a source of protein and energy for ruminants. J.
Anim. Sci. 72:3246.
Hastad, C.W., M.D. Tokach, J.L. Nelssen, R.D. Goodband, S.S. Dritz, J.M. DeRouchey, C.N. Groesbeck, K.R.
Lawrence, N.A. Lenehan, and T.P Keegan. 2004. Energy value of dried distillers grains with solubles in swine diets.
J. Anim. Sci. 82 (Suppl. 2):50.
Klopfenstein, T. and R. Britton. 1987. Heat damage Real or Artifact. In: Dist. Feed Conf. Proceedings. 42:8486.
Knott, J., G.C. Shurson, and J. Goihl. 2004. Effects of the nutrient variability of distillers solubles and grains
within ethanol plants and the amount of distillers solubles blended with distillers grains on fat, protein and
phosphorus content of DDGS. Available: http://www.ddgs.umn.edu/articles-proc-storage-quality/2004-Knott%20Nutrient%20variability.pdf. Accessed: June, 2006.
Licitra G., T.M. Hernandez, and P.J. van Soest. 1996. Standardization of procedures for nitrogen fractionation of
ruminant feeds. Anim. Feed Sci. and Tech. 57:349-358.
Lumpkins, B., A. Batal and N. Dale, 2004. Evaluation of distillers dried grains with solubles as a feed ingredient
for broilers. Poultry Sci. 83:1891-1896.
Lumpkins, B.S. and A.B. Batal. 2005. The bioavailability of lysine and phosphorus in distillers dried grains with
solubles. Poultry Science 84:581-586.
National Research Council. 1994. Nutrient Requirements of Poultry, 9th Revised Edition, National Academy
Press, Washington, DC.
National Research Council. 1996. Nutrient Requirements of Beef Cattle, 7th Revised Edition, National Academy
Press, Washington, DC.
National Research Council. 1998. Nutrient Requirements of Swine, 10th Revised Edition, National Academy
Press, Washington, DC.
National Research Council. 2001. Nutrient Requirements of Dairy Cattle. 7th Rev. Ed. National Academy of Sci.,
Washington, DC.
Noll, S., C. Abe, and J. Brannon. 2003. Nutrient composition of corn distillers dried grains with solubles. Poultry
Science 82(Supplement):71.
Noll, S. 2004. DDGS in poultry diets: Does it make sense. Midwest Poultry Federation Pre-Show Nutrition
Conference, River Centre, St. Paul, MN. March 16, 2004.
Noll, S. L., J. Brannon, J. L. Kalbfleisch, and K. D.Roberson, 2005. Metabolizable energy value for corn distillers
dried grains with solubles in turkey diets. Poultry Sci. 84 (Suppl. 1):
Noll, S., C. Parsons, and B. Walters. 2006. Whats new since September 2005 in feeding distillers co-products to
poultry. Proc. 67th Minnesota Nutrition Confeence & University of Minnesota Research Update Session: Livestock
Production in the New Millennium. pp. 149-154.
Olentine, C. 1986. Ingredient profile: Distillers feeds. Proc. Distillers Feed Conf.
41:13-24.
Pahm, A.A., C. Pedersen, and H.H. Stein. 2006. Evaluation of reactive lysine (homoarginine) as an in vitro
procedure to predict lysine digestibility of distillers dried grains with solubles by growing pigs. J. Anim. Sci. (Suppl.
2) 84:90.
Pederson, C., A. Pahm, and H.H. Stein. 2005. Effectiveness of in vitro procedures to estimate CP and amino acid
digestibility coefficients in dried distillers grain with solubles by growing pigs. J. Anim. Sci. (Suppl. 2) 83:39.
Powers, W.J., H.H. Van Horn, B. Harris, Jr., and C.J. Wilcox. 1995. Effects of variable sources of distillers grains
plus solubles on milk yield and composition. J. Dairy Sci. 78:388-396.
Reese, D.E. and A.J. Lewis. 1989. Nutrient content of Nebraska corn. Nebraska Cooperative Extension Service
EC 89-219, pp. 5-7.
Roberson, K.D., J.L. Kalbfleisch, W. Pan and R.A. Charbeneau, 2005. Effect of corn distillers dried grains with
solubles at various levels on performance of laying hens and yolk color. Intl J. Poultry Sci. 4(2):44-51.

09 Nutrient Composition of DDGS: Variability & Measurement

16

Schasteen, C., J. Wu, and C. Parsons. 2005. Enzyme-based protein digestibility (IDEA) assay accurately
predicts poultry in vivo lysine digestibility for distillers dried grain and solubles (DDGS). J. Anim. Sci. (Suppl. 2)
83:39.
Schingoethe, D.J. 2004. Corn Co-products for Cattle. Proceedings from 40th Eastern Nutrition Conference, May
11-12, Ottawa, ON, Canada. pp 30-47.
Shurson, G.C., C. Santos, J. Aguirre, and S. Hernndez. 2003. Effects of Feeding Babcock B300 Laying Hens
Conventional Sanfandila Layer Diets Compared to Diets Containing 10% Norgold DDGS on Performance and Egg
Quality. A commercial field trial sponsored by the Minnesota Corn Research and Promotion Council and the
Minnesota Department of Agriculture.
Spiehs, M.J., G.C. Shurson, and M.H. Whitney. 1999. Energy, nitrogen, and phosphorus digestibility of growing
and finishing swine diets containing distillers dried grains with solubles. J. Anim. Sci. 77:188 (Suppl. 1).
Spiehs, M.J., M.H. Whitney, and G.C. Shurson. 2002. Nutrient database for distillers dried grains with solubles
produced from new ethanol plants in Minnesota and South Dakota. J. Anim. Sci. 80:2639.
Stein H.H., M.L. Gibson, C. Pedersen, and M.G. Boersma. 2006. Amino acid and energy digestibility in ten
samples of distillers dried grain with solubles fed to growing pigs. J. Anim. Sci. 84: 853-860.
Stein H.H., A.A. Pahm, and C. Pedersen. 2005. Methods to determine amino acid digestibility in corn byproducts. In: Proceedings of the 66th Minnesota Nutrition Conference. St. Paul. MN. USA. pp. 35-49.
Tjardes, J. and C. Wright. 2002. Feeding corn distillers co-products to beef cattle. SDSU Extension Extra. ExEx
2036, August 2002. Dept. of Animal and Range Sciences. pp. 1-5.
Urriola, P.E. 2006. Distillers dried grains with solubles digestibility, in vivo estimation and in vitro prediction.
M.S. Thesis, University of Minnesota.
Urriola, P.E., M.H. Whitney, N.S. Muley, and G.C. Shurson. 2006. Evaluation of regional differences in nutrient
composition and physical characteristics among six U.S. soybean meal sources. J. Anim. Sci. (Suppl. 2) 84:24.
van Barneveld, R.J., E.S Batterham, and B.W. Norton. 1994. A comparison of ileal and fecal digestibilities of
amino acids in raw and heat-treated field peas (Pisum sativum cultivar Dundale). Br. J. Nutr. 72:221-241.
van Soest, P.J. and V.C. Mason. 1991. The influence of the Maillard reaction upon the nutritive value of fibrous
feeds. Anim. Feed Sci. and Tech. 32:45-53.
Whitney, M.H., M.J. Spiehs, and G.C. Shurson. 2001. Availability of phosphorus in distillers dried grains with
solubles for growing swine. J. Anim. Sci. 79:108 (Suppl. 1).
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be formulated by a
qualified nutritionist. The USGC has no control over the nutritional content of any specific product which may be
selected for feeding. Potential buyers should consult an appropriate nutritionist for specific recommendations.
USGC makes no warranties that these recommendations are suitable for any particular herd or for any particular
animal. The USGC disclaims any liability for itself or its members for any problems encountered in the use of these
recommendations. By reviewing this material, buyers agree to these limitations and waive any claims against USGC
for liability arising out of this material.

09 Nutrient Composition of DDGS: Variability & Measurement

17

Fac to rs th at Im pac t
D D G S P r i ci n g &
Tra n s p o rtati o n

User Handbook

Factors to Consider in Buying DDGS:


Supplier, Price and Transportation
Logistics
The National Grain and Feed Association prepared the following recommendations for
companies buying DDGS. Given the rapidly evolving nature of the Biofuels industry and its coproducts, the following are several quality-assurance procedures that users of DDGS may wish to
consider as part of their quality-assurance program to help ensure that the product received meets
necessary quality criteria:
Know your DDGS Suppliers: The nutrient composition of DDGS can vary significantly
from plant to plant. Consider using an approved-supplier process to select those that can
provide DDGS products that meet desired quality parameters. This may include obtaining
laboratory analyses of DDGS, visiting the suppliers manufacturing site to evaluate their
quality-assurance procedures and verifying the suppliers product liability insurance
coverage.
Specify DDGS quality specifications in purchase contracts: Consider incorporating into
purchase contracts specific DDGS quality specifications important to your business. When
agreement between buyer and seller on precise nutrient composition is important, consider
including specified analytical methods for those nutrients in the purchase contract.
Sample and check the DDGS upon delivery: Consider obtaining a representative sample of
DDGS before unloading, using established sampling procedures, to determine if it meets
quality specifications before unloading. Some in-house DDGS inspection procedures to
consider include: 1) inspecting color (a dark color may indicate overheating); 2) checking
odor (a burnt smell also may indicate overheating); and 3) monitoring bulk density and
particle size.
Monitor DDGS for mycotoxins: As noted previously, the mycotoxin content present in
DDGS is three times the level that may be present in the corn used during the ethanol
production process. Consider testing periodically for mycotoxin content in DDGS to ensure
that excessive levels are not present.
Monitor DDGS Nutrient profiles: Consider routinely asking for nutrient profile
information from DDGS suppliers. Follow an established assay schedule to ensure that
accurate nutrient values for DDGS are used in formulations.
Taken from Feed and Feeding Digest; Volume 58, Number 1, February 1, 2007; Published by
National Grain and Feed Association.

Factors Affecting DDGS Price


Distillers dried grains with solubles (DDGS) is a very unique feed ingredient and is
considered to be a mid-protein, mid-energy feedstuff. It partially replaces corn, soybean meal
and phosphorus supplements in animal feeds. However, DDGS price does not exactly follow
corn and soybean meal prices. Figure 1 shows USDA historical wholesale prices for DDGS

10 - Factors that Impact DDGS Pricing &Transportation

($/short ton) compared to monthly average closing prices of near-month corn and soybean meal
futures from the Chicago Board of Trade. The USDA pricing point for DDGS is Lawrenceburg,
Indiana. Notice that DDGS prices tend to follow corn price more closely than soybean meal
prices. Overall trends in the corn and soybean meal markets will affect the DDGS price, but daily
volatility in the corn or soybean meal market on the Chicago Board of Trade does not always
translate into daily volatility in the DDGS market. If corn and/or soybean meal prices are
generally high relative to DDGS prices, DDGS will replace a larger proportion of corn and
soybean meal in animal feeds. On the other hand,.during times of seasonal price increases in the
DDGS market, corn and soybean meal will compete favorably with DDGS and DDGS will not
be used in least cost diet formulations.
A number of factors can affect DDGS pricing. First and foremost, it should be noted that the
highest demand for DDGS is in the United States, where approximately 88% of the distillers
grains produced today are consumed by livestock and poultry. However, the amount of DDGS
exported to other countries is increasing annually. Many producers and marketers of DDGS are
beginning to consider the potential export markets as a very important component in overall
DDGS demand. Strong demand in the early months of the year, coupled with historically short
supplies at the same time, has typically caused higher DDGS prices in the December-May
delivery period. However, even though this has been a historical trend, it is not certain that
DDGS will always be priced higher during this time of year.
Rapid growth of the U.S. ethanol industry has resulted in increasing amounts of DDGS on the
market each month, leading both buyers and sellers to expect there will not be supply shortages,
which had previously occurred in late winter and spring. Most of the distillers grains being
produced in the United States are fed to dairy and beef cattle, and increased usage in beef and
dairy rations, continues to provide additional demand every year. The swine industry is the
fastest growing sector for DDGS use in the United States. The U.S. poultry industry has yet to
reach its potential for DDGS inclusion in the diet, but results from recent and ongoing research
studies show that DDGS is an excellent ingredient for use in poultry feeds. When considering all
of these factors increased usage in the beef and dairy industries, increased acceptance and
usage in swine diets and the potential for increased use in poultry feeds the overall domestic
DDGS demand is increasing and will affect the price of DDGS in the future. The total amount of
DDGS produced is expected to reach 30 million metric tons per year within 10 years, significant
quantities of which will be used in the export market.

10 - Factors that Impact DDGS Pricing &Transportation

Figure 1. USDA historical wholesale prices for DDGS ($/short ton)


compared to monthly average closing prices of near-month corn
and soybean meal futures from the Chicago Board of Trade.
Soybean Meal, Corn, and DDGS Historical Prices
350

300

$/short ton

250

200
DDGS
Corn
Soybean Meal
150

100

50

0
Ja Ma Ma J Se No Ja Ma Ma J Se No Ja Ma Ma J Se No Ja Ma Ma J Se No Ja Ma Ma J
n- r- y- ul- p- v- n- r- y- ul- p- v- n- r- y- ul- p- v- n- r- y- ul- p- v- n- r- y- ul02 02 02 02 02 02 03 03 03 03 03 03 04 04 04 04 04 04 05 05 05 05 05 05 06 06 06 06

Transporting DDGS
Barges and Ocean Vessels
One of the most cost effective freight options available is to ship DDGS on the U.S. river
system via barges, and then load DDGS onto ocean vessels. Barge freight trades as a percent of
tariff. For example, the current tariff from St. Louis, Missouri, to New Orleans, Louisiana is
approximately USD $4.40 per metric ton. If freight from St. Louis to New Orleans is trading in
the range of 300% to 350% of that tariff, the current actual cost would be $13.20 to $15.40 per
metric ton. Percentage rates often fluctuate, so the numbers used in this example in no way
constitute a firm freight quote. They are merely given so the potential DDGS purchaser user has
an idea of how barge freight is calculated. Longer trips (e.g. Minneapolis, Minnesota, to New
Orleans) will have a higher tariff and probably a higher percentage rate. The United States has
5,000 miles of waterway navigable by barges and tug boats. Different tariffs and percentage rates
are traded for each navigable river in the United States. Barges traded to New Orleans are
usually offered as CIF NOLA (Cost, Insurance and Freight to New Orleans). DDGS is loaded
onto barges in the interior United States and shipped to the Port of New Orleans and surrounding
areas where it is transferred into holds on ocean-going vessels. This transfer is usually done via
mid-stream loaders. Both barges and vessels pull alongside the midstream loader where the
transfer is made. Vessels vary in size, but the most common vessel types are Handysize,
Handymax and Panamax vessels. The Handysize will hold 20,000-30,000 tons of cargo, whereas
the Handymax holds 35,000-49,000 tons, and the Panamax holds 50,000-75,000 tons of cargo.

10 - Factors that Impact DDGS Pricing &Transportation

One Panamax vessel will hold an amount of DDGS approximately equivalent to the amount
contained in 37 barges or 555 rail cars. It is important to determine the importers discharge port
requirements of the vessel you intend to charter, as different ports may require that the vessels
are geared (have their own cranes), may require different arrival draft limits, vessel length limits,
and width limitations. Ocean freight trades like a commodity and the rates change on a daily
basis. These rates depend on a number of factors including, but not limited to:
market conditions
type of vessel needed
port drafts
port charges
load terms
discharge terms
Factors affecting the overall ocean freight market include:
supply and demand issues
cost of vessel construction and operation
new vessel construction vs. vessel retirements
seasonal demand (e.g. grain harvest in North and South America)
Chinas demand for all raw materials
length of voyage
turnaround time
market psychology or expectations
Freight chartering options include:
Voyage Charters point A to point B shipments
o less risky for cost calculations
Time Charters gives more flexibility because the vessel is chartered for a specified
amount of time rather than by the voyage
o this scenario gives potentially higher risk and potentially higher reward.
Containers
Approximately 55% of the worlds container trade is attributable to Asia. The fact that the
United States is currently the worlds largest container importer puts the United States in a very
unique situation. Containers filled with electronics, textiles, auto parts and other goods, arrive in
the United States primarily from Asia and they need to be shipped back to that region in
order to be re-loaded with the consumer goods mentioned above. Roughly 60 out of every 100
containers leave the United States empty to return to Asia. The steamship lines would obviously
prefer to generate some revenue on the backhaul to Asia, rather than sending empty containers
back to Asia, which does not generate revenue for them. This backhaul is where DDGS, along
with other agricultural products, have found their niche in the freight market. The largest
surpluses of empty containers in the interior United States are found in Chicago, Illinois, and
Kansas City, Missouri, followed by Memphis, Tennessee. The typical container export process is
as follows:
1. DDGS is shipped from the ethanol plant to a facility dedicated to container loading.
These facilities are typically located close to large container collection yards where the
empty containers are stored.

10 - Factors that Impact DDGS Pricing &Transportation

2. In some cases, ethanol plants are loading the containers on site, thereby circumventing
the costs associated with a third party container loader.
3. Once the container is loaded with DDGS, it is trucked to the container collection yard and
placed onto a rail chassis.
4. From there the container is transported to a U.S. port to be loaded onto a container vessel.
Long Beach, California, handles more containers than any other U.S. port. Typical transit
time from Chicago to Long Beach is 7-10 days. Typical transit time from Long Beach to
Asian ports is 16-18 days.
Shipping via containers is an excellent option for the discriminating buyer who is looking to
purchase DDGS from a limited number of sources or ethanol plants.
Rail
Hopper rail cars are used to export DDGS to Mexico and Canada. Rail shipments of DDGS to
Mexico are growing exponentially every year, and the number of rail car shipments to Canada is
also increasing. Rail exports are considered to be the easiest to manage considering the number
of steps involved and the time in transit. Rail cars are loaded at the ethanol plant, billed with the
railroad and shipped to destination. Cars must be inspected and cleaned once they arrive at the
border. Once inspected and cleaned they cross the border and make their way to the final
destination. Some exporters allow their private cars to be shipped across the border and they are
using this as a revenue source. The principal railroads serving the United States are Union
Pacific (UP) and Burlington Northern Santa Fe (BNSF). Mexicos main rail lines are Ferromex
(FXE) and Kansas City Southern de Mexico (KCSM), formerly TFM. Canadas principal rail
lines are the Canadian National Railway (CN) and Canadian Pacific Railway (CPR).

Transportation and logistics planning


In general, high volume shipping programs will run on schedule. However, it is wise to plan
for the worst case and hope for the best. Rail cars can be delayed by weather, track problems,
rail cars that need repairs en-route, accidental routing to another destination in a switch yard, and
train derailment. Barges can be delayed by river freezings, flooding and droughts that lead to
overdraft problems. Containers are the most reliable of the shipping schedules, but if they are not
loaded on time to meet the vessel at the coast, they too can be delayed several weeks. In
summary; an importer of DDGS should consider the possibility of delays, up to 30-45 days, and
therefore keep some inventory at their facility as a cushion for these delays. In most cases
neither the sellers nor the freight carriers are responsible or liable for delays in delivery to the
destination.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be formulated by a
qualified nutritionist. The USGC has no control over the nutritional content of any specific product which may be
selected for feeding. Potential buyers should consult an appropriate nutritionist for specific recommendations.
USGC makes no warranties that these recommendations are suitable for any particular herd or for any particular
animal. The USGC disclaims any liability for itself or its members for any problems encountered in the use of these
recommendations. By reviewing this material, buyers agree to these limitations and waive any claims against USGC
for liability arising out of this material.

10 - Factors that Impact DDGS Pricing &Transportation

Ethanol Production
a n d i ts C o - P r o d u c ts

User Handbook

Production Processes Used to Produce


Ethanol and Distillers Co-Products
Conversion of Glucose to Ethanol
In the United States, corn is the predominant source of starch (glucose) used to produce
ethanol. With the exception of sugar cane, corn provides the highest ethanol yields compared to
any other feedstock being used (Table 1). However, research is underway to develop methods to
convert carbohydrates from cellulosic feedstocks such as softwood (Arwa et al., 2005) and nonstarch polysaccharides (Arthur, 2006), as well as using sugar beets (Savvides et al., 2000) as a
source of glucose.
Table 1: Starch content and ethanol yield of various feedstocks
Moisture (%) Starch (%)
Ethanol Yield (L/MT)
Feedstock
Starch
100.0
720
Sugar cane
654
Barley
9.7
67.1
399
Corn
13.8
71.8
408
Oats
10.9
44.7
262
Wheat
10.9
63.8
375
Source: Saskatchewan Agriculture and Food (1993)

Dry-Grind Ethanol Plants


Particle size reduction of grain
As shown in Figure 1, the initial step in ethanol production in a dry-grind ethanol plant is to
reduce the particle size of corn by grinding with a hammer mill. Hammer mills crush the corn
grain using high speed, rotating hammer tips. The fineness of the ground corn is determined
mainly by the rotor volume, hammer tip speed, number of hammers and the screen opening size
(Dupin et al., 1997). The screens used in the hammer are normally in the range of 1/8 to 3/16
inches in diameter. Particle size of the grain can affect ethanol yield (Kelsall and Lyons, 1999),
and therefore, ethanol producers tend to use finely ground corn to maximize ethanol yield. As
shown in Table 2, an extra 0.20 gallons (0.85 liters) of ethanol can be produced if the corn is
ground through a 3/16 inch screen compared to a 5/16 inch screen.

11 - Ethanol Production and its Co-Products

Figure 1: Dry-grind ethanol production processes and co-products (Erickson et al., 2005)
Grinding

Corn
Distillation

Slurry Mixing
Fermentation

Liquefaction

Whole Stillage

Ethanol

Centrifuge
Thin Stillage
Evaporation
Condensed
Distillers
Solubles

Coarse
Solids

Rotary Dryer

Wet
Distillers
Grains
Dried
Distillers
Grains

Distillers Dried Grains with Solubles DDGS

Table 2: Ethanol yield from ground corn of different particle size


Particle Size
Ethanol Yield (gallons/bushel)
Fine grind corn, 3/16 in.
2.65
Coarse grind corn, 5/16 in.
2.45
Cooking and saccharification
Water and recycled stillage are added to the ground corn which act as conditioners to begin
leaching of soluble protein, sugars and non-starch bound lipids (Chen et al. 1999).
The slurry mixture is then cooked to hydrolyze starch into glucose along with amylolytic
enzymes in order for yeast (Saccharomyces cerevisiae) to convert glucose to ethanol.
Temperatures typically used during the cooking process are 40-60C in the pre-mixing tank, 90165C for cooking and 60C for liquefaction (Kelsall and Lyons, 1998). Gelatinization of starch
starts to occur between 50-70C. A critical step in converting starch to glucose involves the
completeness of starch gelatinization (Lin and Tanaka, 2003). During gelatinization, nearly all of
the amylose in the starch granules is leached out (Han and Hamaker, 2001), which increases
viscosity due to swollen granules and gels consisting of solubilized amylose (Hermansson and
Kidman, 1995).
Complete hydrolysis of the starch polymer requires a combination of enzymes. Amylases are
the most widely used, thermostable enzymes in starch industry (Sarikaya et al., 2000). These

11 - Ethanol Production and its Co-Products

include -amylases or glucoamylase (Poonam and Dalel, 1995). Enzymes must be thermostable
in order for hydrolysis of the starch to occur immediately after gelatinization. Enzymes account
for 10-20% of the ethanol production cost (Gregg et al., 1998).
Some ethanol plants use batch cooking systems whereas others use continuous cooking
systems (Kelsall and Lyons, 1999). In a batch cooking system, a known quantity of corn meal is
mixed with a known quantity of water and recycled stillage. In the continuous cooking process,
corn meal, water, and recycled stillage are continuously added into a premix tank. The
temperature of the premix tank is maintained just below that needed for gelatinization and the
mash is continuously pumped through a jet cooker. The temperature of cooker is set at 120C,
and the mash passes from cooker into the top of a vertical column. The mash moves down the
column in about 20 minutes and is then passed into flash chamber for liquefaction at 80-90C.
High temperature-tolerant amylase is added at 0.05-0.08% to bring about liquefaction. The
retention time in the liquefaction/flash chamber is about 30 minutes. The pH of the system is
maintained within 6.0-6.5. Batch systems use less enzymes compared to continuous systems and
are also more energy efficient. The main disadvantage of batch systems is reduced productivity
or feedstock utilization per unit of time.
Fermentation
Fermentation is the process where yeast convert sugars to alcohol. The most commonly used
yeast is Saccharomyces cerevisiae (Pretorius, 2000) because it can produce ethanol to a
concentration as high as 18% in the fermentation broth. Saccharomyces is also generally
recognized as safe (GRAS) as a food additive for human consumption (Lin and Tanaka, 2006).
In model fermentation, about 95% of sugar is converted to ethanol and carbon dioxide, 1% into
cellular matter of the yeasts, and 4% into other products such as glycerol (Boulton et al., 1996).
Yeast accounts for about 10% of the economic cost of ethanol production (Wingren et al., 2003).
Pre-fermentation is done to achieve the desired number of yeast cells for fermentation and is a
process that involves agitation for 10-12 hours to achieve 300-500 million cells/milliliter.
Fermentation takes place at a temperature of about 33C (Thomas et al., 1996), at a pH of about
4.0 (Neish and Blackwood, 1951), and lasts between 48-72 hours (Ingledew, 1998). Carbon
dioxide is produced in addition to ethanol and can either be collected or is released into the air.
The control of normal yeast growth is a key factor in efficient ethanol production. The activity
of the yeasts is highly dependent on the temperature of the fermentation system. Torija et al.
(2003) reported that the optimum temperature for reproduction and fermentation in yeast is 28
and 32C, respectively. Fermentation efficiency of S. cerevisiae at high temperatures (above
35C) is low (Banat et al., 1998). Therefore, a cooling mechanism is required in fermentation
systems.
One of the challenges of managing fermenters in an ethanol plant is preventing contamination
with other microbes. Microbial contamination causes reduced ethanol yield and ethanol plant
productivity (Barbour and Priest, 1988). The most common organisms associated with microbial
contamination are lactobacilli and wild yeasts. These microbes compete with Saccharomyces
cerevisiae for nutrients (trace minerals, vitamins, glucose and free amino nitrogen) and produce
inhibitory end-products like acetic and/or lactic acid. Dekkera/Brettanomyces wild yeasts have
become a concern in fuel alcohol production (Abbott and Ingledew, 2005). Reduction in lactic
acid bacterial contamination is currently achieved by using antibiotics in fuel ethanol plants
(Narendranath and Power, 2005).
11 - Ethanol Production and its Co-Products

Distillation of Ethanol
After fermentation, ethanol is collected using distillation columns. Ethanol collected from the
fermenters is contaminated with water, and is purified using a molecular sieve system to remove
the water and produce pure ethanol.
Dry Milling Co-Products
The water and solids remaining after distillation of ethanol is called whole stillage,comprised
primarily of water, fiber, protein and fat. This mixture is centrifuged to separate coarse solids
from liquid. The coarse solids are also called wet cake and contain about 35% dry matter. Wet
cake can be sold to local cattle feeders without drying, or dried to produce dried distillers grains
(DDG),
The liquid, now called as thin stillage, goes through an evaporator to remove additional
moisture and the resulting co-product is called condensed distillers solubles which contains
approximately 30% dry matter. Condensed distillers solubles can be sold locally to cattle
feeders.
Or, the wet cake can be mixed with condensed distillers solubles and dried to produce
distillers dried grains with solubles (DDGS) which has 88% dry matter.

Wet Milling
Unlike dry-grind ethanol plants that ferment the entire ground corn kernel, wet mills separate
the corn kernel into various fractions which allows for the production of multiple food and
industrial products including ethanol. The corn wet milling industry was developed in the early
19th century with the primary purpose of producing starch for use in food and laundry products
(Kerr, 1950). In the 1920s, wet mills began producing crystalline dextrose (Newkirk, 1923) and,
after World War II, began producing ethanol.
Beginning in the early 1990s, wet mills began producing high fructose corn syrup in addition
to other products. Currently, there are approximately 28 wet milling plants in the United States
a large number of them have been built in the last few decades (Johnson and May, 2003). An
overview of the wet milling process is shown in Figure 2.

11 - Ethanol Production and its Co-Products

Figure 2: Wet milling processes and co-products (Erickson et al., 2005)


Steeping

Corn

Starch-Gluten
Separation

Milling
Cyclone

Starch

Separation
Wet Gluten
Germ
Separation
n

Germ
Oil
Refining
Corn Oil

Drying

Cake
(Fiber)

Fermentation

Syrup
Refining

Dextrose
Corn
Gluten
Meal

Corn
Syrup

Corn
Gluten
Feed

Corn Germ
Meal

Starch

Ethanol
Chemicals

High
Fructose
Corn
Syrup

Grain Cleaning
Corn kernel is initially cleaned to remove broken kernels, chaff, cob pieces and foreign
material. This process is important because broken kernels can release starch in steepwater,
which can gelatinize leading to undesirable viscosity during evaporation of steep water into steep
liquor (May, 1998).
Steeping
Steeping involves soaking of the corn kernels under the controlled temperature (48-50C), time
(35-50 hours), concentration of sulfur dioxide (0.1-0.2%) and lactic acid (Watson, 1984). Water
acts as a conditioner so that milling can be performed under optimal conditions (Bass, 1988).
Steeping softens the corn kernel for milling, inhibits microbial growth and enhances pure starch
recovery (Bartling, 1940).
Milling
After soaking, the corn germ becomes soft and rubbery. Hydrocyclones with counter rotating
discs and intermeshing fingers separate the germ. Since the germ is lighter in weight than the rest
of corn kernel it can be easily separated by centrifugal force. Once removed, the germ is purified
to remove starch and protein extracts by washing with water. Oil is then extracted from the germ
to produce corn oil.

11 - Ethanol Production and its Co-Products

Fiber is separated by pumping the slurry (starch, gluten, fiber and kernel fragments) with
considerable force on 120 wedge-wire screen. Fiber particles and kernel fragments are large in
shape and are screened out to leave starch and protein.
Gluten is separated by high speed centrifuges due to the fact that protein is lighter in weight
compared to starch (May, 1987). Gluten is then thickened in centrifuges, dewatered to 42%
solids by vacuum filtration, and dried to 88% solids for sale as corn gluten meal (Jackson and
Shandera, 1995).
Starch processing
Impurities, in the form of proteins, are removed by washing the starch in fresh water using a
countercurrent process in the centrifuges. The purified starch contains less than 0.4% protein
with less than 0.01% free protein (May, 1987). The protein that is removed consists primarily of
starch-protein complexes which are recycled back to the primary separation step. Purified starch
can then be dried, fermented to produce ethanol or refined to produce corn syrup. The procedure
used to produce ethanol from starch in wet mills are similar to those previously described fro
dry-grind ethanol plants.
Co-Products
Corn steep liquor is a high-energy liquid feed ingredient. It contains about 25% crude protein
on a 50% dry matter basis. This product is sometimes combined with the corn gluten feed or
may be sold separately as a liquid protein source for beef or dairy rations. It also can be used
as a pellet binder and is a source of B-vitamins and minerals.
Corn germ meal contains 20% protein, 2% fat, and 9.5% fiber. It has an amino acid balance
that makes it valuable in poultry and swine diets.
Corn gluten feed is a medium protein ingredient composed of the bran and fibrous portions
of the corn kernel. It may or may not contain the condensed corn extractives. This co-product
can be sold wet or dry. The bran and condensed extractives (sometimes called germ meal) are
combined and dried in a rotary dryer. The dried corn gluten feed is made into pellets to
facilitate handling. It typically contains about 21% protein, 2.5% fat and 8% fiber. Wet corn
gluten feed (45% dry matter) is perishable in 6-10 days and must be fed within that time
period or stored in an anaerobic environment. Corn gluten feed is primarily used in dairy and
beef cattle rations.
Corn gluten meal is a high protein concentrate which typically contains 60% protein, 2.5%
fat and 1% fiber. It is a valuable source of methionine. Corn gluten meal also has a high level
of xanthophylls which make it an attractive ingredient in poultry diets as a source of yellow
pigment.

Literature Cited
Abbott, D.A., and W.M. Ingledew. 2005. The importance of aeration strategy in fuel alcohol fermentations
contaminated with Dekkera/Brettanomyces yeasts. Applied Microbiology and Biotechnology. 69:16-21.
Arthur, J.R., C.K. Williams, B.H. Davison, G. Britovsek, J. Cairney, C.A. Eckert, W.J. Frederick, Jr., J.P. Hallett,
D.J. Leak, C.L. Liotta, J.R. Mielenz, R. Murphy, R. Templer, and T. Tschaplinski. 2006. The path forward for
biofuels and biomaterials. Science. 311:484-489.
Arwa, K., A. Berlin, N. Gilkes, D. Kilburn, R. Bura, J. Robinson, A. Markov, A. Skomarovsky, A. Gusakov, O.
Okunev, A. Sinitsyn, D. Gregg, D. Xie, and J. Saddler. 2005. Enzymatic hydrolysis of steam-exploded and ethanol

11 - Ethanol Production and its Co-Products

organosolvpretreated Douglas-Firby novel and commercial fungal cellulases. Applied Biochemistry and
Biotechnology. 26th Symposium on Biotechnology for Fuels and Chemicals. Volume 121, Issue 1-3, pps. 219-230.
Banat, I.M., P. Nigam, D. Singh, R. Merchant, and A.P. McHale. 1998. Ethanol production at elevated
temperatures and alcohol concentrations: A review; Part-I Yeast In General. World J. Microbiol. Biotechnol.
14:809-821.
Barbour, E.A., and F.G. Priest. 1988. Some effects of Lactobacillus contamination in scotch whisky
fermentations. J. Inst. Brew. 94:89-92.
Bartling, F.W. 1940. Wet process corn milling. No. 5. The steep house. Am. Miller. 68:40-41.
Bass, E.J. 1988. Wheat floor milling. Pages 1-68 In: Wheat Chemistry and Technology, Vol. 2. Y. Pomeranz. Ed.
Am. Assoc. Cereal Chem. St. Paul, MN, USA.
Boulton, B., V.L. Singleton, L.F. Bisson, and R.E. Kunkee. 1996. Yeast and biochemistry of ethanol fermentation.
In Principles and Practices of Winemaking, Boulton B, Singleton VL, Bisson LF, Kunkee RE (eds). Chapman and
Hall. New York, NY, USA. pp. 139-172.
Chen, J.J., S. Lu, and C.Y. Lii. 1999. Effect of milling on physicochemical characteristics of waxy rice in Taiwan.
Cereal Chemistry 76:796-799.
Dupin, I. V. S., B. M. McKinnon, C. Ryan, M. Boulay, A. J. Markides, P. J. Graham, P.
Fang, Q., I. Boloni, E. Haque, and C. K. Spillman. 1997. Comparison of energy efficiency between roller mill and
a hammer mill. Appl. Engineering in Agric.13:631-635.
Erickson G.E., T.J. Klopfenstein, D.C. Adams, R.J. Rasby. 2005. General overview of feeding corn milling coproducts to beef cattle. In: Corn Processing Co-Products Manual. University of Nebraska. Lincoln, NE, USA.
Gregg, D.J., A. Boussaid, and J.N. Saddler. 1998. Techno-economic evaluations of a generic wood-to-ethanol
process: effect of increased cellulose yields and enzyme recycle. Bioresour. Technol.63:7-12.
Han, X.Z., and B.R. Hamaker. 2001. Amylopectin fine structure and rice starch paste breakdown. J. Cereal Sci.
34:279-284.
Hermansson, A.M., and S. Kidman. 1995. Starch A phase-separated biopolymer system. In: S.E. Harding, S.E.
Hill and J.R. Mitchell, Editors, Biopolymer Mixtures, Nottingham University Press, UK. pp. 225-245.
Ingledew, W.M. 1998. Alcohol production by Saccharomyces cerevisiae: A yeast primer. Chapter 5 In: The
alcohol textbook. 3rd ed. K.A. Jacques, T.P. Lyons and D.R. Kelsall Ed. Nottingham University Press. Nottingham,
UK.
Jackson, D.S., and D.L. Shandera, Jr. 1995. Corn wet milling: Separation chemistry and technology. Adv. Food
Nutr. Res. 38:271-300.
Johnson, L.A. and J.B. May. 2003. Wet milling: The basis for corn refineries. In: Corn: Chemistry and
Technology. Ed. S.A Watson. pp. 449-495. Am. Assoc. Cereal Chem., St. Paul, MN, USA.
Kelsall, D.R., and T.P. Lyons. 1999. Grain dry milling and cooking for alcohol production: designing for 23%
ethanol and maximum yield. Chapter 2. In: The alcohol textbook. 3rd ed. K.A. Jacques, T.P. Lyons and D.R. Kelsall
Ed. Nottingham University Press. Nottingham, UK.
Kerr, R.W. 1950. Chemistry and industry of starch. Academic Press, New York, NY, USA. p. 29.
Lin, Y., and S. Tanaka. 2006. Ethanol fermentation from biomass resources: current state and prospects. Appl.
Microbiol. Biotechnol. 69: 627-642.
Narendranath, N.V., and R. Power. 2005. Relationship between pH and medium dissolved solids in terms of
growth and metabolism of Lactobacilli and Saccharomyces cerevisiae during ethanol production. Applied and
Environmental Microbiology. 71: 2239-2243.
Neish, A.C., and A.C. Blackwood. 1951. Dissimilation of glucose by yeast at poised hydrogen ion concentrations.
Canadian Journal of Technology. 29:123-129.
Newkirk, W.B. 1923. Method of making grape sugar. U.S. Patent 1:471,347.
Poonam, N. and S. Dalel. 1995. Enzyme and microbial systems involved in starch processing. Enzyme Microb.
Technol. 17:770-778.
Pretorius, I.S. 2000. Tailoring wine yeast for the new millennium: Novel approaches to the ancient art of
winemaking. Yeast 16:675-729.
Renewable Fuels Association. 2006a. http://www.ethanolrfa.org/industry/locations/. Accessed October, 2006.

11 - Ethanol Production and its Co-Products

Renewable Fuels Association. 2006b. http://www.ethanolrfa.org/industry/resources/coproducts/. Accessed


October, 2006.
Sarikaya, E., T. Higassa, M. Adachi, and B. Mikami. 2000. Comparison of degradation abilities of - and amylases on raw starch granules. Proc. Biochem. 35:711-715.
Saskatchewan Agriculture and Food. 1993. Establishing an Ethanol Business.
Savvides, A.L., A. Kallimanis, A. Varsaki, A.I. Koukkou, C. Drainas, M.A. Typas, and A.D. Karagouni. 2000.
Simultaneous ethanol and bacterial ice nuclei production from sugar beet molasses by a Zymomonas mobilis CP4
mutant expressing the inaZ gene of Pseudomonas syringae in continuous culture. J. Appl. Microbiol.89: 1002-1008.
Thomas, K.C., S.H. Hynes, and W.M. Ingledew. 1996. Practical and theoretical considerations in the production
of high concentrations of alcohol by fermentation. Proc. Biochem. 31:321-331.
Torija, Ma. J., N. Rozs, M. Poblet, J.M. Guillamn, and A. Mas. 2003. Effects of fermentation temperature on
the strain population of Saccharomyces cerevisiae . International Journal of Food Microbiology. 80: 47-53.
Watson, S.A. 1984. Corn and sorghum starches: Production. In: Starch: Chemistry and Technology. Ed. R.L.
Whistler, J.M. BeMiller, and E.F. Paschall.417-467. Academic Press, Orlando, FL, USA.
Wingren, A.M., Galbe, and G. Zacchiu. 2003. Techno-Economic Evaluation of Producing Ethanol from
Softwood: Comparison of SSF and SHF and Identification of Bottlenecks. Biotechnol. Prog. 19:1109-1117.
The U.S. Grains Council (USGC) provides these feeding recommendations to assist potential buyers in
understanding generally-accepted feeding levels. However, all rations for specific herds should be formulated by a
qualified nutritionist. The USGC has no control over the nutritional content of any specific product which may be
selected for feeding. Potential buyers should consult an appropriate nutritionist for specific recommendations.
USGC makes no warranties that these recommendations are suitable for any particular herd or for any particular
animal. The USGC disclaims any liability for itself or its members for any problems encountered in the use of these
recommendations. By reviewing this material, buyers agree to these limitations and waive any claims against USGC
for liability arising out of this material.

11 - Ethanol Production and its Co-Products

Fr e q u e n t ly A s ke d
Q u e st i o n s A b o u t D D G S

User Handbook

Frequently Asked Questions About DDGS


What is the average protein content of high quality DDGS?
In a University of Minnesota DDGS sampling survey evaluating 32 different DDGS sources,
the average crude protein content was 27.6% on an as fed basis, with a range from 25.6-29.4%.
Recently, a few ethanol plants are using new processes to produce ethanol and higher protein
DDGS which may contain as much as 40-50% crude protein.

Is there DDGS available that has crude protein levels of 40% or more?
Yes. However, this high protein DDGS currently represents a very small percentage of the total
DDGS production.

What are the advantages/disadvantages of DDGS that is advertised as


having 40% protein or more compared to DDGS with an average
protein of 26-28%?
High protein (more than 40%) DDGS offers the greatest advantage in ruminant diets because
the rumen microbes can convert the nitrogen in this protein into microbial protein that meets the
animals amino acid requirements. High protein DDGS is lower in fat, which may allow higher
feeding levels to lactating dairy cows without concern about milk fat depression. High protein
DDGS is also lower in phosphorus than typical DDGS which is also an advantage for ruminants
because it will reduce the amount of phosphorus excreted in manure.
However, as the protein content of DDGS increases, fat, fiber and phosphorus content
decrease, and these factors may limit the value of high protein DDGS in swine and poultry diets.

Can any DDGS replace soybean meal on a one-to-one basis in a


layer/broiler/swine/ruminant diet? Why or why not?
No. Each individual feed ingredient is a package of nutrients in various quantities and
proportions. The three most expensive nutrients in a livestock and poultry feeds are energy,
amino acids and phosphorus. Depending on relative ingredient prices, DDGS partially replaces
some of the energy, amino acid and phosphorous sources in commercial livestock and poultry
diets. In typical corn/soybean rations DDGS can partially replace corn and soybean meal. But
where a greater variety of energy and protein sources are available, DDGS may replace other
ingredients without reducing the soybean meal in the ration.
The trade-off between soybean meal and DDGS in swine and poultry rations is complex:
The energy value of DDGS is equal to, or greater than dehulled soybean meal in livestock
and poultry diets.
The protein content of DDGS is typically averages about 27% whereas soybean meal
contains 44-48% crude protein.
The amino acids that are most likely to be limiting in corn-soybean meal based swine and
poultry diets are lysine, methionine, threonine and tryptophan. Soybean meal is
substantially higher in these essential amino acids and they are more digestible than
in DDGS.
12 - Frequently Asked Questions About DDGS

Soybean meal contains about the same concentration of phosphorus as DDGS, but the
majority of the phosphorus in DDGS is in a chemical form that is easily digested and
utilized by swine and poultry compared to the indigestible form of phosphorus (phytic
acid) found in soybean meal. This nutritional advantage for DDGS allows nutritionists to
significantly reduce the amount of inorganic phosphorus supplementation needed in the
diet, diet cost and phosphorus concentrations in manure, while supporting optimum swine
and poultry performance.

Does DDGS need to be treated with propionic acid or mold inhibitors


in order to extend the shelf life of DDGS?
Preservatives and mold inhibitors are commonly added to wet distillers grains (~50%
moisture) to prevent spoilage and extend shelf life. However, since the moisture content of
DDGS is usually between 10-12%, there is minimal risk of spoilage during transit and storage
unless water leaks into transit vessels or storage facilities. No research studies have been
conducted to demonstrate that preservatives and mold inhibitors are necessary to prevent
spoilage and extend shelf life of DDGS.
In a U.S. Grains Council field trial, DDGS was shipped from an ethanol plant in South Dakota
in a 40 ft. container to Taiwan. Upon arrival in Taiwan, DDGS was put into 50 kg bags and
stored in a covered steel pole barn for 10 weeks during the course of the dairy feeding trial on a
commercial dairy farm located about 20 km south of the Tropic of Cancer. Environmental
temperatures averaged 90+ degrees F and humidity was in excess of 90% during this storage
period. Samples of DDGS were collected upon arrival at the farm and again after the 10 week
storage period. There was no change in peroxide value (measure of oxidative rancidity of oil)
during this trial.

Does DDGS contain alcohol?


No. The distillation process used in ethanol plants is very complete and, because alcohol is
very volatile (evaporates easily), any alcohol remaining is lost during the drying process used to
produce DDGS.

How can I qualify a DDGS supplier?


Due to the variability in processes used by ethanol plants to produce ethanol and DDGS, there
can be significant variation in nutrient content and digestibility among sources. This variation in
nutrient content and digestibility makes it unwise for nutritionists to formulate diets using
typical nutrient values. Therefore, many DDGS users have chosen to contact direct marketers
of DDGS, request nutrient information and samples from specific ethanol plants of interest,
develop a preferred supplier list of ethanol plants that meet their quality criteria, and purchase
and use only DDGS from those sources.

How do you value DDGS in relation to cost?


The best method of determining DDGS value in various types of livestock and poultry diets is
to obtain a complete nutrient profile and the digestibility coefficients of the source being
considered, the price at which it can be purchased and offering it as an ingredient to compete
with the nutrient profiles and prices of other ingredients being used in a least cost diet
12 - Frequently Asked Questions About DDGS

formulation software program. Alternatively, value can be calculated by knowing the protein
(amino acid), fat and phosphorus content per ton of DDGS and using the cost and concentrations
of these nutrients in competing ingredients commonly used (e.g. soybean meal, choice white
grease and dicalcium phosphate). However, this approach does not account for digestibility of
nutrients which may be lower or higher in DDGS compared to other nutrient sources.

What should be written in the certificate of analyses of DDGS?


Typically, DDGS is traded on a protein, fat or pro-fat combination for nutrient guarantees.
However, more DDGS customers are asking for additional guarantees depending upon the
feeding application where they intend to use it. Additional guarantees are negotiated between the
buyer and seller. It is extremely important to agree on the laboratory and testing method that will
be used for any nutrient analysis being guaranteed or checked because the testing procedure can
have a significant influence on whether or not a guarantee is met.

How much aflatoxin is in DDGS?


Most of the corn used to produce DDGS is grown in the upper Midwest of the United States
where there is minimal risk of aflatoxin production except under unusual growing conditions
(high temperature and high humidity). When these growing conditions occur, they are usually in
relatively small, isolated regions. On average, growing conditions conducive to aflatoxin
production in the upper Midwest occur one out of 10 years. If aflatoxin is detected in a given
area, most ethanol plants receiving corn from that area will use a black light to screen
incoming corn and many will set maximum allowable levels to avoid concentrating aflatoxin in
DDGS. If corn containing aflatoxin or other mycotoxins is used to produce ethanol and DDGS,
those mycotoxins are concentrated by three-fold compared to the initial level found in the corn.

How do you manage the caking of DDGS in containers?


Research studies are being conducted to determine the factors that cause flowability problems
and potential solutions to reduce these problems. It appears that several factors contribute to
whether or not DDGS sets up in a container including: fine particle size, warm temperature
when loading, moisture content, proportion of solubles added to the grains fraction before drying
and the number of times it has been handled and unloaded during transit.

Does DDGS improve egg yolk and poultry skin pigmentation when it
is added to the diet?
Yes. Several studies have been conducted in the past few years showing that egg yolk and
poultry skin pigmentation improves when DDGS is added to the diet. Currently, there are limited
data on xanthophyll content of DDGS but initial sampling indicates it can range from very little
(dark colored DDGS) to approximately 40-50 parts per million (ppm) (light golden colored
DDGS). Although the level of xanthophyll is significantly less than found in corn gluten meal
(180 to 200 ppm), it still contributes a significant amount of pigment to poultry diets and as a
result, less synthetic pigment must be added to the diet to achieve the desired level of
pigmentation. This can represent a significant savings in diet cost.

12 - Frequently Asked Questions About DDGS

U.S. DDGS
S u p p l i e r s L i st

User Handbook

U.S. Suppliers Of
Distillers Dried Grains with Solubles
(DDGS)
AG Processing, Inc (AGP)
Contact: Jeremy Frandson; (402) 492-3322; jfrandson@agp.com
Historically, AGP has been a wet distillers grains marketer, but is now marketing DDGS from
two ethanol plants in Iowa and Nebraska. Their export capabilities include barge, rail car and
container shipments. AGP owns an export facility located in the Port of Grays Harbor in the state
of Washington.

Archer Daniels Midland Company (ADM)


Contact: Greg Morris; (217) 451-4322; Greg_morris@admworld.com
Contact: Matt Woodhouse; (217) 451-7469; m_woodhouse@admworld.com
Website: www.admworld.com
ADM is a producer and exporter of DDGS. They have production facilities in Peoria, Illinois,
and Walhalla, North Dakota. Two additional facilities are under construction in Cedar Rapids,
Iowa, and Columbus, Nebraska. In addition, ADM has marketing agreements from unnamed
ethanol plants. ADM has four gulf export elevators as well as one floating rig to load DDGS or
combination commodity vessels. ADM has three exclusive container loading facilities and also
utilizes two additional public container loading facilities. Export capabilities include vessel,
barge, rail car and container shipments. ADM is the largest U.S. producer and exporter of bulk
DDGS.

Aventine Renewable Energy Inc.


Contact: Greg Kauls;(309) 347 9713; greg.kauls@aventinerei.com
Website: http://www.aventinerei.com/index.html
Aventine Renewable Energy is an ethanol and DDGS producer. Aventine markets both dry and
wet DDGS and sells into rail, barge and truck markets. Please see their website and contact them
directly for more information.

Cargill, Inc.
Contact: Larry Holy; (952) 742-0193; larry_j_holy@cargill.com
Contact: Todd Nicklaus; todd_nicklaus@cargill.com
Website: www.cargill.com
Cargill has marketing agreements with dry grind ethanol facilities. They are also investors in
those facilities. Cargill takes market positions from other DDGS marketers in the United States.
Cargill sells a combination 36% protein-fat DDGS. Export capabilities include vessel, barge,
container, rail cars and truck shipments. Cargill markets approximately 1 million metric tons of
product per year.

13 U.S. DDGS Suppliers List

Commodity Specialists Company (CSC)


Contact: Steve Markham; (800) 769-1066; smarkham@csc-world.com
Contact: Sean Broderick; (800) 769-1066; sbroderick@csc-world.com
Website: www.csc-world.com
CSC is currently marketing DDGS for 21 ethanol plants in the United States. CSC is constantly
seeking new opportunities to market DDGS for new ethanol plants that are in planning or under
construction. Last fiscal year, CSC marketed approximately 1.76 million metric tons of both wet
and dried distillers grains. They have the capability to load trucks and rail cars from their
existing plants. Product can be shipped from their plants to container loading facilities as well as
barge loading facilities for export to foreign markets. CSC has the capability to ship product via
rail to Duluth, Minnesota, for loading onto ocean vessels for export. The following list includes
ethanol plants for which CSC does the DDGS marketing that are best suited for the export
market:
Adkins Energy, LLC - Lena, Illinois
Golden Grain Energy LLC - Hampton, Iowa
Glacial Lakes Energy - Watertown, South Dakota
Dakota Ethanol - Wentworth, South Dakota
Bushmills Energy - Atwater, Minnesota
Lincolnway Energy - Nevada, Iowa
Corn Plus (DDG, no solubles added) - Winnebago, Minnesota

CGB
Contact: Mitchell McGee; (985) 867-3554; mitchell.mcgee@cgb.com
Website: www.cgb.com
CGB Consolidated Grain & Barge is a full service Ag commodity trading company that
specializes in trading and exporting of DDGS. Please see their website and contact them directly
for more information.

Dakota Gold Marketing


Contact: Mike Salonen; (605) 332-2200; msalonen@dakotagoldmarketing.com
Website: www.dakotagoldmarketing.com
Dakota Gold is currently marketing DDGS for 19 ethanol plants in the United States. Dakota
Gold markets approximately 1.8 million metric tons of both wet and dried distillers grains per
year. They have a partnership in Southeast Asia with Delst Asia. Their logistical capabilities
include truck, rail and barge shipments. Product may also be shipped via rail to container loading
facilities for export. Dakota Gold markets a product called HP (High Protein) DDGS. The
modified processes used in one of their ethanol plants results in a higher protein, lower energy
DDGS. The following ethanol plants for which Dakota Gold markets DDGS are capable of
supplying DDGS for the export market:
Northstar Ethanol - Lake Crystal, Minnesota
Voyager Ethanol - Emmetsburg, Iowa
Sioux River Ethanol - Hudson, South Dakota
Otter Creek Ethanol - Ashton, Iowa
Iowa Ethanol - Hanlontown, Iowa
James Valley Ethanol - Groton, South Dakota
13 U.S. DDGS Suppliers List

Great Plains Ethanol - Chancellor, South Dakota


Ethanol 2000 - Bingham Lake, Minnesota
Tall Corn Ethanol (HP DDGS is available) - Coon Rapids, Iowa
Northern Lights Ethanol - Milbank, South Dakota
Northeast Missouri Grain - Macon, Missouri
EXOL - Albert Lea, Minnesota

FC Stone
Contact: Chris Aberle; (386) 409-3999; chrisa@fcstone.com
Website: www.fcstone.com
FC Stone is a broad-based commodity risk management and trading company headquartered in
Des Moines, Iowa. They offer DDGS marketing services to ethanol plants. Please see their
website and contact them directly for more information.

J.D. Heiskell & Co.


Contact: Brent Lorensen; (402) 289-5700; blorensen@heiskell.com
Website: www.heiskell.com
J.D. Heiskell & Co. is a rail exporter of DDGS. Please see their website and contact them
directly for more information.

Land O Lakes, LLC (LOL)


Contact: Duane Kriener; (651) 765-5644; dpkriener@landolakes.com
Website: www.landolakes.com
LOL currently has exclusive agreements with eight ethanol plants for the marketing of their
DDGS. At present, they plan to market DDGS for eight more plants that are under construction.
They have the capability to export DDGS in bulk via vessel, barge, rail car and container.
Product can also be bagged at the customers request.

Louis Dreyfus Commodities Inc.


Contact: Kim Hawks; (203) 761-2156; kim.hawks@ldcommodities.com
Website: www.louisdreyfus.com
Louis Dreyfus Commodities Inc. will commence production of Ethanol and DDGS in October
of 2007. They are also a full service commodity export and trading company involved in all
major commodities. Their trade in DDGS will include exports as well as U.S. domestic truck and
rail.

MB Commodities Inc.
Contact: Pete Thiele; (303) 247-0010
MB Commodities is a full service cash commodities brokerage that brokers most major U.S. Ag
commodities such as (DDGS, Corn, Soybeans, Meals, and Wheat). They broker DDGS for most
major U.S. suppliers as well as international importers and domestic U.S. end users in the feed
industry.

Mills Brothers
Contact: Erick Mills; (206) 575-3000; erick@millsbros.com
13 U.S. DDGS Suppliers List

Website: www.millsbros.com
Mills Brothers is a west coast based principal trading company that specializes in exports
DDGS in containers. Please see their website and contact them directly for more information.

Pattison Bros.
Contact: Russ Leuck; (563) 425-3569; leuckr@pattisonbros.com
Website: www.pattisonbros.com
Pattison Bros. markets golden DDGS with a 36% combination protein/fat as well as a higher
protein (35%-45%) DDGS throughout the world. Pattison can export via container, vessel, barge
and rail car shipments. Pattison has been marketing agricultural products to Southeast Asia for
more than 15 years and understands the importance of quality. The key to DDGS quality is
buying from known origins to achieve consistency: Consistent Quality, Consistent Service and
Consistent Price.

Quality Technology International, Inc. (QTI)


Contact: Daniel Hammes; (847) 649-9300, Ext. 17; danh@qtitech.com
Website: www.qtitech.com
QTI is currently involved in marketing improved versions of DDGS. These improved
versions will have higher protein content (45%-60%), lower fat content and lower fiber content.
QTI is currently investigating potential customers and markets that have feeding applications for
these improved DDGS products. Their focus will be primarily on identifying customers who
want to use these products in diets for layers, pet foods and aquaculture. These products will be
made in southern Wisconsin, Indiana and Ohio. Export capabilities will include barge, rail and
container shipments.

The Rice Company


Contact: Don Morrison; (203) 256-0333; krohn@riceco.com
Website: www.riceco.com
The Krohn Division of The Rice Company has originated DDGS for export from the United
States for over 25 years. They market product for several processors and ship to end users around
the world in bulk vessels. They also sell loaded vessels in the U.S. Gulf and market product in
containers and rail cars. They have distribution networks in several foreign locations.

The Scoular Company


Contact: Thomas Kopp; (612) 851-3703; tkopp@scoular.com
Website: www.scoular.com
Scoulars export capabilities include barge, rail car and container shipments. They have their
own container freight and container transloading division, TSC Container Freight, which is a
Scoular subsidiary. Scoular has marketing agreements with ethanol plants for the sale of DDGS.

13 U.S. DDGS Suppliers List

United Bio Energy Ingredients, LLC (UBE)


Contact: Randy Ives; (800) 796-7890; randy.ives@usbioenergy.net
Contact: Byron Stewart; (316) 616-3515; byron.stewart@usbioenergy.net
Websites: www.usbioenergy.net and www.unitedbioenergy.com
UBE is involved in ethanol and DDGS production. They have exclusive DDGS marketing
agreements for 16 ethanol plants in operation with an annual DDGS production capacity of 2.6
million tons UBE has seven of their own ethanol plants under construction. They provide core
services involving ethanol plant general management, operations management, project
management, ethanol and distillers grains marketing, grain origination and overall risk
management. Of the 16 ethanol plants that market their DDGS through UBE, 11 are capable of
shipping product to export markets. The seven additional ethanol plants under construction by
U.S. Bio Energy/UBE will be capable of supplying product to export markets. Rated annual
DDGS production capacity for the seven plants under construction is an additional 2.0 million
tons. Transportation capabilities include truck, rail, barge and container shipments. The
following is a list of operating ethanol plants for which UBE does the marketing capable of
supplying DDGS to the export market:
Big River Resources- West Burlington, Iowa
Amaizing Energy - Denison, Iowa
Hawkeye Renewables - Iowa Falls, Iowa
Golden Triangle - Craig, Missouri
EKAE (blend of milo and corn DDGS) - Garnett, Kansas
Illinois River Energy - Rochelle, Illinois
Hawkeye Renewables - Fairbank, Iowa
U.S. Bio Energy - Albert City, Iowa
U.S. Bio Energy - Woodbury, Michigan
U.S. Bio Energy - Janesville, Minnesota
Western Wisconsin - Wheeler, Wisconsin

U.S. Commodities
Contact: Mr. Jamie Williams; (612) 486-3880; jwilliams@agmotion.com
Website: www.uscommodities-ag.com
U.S. Commodities is a principal trading and exporting company that specializes in DDGS
among other major U.S. commodities. Please see their website and contact them directly for
more information.

Verasun Energy
Contact: Allan Assmann; (605) 693-6821; aassmann@verasun.com
Website: www.verasun.com
Verasun currently produces and markets DDGS produced at two ethanol plants. Their plant in
Aurora, South Dakota, processes 35 million bushels of corn and produces approximately 290,000
metric tons of DDGS per year. Their plant in Fort Dodge, Iowa, processes 39 million bushels of
corn and produces approximately 319,000 metric tons of DDGS annually.

13 U.S. DDGS Suppliers List

D D G S G lo s s a ry
of Terms

User Handbook

Glossary of Terms
Absorption (in animal nutrition) the movement of nutrients from the digestive tract into the
blood or lymph system.
Acidosis an undesirable condition that can occur in ruminant animals when fed diets high in
readily fermentable carbohydrates such as starch.
Additive an ingredient added in small quantities to the diet for the purpose of fortifying it with
trace nutrients or medicines.
ADF (acid detergent fiber) the fraction of a feedstuff that is not soluble in an acidic detergent
in a laboratory procedure used to determine some components of fiber.
ADG (average daily gain) the rate of body weight gain of an animal expressed on a daily
basis.
ADICP (acid detergent insoluble crude protein) a measure of by-pass or ruminally
undegradable protein of a feed ingredient.
Adipose fat tissue in an animal or carcass.
ADIN (acid detergent insoluble nitrogen) a measure of the insoluble portion of nitrogen in a
feed ingredient; used to calculate ADICP.
Ad libitum (feeding) unlimited access to feed or water.
Aerobic Living or functioning in the presence of oxygen.
Aflatoxin a carcinogenic mycotoxin produced by molds under specific environmental
conditions in growing and stored grains.
Aleurone the protein portion of the endosperm of a seed.
Amino acids nitrogen containing organic molecules that are the building blocks of proteins,
and essential components of nutrition.
Amylase an enzyme that can hydrolyze starch to maltose or glucose.
Anaerobic living or functioning in the absence of oxygen.
Antibiotic a substance produced by a microorganism that has an inhibitory effect on other
microorganisms.
Anti-nutritional factors chemical components of feed ingredients that reduce the nutritional
value of a feed ingredient.
Antioxidant a substance that prevents fats from becoming rancid through oxidation.
Apparent digestibility the amount of a nutrient that is absorbed from the gastrointestinal tract.
Arginine an essential amino acid.
As fed as consumed by the animal.
Ash the residue remaining after complete incineration at 500 to 600 C of a feed; comprised
of metallic oxides.

14 Glossary of Terms

Assay the determination of the chemical components of a feed ingredient or complete feed.
Availability (nutrient) the proportion of a nutrient that is utilized by the animal.
Bacteria single celled living organisms that multiply by simple division. Some are beneficial
while others can cause illness.
Balanced diet a combination of feed ingredients that provide the essential nutrients in the
required amounts to meet the animals needs.
Barrow castrated male pig.
Beta-carotene a precursor source of vitamin A found in some plants and plant products.
Biopsy the removal and examination of tissue or other material from the living body.
Boar intact, uncastrated male pig.
Bran seed coat of cereal grains.
Brewers grains a grain co-product of the brewing industry.
Beer (in ethanol production) a term that refers to the fermented mash that contains ethanol.
By-pass protein protein not broken down by microbes in the rumen and available for further
digestion in the small intestine.
Calorie a unit of energy measurement defined as the amount of heat required to raise the
temperature of one gram of water from 14.5 to 15.5 C.
Carbohydrates organic substances containing carbon, hydrogen and oxygen; many different
kinds are found in plant tissues and include starch, sugar, cellulose, hemicellulose, pectins and
gums.
Carcinogen substances that can cause cancer.
Carotene a yellow organic compound that is a precursor for vitamin A.
Cecum a section of the gastrointestinal tract that follows the small intestine and precedes the
large intestine which contains a significant amount of bacteria that break down fiber not digested
in the small intestine.
Cellulose a polymer of glucose that has a linkage between glucose molecules resistant to
hydrolysis in pigs and poultry, but can be broken down by microbes in the rumen of cattle and
sheep and converted into energy.
Co-product secondary products produced in addition to principle products.
Co-products, ethanol dry mill

The water and solids remaining after distillation of ethanol is called whole
stillage,comprised primarily of water, fiber, protein and fat. This mixture is centrifuged
to separate coarse solids from liquid. The coarse solids are also called wet cake and
contain about 35% dry matter. Wet cake can be sold to local cattle feeders without
drying, or dried to produce dried distillers grains (DDG).

The liquid, now called as thin stillage, goes through an evaporator to remove additional
moisture and the resulting co-product is called condensed distillers solubles which

14 Glossary of Terms

contains approximately 30% dry matter. Condensed distillers solubles can be sold
locally to cattle feeders.

Or, the wet cake can be mixed with condensed distillers solubles and dried to produce
distillers dried grains with solubles (DDGS) which has 88% dry matter.

Colon the lower portion of the large intestine.


Complete feed a single feed mixture which may be used as the only source of the nutrients
required by an animal except water.
Condense a process to reduce an item such as stillage to a denser form by removing moisture.
Condensed distillers solubles see Co-products, ethanol dry milling.
Corn germ meal a co-product from wet milling ethanol plants that contains about 20% crude
protein, 2% fat and 9% fiber with an amino acid balance that makes it a useful feed ingredient in
swine and poultry diets.
Corn steep liquor a high energy liquid co-product produced in wet milling ethanol plants that
is sometimes combined with corn gluten feed or sold separately as a liquid protein source for
beef and dairy cattle.
Crude fat the portion of a feed or feed ingredient that is soluble in ether and is often referred
to as ether extract.
Crude fiber the less digestible portion of a feed ingredient composed of cellulose,
hemicelluolose, lignin and other complex carbohydrates.
Crude protein an estimate of the protein in a feed or feed ingredient, calculated by measuring
the nitrogen content (proteins contain about 16% nitrogen) and multiplying by a factor of 6.25 to
obtain the crude protein percentage.
Cystine a sulfur containing amino acid that can replace up to 50% of the swine requirement for
methionine.
Deamination removal of the amino group from an amino acid.
Deoxynivalenol (DON) a mycotoxin sometimes abbreviated as DON and often referred to as
vomitoxin because it cause reduced feed intake and feed refusal at low concentration in the diet
and vomiting at higher dietary concentrations.
Diet a regulated selection or mixture of feed ingredients provided on a continuous basis or
prescribed schedule.
Digestibility a measure of the extent that the nutrients in a feed are digested and absorbed by
an animal.
Digestible energy gross energy of the feed minus the energy remaining in feces.
Digestion the process that occurs in the gastrointestinal tract of an animal that breaks down
complex nutrients into forms that can be absorbed by an animal.
Distillers dried grains with solubles in drymill ethanol production, a blend of the wet cake
and at least 75% of condensed solubles, dried to a moisture content of approximately 10%. See
Co-products, ethanol dry milling.

14 Glossary of Terms

DL-methionine a source of synthetic methionine.


Dressing percent also known as carcass yield and is the portion of the carcass remaining after
the removal of most internal organs, feet and in most cases, the head.
Drug as defined by the U.S. Food and Drug Administration is a substance intended for the use
in the diagnosis, cure, mitigation, treatment or prevention of disease in humans and other
animals.
Dry grind refers to an ethanol production process that involves grinding the whole corn kernel
and fermenting the resultant corn meal without separating out the component parts.
Dry matter the portion of a feed remaining after water is removed by drying in an oven.
Duodenum the first portion of the small intestine.
Endogenous (in nutrition) compounds such as enzymes and hormones that are internally
produced by the body.
Endosperm part of the seed which provides food for the developing embryo.
Enzyme a protein formed in animal or plant cells that act as biological catalysts to increase the
rate of chemical reactions.
Essential amino acid an amino acid that cannot be synthesized in the body in sufficient
quantities for the bodys needs and must be supplied in the diet.
Ether extract used to measure the amount of fats and oils in feeds and feed ingredients based
on their solubility in ether.
Excreta the products of excretion from an animals body which are primarily feces and urine.
Exogenous (in nutrition) originating from outside of the body.
Fat soluble vitamins vitamins A, D, E and K (menadione).
Fatty acids components of a fat molecule that have different carbon lengths and may be
unsaturated or saturated.
Feed conversion the amount of feed required by an animal for a unit of weight gain.
Fermentation chemical changes brought about by enzymes produced by various
microorganisms.
Flowability the ability of a mass of feed particles or grains to move by gravity out of storage
or transport containers.
Fumonisin a mycotoxin produced by specific molds that can be present in feed ingredients and
reduce animal health and performance.
Fractionation processes used in dry-grind ethanol plants to separate various components of
the corn kernel to improve ethanol yield and produce a variety of co-products with different
nutritional composition.
Gastric refers to the stomach of animals.
Gastrointestinal refers to the stomach and the rest of the intestinal tract used in digestion and
absorption of nutrients.

14 Glossary of Terms

Germ the embryo of a seed.


Glycerol a three carbon component of fat.
Gross energy (GE) the total heat of combustion of a feed or feed ingredient burned in a bomb
calorimeter.
Ground, grinding a mechanical process to reduce particle size by impact, shearing or attrition.
Hulls the outer covering of seed kernels.
Hydrogenation the chemical addition of hydrogen to any unsaturated compound (double
bond), often to fatty acids.
Hydrolysis the chemical process where a compound is split into simpler units with the uptake
of water.
Ileum the lower portion of the small intestine.
International units (IU) an arbitrary scale used to compare the biological activity of some
vitamins.
In vitro refers to things that occur outside the animals body in an artificial environment such
as a test tube.
In vivo refers to things that occur within the animals body.
Iodine number the amount of iodine (in grams) that can be taken up by 100 grams of fat or
fatty acids and is a measure of unsaturation.
Jejunum the middle portion of the small intestine.
Kcal (kilocalorie) is a unit of energy equal to 1,000 calories.
Kjeldahl a method to determine the nitrogen content of a feed ingredient to be used in
calculating and estimating crude protein.
Lesion an unhealthy change in color, size or structure of body tissues.
Lignin an indigestible inorganic component of fiber.
Linoleic acid an essential fatty acid.
Lipid fat.
Liquifaction the process of converting solids into liquid.
Macro (major) minerals minerals present or required in large amounts relative to the animals
requirement and include (calcium, phosphorus, sodium, potassium, magnesium, sulfur and
chloride).
Maillard products a group of poorly digestable protein-carbohydrate complexes that are
produced in feed ingredients that are subjected to significant amounts of heating and are
characterized by darkening of color (browning), burned flavor and burned smell.
Mash a mixture of water and corn meal prior to fermentation in a dry grind ethanol plant.
Meal a grain or feed ingredient or diet that has been ground or otherwise reduced in particle
size.

14 Glossary of Terms

Megacalorie (Mcal) unit of energy equal to 1,000,000 calories or 1,000 kilocalories.


Metabolism the net effect of biochemical changes in the body including building up
(anabolism) and breaking down (catabolism).
Metabolizable energy (ME) gross energy minus fecal and urinary energy from feeding a
complete feed or feed ingredient.
Micro (trace) minerals minerals present or required in small amounts in feeds and feed
ingredients relative to the animals requirement and include (iron, copper, zinc, iodine, selenium
and manganese).
Modified wet cake a blend of partially dried wet distillers grains with condensed distillers
solubles which has dry matter of approximately 50%. See also Co-products, ethanol dry milling.
Monogastric refers to animals such as swine and poultry that have a single, simple stomach.
Mycotoxicosis poisoning of an animal that occurs when consuming significant quantities of
mycotoxins.
Mycotoxins toxic substances produced by specific types of molds under specific types of
climatic and environmental conditions.
NDF (neutral detergent fiber) fiber components in plant and grain cell walls that is
undigestible for monogastric animals.
Net energy (NE) metabolizable energy minus the heat increment.
NFE (nitrogen free extract) is a calculated estimate of the carbohydrate fraction of a feed
ingredient by subtracting moisture, fat, fiber, protein and ash from 100%.
NPN (non-protein nitrogen) any one of a group of nitrogen containing compounds that are
not true proteins that can be precipitated from a solution (e.g. ammonia and urea).
Nutrient any chemical substance that provides nourishment to the body.
Ochratoxin a mycotoxin produced by aspergillus mold which attacks the kidneys, reduces
growth performance and may cause birth defects.
Oleic acid an 18 carbon fatty acid that contains one double bond and is found in animal and
vegetable fat.
Oxidation the union of a substance with oxygen.
Palmitic acid a saturated fatty acid with 16 carbons.
pH a measure of the acidity or alkalinity of a substance; pH = 7 is neutral.
Phytic acid alternative chemical forms of phytate or phytin and are naturally occurring bound
phosphorus compounds in grains and grain co-products that have low digestibility and
availability for monogastric animals.
Phytase is a commercially available enzyme added to monogastric diets to improve
digestibility of phosphorus in the phytic acid form in grains and grain co-products for
monogastric animals.
Ppm (parts per million) a unit of concentration for compounds found in small amounts in feeds
and feed ingredients and is equal to mg/kg.
14 Glossary of Terms

Premix a mixture of the proper proportions of vitamins and trace minerals that when added to
animal diets will meet the requirements for those nutrients.
Propionic acid one of the volatile fatty acids commonly found in rumen contents.
Proximate analysis a combination of analytical procedures used to describe feeds and feed
ingredients.
Rancid a term used to describe fats that have undergone partial decomposition.
Ration a fixed portion of feed, usually expressed as the amount of a diet allowed daily.
Rumen the second compartment of a ruminant stomach.
Ruminant any group of hoofed mammals that have a four compartment, complex stomach and
that chew their cud while ruminating.
Rumination the process of regurgitating previously eaten feed, reswallowing the liquids and
rechewing the solids (cud).
RUP (ruminally undegradable protein) sometimes referred to as by-pass protein, which is
protein that is not degraded by microbes in the rumen and enters the small intestine of ruminants.
Generally, undegradable protein is heat damaged protein.
Saccharification is a process involving hydrolysis (break down) of starch using water and
enzymes in ethanol production.
Saturated fat a fat that contains no fatty acids with double bonds and is solid at room
temperature.
Silage feed resulting from storage and fermentation of wet crops under anaerobic storage
conditions.
Solubles (syrup) see Co-products, ethanol dry milling. In drymill ethanol production, the
liquid portion of stillage separated from the coarse grain by centrifugation and concentrated to
about 30% solids by evaporation.
Starch a white, tasteless, odorless polysaccharide carbohydrate found in large quantities in
corn, sorghum, wheat and other grains that yields glucose upon hydrolysis.
Steeping in wetmill corn processing, a process that involves soaking corn kernels under
controlled conditions for temperature, time and concentration of sulfuric acid and lactic acid to
soften the corn kernel before separating the germ, bran, gluten and starch in wet milling ethanol
production.
Stillage see Co-products, ethanol dry milling.
Stomach the part of the digestive tract where chemical digestion is initiated in most animal
species.
Syrup see Co-products, ethanol dry milling.
TDN (total digestible nutrients) a value that indicates the relative energy value of a feed for
an animal.
Trace minerals see micro minerals.
Ulcer erosion or disintegration of stomach tissue.
14 Glossary of Terms

Unsaturated fat a fat containing from one to three fatty acids that contain one or more double
bonds.
Urea a synthetic, highly concentrated nitrogen product sometimes used as a nitrogen source in
rations for ruminants.
VFA volatile fatty acids which include propionic, acetic and butyric acids.
Volatile fatty acids short chain fatty acids produced in the rumen of cattle and the cecum and
colon of monogastrics that provide energy value to the animal.
Wet cake see Co-products, ethanol dry milling.
Wet distillers grains see Co-products, Ethanol dry milling.
Wet milling processes used to separate various components of the corn kernel into associated
fractions including high fructose corn syrup, corn oil, starch and fiber.
Zearalenone is a mycotoxin produced by fusarium molds under specific climatic and
environmental conditions; it has estrogenic effects, causing reproduction problems in animals.

14 Glossary of Terms

DDGS & Ethanol Production-Related


Website Links

Distillers Grains Technology Council: http://www.distillersgrains.org/grains/.

National Corn Growers Association (NCGA): http://www.ncga.com/.

Renewable Fuels Association (RFA): http://www.ethanolrfa.org/.

Ethanol Producer Magazine: http://www.ethanolproducer.com/index.jsp.

The Online Distillery Network: http://www.distill.com/offlinks.html.

ep Overviews Publishing Inc.: http://www.epoverviews.com/.

University of Minnesota: http://www.ddgs.umn.edu/.

United States Department of Agriculture, Foreign Agriculture Service (FAS):


o http://www.fas.usda.gov/ustrade/
o http://www.fas.usda.gov/

15 Website Links

Você também pode gostar