You are on page 1of 25

International Dairy Journal 21 (2011) 377e401

Contents lists available at ScienceDirect

International Dairy Journal


journal homepage: www.elsevier.com/locate/idairyj

Review

Milk intelligence: Mining milk for bioactive substances associated


with human health
S. Mills a, R.P. Ross a, b, c, C. Hill a, c, d, G.F. Fitzgerald a, c, d, C. Stanton a, b, c, *
a

Food for Health Ireland, Moorepark Food Research Centre, Fermoy, Co. Cork, Ireland
Teagasc, Moorepark Food Research Centre, Fermoy, Co. Cork, Ireland
c
Alimentary Pharmabiotic Centre, University College Cork, Cork, Ireland
d
Department of Microbiology, University College Cork, Ireland
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 18 June 2010
Received in revised form
16 December 2010
Accepted 20 December 2010

Milk has evolved as a complete food for the mammalian nourishment of its young. However, research is
unveiling an ever-accumulating range of bioactivities associated with milk substituents, emphasizing
a role in programming human health. One good example is the increased complexity of carbohydrates in
colostrum that may have a controlling inuence on the selection of gut microbiota in infants at a very early
stage of life. Milk can also affect processes outside the human gut e a proven example is the hypotensive
effect of milk bioactive peptides through angiotensin-I-converting enzyme (ACE) inhibition. However,
even more intriguing is the potential of milk constituents to inuence immune and neural networks
thereby affecting infection rates or mood, respectively. With the advent of bovine and human sequencing
omic technologies, scientists are set to unlock many of the mysteries/mechanisms of how milk is good for
you in ways that up to now were impossible to comprehend.
2011 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .378
Milk proteins and peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .378
2.1.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
2.2.
Antihypertensive peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
2.3.
Antithrombotic peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
2.4.
Opioid peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
2.5.
Casein phosphopeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
2.6.
Immunomodulatory peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
2.7.
Antimicrobial peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
2.8.
Cytomodulatory peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
2.9.
Cysteine protease inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
2.10.
Mammary-associated serum amyloid A3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
3.1.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
3.2.
Prebiotic effect of human milk oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
3.3.
Milk oligosaccharides and infection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 384
3.4.
Milk oligosaccharides and the inflammatory immune response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
3.5.
Sialic acid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
Milk lipids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 386
4.1.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 386
4.2.
Milk fat globule membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 386
4.3.
Phospholipids and cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387

* Corresponding author. Teagasc, Moorepark Food Research Centre, Fermoy, Co. Cork, Ireland. Tel.: 353 25 42606; fax: 353 25 42340.
E-mail address: Catherine.stanton@teagasc.ie (C. Stanton).
0958-6946/$ e see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.idairyj.2010.12.011

378

5.
6.
7.

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

4.4.
Phospholipids and the immune system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
4.5.
Phospholipids and the central nervous system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
4.6.
Polyunsaturated fatty acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 388
4.7.
Conjugated linoleic acid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 388
4.8.
Increasing the conjugated linoleic acid content of milk and dairy foods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 388
4.9.
Increasing eicosapentaenoic acid and docosahexaenoic acid content of milk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
4.10.
Butyrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390
4.11.
Medium chain fatty acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390
4.12.
Antimicrobial lipids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391
Dairy calcium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .391
Microbial diversity inscribed in human milk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 392
Conclusion and outlook for the future . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 392
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393

1. Introduction
The relationship between diet and health is now well known to
be one of the keys to preventing disease and promoting wellbeing.
Indeed, it is on this basis that there has been major growth in the
market for functional foods. These are foods that exert a positive
inuence on human health over and above their nutritive value.
Dairy products hold a major share in this market, owed partly to the
fact that the manufacturing process of fermented dairy products
involves the addition of lactic acid bacteria (LAB), many of which
are now known to possess probiotic properties (Fuller, 1989) or
which themselves produce secondary metabolites with associated
health-promoting effects (Stanton, Ross, Fitzgerald, & Van Sindern,
2005). In addition, it is well known that milk itself is a rich source of
bioactive components that positively inuence host health. This is
probably not surprising given that milk can be considered the
ultimate food, being a complete food for newborns and the sole
food during the early stages of development.
Bovine milk is composed of approximately 5% lactose, 3.2%
protein, 4% lipid and 0.7% mineral salts (Sverin & Wenshui, 2005)
although the exact composition varies in response to a number of
factors including animal nutrition and stage of lactation. While milk
has always been known to have a pivotal role in bone and dental
health (Bonjour, Chevalley, Ferrari, & Rizzoli, 2005; Huncharek,
Muscat, & Kupelnick, 2008; Merritt, Qi, & Shi, 2006; Moore,
Bradlee, Gao, & Singer, 2008), more recent evidence has revealed
that it contains a plethora of components that encode functional
health benets far beyond that expected based on the nutritional
content alone. These encode specic lipids, complex carbohydrates
and peptide sequences encrypted within milk proteins that exert
activities that affect blood pressure or mood.
Milk provides the ultimate model for functional food development being endowed with nutritional, immunological and
biologically active components (Wong, de Souza, Kendall, Emam, &
Jenkins, 2006). Indeed, many milk components are now being
exploited as health-promoting ingredients in other food systems.
An extreme example is the production of recombinant human
milk proteins in plants with a view to developing food products
with enhanced nutrition for formula-fed infants as well as older
children and adults (Arakawa, Chong, Slattery, & Langridge, 1999).
Milk oligosaccharides have been shown to modulate both the
intestinal microbiota and the immune response, thus playing key
roles in defence against infection and allergy development in
formula-fed infants. In addition, milk lipids also possess therapeutic properties and efforts have been made to enhance the
presence of certain fatty acids in bovine milk, in particular the
bioactive lipid conjugated linoleic acid (CLA). Another particularly
exciting development is the Milk Genome Consortium that is
devoted to the understanding of how the biomolecules of milk are

manufactured and regulated and how they achieve their specic


functions (Ward & German, 2004).
This review aims to discuss the specic components of bovine
and human milk that have been associated with human health
through mechanisms that exceed basic nutrition. In particular, we
have focused on some specic examples derived from the protein,
namely bioactive peptides, the polysaccharide and lipid constituents of milk as well as dairy calcium. Finally, this review looks at
some exciting data highlighting the microbial diversity inscribed in
human milk for the benet of the infant.
2. Milk proteins and peptides
2.1. Introduction
The major protein fractions in bovine milk consist of caseins
and whey proteins including immunoglobulins, a-lactalbumin,
b-lactoglobulin, bovine serum albumin, immunoglobulin, lactoferrin
as well as proteoseepeptone fractions and transferrin (Table 1).
Lower amounts of other minor proteins and peptides also exist that
have for example, hormonal or other physiological activities, e.g.,
hormone releasing factor and the endogenous antibacterial system
(Sverin & Wenshui, 2005). In the last couple of decades, much
Table 1
Concentration and biological activity of major bovine milk proteins.a
Protein

Concentration
(g L1)

Function

Total caseins

26.0

a-casein
b-casein
k-casein
b-Lactoglobulin

13.0
9.3
3.3
6.3
3.2

a-Lactalbumin

1.2

Immunoglobulins
(A, M, and G)
Serum albumin
Lactoferrin

0.7

Ion carrier (Ca, PO4, Fe, Zn, Cu),


precursor of bioactive peptides
Precursor of bioactive peptides
Precursor of bioactive peptides
Precursor of bioactive peptides
Anticarcinogenic, weight management
Retinol carrier, binding fatty acids,
possible antioxidant
Lactose synthesis in mammary gland,
Ca carrier, immunomodulation
Immune protection

0.4
0.1

Lactoperoxidase
Lysozyme

0.003
0.0004

Miscellaneous
Proteoseepeptone

0.8
1.2

Glycomacropeptide

1.2

Total whey protein

Antimicrobial, antioxidative,
immunomodulation, iron absorption
Antimicrobial
Antimicrobial, synergistic effect
with immunoglobulins and lactoferrin
Function unknown but precursor
of bioactive protein and peptide in vitro
Antiviral, bidogenic

Adapted from Sverin and Wenshui (2005); see Walstra and Jenness (1984),
Yamauchi (1992) and Korhonen, Pihlanto-Leppala, Rantamaki, and Tupasela
(1998) for more details.

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

research has concentrated on the therapeutic potential of milk


proteins (Madureira, Pereira, Gomes, Pintado, & Malcata, 2007;
Teschemacher, Koch, & Brantl, 1997; Zimecki & Kruzel, 2007). For
example, it has been demonstrated that whey protein is superior to
other dietary proteins for suppression of tumor development, due to
components such as lactoferrin, b-lactoglobulin, a-lactalbumin and
serum albumin (Parodi, 2007). Indeed, under acidic conditions in the
presence of oleic acid, a-lactalbumin and oleic acid form a complex,
termed HAMLET (human a-lactalbumin made lethal to tumor cells)
which has been shown to inhibit a wide array of tumors via an
apoptosis-like event (Svanborg, Agerstam, Aronson, Bjerkvig, &
Duringer, 2003). The bovine counterpart of HAMLET, termed BAMLET, was recently shown to display potent cytotoxic activity against
eight cancer cell lines tested via a mechanism involving lysosomal
membrane permeabilisation (Rammer et al., 2010). A specialized
whey fraction (high in leucine, bioactive peptides and milk calcium)
with the commercial name Prolibra has been shown to promote
loss of body fat mass and greater preservation of lean muscle in
a randomized human clinical trial lasting 12 weeks (Frestedt, Zenk,
Kuskowski, Ward, & Bastian, 2008). Indeed, the utilisation of whey
as a functional food ingredient for weight management is attracting
much interest (Luhovyy, Akhavan, & Anderson, 2007). The special
membrane structure surrounding milk micro-lipid droplets,
composed of a lipid bilayer and proteins, termed the milk fat globule
membrane (MFGM), has also shown potential as a therapeutic agent
against many pathological conditions (Spitsberg, 2005). One of the
proteins isolated from MFGM, termed fatty acid binding protein
(FABP) has been shown to inhibit some breast cancer cell lines
(Spitsberg & Gorewit, 1997a, 2002; Spitsberg, Matitashvili, &
Gorewit, 1995). In addition, the onco-suppressor protein BRCA1
has also been found in both bovine and human MFGM (Spitsberg &
Gorewit, 1997b). Wang, Hirmo, Millen, and Wadstrom (2001c) also
demonstrated that glycoproteins of MFGM were capable of inhibiting the infection of Helicobacter pylori in a BALB/cA mouse model,
and the hemagglutination and adhesion of H. pylori in HeLa S3

379

monolayers. A 19 kDa protein derived from the degradation of proteoseepeptone component 3 following fermentation of fat-free milk
by a number of LAB enhanced monoclonal antibody production of
human hybridoma cells and stimulated immunoglobulin production
of human peripheral blood cells (Sugahara et al., 2005). Recently, the
basic protein fraction of milk termed milk basic protein (MBP) has
also been shown to have a direct effect on the strengthening of bones
in healthy human volunteers (Toba et al., 2001; Uenishi et al., 2007).
Indeed, various proteomic approaches are now being applied to
decipher the subcellular organization of the MFGM as discussed in
a review by Cavaletto, Giuffrida, and Conti (2008) and new studies
are revealing that differences in protein expression even occur in
MFGM between colostrum and day 7 milk (Reinhardt & Lippolis,
2008).
Bioactive peptides which exert numerous physiological
responses can also be generated from milk proteins in the gastrointestinal tract. Such bioactive peptides are latent or encrypted
within native protein precursors, thus proteolysis is required for
their release (Gobbetti, Minervini, & Rizzello, 2004). Indeed,
bioactive peptides have been generated from most of the major
proteins in both bovine and human milk (Sverin & Wenshui,
2005). Bioactive milk peptides were rst described in 1950, when
Mellander (1950) reported that ingestion of casein-derived phosphorylated peptides led to enhanced vitamin D-independent
calcication in rachitic infants. While bioactive peptides can be
generated from a variety of foods, milk proteins are generally
regarded as a very rich source and as a result have become
fundamental constituents of several commercially available functional food products and ingredients (Table 2). Bioactive peptides
from milk can be divided into the following categories based on
their physiological effect on the body or the protein from which
they have been derived; antihypertensive, antithrombotic, opioid,
casein phosphopeptides (CPPs), antimicrobial, cytomodulatory,
immunomodulatory, and miscellaneous peptides (Hayes, Stanton,
Fitzgerald, & Ross, 2007a).

Table 2
Milk-derived bioactive peptides in commercially available functional foods and ingredients.a
Bioactive peptide

Health claim

Product type

Brand name

Manufacturer

VPP, IPP from b-casein and k-casein


VPP, IPP from b-casein and k-casein

Reduction of blood pressure


Reduction of blood pressure

Calpis
Evolus

Calpis Co., Japan


Valio, Finland

Whey peptides

Reduction of blood pressure

BioZate

Davisco, USA

Casein-derived dodecapeptide FFVAPFPEVFGK


Casein-derived dodecapeptide FFVAPFPEVFGK
Glycomacropeptide k-casein f (106-169)

C12 Peption
Casein DP Peptio Drink
BioPURE-GMP

DMV, The Netherlands


Kanebo, Japan
Davisco, USA

as1-casein f (91e100) YLGYLEQLLR

Reduction of blood pressure


Reduction of blood pressure
Anticariogenic, antimicrobial,
antithrombic
Reduce stress

Sour milk
Fermented milk,
calcium enriched
Hydrolysed whey
protein isolate
Ingredient
Soft drink
Whey protein isolate

CSPHP ProDiet F200

Ingredia, France

Casein phosphopeptide
Casein phosphopeptide
Casein phosphopeptide
Casein phosphopeptide
Glutamine-rich peptides

Helps mineral absorption


Helps mineral absorption
Helps mineral absorption
Helps mineral absorption
Immunomodulatory

Capolac
Tekkotsu Inryou
Kotsu Kotsu Calcium
CE90CPP
Glutamine Peptide

Arla Foods, Denmark


Suntory, Japan
Asahi, Japan
DMV, The Netherlands
DMV, The Netherlands

as1-casein f (1e6) RPKHPI, f (1e7)

Reduction of blood pressure

Milk drink,
confectionary
Ingredient
Soft drink
Soft drink
Ingredient
Dry milk protein
hydrolysate
Fermented low
fat hard cheese
Ingredient
Ingredient

Festivo
Cysteine Peptide
PeptoPro

MTT Agrifood
Research, Finland
DMV, The Netherlands
DSM, The Netherlands

Ingredient

Vivinal Alpha

Chewing gum
Margarine

Recaldent
Evolus Double
Effect Spread

RPKHPIK, f (1e9) RPKHPIKHQ,


Milk-derived peptide
Casein-derived peptide
Whey-derived peptide

Boost energy, improve sleep quality


Improves athletic performance
and muscle recovery
Aids relaxation and sleep

Casein phosphopeptides
Milk-derived peptides

Anticariogenic
Reduction of blood pressure

Adapted and updated from Korhonen (2009), Korhonen and Pihlanto (2006), Hartmann and Meisel (2007); f denotes peptide fragment.

Borcula Domo Ingredients


(BDI), The Netherlands
Cadbury Enterprises
Valio, Finland

380

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

2.2. Antihypertensive peptides


Antihypertensive peptides inhibit ACE, a key enzyme involved in
the regulation of blood pressure (BP). These ACE inhibitors are
thought to be competitive inhibitors of ACE by preventing ACE from
synthesising the potent vasoconstrictor, angiotensin-II. In addition,
ACE also hydrolyses bradykinin, a vasodilator (Seppo, Jauhiainen,
Poussa, & Korpela, 2003). Milk fermentations using LAB or their
proteinases have been described as a strategy to release ACEinhibitory peptides from milk proteins, especially caseins (Hayes,
Ross, Fitzgerald, Hill, & Stanton, 2006; Hayes et al., 2007a, 2007b).
ACE inhibitory and antihypertensive peptides from dairy origin
contain up to 10 amino acids (Korhonen & Pihlanto, 2003; Meisel &
Bockelmann, 1999; Pihlanto, 2001; Saito, 2008). These peptides
have already been marketed (Table 2). For example, the ACEinhibitory peptides Val-Pro-Pro (Clare, Catignani, & Swaisgood,
2003; Hamel, Kielwein, & Teschemacher, 1985; Juillard et al.,
1995) and Ile-Pro-Pro (Chabance et al., 1995; Drouet et al., 1990)
derived from casein following fermentation with the strains
Lactobacillus helveticus and Saccharomyces cerevisiae, respectively,
are found in the commercial sour milk product, Calpis (Calpis, Co.
Ltd., Tokyo). The antihypertensive effect of Calpis was shown to
yield a systolic BP decrease of 17.7 mm Hg following administration
of 5 mL kg1 body weight of Calpis sour milk drink over an 8-h
period in spontaneously hypertensive rats (Yamamoto, Akino, &
Takano, 1994). Likewise, a signicant reduction in BP was
obtained in mildly hypertensive patients following oral consumption of 95 mL Calpis over an 8-week period (Hata et al., 1996).
More recently, casein hydrolysate containing Val-Pro-Pro and IlePro-Pro has been shown to improve vascular endothelial dysfunction in subjects with mild hypertension (Hirota et al., 2007).
Various traditional Swiss cheese varieties have also recently been
shown to contain, on average, similar concentrations of Val-Pro-Pro
and Ile-Pro-Pro to the recently developed fermented milk products
with BP-lowering properties (Table 2) (Butikofer, Meyer, Sieber,
Walther, & Wechsler, 2008). Moreover, bioavailability studies by
Foltz et al. (2007) demonstrated that the tri-peptide, Ile-Pro-Pro,
selectively escapes from intestinal degradation and reaches the
circulation undegraded. It was also demonstrated that Val-Pro-Pro
and Ile-Pro-Pro have the potential to inhibit ACE in a very similar
fashion to the current synthetic ACE inhibitors Captopril, Enalaprilat and Lisinopril, by hydrogen-bonding with similar residues in
the ACE catalytic site (Pina & Roque, 2008). Interestingly, antihypertensive peptides have now been produced using recombinant
DNA technologies, whereby recombinant fusion proteins have been
expressed in Escherichia coli, which are then puried and cleaved by
proteinase from a selected strain of L. helveticus (Losacco, Gallerani,
Gobbetti, Minervini, & De Leo, 2007). This technology will enable
scientists to design new peptides with even stronger inhibitory
activity and new therapeutic properties. Additionally, ACE inhibitors derived from casein or casokinins have been identied within
the sequences of human b- and k-casein (Kohmura, Nio, & Ariyoshi,
1990; Kohmura et al., 1989) and have also been generated by tryptic
digestion of a- and b-casein (Maruyama & Suzuki, 1982). Interestingly, as suggested by Hayes et al. (2007a) delivery of these
bioactive peptides to target organs may be enhanced by using
larger peptide sequences, which upon in vivo digestion with
proteinase and peptidase activities should release the antihypertensive peptide in the gastrointestinal tract.
2.3. Antithrombotic peptides
Antithrombotic peptides that interfere with the formation of
thrombi have also been identied in milk (Fiat, Migliore, & Jolles,
1993; Jauhiainen & Korpela, 2007; Zimecki & Kruzel, 2007).

Indeed, enzymatic hydrolysis of k-casein has resulted in some of the


most antithrombotic peptides to date from food sources. Thrombosis is a pathological condition that results in clots or thrombus
formation in arteries, veins or the chambers in the heart. From
a dairy point of view, it is interesting to note that comparisons can be
drawn between the mechanisms involved in milk clotting, dened
by the interaction of k-casein with chymosin and the mechanisms of
blood clotting, dened by the interaction of brinogen with
thrombin (Jolles, 1975; Jolles & Henschen, 1982; Rutherfurd & Gill,
2000). Having found structural similarities between bovine
k-casein and the human brinogen g-chain, Jolles, LoucheuxLefebvre, and Henschen (1978) hypothesised that both may have
evolved from a common ancestor during the past 450 million years.
Indeed, Jolles et al. (1978) demonstrated that several long k-casein
sections corresponding to 80% of the whole protein molecule have
counterparts in the brinogen g-chain, with 31e42% of the amino
acid residues occupying the same positions. In addition, several
common features were observed in the secondary structures of the
k-caseins and the brinogen g-chain (Jolles et al., 1978). Moreover,
the C-terminal sequence of the brinogen g-chain fragment (f)
(400e411) corresponding to the sequence HHLGGAKQAGDV
inhibits platelet aggregation and brinogen binding to ADP-activated platelets (Andrieux et al., 1989). The main antithrombotic
peptide isolated from bovine k-casein corresponding to f (106e116)
with the amino acid sequence MAIPPKKNQDK termed casoplatelin,
was also shown to inhibit ADP-induced platelet aggregation and
brinogen binding in a concentration-dependant manner (Jolles
et al., 1986) compounding the functional similarity between kcasein and brinogen.
Interestingly, it is thought that milk protein-derived antithrombotic peptides are absorbed into the bloodstream. For example,
two peptides from human and bovine k-caseinoglycopeptide have
been identied in the plasma of 5 day old newborns following
ingestion of a cow milk-based formula (Chabance et al., 1995). It is
also interesting to note that k-casein f (152e160) and f (155e160)
isolated as ACE inhibitors may also possess antithrombotic activity
(Gobbetti et al., 2004).
2.4. Opioid peptides
Opioid peptides are those peptides having pharmacological
similarities to opium. The caseins (as1-, as2,- b- and k-) and whey
proteins are potential sources of such opioid peptides. However, the
major opioid peptides are fragments of b-casein, called b-casomorphins (Clare & Swaisgood, 2000; Teschemacher et al., 1997). Similar
proteins have also been reported from human b-casein fractions
(Greenberg, Groves, & Dower, 1984; Yoshikawa, Yoshimura, & Chiba,
1984). On the other hand, all k-casein fragments, known as casoxins
behave as opioid antagonists (Sverin & Wenshui, 2005). Opioid
peptides have also been found encrypted within the primary
sequence of whey proteins such as lactoferrin, b-lactoglobulin, and
bovine serum albumin (Belem, Gibbs, & Lee, 1999; Rokka, Syvaoja,
Tuominen, & Korhonen, 1997). b-casomorphins are resistant to the
action of gastrointestinal enzymes (Read, Lord, Brantl, & Koch, 1990)
and have been associated with the following activities; antihypertensive, immunomodulatory, antidepressant, anti-secretory and
anti-diarrheal (Pihlanto, 2001). Opioid peptides are thought to be
biologically very potent; potentially, micromolar amounts may be
sufcient to exert physiological effects (Meisel, 1997; Meisel &
Fitzgerald, 2000). Systemic administration of a low dose
(1 mg kg1, i.p.) of bovine b-casomorphin-5 was shown to improve
the disturbance of learning and memory in mice (Sakaguchi, Koseki,
Wakamatsu, & Matsumura, 2006). In addition, b-casomorphin-7 was
shown to signicantly contribute to mucin production from both rat
and human intestinal mucin-producing cells using real time-PCR and

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

ELISA studies (Zoghbi et al., 2006). Since intestinal mucins play


a protective role in the gut, consumption of products containing bcasomorphin-7 could help to improve intestinal health by preventing the adherence of pathogens to the intestinal surface and thus
eliminate the onset of intestinal infections.
2.5. Casein phosphopeptides
CPPs refer to casein-derived phosphorylated peptides, which
contain single and multiple phosphoryl residues and are released
by enzymatic hydrolysis of as1-, as2-, b- and k-caseins both in vivo
and in vitro (Adamson & Reynolds, 1995, 1996; Clare & Swaisgood,
2000). These peptides efciently bind divalent cations such as Fe,
Mn, Cu and Se, due to the high content of negative charge and thus
may act as biocarriers. Several studies have investigated if CPPs in
the diet could increase calcium absorption. However, the results
have been conicting between human and animal studies (ScholzAhrens & Schrezenmeir, 2000). For example, 1.0 mg of CPP
administered extrinsically by gastric intubation to young male rats,
dramatically increased calcium absorption from calcium-fortied
milk compared with the control (Tsuchita, Suzuki, & Kuwata, 2001).
Also using an animal model, Erba, Ciappellano, and Testolin (2002)
demonstrated a similar positive effect but found that the ratios
between CPPs and calcium in the rat intestinal lumen was highly
benecial, with a ratio of 15 identied as the most efcient at
increasing mineral transport, where values above and below were
much less effective. Most recently, the addition of CPP was shown
to have a benecial effect on the absorption of calcium in growing
rats from CaCO3 added to both bovine and caprine milk (MoraGutierrez, Farrell, Attaie, McWhinney, & Wang, 2007). It was also
demonstrated that CPPs induce the inux of calcium into human
HT-29 cells, a rise which was dependant on CPP dose and did not
inuence ATP-induced release of calcium from intracellular stores
(Ferraretto, Signorile, Gravaghi, Fiorilli, & Tettamanti, 2001).
Human studies have been more controversial with regards the
positive effects of CPPs on calcium transport however. Indeed,
administration of 1 g of CPPs did not affect calcium metabolism
acutely in nine postmenopausal women following adminstration of
either CPP-enriched milk or CPP-enriched fermented milk (Narva,
Karkkainen, Poussa, Lamberg-Allardt, & Korpela, 2003). While the
effect of high doses of 2 well-dened CPP-enriched preparations on
15 volunteers consuming calcium lactate drinks did indicate
a positive effect of the CPPs on calcium absorption, it was concluded
that the differences in calcium absorption were unlikely to have any
biological signicance (Teucher et al., 2006). On the other hand,
CPPs have been shown to have anticariogenic properties and to
prevent enamel demineralisation (Aimutis, 2004; Grenby, Andrews,
Mistry, & Williams, 2001). Most recently, a highly diluted CPPamorphous calcium phosphate (CPP-ACP) preparation demonstrated potential as a tooth transport medium by preserving the
viability of an L929 broblastic cell line (Cehreli, Gurpinar, Onur, &
Dagli, 2008). CPPs are currently being used in commercial products including mineral drinks, nutritional supplements for children,
confectionary and products for dental care (Cross, Huq, & Reynolds,
2007; Luo & Wong, 2005) (Table 2).
2.6. Immunomodulatory peptides
Several immunomodulatory peptides have been found in bovine
and human milk (Gill, Doull, Rutherfurd, & Cross, 2000; Politis &
Chronopoulou, 2008). Indeed, casein-derived immunopeptides
have been shown to stimulate the phagocytic activities of both
human and murine macrophages and to protect against Klebsiella
pneumoniae infection in mice (Smacchi & Gobetti, 2000). Trypsinhydrolysed human milk was also found to contain

381

immunostimulating activity (Jolles et al., 1981). Immunomodulating peptides may play a key role in the proliferation and maturation
of T cells and natural killer cells in the newborn providing protection against a large number of bacteria, particularly enteric bacteria
(Clare et al., 2003; Clare & Swaisgood, 2000). Interestingly, the
peptide isracidin, f (1e23) of as1-casein obtained from the action of
chymosin, was shown to display antibacterial activity against
Staphylococcus aureus and Candida albicans, and also gave protection against mastitis following injection into the udder of sheep and
cow (Lahov & Regelson, 1996). Most recently, forced feeding
experiments on healthy animals with a malleable protein matrix
(MPM), composed of whey fermented by a LAB, capsular polysaccharides, vitamins, minerals and peptides generated during the
fermentation process was shown to stimulate the immune system
as seen by the increase in polymorphonuclear cell counts and
intracellular glutathione levels (Beaulieu, Dubuc, Beaudet, Dupont,
& Lemieux, 2007).
2.7. Antimicrobial peptides
Milk is a rich source of antimicrobial proteins and peptides,
capable of exerting antimicrobial activities comparable to antibiotics (Clare et al., 2003; Joerger, 2003; Lopez-Exposito & Recio,
2008; Sverin & Wenshui, 2005). This effect is due to the synergistic activity of naturally occurring peptides and defence proteins
besides immunoglobulins, such as lactoferrin, lactoperoxidase and
lysozyme and is greater than any individual contribution (Clare
et al., 2003; Sverin & Wenshui, 2005). The potent properties of
these agents can be reected by the example of bovine lactoferrin,
which has displayed strong antiviral activity against HIV and the
human cytomegalovirus, the latter of which is thought to act
synergistically in patients with acquired immunodeciency
syndrome (Floris, Recio, Berkhout, & Visser, 2003).
Most antimicrobial peptides reported to date have been released
from the parent milk protein following heat and/or alkali treatment
or enzymatic hydrolysis. For example, both bovine and human lactoferricin, corresponding to bovine lactoferrin f (17e41) and human
lactoferrin f (1e47) display antimicrobial activity against a broad
range of Gram-positive and Gram-negative bacteria, including the
food pathogen, Listeria monocytogenes (Clare et al., 2003; Floris et al.,
2003). Casein-derived antimicrobial peptides resulting after
fermentation by Lactobacillus acidophilus DPC6026 demonstrated
antibacterial activity against the pathogenic strains E. coli and
Enterobacter sakazakii, the latter of which can be problematic in milkbased infant formulas (Hayes et al., 2006). A synthetic 23-residue
peptide was generated from proteoseepeptone 3 f (113e135) termed
lactophoricin, which displayed antimicrobial activity against Grampositive and Gram-negative bacteria (Campagna, Mathot, Fleury,
Girardet, & Gaillard, 2004).
2.8. Cytomodulatory peptides
Cytomodulatory peptides which have been shown to inhibit cell
growth and stimulate the activity of immunocompetent and
neonatal intestinal cells, respectively, have been isolated from
a variety of fermented dairy products (Hayes et al., 2007a; Parodi,
2007). Several studies have displayed the cytomodulatory effects
of milk either in vivo or in vitro (Bif, Coradini, Larsen, Riva, & Di
Fronzo, 1997; Kampa et al., 1997; LeBlanc, Matar, Valdez, LeBlanc,
& Perdigon, 2002; de Moreno de LeBlanc, Matar, LeBlanc, &
Perdigon, 2005; Roy, Watanabe, & Tamai, 1999). Such peptides
can be classied as potential anti-carcinogens and have been
reported to act as specic signals that can inhibit the viability of
cancer cells (Meisel, 2005). Whey protein concentrate has been
attributed with anticancer activities, linked to its ability to donate

382

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

cysteine to the glutathione antioxidant system (GSH), one of the


principal cellular protection mechanisms (Bounous, 2000).
Remarkably, the combination of a whey protein isolate, termed
Immunocal, with the anti-cancer drug baicalein enhanced the
cytotoxicity of baicalein by inducing more apoptosis in the human
hepatoma cell line Hep G2 (Tsai, Chang, Chen, & Lu, 2000). Most
recently, waste whey peptides from Mozzarella di Bufala Campana
cheese were shown to exert a signicant anti-proliferative effect on
a Caco-2 cell line (De Simone et al., 2009).
2.9. Cysteine protease inhibitors
The activity of some milk proteins as cysteine protease inhibitors
may also provide valuable therapeutic activities. Indeed, the
cysteine protease inhibitors of milk have been associated with
bactericidal activity and protection from bone resorption (Sano
et al., 2005). During bone resorption, osteoclasts secrete proteases
to digest bone matrix proteins such as collagen (Drake et al., 1996).
A 12 kDa cysteine protease inhibitor was puried from MBP and was
subsequently identied as bovine cystatin C (Matsuoka et al., 2002).
Moreover, both lactoferrin and b-casein have demonstrated
cysteine protease inhibition (Ohashi et al., 2003). Indeed, the 17mer at the C-terminus of lactoferrin showed 90% homology with the
sequences of a common active site of the cystatin family. The ability
to inhibit cysteine proteases has thus been postulated as an antimicrobial mechanism for both milk proteins (Ohashi et al., 2003).

factor (Gyorgy, Jeanloz, Von Nicholai, & Zilliken, 1974). Since then,
a large array of milk oligosaccharides was identied and emerging
techniques enable complete analysis of the oligosaccharide content
of human milk in particular (Ninoneuvo et al., 2006). Indeed,
oligosaccharides are a major component of human milk (Table 3). It
has been estimated that human milk contains 7e12 g L1 oligosaccharides (Boehm & Stahl, 2007), about 5e10% of the lactose
concentration. Comparison of the oligosaccharide content of
human milk with milk from other species indicates that human
milk is unique in terms of the complexity and content of its
oligosaccharides (Boehm & Stahl, 2007; Egge, 1993; Kunz, Rudloff,
Baier, Klein, & Strobel, 2000). Human milk oligosaccharides
(HMOs) achieve maximum concentration in the colostrum (above
20 g L1) after which they reach stability in the mature milk (about
12e14 g L1) (Coppa et al., 1993, 1999). Bovine milk on the other
hand harbours very low levels of oligosaccharides, in the region of
1 g L1 (Montreuil, 1960) (Table 3).
The monomers of milk oligosaccharides include D-glucose, Dgalactose, N-acetylglucosamine, L-fucose and sialic acid (N-acetylneuraminic acid), thus a large number of core structures can be
formed (Stahl, Thurl, Zeng, Karas, & Hillenkamp, 1994). However,
the variety is further increased due to a-glycosidic linkages of
fucose and/or sialic acid to the core molecules. The linkage of fucose
is genetically connected to the secretor/Lewis blood group status of
the individual mother (Boehm & Stahl, 2007; Kunz et al., 2000).
Indeed, the addition of fucose to an oligosaccharide by an a1,2
linkage is catalysed primarily by the secretor gene, Se (FUT2); the

2.10. Mammary-associated serum amyloid A3


Another example of a milk antimicrobial peptide is one which is
derived from mammary-associated serum amyloid A3 (M-SAA3),
which is highly abundant in colostrum of mammals but is present in
lower levels in mature milk (McDonald, Larson, Mack, & Weber,
2001). The proposed action of M-SAA3 is not through direct
killing, but via the induction of a host response affecting pathogen
binding to the gut surface. N-terminal sequencing data revealed
a conserved amino acid motif in the M-SAA3 proteins that has been
shown to enhance the mRNA expression of the specic mucin MUC3
by human intestinal cells (Larson, Wei, Weber, Mack, & McDonald,
2003), a phenomenon already discussed for the opioid peptide bcasomorphin-7. MUC3 prevents enteropathogen adherence to
epithelial cells. Indeed, peptides containing the species-conserved
amino acid motif have been shown to reduce the adherence of
enteropathogenic E. coli to human intestinal cells in culture by up to
80%. However, while the human 42-mer M-SAA3 protein and the
10-mer peptide (f (2e11) of the N-terminal region) were recently
shown to exhibit anti-infective activity in vitro, neither the protein
nor the peptide prevented enteric infection in the murine models
tested, which may be due to enzymatic degradation in the gastrointestinal tract or inactivity in the murine model (Gardiner et al.,
2009).
3. Oligosaccharides
3.1. Introduction
As early as 1905, the predominance of lactobacilli in the intestinal microbiota of breast-fed infants (in particular Lactobacillus
bidus now known as Bidobacterium bidum) was noted (Tissier,
1905). In contrast, E. coli was found to dominate in both precociously weaned infants and adults. It was hypothesised that breastfeeding afforded a protection against pathogen colonisation of the
gut, by promoting the growth and colonisation of L. bidus. Almost
70 years later, a glycan isolated from human milk was shown to
stimulate the growth of L. bidus, and was termed the bidus

Table 3
Compositions and concentrations of oligosaccharides in bovine milk, bovine colostrum and human milk.a
Oligosaccharide

Bovine milk Bovine colostrum


(g L1)
(g L-1)

Lactose

40e50

40e50

Trace
e
e
e
e

e
e
e
e
e
NRb
NR
NR
NR
NR
NR
NR

0.03e0.06

0.019
0.095

Trace
Trace
Trace

e
e
e
0.047
Trace (3
0.028
Trace (3
Trace (1
Trace (2
NR
e

Neutral oligosaccharides
Lacto-N-tetraose
Lacto-N-fucopentaose I
Lacto-N-fucopentaose II
Lacto-N-fucopentaose III
Lacto-N-difucohexaose I
Lacto-N-novopentaose
N-acetylgalactosaminylglucose
N-acetylgalactosyl-lactose
a30 -galactosyl-lactose
b30 -galactosyl-lactose
60 -galactosyl-lactose
N-acetyl-lactoseamine
N-acetyl-galactosaminyl-lactose
Acidic oligosaccharides
NeuAc(a2e6)lactose
NeuAc(a2e3)lactose
N-glycoylneuraminyl-lactose
NeuAc-lacto-N-tetraose a
NeuAc-lacto-N-tetraose c
NeuAc2-lacto-N-tetraose
6-Sialyl-lactosamine
3-Sialyl-galactosyl-lactose
Disialyl-lactose
Sialyl-lactose-1-phosphate
Sialyl-lactose-6-phosphate
3-Glycolyl-neuraminyl-lactose
6-Glycolyl-neuraminyl-lactose
GlcNAcb(1e3)Galb(1e4)
GlcNAcb(1e3)Galb(1e4)Glc

Human milk
(g L-1)
55e70

0.5e1.5
1.2e1.7
0.3e1.0
0.01e0.2
0.1e0.2

0.3e0.5
0.1e0.3
0.03e0.2
0.1e0.6
0.2e0.6

mmol L1)
mmol L1)
mmol L1)
mmol L1)
e

a
Adapted from Mehra and Kelly (2006); for details see Kunz and Rudloff (2002),
Gopal and Gill (2000), Nakamura and Urashima (2004).
b
NR: oligosaccharides detected and structurally characterized, but concentration
not reported.

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

addition of fucose by an a1,3 or a1,4 linkage is catalysed by fucosyltransferases produced by the Lewis gene, Le (FUT3) or other a1,3
transferase genes (FUT4, 5, 6, 7, and 9) of this family (Oriol,
Mollicone, Cailleau, Balanzino, & Breton, 1999). Synthesis of all of
these molecules transpires in the breast starting with lactose at the
reducing end (Boehm & Stahl, 2003; Kunz et al., 2000). Interestingly, milk fucosyloligosaccharide expression has been shown to
change both qualitatively and quantitatively over the course of
lactation showing a relative decrease over the rst year (Chaturvedi
et al., 2001). Moreover, concentrations of certain HMOs in colostrum have recently been shown to uctuate even between day 1
and day 3 of lactation (Asakuma et al., 2007; 2008). HMOs, which
can be divided into neutral and acidic fractions, have been shown to
exert a variety of physiological functions (Coppa, Bruni, Morelli,
Soldi, & Gabrielli, 2004; Coppa, Zampini, Galeazzi, & Gabrielli,
2006; German, Freeman, Lebrilla, & Mills, 2008; Hickey, 2009;
Kunz & Rudloff, 2008; Kunz et al., 2000; Newburg, 2009;
Newburg, Ruiz-Palacois, & Morrow, 2005). Moreover, a recent
study has also demonstrated that HMOs have the potential to
inuence various stages in gastrointestinal development in vitro by
inducing growth inhibition or differentiation depending on the
oligosaccharide and cell line tested (Kuntz, Rudloff, & Kunz, 2008).
3.2. Prebiotic effect of human milk oligosaccharides
Interestingly, an analysis of the fecal microbial communities of
106 individual mammals indicated that host diet had a major
inuence on bacterial diversity (Ley et al., 2008). In keeping with
this notion, it has been suggested that a clue to the ability of the
host to select for different intestinal bacteria may come from
understanding the role of HMOs in the infant gut as prebiotics
(Ninonuevo et al., 2007). Indeed, selection pressures in the gut,
such as carbohydrate availability have shaped the intestinal metagenome such that the gut microbiota in adults is dominated by
members of just 2 divisions of bacteria- the Bacteroidetes and the
Firmicutes and one member of Archaea, Methanobrevibacter smithii
(Eckburg et al., 2005). Interestingly, a metagenomics approach
studying genomic and genetic diversity in our distal gut microbiome demonstrated the abundance of Bidobacterium phylotypes
(Gill et al., 2006). Moreover, a recently dened minimal human gut
metagenome, described as the bacterial functions involved in gut
homeostasis, includes functions involved in compex polysaccharide
degradation and synthesis of short chain fatty acids (SCFAs)
amongst others (Qin et al., 2010).
HMOs are one of the rst selective pressures enforced on the gut
microbiota and in doing so exert a prebiotic effect which has been
conrmed in many studies (Coppa et al., 2006). Remarkably, at
birth the human gastrointestinal tract is sterile however, various
factors affect microbial colonisation (Mackie, Sghir, & Gaskins,
1999; Mountzouris & Gibson, 2003; Orrhage & Nord, 1999)
including mode of delivery (Gronlund, Lehtonen, Eerola, & Kero,
1999), type of feeding (Orrhage & Nord, 1999) and antibiotic
therapy (Burman, Berglund, Huovinen, & Tullus, 1999; Kalenic,
Francetic, Polak, Zele-Starcevic, & Bencic, 1993). The intestinal
microbiota of the breast-fed infant is represented mainly by bidobacteria and lactobacilli largely due to the fact that bidobacteria
can metabolise the neutral oligosaccharide fraction of human milk
(Harmsen et al., 2000). In supporting this, a recent study has shown
that the strain Bidobacterium infantis can grow on puried HMOs
as a sole carbon source (Ward, Ninonuevo, Mills, Lebrilla, &
German, 2006). Of ve bidobacteria strains tested in a follow-up
study, Bidobacterium longum biovar infantis was shown to achieve
highest cell density suggesting that HMOs may even selectively
promote the growth of certain bidobacteria strains (Ward,
Ninonuevo, Mills, Lebrilla, & German, 2007). LoCascio et al.

383

(2007) demonstrated that the infant gut isolate B. longum biovar


infantis ATCC 15697 possessed fucosidase and sialidase activities,
not present in other tested strains and preferentially consumed
small mass oligosaccharides, representing 64% of the total HMOs
available. Indeed, the sequenced genome of this strain revealed the
presence of a milk-active gene suite reecting potential adaptation
to the infant host that included a 43 kilobase pair cluster dedicated
to the uptake and processing of HMOs (Sela et al., 2008). Several
bidobacterial strains have been shown to harbour a lacto-N-biosidase enzyme that liberates lacto-N-biose I from lacto-N-tetraose,
a major component of HMOs, which was not observed in other
bacteria such as clostridia, bacteroides and lactobacilli (Wada et al.,
2008). On the contrary, bidobacteria and lactobacilli represent
only 40e60% of the intestinal microbiota of bottle-fed infants, with
the remaining composed of Enterobacteriaceae and Bacteroides
(Harmsen et al., 2000; Ouwehand, Isolauri, & Salminen, 2002;
Rubaltelli, Biadaioli, Pecile, & Nicoletti, 1998).
The effects of an intestinal microbiota dominated by bidobacteria in breast-fed infants are extremely benecial and include
inhibition of pathogen colonisation (Knol et al., 2005a), modulation
of mucosal physiology and barrier function, and modulation of the
systemic immunologic and inammatory responses as emphasized
by the Committee on Nutrition of the European Society of
Pediatric Gastroenterology, Hepatology and Nutrition in 2004
(ESPGHAN Committee on Nutrition, 2004). Furthermore, it has
been shown that the composition of the gut metagenome can have
consequences for an individuals predisposition to obesity, a trend
which has been observed in both animal models and humans
(Backhed & Crawford, 2009; Backhed et al., 2004; Cani & Delzenne,
2009; DiBaise et al., 2008; Tsukumo, Carvalho, Carvalho-Filho, &
Saad, 2009). Indeed, an increase in the proportion of Firmicutes
relative to Bacteroidetes has been detected in both obese humans
and animals. Several mechanisms linking gut microbiota to energy
metabolism in the host have been proposed including energy
harvest from the diet, synthesis of gut peptides involved in energy
homeostasis, and regulation of fat storage (Cani & Delzenne, 2009).
Almost 40 oligosaccharides have been identied in bovine milk,
which is used in the manufacture of infant formula, and in
comparison to HMOs these oligosaccharides are composed of
shorter oligomeric chains and are signicantly more anionic with
approx. 70% being sialylated (Tao et al., 2008). Since analogues of
HMOs are not commercially available, oligosaccharides from other
sources have thus been assessed for their bidogenic effects in
humans and subsequent use in infant formula. The prebiotic
mixture fructo- and galacto-oligosaccharides (FOS and GOS,
respectively) have proven highly effective as supplements for
promoting the growth of bidobacteria and decreasing the numbers
of pathogenic microorganisms (Boehm et al., 2005; Bruzzese,
Volpicelli, Squaglia, Tartaglione, & Guarino, 2006; Fanaro et al.,
2005a; Miniello, Moro, & Armenio, 2003; Moro & Arslanoglu, 2005).
FOS are of vegetable origin and are found in onions, asparagus,
artichokes and tomatoes (Bornet, 1994; Crittenden & Playne, 1996),
whereas GOS are derived from bovine milk (Montserrat &
Santamaria-Orleans, 2001). Both FOS and GOS harbour the properties of soluble alimentary bres, entering the intestinal tract
intact where they are then fermented by the colonic microbiota
(Montserrat & Santamaria-Orleans, 2001). Analysis of the transcriptional response of B. longum following growth in human milk,
formula milk containing GOS and FOS and semi-synthetic medium
with glucose demonstrated substantial similarity in the transcriptomes for human milk and formula compared with that of the
semi-synthetic medium (Gonzalez, Klaassens, Malinen, de Vos, &
Vaughan, 2008). However, only a few genes implicated in carbohydrate metabolism were similarly upregulated during growth in
human milk and formula milk containing GOS and FOS, which, as

384

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

the authors suggest, may provide guidance for optimizing formula


milk to better simulate the bidogenic effects of human breast
milk. Interestingly, Bouhnik et al. (1996) demonstrated that ingestion of FOS at a level of 12.5 g day1 led to an increase in colonic
bidobacteria in healthy volunteers. Likewise, Boehm et al. (2002)
demonstrated that supplementation of preterm formula with
10 g L1 of GOS and FOS stimulated the growth of bidobacteria in
preterm infants resulting in stool samples with similar characteristics to those from preterm infants who had received human milk.
Indeed, several studies have reported similar effects of feeding
infants GOS/FOS formula, resulting in an intestinal microecology
with similarities to that of breast-fed infants (Bakkar-Zierikzee
et al., 2005; Knol et al., 2005b). Most recently, Costalos, Kapiki,
Apostolou, and Papathoma (2008) demonstrated that infants fed
a formula containing GOS/FOS produced softer stools with
a microbiota having a signicantly lower proportion of both clostridia and E. coli and a higher proportion of bidobacteria as
compared with infants given a standard infant formula. Interestingly, infant formula supplemented with acidic oligosaccharides
derived from pectin hydrolysate did not exert any signicant
inuence on the intestinal microecology of term infants (Fanaro
et al., 2005b). The subsequent health benets of consuming the
GOS/FOS mixture in infants have also been documented. Indeed,
consumption of a supplemented hypoallergenic formula containing
8 g L1 of GOS/FOS for 6 months was shown to reduce the incidence
of atopic dermatitis in a double-blind, randomized, placebocontrolled trial involving 259 infants at risk for atopy during the
rst six months of age (Moro et al., 2006) possibly due to their
bidogenic effects. In addition, consumption of the oligosaccharide
prebiotics by healthy term infants with a parental history of atopy
was shown to reduce the number of infectious episodes and the
incidence of recurring infections, particularly respiratory infections
during the rst 6 months of life in a double-blind, placebocontrolled trial (Arslanoglu, Moro, & Boehm, 2007). In a follow-up
study, it was demonstrated that infants in the GOS/FOS group had
fewer episodes of respiratory tract infections (P < 0.1), fewer fever
episodes (P < 0.00001), and fewer antibiotic prescriptions
(P < 0.05) during the rst two years of life, demonstrating that the
protective effects lasted beyond the 6 month intervention period
(Arslanoglu et al., 2008). In addition, consumption of supplemented
formula containing GOS/FOS for 6 months was shown to induce
a benecial antibody prole following vaccination at 3 months with
Hexavac against a.o. diphteria, tetanus, polio (DTP) by reducing the
total Ig response and modulating the immune response towards
cows milk protein, while leaving the response to vaccination intact
(van Hoffen et al., 2009).
The resulting by-products from the fermentation of prebiotic
substances by the gut microbiota, in particular SCFAs are capable of
exerting numerous physiological effects and are thus essential
biological components in maintaining health and evading disease
(Bruzzese et al., 2006; Roy, Kien, Bouthillier, & Levy, 2006). Moreover, a reduction of SCFA concentrations in the gut has also been
observed in patients with inammatory bowel disease (IBD)
(Treem, Ahsan, Shoup, & Hyams, 1994). Butyrate is perhaps the
most inuential SCFA in terms of its physiological effects, serving as
the principal energy source for colonic epithelial cells and as
regulator to a number of genes associated with cell differentiation,
proliferation and apoptosis. Thus, butyrate has a profound inuence on the aetiology of various colonic diseases including cancer,
IBD and the regulation of lipid metabolism (Scheppach & Weiler,
2004). Interestingly, a recent study demonstrated that the
eukaryotic G-protein-coupled receptor, GPR43 is the relevant
receptor for SCFA effects on immune cells (Maslowski et al., 2009).
Indeed, using murine models, stimulation of GPR43 by SCFAs was
necessary for the normal resolution of certain inammatory

responses whereas GPR43-decient mice (Gpr43/) demonstrated


exacerbated inammation in models of colitis, asthma and arthritis
in the presence of SCFAs (Maslowski et al., 2009). Since butyrate is
also produced in bovine milk, the physiological effects of this fatty
acid are discussed in more detail in the Lipids section.
3.3. Milk oligosaccharides and infection
It has long been recognised that breast-fed infants suffer from
lower rates of enteric infections such as infectious diarrhoea,
a factor that has been primarily attributed to both the secretory
antibodies and prebiotic factors in human milk. However, while the
prebiotic effect exerts a competitive inhibition on enteric pathogenic bacteria and undoubtedly ne-tunes the immune response
against infection, the inhibition of pathogen binding to host cell
ligands afforded by oligosaccharides is believed to be the primary
means of protection for breast-fed infants against enteric
pathogens.
During pathogenesis, the majority of enteric pathogens use host
cell-surface glycans to bind to their target cells. In many cases, this
adhesion is mediated by lectins present on the surface of the
pathogenic microorganism. It is now known that soluble oligosaccharides from milk competitively inhibit the ability of pathogens
to bind to receptors in the gut by binding bacterial lectins (Newburg
et al., 2005). The bacteria are then swept away by ciliary action,
leaving the host mucosa untouched.
Several in vitro studies have yielded results demonstrating the
protective capabilities of HMOs. For example, the a1,2-linked
fucosylated glyconjugates expressed in human milk have been
shown to inhibit binding of Campylobacter to intestinal cells in vivo,
a bacterium known to be a signicant cause of diarrhoea worldwide
(Ruiz-Palacois, Cervantes, Ramos, Chavez-Munguia, & Newburg,
2003). Fucosylated oligosaccharides have also demonstrated an
ability to bind to members of the calicivirus family of enteric
pathogens which have been recognised as a major cause of diarrhoea in humans, especially infants (Marionneau et al., 2002;
Ruvoen-Clouet, Ganiere, Andre-Fontaine, Blanchard, & Le Pendu,
2000). Rotavirus, on the other hand, was shown to be inhibited
by a mucin-associated 46 kDa milk glycoprotein called lactadherin
(Newburg et al., 2005; Yolken et al., 1992). Removal of sialic acid
from lactadherin renders the glycoprotein ineffectual against
inhibiting rotavirus, suggesting that the glycan portion of the
molecule may be responsible for inhibition. Similarly, HMOs were
shown to strongly inhibit hemagglutination mediated by enterotoxigenic E. coli (ETEC) and uropathogenic E. coli (UPEC) strains,
which decreased when the oligosaccharides were desialylated
(Martin-Sosa, Martin, & Hueso, 2002).
Interestingly, although bovine milk oligosaccharides were less
efcient at inhibiting the hemagglutination of the ETEC strains,
they were considered good inhibitors of the UPEC strains (MartinSosa et al., 2002). HMOs have also demonstrated an ability to
inhibit the activities of bacterial toxins. Research by Otnaess,
Laegreid, and Ertresvag (1983) demonstrated that the activities of
the labile toxin of E. coli and cholera toxin were inhibited by the
ganglioside fraction of human milk. Various treatments of the
inhibitory component indicated that it was indeed an oligosaccharide or glyconjugate (glycan). Later, it was demonstrated that
the oligosaccharides of human milk containing a1,2-linked fucose
structures were also capable of binding the stable toxin of E. coli
(Newberg, Pickering, McCleur, & Cleary, 1990). Interestingly, studies
using T84 cells derived from human enterocytes indicated that the
oligosaccharide inhibits stable toxin by binding its cellular
receptor; the extracellular domain of guanylate cyclase, an enzyme
responsible for producing guanosine monophosphate (GMP).
Elevated levels of GMP as a result of stable toxin binding result in

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

loss of chloride and bicarbonate transport, culminating in the efux


of uid and electrolytes on a cellular level, which transpires as
secretory diarrhoea in the host. However, binding of the extracellular domain of guanylate cyclase by the oligosaccharide competitively inhibits binding by stable toxin of E. coli (Crane, Azar, Stam, &
Newburg, 1994). Newburg, Chaturvedi, Crane, Cleary, and Pickering
(1995) estimated that the inhibitory oligosaccharide is active at
a concentration of approximately 30 parts per billion, the concentration at which it occurs in human milk.
Studies in human populations have demonstrated that the
protective capabilities of HMOs observed in vitro are also active in
vivo. Morrow et al. (2004) collected data and samples from 93
breast-feeding mother-infant pairs from birth to 2 years with infant
feeding and diarrhoea data collected weekly. Milk samples
obtained 1e5 weeks postpartum were analysed for oligosaccharide
content. Moderate to severe diarrhoea of all causes occurred less
often in infants whose milk contained high levels of 2-linked
fucosyloligosaccharides as a percent of milk oligosaccharide. Other
studies have also demonstrated the signicance of 2-linked fucosyloligosaccharides for inhibition of pathogen binding during
breast feeding against infectious diarrhoea (Morrow, Ruiz-Palacois,
Jiang, & Newburg, 2005; Newburg et al., 2003). Most recently, it was
demonstrated that early consumption of HMOs through breast
feeding was also associated with less reported respiratory and
gastrointestinal illness in infants (Stepans et al., 2006).
The prevention of bacterial adhesion as observed by human milk
glycans is an attractive option for the development of new therapeutic agents targeting infections of mucosal surfaces in particular,
e.g., infections of the intestine, urinary tract, nasopharyngeal tract,
etc. The concept that the infectious agent is neutralised through
binding of the oligosaccharide is appealing, as the infectious agent
is still available to provoke an immune response. However,
manufacturing and delivery challenges must yet be addressed, but
based on the information encoded in human milk glycans, carbohydrate-based therapies offer great potential for future antimicrobial drugs as discussed by Sharon and Ofek (2000) and
Bavington and Page (2005).
3.4. Milk oligosaccharides and the inammatory immune response
Interestingly, milk oligosaccharides have been shown to play
a signicant role in the activities of the inammatory immune
response. Inammation is a complex multi-step process providing
a non-specic defence mechanism against tissue injury. During the
inammatory process, injured tissue cells release chemical signals
called inammatory mediators that activate the endothelium of
nearby capillaries. This activation results in the release of cellsurface adhesion molecules referred to as selectins, which are
glycoprotein in nature, and essential for the formation of plateletneutrophil complexes (PNC) (Jungi, Spycher, Nydegger, & Barandun,
1986; Kannagi, 2002; Ley, 2003; Rhee et al., 2003; Varki, 1994).
These heterogenous cell aggregates represent a large subpopulation of neutrophils with a greater capacity for phagocytosis and
increased production of reactive oxygen species (ROS) (Peters et al.,
1999). Selectins are also essential components in the cell adhesion
cascade leading to the extravasation of blood cells into the tissue
during inammatory processes (McEver, 1997). PNC formation
requires P-selectin on activated platelets and P-selectin glycoprotein ligand 1 (PSGL-1) on neutrophils (Larson et al., 1989). This
interaction induces signalling pathways leading to an increased
expression of adhesion molecules, including b 2 integrin CD11b/
CD18 (Piccardoni et al., 2001). However, the excessive recruitment
of such cell aggregrates can have deleterious consequences, leading
to the destruction of healthy tissue as observed in many diseases
(Carden & Granger, 2000; Laroux & Grisham, 2001; Schon et al.,

385

2002). Milk oligosaccharides have been shown to harbour


epitopes similar to selectin-binding ligands. Furthermore, since it
has been conrmed that lactose-derived oligosaccharides in infants
survive the gastrointestinal tract and are excreted in the urine
(Rudloff, Pohlentz, Diekmann, Egge, & Kunz, 1996), such oligosaccharides may have a role to play in anti-inammatory processes.
Interestingly, sialyl-Lewis ligands on milk oligosaccharides were
shown to have the greatest capacity for binding selectins in vitro
(Rudloff, Stefan, Pohlentz, & Kunz, 2002).
In an attempt to decipher the exact mechanisms involved Bode
Rudloff, Kunz, Strobel, and Klein (2004b) demonstrated that the
acidic fraction of HMOs competes with PSGL-1 for P-selectin binding
thus reducing PNC formation. Indeed, the acidic fraction reduced
PNC formation up to 20%, which was similar to the effect seen with
sialyl-Lewis X, a physiological binding determinant for selectins. A
dose-dependent decrease in b 2 integrin expression was also
observed in associated neutrophils. Moreover, several active
components within the acidic component of HMOs were identied
as determinants for selectin binding, such as 30 -sialyl-lactose and 30 sialyl-3-fucosyl-lactose (Bode et al., 2004a). Interestingly, a study by
Schumacher, Bendas, Stahl, and Beermann (2006) identied the
acidic fraction of HMOs as effective agents in interfering with
P-selectin ligand binding, suggesting that molecular charge is
important for binding interactions. However, total HMOs were not as
effective as the acidic fraction alone, suggesting that HMOs modulate
rather than block the function of P-selectin (Schumacher et al., 2006).
SCFAs resulting from the degradation of HMOs (particularly
acetate and propionate) are now known to bind GPR43, which is
expressed by cell types involved in innate immunity, and in doing
so regulate immune and inammatory responses (Maslowski et al.,
2009). Indeed, re-colonisation of germ-free mice, suffering from
induced colitis, with gut microbiota showed a marked reduction in
inammation. Likewise, treatment of these germ-free mice with
acetate markedly improved disease indices. Therefore, SCFA-GPR43
interactions provide another mechanism whereby HMOs can
regulate the immune and inammatory responses.
Several modern-day diseases have resulted from an overshooting immune response. The lower incidence of infectious and
inammatory diseases in breast-fed versus formula-fed infants
suggests that HMOs may play a signicant role in anti-inammatory processes. For example, necrotising enterocolitis (NEC),
a disease primarily observed in premature infants results from
pathogen colonisation and adhesion to the intestinal surface
resulting in tissue necrosis. However, excessive leukocyte inltration, followed by the production of ROS have also been shown to
play a major role in its progression (Bode et al., 2004a; Musemeche,
Caplan, Hsueh, Sun, & Kelly, 1991). The lower incidence of NEC in
breast-fed infants suggests that HMOs might inuence the onset of
the disease by preventing an amplied immune response.
3.5. Sialic acid
Another milk oligosaccharide that has been the recipient of
considerable attention in recent years is N-acetylneuraminic acid,
also known as sialic acid. Several oligosaccharides found in milk
tend to be sialylated, consisting of a core structure with a lactose
unit at the reducing end and at least one N-acetylneuraminic acid
unit at the non-reducing end (Wang & Brand-Miller, 2003). The
most abundant sialyloligosaccharides of human milk found at
several stages of lactation were 30 -sialyl-3-fucosyl-lactose and sialyllacto-N-tetraoses (a, b and c), while 60 -sialyllactosamine and 30 sialyllactose were the most abundant sialyloligosaccharides in
bovine milk (Martin-Sosa, Martin, Garcia-Pardo, & Hueso, 2003).
Interestingly, sialic acid is also found in high concentrations in
the mammalian nervous system, with the majority being present in

386

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

gangliosides (65%) followed by glycoproteins (32%) and the


remaining 3% as free sialic acid (Brunngraber, Witting, Haberland, &
Brown, 1972). As suggested by Wang and Brand-Miller (2003), the
sialic acid moieties of gangliosides and glycoproteins in the brain
frontal cortex play both a structural and functional role and probably participate in a variety of cellular events, such as cell recognition, cell-to-cell contact, receptor binding and modulation,
immunological properties and biosignal transduction. Gangliosides
are complex glycosphingolipids, which make up 10% of the total
lipid mass in the brain, containing different numbers of negatively
charged sialic acid moieties (Wang & Brand-Miller, 2003). While
the exact functions of gangliosides still remain relatively poorly
understood, it has been hypothesised that they are involved in the
formation of memory (Wang & Brand-Miller, 2003). Brain glycoproteins are also related to brain function in the formation of
memory and learning (Bogoch, 1977; Schmidt, 1989). The repulsive
effects of the negatively charged sialic acid may serve to prevent
cell aggregation (Schauer & Kamerling, 1997). On the other hand,
cell adhesion may be facilitated through positively charged
substances or calcium (Wang & Brand-Miller, 2003). Indeed,
binding of calcium to ganglioside sialic acid is of great importance
in the function of nervous tissues (Rosenberg, 1995; Schauer &
Kamerling, 1997). Thus, it has been postulated that sialic acid is
the actual receptor for neurotransmitters in the central nervous
system (CNS) (von Itzstein & Thomson, 1997).
Several studies have demonstrated that children who have been
breast-fed as babies display higher intelligence skills than children
who were formula-fed (Fergusson, Beautrais, & Silva, 1982; Lucas,
Morley, & Cole, 1998; Lucas, Morley, Cole, Lister, & Leeson-Payne,
1992; Rodgers, 1978; Smith, Durkin, Hinton, Bellinger, & Kuhn,
2003). In an intriguing study, Wang, McVeagh, Petocz, and BrandMiller (2003) examined the brain frontal cortex gray matter of 25
infants who died of sudden infant death syndrome and whose
feeding pattern had been previously recorded in 21 cases. It was
found that ganglioside-bound and protein-bound sialic acid
concentrations were 32% and 22% higher, respectively, in the frontal
cortex gray matter of breast-fed infants. Moreover, it was demonstrated that the concentration of sialic acid in saliva of breast-fed
infants was twice the level of sialic acid found in the saliva of
formula-fed infants (Wang et al., 2001b).
In a more recent study, it was shown that feeding piglets
protein-bound sialic acid during early development enhanced
learning and memory and increased the expression of 2 genes
associated with learning in developing piglets, namely a2,8-sialyltransferase II and IV (Wang et al., 2007b). Furthermore, within
2 min of injection of 3-day old male domestic piglets with N-acetylneuraminic acid-6-14C, 80% of the activity was removed from
the blood. At 120 min, the brain contained signicantly more
radioactivity than the liver, pancreas, heart and spleen, but less
than the kidneys, proving that an exogenous source of sialic acid is
capable of crossing the blood-brain barrier (Wang, Downing,
Petocz, Brand-Miller, & Bryden, 2007a).
The potential signicance of sialic acid for early brain development has led to an assessment of the sialic acid contents of both
bovine milk and infant formulas. Interestingly, Asakuma et al.
(2007) demonstrated that the concentration of 30 -sialyllactoses in
human colostrum was signicantly higher on day 1 of the rst three
days of lactation than on day 2 and 3, but the levels of 60 - sialyllactoses and sialyllactose-N-tetraose were higher on day 3 than
on day 1, indicating that during the rst 3 days of lactation the
concentration of sialyloligosaccharides changes in accordance with
the physiological demands of the newborn infant. However, infant
formulas were shown to contain less than 25% sialic acid than that
found in mature human milk (Wang, Brand-Miller, McVeagh, &
Petocz, 2001a). Thus, further studies are required to assess the

effectiveness and safety of sialic acid supplemented formulas.


Interestingly, an assessment of bovine colostrum and milk during
prepartum and early lactation suggested that bovine colostrum,
especially that collected after parturition may be a suitable source
of 30 sialyllactoses as an additive by the food or pharmaceutical
industries (Nakamura et al., 2003).
Isomeric differences exist amongst mammalian sialic acids,
which may have consequences for human health and disease.
Indeed, while humans produce N-acetylneuraminic acid other
mammals synthesise N-glycolylneuraminic acid (Neu5Gc), a mutation estimated to have occurred w2e3 million years ago, just
before the emergence of the genus Homo (Wood & Collard, 1999). A
Neu5Gc-binding pathogen such as Non-Human Hominid (NHH)
malaria, combined with genetic drift caused by ancestral demography, has been proposed as a possible selection mechanism, suggesting that sialic acid-related genes are a hot-spot for genetic and
physiological changes in human evolution (Varki, 2010). However,
humans can acquire Neu5Gc from dietary sources including red
meats and, to a lesser extent, milk products (Tangvoranuntakul
et al., 2003); this is detected as foreign by the immune system
and may be recognised by Neu5Gc-binding pathogens. Thus, it has
been suggested that Neu5Gc antibodies contribute to red meat
aggravation of diseases by stimulating chronic inammation, and
that Neu5Gc-containing epithelial cells may play a role in red meatrelated food poisoning (Varki, 2010). Moreover, human carcinomas
have been shown to accumulate dietary Neu5Gc (Yin et al., 2006)
and anti-Neu5Gc antibodies have been shown to enhance Neu5Gcpositive tumors in mice (Hedlund, Padler-Karavani, Varki, & Varki,
2008). However, it has been suggested that high levels of these
antibodies may kill tumor cells and even protect certain individuals
with very high levels from some cancers (Varki, 2010).
While the concepts proposed have been deduced from sound
scientic evidence many of these are speculative and require
further investigation. But as with all biological systems, and milk is
no exception, one must take into account the interplay between all
substituents and the host. Thus, while consumption of Neu5Gc
from bovine milk may contribute to human disease phenotypes,
milk itself is composed of a multitude of health-promoting factors,
which one could speculate, could also negate, or at least minimize
such deleterious effects.
4. Milk lipids
4.1. Introduction
Both bovine and human milk fat are rich sources of bioactive
substances offering a plethora of ingredients for the development
of functional food products (German & Dillard, 2006; Haug et al.,
2007). Structurally, milk fat occurs as micro droplets surrounded
by the MFGM, containing mainly triacylglycerols, esters and retinol
esters in the core, and primarily phospholipids and cholesterol in
the membrane (Argov, Lemay, & German, 2008; Jensen, 1999;
Mather, 2000). In the case of human milk, 40e55% of the total
energy intake of the offspring is obtained from milk fat (German &
Dillard, 2006), while it also supplies essential nutrients such as
essential fatty acids and the fat soluble vitamins, A, D, E and K.
Triacylglycerols represent 97e98% of the total lipids in the milk of
most species, while the level of cholesterol in milk is low when
compared with other food sources (Fox & McSweeney, 1998).
4.2. Milk fat globule membrane
In a review by Spitsberg (2005), the associated health benets of
bovine MFGM were discussed in detail. Interestingly, components
of the MFGM have exhibited anticancer, anticholesterolemic and

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

antimicrobial effects, with some of these effects attributable to the


protein fraction. Recently, it was demonstrated that milk fat globules, MFGM and the glycosphingolipids lactosylceramide, and the
gangliosides GM3 and GD3 were capable of binding ETEC strains
and inhibiting bacterial haemagglutination (Sanchez-Juanes,
Alonso, Zancada, & Hueso, 2009). Moreover, feeding cows a diet
rich in polyunsaturated fatty acids (PUFAs) has been shown to affect
the composition of the MFGM, reecting changes in fatty acid
synthesis in the mammary gland resulting in a higher amount of
phospholipids (18%), which was related to a smaller size of milk fat
globules, an increase in the concentration of sphingomyelin (30%),
and a higher content of stearic acid (1.7-fold), unsaturated fatty
acids (1.36-fold) and C18:1 trans fatty acids (3.7-fold) (Lopez et al.,
2008).
4.3. Phospholipids and cancer
While much of the health benets of MFGM have been associated
with proteins, the phospholipid content has also received signicant
attention for its associated health effects. Indeed, Spitsberg (2005)
has proposed that the consumption of MFGM as a neutraceutical
or dairy food, or the consumption of food products enriched with
MFGM conveys health benets due to the presence of phospholipids.
The three main phospholipids of MFGM include sphingomyelin,
phosphatidyl choline and phosphatidyl ethanolamine (Spitsberg,
2005). Several cellular processes are affected by phospholipids,
from growth and development to memory, stress responses, and
myelination in the CNS (McDaniel, Maier, & Einstein, 2003; Oshida
et al., 2003b). Furthermore, the sphingomyelin component of the
MFGM has exhibited anticancer effects (Berra, Colombo,
Sottocornola, & Giacosa, 2002; Hertervig, Nilsson, Cheng, & Duan,
2003; Lemonnier et al., 2003; Parodi, 2001) and has been shown to
be an effective inhibitor of intestinal absorption of cholesterol in rats
(Noh & Koo, 2004).
Sphingomyelin belongs to one of the most complex classes of
lipids in nature, namely the sphingolipids. Such lipids contain an
amide-linked fatty acid and a long chain (sphingoid) base. They are
located in cellular membranes, lipoproteins (especially low density
lipoproteins, LDL) and other lipid-rich structures such as skin
(Akalin, Gonc, & Unal, 2006). It is now known that these complex
sphingolipids and their breakdown products (ceramides and
sphingosines) are highly bioactive compounds that play essential
roles in cell growth, survival and death (Akalin et al., 2006; Cuvillier
& Levade, 2003; Vesper et al., 1999). Interestingly, orally administered sphingolipids have been shown to prevent colon cancer
development in different rodent models (Schmelz, 2004). These
dietary sphingolipids are digested throughout the small intestine
and colon by intestinal or bacterial enzymes (Schmelz, Crall,
Larocque, Dillehay, & Merrill, 1994).
Earlier studies have also shown that feeding sphingomyelin to
mice in amounts equivalent to those found in foods such as dairy
products (0.025e0.1%) (Vesper et al., 1999) inhibits the early stages
of colon carcinogenesis and decreases the ratio of malignant
adenocarcinomas to benign adenomas (Dillehay, Webb, Schmelz, &
Merrill, 1994; Schmelz et al., 1996). In addition to sphingomyelin,
the glycosphingolipids of milk, namely glucosylceramide, lactosylceramide, and ganglioside GD3 have been shown to inhibit the
early stages of chemically induced colon cancer in female CF1 mice
(Schmelz, Sullards, Dillehay, & Merrill, 2000). The genetic mutations
associated with colon tumors are often the result of defects in the
adenomatous polyposis coli (APC)/b-catenin regulatory system.
Feeding C57B1/6JMin/ mice (harbouring a truncated APC gene
product) diets supplemented with ceramide, sphingomyelin, glucosylceramide, lactosylceramide, and ganglioside GD3 reduced the
number of tumors in all regions of the intestine (Schmelz et al.,

387

2001). It was concluded that dietary sphingolipids, via their digestion products, bypass or correct defect(s) in the (APC)/b-catenin
regulatory pathway (Schmelz et al., 2001).
Similar effects were also observed in human colon cancer cell
lines. Studies using HT-29 and HCT-116 human colon cancer cells
demonstrated that both sphingosine and ceramide inhibited
growth and caused death of the colon cancer cells in time and
concentration-dependent manners, where the 4,5-trans double
bond of ceramide was found to be essential for its inhibitory effects
(Ahn & Schroeder, 2002). Apoptosis has been suggested as the
mechanism of action since both the sphingoid bases and ceramide
were shown to cause chromatin and nuclear condensation, and
fragmentation of DNA. Moreover, the sphingoid bases were shown
to arrest the cell cycle at the G2/M phase and cause accumulation in
the S phase.
Sphingolipid metabolites have shown potential as chemotherapeutic and chemopreventive agents. The sphingoid bases sphingosine and sphinganine were shown to preferentially inhibit breast
cancer cells and eliminate stem cells from which most breast cancer
cells arise (Ahn, Chang, & Schroeder, 2006). In contrast, sphingosine-1-phosphate has been shown to stimulate cell growth and
suppress apoptosis by acting through G-protein coupled receptors
present on mammalian cells, thus stimulating cell proliferation,
angiogenesis and inhibiting apoptosis (Oskouian & Saba, 2007;
Spiegel & Milstien, 2003). The enzyme responsible for sphingosine-1-phosphate synthesis, namely sphingosine kinase 1, behaves
as an oncogene in experimental systems (Oskouian & Saba, 2007).
Moreover, sphingosine-1-phosphate-specic antibodies slow
tumor progression and angiogenesis in mice murine xenograft and
allograft models (Oskouian & Saba, 2007). As suggested by
Oskouian and Saba (2007), knowledge of this nature may have
implications regarding colon cancer screening, dietary chemoprevention and therapeutics.
4.4. Phospholipids and the immune system
Sphingolipids play an essential role in the development, activation and regulation of the immune system (Cinque et al., 2003).
As described by Cinque et al. (2003), the breakdown products of
sphingolipid metabolism including ceramide, sphingosine, ceramide-1-phosphate and sphingosine-1-phosphate serve as second
lipid messengers, which are all involved in a common signalling
pathway that controls the main stages of immune cell development, differentiation, activation, proliferation and function. Interestingly, treatment of bone marrow-derived dendritic cells with the
human milk-derived gangliosides, GD3 and GM3, decreased the
production of interleukin-6 (IL-6), IL-10, IL-12 and tumor necrosis
factor-alpha. Furthermore, the expression of CD40, CD80, CD86 and
major histocompatibility complex class II on dendritic cells was
suppressed (Bronnum, Seested, Hellgren, Brix, & Frokiaer, 2005),
suggesting that dietary sphingolipids may also have a dramatic
effect on immune regulation.
4.5. Phospholipids and the central nervous system
Sphingolipids play a signicant role in neuronal cell function by
regulating rates of neuronal growth, differentiation and death
(Buccoliero & Futerman, 2003). Likewise, dietary sphingolipids may
also have the potential to exert dramatic effects on the activities of
the CNS. Interestingly, studies using L-cycloserine treated rats
(which lack the ability to produce serine palmitoyltransferase
(SPT), a rate-limiting enzyme for sphingolipid biosynthesis)
demonstrated that feeding the rats bovine sphingomyelin-supplemented diets contributed to CNS myelination with experimental
inhibition of SPT corrected (Oshida et al., 2003a). Given the potent

388

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

biological activities of sphingolipids, there is no doubt that these


food components could be categorized as functional food ingredients. However, while cell culture studies and a few animal studies
have provided promising results, more animal and human trials are
urgently required to thoroughly investigate the physiological efcacies of these agents.
While the nutritional and functional characteristics of the
MFGM are well known it should be recognised that raw milk is
exposed to various processing parameters to ensure safety and
prolong shelf-life, including heat treatments and homogenization,
which have profound effects on the physical structure of milk fat
(Michalski, 2007; Michalski & Januel, 2006). Homogenization, for
example, results in a dramatic reduction in fat globule size (Walstra,
2003). Using confocal laser scanning microscopy, a recent study
demonstrated that processing (centrifugation, pasteurizationhomogenisation and churning) changed the composition of the
MFGM through the loss of phospholipids and the adsorption of
caseins and whey proteins onto the surface (Gallier, Gragson,
Jimenez-Flores, & Everett, 2010). The effects of such structural
changes on the health properties associated with the MFGM are
still largely unknown. Therefore, in vitro and in vivo studies that
elucidate the effects of processing on the bioactivities of MFGM and
its substituents will provide valuable information which will
undoubtedly lead to dairy products/ingredients which harbour the
full complement of bioactivities found in unadulterated raw milk.
4.6. Polyunsaturated fatty acids
Both bovine and human milk fat contain PUFAs including low
levels of the essential n-6 and n-3 series, linoleic acid (LA, 18:2n-6)
and a-linolenic acid (ALA, 18:3n-3). LA and ALA concentrations
have been reported to range between 1e3 and 0.5e2 wt%,
respectively, in bovine milk (Jensen, 2002; Kaylegian & Lindsay,
1995), whereas concentrations of 10.75  4.22 and 0.35 
0.05 wt%, respectively, have been reported in human milk (Gibson
& Kneebone, 1981). These fatty acids cannot be synthesised by the
body but are required for survival of humans and other mammals
and hence have to be obtained from the diet (Das, 2006). LA is the
most common PUFA incorporated into the dynamic phospholipids
required for membrane structure and lipoproteins [particularly
phospholipid-rich high density lipoprotein (HDL)] involved in lipid
transport (Wijendran & Hayes, 2004). Both LA and ALA can be
further metabolised by D6-desaturation, elongation and D5-desaturation to form arachidonic acid (ARA, 20:4n-6) and eicosapentaenoic acid (EPA, 20:5n-3), respectively, which form
precursors for the synthesis of prostaglandins and leukotrienes
which control numerous cell activities responsible for healthy
living (Innis, 2004, 2007b; Wall, Ross, Fitzgerald, & Stanton, 2010).
EPA is then further metabolised to form docosahexaenoic acid
(DHA, 22:6n-3) (Innis, 2003), which unlike the n-6 fatty acids, has
a specic distribution in tissues and phospolipids and is extremely
important in the development of the CNS (Innis, 2007a; Wurtman,
2008).
Interestingly, the n-6 and particularly the n-3 fatty acids have
demonstrated potential health benets both in vitro and in vivo
(Connor, 2000) by reducing the risk of cardiovascular disease (CVD)
(Fedacko et al., 2007; Siddiqui, Harvey, & Zaloga, 2008), type-2
diabetes, hypertension (Wijendran & Hayes, 2004; Willett, 2007;
Zhao et al., 2007), cancer (Dupertuis, Meguid, & Pichard, 2007;
Larsson, Kumlin, Ingelman-Sundberg, & Wolk, 2004) and certain
neurological disfunctions (Alessandri et al., 2004; Hamilton,
Hillard, Spector, & Watkins, 2007), and have recently been shown
to improve bone density in ovariectomized rats (Poulsen, Moughan,
& Kruger, 2007) and eye lens nuclear density in women aged 52e73
years (Lu et al., 2007).

The n-3 fatty acids have also exhibited therapeutic potential in


the treatment of inammatory diseases such as rheumatoid
arthritis, and in the alleviation of symptoms of mental health such
as depression and dementia (Ruxton, Reed, Simpson, & Millington,
2007). The n-3 fatty acid, DHA, has also shown promise as
a potential treatment for late-onset Alzheimers disease (Ma et al.,
2007). Indeed, genetic polymorphisms that reduce expression of
the protein LRII are associated with increased risk of late-onset
Alzheimers disease. However, DHA was shown to signicantly
increase LRII in multiple systems, including primary rat neurons,
aged non-Tg mice and an aged DHA-depleted APPsw AD mouse
model and in a human neuronal cell line (Ma et al., 2007).
4.7. Conjugated linoleic acid
In ruminant animals, dietary LA can be converted to CLA as
a consequence of biohydrogenation by the rumen bacteria coupled
with endogenous synthesis by the animals own tissues (Griinari
et al., 2000; Lock & Bauman, 2004) where the cis-9, trans-11
isomer accounts for the majority (over 90%) of total milk fat CLA
(Stanton, Murphy, McGrath, & Devery, 2003). CLA is thus produced
when dietary LA is converted to vaccenic acid by the microbial
enzyme linoleic acid isomerase, in the process called biohydrogenation (Stanton et al., 2003). Vaccenic acid is then converted to CLA in the mammary gland via the enzyme D9-desaturase
and is thus excreted in milk (Griinari & Bauman, 1999). CLA can also
be generated directly from LA as a result of microbial biohydrogenation, where it is formed as an intermediate in the
generation of vaccenic acid. Of the 1010e1011 bacteria residing in
the rumen of a dairy cow, only a few species have been shown to
perform these biohydrogenation reactions (Lock & Bauman, 2004).
Remarkably, compared with their parent fatty acids, CLA
isomers have demonstrated superior abilities in vitro and in vivo to
prevent various types of cancer, hypertension, atherosclerosis and
diabetes as well as an ability to improve immune function and body
composition, which have been discussed in detail in several
publications (Benjamin & Spener, 2009; Bhattacharya, Banu,
Rahman, Causey, & Fernandes, 2006; Churruca, FernandezQuintela, & Portillo, 2009; Kelley, Hubbard, & Erickson, 2007;
Mitchell & McLeod, 2008; Nagao & Yanagita, 2005; Pariza, Park, &
Cook, 2001; Silveira, Carraro, Monereo, & Tebar, 2007; Watras,
Buchholz, Close, Zhang, & Schoeller, 2007). Although there are 28
different CLA isomers, health benets have been mainly attributed
to two, namely, cis-9, trans-11 and trans-10, cis-12 CLA. In addition,
minute amounts of ALA biohydrogenation intermediates (conjugated alpha-linolenic acids, CLNA) have also been identied in milk
fat (Destaillats, Trottier, Galvez, & Angers, 2005; Plourde,
Destaillats, Chouinard, & Angers, 2007). Recent evidence has
highlighted the health-promoting properties of these conjugated
PUFAs (Nagao & Yanagita, 2005). These include C18:3 c9, t11, t13
CLNA (Tsuzuki, Tokuyama, Igarashi, & Miyazawa, 2004; Yasui,
Hosokawa, Kohno, Tanaka, & Miyashita, 2006; Yasui et al., 2005)
exhibiting anticancer properties, c9, t11, c13 CLNA that has
exhibited a hypolipidemic effect (Arao, Yotsumoto, Han, Nagao, &
Yanagita, 2004), and c9, t11, c15 CLNA that has been shown to
inhibit the growth of SW480 colon cancer cells (Coakley et al.,
2008; Hennessy et al., 2011).
4.8. Increasing the conjugated linoleic acid content of milk and
dairy foods
With regard to dietary CLA, it has been estimated that dairy
products and red meat provide about 70% and 25%, respectively of
total dietary intake in humans (Ritzenthaler et al., 2001). Understandably, however, the levels of these fatty acids in milk can vary

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

depending to a large extent on animal nutrition. Major research


efforts have therefore focused on methods to achieve higher CLA
content in milk through animal nutrition that have been reviewed
by Lock and Bauman (2004) and Stanton et al. (2003). Normally,
diets consumed by lactating dairy cows are low in fat content,
containing only about 4e5% lipid (Lock & Bauman, 2004).
Increasing the CLA content of milk, in particular the cis-9, trans-11
isomer, generally requires an increase in rumen vaccenic acid
output, thus allowing for increased endogenous synthesis in the
mammary gland (Lock & Bauman, 2004; Stanton et al., 2003).
Increasing rumen vaccenic acid concentration can be achieved by
increasing the supply of 18-carbon PUFA precursors. Likewise,
inhibiting the reductase enzymes in the rumen responsible for the
conversion of vaccenic acid to stearic acid will also ensure the
increased accumulation and absorption of vaccenic acid by
the animal that can be endogenously converted to CLA in the
mammary gland via D9-desaturase (Griinari & Bauman, 1999).
Plant oils and seeds, marine oils, animal fats, forage, ionophores,
synthetic CLA supplements and combination diets have all been
investigated as means of increasing milk fat CLA concentrations
(Chilliard, Ferlay, & Doreau, 2001; Hennessy, Ross, Stanton, Devery,
& Murphy, 2007; Murphy, Coakley, & Stanton, 2008; Stanton et al.,
2003). Physiological factors also affect milk fat content of CLA, even
between individual animals, presumably related to individual
variations in expression of the D9-desaturase gene and rumenic
outow of vaccenic acid and CLA.
The ultimate aim of increasing the CLA content of milk fat is to
generate dairy products with increased CLA concentration and thus
improved nutritive and therapeutic value when consumed. In
a recent study, CLA-enriched cheese was successfully manufactured
using milk from cows that had been fed on pasture with sunower
oil (Table 4 and Fig. 1, adapted from Coakley et al., 2007). The content
of cis-9, trans-11 CLA in the cheese derived from the pasture-based
sunower oil milk was 1.93 g 100 g1 of fatty acid methyl esters
(FAME), which remained stable throughout the 6 months of
ripening, as opposed to 0.78 g 100 g1 of FAME for the control cheese.
Moreover, sensory evaluation of CLA-enriched milk and cheese by an
open and trained panel clearly demonstrated that the control and
test samples were comparable in terms of mouth-feel, colour, avour
and quality (Khanal et al., 2005). Likewise, manufacture of UHT milk,
butter and cheese with CLA-enriched milk demonstrated that the
high content of CLA did not negatively affect the overall impression
and avour of the three products (Jones et al., 2005a). Processing has
little effect on CLA, so its content in dairy products is directly related
to the CLA concentration of the starting milk (Parodi, 2003).
However, a more recent study has demonstrated that thermal
treatment and refrigerated storage resulted in signicant decreases
or disappearances of the minor CLA isomers coupled with an
increase in trans, trans isomers (Rodriguez-Alcala & Fontecha, 2007).
In terms of the effects of ingestion of CLA-enriched foods on plasma
fatty acid proles, a recent animal study using growing pigs indicated that consumption of butter enriched with cis-9, trans-11 CLA
over a three week period resulted in increased serum concentrations
of ALA and decreased myristic and palmitic acid, with no signicant
effect on the lipoprotein prole (Haug et al., 2008).
In addition, milk-fat CLA must also retain its biological activity.
In a study by Miller, Stanton, Murphy, and Devery (2003), milk-fat
CLA was shown to be as effective as synthetic isomers of CLA at
modulating synthetic CLA-responsive biomarkers in MCF-7 and
SW480 cell lines. However, nutritional strategies to increase the
CLA content of milk also result in milk fat containing increased
concentrations of trans-18:1 (particularly trans-11 18:1, vaccenic
acid), lowered saturated fatty acid concentrations and slightly
increased n-3 PUFA content (Jones et al., 2005b). The greater vaccenic acid content of the CLA-enriched products reects increased

389

Table 4
Effect of treatment on the milk fatty acid compositiona of cows indoors on the
control diet and cows outdoors on pasture receiving sunower oil supplementation
at day 14 (adapted from Coakley et al., 2007).
Fatty acid

Indoors
control
diet

Outdoors on
pasture with
sunower oil
supplementation

C4:0
C6:0
C8:0
C10:0
C12:0
C14:0
C14:1 total
C15:0
C16:0
C16:1 total
C17:0
C17:1
C18:0
C18:1 trans-9 18:trans-11
C18:1 trans-13 18:cis-9
C18:2 cis-9, cis-12 linoleic acid
C20:0
C18:3 cis-9, cis-12, cis-15 linolenic acid
C18:2 cis-9, trans-11 CLA
Other fatty acids

1.46
1.50
1.18
2.92
3.56
12.82
1.45
1.66
35.59
2.30
0.60
0.52
7.51
1.37
18.27
1.14
0.12
0.53
0.46
4.38

1.51
1.22
0.85
1.86
2.30
9.35
1.41
1.13
21.62
2.09
0.61
0.35
12.63
8.05
25.63
2.53
0.10
0.53
2.22
4.50

Values are g 100 g1 fatty acid methyl ester.

ruminal biohydrogenation required to increase synthesis of cis-9,


trans-11 CLA. However, dietary trans fatty acids have been associated with atherogenic risk factors including increased plasma
cholesterol and triglyceride levels as well as plasma markers of
inammation and endothelial dysfunction (Lopez-Garcia et al.,
2005; Mensink & Katan, 1990). A comparison of the effects of
vaccenic and elaidic acid (trans-9 18:1) consumption by male
hamsters indicated that the ratio of LDL/HDL-cholesterol in plasma
was signicantly higher in those fed vaccenic acid (Meijer, van Tol,
van Berkel, & Weststrate, 2001). In contrast to this Bassett et al.
(2010) found that a vaccenic acid-rich butter protected against
atherosclerosis in LDL receptor decient (LDLr/) mice, whereas
a hydrogenated elaidic acid-rich diet stimulated the condition.
Likewise, cholesterol-fed hamsters supplemented with a vaccenic
acid-enriched butter resulted in a plasma lipoprotein cholesterol
prole that is associated with a reduced risk of atherosclerosis
(Lock, Horne, Bauman, & Salter, 2005) Moreover, humans can
convert vaccenic acid to cis-9, trans-11 CLA (Turpeinen et al., 2002).
In a double-blind, cross-over study, Burdge et al. (2005) examined
the level of cis-9, trans-11 CLA and the vaccenic acid content of
plasma and leukocyte lipids in healthy men following consumption
of dairy products naturally enriched with these fatty acids. While
the cis-9, trans-11 CLA content of both plasma and cellular lipids
was increased, a signicant increase was also observed in vaccenic
acid content. However, in a more recent study, it was demonstrated
that consumption of dairy products naturally enhanced with cis-9,
trans-11 CLA and trans-11 18:1 by healthy middle-aged men did not
cause signicant changes in CVD risk variables, as assessed by
investigating blood lipid proles, the atherogenicity of LDL,
markers of inammation and insulin resistance, following a 6-week
consumption period (Tricon et al., 2006).
4.9. Increasing eicosapentaenoic acid and docosahexaenoic
acid content of milk
Enhancing the EPA and DHA concentration of milk fat is a worthy
approach to enhance the fatty acid prole of dairy products. Both
EPA and DHA are either absent or present at minimal levels in

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

Cis-9, trans-11 CLA (g 100 g-1 of FAME)

390

2.0

1.5

1.0

0.5

0.0
0

50

100
150
Ripening time (days)

200

250

Fig. 1. Cis-9, trans-11 conjugated linoleic acid (CLA) concentration in cheese manufactured from (>) milk from cows fed on pasture and receiving sunower oil supplementation at
day 14 and (B) milk from cows fed indoors with a control diet (adapted from Coakley et al., 2007).

traditional dairy cow diets. One potential means of increasing the


content of both EPA and DHA in milk fat is to exploit sh oils, sh byproducts and marine algae as dairy feedstuffs, since these are rich
sources of both fatty acids. The term transfer efciency refers to the
extent to which dietary supplementation enhances the secretion of
benecial fatty acids in milk. However, the transfer efciency of EPA
and DHA to milk fat is low (Lock & Bauman, 2004). For example,
before sh oil supplementation, the content of EPA and DHA in milk
fat averaged less than 0.1% of total fatty acids but only increased to
0.2e0.3% of total fatty acids after supplementation (Chilliard et al.,
2001). However, transfer efciencies can be improved when sh
oil is directly administered postruminally or fed in a rumen-protected form (Lock & Bauman, 2004). Indeed, McConnell (2004)
reported transfer efciencies of 30 and 25% for EPA and DHA,
respectively when 150 g d1 of fractionated sh oil was infused into
the abomasum. Transfer efciencies can also be improved by
feeding sh meal as opposed to sh oils, the latter which are protected from rumen biohydrogenation by the matrix within which
the oil is encased (Lock & Bauman, 2004).
Overall, two hypothesis have been put forward for the low
transfer efciencies for EPA and DHA into milk (Lock & Bauman,
2004). The rst hypothesis suggests that EPA and DHA are biohydrogenated by rumen bacteria; the second hypothesis suggests
that EPA and DHA partition into plasma lipids that are less available
to the mammary gland.

4.10. Butyrate
Milk fats contain a large variety of SCFAs (Christie, 1995) with 1e6
carbon atoms in the acyl chain. Butyric acid (C4:0), exerts an array of
effects on the colonic mucosa and represents approximately 10% of
all fatty acids found in bovine milk, being present at approximately
2e5 wt % (Jensen, 2002; Kaylegian & Lindsay, 1995). Interestingly,
butyrate is generated by bacteria in the rumen from carbohydrates
and then transported via the blood to the mammary gland where it is
reduced to butanoic acid (Jensen, 1999). In humans, butyrate
provides the principal source of energy for colonic epithelial cells but
also regulates a number of genes associated with cell differentiation,
proliferation and apoptosis (Hamer et al., 2008; Scheppach et al.,

1992). Butyrate is also produced in the intestine following the


microbial breakdown of prebiotic substances and non-digestible
carbohydrates. Several studies have yielded results demonstrating
the anti-proliferative, anti-inammatory and apoptotic properties of
butyrate, these have been reviewed by Mills, Ross, Fitzgerald, and
Stanton (2009).
While ingested butyrate does not reach the large intestine,
studies suggest that dietary butyrate can exhibit biological effects
at locations other than the colon. Orally ingested milk butyrate
undergoes lipase-mediated hydrolysis in the stomach which is
complete on reaching the proximal small intestine (Parodi, 1997).
Here, butyrate is immediately absorbed, processed and released
into the blood stream for transport to the liver where most is
metabolized (Parodi, 1997; Smith, Yokoyama, & German, 1998).
Interestingly, Yanagi, Yamashita, and Imai (1993) demonstrated
that margarine supplemented with sodium butyrate signicantly
reduced the incidence of mammary tumors in a rat model. Ingestion of butyrate as a natural component of anhydrous milk fat
(containing 0.8% butyrate) was found to be effective at inhibiting
mammary tumorigenesis in rats (Belobrajdic & McIntosh, 2000).
Such studies provide promising evidence that dietary sources of
butyrate can inuence carcinogenesis at other sites in the body.

4.11. Medium chain fatty acids


Medium chain fatty acids (MCFA) that contain 8e12 carbons and
are saturated are constituents of milk and have been associated
with benecial effects linked with metabolic activities. It has been
postulated that MCFA may help to reduce the risk of developing
features of metabolic syndrome (Pfeuffer & Schrezenmeir, 2002;
2007), a cluster of metabolic disorders including dyslipidaemia,
hypertension, obesity and glucose intolerance, where insulin
resistance is the core phenomenon and co-occurrence is associated
with increased CVD risk (Pfeuffer & Schrezenmeir, 2007). Earlier
studies have shown that dietary substitution of medium chain
triglycerides (MCT) for long chain triglycerides (LCT) can inuence
energy balance and thus it was postulated that they may promote
weight reduction (Dulloo, Fathi, Mensi, & Girardier, 1996; Hill et al.,
1989).

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

Unlike long chain fatty acids (LCFAs), MCFAs are hydrolysed


more rapidly and metabolized more completely after absorption
across the epithelial barrier (Babayan, 1987; Bach & Babayan, 1992).
They are transported directly to the liver via the portal circulation
unlike the LCFA that are preferentially incorporated into chylomicrons as LCT and transported via lymph (Guillot, Vaugelade,
Lemarchal, & Rerat, 1993; Papamandjaris, MacDougall, & Jones,
1998). Once transported to the liver, the MCFA can follow various
catabolic pathways including beta-oxidation, omega-oxidation, and
peroxisomal oxidation (Papamandjaris et al., 1998). Thus, few MCFA
are recovered in triglyceride (Christensen, Gronn, Hagve, &
Christophersen, 1991), phospholipid or cholesterol ester fractions
(Mascioli et al., 1989).
Tsuji et al. (2001) demonstrated that a daily intake of 10 g of
MCT for a 12-week period signicantly reduced body weight, body
fat, abdominal subcutaneous fat and waist and hip sizes in subjects
with a body mass index (BMI) of 23 kg m2, suggesting that intake
of MCT could be effective for preventing obesity in subjects with
a high BMI. In follow-up studies the differences in the dynamics of
postprandial serum lipids between two groups of subjects with
BMI  23 kg m2 and <23 kg m2 after intake of MCT or LCT at
a single dose of 10 g demonstrated that the response of triglyceride
after LCT intake was greater for subjects with BMI  23 kg m2 than
for those with BMI < 23 kg m2 (Kasai et al., 2003a). Moreover,
subjects with BMI  23 kg m2 showed a smaller triglyceride
response after receiving MCT than after receiving LCT and remnantlike lipoprotein cholesterol levels were lower in subjects who
received MCT (Kasai et al., 2003b). Interestingly, the use of
a mixture of both medium- and long-chain triacylglycerols as
substitutes for common edible vegetable oils has resulted in the
FOSHU (Food for Specied Health Use) product with the trade
name Healthy Resseta which has been shown to have a suppressing effect on body fat accumulation and to increase diet induced
thermogenesis and is now widely sold in Japan (Aoyama, Nosaka, &
Kasai, 2007; Ogawa et al., 2007).
4.12. Antimicrobial lipids
Components of human and bovine milk fat have been shown to
exert antimicrobial properties (German & Dillard, 2006; Isaacs,
2005). For example, a number of milk fatty acids were highly
active against the enveloped viruses, where the antiviral fatty acids
were shown to affect the viral envelope, causing leakage, and at
higher concentrations, complete envelope disintegration (Thormar,
Isaacs, Brown, Barshatzky, & Pessolano, 1987). Likewise, synthetic
lipids adapted from naturally occurring compounds found in
human breast milk demonstrated an ability to kill the sexually
transmitted Chlamydia trachomatis by disrupting the chlamydial
inner membrane (Lampe, Ballweber, Isaacs, Patton, & Stamm, 1998).
In another study, lauric acid (C12:0), and linoleic and linolenic
acids were shown to be bactericidal and decreased the invasiveness
of the food-borne pathogen L. monocytogenes in a Caco-2 enterocyte-like cell line (Petrone et al., 1998). Capric acid (C10:0) and
lauric acid and digestion products of sphingolipids from bovine
milk triglycerides and membrane lipids were also shown to be
effective bactericidal agents against a variety of pathogens as well
as L. monocytogenes, such as E. coli 0157, Salmonella enteritidis,
Campylobacter jejuni, and Clostridium perfringens (Sprong, Hulstein,
& Van der Meer, 2001). In a follow-up study, L. monocytogenes and
C. jejuni were shown to be very sensitive to capric acid and lauric
acid and the C18:0 fatty acids, together with digestion products of
sphingolipids, whereas E. coli 0157 and S. enteritidis were less
vulnerable (Sprong, Hulstein, & Van der Meer, 2002). However, the
antimicrobial effect of these fatty acids may be limited to C10:0 and
C12:0 fatty acids in vivo as determined following testing in rats

391

(Sprong et al., 2002). Analysis of the anti-microbial properties of


bovine milk fat after partial hydrolysis by calf pregastric lipase
demonstrated that lauric acid was most potent against the Gram
positive enterococci, while caprylic acid (C8:0) was most potent
against the Gramnegative coliforms (Sun, OConnor, & Roberton,
2002). More recently, the products from lipase-catalysed hydrolysis of bovine milk fat were shown to kill H. pylori in vitro (Sun,
OConnor, MacGibbon, & Roberton, 2007).
Interestingly, free fatty acids from bovine whey cream have been
shown to inhibit the germination of C. albicans in vitro, which was
mainly attributed to lauric acid, myristoleic acid (C14:1n-5), LA and
ARA (Clement, Tremblay, Lange, Thibodeau, & Belhumeur, 2007). A
more recent study demonstrated that capric acid, lauroleic acid
(C12:1), 11-methyldodecanoic acid (iso-C13:0), myristoleic acid
(C14:1n-5), and g-linolenic acid (C18:3n-6) from bovine whey cream
also exhibited antifungal activities against Aspergillus fumigates as
well as C. albicans (Clement, Tremblay, Lange, Thibodeau, &
Belhumeur, 2008).
5. Dairy calcium
Both animal and human studies have clearly indicated that
dietary calcium can play a signicant role in weight management;
remarkably, however, the effect from dairy sources is distinctly
more dramatic than the effect from non-dairy sources (Zemel, 2003;
2004; 2005). It has been suggested that dietary calcium exerts this
effect on weight through the calcitrophic hormones, parathyroid
hormone and 1,25(OH)2D (Zemel, 2004). These hormones have
been shown to respond to low-calcium diets and exert coordinated
regulatory effects on human adipocyte lipogenic and lipolytic
systems (Zemel, 2004). In addition, high calcium diets have been
shown to increase fecal fat excretion (Jacobsen, Lorenzen, Toubro,
Krog-Mikkelsen, & Astrup, 2003; Papakonstantinou, Flatt, Huth, &
Harris, 2003). These fat-reducing effects have been observed in
both animal (Causey & Zemel, 2003; Metz, Karanja, Torok, &
McCarron, 1988; Papakonstantinou et al., 2003; Parra, Bruni,
Palou, & Serra, 2008; Shi, Dirienzo, & Zemel, 2001; Sun & Zemel,
2004; Zemel & Geng, 2001; Zemel & Morgan, 2002; Zemel, Shi,
Greer, Dirienzo, & Zemel, 2000; Zemel, Sun, & Geng, 2001) and
human trials (Zemel, Thompson, Milstead, Morris, & Campbell,
2004) that have been discussed in detail in the following reviews
by Zemel (2003; 2004, 2005).
Dairy sources of calcium have been reported to be 50e100%
more effective than supplemental calcium (Zemel, 2005), which
can probably be best appreciated from the results of the following
clinical trials. In the rst clinical trial (Zemel et al., 2004), 32 obese
adults were maintained on balanced caloric-decit diets
(500 kcal day1 decit) and randomized to control diet
(400e500 mg of calcium day1), high calcium (control diet supplemented with 800 mg of calcium day1) or high dairy (3e4
servings of milk, yoghurt and/or cheese day1, representing a total
calcium intake of 1200e1300 mg day1). Over a 24-week study,
control subjects lost 5.4% of their body weight, subjects on the high
calcium diet lost 8.6%, while those on the dairy diet lost 10.9%
(p < 0.01). Fat loss followed a similar trend. In addition, there was
notable change in the distribution of body fat loss, as fat loss from
the trunk region represented 19% of the total fat lost on the lowcalcium diet, which increased to 50% and to 66% on the high
calcium and high dairy diets, respectively (Zemel et al., 2004). A
follow-up study of 34 obese subjects also supported these ndings
(Zemel et al., 2005a) which was replicated in a six-month clinical
trial in obese African Americans (Zemel, Richards, Milstead, &
Campbell, 2005b). Likewise, dairy calcium intake over 18 months
by young female subjects resulted in reduced fat mass accumulation, although the effect was small (Eagan, Lyle, Gunther, Peacock, &

392

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

Teegarden, 2006). Moreover, a diet rich in dairy calcium was also


shown to enhance weight loss in type-2 diabetic patients (Shahar,
Abel, Elhayany, Vardi, & Fraser, 2007).
A study involving 49 Caucasian postmenopausal women indicated that increased calcium intake was associated with lower body
fat and higher dairy intake was associated with lower fat mass
(Heiss, Shaw, & Carothers, 2008). More recently, a meta-analysis of
randomized controlled trials indicated that increasing dairy
calcium intake by 1241 mg day1 increased faecal fat excretion by
5.2 g day1, which could have signicant consequences for the
prevention of weight (re-)gain (Christensen et al., 2009).
The greater effect observed for dairy calcium has been attributed
to additional bioactive components which have been assigned to
the whey portion of milk (Causey & Zemel, 2003; Ha & Zemel, 2003;
Zemel, 2005). ACE-inhibitory activity (previously discussed) is one
such component which has been postulated as increasing the fatreducing effects of dairy calcium, since angiotensin-II regulates, in
part, adipocyte lipogenesis (Pfeuffer & Schrezenmeir, 2007).
Moreover, branched chain amino acids (BCAA, leucine, isoleucine
and valine) have also been implicated, since in addition to protein
synthesis, these amino acids also play specic metabolic roles as
energy substrates and in the regulation of muscle protein synthesis
(Layman, 2003). As already discussed in this review, MCFAs have
been shown to have a positive effect on weight reduction. Thus, the
weight-reducing effects of milk possibly result from a collection of
effects rather than being exerted from a single mechanism.
However, not all studies have concluded that dietary calcium or
indeed dairy calcium has a positive effect on obesity and this
controversy relating to results from clinical trials has been discussed in the following reviews; Barba and Russo (2006), Barr
(2003) and Teegarden (2005). Indeed, some studies have failed to
identify the benecial effects associated with reductions in weight
loss following calcium consumption (supplemental or dairy). For
example, in a study of an obesity-prone population consisting of 65
Indian Pima adults (35 men and 30 women aged 33  8 years) and
78 Pima Indian children (36 boys and 42 girls, aged 10  0.3 years),
no association was observed between dietary calcium intake and
body size or adiposity (Venti, Tataranni, & Salbe, 2005). The authors
concluded that the high-fat energy diet consumed by the volunteers possibly overwhelmed the anti-obesity effect of calcium. A
one-year intervention trial of 155 young (18e30 years), healthy,
normal weight women consuming a medium dairy diet (calcium
intake of w1000e1100 mg day1) or high dairy diet (calcium intake
of 1300e1400 mg day1) indicated that the increased intake of
dairy products did not alter body weight or fat mass (Gunther et al.,
2005). Longer term trials are required to thoroughly investigate the
effects of dietary calcium as changes in fat mass may be too small to
detect in short periods of time. In addition, several other dietary
factors must be considered such as total energy intake, dietary
protein content and vitamin D status. Moreover, nutrigenomics is
proving that different populations and indeed different individuals
exert variable responses to the same dietary components. Barba
and Russo (2006) have suggested that the complex composition
of dairy foods may be responsible for this biological variability.
Overall, the evidence suggests that dairy foods have an important
role to play in weight regulation. In the future, a better understanding of the mechanisms involved in dairy calcium and weight
loss will enable reliable recommendations for dairy calcium
consumption which will potentially help reduce the development
of obesity for many individuals.
6. Microbial diversity inscribed in human milk
The origin of lactobacilli and bidobacteria that colonise the
neonatal gut has in the past been attributed to contamination of the

infant with maternal microorganisms upon passage through the


birth canal. However, recent studies have provided clear evidence
that human milk is a direct source of both lactobacilli (Martin,
Heilig, Zoetandal, Smidt, & Rodriguez, 2007a) and bidobacteria
(Gueimonde, Laitinen, Salminen, & Isolauri, 2007; Martin et al.,
2009), and LAB colonisation may not be signicantly related to
the mode of delivery (Martin et al., 2007b). Indeed, Martn et al.
(2006) successfully tracked a Lactobacillus salivarius isolate, which
was shown to have potential probiotic properties, from the faeces
of a one-month-old breast-fed infant to the breast milk of the
respective mother through the use of DNA nger-printing. Moreover, in a related study, analysis of the probiotic potential of three
more Lactobacillus isolates from breast milk indicated that each
strain harboured probiotic properties comparable to strains
commonly used in commercial probiotic products (Martn et al.,
2005) and all four strains were shown to possess potent antimicrobial activities (Oliveras, Diaz-Ropero, Martn, Rodriguez, & Xaus,
2006). In the case of bidobacteria, B. longum was found to be the
dominant species in human breast milk followed by Bidobacterium animalis, B. bidum and Bidobacterium catenulatem following
the analysis of 20 human milk samples (Gueimonde et al., 2007).
The source of such bacteria in the milk perhaps represents
a unique form of mother-infant communication (Perez et al., 2007).
Indeed, in a compelling study by Perez et al. (2007), it was
demonstrated that human breast milk cells contain a limited
number of viable bacteria but a range of bacterial DNA signatures
which are also found in maternal peripheral blood mononuclear
cells. The results of the study suggest that intestinally derived
bacterial components are transported to the lactating breast within
mononuclear cells. The authors proposed that this directs the
neonatal immune system to recognize specic bacterial molecular
patterns and to respond appropriately to pathogens and
commensals (Perez et al., 2007). In support of this, bacterial DNA
has been shown to stimulate innate immunity in pregnant mice,
improve maternal survival rates and prevent pathogen transmission to the fetus (Ito, Ishii, Shirota, & Klinman, 2004). Thus, it
appears that through human milk the infant is immediately
equipped with a strategy for immune surveillance at the earliest
stages of life.
7. Conclusion and outlook for the future
Milk provides a plethora of bioactive ingredients for incorporation into functional food products. This has come at a time when
consumers want more from food than just basic nutrition, but
rather a capacity to alleviate the onset of lifestyle diseases through
diet. Using milk as a model, system could prove invaluable for
developing designer foods or even designer milk, a concept that
has been discussed in a review by Sabikhi (2007). However, it
should be noted that many of the physiological effects observed for
the bioactive components of milk have only been proven in vitro or
in animal models and have yet to be proven in humans. Another
major challenge facing food scientists and manufacturers alike is
the cost-effective large-scale production of milk bioactive ingredients. For example, while the potential for milk proteins and
peptides as ingredients in functional food products has been well
documented on a vast scale, their large-scale production and
commercialisation is still limited. Major efforts are now underway
to develop methods to ensure optimal activity of these agents in
food systems and their subsequent utilisation in the body. One
group has recently cloned the 11-residue antimicrobial peptide
from bovine lactoferrin (BL-11) and the 12-residue hypotensive
peptide from as1-casein (C-12) in the dairy starter culture Streptococcus thermophilus, which is utilised in the production of yoghurt
and various cheeses (Renye & Somkuti, 2008). Multiple repeats of

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

the hypocholesterolemic peptide IIAEK, derived from bovine milk


b-lactoglobulin, have been introduced into the ve variable regions
of soybean proglycinin A1aB1b and when expressed in E. coli have
demonstrated the large-scale production of a small peptide of
fewer than 10 amino acids (Prak, Maruyama, Maruyama, & Utsumi,
2006; Prak & Utsumi, 2009). The potent antihypertensive peptide
derived from ovalbumin, RPLKPW (novokinin), has recently been
incorporated into the soybean through transgenesis (Yamada et al.,
2008). Through such approaches milk bioactive peptide sequences
may even be incorporated into food proteins of non-dairy origin.
However, using molecular genetic approaches to increase
bioavailability or production of therapeutic peptides is perhaps an
unrealistic goal, considering the strict regulations governing the
use of genetically modied organisms in the food industry in
certain parts of the world. The existing modern technologies
applicable for the isolation of bioactive native proteins and peptides
from milk are beyond the scope of this review but have been discussed in detail in reviews by Korhonen and Pihlanto (2003, 2007).
HMOs are essential components for the correct functioning of
numerous physiological systems: they modulate the growth of the
intestinal microbiota; they prevent adhesion of pathogenic microorganisms to the intestinal mucosa; through selectin binding, they
behave as anti-inammatory mediators by reducing the binding of
platelets to neutrophils; they may play a role in the development of
the infant brain, through sialylated oligosaccharides. However, the
variation between bovine milk oligosaccharides and HMOs is
extensive and studies are indicating that the diversity and vast
number of HMOs are essential to the correct overall development of
the infant. Strategies to emulate the oligosaccharide content of
human milk are ongoing (Espinosa, Tamez, & Prieto, 2007),
although to date FOS and GOS have proven successful as prebiotic
ingredients for use in infant formula. Goats milk may also hold
potential as a natural source of lactose-derived oligosaccharides as
supplements for infant formulas and for the development of
functional foods, having been successfully isolated by membrane
technology (Martinez-Ferez et al., 2006). The production of HMOs
on a large-scale has been achieved in the past through the use of
metabolically engineered bacteria by direct fermentation of inexpensive carbon sources (Priem, Gilbert, Wakarchuk, Heyraud, &
Samain, 2002). Future research must now focus on means of
determining which structures exert an advantageous biological role
and an economic means of harvesting such oligosaccharides, which
are currently nding themselves at the core of a new food additives
generation (Barreteau, Delattre, & Michaud, 2006) and as potential
ingredients in drug development (Kobata, 2003).
While milk fat serves as a source of energy for the neonate of
each species, it is also a source of bioactive agents which can
inuence all aspects of physiology from the immune system to the
CNS. In addition, its components exert both antibacterial and
antiviral activities. Transgenesis may in the future provide an
approach towards altering the fatty acid composition of milk.
Indeed, the expression of a rat stearoyl-CoA desaturase gene under
the control of the bovine b-lactoglobulin promoter in transgenic
dairy goats altered the fatty acid composition of milk, resulting in
a less saturated and more monounsaturated prole (Reh et al.,
2004). However, milk fat, unlike other components, is more
pliable in that its content can be manipulated through diet. Indeed,
the ability to manipulate the content of certain fatty acids, such as
CLA, has enabled scientists to directly enhance the therapeutic
properties of milk and dairy products as a result.
In conclusion, it appears that almost every component of milk
provides benecial therapeutic effects. This may not be surprising
given that milk is the natural food having evolved under selective
pressure to meet the nutritional needs of its species. Now scientists
are going beyond the components of milk itself and examining the

393

genomics of milk, the genes that code the composition of milk in


the Milk Genomics Consortium (German et al., 2006), with the
expectation that such an approach will reveal the principal biological denition of mammalian nutrition. The future for milk
research therefore looks brighter than ever and no doubt will
continue in this direction as scientists learn more about this
complex, yet highly organized food which should serve as a model
system for creating superior functional food products.

References
Adamson, N. J., & Reynolds, E. C. (1995). Characterisation of tryptic casein phosphopeptides prepared under industrially relevant conditions. Biotechnology and
Bioengineering, 45, 196e204.
Adamson, N. J., & Reynolds, E. C. (1996). Characterisation of casein phosphopeptides
prepared using alcalase: determination of enzyme specicity. Enzyme and
Microbial Technology, 19, 202e207.
Ahn, E., & Schroeder, J. J. (2002). Sphingoid bases and ceramide induce apoptosis in
HT-29 and HCT-116 human colon cancer cells. Experimental Biology and Medicine, 227, 345e353.
Ahn, E. H., Chang, C. C., & Schroeder, J. J. (2006). Evaluation of sphinganine and
sphingosine as human breast cancer chemotherapeutic and chemopreventive
agents. Experimental Biology and Medicine, (Maywood), 231, 1664e1672.
Aimutis, W. R. (2004). Bioactive properties of milk proteins with particular focus on
anticariogenesis. Journal of Nutrition, 134, 989Se995S.
Akalin, S., Gonc, S., & Unal, G. (2006). Functional properties of bioative components
of milk fat in metabolism. Pakistan Journal of Nutrition, 5, 194e197.
Alessandri, J. M., Guesnet, P., Vancassel, S., Astorg, P., Denis, I., Langelier, B., et al.
(2004). Polyunsaturated fatty acids in the central nervous system: evolution of
concepts and nutritional implications throughout life. Reproduction Nutrition
Development, 44, 509e538.
Andrieux, A., Hudry-Clergeon, G., Ryckewaert, J. J., Chapel, A., Ginsberg, M. H.,
Plow, E. F., et al. (1989). Amino acid sequences in brinogen mediating its
interaction with its platelet receptor, GPIIbIIIa. Journal of Biological Chemistry,
264, 9258e9265.
Aoyama, T., Nosaka, N., & Kasai, M. (2007). Research on the nutritional characteristics of medium-chain fatty acids. Journal of Medical Investigation, 54, 385e388.
Arakawa, T., Chong, D. K. X., Slattery, C. W., & Langridge, W. H. R. (1999). Improvements in human health through production of human milk proteins in transgenic
food plants. Advances in Experimental and Medical Biology, 464, 149e159.
Arao, K., Yotsumoto, H., Han, S. Y., Nagao, K., & Yanagita, T. (2004). The 9cis,
11trans,13cis isomer of conjugated linolenic acid reduces apoliprotein B100
secretion and triacylglycerol synthesis in HepG2 cells. Bioscience, Biotechnology,
and Biochemistry, 68, 2643e2645.
Argov, N., Lemay, D. G., & German, J. B. (2008). Milk fat globule structure and
function: nanoscience comes to milk production. Trends in Food Science and
Technology, 19, 617e623.
Arslanoglu, S., Moro, G. E., & Boehm, G. (2007). Early supplementation of prebiotic
oligosaccharides protects formula-fed infants against infections during the rst
6 months of life. Journal of Nutrition, 137, 2420e2424.
Arslanoglu, S., Moro, G. E., Schmitt, J., Tandoi, L., Rizzardi, S., & Boehm, G. (2008).
Early dietary intervention with a mixture of prebiotic oligosaccharides reduces
the incidence of allergic manifestations and infections during the rst two years
of life. Journal of Nutrition, 138, 1091e1095.
Asakuma, S., Akahori, M., Kimura, K., Watanabe, Y., Nakamura, T., Tsunemi, M., et al.
(2007). Sialyl oligosaccharides of human colostrum: changes in concentration
during the rst three days of lactation. Bioscience, Biotechnology, and Biochemistry, 71, 1447e1451.
Asakuma, S., Urashima, T., Akahori, M., Obayashi, H., Nakamura, T., Kimura, K., et al.
(2008). Variation of major neutral oligosaccharides levels in human colostrum.
European Journal of Clinical Nutrition, 62, 488e494.
Babayan, V. K. (1987). Medium chain triglycerides and structured lipids. Lipids, 22,
417e420.
Bach, A. C., & Babayan, V. K. (1992). Medoim-chain triglycerides: an update.
American Journal of Clinical Nutrition, 36, 950e962.
Backhed, F., & Crawford, P. A. (2009). Coordinated regulation of the metabolome and
lipidome at the host-microbial interface. Biochimica et Biophysica Acta, 1801,
240e245.
Backhed, F., Ding, H., Wang, T., Hooper, L. V., Koh, G. Y., Nagy, A., et al. (2004). The
gut microbiota as an environmental factor that regulates fat storage. Proceedings of the National Academy of Sciences, USA, 101, 15718e15723.
Bakkar-Zierikzee, A. M., Alles, M. S., Knol, J., Kok, F. J., Tolboom, J. T. M., &
Bindels, J. G. (2005). Effects of infant formula containing a mixture of galactoand fructooligosaccharides or viable Bidobacterium animalis on the intestinal
microora during the rst 4 months of life. British Jounal of Nutrition, 94,
783e790.
Barba, G., & Russo, P. (2006). Dairy foods, dietary calcium and obesity: a short
review of the evidence. Nutrition, Metabolism and Cardiovascular Diseases, 20,
691e762.
Barr, S. I. (2003). Increased dairy product or calcium intake: is body weight or
composition affected in humans? Journal of Nutrition, 133, 245Se248S.

394

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

Barreteau, H., Delattre, C., & Michaud, P. (2006). Production of oligosaccharides as


promising new food additive generation. Food Technology and Biotechnology, 44,
323e333.
Bassett, C. M., Edel, A. L., Patenaude, A. F., McCullough, R. S., Blackwood, D. P.,
Chouinard, P. Y., et al. (2010). Dietary vaccenic acid has antiatherogenic effects
in LDLr-/- mice. Journal of Nutrition, 140, 18e24.
Bavington, C., & Page, C. (2005). Stopping bacterial adhesion: a novel approach to
treating infections. Respiration, 72, 335e344.
Beaulieu, J., Dubuc, R., Beaudet, N., Dupont, C., & Lemieux, P. (2007). Immunomodulation by a malleable matrix composed of fermented whey proteins and
lactic acid bacteria. Journal of Medicinal Food, 10, 67e72.
Belem, M. A. F., Gibbs, B. F., & Lee, B. H. (1999). Proposing sequences for peptides
derived from whey fermentation with potential bioactive sites. Journal of Dairy
Science, 82, 486e493.
Belobrajdic, D. P., & McIntosh, G. H. (2000). Dietary butyrate inhibits NMU-induced
mammary cancer in rats. Nutrition and Cancer, 36, 217e223.
Benjamin, S., & Spener, F. (2009). Conjugated linoleic acids as functional food: an
insight into their health benets. Nutrition and Metabolism (London), 6, 36.
Berra, B., Colombo, I., Sottocornola, E., & Giacosa, A. (2002). Dietary sphingolipids in
colorectal cancer prevention. European Journal of Cancer, 1, 193e197.
Bhattacharya, A., Banu, J., Rahman, M., Causey, J., & Fernandes, G. (2006). Biological
effects of conjugated linoleic acids in health and disease. Journal of Nutrional
Biochemistry, 17, 789e810.
Bif, A., Coradini, D., Larsen, R., Riva, L., & Di Fronzo, G. (1997). Antiproliferative
effect of fermented milk on the growth of a human breast cancer cell line.
Nutrition and Cancer, 28, 93e99.
Bode, L., Kunz, C., Muhly-Reinholz, M., Mayer, K., Seeger, W., & Rudloff, S. (2004a).
Inhibition of monocyte, lymphocyte, and neutrophil adhesion to endothelial
cells by human milk oligosaccharides. Journal of Thrombosis and Haemostasis,
92, 1402e1410.
Bode, L., Rudloff, S., Kunz, C., Strobel, S., & Klein, N. (2004b). Human milk oligosaccharides reduce platelet-neutrophil complex formation leading to a decrease in
neutrophil beta 2 integrin expression. Journal of Leukocye Biology, 76, 820e826.
Boehm, G., & Stahl, B. (2003). Oligosaccharides. In T. Mattila-Sandholm (Ed.),
Functional dairy products (pp. 203e243). Cambridge, UK: Woodhead publishers.
Boehm, G., & Stahl, B. (2007). Oligosaccharides from milk. Journal of Nutrition, 137,
847Se849S.
Boehm, G., Lidestri, M., Casetta, P., Jelinek, J., Negretti, F., Stahl, B., et al. (2002).
Supplementation of a bovine milk formula with an oligosaccharide mixture
increases counts of faecal bidobacteria in preterm infants. Archives of Disease
in Childhood. Fetal and Neonatal Edition, 86, F178eF181.
Boehm, G., Stahl, B., Jelinek, J., Knol, J., Miniello, V., & Moro, G. E. (2005). Prebiotic
carbohydrates in human milk and formulas. Acta Pediatrica Supplement, 94,
18e21.
Bogoch, S. (1977). Recognins and their chemoreciprocals. In F. V. DeFeudis, &
J. M. Delgado (Eds.), Behavioural neurochemistry (pp. 270). New York, NY, USA:
Spectrum Publishers.
Bonjour, J. P., Chevalley, T., Ferrari, S., & Rizzoli, R. (2005). Milk and bone health:
essential role of calcium and proteins. Cahiers de Nutrition et de Dietetique, 40,
1S12e11S19.
Bornet, F. R. J. (1994). Undigestible sugars in food products. American Journal of
Clinical Nutrition, 59, 763Se769S.
Bouhnik, Y., Flouri, B., Riottot, M., Bisetti, N., Gailing, M. F., Guibert, A., et al. (1996).
Effects of fructo-oligosaccharides ingestion on fecal bidobacteria and selected
metabolic indexes of colon carcinogenesis in healthy humans. Nutrition and
Cancer, 26, 21e29.
Bounous, G. (2000). Whey protein concentrate (WPC) and glutathione modulation
in cancer treatment. Anticancer Research, 20, 4785e4792.
Bronnum, H., Seested, T., Hellgren, L. I., Brix, S., & Frokiaer, H. (2005). Milk-derived
GM(3) and GD(3) differentially inhibit dendritic cell maturation and effector
functionalities. Scandinavian Journal of Immunology, 61, 551e557.
Brunngraber, E. G., Witting, L. A., Haberland, C., & Brown, B. (1972). Glycoproteins in
Tay-sachs disease: isolation and carbohydrate composition of glycopeptides.
Brain Research, 38, 151e162.
Bruzzese, E., Volpicelli, M., Squaglia, M., Tartaglione, A., & Guarino, A. (2006). Impact
of prebiotics on human health. Digestive and Liver Disease, 38, S283eS287.
Buccoliero, R., & Futerman, A. H. (2003). The roles of ceramide and complex
sphingolipids in neuronal cell function. Pharmacological Research, 47, 409e419.
Burdge, G. C., Tricon, S., Morgan, R., Kliem, K. E., Childs, C., Jones, E., et al. (2005).
Incorporation of cis-9, trans-11 conjugated linoleic acid and vaccenic acid
(trans-11 18:1) into plasma and leucocyte lipids in healthy men consuming
dairy products naturally enriched in these fatty acids. British Journal of Nutrition,
94, 237e243.
Burman, L. G., Berglund, B., Huovinen, P., & Tullus, K. (1999). Effect of ampicillin
versus cefuroxime on the emergence of beta-lactam resistance in fecal Enterobacter cloacae isolates from neonates. Journal of Antimicrobial Chemotherapy,
31, 111e116.
Butikofer, U., Meyer, J., Sieber, R., Walther, B., & Wechsler, D. (2008). Occurrence of
the angiotensin-converting enzyme inhibiting tripeptides Val-Pro-Pro and IlePro-Pro in different cheese varieties of Swiss origin. Journal of Dairy Science, 91,
29e38.
Campagna, S., Mathot, A. G., Fleury, Y., Girardet, J. M., & Gaillard, J. L. (2004).
Antibacterial activity of lactophoricin, a synthetic 23-residues peptide derived
from the sequence of bovine milk component-3 of proteose peptone. Journal of
Dairy Science, 87, 1621e1626.

Cani, P. D., & Delzenne, N. M. (2009). The role of the gut microbiota in energy
metabolism and metabolic disease. Current Pharmaceutical Design, 15,
1546e1558.
Carden, D. L., & Granger, D. N. (2000). Pathophysiology of ischaemia-reperfusion
injury. Journal of Pathology, 190, 255e266.
Causey, K. R., & Zemel, M. B. (2003). Dairy augmentation of the anti-obesity effect of
Ca in aP2-agouti transgenic mice. Faseb Journal, A746. 453.457.
Cavaletto, M., Giuffrida, M. G., & Conti, A. (2008). Milk fat globule membrane
components - a proteomic approach. Advances in Experimental Medicine and
Biology, 606, 129e141.
Cehreli, S. B., Gurpinar, A. O., Onur, A. M., & Dagli, F. T. (2008). In vitro evaluation of
casein phosphopeptide-amorphous calcium phosphate as a potential tooth
transport medium: viability and apoptosis in L929 broblasts. Dental Traumatology, 24, 314e319.
Chabance, B., Jolles, P., Izquierdo, C., Mazoyer, E., Francoual, C., Drouet, L., et al.
(1995). Characterisation of an antithrombotic peptide from kappa-casein in
newborn plasma after milk ingestion. British Journal of Nutrition, 73, 583e590.
Chaturvedi, P., Warren, C. D., Altaye, M., Morrow, A. L., Ruiz-Palacois, G.,
Pickering, L. K., et al. (2001). Fucosylated human milk oligosaccharides vary
between individuals and over the course of lactation. Glycobiology, 11, 365e372.
Chilliard, Y., Ferlay, A., & Doreau, M. (2001). Effect of different types of forages,
animal fat or marine oils in cows diet on milk fat secretion and composition,
especially conjugated linoleic acid (CLA) and polyunsaturated fatty acids. Livestock Production Science, 70, 31e48.
Christensen, E., Gronn, M., Hagve, T. A., & Christophersen, B. O. (1991). Omegaoxidation of fatty acids studied in isolated liver cells. Biochimica et Biophysica
Acta, 1081, 167e173.
Christensen, R., Lorenzen, J. K., Svith, C. R., Bartels, E. M., Melanson, E. L., Saris, W. H.,
et al. (2009). Effect of calcium from dairy and dietary supplements on faecal fat
excretion: a meta-analysis of randomized controlled trials. Obesity Reviews, 10,
475e486.
Christie, W. W. (1995). Composition and structure of milk lipids. In P. F. Fox (Ed.),
Advanced dairy chemistry (pp. 136). London, UK: Chapman and Hall.
Churruca, I., Fernandez-Quintela, A., & Portillo, M. P. (2009). Conjugated linoleic acid
isomers: differences in metabolism and biological effects. Biofactors, 35,
105e111.
Cinque, B., Di Marzio, L., Centi, C., Di Rocco, C., Riccardi, C., & Grazia Cifone, M.
(2003). Sphingolipids and the immune system. Pharmacological Research, 47,
421e437.
Clare, D. A., & Swaisgood, H. E. (2000). Bioactive milk peptides: a prospectus. Journal
of Dairy Science, 83, 1187e1195.
Clare, D. A., Catignani, G. L., & Swaisgood, H. E. (2003). Biodefense properties of
milk: the role of antimicrobial proteins and peptides. Current Pharmaceutical
Design, 9, 1239e1255.
Clement, M., Tremblay, J., Lange, M., Thibodeau, J., & Belhumeur, P. (2007). Wheyderived free fatty acids suppress the germination of Candida albicans in vitro.
FEMS Yeast Research, 7, 276e285.
Clement, M., Tremblay, J., Lange, M., Thibodeau, J., & Belhumeur, P. (2008). Purication and identication of bovine cheese whey fatty acids exhibiting in vitro
antifungal activity. Journal of Dairy Science, 91, 2535e2544.
Coakley, M., Banni, S., Johnson, M. C., Mills, S., Devery, R., Fitzgerald, G., et al. (2008).
Inhibitory effect of conjugated alpha-linolenic acid from bidobacteria of
intestinal origin on SW480 cancer cells. Lipids, 44, 249e256.
Coakley, M., Barrett, E., Murphy, J. J., Ross, R. P., Devery, R., & Stanton, C. (2007).
Cheese manufacture with milk with elevated conjugated linoleic acid levels
caused by dietary manipulation. Journal of Dairy Science, 90, 2919e2927.
Connor, W. E. (2000). Importance of n-3 fatty acids in health and disease. American
Journal of Clinical Nutrition, 71, 171Se175S.
Coppa, G. V., Bruni, S., Morelli, L., Soldi, S., & Gabrielli, O. (2004). The rst prebiotics
in humans: human milk oligosaccharides. Journal of Clinical Gastroenterology,
38, S80eS83.
Coppa, G. V., Gabrielli, O., Pierani, P., Catassi, C., Carlucci, A., & Giorgi, P. L. (1993).
Change in carbohydrate composition in human milk over 4 months of lactation.
Pediatrica, 91, 637e641.
Coppa, G. V., Pierani, P., Zampini, L., Carloni, I., Carlucci, A., & Gabrielli, O. (1999).
Oligosaccharides in human milk during different phases of lactation. Acta
Pediatrica, 430, S89eS94.
Coppa, G. V., Zampini, L., Galeazzi, T., & Gabrielli, O. (2006). Prebiotics in human
milk: a review. Digestive and Liver Disease, 38, S291eS294.
Costalos, C., Kapiki, A., Apostolou, M., & Papathoma, E. (2008). The effect of
a prebiotic supplemented formula on growth and stool microbiology of term
infants. Early Human Development, 84, 45e49.
Crane, J. K., Azar, S. S., Stam, A., & Newburg, D. S. (1994). Oligosaccharides from
human milk block binding and activity of the Escherichia coli heat-stable
enterotoxin (STa) in T84 intestinal cells. Journal of Nutrition, 124, 2358e2364.
Crittenden, R. G., & Playne, M. J. (1996). Production, properties and applications of
food-grade oligosaccharides. Trends in Food Science and Technology, 7, 353e361.
Cross, K. J., Huq, N. L., & Reynolds, E. C. (2007). Casein phosphopeptides in oral
healthechemistry and clinical applications. Current Pharmaceutical Design, 13,
793e800.
Cuvillier, O., & Levade, T. (2003). Enzymes of sphingosine metabolism as potential
pharmacological targets for therapeutic intervention in cancer. Pharmacological
Research, 47, 439e445.
Das, U. N. (2006). Essential fatty acids: biochemistry, physiology and pathology.
Biotechnology Journal, 1, 420e439.

S. Mills et al. / International Dairy Journal 21 (2011) 377e401


De Simone, C., Picariello, G., Mamone, G., Stiuso, P., Dicitore, A., Vanacore, D., et al.
(2009). Characterisation and cytomodulatory properties of peptides from
Mozzarella di Bufala Campana cheese whey. Journal of Peptide Science, 15,
251e258.
Destaillats, F., Trottier, J. P., Galvez, J. M., & Angers, P. (2005). Analysis of alphalinolenic acid biohydrogenation intermediates in milk fat with emphasis on
conjugated linolenic acids. Journal of Dairy Science, 88, 3231e3239.
DiBaise, J. K., Zhang, H., Crowell, M. D., Krajmalnik-Brown, R., Decker, G. A., &
Rittmann, B. E. (2008). Gut microbiota and its possible relationship with
obesity. Mayo Clinic Proceedings, 83, 460e469.
Dillehay, D. L., Webb, S. K., Schmelz, E. M., & Merrill, A. H., Jr. (1994). Dietary
sphingomyelin inhibits 1,2-dimethylhydrazine-induced colon cancer in CF1
mice. Journal of Nutrition, 124, 615e620.
Drake, F. H., Dodds, R. A., James, I. E., Connor, J. R., Debouck, C., Richardson, S., et al.
(1996). Cathepsin K, but not cathepsins B, L, or S, is abundantly expressed in
human osteoclasts. Journal of Biological Chemistry, 271, 12511e12516.
Drouet, L., Bal dit Sollier, C., Cisse, M., Pignaud, G., Mazoyer, E., Fiat, A. M., et al.
(1990). The antithrombotic effect of KRDS, a lactotransferrin peptide, compared
with RGDS. Nouvelle Revue Francaise dHematologie, 32, 59e62.
Dulloo, A. G., Fathi, M., Mensi, N., & Girardier, L. (1996). Twenty-four-hour energy
expenditure and urinary catecholamines of humans consuming low-tomoderate amounts of medium-chain triglycerides: a dose-response study in
a human respiratory chamber. European Journal of Clinical Nutrition, 50, 152e158.
Dupertuis, Y. M., Meguid, M. M., & Pichard, C. (2007). Colon cancer therapy: new
perspectives of nutritional manipulations using polyunsaturated fatty acids.
Current Opinion in Clinical Nutrition and Metabolic Care, 10, 427e432.
Eagan, M. S., Lyle, R. M., Gunther, C. W., Peacock, M., & Teegarden, D. (2006). Effect of
1-year dairy product intervention on fat mass in young women: 6-month
follow-up. Obesity (Silver Spring), 14, 2242e2248.
Eckburg, P. B., Bik, E. M., Bernstein, C. N., Purdom, E., Dethlefsen, L., Sargent, M., et al.
(2005). Diversity of the human intestinal microbial ora. Science, 308,
1635e1638.
Egge, H. (1993). The diversity of oligosaccharides in human milk. In B. Renner, &
G. Sawatski (Eds.), New perspectives in infant nutrition (pp. 12e26). Stuttgart,
Germany: Thieme.
Erba, D., Ciappellano, S., & Testolin, G. (2002). Effect of the ratio of casein phosphopeptides to calcium (w/w) on passive calcium transport in the distal small
intestine of rats. Nutrition, 18, 743e746.
ESPGHAN Committee on Nutrition. (2004). Prebiotic oligosaccharides in dietetic
products for infants: a commentary by the ESPGHAN Committee on Nutrition.
Journal of Pediatric Gastroenterology and Nutrition, 39, 465e473.
Espinosa, R. M., Tamez, M., & Prieto, P. (2007). Efforts to emulate human milk
oligosaccharides. British Journal of Nutrition, 98(Suppl. 1), S74eS79.
Fanaro, S., Boehm, G., Garssen, J., Knol, J., Mosca, F., Stahl, B., et al. (2005a). Galactooligosaccharides and long-chain fructooligosaccharides as prebiotics in infant
formulas: a review. Acta Pediatrica Supplement, 94, S22eS26.
Fanaro, S., Jelinek, J., Stahl, B., Boehm, G., Kock, R., & Vigi, V. (2005b). Acidic oligosaccharides from pectin hydrolysate as new component for infant formulae:
effect on intestinal ora, stool characteristics, and pH. Journal of Pediatric
Gastroenterology and Nutrition, 41, 186e190.
Fedacko, J., Pella, D., Mechirova, V., Horvath, P., Rybar, R., Varjassyova, P., et al.
(2007). n-3 PUFAs-From dietary supplements to medicines. Pathophysiology, 14,
127e132.
Fergusson, D. M., Beautrais, A. L., & Silva, P. (1982). Breast-feeding and cognitive
development in the rst seven years of life. Social Science and Medicine, 16,
1705e1708.
Ferraretto, A., Signorile, A., Gravaghi, C., Fiorilli, A., & Tettamanti, G. (2001). Casein
phosphopeptides inuence calcium uptake by cultured human intestinal HT-29
tumor cells. Journal of Nutrition, 131, 1655e1661.
Fiat, A. M., Migliore, D., & Jolles, P. (1993). Biologically active peptides from milk
proteins with emphasis on two examples concerning antithrombotic and
immunostimulating activities. Journal of Dairy Science, 76, 301e310.
Floris, R., Recio, I., Berkhout, B., & Visser, S. (2003). Antibacterial and antiviral effects
of milk proteins and derivatives thereof. Current Pharmaceutical Design, 9,
1257e1275.
Foltz, M., Meynen, E. E., Bianco, V., van Platerink, C., Koning, T. M., & Kloek, J. (2007).
Angiotensin converting enzyme inhibitory peptides from a lactotripeptideenriched milk beverage are absorbed intact into the circulation. Journal of
Nutrition, 137, 953e958.
Fox, P. F., & McSweeney, P. L. H. (1998). Milk lipids; Dairy chemistry and biochemistry.
New York, NY, USA: Black Academic & Professional.
Frestedt, J. L., Zenk, J. L., Kuskowski, M. A., Ward, L. S., & Bastian, E. D. (2008).
A whey-protein supplement increases fat loss and spares lean muscle in obese
subjects: a randomized human clinical study. Nutriton and Metabolism (London),
5, 8e14.
Fuller, R. (1989). Probiotics in man and animals. Journal of Applied Bacteriology, 66,
365e378.
Gallier, S., Gragson, D., Jimenez-Flores, R., & Everett, D. (2010). Using confocal laser
scanning microscopy to probe the milk fat globule membrane and associated
proteins. Journal of Agricultural and Food Chemistry, 58, 4250e4257.
Gardiner, G. E., OFlaherty, S., Casey, P. G., Weber, A., McDonald, T. L., Cronin, M., et al.
(2009). Evaluation of colostrum-derived human mammary-associated serum
amyloid A3 (M-SAA3) protein and peptide derivatives for the prevention of
enteric infection: in vitro and in murine models of intestinal disease. FEMS
Immunology and Medical Microbiology, 55, 404e413.

395

German, J. B., & Dillard, C. J. (2006). Composition, structure and absorption of milk
lipids: a source of energy, fat-soluble nutrients and bioactive molecules. Critical
Reviews in Food Science and Nutrition, 46, 57e92.
German, J. B., Freeman, S. L., Lebrilla, C. B., & Mills, D. A. (2008). Human milk oligosaccharides: evolution, structures and bioselectivity as substrates for intestinal
bacteria. Nestle Nutrition Workshop Series: Pediatric Program, 62, 205e222.
German, J. B., Schanbacher, F. L., Lonnerdal, B., Medrano, J. F., McGuire, M. A.,
MacManaman, J. L., et al. (2006). International milk genomics consortium.
Trends in Food Science and Technology, 17, 656e661.
Gibson, R. A., & Kneebone, G. M. (1981). Fatty acid composition of human colostrum
and mature breast milk. American Journal of Clinical Nutrition, 34, 252e257.
Gill, H. S., Doull, F., Rutherfurd, K. J., & Cross, M. L. (2000). Immunoregulatory
peptides in bovine milk. British Journal of Nutrition, 84, S111eS117.
Gill, S. R., Pop, M., Deboy, R. T., Eckburg, P. B., Turnbaugh, P. J., Samuel, B. S., et al.
(2006). Metagenomic analysis of the human distal gut microbiome. Science, 312,
1355e1359.
Gobbetti, M., Minervini, F., & Rizzello, C. G. (2004). Angiotensin I-convertingenzyme-inhibitory and antimicrobial bioactive peptides. International Journal of
Dairy Technology, 57, 173e188.
Gonzalez, R., Klaassens, E. S., Malinen, E., de Vos, W. M., & Vaughan, E. E. (2008).
Differential transcriptional response of Bidobacterium longum to human milk,
formula milk, and galactooligosaccharide. Applied and Environmental Microbiology, 74, 4686e4694.
Gopal, P. K., & Gill, H. S. (2000). Oligosaccharides and glycoconjugates in bovine
milk and colostrum. British Journal of Nutrition, 84(Suppl. 1), S69eS74.
Greenberg, R., Groves, M. L., & Dower, H. J. (1984). Human beta-casein. Amino acid
sequence and identication of phosphorylation sites. Journal of Biological
Chemistry, 259, 5132e5138.
Grenby, T. H., Andrews, A. T., Mistry, M., & Williams, R. J. H. (2001). Dental cariesprotective agents in milk products: Investigations in vitro. Journal of Dentistry,
29, 83e92.
Griinari, J. M., & Bauman, D. E. (1999). Biosynthesis of conjugated linoleic acid and
its incorporation into meat and milk in ruminants. In M. P. Yurawecz,
M. M. Mossoba, J. K. G. Kramer, M. W. Pariza, & G. J. Nelson (Eds.), Advances in
conjugated linoleic acid (pp. 180e200). Champaign, IL, USA: AOCS Press.
Griinari, J. M., Corl, B. A., Lacy, S. H., Chouinard, P. Y., Nurmela, K. V., & Bauman, D. E.
(2000). Conjugated linoleic acid is synthesized endogenously in lactating dairy
cows by Delta(9)-desaturase. Journal of Nutrition, 130, 2285e2291.
Gronlund, M. M., Lehtonen, O. P., Eerola, E., & Kero, P. (1999). Fecal microora in
healthy infants born by different methods of delivery: permanent changes in
intestinal ora after caesarean delivery. Journal of Pediatric Gastroenterology and
Nutrition, 28, 19e25.
Gueimonde, M., Laitinen, K., Salminen, S., & Isolauri, E. (2007). Breast milk: a source
of bidobacteria for infant gut development and maturation? Neonatalogy, 92,
64e66.
Guillot, E., Vaugelade, P., Lemarchal, P., & Rerat, A. (1993). Intestinal absorption and
liver uptake of medium-chain fatty acids in non-anaesthetized pigs. British
Journal of Nutrition, 69, 431e442.
Gunther, C. W., Legowski, P. A., Lyle, R. M., McCabe, G. P., Eagan, M. S., Peacock, M.,
et al. (2005). Dairy products do not lead to alterations in body weight or fat
mass in young women in a 1-y intervention. American Journal of Clinical
Nutrition, 81, 751e756.
Gyorgy, P., Jeanloz, R. W., Von Nicholai, H., & Zilliken, F. (1974). Undialyzable growth
factor for Lactobacillus bidus var. pennsylvanicus. Protective effect of sialic acid
bound to glycoproteins and oligosaccharides against bacterial degradation.
European Journal of Biochemistry, 43, 29e33.
Ha, E., & Zemel, M. B. (2003). Functional properties of whey, whey components, and
essential amino acids: mechanisms underlying health benets for active people
(review). Journal of Nutritional Biochemistry, 14, 251e258.
Hamel, U., Kielwein, G., & Teschemacher, H. (1985). b-Casomorphin immunoreactive
materials in cows milk incubated with various bacterial species. Journal of Dairy
Research, 52, 139e148.
Hamer, H. M., Jonkers, D., Venema, K., Vanhoutvin, S., Troost, F. J., & Brummer, R. J.
(2008). Review article: the role of butyrate on colonic function. Alimentary
Pharmacology and Therapeutics, 104, 104e119.
Hamilton, J. A., Hillard, C. J., Spector, A. A., & Watkins, P. A. (2007). Brain uptake and
utilisation of fatty acids, lipids and lipoproteins: applications to neurological
disorders. Journal of Molecular Neuroscience, 33, 2e11.
Harmsen, H. J., Wildeboer-Veloo, A. C., Raangs, G. C., Wagendorp, A. A., Klijn, N.,
Bindells, J. G., et al. (2000). Analysis of intestinal ora development in breastfed and formula-fed infants by using molecular identication and detection
methods. Journal of Pediatric Gastroenterology and Nutrition, 30, 61e67.
Hartmann, R., & Meisel, H. (2007). Food-derived peptides with biological activity:
from research to food applications. Current Opinion in Biotechnology, 18,
163e169.
Hata, Y., Yamamoto, M., Ohni, M., Nakajima, K., Nakamura, Y., & Takano, T.
(1996). A placebo-controlled study of the effect of sour milk on blood
pressure in hypertensive subjects. American Journal of Clinical Nutrition, 64,
767e771.
Haug, A., Hostmark, A. T., & Harstad, O. M. (2007). Bovine milk in human
nutritionea review. Lipids in Health and Disease, 6, 25.
Haug, A., Sjogren, P., Holland, N., Muller, H., Kjos, N. P., Taugbol, O., et al. (2008).
Effects of butter naturally enriched with conjugated linoleic acid and vaccenic
acid on blood lipids and LDL particle size in growing pigs. Lipids in Health and
Disease, 7, 31e36.

396

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

Hayes, M., Ross, R. P., Fitzgerald, G. F., Hill, C., & Stanton, C. (2006). Casein-derived
antimicrobial peptides generated by Lactobacillus acidophilus DPC6026. Applied
and Environmental Microbiology, 72, 2260e2264.
Hayes, M., Stanton, C., Fitzgerald, G. F., & Ross, R. P. (2007a). Putting microbes to
work: dairy fermentation, cell factories and bioactive peptides. Part II: bioactive
peptide functions. Biotechnology Journal, 2, 435e449.
Hayes, M., Stanton, C., Slattery, H., O Sullivan, O., Hill, C., Fitzgerald, G. F., et al.
(2007b). Casein fermentate of Lactobacillus animalis DPC6134 contains a range
of novel propeptide angiotensin-converting enzyme inhibitors. Applied and
Environmental Microbiology, 73, 4658e4667.
Hedlund, M., Padler-Karavani, V., Varki, N. M., & Varki, A. (2008). Evidence for
a human-specic mechanism for diet and antibody-mediated inammation in
carcinoma progression. Proceedings of the National Academy of Sciences USA, 105,
18936e18941.
Heiss, C. J., Shaw, S. E., & Carothers, L. (2008). Association of calcium intake and
adiposity in postmenopausal women. Journal of the American College of Nutrition, 27, 260e266.
Hennessy, A. A., Ross, R. P., Stanton, C., Devery, R., & Murphy, J. J. (2007). Development of dairy based functional foods enriched in conjugated linoleic acid
with special reference to rumenic acid. In M. Saarela (Ed.), Functional dairy
products (pp. 443e494). New York, NY, USA: CRC Press.
Hennessy, A. A., Ross, R. P., Devery, R., & Stanton, C. (2011). The health promoting
properties of the conjugated isomers of a-linolenic acid. Lipids, doi:10.1007/
s11745-010-3501-5.
Hertervig, E., Nilsson, A., Cheng, Y., & Duan, R. D. (2003). Puried intestinal alkaline
sphingomyelinase inhibits proliferation without inducing apoptosis in HT-29
colon carcinoma cells. Journal of Cancer Research and Clinical Oncology, 129,
577e582.
Hickey, R. (2009). Harnessing milk oligosaccharides for nutraceutical applications.
In M. Corredig (Ed.), Dairy derived ingredients: Food and nutraceutical uses
(pp. 308e343). Cambridge, UK: Woodhead Publishing.
Hill, J. O., Peters, J. C., Yang, D., Sharp, T., Kaler, M., Abumrad, N. N., et al. (1989).
Thermogenesis in humans during overfeeding with medium-chain triglycerides. Metabolism, 38, 641e648.
Hirota, T., Ohki, K., Kawagishi, R., Kajimoto, Y., Mizuno, S., Nakamura, Y., et al. (2007).
Casein hydrolysate containing the antihypertensive tripeptides Val-Pro-Pro and
Ile-Pro-Pro improves vascular endothelial function independent of blood
pressure-lowering effects: contribution of the inhibitory action of angiotensinconverting enzyme. Hypertension Research, 30, 489e496.
van Hoffen, E., Ruiter, B., Faber, J., MRabet, L., Knol, E. F., Stahl, B., et al. (2009).
A specic mixture of short-chain galacto-oligosaccharides and long-chain
fructo-oligosaccharides induces a benecial immunoglobulin prole in infants
at high risk for allergy. Allergy, 64, 484e487.
Huncharek, M., Muscat, J., & Kupelnick, B. (2008). Impact of dairy products and
dietary calcium on bone-mineral content in children: results of a meta-analysis.
Bone, 43, 312e321.
Innis, S. M. (2003). Perinatal biochemistry and physiology of long-chain polyunsaturated fatty acids. Journal of Pediatrics, 143, S1eS8.
Innis, S. M. (2004). Polyunsaturated fatty acids in human milk: an essential role
in infant development. Advances in Experimental Medicine and Biology, 554,
27e43.
Innis, S. M. (2007a). Dietary (n-3) fatty acids and brain development. Journal of
Nutrition, 137, 855e859.
Innis, S. M. (2007b). Human milk: maternal dietary lipids and infant development.
Proceedings of the Nutrition Society, 66, 397e404.
Isaacs, C. E. (2005). Human milk inactivates pathogens individually, additively, and
synergistically. Journal of Nutrition, 135, 1286e1288.
Ito, S., Ishii, K. J., Shirota, H., & Klinman, D. M. (2004). CpG oligodeoxynucleotides
improve the survival of pregnant and fetal mice following Listeria monocytogenes infection. Infection and Immunity, 72, 3543e3548.
von Itzstein, M., & Thomson, R. J. (1997). Sialic acids and sialic acid-recognising
proteins: drug discovery targets and potential glycopharmaceuticals. Current
Medicinal Chemistry, 4, 185e210.
Jacobsen, R., Lorenzen, J. K., Toubro, S., Krog-Mikkelsen, I., & Astrup, A. (2003).
Effect of short-term high dietary calcium intake on 24-h energy expenditure,
fat oxidation, and fecal fat excretion. International Journal of Obesity, 29,
292e301.
Jauhiainen, T., & Korpela, R. (2007). Milk peptides and blood pressure. Journal of
Nutrition, 137, 825Se829S.
Jensen, R. G. (1999). Lipids in human milk. Lipids, 34, 1243e1271.
Jensen, R. G. (2002). The composition of bovine milk lipids: January 1995 to
December 2000. Journal of Dairy Science, 85, 295e350.
Joerger, R. D. (2003). Alternatives to antibiotics: bacteriocins, antimicrobial peptides
and bacteriophages. Poultry Science, 82, 82640e82647.
Jolles, P. (1975). Structural aspects of the milk clotting process. Comparative
features with the blood clotting process. Molecular and Cellular Biochemistry, 7,
73e85.
Jolles, P., & Henschen, A. (1982). Comparison between the clotting of blood and
milk. Trends in Biochemical Sciences, 7, 325e328.
Jolles, P., Levy-Toledano, S., Fiat, A. M., Soria, C., Gillessen, D., Thomaidis, A., et al.
(1986). Analogy between brinogen and casein. Effect of an undecapeptide
isolated from kappa-casein on platelet function. European Journal of Biochemistry, 158, 379e382.
Jolles, P., Loucheux-Lefebvre, M. H., & Henschen, A. (1978). Structural relatedness of
k-casein and brinogen g-chain. Journal of Molecular Evolution, 11, 271e277.

Jolles, P., Parker, F., Floch, F., Migliore, D., Alliel, P., Zerial, A., et al. (1981). Immunostimulating substances from human casein. Immunopharmacology and
Immunotoxicology, 3, 363e369.
Jones, E. L., Shingeld, K. J., Kohen, C., Jones, A. K., Lupoli, B., Grandison, A. S., et al.
(2005a). Chemical, physical, and sensory properties of dairy products enriched
with conjugatd linoleic acid. Journal of Dairy Science, 88, 2923e2937.
Jones, E. L., Shingeld, K. J., Kohen, C., Jones, A. K., Lupoli, B., Grandison, A. S., et al.
(2005b). Chemical, physical, and sensory properties of dairy products enriched
with conjugated linoleic acid. Journal of Dairy Science, 88, 2923e2937.
Juillard, V., Laan, H., Kunji, E. R., Jeronimus-Stratingh, C. M., Bruins, A. P., &
Konings, W. N. (1995). The extracellular PI-type proteinase of Lactococcus lactis
hydrolyzes beta-casein into more than one hundred different oligopeptides.
Journal of Bacteriology, 177, 3472e3478.
Jungi, T. W., Spycher, M. O., Nydegger, U. E., & Barandun, S. (1986). Platelet-leukocyte
interaction: selective binding of thrombin-stimulated platelets to human
monocytes, polymorphonuclear leukocytes, and related blood cells. Blood, 67,
629e636.
Kalenic, S., Francetic, I., Polak, J., Zele-Starcevic, L., & Bencic, Z. (1993). Impact of
ampicillin and cefuroxime on bacterial colonisation and infection in patients on
a neonatal intensive care unit. Journal of Hospital Infection, 23, 35e41.
Kampa, M., Bakogeorgou, E., Hatzoglou, A., Damianaki, A., Martin, P. M., &
Castanas, E. (1997). Opioid alkaloids and casomorphin peptides decrease the
proliferation of prostatic cancer cell lines (LNCaP, PC3 and DU145) through
a partial intersection with opioid receptors. European Journal of Pharmacology,
335, 255e265.
Kannagi, R. (2002). Regulatory roles of carbohydrate ligands for selectins in the
homing of lymphocytes. Current Opinion in Structural Biology, 12, 599e608.
Kasai, M., Maki, H., Nosaka, N., Aoyama, T., Ooyama, K., Uto, H., et al. (2003a). Effect
of medium-chain triglycerides on the postprandial triglyceride concentration in
healthy men. Bioscience, Biotechnology and Biochemistry, 67, 46e53.
Kasai, M., Maki, H., Suzuki, Y., Nosaka, N., Aoyama, T., Inuzaka, H., et al. (2003b).
Effect of medium chain triglycerides on postprandial concentrations of
remnant-like particles (RLP) in healthy men. Journal of Oleo Science, 52,
197e204.
Kaylegian, K. E., & Lindsay, R. C. (1995). Milk fat usage and modifcation. In
K. E. Kaylegian, & R. C. Lindsay (Eds.), Handbook of milkfat fractionation technology and applications (pp. 1e18). Champaign, IL, USA: AOCS Press.
Kelley, N. S., Hubbard, N. E., & Erickson, K. L. (2007). Conjugated linoleic acid
isomers and cancer. Journal of Nutrition, 137, 2599e2607.
Khanal, R. C., Dhiman, T. R., Ure, A. L., Brennand, C. P., Boman, R. L., & McMahon, D. J.
(2005). Consumer acceptability of conjugated linoleic acid-enriched milk and
cheddar cheese from cows grazing on pasture. Journal of Dairy Science, 88,
1837e1847.
Knol, J., Boehm, G., Lidestri, M., Negretti, F., Jelinek, J., Agosti, M., et al. (2005a).
Increase of faecal bidobacteria due to dietary oligosaccharides induces
a reduction of clinically relevant pathogen germs in the faeces of formula-fed
preterm infants. Acta Pediatrica, 94, S31eS33.
Knol, J., Scholtens, P., Kafka, C., Steenbakkers, J., Gro, S., Helm, K., et al. (2005b).
Colon microora in infants fed formula with galacto- and fructo-oligosaccharides: more like breast-fed infants. Journal of Pediatric Gastroenterology and
Nutrition, 40, 36e42.
Kobata, K. (2003). Possible application of milk oligosaccharides for drug development. Chang Gung Medical Journal, 26, 620e636.
Kohmura, M., Nio, N., & Ariyoshi, Y. (1990). Inhibition of angiotensin I-converting
enzyme by synthetic peptides of human k-casein. Agricultural and Biological
Chemistry, 54, 835e836.
Kohmura, M., Nio, N., Kubo, K., Minoshima, Y., Munekata, E., & Ariyoshi, Y. (1989).
Inhibition of angiotensin I-converting enzyme by synthetic peptides of human
b-casein. Agricultural and Biological Chemistry, 53, 2107e2114.
Korhonen, H. (2009). Milk-derived bioactive peptides: from science to applications.
Journal of Functional Foods, 1, 177e187.
Korhonen, H., & Pihlanto, A. (2003). Food-derived bioactive peptideseopportunities
for designing future foods. Current Pharmaceutical Design, 9, 1297e1308.
Korhonen, H., & Pihlanto, A. (2006). Bioactive peptides: production and functionality. International Dairy Journal, 16, 945e960.
Korhonen, H., & Pihlanto, A. (2007). Technological options for the production of
health-promoting proteins and peptides derived from milk and colostrum.
Current Pharmaceutical Design, 13, 829e843.
Korhonen, H., Pihlanto-Leppala, A., Rantamaki, P., & Tupasela, T. (1998). The functional and biological properties of whey proteins: prospects for the development of functional foods. Agricultural and Food Science, Finland, 7, 283e296.
Kuntz, S., Rudloff, S., & Kunz, C. (2008). Oligosaccharides from human milk inuence
growth-related characteristics of intestinally transformed and non-transformed
intestinal cells. British Journal of Nutrition, 99, 462e471.
Kunz, C., & Rudloff, S. (2002). Health benets of milk-derived carbohydrates. Bulletin
of the International Dairy Federation, 375, 72e79.
Kunz, C., & Rudloff, S. (2008). Potential anti-inammatory and anti-infectious
effects of human milk oligosaccharides. Advances in Experimental and Medical
Biology, 606, 455e465.
Kunz, C., Rudloff, W., Baier, N., Klein, N., & Strobel, S. (2000). Oligosaccharides in
human milk: structural, functional, and metabolic aspects. Annual Reviews of
Nutrition, 20, 699e722.
Lahov, E., & Regelson, W. (1996). Antibacterial and immunostimulating caseinderived substances from milk: caseicidin, isracidin peptides. Food and Chemical
Toxicology, 34, 131e145.

S. Mills et al. / International Dairy Journal 21 (2011) 377e401


Lampe, M. F., Ballweber, L. M., Isaacs, C. E., Patton, D. L., & Stamm, W. E. (1998).
Killing of Chlamydia trachomatis by novel antimicrobial lipids adapted from
compounds in human breast milk. Antimicrobial Agents and Chemotherapy, 42,
1239e1244.
Laroux, F. S., & Grisham, M. B. (2001). Immunological basis of inammatory bowel
disease: role of the microcirculation. Microcirculation, 8, 283e301.
Larson, E., Celi, A., Gilbert, G. E., Furie, B. C., Erban, J. K., Bonfanti, R., et al. (1989).
PADGEM protein: a receptor that mediates the interaction of activated platelets
with neutrophils and monocytes. Cell, 59, 305e312.
Larson, M. A., Wei, S. H., Weber, A., Mack, D. R., & McDonald, T. L. (2003). Human
serum amyloid A3 peptide enhances intestinal MUC3 expression and inhibits
EPEC adherence. Biochemical and Biophysical Research Communications, 300,
531e540.
Larsson, S. C., Kumlin, M., Ingelman-Sundberg, M., & Wolk, A. (2004). Dietary longchain n-3 fatty acids for the prevention of cancer: a review of potential
mechanisms. Americal Journal of Clinical Nutrition, 79, 935e945.
Layman, D. K. (2003). The role of leucine in weight loss diets and glucose homeostasis. Journal of Nutrition, 133, 261Se267S.
LeBlanc, J. G., Matar, C., Valdez, J. C., LeBlanc, J., & Perdigon, G. (2002). Immunomodulating effects of peptidic fractions issued from milk fermented with
Lactobacillus helveticus. Journal of Dairy Science, 85, 2733e2742.
Lemonnier, L. A., Dillehay, D. L., Vespremi, M. J., Abrams, J., Brody, E., & Schmelz, E. M.
(2003). Sphingomyelin in the suppression of colon tumors: prevention versus
intervention. Archives of Biochemistry and Biophysics, 419, 129e138.
Ley, K. (2003). The role of selectins in inammation and disease. Trends in Molecular
Sciences, 9, 263e268.
Ley, R. E., Hamady, M., Lozupone, C., Turnbaugh, P. J., Ramey, R. R., Bircher, J. S., et al.
(2008). Evolution of mammals and their gut microbes. Science, 320, 1647e1651.
LoCascio, R. G., Ninonuevo, M. R., Freeman, S. L., Sela, D. A., Grimm, R., Lebrilla, C. B.,
et al. (2007). Glycoproling of bidobacterial consumption of human milk
oligosaccharides demonstrates strain specic, preferential consumption of
small chain glycans secreted in early human lactation. Journal of Agricultural and
Food Chemistry, 55, 8914e8919.
Lock, A. L., & Bauman, D. E. (2004). Modifying milk fat composition of dairy cows to
enhance fatty acids benecial to human health. Lipids, 39, 1197e1206.
Lock, A. L., Horne, C. A., Bauman, D. E., & Salter, A. M. (2005). Butter naturally
enriched in conjugated linoleic acid and vaccenic acid alters tissue fatty acids
and improves the plasma lipoprotein prole in cholesterol-fed hamsters. Journal of Nutrition, 135, 1934e1939.
Lopez, C., Briard-Bion, V., Menard, O., Rousseau, F., Pradel, P., & Besle, J. M. (2008).
Phospholipid, sphingolipid, and fatty acid compositions of the milk fat globule
membrane are modied by diet. Journal of Agricultural and Food Chemistry, 56,
5226e5236.
Lopez-Exposito, I., & Recio, I. (2008). Protective effect of milk peptides: antibacterial
and antitumor properties. Advances in Experimental Medicine and Biology, 606,
271e293.
Lopez-Garcia, E., Schulze, M. B., Meigs, J. B., Manson, J. E., Rifai, N., Stampfer, M. J.,
et al. (2005). Consumption of trans fatty acids is related to plasma biomarkers
of inammation and endothelial dysfunction. Journal of Nutrition, 135, 562e566.
Losacco, M., Gallerani, R., Gobbetti, M., Minervini, F., & De Leo, F. (2007). Production
of active angiotensin-I converting enzyme inhibitory peptides derived from
bovine beta-casein by recombinant DNA technologies. Biotechnology Journal, 2,
1425e1434.
Lu, M., Taylor, A., Chylack, L. T., Jr., Rogers, G., Hankinson, S. E., Willett, W. C., et al.
(2007). Dietary linolenic acid intake is positively associated with ve-year
change in eye lens nuclear density. Journal of the American College of Nutrition,
26, 133e140.
Lucas, A., Morley, R., & Cole, T. J. (1998). Randomised trial of early diet in preterm
babies and later intelligence quotient. British Medicial Journal, 317, 1481e1487.
Lucas, A., Morley, R., Cole, T. J., Lister, G., & Leeson-Payne, C. (1992). Breast milk and
subsequent intelligence quotient in children born preterm. Lancet, 339,
261e264.
Luhovyy, B. L., Akhavan, T., & Anderson, G. H. (2007). Whey proteins in the regulation of food intake and satiety. Journal of the American College of Nutrition, 26,
704Se712S.
Luo, S. J., & Wong, L. L. (2005). Oral care chewing gums and methods of use. United
States Patent 6846500.
Ma, Q. L., Teter, B., Ubeda, O. J., Morihara, T., Dhoot, D., Nyby, M. D., et al. (2007).
Omega-3 fatty acid docosahexaenoic acid increases SorLA/LR11, a sorting
protein with reduced expression in sporadic Alzheimers disease (AD): relevance to AD prevention. Journal of Neuroscience, 27, 14299e14307.
Mackie, R. I., Sghir, A., & Gaskins, H. R. (1999). Developmental microbial ecology of
the neonatal gastrointestinal tract. American Journal of Clinical Nutrition, 69,
1035Se1045S.
Madureira, A. R., Pereira, C. I., Gomes, A. M. P., Pintado, M. E., & Malcata, F. X. (2007).
Bovine whey proteins - overview of their main biological properties. Food
Research International, 40, 1197e1211.
Marionneau, S., Ruvoen, N., Le Moullac-Vaidye, B., Clement, M., Cailleau-Thomas, A.,
Ruiz-Palacois, G., et al. (2002). Norwalk virus binds to histo-blood group antigens present on gastroduodenal epithelial cells of secretor individuals.
Gastroenterology, 122, 1967e1977.
Martin, R., Heilig, G. H. J., Zoetandal, E. G., Smidt, H., & Rodriguez, J. M. (2007a).
Diversity of the Lactobacillus group in breast milk and vagina of healthy women
and potential role in the colonization of the infant gut. Journal of Applied
Microbiology, 103, 2638e2644.

397

Martin, R., Heilig, H. G. H. J., Zoetandal, E. G., Jimenez, E., Fernandez, L., Smidt, H.,
et al. (2007b). Cultivation-independent assessment of the bacterial diversity of
breast milk among healthy women. Research in Microbiology, 158, 31e37.
Martin, R., Jimenez, E., Heilig, H., Fernandez, L., Marin, M. L., Zoetendal, E. G., et al.
(2009). Isolation of bidobacteria from breast milk and assessment of the
bidobacterial population by PCR-denaturing gradient gel electrophoresis and
quantitative real-time PCR. Applied and Environmental Microbiology, 75,
965e969.
Martn, R., Jimnez, E., Olivares, M., Marn, M. L., Fernndez, L., Xaus, J., et al. (2006).
Lactobacillus salivarius CECT 5713, a potential probiotic strain isolated from
infant feces and breast milk of a mother-child pair. International Journal of Food
Microbiology, 112, 35e43.
Martn, R., Olivares, M., Marn, M. L., Fernndez, L., Xaus, J., & Rodrguez, J. M. (2005).
Probiotic potential of 3 lactobacilli strains isolated from breast milk. Journal of
Human Lactation, 21, 8e17.
Martinez-Ferez, A., Rudloff, S., Guadix, E., Henkel, C. A., Pohlentz, G., Boza, J. J., et al.
(2006). Goats milk as a natural source of lactose-derived oligosaccharides:
isolation by membrane technology. International Dairy Journal, 16, 173e181.
Martin-Sosa, S., Martin, M. J., & Hueso, P. (2002). The sialylated fraction of milk
oligosaccharides is partially responsible for binding to enterotoxigenic and uropathogenic Escherichia coli human strains. Journal of Nutrition, 132, 3067e3072.
Martin-Sosa, S., Martin, M. J., Garcia-Pardo, L. A., & Hueso, P. (2003). Sialyloligosaccharides in human and bovine milk and in infant formulas: variations with
the progression of lactation. Journal of Dairy Science, 86, 52e59.
Maruyama, S., & Suzuki, H. (1982). A peptide inhibitor of angiotensin I-converting
enzyme in the tryptic hydrolysate of casein. Agricultural and Biological Chemistry, 46, 1393e1394.
Mascioli, E. A., Lopes, S., Randall, S., Porter, K. A., Kater, G., Hirschberg, Y., et al.
(1989). Serum fatty acid proles after intravenous medium chain triglyceride
administration. Lipids, 24, 793e798.
Maslowski, K. M., Vieira, A. T., Ng, A., Kranich, J., Sierro, F., Yu, D., et al. (2009).
Regulation of inammatory responses by gut microbiota and chemoattractant
receptor GPR43. Nature, 461, 1282e1286.
Mather, I. H. (2000). A review and proposed nomenclature for major proteins of the
milk-fat globule membrane. Journal of Dairy Science, 83, 203e247.
Matsuoka, Y., Serizawa, A., Yoshioka, T., Yamamura, J., Morita, Y., Kawakami, H., et al.
(2002). Cystatin C in milk basic protein (MBP) and its inhibitory effect on bone
resorption in vitro. Bioscience, Biotechnology and Biochemistry, 66, 2531e2536.
McConnell, C. (2004). The effects of omega-3 fatty acids on milk fat synthesis and
composition in dairy cows. M.S. thesis, Ithaca, NY, USA: Cornell University.
McDaniel, M. A., Maier, S. F., & Einstein, G. O. (2003). Brain-specic nutrients:
a memory cure? Nutrition, 19, 955e956.
McDonald, T. L., Larson, M. A., Mack, D. R., & Weber, A. (2001). Elevated extrahepatic
expression and secretion of mammary-associated serum amyloid A 3 (M-SAA3)
into colostrum. Veterinary Immunology and Immunopathology, 83, 203e211.
McEver, R. P. (1997). Selectin-carbohydrate interactions during inammation and
metastasis. Glyconjugate Journal, 14, 585e591.
Mehra, R., & Kelly, P. (2006). Milk oligosaccharides: structural and technological
aspects. International Dairy Journal, 16, 1334e1340.
Meijer, G. W., van Tol, A., van Berkel, T. J., & Weststrate, J. A. (2001). Effect of dietary
elaidic versus vaccenic acid on blood and liver lipids in the hamster. Atherosclerosis, 157, 31e40.
Meisel, H. (1997). Biochemical properties of bioactive peptides derived from milk
proteins: potential nutraceuticals for food and pharmaceutical applications.
Livestock Production Science, 50, 125e138.
Meisel, H. (2005). Biochemical properties of peptides encrypted in bovine milk
proteins. Current Medicinal Chemistry, 12, 1905e1919.
Meisel, H., & Bockelmann, W. (1999). Bioactive peptides encrypted in milk proteins:
proteolytic activation and thropho-functional properties. Antonie Van Leeuwenhoek, 76, 207e215.
Meisel, H., & Fitzgerald, R. J. (2000). Opioid peptides encrypted in intact milk
protein sequences. British Journal of Nutrition, 84, S27eS31.
Mellander, O. (1950). The physiological importance of the casein phosphopeptide
calcium salts. II. Peroral calcium dosage of infants. Acta Societatis Medicorum
Upsaliensis, 55, 247e255.
Mensink, R. P., & Katan, M. B. (1990). Effect of dietary trans fatty acids on highdensity and low-density lipoprotein cholesterol levels in healthy subjects. New
England Journal of Medicine, 323, 439e445.
Merritt, J., Qi, F., & Shi, W. (2006). Milk helps build strong teeth and promotes oral
health. Journal of the California Dental Association, 34, 361e366.
Metz, J. A., Karanja, N., Torok, J., & McCarron, D. A. (1988). Modication of total body
fat in spontaneously hypertensive rats and Wistar-Kyoto rats by dietary calcium
and sodium. American Journal of Hypertension, 1, 58e60.
Michalski, M. C. (2007). On the supposed inuence of milk homogenization on the
risk of CVD, diabetes and allergy. British Journal of Nutrition, 97, 598e610.
Michalski, M., & Januel, C. (2006). Does homogenization affect the human
health properties of cows milk? Trends in Food Science and Technology, 17,
423e437.
Miller, A., Stanton, C., Murphy, J., & Devery, R. (2003). Conjugated linoleic acid
(CLA)-enriched milk fat inhibits growth and modulates CLA-responsive
biomarkers in MCF-7 and SW480 human cancer cell lines. British Journal of
Nutrition, 90, 877e885.
Mills, S., Ross, R. P., Fitzgerald, G., & Stanton, C. (2009). Microbial production of
bioactive metabolites. In A. Y. Tamime (Ed.), Dairy fats and related products
(pp. 257e285). Oxford, UK: Wiley-Blackwell.

398

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

Miniello, V. L., Moro, G. E., & Armenio, L. (2003). Prebiotics in infant formulas: new
perspectives. Acta Pediatrica, 9, S68eS76.
Mitchell, P. L., & McLeod, R. S. (2008). Conjugated linoleic acid and atherosclerosis:
studies in animal models. Biochemistry and Cell Biology, 86, 293e301.
Montreuil, J. (1960). Les glucides du lait. Bulletin de la Societe de Chimie Biologique,
42, 1399e1427.
Montserrat, R.-U., & Santamaria-Orleans, A. (2001). Oligosaccharides: application in
infant food. Early Human Development, 65, S43eS52.
Moore, L. L., Bradlee, M. L., Gao, D., & Singer, M. R. (2008). Effects of average
childhood dairy intake on adolescent bone health. Journal of Pediatrics, 153,
667e673.
Mora-Gutierrez, A., Farrell, H. M., Attaie, R., McWhinney, V. J., & Wang, C. (2007).
Inuence of bovine and caprine casein phosphopeptides differing in alphas1casein content in determining the absorption of calcium from bovine and
caprine calcium-fortied milks in rats. Journal of Dairy Research, 74,
356e366.
de Moreno de LeBlanc, A., Matar, C., LeBlanc, N., & Perdigon, G. (2005). Effects of
milk fermented by Lactobacillus helveticus R389 on a murine breast cancer
model. Breast Cancer Research, 7, R477e486.
Moro, G. E., & Arslanoglu, S. (2005). Reproducing the bidogenic effect of human
milk in formula-fed infants: why and how? Acta Paediatrica, 94, S14eS17.
Moro, G., Arslanoglu, S., Stahl, B., Jelinek, J., Wahn, B., & Goehm, G. (2006). A mixture
of prebiotic oligosaccharides reduces the incidence of atopic dermatitis during
the rst six months of age. Archives of Disease in Childhood, 91, 814e819.
Morrow, A. L., Guillermo, M., Ruiz-Palacois, G. M., Altaye, M., Jiang, X., Guerrero, L.,
et al. (2004). Human milk oligosaccharides are associated with protection
against diarrhea in breast-fed infants. Journal of Pediatrics, 145, 297e303.
Morrow, A. L., Ruiz-Palacois, G. M., Jiang, X., & Newburg, D. S. (2005). Human milk
glycans that inhibit pathogen binding protect breast-feeding infants against
infectious diarrhoea. Journal of Nutrition, 135, 1304e1307.
Mountzouris, K. C., & Gibson, G. R. (2003). Colonisation of the gastrointestinal tract.
Annales Nestle, 61, 43e54.
Murphy, J. J., Coakley, M., & Stanton, C. (2008). Supplementation of dairy cows with
a sh oil containing supplement and sunower oil to increase the CLA content
of milk produced at pasture. Livestock Science, 116, 332e337.
Musemeche, C., Caplan, M., Hsueh, W., Sun, X., & Kelly, A. (1991). Experimental
necrotizing enterocolitis: the role of polymorphonuclear neutrophils. Journal of
Pediatric Surgery, 26, 1047e1050.
Nagao, K., & Yanagita, T. (2005). Conjugated fatty acids in food and their health
benets. Journal of Bioscience and Bioengineering, 100, 152e157.
Nakamura, T., & Urashima, T. (2004). The milk oligosaccharides of domestic farm
animals. Trends in Glycoscience and Glycotechnology, 16, 135e142.
Nakamura, T., Kawase, H., Kimura, K., Watanabe, Y., Ohtani, M., Arai, I., et al.
(2003). Concentrations of sialyloligosaccharides in bovine colostrum and milk
during the prepartum and early lactation. Journal of Dairy Science, 86,
1315e1320.
Narva, M., Karkkainen, M., Poussa, T., Lamberg-Allardt, C., & Korpela, R. (2003).
Caseinphosphopeptides in milk and fermented milk do not affect calcium
metabolism acutely in postmenopausal women. Journal of the American College
of Nutrition, 22, 88e93.
Newberg, D. S., Pickering, L. K., McCleur, R. H., & Cleary, T. G. (1990). Fucosylated
oligosacharides of human milk protect suckling mice from heat-stabile
enterotoxin of Escherichia coli. Journal of Infectious Diseases, 162, 1075e1080.
Newburg, D. S. (2009). Neonatal protection by an innate immune system of human
milk consisting of oligosaccharides and glycans. Journal of Animal Science, 87,
26e34.
Newburg, D. S., Chaturvedi, P., Crane, J. K., Cleary, T. G., & Pickering, L. K. (1995).
Fucosylated oligosaccharide(s) of human milk inhibits stable toxin of Escherichia coli. In V. P. Agrawal, C. B. Sharma, A. Sah, & M. D. Zingde (Eds.), Complex
carbohydrates and advances in biosciences (pp. 199e226). Muzaffarnagar, India:
Society Biosci.
Newburg, D. S., Ruiz-Palacois, G. M., & Morrow, A. L. (2005). Human milk glycans
protect infants against enteric pathogens. Annual Reviews of Nutrition, 25,
37e58.
Newburg, D. S., Ruiz-Palacois, G. M., Altaye, M., Chaturvedi, P., Meinzen-Derr, J., de
Lourdes Guerrero, M., et al. (2003). Innate protection conferred by fucosylated
oligosaccharides of human milk against diarrhea in breastfed infants. Glycobiology, 14, 253e263.
Ninoneuvo, M. R., Park, Y., Yin, H., Zhang, J., Ward, R. E., Clowers, B. H., et al. (2006).
A strategy for annotating the human milk glycome. Journal of Agricultural and
Food Chemistry, 54, 7471e7480.
Ninonuevo, M. R., Ward, R. E., LoCascio, R. G., German, J. B., Freeman, S. L.,
Barboza, M., et al. (2007). Methods for the quantitation of human milk oligosaccharides in bacterial fermentation by mass spectrometry. Analytical
Biochemistry, 361, 15e23.
Noh, S. K., & Koo, S. L. (2004). Milk sphingomyelin is more effective than egg
sphingomyelin in inhibiting intestinal absorption of cholesterol and fat in rats.
Journal of Nutrition, 134, 2611e2616.
Ogawa, A., Nosaka, N., Kasai, M., Aoyama, T., Okazaki, M., Igarashi, O., et al. (2007).
Dietary medium- and long-chain triacylglycerols accelerate diet-induced thermogenesis in humans. Journal of Oleo Science, 56, 283e287.
Ohashi, A., Murata, E., Yamamoto, K., Majima, E., Sano, E., Le, Q. T., et al. (2003). New
functions of lactoferrin and beta-casein in mammalian milk as cysteine
protease inhibitors. Biochemical and Biophysical Research Communications, 306,
98e103.

Oliveras, M., Diaz-Ropero, M. P., Martn, R., Rodriguez, J. M., & Xaus, J. (2006).
Antimicrobial potential of four Lactobacillus strains isolated from breast milk.
Journal of Applied Micobiology, 101, 72e79.
Oriol, R., Mollicone, R., Cailleau, A., Balanzino, L., & Breton, C. (1999). Divergent
evolution of fucosyltransferase genes from vertebrates, invertebrates, and
bacteria. Glycobiology, 9, 323e334.
Orrhage, K., & Nord, C. E. (1999). Factors controlling the bacterial colonization of the
intestine in breast-fed infants. Acta Pediatrica, 430, S43eS54.
Oshida, K., Shimizu, T., Takase, M., Tamura, Y., Shimizu, T., & Yamashiro, Y. (2003a).
Effects of dietary sphingomyelin on central nervous system myelination in
developing rats. Pediatric Research, 53, 589e593.
Oshida, K., Shimuzu, T., Takase, M., Tamura, Y., Shimizu, T., & Yamashiro, Y. (2003b).
Effect of dietary sphingomyelin on central nervous system myelination in
developing rats. Pediatric Research, 53, 580e592.
Oskouian, B., & Saba, J. (2007). Sphingosine-1-phosphate metabolism and intestinal
tumorigenesis: lipid signaling strikes again. Cell Cycle, 6, 522e527.
Otnaess, A. B., Laegreid, A., & Ertresvag, K. (1983). Inhibition of enterotoxin from
Escherichia coli and Vibrio cholerae by gangliosides from human milk. Infection
and Immunity, 40, 563e569.
Ouwehand, A., Isolauri, E., & Salminen, S. (2002). The role of the intestinal microora for the development of the immune system in early childhood. European
Journal of Nutrition, 41, S132eS137.
Papakonstantinou, E., Flatt, W. P., Huth, P. J., & Harris, R. B. (2003). High dietary
calcium reduces body fat content, digestibility of fat, and serum vitamin D in
rats. Obesity Research, 11, 387e394.
Papamandjaris, A. A., MacDougall, D. E., & Jones, P. J. (1998). Medium chain fatty
acid metabolism and energy expenditure: obesity treatment implications. Life
Sciences, 62, 1203e1215.
Pariza, M. W., Park, Y., & Cook, M. E. (2001). The biologically active isomers of
conjugated linoleic acid. Progress in Lipid Research, 40, 283e298.
Parodi, P. W. (1997). Cows milk fat components as potential anticarcinogenic
agents. Journal of Nutrition, 127, 1055e1060.
Parodi, P. W. (2001). Cows milk components with anti-cancer potential. Australian
Journal of Dairy Technology, 56, 65e73.
Parodi, P. W. (2003). Conjugated linoleic acid in food. In J. L. Sebedio, W. W. Christie,
& R. O. Adlof (Eds.), Advances in conjugated linoleic acid research (pp. 101e122).
Champaign, IL, USA: AOCS Press.
Parodi, P. W. (2007). A role for milk proteins and their peptides in cancer prevention. Current Pharmaceutical Design, 13, 813e828.
Parra, P., Bruni, G., Palou, A., & Serra, F. (2008). Dietary calcium attenuation of body
fat gain during high-fat feeding in mice. Journal of Nutritional Biochemistry, 19,
109e117.
Perez, P. F., Dore, J., Leclerc, M., Levenez, F., Benyacoub, J., Serrant, P., et al. (2007).
Bacterial imprinting of the neonatal immune system: lessons from maternal
cells? Pediatrics, 119, e724e732.
Peters, M. J., Dixon, G., Kotowicz, K. T., Hatch, D. J., Heyderman, R. S., & Klein, N. J.
(1999). Circulating platelet-neutrophil complexes represent a subpopulation of
activated neutrophils primed for adhesion, phagocytosis and intracellular
killing. British Journal of Haematology, 106, 391e399.
Petrone, G., Conte, M. P., Longhi, C., di Santo, S., Superti, F., Ammendolia, M. G., et al.
(1998). Natural milk fatty acids affect survival and invasiveness of Listeria
monocytogenes. Letters in Applied Microbiology, 27, 362e368.
Pfeuffer, M., & Schrezenmeir, J. (2002). Milk lipids in diet and health - medium chain
fatty acids (MCFA). Bulletin of the International Dairy Federation, 377, 32e42.
Pfeuffer, M., & Schrezenmeir, J. (2007). Milk and the metabolic syndrome. Obesity
Reviews, 8, 109e118.
Piccardoni, O., Sideri, R., Manarini, S., Piccoli, A., Martelli, N., de Gaetano, G., et al.
(2001). Platelet/polymorphonuclear leukocyte adhesion: a new role for SRC
kinases in Mac-1 adhesive function triggerred by P-selectin. Blood, 98,
108e116.
Pihlanto, A. (2001). Bioactive peptides derived from bovine whey proteins: opioid
and ace-inhibitory peptides. Trends in Food Science and Technology, 11, 347e356.
Pina, A. S., & Roque, A. C. (2008). Studies on the molecular recognition between
bioactive peptides and angiotensin-converting enzyme. Journal of Molecular
Recognition, 22, 162e168.
Plourde, M., Destaillats, F., Chouinard, P. Y., & Angers, P. (2007). Conjugated alphalinolenic acid isomers in bovine milk and muscle. Journal of Dairy Science, 90,
5269e5275.
Politis, I., & Chronopoulou, R. (2008). Milk peptides and immune response in the
neonate. Advances in Experimental and Medical Biology, 606, 253e269.
Poulsen, R. C., Moughan, P. J., & Kruger, M. C. (2007). Long-chain polyunsaturated
fatty acids and the regulation of bone metabolism. Experimental Biology and
Medicine (Maywood), 232, 1275e1288.
Prak, K., & Utsumi, S. (2009). Production of a bioactive peptide (IIAEK) in Escherichia
coli using soybean proglycinin A1ab1b as a carrier. Journal of Agricultural and
Food Chemistry, 57, 3792e3799.
Prak, K., Maruyama, Y., Maruyama, N., & Utsumi, S. (2006). Design of genetically
modied soybean proglycinin A1aB1b with multiple copies of bioactive peptide
sequences. Peptides, 27, 1179e1186.
Priem, B., Gilbert, M., Wakarchuk, W. W., Heyraud, A., & Samain, E. (2002). A new
fermentation process allows large-scale production of human milk oligosaccharides by metabolically engineered bacteria. Glycobiology, 12, 235e240.
Qin, J., Li, R., Raes, J., Arumugam, M., Burgdorf, K. S., Manichanh, C., et al. (2010).
A human gut microbial gene catalogue established by metagenomic
sequencing. Nature, 464, 59e65.

S. Mills et al. / International Dairy Journal 21 (2011) 377e401


Rammer, P., Groth-Pedersen, L., Kirkegaard, T., Daugaard, M., Rytter, A.,
Szyniarowski, P., et al. (2010). BAMLET activates a lysosomal cell death program
in cancer cells. Molecular Cancer Therapeutics, 9, 24e32.
Read, L. C., Lord, A. P., Brantl, V., & Koch, G. (1990). Absorption of beta-casomorphins
from autoperfused lamb and piglet small intestine. American Journal of Physiology, 259, G443eG452.
Reh, W. A., Maga, E. A., Collette, N. M., Moyer, A., Conrad-Brink, J. S., Taylor, S. J., et al.
(2004). Hot topic: using a stearoyl-CoA desaturase transgene to alter milk fatty
acid composition. Journal of Dairy Science, 87, 3510e3514.
Reinhardt, T. A., & Lippolis, J. D. (2008). Developmental changes in the milk fat
globule membrane proteome during the transition from colostrum to milk.
Journal of Dairy Science, 91, 2307e2318.
Renye, J. A., Jr., & Somkuti, G. A. (2008). Cloning of milk-derived bioactive peptides
in Streptococcus thermophilus Biotechnology Letters, 30, 723e730.
Rhee, J. S., Santoso, S., Herrmann, L., Bierhaus, A., Kanse, S. M., May, A. E., et al.
(2003). New aspects of integrin-mediated leukocyte adhesion in inammation:
regulation by hemostatic factors and bacterial products. Current Molecular
Medicine, 3, 387e392.
Ritzenthaler, K. L., McGuire, M. K., Falen, R., Shultz, T. D., Dasgupta, N., &
McGuire, M. A. (2001). Estimation of conjugated linoleic acid intake by written
dietary assessment methodologies underestimates actual intake evaluated by
food duplicate methodology. Journal of Nutrition, 131, 1548e1554.
Rodgers, B. (1978). Feeding in infancy and later ability and attainment: a longitudional study. Developmental Medicine and Child Neurology, 20, 421e426.
Rodriguez-Alcala, L. M., & Fontecha, J. (2007). Hot topic: fatty acid and conjugated
linoleic acid (CLA) isomer composition of commercial CLA-fortied dairy products:
evaluation after processing and storage. Journal of Dairy Science, 90, 2083e2090.
Rokka, T., Syvaoja, E. L., Tuominen, J., & Korhonen, H. (1997). Release of bioactive
peptides by enzymatic proteolysis of Lactobacillus GG fermented UHT milk.
Milchwissenschaft, 52, 675e677.
Rosenberg, A. (1995). Biology of sialic acids. New York, NY, USA: Plenum Press.
Roy, C. C., Kien, C. L., Bouthillier, L., & Levy, E. (2006). Short-chain fatty acids: ready
for prime time. Nutrition in Clinical Practice, 21, 351e366.
Roy, M. K., Watanabe, Y., & Tamai, Y. (1999). Induction of apoptosis in HL-60 cells by
skimmed milk digested with a proteolytic enzyme from the yeast Saccharomyces cerevisiae. Journal of Bioscience and Bioengineering, 88, 426e432.
Rubaltelli, F. F., Biadaioli, R., Pecile, P., & Nicoletti, P. (1998). Intestinal ora in breastand bottle-fed infants. Journal of Perinatal Medicine, 26, 186e191.
Rudloff, S., Pohlentz, G., Diekmann, L., Egge, H., & Kunz, C. (1996). Urinary excretion
of lactose and oligosaccharides in preterm infants fed human milk or infant
formula. Acta Paediatrica, 85, 598e603.
Rudloff, S., Stefan, C., Pohlentz, G., & Kunz, C. (2002). Detection of ligands for
selectins in the oligosaccharide fraction of human milk. European Journal of
Nutrition, 41, 85e92.
Ruiz-Palacois, G. M., Cervantes, L. E., Ramos, P., Chavez-Munguia, B., &
Newburg, D. S. (2003). Campylobacter jejuni binds intestinal H(O) antigen (Fuca
1,2Gal b1,4GlcNAc), and fucosyloligosaccharides of human milk inhibits its
binding and infection. Journal of Biological Chemistry, 278, 14112e14120.
Rutherfurd, K. J., & Gill, H. S. (2000). Peptides affecting coagulation. British Journal of
Nutrition, 84(Suppl. 1), S99eS102.
Ruvoen-Clouet, N., Ganiere, J. P., Andre-Fontaine, G., Blanchard, D., & Le Pendu, J.
(2000). Binding of rabbit hemorrhagic disease virus to antigens of the ABH
histo-blood group family. Journal of Virology, 74, 11950e11954.
Ruxton, C. H., Reed, S. C., Simpson, M. J., & Millington, K. J. (2007). The health
benets of omega-3 polyunsaturated fatty acids: a review of the evidence.
Journal of Human Nutrition and Dietetics, 20, 275e285.
Sabikhi, L. (2007). Designer milk. Advances in Food and Nutrition Research, 53,
161e198.
Saito, T. (2008). Antihypertensive peptides derived from bovine casein and whey
proteins. Advances in Experimental Medicine and Biology, 606, 295e317.
Sakaguchi, M., Koseki, M., Wakamatsu, M., & Matsumura, E. (2006). Effects of
systemic administration of beta-casomorphin-5 on learning and memory in
mice. European Journal of Pharmacology, 530, 81e87.
Sanchez-Juanes, F., Alonso, J. M., Zancada, L., & Hueso, P. (2009). Glycosphingolipids
from bovine milk and milk fat globule membranes: a comparative study.
Adhesion to enterotoxigenic Escherichia coli strains. Biological Chemistry, 390,
31e40.
Sano, E., Miyauchi, R., Takakura, N., Yamauchi, K., Murata, E., Trang Le, Q., et al.
(2005). Cysteine protease inhibitors in various milk preparations and its
importance as a food. Food Research International, 38, 427e433.
Schauer, R., & Kamerling, J. P. (1997). Chemistry, biochemistry and biology of sialic
acids. In H. Schachter, J. Montreuil, & J. F. G. Vliegenthart (Eds.), Glycoproteins II
(pp. 243e372). Amsterdam, The Netherlands: Elsevier.
Scheppach, W., & Weiler, F. (2004). The butyrate story: old wine in new bottles?
Current Opinion in Clinical Nutrition and Metablic Care, 7, 563e567.
Scheppach, W., Bartram, P., Richter, A., Richter, F., Liepold, H., Dusel, G., et al. (1992).
Effect of short chain fatty acids on human colonic mucosa in vitro. Journal of
Parenteral and Enteral Nutrition, 16, 43e48.
Schmelz, E. M. (2004). Sphingolipids in the chemoprevention of colon cancer.
Frontiers in Bioscience, 9, 2632e2639.
Schmelz, E. M., Crall, K. J., Larocque, R., Dillehay, D. L., & Merrill, A. H., Jr. (1994).
Uptake and metabolism of sphingolipids in isolated intestinal loops of mice.
Journal of Nutrition, 124, 702e712.
Schmelz, E. M., Dillehay, D. L., Webb, S. K., Reiter, A., Adams, J., & Merrill, A. H., Jr.
(1996). Sphingomyelin consumption suppresses aberrant colonic crypt foci and

399

increases the proportion of adenomas versus adenocarcinomas in CF1 mice


treated with 1,2-dimethylhydrazine: implications for dietary sphingolipids and
colon carcinogenesis. Cancer Research, 56, 4936e4941.
Schmelz, E. M., Roberts, P. C., Kustin, E. M., Lemonnier, L. A., Sullards, M. C.,
Dillehay, D. L., et al. (2001). Modulation of intracellular beta-catenin localization
and intestinal tumorigenesis in vivo and in vitro by sphingolipids. Cancer
Research, 61, 6723e6729.
Schmelz, E. M., Sullards, M. C., Dillehay, D. L., & Merrill, A. H., Jr. (2000). Colonic cell
proliferation and aberrant crypt foci formation are inhibited by dairy glycosphingolipids in 1, 2-dimethylhydrazine-treated CF1 mice. Journal of Nutrition,
130, 522e527.
Schmidt, R. (1989). Glycoprotenis involved in long-lasting placticity in the teleost
brain. Fortschritte der Zoologie, 37, 327e339.
Scholz-Ahrens, K. E., & Schrezenmeir, J. (2000). Effects of bioactive substances in
milk on mineral and trace element metabolism with special reference to casein
phosphopeptides. British Journal of Nutrition, 84(Suppl. 1), S147e153.
Schon, M. P., Krahn, T., Schon, M., Rodriquez, M.-L., Antonicek, H., Schultz, J. E.,
et al. (2002). Efomycine M, a new specic inhibitor of selectin, impairs
leukocyte adhesion and alleviates cutaneous inammation. Nature Medicine,
8, 366e372.
Schumacher, G., Bendas, G., Stahl, B., & Beermann, C. (2006). Human milk oligosaccharides affect P-selectin binding capacities: in vitro investigation. Nutrition,
22, 620e627.
Sela, D. A., Chapman, J., Adeuya, A., Kim, J. H., Chen, F., Whitehead, T. R., et al. (2008).
The genome sequence of Bidobacterium longum subsp. infantis reveals adaptations for milk utilization within the infant microbiome. Proceedings of the
National Academy of Sciences USA, 105, 18964e18969.
Seppo, L., Jauhiainen, T., Poussa, T., & Korpela, R. (2003). A fermented milk high in
bioactive peptides has a blood pressure-lowering effect in hypertensive
subjects. American Journal of Clinical Nutrition, 77, 326e330.
Sverin, S., & Wenshui, X. (2005). Milk biologically active components as neutraceuticals. Critical Reviews in Food Science and Nutrition, 45, 645e656.
Shahar, D. R., Abel, R., Elhayany, A., Vardi, H., & Fraser, D. (2007). Does dairy calcium
intake enhance weight loss among overweight diabetic patients? Diabetes Care,
30, 485e489.
Sharon, N., & Ofek, I. (2000). Safe as mothers milk: carbohydrates as future antiadhesion drugs for bacterial diseases. Glyconjugate Journal, 17, 659e664.
Shi, H., Dirienzo, D., & Zemel, M. B. (2001). Effects of dietary calcium on adipocyte
lipid metabolism and body weight regulation in energy-restricted aP2-agouti
transgenic mice. Faseb Journal, 15, 291e293.
Siddiqui, R. A., Harvey, K. A., & Zaloga, G. P. (2008). Modulation of enzymatic
activities by n-3 polyunsaturated fatty acids to support cardiovascular health.
Journal of Nutritional Biochemistry, 19, 417e437.
Silveira, M. B., Carraro, R., Monereo, S., & Tebar, J. (2007). Conjugated linoleic acid
(CLA) and obesity. Public Health Nutrition, 10, 1181e1186.
Smacchi, E., & Gobetti, M. (2000). Bioactive peptides in dairy products: synthesis
and interaction with proteolytic enzyme. Food Microbiology, 17, 129e141.
Smith, J. G., Yokoyama, W. H., & German, J. B. (1998). Butyric acid from the diet:
actions at the level of gene expression. Critical Reviews in Food Science and
Nutrition, 38, 259e297.
Smith, M. M., Durkin, M., Hinton, V. J., Bellinger, D., & Kuhn, L. (2003). Inuence of
breastfeeding on cognitive outcomes at age 6e8 years: follow-up of very low
birth weight infants. American Journal of Epidemiology, 158, 1075e1082.
Spiegel, S., & Milstien, S. (2003). Sphingosine-1-phosphate: an enigmatic signalling
lipid. Nature Reviews Molecular Cell Biology, 4, 397e407.
Spitsberg, V. L. (2005). Bovine milk fat globule membrane as a potential neutraceutical. Journal of Dairy Science, 88, 2289e2294.
Spitsberg, V. L., & Gorewit, R. C. (1997a). Anti-cancer proteins found in milk. In CALS
News, Vol. 3. Ithaca, NY, USA: Cornell University.
Spitsberg, V. L., & Gorewit, R. C. (1997b). Breast ovarian cancer susceptibility protein
(BRCA1) in milk, tissue and cells. Journal of Dairy Science, 80, (Suppl 1), 60.
Spitsberg, V. L., & Gorewit, R. C. (2002). Isolation, purication and characterisation of fatty acid binding protein from milk fat globule membrane: effect
of bovine growth hormone treatment. Pakistan Journal of Nutrition, 1,
43e48.
Spitsberg, V. L., Matitashvili, E., & Gorewit, R. C. (1995). Association of fatty acid
binding protein and glycoprotein CD36 in the bovine mammary gland. European Journal of Biochemistry, 230, 872e878.
Sprong, R. C., Hulstein, M. F. E., & Van der Meer, R. (2002). Bovine milk fat
components inhibit food-borne pathogens. International Dairy Journal, 12,
209e215.
Sprong, R. C., Hulstein, M. F., & Van der Meer, R. (2001). Bactericidal activities of
milk lipids. Antimicrobial Agents and Chemotherapy, 45, 1298e1301.
Stahl, B., Thurl, S., Zeng, J., Karas, M., & Hillenkamp, F. (1994). Oligosaccharides from
human milk as revealed by matrix-assisted laser desorption/ionisation mass
spectrometry. Analytical Biochemistry, 223, 218e226.
Stanton, C., Murphy, J., McGrath, E., & Devery, R. (2003). Aninal feeding strategies for
conjugated linoleic acid enrichment of milk. In J. L. Sebedio, W. W. Christie, &
R. O. Adlof (Eds.), Advances in conjugated linoleic acid research (pp. 123e145).
Champaign, IL, USA: AOCS Press.
Stanton, C., Ross, R. P., Fitzgerald, G. F., & Van Sindern, D. (2005). Fermented functional foods based on probiotics and their biogenic metabolites. Current Opinion
in Biotechnology, 16, 198e203.
Stepans, M. B., Wilhelm, S. L., Hertzog, M., Rodehorst, T. K., Blaney, S., Clemens, B.,
et al. (2006). Early consumption of human milk oligosaccharides is inversely

400

S. Mills et al. / International Dairy Journal 21 (2011) 377e401

related to subsequent risk of respiratory and enteric disease in infants.


Breastfeeding Medicine, 1, 207e215.
Sugahara, T., Onda, H., Shinohara, Y., Horii, M., Akiyama, K., Nakamoto, K., et al.
(2005). Immunostimulation effects of proteose-peptone component 3 fragment
on human hybridomas and peripheral blood lymphocytes. Biochimica et Biophysica Acta, 1725, 233e240.
Sun, C. Q., OConnor, C. J., MacGibbon, A. K. H., & Roberton, A. M. (2007). The
products from lipase-catalysed hydrolysis of bovine milk fat kill Helicobacter
pylori in vitro. FEMS Immunology and Medical Microbiology, 49, 235e242.
Sun, C. Q., OConnor, C. J., & Roberton, A. (2002). The antimicrobial properties of
milkfat after partial hydrolysis by calf pregastric lipase. Chemico-Biological
Interactions, 140, 185e198.
Sun, X., & Zemel, M. B. (2004). Calcium and dairy products inhibit weight and fat
regain during ad libitum consumption following energy restriction in Ap2agouti transgenic mice. Journal of Nutrition, 134, 3054e3060.
Svanborg, C., Agerstam, H., Aronson, A., Bjerkvig, R., & Duringer, C. (2003). HAMLET
kills tumor cells by an apoptosis-like mechanism-cellular, molecular, and
therapeutic aspects. Advances in Cancer Research, 88, 1e29.
Tangvoranuntakul, P., Gagneux, P., Diaz, S., Bardor, M., Varki, N., Varki, A., et al. (2003).
Human uptake and incorporation of an immunogenic nonhuman dietary sialic
acid. Proceedings of the National Academy of Sciences USA, 100, 12045e12050.
Tao, N., DePeters, E. J., Freeman, S., German, J. B., Grimm, R., & Lebrilla, C. B. (2008).
Bovine milk glycome. Journal of Dairy Science, 91, 3768e3778.
Teegarden, D. (2005). The inuence of dairy product consumption on body
composition. Journal of Nutrition, 135, 2749e2752.
Teschemacher, H., Koch, G., & Brantl, V. (1997). Milk protein-derived opioid receptor
ligands. Biopolymers, 43, 99e117.
Teucher, B., Majsak-Newman, G., Dainty, J. R., McDonagh, D., FitzGerald, R. J., &
Fairweather-Tait, S. J. (2006). Calcium absorption is not increased by caseinophosphopeptides. American Journal of Clinical Nutrition, 84, 162e166.
Thormar, H., Isaacs, C. E., Brown, H. R., Barshatzky, M. R., & Pessolano, T. (1987).
Inactivation of enveloped viruses and killing of cells by fatty acids and monoglycerides. Antimicrobial Agents and Chemotherapy, 31, 27e31.
Tissier, H. (1905). Repartition des microbes dans lintestin du nourisson. Annales de
linstitut Pasteur, 19, 109.
Toba, Y., Takada, Y., Matsuoka, Y., Morita, Y., Motouri, M., Hirai, T., et al. (2001). Milk
basic protein promotes bone formation and suppresses bone resorption in
healthy adult men. Bioscience, Biotechnology and Biochemistry, 65, 1353e1357.
Treem, W. R., Ahsan, N., Shoup, M., & Hyams, J. S. (1994). Fecal short-chain fatty
acids in children with inammatory bowel disease. Journal of Pediatric Gastroenterology and Nutrition, 18, 159e164.
Tricon, S., Burdge, G. C., Jones, E. L., Russell, J. J., El-Khazen, S., Moretti, E., et al.
(2006). Effects of dairy products naturally enriched with cis-9,trans-11 conjugated linoleic acid on the blood lipid prole in healthy middle-aged men.
American Journal of Clinical Nutrition, 83, 744e753.
Tsai, W. Y., Chang, W. H., Chen, C. H., & Lu, F. J. (2000). Enhancing effect of patented
whey protein isolate (Immunocal) on cytotoxicity of an anticancer drug.
Nutrition and Cancer, 38, 200e208.
Tsuchita, H., Suzuki, T., & Kuwata, T. (2001). The effect of casein phosphopeptides on
calcium absorption from calcium-fortied milk in growing rats. British Journal
of Nutrition, 85, 5e10.
Tsuji, H., Kasai, M., Takeuchi, H., Nakamura, M., Okazaki, M., & Kondo, K. (2001).
Dietary medium-chain triacylglycerols suppress accumulation of body fat in
a double-blind, controlled trial in healthy men and women. Journal of Nutrition,
131, 2853e2859.
Tsukumo, D. M., Carvalho, B. M., Carvalho-Filho, M. A., & Saad, M. J. (2009).
Translational research into gut microbiota: new horizons in obesity treatment.
Arquivos Brasileiros de Endocrinologia e Metabologia, 53, 139e144.
Tsuzuki, T., Tokuyama, Y., Igarashi, M., & Miyazawa, T. (2004). Tumor growth
suppression by a-elostearic acid, a linolenic acid isomer with a conjugated
triene system, via lipid peroxidation. Carcinogenesis, 25, 1417e1425.
Turpeinen, A. M., Mutanen, M., Aro, A., Salminen, I., Basu, S., Palmquist, D. L., et al.
(2002). Bioconversion of vaccenic acid to conjugated linoleic acid in humans.
American Journal of Clinical Nutrition, 76, 504e510.
Uenishi, K., Ishida, H., Toba, Y., Aoe, S., Itabashi, A., & Takada, Y. (2007). Milk basic
protein increases bone mineral density and improves bone metabolism in
healthy young women. Osteoporosis International, 18, 385e390.
Varki, A. (1994). Selectin ligands. Proceedings of the National Academy of Sciences
USA, 91, 7390e7397.
Varki, A. (2010). Colloquium paper: uniquely human evolution of sialic acid genetics
and biology. Proceedings of the National Academy of Sciences USA, 107(Suppl. 2),
8939e8946.
Venti, C. A., Tataranni, P. A., & Salbe, A. D. (2005). Lack of relationship between
calcium intake and body size in an obesity-prone population. Journal of the
American Dietetic Association, 105, 1401e1407.
Vesper, H., Schmelz, E. M., Nikolova-Karakashian, M. N., Dillehay, D. L., Lynch, D. V.,
& Merill, A. H. J. (1999). Sphingolipids in food and the emerging importance of
sphingolipids to nutrition. Journal of Nutrition, 129, 1239e1250.
Wada, J., Ando, T., Kiyohara, M., Ashida, H., Kitaoka, M., Yamaguchi, M., et al. (2008).
Bidobacterium bidum lacto-N-biosidase, a critical enzyme for the degradation
of human milk oligosaccharides with a type 1 structure. Applied and Environmental Microbiology, 74, 3996e4004.
Wall, R., Ross, R. P., Fitzgerald, G. F., & Stanton, C. (2010). Fatty acids from sh: the
anti-inammatory potential of long-chain omega-3 fatty acids. Nutrition
Reviews, 68, 280e289.

Walstra, P. (2003). Physical chemistry of foods. New York, NY, USA: Marcel Dekker.
Walstra, P., & Jenness, R. (1984). Dairy chemistry and physics. New York, NY, USA:
John Wiley & Sons Inc.
Wang, B., & Brand-Miller, J. (2003). The role and potential of sialic acid in human
nutrition. European Journal of Clinical Nutrition, 57, 1351e1369.
Wang, B., Brand-Miller, J., McVeagh, P., & Petocz, P. (2001a). Concentration and
distribution of sialic acid in human milk and infant formulas. American Journal
of Clinical Nutrition, 74, 510e515.
Wang, B., Downing, J. A., Petocz, P., Brand-Miller, J., & Bryden, W. L. (2007a).
Metabolic fate of intravenously administered N-acetylneuraminic acid-6-14C in
newborn piglets. Asia Pacic Journal of Clinical Nutrition, 16, 110e115.
Wang, B., McVeagh, P., Petocz, P., & Brand-Miller, J. (2003). Brain ganglioside and
glycoprotein sialic acid in breastfed compared with formula-fed infants.
American Journal of Clinical Nutrition, 78, 1024e1029.
Wang, B., Miller, J. B., Sun, Y., Ahmad, Z., McVeagh, P., & Petocz, P. (2001b).
A longitudinal study of salivary sialic acid in preterm infants: comparison
of human milk-fed versus formula-fed infants. Journal of Pediatrics, 138, 914e916.
Wang, B., Yu, B., Karim, M., Hu, H., Sun, Y., McGreevy, P., et al. (2007b). Dietary sialic
acid supplementation improves learning and memory in piglets. American
Journal of Clinical Nutrition, 85, 561e569.
Wang, X., Hirmo, S., Millen, R., & Wadstrom, T. (2001c). Inhibition of Helicobcater
pylori infection by bovine milk glycoconjugates in a BALB/ca mouse model.
FEMS Immunology and Medical Microbiology, 20, 275e281.
Ward, R. E., & German, J. B. (2004). Understanding milks bioactive components:
a goal for the genomics toolbox. Journal of Nutrition, 134, S962eS967.
Ward, R. E., Ninonuevo, M., Mills, D. A., Lebrilla, C. B., & German, J. B. (2006).
In vitro fermentation of breast milk oligosaccharides by Bidobacterium
infantis and Lactobacillus gasseri. Applied and Environmental Microbiology, 72,
4497e4499.
Ward, R. E., Ninonuevo, M., Mills, D. A., Lebrilla, C. B., & German, J. B. (2007). In vitro
fermentability of human milk oligosaccharides by several strains of bidobacteria. Molecular Nutrition and Food Research, 51, 1398e1405.
Watras, A. C., Buchholz, A. C., Close, R. N., Zhang, Z., & Schoeller, D. A. (2007). The
role of conjugated linoleic acid in reducing body fat and preventing holiday
weight gain. International Journal of Obesity (London), 31, 481e487.
Wijendran, V., & Hayes, K. C. (2004). Dietary n-6 and n-3 fatty acid balance and
cardiovascular health. Annual Review of Nutrition, 24, 597e615.
Willett, W. C. (2007). The role of dietary n-6 fatty acids in the prevention of
cardiovascular disease. Journal of Cardiovascular Medicine (Hagerstown), 8,
S42eS45.
Wong, W. M. J., de Souza, R., Kendall, C. W. C., Emam, A., & Jenkins, D. J. A. (2006).
Colonic health: fermentation and short chain fatty acids. Journal of Clinical
Gastroenterology, 40, 235e243.
Wood, B., & Collard, M. (1999). The human genus. Science, 284, 65e71.
Wurtman, R. J. (2008). Synapse formation and cognitive brain development: effect of
docosahexaenoic acid and other dietary constituents. Metabolism, 57(Suppl. 2),
S6eS10.
Yamada, Y., Nishizawa, K., Yokoo, M., Zhao, H., Onishi, K., Teraishi, M., et al. (2008).
Anti-hypertensive activity of genetically modied soybean seeds accumulating
novokinin. Peptides, 29, 331e337.
Yamamoto, N., Akino, A., & Takano, T. (1994). Antihypertensive effects of different
kinds of fermented milk in spontaneously hypertensive rats. Bioscience,
Biotechnology and Biochemistry, 58, 776e778.
Yamauchi, K. (1992). Biologically functional proteins of milk and peptides derived
from milk proteins. Bulletin of the International Dairy Federation, 272, 51e58.
Yanagi, S., Yamashita, M., & Imai, S. (1993). Sodium butyrate inhibits the enhancing
effect of high fat diet on mammary tumorigenesis. Oncology, 50, 201e204.
Yasui, Y., Hosokawa, M., Kohno, H., Tanaka, T., & Miyashita, K. (2006). Troglitazone
and 9cis, 11trans, 13trans-conjugated linolenic acid: comparison of their antiproliferative and apoptosis-inducing effects on different colon cancer cell lines.
Chemotherapy, 52, 220e225.
Yasui, Y., Hosokawa, M., Sahara, T., Suzuki, R., Ohgiya, S., Kohno, H., et al. (2005).
Bitter gourd seed fatty acid rich in 9c, 11t, 13t-conjugated linolenic acid
induces apoptosis and up-regulates the GADD45, p53 and PPARg in human
colon cancer Caco-2 cells. Prostaglandins, Leukotrienes and Essential Fatty
Acids, 73, 113e119.
Yin, J., Hashimoto, A., Izawa, M., Miyazaki, K., Chen, G. Y., Takematsu, H., et al.
(2006). Hypoxic culture induces expression of sialin, a sialic acid transporter,
and cancer-associated gangliosides containing non-human sialic acid on human
cancer cells. Cancer Research, 66, 2937e2945.
Yolken, R. H., Peterson, J. A., Vonderfecht, S. L., Fouts, E. T., Midthun, K., &
Newberg, D. S. (1992). Human milk mucin inhibits rotavirus replication and
prevents experimental gastroenteritis. Journal of Clinical Investigation, 90,
1984e1991.
Yoshikawa, M., Yoshimura, T., & Chiba, H. (1984). Opioid peptides from human bcasein. Agricultural and Biological Chemistry, 48, 3185e3187.
Zemel, M. B. (2003). Role of dietary calcium and dairy products in modulating
adiposity. Lipids, 38, 139e146.
Zemel, M. B. (2004). Role of calcium and dairy products in energy partitioning and
weight management. American Journal of Clinical Nutrition, 79, 907Se912S.
Zemel, M. B. (2005). The role of dairy foods in weight management. Journal of the
American College of Nutrition, 24, 537Se546S.
Zemel, M. B., & Geng, X. (2001). Dietary calcium and yoghurt accelarate body fat loss
secondary to caloric restriction in aP2-agouti transgenic mice. Abstract. Obesity
Research, 9, 146S.

S. Mills et al. / International Dairy Journal 21 (2011) 377e401


Zemel, M. B., & Morgan, K. (2002). Interaction between calcium, dairy and dietary
macronutrients in modulating body composition in obese mice. Abstract. Faseb
Journal, 16, A369.
Zemel, M. B., Richards, J., Mathis, S., Milstead, A., Gebhardt, L., & Silva, E. (2005a).
Dairy augmentation of total and central fat loss in obese subjects. International
Journal of Obesity (London), 29, 391e397.
Zemel, M. B., Richards, J., Milstead, A., & Campbell, P. (2005b). Effects of calcium and
dairy on body composition and weight loss in African-American adults. Obesity
Research, 13, 1218e1225.
Zemel, M. B., Shi, H., Greer, B., Dirienzo, D., & Zemel, P. C. (2000). Regulation of
adiposity by dietary calcium. Faseb Journal, 14, 1132e1138.
Zemel, M. B., Sun, X., & Geng, X. (2001). Effects of a calcium-fortied breakfast
cereal on adiposity in a transgenic mouse model of obesity. Abstract. Faseb
Journal, 15, A598, 480.597.

401

Zemel, M. B., Thompson, W., Milstead, A., Morris, K., & Campbell, P. (2004). Calcium
and dairy acceleration of weight and fat loss during energy restriction in obese
adults. Obesity Research, 12, 582e590.
Zhao, G., Etherton, T. D., Martin, K. R., Gillies, P. J., West, S. G., & Kris-Etherton, P. M.
(2007). Dietary a-linolenic acid inhibits proinammatory cytokine production
by peripheral blood mononuclear cells in hypercholesterolemic subjects.
Amercian Journal of Clinical Nutrition, 85, 385e391.
Zimecki, M., & Kruzel, M. L. (2007). Milk-derived proteins and peptides of potential
therapeutic and nutritive value. Journal of Experimental Therapeutics and
Oncology, 6, 89e106.
Zoghbi, S., Trompette, A., Claustre, J., El Homsi, M., Garzon, J., Jourdan, G., et al.
(2006). b-Casomorphin-7 regulates the secretion and expression of gastrointestinal mucins through a mu-opioid pathway. American Journal of Physiology Gastrointestinal and Liver Physiology, 290, G1105eG1113.

You might also like