Вы находитесь на странице: 1из 13

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/8432924

Yogurt and gut function


ARTICLE in AMERICAN JOURNAL OF CLINICAL NUTRITION SEPTEMBER 2004
Impact Factor: 6.77 Source: PubMed

CITATIONS

READS

159

438

3 AUTHORS, INCLUDING:
Oskar Adolfsson

Simin Nikbin Meydani

AC Immune SA

Tufts University

22 PUBLICATIONS 680 CITATIONS

230 PUBLICATIONS 8,951 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Oskar Adolfsson


Retrieved on: 27 March 2016

Review Article

Yogurt and gut function1,2


Oskar Adolfsson, Simin Nikbin Meydani, and Robert M Russell

KEY WORDS
Yogurt, gut function, gut immunity, gastrointestinal diseases, gut microflora

INTRODUCTION

Components of the human intestinal microflora and of the


food entering the intestine may have harmful or beneficial effects
on human health. Abundant evidence implies that specific bacterial species used for the fermentation of dairy products such as
yogurt and selected from the healthy gut microflora have powerful antipathogenic and antiinflammatory properties. These microorganisms are therefore involved with enhanced resistance to
colonization of pathogenic bacteria in the intestine, which has led
to the introduction of novel modes of therapeutic and prophylactic interventions based on the consumption of monocultures
and mixed cultures of beneficial live microorganisms as probiotics. Probiotics are defined as living microorganisms, which
on ingestion in sufficient numbers, exert health benefits beyond
inherent basic nutrition (1).
Yogurt is one of the best-known of the foods that contain
probiotics. Yogurt is defined by the Codex Alimentarius of 1992

as a coagulated milk product that results from the fermentation of


lactic acid in milk by Lactobacillus bulgaricus and Streptococcus thermophilus (2). Other lactic acid bacteria (LAB) species
are now frequently used to give the final product unique characteristics. As starter cultures for yogurt production, LAB species
display symbiotic relations during their growth in milk medium
(3). Thus, a carefully selected mixture of LAB species is used to
complement each other and to achieve a remarkable efficiency in
acid production. Furthermore, to increase the number of LAB
that survive the low pH and high acidity of the gastrointestinal
environment, some LAB species that are indigenous to the human intestine have been used in yogurt production. To meet the
National Yogurt Associations criteria for live and active culture yogurt, the finished yogurt product must contain live LAB
in amounts 108 organisms/g at the time of manufacture (3), and
the cultures must remain active at the end of the stated shelf life,
as ascertained with the use of a specific activity test.
In many modern societies, fermented dairy products make up
a substantial proportion of the total daily food consumption.
Furthermore, it has long been believed that consuming yogurt
and other fermented milk products provides various health benefits (4). Studies from the 1990s on the possible health properties
of yogurt added to this belief (1, 5).
Probiotic therapy is based on the notion that there is such a
thing as a normal healthy microflora, but normal healthy microflora has not been defined except perhaps as microflora without a pathogenic bacterial overgrowth. The development of
novel means of characterizing and modifying the gut microflora
has opened up new perspectives on the role of the gut microflora
in health and disease. Numerous studies suggested beneficial
therapeutic effects of LAB on gut health. However, results have
been inconsistent, which may be due to differences in the strains
of LAB, routes of administration, and investigational procedures
used in these studies.
Several LAB species are currently used in the production of
yogurt. This review focuses on the current evidence suggesting
that yogurt and specific LAB species that are used for the fermentation of milk may or may not have valuable healthpromoting properties or therapeutic effects on various gastrointestinal functions and diseases.
1
From the Jean Mayer USDA Human Nutrition Research Center on Aging
at Tufts University, Boston.
2
Address reprint requests to SN Meydani, Nutritional Immunology Laboratory, JM USDA-HNRCA at Tufts University, 711 Washington Street,
Boston, MA 02111. E-mail: simin.meydani@tufts.edu.
Received October 3, 2003.
Accepted for publication February 12, 2004.

Am J Clin Nutr 2004;80:24556. Printed in USA. 2004 American Society for Clinical Nutrition

245

Downloaded from www.ajcn.org by guest on June 6, 2011

ABSTRACT
In recent years, numerous studies have been published on the health
effects of yogurt and the bacterial cultures used in the production of
yogurt. In the United States, these lactic acidproducing bacteria
(LAB) include Lactobacillus and Streptococcus species. The benefits of yogurt and LAB on gastrointestinal health have been investigated in animal models and, occasionally, in human subjects. Some
studies using yogurt, individual LAB species, or both showed promising health benefits for certain gastrointestinal conditions, including lactose intolerance, constipation, diarrheal diseases, colon cancer, inflammatory bowel disease, Helicobacter pylori infection, and
allergies. Patients with any of these conditions could possibly benefit
from the consumption of yogurt. The benefits of yogurt consumption
to gastrointestinal function are most likely due to effects mediated
through the gut microflora, bowel transit, and enhancement of gastrointestinal innate and adaptive immune responses. Although substantial evidence currently exists to support a beneficial effect of
yogurt consumption on gastrointestinal health, there is inconsistency
in reported results, which may be due to differences in the strains of
LAB used, in routes of administration, or in investigational procedures or to the lack of objective definition of gut health. Further
well-designed, controlled human studies of adequate duration are
needed to confirm or extend these findings.
Am J Clin Nutr
2004;80:24556.

246

ADOLFSSON ET AL

NUTRITIONAL VALUE OF YOGURT

The nutrient composition of yogurt is based on the nutrient


composition of the milk from which it is derived, which is affected by many factors, such as genetic and individual mammalian differences, feed, stage of lactation, age, and environmental
factors such as the season of the year. Other variables that play a
role during processing of milk, including temperature, duration
of heat exposure, exposure to light, and storage conditions, also
affect the nutritional value of the final product. In addition, the
changes in milk constituents that occur during lactic acid fermentation influence the nutritional and physiologic value of the
finished yogurt product. The final nutritional composition of
yogurt is also affected by the species and strains of bacteria used
in the fermentation, the source and type of milk solids that may
be added before fermentation, and the temperature and duration
of the fermentation process.

change that occurs during the fermentation process is the hydrolysis


of 20 30% of the disaccharide lactose to its absorbable monosaccharide components, glucose and galactose (2). In addition, a portion of the glucose is converted to lactic acid. Depending on other
ingredients added, this hydrolysis results in lower lactose concentrations in yogurt than in milk, which in part explains why yogurt is
tolerated better than milk by persons with lactose maldigestion (13
15). However, other factors also seem to play a role. For example,
lactose-intolerant subjects exhibited better tolerance for yogurt with
a relatively high amount of lactose than for milk containing a similar
amount of lactose (13, 15). In another example, bacteria present in
yogurt, such as L. bulgaricus and S. thermophilus, expressed functional lactase, the enzyme that breaks down lactose (16). This expression may also contribute to better tolerance of lactose in yogurt
than of lactose in milk by persons with lactose maldigestion (15).
Protein

B vitamins

The protein content of commercial yogurt is generally higher


than that of milk because of the addition of nonfat dry milk during
processing and concentration, which increases the protein content of the final product. It has been argued that protein from
yogurt is more easily digested than is protein from milk, as
bacterial predigestion of milk proteins in yogurt may occur (8,
17). This argument is supported by evidence of a higher content
of free amino acids, especially proline and glycine, in yogurt than
in milk. The activity of proteolytic enzymes and peptidases is
preserved throughout the shelf life of the yogurt. Thus, the concentration of free amino groups increases up to twofold during
the first 24 h and then doubles again during the next 21 d of
storage at 7 C (18). Some bacterial cultures have been shown to
have more proteolytic activity than do others. For example, L.
bulgaricus was shown to have a much higher proteolytic activity
during milk fermentation and storage than does S. thermophilus,
as indicated by elevated concentrations of peptides and free
amino acids after milk fermentation (19).
During fermentation, both heat treatment and acid production
result in finer coagulation of casein, which may also contribute to
the greater protein digestibility of yogurt than of milk. Proteins in
yogurt are of excellent biological quality, as are those in milk,
because the nutritional value of milk proteins is well preserved
during the fermentation process (20). Both the caseins and the
whey proteins in yogurt are rich sources of all the essential amino
acids, and the intestinal availability of nitrogen has been reported
as being high (93%; 21, 22). Labeling of milk proteins with the
stable isotope 15N has made it possible to discriminate between
exogenous and endogenous nitrogen fractions in serum after
ingestion of 15N-labeled milk or 15N-labeled yogurt proteins. In
a study of human subjects, Gaudichon et al (23) found that proteins from both milk and yogurt were rapidly hydrolyzed after
ingestion, but the gastroduodenal transfer of dietary nitrogen was
slower when yogurt was fed than when milk was fed.

Lactose

Lipids

Dairy products and foods prepared with the use of dairy ingredients are an exclusive source of the disaccharide lactose in human
diets. Before absorption, lactose is hydrolyzed by the intestinal
brush border -galactosidase (lactase) into glucose and galactose.
These monosaccharides are absorbed and used as energy sources.
Before fermentation, the lactose content of the yogurt mix generally is 6% (3). One example of a significant bacteria-induced

Milk fat also goes through biochemical changes during the


fermentation process. Minor amounts of free fatty acids are released as a result of lipase activity (3). Because most of the yogurt
sold in the United States is of the low-fat and nonfat varieties,
hydrolysis of lipids contributes little to the attributes of most
yogurt products. However, yogurt has been shown to have a higher
concentration of conjugated linoleic acid (CLA), a long-chain

Downloaded from www.ajcn.org by guest on June 6, 2011

Dairy products have generally been considered an excellent


source of high-quality protein, calcium, potassium, phosphorus,
magnesium, zinc, and the B vitamins riboflavin, niacin, vitamin
B-6, and vitamin B-12 (6). A much greater loss of vitamins than
of minerals may occur during the processing of yogurt because
vitamins are more sensitive to changes in environmental factors
than are minerals. Some of the factors that are important during
the processing of milk and that are known to have adverse effects
on the vitamin content of dairy products in general include heat
treatment and pasteurization, ultrafiltration, agitation, and oxidative conditions. In addition, bacterial cultures used during the
fermentation process of yogurt can influence the vitamin content
of the final product (6).
LAB species do require B vitamins for growth, but some
cultures are capable of synthesizing B vitamins (6). An example
of a B vitamin that is utilized by LAB is vitamin B-12 (7, 8).
Vitamins required for the growth of LAB cultures vary from one
strain to another. Significant losses of vitamin B-12 can be corrected by the careful use of supplementary LAB cultures that are
capable of synthesizing vitamin B-12 (9).
Folate is the best example of a B vitamin that some LAB
species synthesize (10, 11). Depending on the bacterial strains
used, the folate content of yogurt can vary widely, ranging from
4 to 19 g/100 g (8). The major form of folate present in milk is
5-methyl-tetrahydrofolate (12). In a recent study, bacterial isolates from various species used for milk fermentation and yogurt
production were examined for their ability to synthesize or utilize
folate (11). S. thermophilus and Bifidobacteria were folate producers, whereas Lactobacilli depleted folate from the milk media. A combination of folate-producing cultures resulted in even
greater folate content of the final fermented product. Further
studies on the effect of changes in the vitamin B content of milk
on fermentation would be of great practical significance.

YOGURT AND GUT FUNCTION

biohydrogenated derivative of linoleic acid, than does the milk from


which the yogurt was processed (24). A fermented dairy product
from India, referred to as dahi, has also been shown to have higher
CLA content than does nonfermented dahi (25). The major sources
of CLA in our diets are animal products from ruminants, in which
CLA is synthesized by rumen bacteria. Increased consumption of
dairy fat was shown to be associated with increased concentrations
of CLA in both human adipose tissue (26) and human milk (27). It
was hypothesized that biohydrogenation also occurs during fermentation of milk and results in higher concentrations of CLA in the final
product (28).
CLA was reported to have immunostimulatory and anticarcinogenic properties (29). In a recent study of breast and colon
cancer cells, Kemp et al (30) showed that the anticarcinogenic
properties of CLA may be due to the ability of some CLA isomers
to inhibit the expression of cyclins and thus halt the progression
of the cell cycle from G1 to S phase. In addition, CLA induced the
expression of the tumor suppressor p53.
Minerals

These studies may suggest that the bioavailability of calcium in


yogurt is greater and yogurt may increase bone mineralization
more than do nonfermented milk products. However, there are
currently no published studies that show a superior effect of
yogurt on bone mineralization in human subjects.

MECHANISTIC RATIONALE FOR POTENTIAL


BENEFITS OF YOGURT ON GUT FUNCTION AND
HEALTH

It has been suggested that yogurt and LAB contribute to several facets of gastrointestinal health: the makeup of the gastrointestinal flora, the immune response, and laxation.
Gut microflora
Lactobacilli are among the components of microbial flora in
both the small and large intestines. The ability of nonpathogenic
intestinal microflora, such as LAB, to associate with and bind to
the intestinal brush border tissue is thought to be an important
attribute that prevents harmful pathogens from accessing the
gastrointestinal mucosa (39). For LAB to have an effect, they
must adapt to the host intestinal environment and be capable of
prolonged survival in the intestinal tract (40 43). LAB survival
is influenced by gastric pH as well as by exposure to digestive
enzymes and bile salts (42), and LAB species differ in their
ability to survive in the gastrointestinal environment (43).
When 4 strains of Bifidobacterium (B. infantis, B. bifidum, B.
adolescentis, and B. longum) were compared, B. longum was the
most resistant to the effects of gastric acid (44). Bifidobacterium
animalis was reported to have a high survival rate during intestinal transit in human subjects (45).
The effect of feeding yogurt fermented with S. thermophilus,
L. bulgaricus, and Lactobacillus casei on the fecal microflora of
healthy infants aged 10 18 mo was investigated by GuerinDanan et al (46). Whereas the number of infants with fecal
Lactobacillus increased after the feeding, the total numbers of
anaerobes, Bifidobacteria, bacteroides, and enterobacteria were
not affected by yogurt intake. In a group of elderly patients with
atrophic gastritis and hypochlorhydria, Lactobacillus gasseri
survived passage through the gastrointestinal tract, but S. thermophilus and L. bulgaricus were not recovered (43). Bifidobacterium sp has also been shown to survive passage through the
gastrointestinal tract: fecal concentrations were detectable for
8 d after the cessation of intake (47).
Another important factor that limits the survival of lactobacilli
within the upper gastrointestinal tract is the inherent ability of the
organisms to adhere to intestinal epithelial cells (42). With the
use of scanning electron microscopy, Plant and Conway (48)
screened 16 strains of Lactobacillus for their capacity to associate with Peyers patches and the lymphoid villous intestinal tissues in mice. Two of the 16 strains investigated, Lactobacillus
acidophilus and L. bulgaricus, are of interest because they relate
to yogurt. It was found, in both in vitro and in vivo models using
BALB/c mice, that L. bulgaricus did not associate with Peyers
patches or with the lymphoid villous intestinal tissues. L. acidophilus had a low degree of association with Peyers patches
and no association to the lymphoid villous intestinal tissue. Nevertheless, the authors stated that the strains of Lactobacillus
tested showed high rates of survival when Lactobacillus was
administered orally.

Downloaded from www.ajcn.org by guest on June 6, 2011

In addition to being a good source of protein, yogurt is an


excellent source of calcium and phosphorus. In fact, dairy products such as milk, yogurt, and cheese provide most of the highly
bioavailable calcium in the typical Western diet. Because of the
lower pH of yogurt compared with that of milk, calcium and
magnesium are present in yogurt mostly in their ionic forms.
One of the major functions of calcium is the role it plays in bone
formation and mineralization. The calcium requirements during
growth, pregnancy, and lactation are increased. However, the average calcium intake of women of childbearing age is consistently
less than is recommended (31). In addition, calcium intake of
women tends to fall even lower during the postmenopausal years
(32). This is especially important for postmenopausal women,
who are at increased risk of bone loss and osteoporosis. Dietary
fiber has an adverse effect on calcium absorption, whereas lactose may enhance the absorption of calcium (33). In the rat
model, calcium retention was greater with consumption of a diet
in which lactose made up half the total carbohydrates ingested
than with consumption of the control diet (34). Schaafsma et al
(35), investigating the effect of dairy products on mineral absorption by using rat models, reported that lactose enhances the
absorption of calcium, magnesium, and zinc. Because yogurt has
a lactose content lower than that of milk, the bioavailability of
these minerals may be negatively affected, although the effect is
likely to be small.
The acidic pH of yogurt ionizes calcium and thus facilitates
intestinal calcium uptake (36). The low pH of yogurt also may
reduce the inhibitory effect of dietary phytic acid on calcium
bioavailability. Vitamin D plays a major regulatory role in intestinal calcium absorption. The active, saturable, transcellular
route of calcium absorption in the duodenum and proximal jejunum requires calbindin-D, a vitamin D dependent calciumbinding protein (37). In the United States, milk and infant formula are fortified with vitamin D, and hence they serve as good
dietary sources, with 2.5 g (100 IU) vitamin D/237-mL serving.
However, other dairy products, such as yogurt, typically are not
fortified with vitamin D.
Few studies have investigated the effect of yogurt-derived
calcium on bone mineralization in animals (34, 38). Kaup et al
(34) reported that yogurt-fed rats showed greater bone mineralization than did rats fed a diet containing calcium carbonate.

247

248

ADOLFSSON ET AL

The ability of LAB to decrease the gastrointestinal invasion of


pathogenic bacteria has also been described (39, 49). Bernet et al
(39) reported a dose-dependent L. acidophilusmediated inhibition of the adherence of enteropathogenic Escherichia coli and
Salmonella typhimurium to the enterocyte cell-line Caco-2. In
addition, L. acidophilus inhibited the entry of E. coli, S. typhimurium, and Yersinia pseudotuberculosis into Caco-2 cells. In
another report (49), the same authors described similar inhibitory
effects when 2 different strains of Bifidobacteria (B. breve and B.
infantis) were used. In addition, long-term feeding of yogurt does
not result in a significant change in the results of breath-hydrogen
tests, which indicates the absence of a significant change in the
intestinal survival of the yogurt organisms (50). Furthermore, it
is possible that the ability of LAB to compete with pathogens for
adhesion to the intestinal wall is influenced by their membrane
fluidity. This possibility was suggested by studies indicating that
the type and quantities of polyunsaturated fatty acids in the extracellular milieu influence the adhesive properties of LAB to the
epithelium (51, 52).
Gut-associated immune response

Downloaded from www.ajcn.org by guest on June 6, 2011

The mucosal lymphoid tissue of the gastrointestinal tract plays


an important role as a first line of defense against ingested pathogens. The interactions of LAB with the mucosal epithelial lining
of the gastrointestinal tract, as well as with the lymphoid cells
residing in the gut, have been suggested as the most important
mechanism by which LAB enhances gut immune function. Several factors have been identified as contributing to the immunomodulating and antimicrobial activities of LAB, including the
production of low pH, organic acids, carbon dioxide, hydrogen
peroxide, bacteriocins, ethanol, and diacetyl; the depletion of
nutrients; and competition for available living space (1, 5, 53).
The gastrointestinal tract is a complex immune system tissue.
The main site of the mucosal immune system in the gut is referred
to as gut-associated lymphoid tissue (GALT), which can be divided into inductive and effector sites. In the small intestine, the
inductive sites are in the Peyers patches, which consist of large
lymphoid follicles in the terminal small intestine. The bestdefined effector component of the mucosal adaptive immune
system is secretory immunoglobulin A (sIgA). sIgA is the main
immunoglobulin of the humoral immune response, which together with the innate mucosal defenses provides protection
against microbial antigens at the intestinal mucosal surface (54).
In a healthy person, sIgA inhibits the colonization of pathogenic
bacteria in the gut, as well as the mucosal penetration of pathogenic antigens. At least 80% of all the bodys plasma cells, the
source of sIgA, are located in the intestinal lamina propria
throughout the length of the small intestine. IgA is the most
abundantly produced immunoglobulin in the human body. The
production of intestinal sIgA requires the presence of commensal
microflora (55), which indicates that the production of intestinal
sIgA is induced in response to antigenic stimulation. It is not yet
clear, however, how lamina propria B cells are activated to become IgA-secreting plasma cells or how the intestinal microflora
influence this process. Most studies on the effect of fermented
milk or specific LAB on gut immune function have centered on
their immune adjuvant effects in the gut.
The ability of LAB to modulate IgA concentrations in the gut
has also been the subject of several studies. Orally administered
L. acidophilus and L. casei and the feeding of yogurt increased
both IgA production and the number of cells secreting IgA in the

small intestine of mice in a dose-dependent manner (5). Similarly, a report by Puri et al (56) indicated that S. typhimuriuminduced serum IgA concentrations were significantly higher in
mice fed yogurt over a period of 4 wk than in milk-fed control
mice. This report suggests that the IgA secreted by the challenged
intestinal B cells enters the circulation and increases the concentrations of IgA in the serum. Thus the IgA-enhancing effect of
yogurt intake may have both an effect on the gut and a systemic
effect. The same study also showed that intestinal lymphocytes
from mice fed yogurt had a higher mitogen-induced proliferative
response after a challenge with S. typhimurium than did those
from control-fed mice.
In a study using human subjects, Link-Amster et al (57)
showed that the specific anti-IgA titer to S. typhimurium was 4
times greater in subjects fed fermented milk containing L. acidophilus than in control subjects fed diets without fermented
milk. Total sIgA concentrations also increased in subjects fed
fermented milk.
Macrophages play an important role as a part of the innate
immune response in the gut, and they represent one of the first
lines of nonspecific defense against bacterial invasion. The effects of feeding milk fermented with either L. casei or L. acidophilus or both on the specific and nonspecific host defense
mechanisms in Swiss mice were investigated by Perdigon et al
(58). They showed that feeding milk fermented with L. casei, L.
acidophilus, or both for 8 d increased the in vitro and in vivo
phagocytic activity of peritoneal macrophages and the production of antibodies to sheep red blood cells. The activation of the
immune system began on day 3, peaked on day 5, and decreased
somewhat on day 8 of feeding. Phagocytic activity was further
boosted in mice given a single dose of fermented milk on day 11
of feeding.
Modulation of cytokine production by yogurt and LAB has
also been the focus of several studies. In addition to interleukin
(IL)-1 and tumor necrosis factor (TNF) , which are mainly
produced by macrophages, T lymphocytes are the source of most
cytokines investigated in those reports. T cells are frequently
classified into 2 categoriestype 1 (Th1) and type 2 (Th2) helper
T cells. On activation, these cells produce 2 diverse patterns of
cytokines (59). Th1 cells are the main producers of interferon-
(IFN-) and IL-2, and Th2 cells produce IL-4, IL-5, IL-6, and
IL-10. The Th1 cytokines boost cell-mediated immunity, and the
Th2 cytokines augment humoral immunity. IFN- plays a critical role in the induction of other cytokines and in mediation of
macrophage and natural killer cell activation.
Several reports indicated that consumption of yogurt or intake of
LAB by themselves modulates the production of several cytokines,
such as IL-1, IL-6, IL-10, IL-12, IFN-, and TNF- (60 63).
Moreover, the production of IFN- in an in vitro culture system
using human lymphocytes was reported to be greater with cultures
in the presence of LAB (L. bulgaricus and S. thermophilus) than
with those without LAB (64). Yogurt containing live L. bulgaricus
and S. thermophilus was also reported to augment IFN- production
by purified T cells from young adults after 4 mo feeding (62).
Effects of yogurt consumption on the modulation of cytokine
production in the human gastrointestinal tract, whether by cells
of the GALT or by others, have not been investigated. These
types of studies, although feasible with the use of biopsy samples
from the intestines of healthy subjects (65), are difficult to carry
out, and good animal models currently do not exist.

YOGURT AND GUT FUNCTION

Even though cytokines play diverse roles in regulating


immune functions, some cytokines, eg, IL-1, IL-6, and TNF-,
have been given more attention than others because they have
traditionally been classified as proinflammatory and as such are
known to be associated with inflammatory conditions such as
Crohn disease and ulcerative colitis (66). Another diverse family
of immune modulators that play important roles in the health of
the gastrointestinal tract consists of chemokines and their receptors (67). Currently, only limited data have been published on the
effect of yogurt or its components on chemokine modulation in
the gastrointestinal tract. The effects of different strains of Lactobacillus on chemokine production by the intestinal epithelial
cell-line, HT-29, were investigated by Wallace et al (68). All 3
LAB species investigatedL. acidophilus, Lactobacillus rhamnosus, and Lactobacillus delbrueckii had suppressive effects
on the production of 2 chemokines, RANTES (a member of the
IL-8 superfamily of cytokines) and IL-8, by activated HT-29
cells. As is the case with proinflammatory cytokines, these chemokines are necessary for normal immune function. However, a
high production of these chemokines during an inflammatory
condition is believed to exacerbate the inflammatory response.

Few reports have discussed the effects of yogurt and LAB on


laxation. In the studies published, however, both significant effects (G Wilhelm, unpublished observations, 1993; 69) and no
effects (70) of yogurt or LAB on laxation and gastrointestinal
transit time were described.
Strandhagen et al (69) reported that the transit time for 50%
(t50) of gastric content was significantly greater for ropy milk, an
L. bulgaricus and S. thermophilusfermented milk product
indigenous to Sweden, than for unfermented milk. Another study
showed that milk fermented with L. bulgaricus and S. thermophilus reduced intestinal transit time in human subjects with habitual
constipation (G Wilhelm, unpublished observations, 1993). In
the same study, subjects consuming fermented milk also had
improved bowel function. The number of defecations increased
from 3/wk during a control period to 7/wk when fermented milk
was consumed. When milk fermented with L. acidophilus was
consumed, the number of defecations increased further to 15/wk.
Studies were conducted of the effects of a commercially available yogurt fermented with B. animalis on orofecal gut transit
time (71, 72). In a double-blind, randomized, crossover design,
B. animalis reduced the colonic transit time in a group of healthy
women aged 18 45 y (72). Likewise, in a group of elderly
subjects experiencing lengthy orofecal gut transit time but otherwise free of any gastrointestinal pathology, B. animalis intake
provided led to a significant reduction in transit time (71). Thus,
the effect of LAB ingestion on orofecal gut transit time appears
to be dependent on the bacterial strain used and the population
being studied.
YOGURT AND DISEASES OF THE
GASTROINTESTINAL TRACT

Lactase deficiency and lactose maldigestion


Lactase deficiency among adults is the most common of all
known enzyme deficiencies. More than half of the worlds adult
population is lactose intolerant. In developmental terms, this may
not necessarily be considered abnormal, because humans are the

only known mammal in whom lactase activity in the small


intestine is sustained after weaning. In the case of lactose maldigestion, undigested lactose remains in the intestinal lumen,
and, as it reaches the colon, it is fermented by colonic bacteria.
Byproducts of this process include short-chain fatty acids such as
lactate, butyrate, acetate, and propionate. These fatty acids associate with electrolytes and lead to an osmotic load that can
induce diarrhea. Furthermore, fermentation of lactose by colonic
bacteria produces methane, hydrogen, and carbon dioxide. These
gases may stay in the lumen and eventually will both be excreted
as flatus, diffusing into the circulation, and be exhaled via the
lungs. Exhaled hydrogen after a lactose load has been used as an
indirect but measurable indicator of lactose maldigestion. In
addition to lactose, some sources of dietary fiber and other unabsorbed carbohydrates can serve as substrates for colonic fermentation that results in increased hydrogen production.
Inability to digest lactose varies widely among ethnic and
geographic populations (73, 74). In the United States, the prevalence of primary lactose intolerance in adults is 53% among
Mexican Americans, 75% among African Americans, and 15%
among whites. The prevalence among adults in South America
and Africa is 50% and that in some Asian countries is close to
100%. Lactose intolerance varies greatly between European
countries, from 2% prevalence in Scandinavian adults to
70% among Southern Italian adults (74).
Lactose maldigestion may develop secondary to inflammation
or as a result of functional loss of the small intestinal mucosa (14),
which can result from conditions such as Crohn disease, celiac
sprue, short bowel syndrome, or bacterial and parasitic infections. In addition, lactose maldigestion may develop as a consequence of severe protein calorie malnutrition. The disorder is
clinically expressed by symptoms of abdominal cramps, diarrhea, and flatulence after milk ingestion. However, most persons
who have symptoms of lactose intolerance can endure small
amounts (210 g) of lactose in a meal without becoming symptomatic (14).
It is well known that, for many lactose-intolerant people, fermented milk products are better accepted than are unfermented
milk products. There may be more than one reason for this.
During fermentation of milk, lactose is partially hydrolyzed,
which results in a lower lactose content in yogurt than in milk (2).
However, this reduction in lactose may not be significant, because milk solids are usually added during processing. The
greater tolerance of lactose from yogurt than of that from milk
among lactose-intolerant subjects may be due to the endogenous
lactase activity of yogurt organisms (13, 15, 75). Kolars et al (15)
used a series of breath hydrogen tests as well as a subjective
assessment to ascertain whether subjects who were identified as
lactose-intolerant digested and absorbed lactose in yogurt better
than they digested and absorbed lactose in milk. The area under
the curve for breath hydrogen was smaller after yogurt consumption than after consumption of milk or lactose in water, which
indicates better digestion and absorption of lactose from yogurt
than of that from either milk or lactose in water. Subjective
assessment by the subjects in the study of Kolars et al also
indicated that lactose in yogurt was better tolerated than the same
amount of lactose from milk or in water. Using breath hydrogen
measurement, Savaiano et al (75) investigated the effects of 3
varieties of cultured milk products on the digestion of lactose by
9 lactase-deficient human subjects. When yogurt, cultured milk
(buttermilk), and sweet acidophilus milk were compared, yogurt

Downloaded from www.ajcn.org by guest on June 6, 2011

Laxation

249

250

ADOLFSSON ET AL

Diarrheal diseases
Diarrhea is a common problem among children worldwide and
has been reported to contribute substantially to pediatric physician visits and hospitalizations in the United States (81). Since
the early 20th century, it has been hypothesized that live bacterial
cultures, such as those used for the fermentation of dairy products, may offer benefits in preventing and treating diarrhea (4).
A recent meta-analysis of randomized, controlled studies by
Van Neil et al (82) found that therapy using Lactobacillus strains
offered a safe and effective means of treating acute infectious
diarrhea in children. Both the duration and frequency of diarrheal
episodes were reduced when compared with those in control
subjects. The benefit of Lactobacillus therapy was seen in diarrheal
diseases caused by various pathogens. The effect of supplementing
formula with B. bifidum and S. thermophilus on preventing the onset
of acute viral diarrhea in infants was examined in a double-blind,

placebo-controlled trial (83). The infants receiving bacterial therapy


developed diarrhea and shed rotavirus less than did the infants fed
the control formula. Evidence of the beneficial effect of LAB on the
occurrence of diarrhea of bacterial origin is more contradictory because both benefits (84, 85) and no effects (86, 87) of feeding LAB
were reported.
Several studies investigated the effects of probiotic bacteria on
diarrhea associated with the use of antibiotics. The most likely
cause of diarrhea associated with antibiotic use is the negative
influence of antibiotics on the bacterial steady state of the intestines (88). Most cases of antibiotic-associated diarrhea are mild,
and they end shortly after antibiotic therapy is discontinued. A
less common but more serious type of antibiotic-associated diarrhea is due to antibiotic-mediated overgrowth of pathogenic
bacterial species such as Clostridium difficile that is associated
with pseudomembranous colitis (89).
A recent meta-analysis evaluated the ability of several different probiotic LAB species to prevent antibiotic-associated diarrhea (90). Of the 9 studies that were included in the analysis, 4
used Lactobacilli strains or a combination of Lactobacilli and
Bifidobacteria (9194). Of those 4 studies, 2 showed a significant benefit of probiotic use in comparison with placebo (93, 94).
The authors concluded that probiotic bacteria supplied in capsules or as yogurt-based products may be useful in preventing
antibiotic-associated diarrhea. However, none of these studies
provide evidence for a role of probiotic bacteria in the treatment
of such diarrhea.
The mechanisms by which LAB may provide a beneficial
effect against some forms of diarrheal disease are unknown. It
has been suggested that the beneficial effect may stem from the
ability of LAB to reestablish the intestinal microflora, to increase
the intestinal barrier by competing with pathogenic bacteria for
adhesion to the enterocytes, or to increase mucosal IgA response
to pathogens.
Colon cancer
According to the National Cancer Institute, cancer of the colon
is the second leading cancer diagnosis among both women and
men in the United States (95). Colon cancer is also the second
most common cause of cancer death. Risk factors for colorectal
cancer include both genetic and environmental factors, and several reports have suggested that interactions between dietary
factors, colonic epithelium, and intestinal flora are central to the
development of colon cancer.
The role of diet in the etiology of cancer has been given greater
attention in recent years. Although the relation between colon
cancer and certain food constituents, such as fiber and fat, generated the most interest, the possibility that fermented dairy products may protect against tumor formation in the colon was also
investigated. Epidemiologic evidence suggests a negative correlation between the incidence of certain cancers, including colon
cancer, and the intake of fermented dairy products (96). Moreover, fermented dairy products or the bacteria used for milk
fermentation were shown to have an effect on colon cancer and
certain other tumors in murine models of carcinogenesis (97
100). However, a number of animal studies investigating the
effect of various strains of LAB on colon carcinogenesis showed
inconsistent results.
Wollowski et al (100) investigated the protective effect of several
strains of LAB, traditionally used for milk fermentation, against
1,2-dimethylhydrazine (DMH)induced colon carcinogenesis in

Downloaded from www.ajcn.org by guest on June 6, 2011

had the most beneficial effect on lactose digestion in these subjects. Lactase activity and the number of surviving LAB were
significantly reduced when the yogurt was pasteurized.
The enzyme activity of lactase is generally stable in response
to environmental factors. For example, it was shown that the lactase
activity of yogurt was preserved and even increased when the yogurt
was subjected to an environment that simulated the temperature
and low pH values of the gut (15). As suggested by the authors, this
study supports the notion that lactose in yogurt is autohydrolyzed
once it is in the jejunal environment. Other studies reported that
lactase activity is less stable in response to acidic environment.
Pochart et al (76) reported that lactase activity in yogurt decreased by
80% at a pH of 5.0 in an in vitro model.
However, heating yogurt does significantly decrease lactase
activity, which indicates that yogurt that has been heat treated is
not as beneficial for lactose-intolerant persons is yogurt containing live and active cultures. Thus, there is a growing body of
evidence that yogurt containing live and active cultures is better
tolerated by lactose malabsorbers than are heat-treated fermented milks (50). During the fermentation process, the amount
of lactose present in yogurt is reduced. The lactose content also
varies with the duration of storage after fermentation. In addition,
the bacterial lactase activity corresponds with the survival time of
lactobacilli after ingestion. The enhanced digestion of lactose is
explained partly by the improved lactase activity after yogurt
ingestion and partly by other enzymatic functions, such as the
activity of the lactose transport system (permease) that allows
lactose to enter the probiotic cell (77, 78). Furthermore, animal
studies have suggested that LAB may induce lactase activity of
the gut intestinal endothelial cells (79).
A study by Martini et al (80) supports the microbial mediation
of lactase activity in the gastrointestinal tract. Those authors
showed that lactase activity in yogurt was stable at pH 4.0, but
that microbial cell disruption resulted in 80% loss of lactase
activity and a twofold increase in lactose malabsorption in a
group of lactose maldigesters.
Although the organisms that make up the live cultures in
yogurt are recognized as having functional lactase activity and as
contributing to the digestion of lactose, their survival in the
gastrointestinal tract is short. On average, significant numbers
survive for 1 h after ingestion (15, 50). Regardless of this
somewhat limited survival time, the beneficial effect of LAB on
lactose digestion in those suffering from lactose intolerance is
now widely accepted.

YOGURT AND GUT FUNCTION

Inflammatory bowel disease


Inflammatory bowel disease (IBD) is a term used for certain
chronic immunemediated conditions of the intestinal tract.
These chronic diseases include Crohn disease and ulcerative
colitis, conditions that have comparable symptoms but that affect
the digestive tract in very different ways (66). Ulcerative colitis
involves inflammation of the colon and rectum and not that of the
upper gastrointestinal tract, whereas Crohn disease can affect the
upper intestinal digestive tract and thus can lead to malabsorption
of both macronutrients and micronutrients. The etiologies of
these diseases are unknown, but studies suggest that the intestinal
microflora play a crucial pathogenic role (108). This notion is

supported by animal models of Crohn disease, in which the


presence of intestinal microflora is absolutely required for the
development of disease.
Proinflammatory cytokines, particularly TNF-, have also
been recognized as playing a central role in the pathogenesis of
Crohn disease. However, despite earlier hopes, the results from
studies using TNF- antagonists were disappointing, and there
were some reports of severe complications (109). Nevertheless,
reducing the production or effect of TNF- (or both) in Crohn
disease patients is belived to be beneficial. Bourrel et al (63)
reported that, when inflamed intestinal mucosa from a group of
Crohns disease patients was cocultured in the presence of L.
casei or L. bulgaricus, expression and release of TNF- by
intraepithelial lymphocytes were reduced.
Normally, a healthy mucosal barrier provides a first defense
mechanism against both the intestinal microflora and invading
pathogens. It has been suggested that the proportions of different
intestinal microflora are altered in patients with IBD. For example, colonic biopsy specimens have shown lower concentrations
of Lactobacillus and lower fecal concentrations of both Lactobacillus and Bifidobacterium species in patients with Crohn disease than in healthy subjects (110). This disturbance in intestinal
flora may increase the opportunity for colonization of pathogens
and bring about a subsequent proinflammatory response.
In the case of IBD, a defective mucosal barrier allows for increased uptake of antigens and proinflammatory mediators originating from luminal bacteria. It has been reported that patients with
IBD have diminished mucosal protection as a result of changes in the
composition and thickness of the mucosal layer and alterations in the
glycosylation status of mucosal glycoproteins (111). These changes
in the intestinal mucosa are also associated with decreased intestinal
IgA activity and increased IgG activity, which coincides with reduced state of protection and a proinflammatory condition. With
weakened mucosal barrier and thereby increased adherence of bacterial pathogens to the mucosa, sustained inflammation results, and
that leads to further damage to the gut mucosa. In recent years,
immunosuppressive and immunomodulating therapies, such as the
steroids used since the 1960s, have become more and more frequent
in the treatment of these conditions. Although efficacious, these
types of drugs can increase the prevalence of opportunistic infections as well as the severity of any underlying infection that may be
present (112). Other side effects of these treatments may include
hepatotoxicity, fibrosis, lymphoma, and pathologic suppression of
bone marrow function.
The role of beneficial intestinal microflora in the prevention of
intestinal inflammation was investigated by using gene-targeted
IL-10 knockout (IL-10/) mice (113, 114). These IL-10 deficient mice spontaneously develop ileocolitis with many similarities to Crohn disease in humans. Furthermore, affected mice
respond favorably to immunosuppression or immunomodulatory drugs that are similar to those used to treat human IBD. The
immunoregulatory activity of IL-10 has been studied extensively. It is now well established that IL-10 plays a role in downregulating both the synthesis of inflammatory cytokines and
the presentation of antigens. Thus, IL-10 has been suggested for
use as an immunomodulator for the treatment of Crohn disease.
Targeted in vivo delivery of IL-10 to the affected intestinal epithelium by using genetically engineered Lactococcus lactis has
shown great promise in 2 mouse models of IBD (114).
Madsen et al (113) found that IL-10/ mice had increased
adherence of luminal bacteria to the mucosal layer in the colon

Downloaded from www.ajcn.org by guest on June 6, 2011

rats. Oral treatment with L. bulgaricus for 4 d protected against


DMH-induced DNA damage in the colon. In contrast, there was no
protective effect when S. thermophilus was administered. The authors did not ascertain the mechanisms of protection by L. bulgaricus, but they speculated that thiol-containing breakdown products of
proteins that result from the proteolytic activity of L. bulgaricus may
have produced the effect.
In a previous study using a similar DMH-induced colon cancer
model in rats, Shackelford et al (99) showed that milk fermented
with L. bulgaricus resulted in greater survival than did nonfermented milk. However, in contrast to the findings of Wollowski
et al (100), L. bulgaricus-fermented milk did not reduce the
number of rats that developed colon tumors, whereas S.
thermophilus-fermented milk did do so (99). In a study using
azoxymethane to induce aberrant crypt foci in the colon of rats,
no significant effects were seen with either B. longum or L. casei
(101). Those authors did, however, observe a protective effect of
L. acidophilus and inulin, but only when the total fat content of
the diet was increased.
Using a colon carcinoma cell culture system, Ganjam et al
(102) isolated a yogurt fraction that decreased cell proliferation,
as ascertained with the use of thymidine incorporation. Cell proliferation was not inhibited in response to a similarly isolated
milk fraction or to lactic acid.
Elevated activity of several bacterial fecal enzymes, some of
which are involved in the metabolism of genotoxic nitrates, was
associated with an increased risk of colon cancer (103, 104). The
activity of these enzymes can be altered by diet or antibiotic
intake (10, 105). L. acidophilus (106) and L. gasseri (43) were
shown to reduce the fecal enzyme activity of nitroreductase,
azoreductase, and -glucuronidase in humans, with a reduction
by 50% or 75% in the activities of these enzymes during a period
of Lactobacilli feeding. Likewise, Guerin-Danan et al (46) reported that 10 18-mo-old infants fed yogurt fermented with S.
thermophilus, L. bulgaricus, and L. casei had lower fecal
-glucuronidase activity than did a similar group of infants fed
milk or yogurt not fermented with L. casei.
The mechanism by which LAB may have an effect on colon
carcinogenesis is currently unknown. Some of the mechanisms
that may be involved include enhancement of the hosts gut immune
response, suppression of harmful intestinal bacteria, sequestration
of potential mutagens, production of antimutagenic compounds,
reduction of pH concentrations in the colon, and alteration of other
physiologic conditions (107). Furthermore, it was shown by Pedrosa
et al (43) that the feeding of yogurt or Lactobacillus reduced fecal
enzymes, which convert procarcinogens to carcinogens, such as
azoreductase and nitroreductase.

251

252

ADOLFSSON ET AL

Helicobacter pylori
It has only been 20 y since Helicobacter pylori, a gramnegative, spiral-shaped bacterium that is found in the gastric
mucous layer or adherent to the epithelial lining of the stomach,
was discovered (117). H. pylori relies on the ammoniaproducing surface protein urease for adherence and colonization
to the gastric epithelium. Urease allows H. pylori to survive by
neutralizing the acidic gastric environment (118). H. pylori produces catalase, which may play a role in protecting the bacteria
from free radicals that are released by activated leukocytes. H.
pylori infection is associated with a massive infiltration of neutrophils into the gastric wall and local production of IFN-,
proinflammatory cytokines eg, TNF-, IL-1, and IL-6 and
the chemokine IL-8.
Infection with H. pylori is now known to play a role in peptic
ulcer disease, chronic gastritis, gastric adenocarcinoma, and
mucosa-associated lymphoid tissue lymphoma. The association
between duodenal ulcer disease and H. pylori is also well documented: H. pylori infection is reported in 90% of duodenal
ulcer patients (119). Treatment of this infection involves the use
of proton pump inhibitors, often in combination with antibiotics.
However, the use of antibiotics to treat H. pylori infection has
been associated with adverse effects and frequently leads to
resistance to antibiotic therapy.
Several in vitro and animal studies have shown reduced viability of H. pylori and less adhesion of the bacteria to human
intestinal mucosal cells after treatment with various Lactobacillus strains (120). In series of in vitro assays, Midolo et al (121)
showed that the growth of H. pylori was inhibited by lactic acid
in a pH-independent manner. They also found that 6 strains of L.
acidophilus and L. casei inhibited the growth of H. pylori,

whereas B. bifidus and L. bulgaricus did not. The inhibitory


effect correlated with the concentrations of lactic acid produced
by the LAB examined. In another study, Coconnier et al (122)
reported that conditioned media from L. acidophilus reduced the
viability of H. pylori in vitro, independent of lactic acid concentrations. In addition, the adhesion of H. pylori to human mucosecreting HT-29 cells decreased. Several in vitro studies were conducted to ascertain whether the effects of LAB on H. pylori
survival and function are due to lactic acid or to other antibacterial products generated by LAB, such as bacteriocins. Of the
several bacteriocins tested, lacticins produced by Lactoc. lactis
were shown to have the greatest anti-Helicobacter activity when
used against several strains of H. pylori (123).
Studies that indicate promising inhibitory effects of LAB on H.
pylori survival and function in vitro were extended to in vivo studies
using human patients. Armuzzi et al (124) reported that, when 120
asymptomatic subjects who were positive for H. pylori infection
received an L. casei strain GG supplement over a 14-d period in
addition to a standard 1-wk antibiotic therapy regimen, the eradication of H. pylori was faster than that in control subjects.
Although promising results have been reported, the effects of
LAB on H. pylori infection in humans remain ambiguous. For
example, L. acidophilus and L. gasseri were both shown to
decrease H. pylori infection, as indicated by reduced [13C] urea
breath test values (125, 126), and therapy with L. acidophilus was
shown to reduce gastric mucosal inflammation (125). However,
gastric biopsies did not show eradication of H. pylori. Similarly,
Cats et al (127) reported that viable L. casei was required to
inhibit the growth of H. pylori in vitro, but only a slight nonsignificant trend was observed toward an in vivo suppressive effect
of an L. casei-supplemented milk drink.
Allergic reactions
The effects of yogurt and LAB on allergic reactions in the
gastrointestinal tract have received some interest (128, 129). It
was reported that a delay in the development of Bifidobacterium
and Lactobacillus in the gastrointestinal microflora is a general
finding in children with allergic reactions (128). Isolauri (130)
reported data suggesting that Lactobacillus GG can be used to
prevent food allergies.
Heat treatment was suggested as a way of reducing the ability
of milk proteins to cause allergic reactions, which would make
milk a more suitable source of protein for persons with an immunologic sensitization to cow milk protein (131). However,
Kirjavainen et al (129) used a randomized double-blind design to
investigate in a recent study the effects of heat-inactivated and
viable L. rhamnosus GG on infants with atopic eczema and cow
milk allergy. Milk formula supplemented with viable but not
heat-inactivated L. rhamnosus GG significantly improved atopic
eczema and subjective symptoms of cow milk allergy in subjects
in comparison with the control group. These results suggest that,
in persons with cow milk allergy, the presence of viable LAB
may provide benefits that outweigh the possible detrimental
effects that undenatured milk proteins may have on milk allergy. Furthermore, the immunologic response to native milk
proteins may differ from that to heat-denatured milk proteins. A
recent study using a rat model showed that heat-denaturated
-lactoglobulin induced a local mucosal inflammatory response,
whereas native -lactoglobulin induced an IgE-mediated
systemic response (132). Heat denaturation is likely to result in

Downloaded from www.ajcn.org by guest on June 6, 2011

that preceded the development of colitis. This occurred in parallel to decreased numbers of luminal Lactobacillus. When the
concentrations of Lactobacillus in the gastrointestinal lumen
were restored by rectal delivery of Lactobacillus reuteri or by
oral lactulose therapy, colitis was attenuated. The concentrations
of adherent and translocated bacteria in the mucosal wall also
were reduced.
Another benefit of LAB in Crohn disease may be due to the
stimulation of the IgA response. A report by Malin et al (115)
suggests that oral bacteriotherapy using L. casei can restore
antigen-specific IgA immune response in persons with Crohn
disease. In a previous study from the same laboratory (116), oral
administration of L. casei to patients with viral gastroenteritis
promoted antigen-specific IgA responses and shortened the patient diarrhea.
Although experimental evidence exists indicating beneficial
effects of LAB on Crohn disease and ulcerative colitis, the exact
mechanism through which LAB species antagonize the progression of these diseases is poorly understood. The exact etiology of
IBD is also unknown, but it is likely that, in susceptible persons,
IBD results from an ongoing inflammatory response, which may
be due to a defect in both the regulation of the mucosal proinflammatory response and the function of the intestinal epithelium. Currently, evidence suggests that yogurt and LAB have
modest clinical benefits and are safe for use in patients with these
conditions. Further studies are required to ascertain whether yogurt is beneficial as a prophylactic or a therapeutic regimen for
IBD (or both) and to establish exactly which mechanisms are
involved.

YOGURT AND GUT FUNCTION

SAFETY

Although the safe use of nonsporing anaerobic LAB in fermented foods is widespread and has a long history, there have
been occasional reports associating LAB with clinical infections
(53, 136) because benign microorganisms have been shown to be
infective when a patient is severely debilitated or immunosuppressed (137, 138). Some of the diseases that have been associated with LAB infection include septicemia, infective endocarditis, and dental caries.
Very rarely, cases of lactobacillemia have been reported in
patients with severe underlying illness, many of whom received
a prior antibiotic therapy that may have selected-out for the
organism (139, 140). Moreover, Husni et al (141) reviewed the
cases of 45 patients with clinically significant lactobacillemia
and reported that 11 of the patients were receiving immunosuppressive therapy and 23 had received antibiotics. In none of these
reports was a definitive link made between the consumption of
fermented milk products and infection.
In addition, rare cases of endocarditis have been associated
with L. rhamnosus, a LAB indigenous to the human gastrointestinal tract (142144). However, as with lactobacillemia, no reports to date have been able to identify a connection between
LAB from fermented milk and infection in humans. In most of
these cases, the origin of the Lactobacillus is most likely the host.

There is also a hypothetical risk of the transfer of antimicrobial


resistance from LAB to other microorganisms with which LAB
might come in contact, but this has not yet been described in the
literature.
In the past, Lactobacilli isolated from infections were habitually dismissed as contaminants or secondary invaders. However,
recent evidence suggests that they might function as opportunistic
pathogens in a small number of severely immunosuppressed
persons. Even in these patients, this is a very rare event, and it has
not yet been reported in a large group of immunosuppressed
persons, such as the elderly or persons with AIDS. LAB have a
long history of safe use in foods and also in products that have
been tested in clinical trials. However, as with any new food
ingredient, the safety of a new strain of LAB must be clearly
established before it is introduced into fermented dairy products.
CONCLUDING REMARKS AND RECOMMENDATIONS
FOR FUTURE STUDIES

It has long been believed that the consumption of yogurt and


other fermented milk products provides various health benefits.
Recent studies of the possible health benefits of yogurt in gutassociated diseases substantiate some of these beliefs. Of particular interest are the reduction by yogurt, yogurt bacteria, or
bothin the duration of diarrheal diseases in children, the preventive or therapeutic (or both) effects on IBD and colon cancer
as suggested by epidemiologic evidence and animal studies, and
the possible beneficial effects in increasing the eradication rate of
H. pylori as indicated by in vitro and preliminary human studies.
In addition, there is ever-increasing evidence of the beneficial
effect of yogurt containing live and active cultures on the digestion of lactose in persons with lactose intolerance.
These findings are interesting and should encourage future
studies to 1) substantiate or extend these findings by using animal
models and clinical trials; 2) ascertain whether these effects are
age-specific or can be observed across all age groups: eg, ascertain whether yogurt would have effects similar to those observed
in children on attenuation of the incidence or duration of diarrheal diseases in elderly people, a group that has high morbidity
and mortality from these infections; and 3) investigate the mechanisms through which yogurt exerts its effects and ascertain the
critical components of yogurt involved in its mechanisms of
action. Finally, in recent years, yogurt has been touted as improving gut health. In the absence of a universally accepted
definition or any definition of gut health, it is difficult to substantiate these claims. Studies focused on determining the characteristics of a healthy gut would be extremely helpful in evaluating the effect of yogurt on gut health.
All 3 authors participated in the literature review and the development of
the manuscript outline, and SNM and RMR determined the areas to be
discussed. OA conducted the literature search and organized and wrote the
manuscript. SNM provided corrections. RMR revised the manuscript.
This review was prepared in response to a request from the National
Yogurt Association for a critical and objective review, for which the authors
received an honorarium.

REFERENCES
1. Guarner F, Schaafsma GJ. Probiotics. Int J Food Microbiol 1998;39:
237 8.
2. Bourlioux P, Pochart P. Nutritional and health properties of yogurt.
World Rev Nutr Diet 1988;56:21758.

Downloaded from www.ajcn.org by guest on June 6, 2011

conformational changes that expose or hide (or both) epitopes


and lead to the activation of different subpopulations of immune
cells and thus to different end results.
The mechanisms of the protective effects of LAB on allergic
reactions are not known. A proinflammatory response in the gut
mucosa that is induced by food allergens may impair the function
of the intestinal barrier. It is possible that LAB may prevent
allergic reactions by having a protective effect on the function of
the intestinal barrier, although the mechanism of such an effect is
poorly understood. A more direct link between the function of
GALT and allergic responses is also possible. One of the primary
mechanisms of active cellular suppression of proinflammatory
events in the gut after antigen-specific triggering is the secretion
of suppressive cytokines, such as transforming growth factor
and IL-10. Transforming growth factor is produced by both
CD4 and CD8 GALT-derived T cells and is an important
mediator of the active suppression component of oral tolerance.
Furthermore, IL-4 mediated isotype switching of immunoglobulin from IgM to IgE and IgE-dependent degranulation of mast
cells has been shown to be involved in the pathogenesis of food
allergyrelated enteropathy (133).
Yogurts LAB are known to enhance the production of IFN-
(62, 134), which acts to inhibit isotype switching to IgE. IgEmediated hypersensitivity reaction, also known as type 1 allergy,
is triggered by the cross-linking of antigens with IgE antibodies
that are bound to Fc receptors on mast cells. It was reported that
L. casei inhibited antigen-induced IgE production by mouse
splenocytes (135). In addition, production of the immunosuppressive cytokine IL-10 is induced by LAB (60).
A combination of enhancing and suppressive effects is the
most likely mechanism by which LAB may have their effects.
However, the ways in which LAB or other components of yogurt
influence the production of these immunoregulatory cytokines in
the gut remain to be elucidated, as do the possible mechanisms of
LAB-mediated protection.

253

254

ADOLFSSON ET AL

31.
32.
33.
34.
35.
36.
37.
38.
39.

40.
41.
42.
43.

44.

45.
46.

47.
48.
49.

50.

51.
52.

53.
54.
55.

expression of G1-restriction points in breast and colon cancer cells. J


Nutr 2003;133:3670 7.
Block G, Abrams B. Vitamin and mineral status of women of childbearing potential. Ann N Y Acad Sci 1993;678:244 54.
Ervin RB, Kennedy-Stephenson J. Mineral intakes of elderly adult
supplement and non-supplement users in the third National Health and
Nutrition Examination Survey. J Nutr 2002;132:34227.
Allen LH. Calcium bioavailability and absorption: a review. Am J Clin
Nutr 1982;35:783 808.
Kaup SM, Shahani KM, Amer MA, Peo ER. (Bioavailability of calcium
in yogurt.) Milchwissenschaft 1987;42:513 6 (in German).
Schaafsma GJ, Dekker PR, de Ward H. Nutritional aspects of yogurt. 2.
Bioavailability of essential minerals and trace elements. Neth Milk
Dairy J 1988;42:135 46.
Bronner F, Pansu D. Nutritional aspects of calcium absorption. J Nutr
1999;129:9 12.
Norman AW. Intestinal calcium absorption: a vitamin D-hormone
mediated adaptive response. Am J Clin Nutr 1990;51:290 300.
Pointillart A, Cayron B, Gueguen L. Calcium and phosphorus utilization and bone mineralization in yogurt-fed pigs. Sci Alim 1986;6:15
30.
Bernet MF, Brassart D, Neeser JR, Servin AL. Lactobacillus acidophilus LA 1 binds to cultured human intestinal cell lines and inhibits cell
attachment and cell invasion by enterovirulent bacteria. Gut 1994;35:
4839.
Alm L, Pettersson L. Survival rate of lactobacilli during digestion: an
in vitro study Am J Clin Nutr 1980;33(suppl):S2543 (abstr).
Robins-Browne RM, Path FF, Levine MM. The fate of ingested lactobacilli in the proximal small intestine. Am J Clin Nutr 1981;34:514 9.
Conway PL, Gorbach SL, Goldin BR. Survival of lactic acid bacteria in
the human stomach and adhesion to intestinal cells. J Dairy Sci 1987;
70:112.
Pedrosa MC, Golner BB, Goldin BR, Barakat S, Dallal GE, Russell
RM. Survival of yogurt-containing organisms and Lactobacillus gasseri (ADH) and their effect on bacterial enzyme activity in the gastrointestinal tract of healthy and hypochlorhydric elderly subjects. Am J
Clin Nutr 1995;61:3539.
Clark PA, Martin JH. Selection of bifidobacteria for use as dietary
adjuvants in cultured dairy foods: III. Tolerance to stimulated bile
concentrations of human small intestines. Cult Dairy Prod J 1994;29:
18 21.
Duez H, Pelletier H, Cools S, et al. A colony immunoblotting method
for quantitative detection of a Bifidobacterium animalis probiotic strain
in human faeces. J Appl Microbiol 2000;88:1019 27.
Guerin-Danan C, Chabanet C, Pedone C, et al. Milk fermented with
yogurt cultures and Lactobacillus casei compared with yogurt and
gelled milk: influence on intestinal microflora in healthy infants. Am J
Clin Nutr 1998;67:1117.
Bouhnik Y, Pochart P, Marteau P, Arlet G, Goderel I, Rambaud JC.
Fecal recovery in humans of viable Bifidobacterium sp ingested in
fermented milk. Gastroenterology 1992;102:875 8.
Plant L, Conway P. Association of Lactobacillus spp. with Peyers
patches in mice. Clin Diagn Lab Immunol 2001;8:320 4.
Bernet MF, Brassart D, Neeser JR, Servin AL. Adhesion of human
bifidobacterial strains to cultured human intestinal epithelial cells and
inhibition of enteropathogen-cell interactions. Appl Env Microbiol
1993;59:4121 8.
Lerebours E, N'Djitoyap Ndam C, Lavoine A, Hellot MF, Antoine JM,
Colin R. Yogurt and fermented-then-pasteurized milk: effects of shortterm and long-term ingestion on lactose absorption and mucosal lactase
activity in lactase-deficient subjects. Am J Clin Nutr 1989;49:8237.
Kankaanpa a P, Salminen SJ, Isolauri E, Lee YK. The influence of
polyunsaturated fatty acids on probiotic growth and adhesion. FEMS
Microbiol Lett 2001;194:149 53.
Kankaanpa a P, Yang B, Kallio H, Isolauri E, Salminen S. Effects of
polyunsaturated fatty acids in growth medium on lipid composition and
on physicochemical surface properties of Lactobacilli. Appl Env Microbiol 2004;70:129 36.
Aguirre M, Collins MD. Lactic acid bacteria and human clinical infection. J Appl Bacteriol 1993;75:95107.
Brandtzaeg P, Baekkevold ES, Farstad IN, et al. Regional specialization in the mucosal immune system: what happens in the microcompartments? Immunol Today 1999;20:14151.
Macpherson AJ, Gatto D, Sainsbury E, Harriman GR, Hengartner H,

Downloaded from www.ajcn.org by guest on June 6, 2011

3. Chandan RC, Shahani KM. Yogurt. In: Hui YH, ed. Dairy science and
technology handbook. New York: VCH Publishers, Inc, 1993:157.
4. Metchnikoff E. Sur la flore du corps humain. (On the flora of the human
body. ) Manch Lit Philos Soc 1901;45:138 (in French).
5. Perdigon G, Alvarez S, Rachid M, Aguero G, Gobbato NJ. Immune
system stimulation by probiotics. J Dairy Sci 1995;78:1597 606.
6. Buttriss J. Nutritional properties of fermented milk products. Int J Dairy
Tech 1997;50:217.
7. Reddy KP, Shahani KM, Kulkarni SM. B-complex vitamins in cultured
and acidified yogurt. J Dairy Sci 1976;59:1915.
8. Shahani KM, Chandan RC. Nutritional and healthful aspects of cultured and culture-containing dairy foods. J Dairy Sci 1979;62:168594.
9. Kneifel W, Mayer HK. Vitamin profiles of kefirs made from milks of
different species. Int J Food Sci Technol 1991;26:423 8.
10. Kneifel W, Kaufmann M, Fleischer A, Ulberth F. Screening of commercially available mesophilic dairy starter cultures: biochemical, sensory and morphological properties. J Dairy Sci 1992;75:3158 66.
11. Crittenden RG, Martinez NR, Playne MJ. Synthesis and utilisation of
folate by yoghurt starter cultures and probiotic bacteria. Int J Food
Microbiol 2003;80:21722.
12. Wigertz K, Svensson UK, Ja gerstad M. Folate and folate binding protein content in dairy products. J Dairy Res 1996;64:239 54.
13. Rosado JL, Solomons NW, Allen LH. Lactose digestion from unmodified, low-fat and lactose-hydrolyzed yogurt in adult lactose maldigesters. Eur J Clin Nutr 1992;46:617.
14. Vesa TH, Marteau P, Korpela R. Lactose intolerance. J Am Coll Nutr
2000;19:165S75S.
15. Kolars JC, Levitt MD, Aouji M, Savaiano DA. Yogurtan autodigesting source of lactose. N Engl J Med 1984;310:13.
16. Goodenough ER, Kleyn DH. Influence of viable yogurt microflora on
digestion of lactose by the rat. J Dairy Sci 1976;59:601 6.
17. Rasic JL, Kurmann JA. Yoghurt: scientific grounds, technology, manufacture and preparations. Vol 1 of Rasic JL, Kurmann JA, eds. Fermented fresh milk products and their cultures. Copenhagen: Technical
Dairy Publishing House, 1978.
18. Loones A. Transformation of milk components during yogurt fermentation. In: Chandan RC, ed. Yogurt: nutritional and health properties.
McLean, VA: National Yogurt Association, 1989:95114.
19. Beshkova DM, Simova ED, Frengova GI, Simov ZI, Adilov EF. Production of amino acids by yogurt bacteria. Biotechnol Prog 1998;14:
9635.
20. Hewitt D, Bancroft HJ. Nutritional value of yogurt. J Dairy Res 1985;
52:197207.
21. Bissonnette DJ, Jeejeebhoy KN, eds. Meeting dietary nutrient requirements with cows milk and milk products. Rotterdam: Balkema, 1994.
22. Gaudichon C, Roos N, Mah S, Sick H, Bouley C, Tom D. Gastric
emptying regulates the kinetics of nitrogen absorption from 15Nlabeled milk and 15N-labeled yogurt in miniature pigs. J Nutr 1994;
124:1970 7.
23. Gaudichon C, Mah S, Roos N, et al. Exogenous and endogenous
nitrogen flow rates and level of protein hydrolysis in the human jejunum after [15N] milk and [15N] yogurt ingestion. Br J Nutr 1995;74:
251 60.
24. Shantha NC, Ram LN, OLeary J, Hicks CL, Decker EA. Conjugated
linoleic acid concentrations in dairy products as affected by processing
and storage. J Food Sci 1995;60:695 8.
25. Aneja RP, Murthi TN. Conjugated linoleic acid contents of Indian curd
and ghee. Indian J Dairy Sci 1990;43:231 8.
26. Jiang J, Wolk A, Vessby B. Relation between the intake of milk fat and
the occurrence of conjugated linoleic acid in human adipose tissue.
Am J Clin Nutr 1999;70:217.
27. Park Y, McGuire MK, Behr R, McGuire MA, Evans MA, Schultz TD.
High-fat dairy product consumption increases 9c; 11t18:2 (rumenic
acid) and total lipid concentrations of human milk. Lipids 1999;34:
5439.
28. Boccignone M, Brigidi R, Sarra C. Studi effettuati sulla compo-sizione
in trigliceridi ed acidi grassi liberi nello yogurt preparato dalatte vaccino, pecorinoe, caprino. (Studies on triglyceride and free fatty acid
composition of yogurt prepared from cow, goat, and sheep milk.) Ann
Fac Med Vet (Torino) 1984;28:22333 (in Italian).
29. Whigham LD, Cook ME, Atkinson RL. Conjugated linoleic acid: implications for human health. Pharmacol Res 2000;42:50310.
30. Kemp MQ, Jeffy BD, Romagnolo DF. Conjugated linoleic acid inhibits
cell proliferation through a p53-dependent mechanism: effects on the

YOGURT AND GUT FUNCTION

56.

57.

58.

59.

60.

61.

63.

64.

65.

66.
67.
68.

69.

70.

71.

72.

73.

74.
75.

76.

77.

78.
79.

80.

81.

82.

83.

84.

85.

86.

87.

88.
89.

90.
91.

92.

93.

94.

95.

96.

97.

98.
99.

100.

101.

glucose, and galactose in homofermentative lactobacilli. Appl Environ


Microbiol 1986;51:82531.
Foucaud C, Poolman B. Lactose transport system of Streptococcus
thermophilus. J Biol Chem 1992;267:2208794.
Thoreux K, Balas D, Bouley C, Senegas-Balas F. Diet supplemented
with yoghurt or milk fermented by Lactobacillus casei DN-114 001
stimulates growth and brush-border enzyme activities in mouse small
intestine. Digestion 1998;59:349 59.
Martini MC, Bollweg GL, Levitt MD, Savaiano DA. Lactose digestion
by yogurt -galactosidase: influence of pH and microbial cell integrity.
Am J Clin Nutr 1987;45:432 6.
Glass RI, Lew JF, Gangarosa RE, LeBaron CW, Ho MS. Estimates of
morbidity and mortality rates for diarrheal diseases in American children. J Pediatr 1991;118:S2733.
Van Neil CW, Feudtner C, Garrison MM, Christakis DA. Lactobacillus
therapy for acute infectious diarrhea in children: a meta-analysis. Pediatrics 2002;109:678 84.
Saavedra JM, Bauman NA, Oung I, Perman JA, Yolken RH. Feeding of
Bifidobacterium bifidum and Streptococcus thermophilus to infants in
a hospital for prevention of diarrhoea and shedding of rotavirus. Lancet
1994;344:1046 9.
Gorbach SL, Chang TW, Goldin B. Successful treatment of relapsing
Clostridium difficile colitis with Lactobacillus GG. Lancet 1987;2:
1519(letter).
Biller JA, Katz AJ, Flores AF, Buie TM, Gorbach SL. Treatment of
recurrent Clostridium difficile colitis with Lactobacillus GG. J Pediatr
Gastroenterol Nutr 1995;21:224 6.
Shornikova AV, Isolauri E, Burkanova L, Lukovnikova S, Vesikari T.
A trial in the Karelian Republic of oral rehydration and Lactobacillus
GG for treatment of acute diarrhoea. Acta Paediatr 1997;86:460 5.
Clements ML, Levine MM, Ristaino PA, Daya VE, Huges TP. Exogenous lactobacilli fed to mantheir fate and ability to prevent diarrheal
disease. Prog Food Nutr Sci 1983;7:29 37.
Bartlett JG. Antibiotic-associated diarrhea. Clin Infect Dis 1992;15:
573 81.
Van der Waaij D. The ecology of the human intestine and its consequences for overgrowth by pathogens such as Clostridium difficile.
Annu Rev Microbiol 1989;43:69 87.
DSouza AL, Rajkumar C, Cooke J, Bulpitt CJ. Probiotics in prevention
of antibiotic associated diarrhoea: meta-analysis. BMJ 2002;324:1 6.
Gotz V, Romankiewicz JA, Moss J, Murray HW. Prophylaxis against
ampicillin associated diarrhoea with Lactobacillus preparation. Am J
Hosp Pharm 1979;36:754 7.
Tankanow RM, Ross MB, Ertel IJ, Dickinson DG, McCormick LS,
Garfinkel JF. A double blind, placebo-controlled study of the efficacy
of Lactinex in the prophylaxis of amoxicillin-induced diarrhea. DICP
1990;24:382 4.
Orrhage K, Brismar B, Nord CE. Effects of supplements of Bifidobacterium longum and Lactobacillus acidophilus on intestinal microbiota
during administration of clindamycin. Microb Ecol Health Dis 1994;
7:1725.
Vanderhoof JA, Whitney DB, Antonson DL, Hanner TL, Lupo JV,
Young RJ. Lactobacillus GG in the prevention of antibiotic-associated
diarrhoea in children. J Pediatr 1999;135:356 68.
National Cancer Institute. SEER Cancer Incidence Public-Use Database, 1973-1996, August 1998 Submission. Bethesda, MD: US Department of Health and Human Services, Public Health Service, 1999.
Peters RK, Pike MC, Garabrant D, Mack TM. Diet and colon cancer
in Los Angeles County, California. Cancer Causes Control 1992;3:
45773.
Reddy BS, Rivenson A. Inhibitory effect of Bifidobacterium longum on
colon, mammary, and liver carcinogenesis induced by 2-amino-3methylimidazo[4,5-f]quinoline, a food mutagen. Cancer Res 1993;53:
3914 8.
Ayebo AD, Shahani KM, Dam R. Antitumor component(s) of yogurt:
fractionation. J Dairy Sci 1981;64:2318 23.
Shackelford LA, Rao DR, Chawan CB, Pulusani SR. Effect of feeding
fermented milk on the incidence of chemically induced colon tumors in
rats. Nutr Cancer 1983;5:159 64.
Wollowski I, Ji S, Bakalinsky AT, Neudecker C, Pool-Zobel BL. Bacteria used for the production of yogurt inactivate carcinogens and prevent DNA damage in the colon of rats. J Nutr 1999;129:77 82.
Bolognani F, Rumney CJ, Pool-Zobel BL, Rowland IR. Effect of

Downloaded from www.ajcn.org by guest on June 6, 2011

62.

Zinkernagel RM. A primitive T cell-independent mechanism of intestinal mucosal IgA responses to commensal bacteria. Science 2000;288:
2222 6.
Puri P, Rattan A, Bijlani RL, Mahapatra SC, Nath I. Splenic and intestinal lymphocyte proliferation response in mice fed milk or yogurt and
challenged with Salmonella typhimurium. Int J Food Sci Nutr 1996;
47:391 8.
Link-Amster H, Rochat F, Saudan KY, Mignot O, Aeschlimann JM.
Modulation of a specific humoral immune response and changes in
intestinal flora mediated through fermented milk intakes. FEMS Immunol Med Microbiol 1994;10:55 64.
Perdigon G, de Macias ME, Alvarez S, Oliver G, de Ruiz Holgado AA.
Systemic augmentation of the immune response in mice by feeding
fermented milks with Lactobacillus casei and Lactobacillus acidophilus. Immunology 1988;63:1723.
Mosmann TR, Cherwinski H, Bond MW, Giedlin MA, Coffman RL.
Two types of murine helper T cell clone. I. Definition according to
profiles of lymphokine activities and secreted proteins. J Immunol
1986;136:2348 57.
Miettinen M, Vuopio-Varkila J, Varkila K. Production of human tumor
necrosis factor-alpha, interleukin-6 and interleukin-10 is induced by
lactic acid bacteria. Infect Immun 1996;64:54035.
Solis-Pereyra B, Aattouri N, Lemonnier D. Role of food in the stimulation of cytokine production. Am J Clin Nutr 1997;66(suppl):
521S5S.
Halpern GM, Vruwink KG, van de Water J, Keen CL, Gershwin ME.
Influence of long-term yogurt consumption in young adults. Int J Immunother 1991;7:20510.
Borruel N, Carol M, Casellas F, et al. Increased mucosal tumour necrosis factor alpha production in Crohns disease can be downregulated
ex vivo by probiotic bacteria. Gut 2002;51:659 64.
De Simone C, Bianchi Salvadori B, Negri M, Ferrazzi M, Baldinelli L,
Vesely R. The adjuvant effect of yogurt on production of gammainterferon by Con A stimulated human peripheral blood lymphocytes.
Nutr Rep Int 1986;33:419 33.
Beharka AA, Paiva S, Leka LS, Ribaya-Mercado JD, Russell RM,
Nibkin Meydani S. Effect of age on the gastrointestinal-associated
mucosal immune response of humans. J Gerontol A Biol Sci Med Sci
2001;56:B218 23.
Podolsky DK. Inflammatory bowel disease. N Engl J Med 2002;347:
41729.
Ajuebor MN, Swain MG. Role of chemokines and chemokine receptors
in the gastrointestinal tract. Immunology 2002;105:137 43.
Wallace TD, Bradley S, Buckley ND, Green-Johnson JM. Interactions
of lactic acid bacteria with human intestinal epithelial cells: effects on
cytokine production. Food Prot 2003;66:466 72.
Strandhagen E, Lia A, Lindstrand S, et al. Fermented milk (ropy milk)
replacing regular milk reduces glycemic responce and gastric emptying
in healthy subjects. Scand J Nutr 1994;38:11721.
Nakamura T, Nishida S, Mizutani M, Iino H. Effects of yogurt supplemented with brewers yeast cell wall on constipation and intestinal
microflora in rats. J Nutr Sci Vitaminol (Tokyo) 2001;47:36772.
Meance S, Cayuela C, Turchet P, Raimondi A, Lucas C, Antoine JM.
A fermented milk with a Bifidobacterium probiotic strain DN-173 010
shortened oro-fecal gut transit time in elderly. Microb Ecol Health Dis
2001;13:21722.
Marteau P, Cuillerier E, Meance S, et al. Bifidobacterium animalis
strain DN-173 010 shortens the colonic transit time in healthy women:
a double-blind, randomized, controlled study. Aliment Pharmacol Ther
2002;16:58793.
Rorick MH, Scrimshaw NS. Comparative tolerance of elderly from
differing ethnic backgrounds to lactose-containing and lactose-free
dairy drinks: a double-blind study. J Gerontol 1979;34:191 6.
Sahi T. Genetics and epidemiology of adult-type hypolactasia. Scand J
Gastroenterol 1994;202:720.
Savaiano DA, AbouElAnouar A, Smith DE, Levitt MD. Lactose malabsorption from yogurt, pasteurized yogurt, sweet acidophilus milk,
and cultured milk in lactase-deficient individuals. Am J Clin Nutr
1984;40:1219 23.
Pochart P, Dewit O, Desjeux JF, Bourlioux P. Viable starter culture,
-galactosidase activity, and lactose in duodenum after yogurt ingestion in lactase-deficient humans. Am J Clin Nutr 1989;49:828 31.
Hickey MW, Hillier AJ, Jago GR. Transport and metabolism of lactose,

255

256

102.
103.
104.
105.

106.
107.
108.
109.
110.

112.
113.
114.
115.
116.

117.
118.
119.
120.

121.
122.

123.

lactobacilli, bifidobacteria and inulin on the formation of aberrant crypt


foci in rats. Eur J Nutr 2001;40:293300.
Ganjam LS, Thornton WH, Marshall RT, MacDonald RS. Antiproliferative effects of yogurt fractions obtained by membrane dialysis on
cultured mammalian intestinal cells. J Dairy Sci 1997;80:23259.
Kim DH, Jin YH. Intestinal bacterial beta-glucuronidase activity of
patients with colon cancer. Arch Pharm Res 2001;24:564 7.
Reddy BS, Engle A, Simi B, Goldman M. Effect of dietary fiber on
colonic bacterial enzymes and bile acids in relation to colon cancer.
Gastroenterology 1992;102:1475 82.
Goldin BR, Gorbach SL. Alterations of the intestinal microflora by diet,
oral antibiotics, and Lactobacillus: decreased production of free amines
from aromatic nitro compounds, azo dyes, and glucuronides. J Natl
Cancer Inst 1984;73:689 95.
Goldin BR, Gorbach SL. The effect of milk and lactobacillus feeding on
human intestinal bacterial enzyme activity. Am J Clin Nutr 1984;39:
756 61.
Rafter JJ. The role of lactic acid bacteria in colon cancer prevention.
Scand J Gastroenterol 1995;30:497502.
Sartor RB. Pathogenesis and immune mechanisms of chronic inflammatory bowel diseases. Am J Gastroenterol 1997;92:5S11S.
Kwon HJ, Cote TR, Cuffe MS, Kramer JM, Braun MM. Case reports of
heart failure after therapy with a tumor necrosis factor antagonist. Ann
Intern Med 2003;138:80711.
Fabia R, ArRajab A, Johansson ML, et al. Impairment of bacterial flora
in human ulcerative colitis and experimental colitis in the rat. Digestion
1993;54:248 55.
McCormick DA, Horton LW, Mee AS. Mucin depletion in inflammatory bowel disease. J Clin Pathol 1990;43:143 6.
van Wijngaarden P, Meijssen MA. Tuberculous pleurisy: an unusual
complication during treatment of Crohn disease with azathioprine.
Scand J Gastroenterol 2001;37:1004 7.
Madsen KL, Doyle JS, Jewell LD, Tavernini MM, Fedorak RN. Lactobacillus species prevents colitis in interleukin 10 gene-deficient
mice. Gastroenterology 1999;116:110714.
Steidler L, Hans W, Schotte L, et al. Treatment of murine colitis by
Lactococcus lactis secreting interleukin-10. Science 2000;289:13525.
Malin M, Suomalainen H, Saxelin M, Isolauri E. Promotion of IgA
immune response in patients with Crohns disease by oral bacteriotherapy with Lactobacillus GG. Ann Nutr Metab 1996;40:137 45.
Kaila M, Isolauri E, Soppi E, Virtanen E, Laine S, Arvilommi H.
Enhancement of the circulating antibody secreting cell response in
human diarrhea by a human Lactobacillus strain. Pediatr Res 1992;32:
141 4.
Marshall BJ. Unidentified curved bacillus on gastric epithelium in
active chronic gastritis. Lancet 1983;1:12735.
Labigne A, de Reuse H. Determinants of Helicobacter pylori pathogenicity. Infect Agents Dis 1996;5:191202.
Duggan A. Helicobacter pylori: when is treatment now indicated? Int
Med J 2002;32:4659.
Aiba Y, Suzuki N, Kabir AMA, Takagi A, Koga Y. Lactic acidmediated supression of Helicobacter pylori by the oral administration
of Lactobacillus salivarius as a probiotic in a gnobiotic murine model.
Am J Gastroenterol 1998;93:2097101.
Midolo PD, Lambert JR, Hull R, Luo F, Grayson ML. In vitro inhibition
of Helicobacter pylori NCTC 11637 by organic acids and lactic acid
bacteria. J Appl Bacteriol 1995;79:4759.
Coconnier M, Lievin V, Hemery E, Servin AL. Antagonistic activity
against Helicobacter infection in vitro and in vivo by the human
Lactobacillus acidophilus strain LB. Appl Env Microbiol 1998;64:
4573 80.
Kim TS, Hur JW, Yu MA, et al. Antagonism of Helicobacter pylori by
bacteriocins of lactic acid bacteria. J Food Prot 2003;66:312.

124. Armuzzi A, Cremonini F, Ojetti V, et al. Effect of Lactobacillus GG


supplementation on antibiotic-associated gastrointestinal side effects
during Helicobacter pylori eradication therapy: a pilot study. Digestion
2001;63:17.
125. Michetti P, Dorta G, Wiesel PH, et al. Effect of whey-based culture
supernatant of Lactobacillus acidophilus (johnsonii) La1 on Helicobacter pylori infection in humans. Digestion 1999;60:2039.
126. Sakamoto I, Igarashi M, Kimura K, Takagi A, Miwa T, Koga Y. Suppressive effect of Lactobacillus gasseri OLL 2716 (LG21) on Helicobacter pylori infection in humns. J Antimicrob Chemother 2001;47:
709 10.
127. Cats A, Kuipers EJ, Bosschaert MAR, Pot RGJ, VandenbrouckeGrauls CMJE, Kusters JG. Effect of frequent consumption of a Lactobacillus casei-containing milk drink in Helicobacter pylori-colonized
subjects. Aliment Pharmacol Ther 2003;17:429 35.
128. Kallioma ki M, Isolauri E. Role of intestinal flora in the development of
allergy. Curr Opin Allergy Clin Immunol 2003;3:1520.
129. Kirjavainen PV, Salminen SJ, Isolauri E. Probiotic bacteria in the
management of atopic disease: underscoring the importance of viability. J Pediatr Gastroenterol Nutr 2003;36:2237.
130. Isolauri E. Studies on Lactobacillus GG in food hypersensitivity disorders. Nutr Today 1996;31:28S31S.
131. Gurr MI. The nutritional role of cultured dairy products. Can Inst Food
Sci Technol 1984;17:57 64.
132. Rytknen J, Karttunen TJ, Karttunen R, et al. Effect of heat denaturation on beta-lactoglobulin-induced gastrointestinal sensitization in
rats: denatured LG induces a more intensive local immunologic response than native LG. Pediatr Allergy Immunol 2002;13:269 77.
133. Bischoff SC, Mayer JH, Manns MP. Allergy and the gut. Int Arch
Allergy Immunol 2000;121:270 83.
134. Miettinen M, Matikainen S, Vuopio-Varkila J, et al. Lactobacilli and
streptococci induce interleukin-12 (IL-12), IL-18, and gamma interferon production in human peripheral blood mononuclear cells. Infect
Immun 1998;66:6058 62.
135. Shida K, Makino K, Morishita A, et al. Lactobacillus casei inhibits
antigen-induced IgE secretion through regulation of cytokine production in murine splenocyte cultures. Int Arch Allergy Immunol 1998;
115:278 87.
136. Gasser F. Safety of lactic acid bacteria and their occurrence in human
clinical infection. Bull Inst Pasteur 1994;92:45 67.
137. MacGregor G, Smith AJ, Thakker B, Kinsella J. Yoghurt biotherapy:
contraindicated in immunosuppressed patients? Postgrad Med J 2002;
78:366 7.
138. Schlegel L, Lemerle S, Geslin P. Lactobacillus species as opportunistic
pathogens in immune-compromised patients. Eur J Clin Microbiol
Infect Dis 1998;17:887 8.
139. Bayer AS, Chow AW, Betts D, Guze LB. Lactobacillemiareport of
nine cases. Important clinical and therapeutic considerations. Am J
Med 1978;64:808 13.
140. Horwitch CA, Furseth HA, Larson AM, Jones TL, Olliffe JF, Spach
DH. Lactobacillemia in three patients with AIDS. Clin Infect Dis 1995;
21:1460 2.
141. Husni RN, Gordon SM, Washington JA, Longworth DL. Lactobacillus
bacteremia and endocarditis: review of 45 cases. Clin Infect Dis 1997;
25:1048 55.
142. Mackay AD, Taylor MB, Kibbler CC, Hamilton-Miller JMT. Lactobacillus endocarditis caused by a probiotic organism. Clin Microbiol
Infect 1999;5:290 2.
143. Presterl E, Kneifel W, Mayer HK, Zehetgruber M, Makristathis A,
Graninger W. Endocarditis by Lactobacillus rhamnosus due to yogurt
ingestion? Scand J Infect Dis 2001;33:710 4.
144. Avlami A, Kordossis T, Vrizidis N, Sipsas NV. Lactobacillus rhamnosus endocarditis complicating colonoscopy. J Infect 2001;42:2835.

Downloaded from www.ajcn.org by guest on June 6, 2011

111.

ADOLFSSON ET AL

Вам также может понравиться