Hostname: page-component-7c8c6479df-5xszh Total loading time: 0 Render date: 2024-03-17T14:00:05.144Z Has data issue: false hasContentIssue false

Scientific evidence for health effects attributed to the consumption of probiotics and prebiotics: an update for current perspectives and future challenges

Published online by Cambridge University Press:  07 October 2015

Rafael Chacon Ruiz Martinez
Affiliation:
Department of Food and Experimental Nutrition, School of Pharmaceutical Sciences, University of São Paulo, Av. Prof. Lineu Prestes, 580, CEP 05508-000, São Paulo, Brazil
Raquel Bedani
Affiliation:
Department of Biochemical and Pharmaceutical Technology, School of Pharmaceutical Sciences, University of São Paulo, Av. Prof. Lineu Prestes, 580, CEP 05508-000, São Paulo, Brazil
Susana Marta Isay Saad*
Affiliation:
Department of Biochemical and Pharmaceutical Technology, School of Pharmaceutical Sciences, University of São Paulo, Av. Prof. Lineu Prestes, 580, CEP 05508-000, São Paulo, Brazil
*
*Corresponding author: S. M. I. Saad, fax +55 11 3815 6386, email susaad@usp.br
Rights & Permissions [Opens in a new window]

Abstract

Probiotics and prebiotics, mainly commercialised as food ingredients and also as supplements, are considered highly profitable niche markets. However, in recent years, the food industry has suffered from a series of health claim restrictions on probiotics and prebiotics in many parts of the world, including those made by the European Food Safety Authority. Therefore, we reviewed the core benefits of probiotic and prebiotic consumption on health. A number of studies have examined the prevention and/or management of intestinal infections, respiratory tract infections, CVD, osteoporosis, urogenital infections, cavities, periodontal disease and halitosis, allergic reactions, inflammatory bowel disease and irritable bowel syndrome and Helicobacter pylori gastric infections. In fact, a deeper understanding of the mechanisms involved in human microbiota and immune system modulation by probiotics and prebiotics relies on continuous efforts to establish suitable biomarkers of health and diseases risk factors for the design of clinical trials required for health claim approval. In spite of the promising results, the performance of large, long-term, well-planned, well-aligned clinical studies is crucial to provide more reliability and a more solid basis for the outcomes achieved and to support the potential use of probiotics and prebiotics in clinical practice.

Type
Full Papers
Copyright
Copyright © The Authors 2015 

The growing concern with food habits and their relation to health and longevity has stimulated the development of a large number of studies in the field of food science and nutrition. There has been considerable discussion on the role of the intestinal microbiota in the aetiology of a number of diseases and the effect of diet components on the modulation of the intestinal microbiota and its association with the reduced risk for illness development. As a result, in recent years, the concept of functional foods has been examining food additives that may exert beneficial effects on the composition and/or activity of the host intestinal microbiota, and an important class of functional foods has received considerable attention: probiotics and prebiotics.

Intestinal microbiota, a term used to replace the former name of intestinal microflora, is an ecosystem consisting of different ecological niches composed of a huge diversity of bacterial species and strains( Reference Laparra and Sanz 1 , Reference Aureli, Capurso and Castellazzi 2 ). This microbial population increases throughout the gastrointestinal tract (GIT) showing approximately 103 micro-organisms/ml of the luminal content in the duodenum, 108 micro-organisms/g of the ileal content and up to 1012 micro-organisms/g of the colonic content( Reference Ley, Lozupone and Hamady 3 , Reference Booijink, Zoetendal and Kleerebezem 4 ).

The wide diversity of intestinal microbiota has only recently been recognised because of the development and use of culture-independent molecular methods, which are based on the analysis of the 16S ribosomal RNA. These techniques have indicated that most bacteria in intestinal microbiota from healthy individuals belong to three main phyla: Firmicutes, Bacteroidetes and Actinobacteria( Reference Zoetendal, Rajilic-Stojanovic and De Vos 5 , Reference Tap, Mondot and Levenez 6 ). Each person presents a distinct and highly variable intestinal microbiota, at least at the species level; however, a stable core of intestinal colonists (intestinal microbiota-core) and of genes (microbiome-core) are shared by individuals and may be related to the intestinal function( Reference Phillips 7 Reference Zhang, Guo and Xue 11 ).

The intestinal microbiota is in direct contact with the intestinal mucosa. Both, along with the mucus, form the so-called mucosal barrier, an important defence system against potentially pathogenic and immunogenic factors present in the lumen. In fact, the mucous membrane separates the lumen containing the microbiota, organic food waste and secretions (salivary, gastric, biliary, pancreatic and intestinal) from the lymphoid tissue associated with the intestines( Reference Aureli, Capurso and Castellazzi 2 ). The cells that make up the immune system are mainly concentrated in the lymphatic organs located in the lamina propria of the GIT. The lymphoid tissue associated with the intestine is composed of several follicular structures, Peyer’s patches, T lymphocyte aggregates, antigen-presenting cells and B lymphocytes( Reference Aureli, Capurso and Castellazzi 2 ).

The high metabolic activity of the intestinal microbiota, besides its nutritional role, results in a significant impact on the individual’s health and well-being( Reference Tremaroli and Bäckhed 10 ). The interaction between the intestinal microbiota and the host generates several advantages for both. The main functions of the intestinal microbiota include participation in the intestinal wall formation; colonisation resistance against pathogens; production of SCFA; butyrate, propionate and acetate; production of vitamins, especially vitamin B and vitamin K complex; interactions with the mucosal immune system; and degradation of xenobiotics( Reference Qin, Li and Raes 9 , Reference Hill 12 Reference Saulnier, Kolida and Gibson 16 ).

According to Round & Mazmanian( Reference Round and Mazmanian 17 ), a healthy GIT microbiota contains a balanced composition of several classes of bacteria, including the symbionts known as health-promoting micro-organisms such as Bifidobacterium spp. and Lactobacillus spp., the commensals, micro-organisms that allegedly provide neither benefits nor harm to the host, and pathobionts, potentially pathogenic micro-organisms.

Fluctuations in the composition of the intestinal ecosystem have been associated with various diseases, including imunnoinflammatory disorders, obesity and cancer( Reference Scanlan, Shanahan and O’Mahony 18 Reference Caesar, Fak and Bäckhed 21 ). The composition and metabolic activity of the intestinal microbiota can be modified by various factors, including antibiotic treatment, inflammatory processes, ageing, diet changes, GIT motility, host secretions, among others( Reference Laparra and Sanz 1 , Reference Gerritsen, Smidt and Rijkers 22 ). Changes in the diet, including the amount, type and balance of the main dietary macronutrients (carbohydrates, proteins and fats), may significantly affect the intestinal microbiota diversity, which may influence its functional relation with the host( Reference Ley, Peterson and Gordon 23 Reference Scott, Gratz and Sheridan 25 ).

Thus, there is a growing interest in alternatives that can beneficially modulate the intestinal microbiota, improving the individual’s health and consequently reducing the risk for the development of diseases. The consumption of probiotic micro-organisms and prebiotic ingredients is a promising alternative to beneficially influence the intestinal microbial ecology, maintaining the intestinal homoeostasis and controlling the dysbiosis, and, consequently, improving health( Reference O’Shea, Cotter and Stanton 26 , Reference Peterson, Sharma and Elmén 27 ). Therefore, intestinal microbiota is an important target for interventions with probiotics and prebiotics, administered as supplements or food ingredients, with the specific goal of modulating the microbial community composition, as well as the microbiome functional capacity( Reference McNulty, Yatsunenko and Hsiao 28 , Reference Rauch and Lynch 29 ). In this review paper, the key health benefits attributed to probiotic micro-organisms and prebiotic fibre consumption will be discussed.

Probiotics

Several definitions of probiotics have been proposed over the years. However, a consensus statement on the scope and appropriate use of the term probiotic was recently given by the International Scientific Association for Probiotics and Prebiotics( Reference Hill, Guarner and Reid 30 ): probiotics are defined as ‘live micro-organisms that, when administered in adequate amounts, confer a health benefit on the host’.

The vast majority of potentially probiotic lactic acid bacteria (LAB) belongs to the phylum Firmicutes, a very diverse group of bacteria with low G+C genomic contents and which includes the Aerococcus, Enterococcus, Lactobacillus, Lactococcus, Leuconostoc, Oenococcus, Pediococcus, Streptococcus, Carnobacterium, Tetragenococcus, Vagococcus and Weissella genera( Reference Stolaki, De Vos and Kleerebezem 31 ). The Bifidobacterium genus is considered by many scientists to be a member of the LAB group, as it shares some typical characteristics of this group, such as the production of lactic acid. However, this genus belongs to the phylum Actinobacteria, a group of bacteria with high G+C genomic content and a distinct carbohydrate fermentation system when compared with the LAB belonging to the phylum Firmicutes( Reference Stolaki, De Vos and Kleerebezem 31 ). Although Bifidobacterium and Lactobacillus genera do not belong to the same taxonomic group, in this review paper, merely for didactic purposes, the genus Bifidobacterium will be included when the LAB group is mentioned in the text.

For a micro-organism to be considered as a probiotic, certain criteria must be fulfilled, including the following: (1) even though certain commercially available probiotic strains are not of human origin, it is believed that if a probiotic is isolated from the GIT of human beings it is safe for human consumption and can be more effective within the intestinal ecosystem; (2) probiotic cultures should be recognised as safe (GRAS status – generally recognised as safe) for human consumption through scientific evidence or experiments based on the history of consumption by a significant number of subjects: bacteria of the genera Bifidobacterium spp. and Lactobacillus spp. have a long history of safe consumption without any reported harmful health effects; (3) the preparation of large-scale probiotics should be feasible, and it is very important that these micro-organisms are viable and active in the vehicles in which they are incorporated; (4) probiotics should be resistant to gastric and intestinal juices, as low pH is one of the host defence mechanisms against ingested micro-organisms, including probiotics; (5) probiotics should adhere to human intestinal cells and intestinal mucins, which improve their persistence and allow their growth in the intestine, and may favour the competitive exclusion of potential pathogens of the mucosal surface; (6) probiotics should produce antimicrobial substances against intestinal pathogens in order to restore the healthy microbiota composition; (7) probiotics must be safe when ingested through food consumption and during clinical use, even for immunocompromised individuals; (8) probiotics must have their safety and efficacy established through randomised and placebo-controlled clinical trials( Reference Tannock 32 , Reference Kolida and Gibson 33 ).

The sufficient dose of probiotic micro-organisms to lead to beneficial health effects may vary depending on the strain and the product. In general, products containing probiotic micro-organisms should have a minimum number of viable cells, with proven efficacy established based on human clinical trials, estimated to be between 106 and 108 colony-forming units per gram (CFU/g) of end product or 108–1010 CFU/d (considering 100 g or 100 ml of the ingested food)( Reference Champagne, Ross and Saarela 34 ). Brazilian legislation recommends a minimum probiotic population ranging from 108 up to 109 CFU/daily serving portion of the food product to obtain a beneficial health effect in the gut( 35 ). A similar number of viable probiotic cells (109 CFU) per serving portion, consumed on a daily basis, is also recommended by Health Canada( 36 ) and the Italian Health Ministry( 37 ).

The ideal amount of probiotic micro-organisms to be administered is not easy to be determined. It is believed that it is strain specific and depends on the type of the intended beneficial effect, and thus different effects may require different strains and different probiotic quantities. Of course, the total probiotic microbial population may not be low if the goal is to influence the composition and/or the metabolic activity of the host microbiota( Reference Aureli, Capurso and Castellazzi 2 ). Other factors may be involved in determining the daily probiotic dose, including the daily administration frequency (one to four times); the administration period (before, during or after meals); the duration of administration (from 1 d to several months); the vehicle used for the probiotic delivery (fermented food, drink, capsule, tablet or powder); and the viability of the probiotic strain( Reference Lee and Salminen 38 ).

Prebiotics

Prebiotics are defined as ‘selectively fermentable ingredients that allow specific changes in the composition and/or activity of gastrointestinal microbiota that allow benefits to the host’( Reference Gibson, Probert and Van Loo 39 , Reference Gibson, Scott and Rastall 40 ).

On the basis of this definition, the requirements that a dietary ingredient must meet in order to be characterised as a prebiotic include the following. (1) The fermentability should be demonstrated in in vitro tests that simulate, for example, physiological conditions found in the GIT. Promising substrates should be evaluated in randomised and placebo-controlled clinical studies, in order to confirm the positive outcomes obtained by in vitro studies. (2) The main trait of a prebiotic is to be a selective substrate for one or more beneficial GIT commensal bacteria, which are stimulated to multiply and/or are metabolically activated, beneficially altering the colonic microbiota composition of the host. To confirm the selectivity of a prebiotic, it is of great importance to monitor the changes in the faecal microbiota during supplementation studies with the prebiotic through in vitro and in vivo tests. Although both criteria are essential, selectivity is the most important, as well as the most difficult, to be achieved( Reference Tannock 32 , Reference Gibson, Probert and Van Loo 39 ). Moreover, selectivity consists of a key attribute that distinguishes prebiotics from other dietary fibres( Reference Rastall 41 ).

Although nondigestibility has been excluded from the latest definition of prebiotics, these fibres should not be digestible by human enzymes or be only partially digestible in order to reach, in adequate amounts, more distant segments of the GIT( Reference Kolida and Gibson 33 ). Nowadays, the main well-known prebiotics are non-digestible carbohydrates including fructo-oligosaccharides (FOS), inulin, galacto-oligosaccharides (GOS), trans-galacto-oligosaccharides (TOS) and lactulose. Other non-digestible carbohydrates have been studied for their prebiotic potential including soya bean oligosaccharides, isomalto-oligosaccharides, xylo-oligosaccharides, polydextrose, glucans, cereal-derived arabinoxylans and arabinoxylan oligosaccharides( Reference Gibson, Scott and Rastall 40 , Reference Roberfroid 42 , Reference Broekaert, Courtin and Verbeke 43 ). However, most of the data available in the scientific literature on prebiotic effects are related to inulin and FOS.

Several studies have evaluated the potential of resistant starch (RS), a high-amylose starch, to act as a prebiotic ingredient( Reference Gibson, Scott and Rastall 40 ). A study developed by Crittenden et al.( Reference Crittenden, Morris and Harvey 44 ) reported that various Bifidobacterium strains, including B. adolescentis, B. bifidum, B. breve, B. infantis, B. lactis and B. longum, were able to hydrolyse RS. Although the evidence of RS as a prebiotic compound is somewhat limited, as most studies have been performed using animal models, many of the beneficial effects in the large bowel appear to be caused by SCFA formed by bacterial fermentation( Reference Topping, Fukushima and Bird 45 , Reference Figueroa-González, Quijano and Ramírez 46 ). According to available data, human colonic bacteria may ferment RS to SCFA, mainly acetate, propionate and butyrate( Reference Topping and Clifton 47 ). Nevertheless, further research regarding the potential of RS as a prebiotic compound in human is still necessary, especially regarding the ability to selectively stimulate beneficial micro-organisms.

Prebiotics are available and may be extracted from plant sources. However, most of the prebiotic fibres used as food ingredients are synthesised commercially through enzymatic or chemical methods( Reference Crittenden and Playne 48 ).

As in the case of probiotics, in order to ensure a continuous effect, prebiotics should be consumed daily. Favourable changes in the intestinal microbiota were observed at doses of 4–20 g/d of inulin and/or FOS( Reference Gibson, Scott and Rastall 40 , Reference Rastall 41 ). The daily dose of prebiotics (inulin, FOS and lactulose) per serving portion of the food product recommended by the Brazilian legislation is of 3 g for solid foods and of 1·5 g for liquid foods( 35 ).

The molecular structure of prebiotics is important to determine the physiological effects and also which species of micro-organisms will be able to use them as C and energy sources in the intestine. However, despite the diversity of molecular weights, compositions of sugars and structural connections within the range of prebiotic ingredients, bifidobacteria are the micro-organisms mostly involved in this response. The mechanisms by which prebiotics promote specific growth of bifidobacteria in the intestinal microbiota are still not clear. However, several hypotheses ought to be mentioned, including the following: (1) bifidobacteria may use a wide variety of oligosaccharides and complex carbohydrates as C and energy sources; (2) in the presence of several non-digestible oligosaccharides, bifidobacteria have higher growth rates, when compared with putrefactive or potentially pathogenic bacteria in the intestine environment; (3) although other genera of bacteria (Lactobacillus, Bacteroides and Eubacteria) are able to multiply in vitro using prebiotic sources, bifidobacteria seem to do it in a more efficient way. Furthermore, bifidobacteria are tolerant to SCFA and to the acidification of the intestinal environment. Bifidobacteria produce permeases that are able to internalise non-digestible oligosaccharides, which are then metabolised, thus minimising the release of simple sugars that could be consumed by other intestinal bacteria( Reference Crittenden and Playne 48 ).

Synbiotics

A synbiotic product combines one or more probiotic micro-organism with a prebiotic fibre. These food ingredients together in a product may lead to the previous adaptation of the probiotic to the prebiotic substrate, which might promote a positive interaction between the probiotic and the prebiotic in vivo. In some cases, this may lead to a competitive advantage for the probiotic, if consumed along with the prebiotic fibre. Alternatively, this synbiotic effect can be directed to different ‘target’ regions of the small and large intestines. The consumption of appropriately selected probiotics and prebiotics may increase the beneficial effects of both synergistically, as the stimulus of known probiotic strains leads to the choice of the most profitable combination between substrate and micro-organism( Reference Holzapfel and Schillinger 49 Reference Bielecka, Biedrzycka and Majkowska 52 ).

Both approaches may directly or indirectly be in accordance with the definition of synbiotics. However, according to Kolida & Gibson( Reference Kolida and Gibson 33 ), the synergistic approach tends to be the most important.

The synbiotic concept offers great potential to increase the effectiveness of this class of functional foods, as it explores the advantages that a combination of prebiotics and probiotics may offer, not only to health but also to the stability of the product, during its storage period( Reference Kolida and Gibson 33 , Reference Sanders and Marco 53 ).

Health effects associated with probiotic and prebiotic consumption

Intestinal infections

Most of the studies on the clinical use of probiotics are focused on the GIT. In this site, it is believed that probiotics are able to compete with pathogenic micro-organisms for adhesion sites and nutrients. Besides, they may produce different antimicrobial compounds, a process called ‘colonisation resistance’ or ‘competitive exclusion’( Reference Rolfe 54 , Reference Saad 55 ). In terms of prebiotics, their main characteristics are resistance to digestive enzymes in the human gut, fermentability by the colonic microbiota and bifidogenic and pH-lowering effects. Accordingly, because of the last characteristic, prebiotics could inhibit certain strains of potentially pathogenic bacteria, particularly Clostridium, and prevent diarrhoea( Reference de Vrese and Marteau 56 ).

Intestinal infections are characterised by an imbalance of the normal intestinal microbiota, which leads to increased pathogenic micro-organism populations. To determine the efficacy of probiotic strains in the treatment of these infections, a number of clinical studies have been performed( Reference Saad, Delattre and Urdaci 57 ). Some studies showed that different probiotic strains administered in children at the beginning of the diarrhoea episode were able to reduce the infection duration and/or intensity( Reference Floch, Walker and Madsen 58 , Reference Guandalini 59 ). In Europe, the use of probiotics for acute infectious diarrhoea in children is an accepted therapy( Reference Wolvers, Antoine and Myllyluoma 60 ). Table 1 shows some of the publications cited in this review, which demonstrated beneficial effects of the consumption of probiotics and/or prebiotics against intestinal infections and other pathological conditions.

Table 1 Beneficial effects of probiotics and/or prebiotics on important human pathologic conditions

FOS, fructo-oligosaccharides; GOS: galacto-oligosaccharides; TOS, trans-galacto-oliogosaccharides; scGOS, short-chain galacto-oligosaccharides; lcFOS, long-chain fructo-oligosaccharides; XOS, xylo-oligosaccharides; VSL#3: L. casei, L. plantarum, L. acidophilus, L. delbrueckii spp. bulgaricus, B. longum, B. breve, B. infantis and S. thermophilus.

The results reported for probiotic effects against antibiotic-associated diarrhoea are rather heterogeneous and sometimes even contradictory. Studies conducted by Arvola et al.( Reference Arvola, Laiho and Torkkeli 61 ) and Vanderhoof et al.( Reference Vanderhoof, Whitney and Antonson 62 ) demonstrated that Lactobacillus GG was effective in reducing the adverse effects and diarrhoea commonly associated with the use of antibiotics. However, according to Marchand & Vandenplas( Reference Marchand and Vandenplas 63 ), several papers reported that some probiotic micro-organisms did not present any clinical efficacy in the treatment of this condition. A systematic review of probiotic effectiveness also indicated that findings reported in the scientific literature do not support the use of probiotics for Clostridium difficile infection( Reference Dendukuri, Costa and McGregor 64 ). In contrast to probiotics, there are few clinical trials on prebiotic effects in preventing antibiotic-associated diarrhoea( Reference Roberfroid, Gibson and Hoyles 65 ).

According to Marchand & Vandenplas( Reference Marchand and Vandenplas 63 ), at least three clinical studies have demonstrated the potential of the yeast Saccharomyces boulardii in reducing antibiotic-associated diarrhoea. Thompson( Reference Thompson 66 ) also reported the beneficial effects of S. boulardii on Clostridium difficile-associated diarrhoea (CDAD). Moreover, according to a meta-analysis, probiotic lactobacilli may effectively prevent antibiotic-associated diarrhoea both in children and in elderly people( Reference Kale-Pradhan and Jassal 67 ). In the same way, Hickson et al.( Reference Hickson, D’Souza and Muthu 68 ) observed that a probiotic drink containing Lactobacillus casei DN-114 001, Lactobacillus bulgaricus and Streptococcus thermophilus lowered the risk for developing antibiotic and CDAD among seniors in 22 and 17 % cases, respectively. On the other hand, according to a study conducted by Pozzoni et al.( Reference Pozzoni, Riva and Bellatore 69 ), in which 564 hospitalised patients were followed up for 3 months, the probiotic micro-organism did not prevent antibiotic-associated diarrhoea caused by C. difficile.

Although studies provide evidence that selected probiotics may significantly decrease the risk of diarrhoea in subjects treated with antibiotics, not all antimicrobial agents are equal in causing antibiotic-associated diarrhoea. Therefore, conclusions on the efficacy of probiotics in preventing diarrhoea caused by any particular antibiotic class may not be made( Reference Wolvers, Antoine and Myllyluoma 60 ).

A review published by Floch et al.( Reference Floch, Walker and Madsen 58 ) indicated that probiotics are helpful in the prevention of CDAD in both adults and children, especially when Lactobacillus rhamnosus GG and S. boulardii are combined. Several placebo-controlled studies have suggested that Lactobacillus GG could be used to treat several forms of diarrhoea, including rotavirus diarrhoea, travellers’ diarrhoea and relapsing C. difficile diarrhoea( Reference Gorbach 70 , Reference Goldin and Gorbach 71 ). On the other hand, according to Graul et al.( Reference Graul, Cain and Karpa 72 ), two trials using L. rhamnosus GG for the prevention of CDAD failed to demonstrate that this supplement could reduce the incidence of this disease( Reference Arvola, Laiho and Torkkeli 61 , Reference Thomas, Litin and Osmon 73 ).

In this scenario, although probiotics may be effective in the prevention of infection caused by C. difficile, so far there is no sufficient scientific evidence to unambiguously prove the effectiveness of this approach( Reference Floch, Walker and Madsen 58 , Reference Balakrishnan and Floch 74 ). However, it is important to stress out that the reduction of CDAD risk by reducing the presence of the pathogen toxin is indeed considered a beneficial physiological effect. In reality, the European Food Safety Authority (EFSA)( 75 ) concluded that there is insufficient evidence to establish a cause and effect relationship between the consumption of Actimel® (a fermented milk containing L. casei DN-114 001 plus yogurt bacteria; Danone) and the reduction of risk for developing C. difficile diarrhoea in patients receiving antibiotics by reducing the presence of C. difficile toxins. On the other hand, a randomised and controlled study conducted by Lewis et al.( Reference Lewis, Burmeister and Brazier 76 ) demonstrated that the consumption of the prebiotic oligofructose (with a daily ingestion of 12 g) presented a positive effect in the treatment of CDAD.

Rotavirus is the leading cause of severe diarrhoea in children worldwide( Reference Parashar, Gibson and Bresse 77 ). The organism invades the epithelial cells of the small intestine and multiplies, leading to the destruction of the intestinal mucosa and an increased intestinal permeability( Reference Salim, Phillips and Farthing 78 ). Several studies have shown that the probiotic strains L. rhamnosus GG, Lactobacillus reuteri, L. casei Shirota and Bifidobacterium animalis spp. lactis Bb-12 were able to reduce the duration of rotavirus diarrhoea in about 1 d( Reference Kaila, Isolauri and Soppi 79 Reference Shornnikova, Casas and Mykkanen 82 ). The effectiveness of certain probiotic strains in the treatment/prevention of infection is closely related to both strain and dose of the micro-organism used( Reference Saad, Delattre and Urdaci 57 , Reference Graul, Cain and Karpa 72 , Reference van Baarlen, Troost and Van der Meer 83 ).

Probiotics may also be helpful in the prevention of travellers’ diarrhoea, a condition commonly observed in people travelling to countries with lower economic and social development and warmer climates( Reference Goldin and Gorbach 71 ). Hilton et al.( Reference Hilton, Kolakowski and Singer 84 ) showed that Lactobacillus GG was able to prevent this infection. However, Oksanen et al.( Reference Oksanen, Salminen and Saxelin 85 ) reported limited effect with the use of the same strain. In general, the clinical evidence for travellers’ diarrhoea prevention using probiotic micro-organisms is still limited( Reference Lyra, Lahtinen and Ouwehand 86 ), in particular because of the reduced number of interventions and variability of the experimental protocols used( Reference McFarland 87 ). The wide variety of the potential causes of traveller’s diarrhoea and the difficulty of volunteers to follow study protocols during their trips bring additional challenges for the intervention with probiotics when targeting this type of illness( Reference Lyra, Lahtinen and Ouwehand 86 ).

The role of prebiotics in the prevention of travellers’ diarrhoea has also been investigated through randomised, double-blind, placebo-controlled studies, and results are also inconsistent. According to Drakoularakou et al.( Reference Drakoularakou, Tzortis and Rastall 88 ), a GOS mixture (B-GOS) showed significant potential to prevent the incidence and symptoms of travellers’ diarrhoea in subjects who travelled for at least 2 weeks to a country of low or high risk for the condition. However, Cummings et al.( Reference Cummings, Christie and Cole 89 ) demonstrated that the consumption of 10 g of FOS daily, before and during holiday to medium- and high-risk destinations for travellers’ diarrhoea, was not effective in preventing it. Moreover, Virk et al.( Reference Virk, Mandrekar and Berbari 90 ) observed that the use of an oral synbiotic product containing Enterococcus faecium SF68, Saccharomyces cerevisae CNCM I 4444 and FOS, although safe, was ineffective in preventing travellers’ diarrhoea among people who visited areas with increased risk for the condition and did not decrease its duration or the need for antibiotics when it actually occurred.

Even though some studies have shown promising findings regarding prebiotic effects on intestinal microbiota, there is not enough evidence to recommend prebiotics for the prevention or treatment of diarrhoea. In general, it is important to highlight that there are many different causes of diarrhoea, and the studies using probiotic strains and prebiotic ingredients on the same type of diarrhoea may use different criteria in the selection of volunteers and various outcome parameters, which makes it difficult to draw conclusive remarks about the beneficial effects of these bioactive compounds on this pathology. It seems that probiotics are more effective in the prevention of rotavirus and antibiotic-associated diarrhoea than travellers’ diarrhoea( Reference Lyra, Lahtinen and Ouwehand 86 ). Nevertheless, further studies about the impact of probiotics and prebiotics on the prevention and the management of diarrhoea are mandatory and could propel (or not) their use in medical settings.

Respiratory tract infections

Respiratory tract infections victimise a high proportion of individuals and are related to high morbidity and mortality rates( Reference Vouloumanou, Makris and Karageorgopoulos 91 ). This condition, common among children, is normally self-limiting, and the risk for the development of complications is considered small( Reference Thompson, Vodicka and Blair 92 ). Antibiotics are commonly used in an inappropriate manner for the treatment of the pathology and may therefore select resistant micro-organisms, an increasing global phenomenon( Reference Snow, Mottur-Pilson and Gonzales 93 , Reference Aguilar, Giménez and Barberán 94 ).

Initially, several studies carried out with animal models evidenced a potentially protective effect of probiotics against bacterial and viral respiratory tract infections. This positive outcome was attributed to the immune system stimulation( Reference Cangemi de Gutierrez, Santos and Nader-Macias 95 Reference Racedo, Villena and Medina 97 ).

Vouloumanou et al.( Reference Vouloumanou, Makris and Karageorgopoulos 91 ) published a systematic review on fourteen randomised controlled trials that evaluated the ability of different probiotics (Lactobacillus and Bifidobacterium strains – alone and combined, and a non-pathogenic Enterococcus faecalis strain) for the prevention of respiratory tract infections. The authors concluded that the probiotic strains tested may possibly have a positive effect on the severity and extent of the symptoms, although they may not be useful in reducing the incidence of the illness. More recently, King et al.( Reference King, Glanville and Sanders 98 ) published a systematic review that investigated the effect of probiotics, particularly Lactobacillus and Bifidobacterium strains, on acute respiratory infections in otherwise healthy children and adults. The authors included twenty-one randomised controlled trials – of which twelve were considered to have a low risk of bias – and observed that subjects who received probiotics had reduced number of days of disorder and less absenteeism (days away from day care/school/work), in comparison with patients in the placebo group.

Arslanoglu et al.( Reference Arslanoglu, Moro and Boehm 99 ) conducted a randomised, double-blind, placebo-controlled trial in which healthy infants consumed a prebiotic formula containing FOS and GOS or placebo (maltodextrin) during the first 6 months of life. The authors reported significantly fewer episodes of all types of infections combined, as well as a tendency to have fewer upper respiratory tract infection episodes in the probiotic group, in comparison with the placebo group. Arslanoglu et al.( Reference Arslanoglu, Moro and Boehm 99 ) also pointed out that the cumulative incidence of any recurring respiratory infections was significantly lower in infants who received prebiotics, compared with infants fed placebo. Even though the authors did not clarify the specific mechanism(s) of action involved in the positive outcomes observed, they hypothesised that changes in the intestinal microbiota could have had an important role.

Luoto et al.( Reference Luoto, Ruuskanen and Waris 100 ) investigated the impact of prebiotic and probiotic supplementation for the prevention of rhinovirus infections in preterm infants. The researchers carried out a randomised, double-blind, placebo-controlled trial including ninety-four preterm infants, who received oral prebiotics (GOS and polydextrose mixture, 1:1), a probiotic (L. rhamnosus GG, ATCC 53103) or placebo (microcrystalline cellulose) from 3 up to 60 d of life. Follow-up visits were performed at the ages of 1, 2, 4, 6 and 12 months, whereas an additional telephone call to the parents occurred at the age of 9 months. Throughout the study, the authors observed a significantly lower incidence of respiratory tract infections, including rhinovirus infections, in the prebiotic and probiotic groups, compared with infants fed placebo. However, these researchers also reported that neither the probiotic nor the prebiotic interventions had any influence on the duration and severity of the symptomatic rhinovirus infections.

In another study, Puccio et al.( Reference Puccio, Cajozzo and Meli 101 ) evaluated 187 infants who were not breast-fed after 14 d of life and randomly received an experimental formula containing the probiotic strain B. longum BL99 and a prebiotic mixture of FOS and GOS or a control formula. The authors observed equivalent mean weight gain in both groups, as well as no statistical difference in the incidence of respiratory tract infections.

Although certain published studies have shown encouraging results, there are still some important gaps in our knowledge of the impact of probiotics and/or prebiotics on respiratory tract infections. In general, the research in this field of application has shown heterogeneous findings regarding, for example, probiotic strains, types of prebiotic ingredients, doses, mode and period of administration. Another aspect that could be better explored includes the use of symbiotic formulations to enhance the potential beneficial effects on respiratory tract infections. Thus, more clinical trials are essential to undeniably attest the efficacy of probiotics and prebiotics in the prevention and/or treatment of respiratory tract infections, and more in vitro studies are required to elucidate the mechanisms of action involved.

CVD

CVD are the leading cause of death in the world, and their incidence has been increasing in recent years( 102 ). Among the different risk factors involved in their pathogenesis, hypercholesterolaemia is the main factor( Reference Grundy, Cleeman and Merz 103 ). To reduce the incidence of CVD, it is necessary to decrease cholesterol levels in hypercholesterolaemic subjects, which may be achieved using both pharmacological and non-pharmacological interventions. Many patients prefer non-medicamental treatments for the reduction of the blood cholesterol rate because of the frequent adverse effects related to the use of lipid-lowering drugs, medicine contraindications or the personal preference for natural or alternative therapies( Reference Houston, Fazio and Chilton 104 ).

Mann & Spoerry( Reference Mann and Spoerry 105 ) were the first researchers to observe the hypocholesterolaemic activity of fermented milk in a Maasai tribe, located in Kenya. Since then, many scientists have used animal and human models to evaluate the effects of probiotic micro-organisms on serum cholesterol level, and the probiotic benefits have been emphasised over the last 40 years.

Hepatic biosynthesis and the intestinal absorption are two sources of cholesterol in the human body, and both have an important role in the overall balance. On the basis of in vitro and in vivo studies, certain mechanisms have been proposed to explain the inhibition of cholesterol absorption in the small intestine, namely (i) linkage and incorporation of cholesterol by the bacterial cells; (ii) the suppression of bile acid reabsorption, mediated by the bile acid hydrolase of bacterial origin, which catalyses the deconjugation of bile acid salts, releasing free primary bile acids excreted in faeces in higher quantities; (iii) co-precipitation of cholesterol with deconjugated bile salts; (iv) conversion of cholesterol into coprostanol; and (v) production of SCFA by probiotics( Reference Ooi and Liong 106 , Reference Zhuang, Liu and Zhang 107 ).

More recently, studies have shown that several compounds implicated in cholesterol metabolism may be involved in the cholesterol-reducing ability of probiotic micro-organisms; however, this association is not fully understood( Reference Kumar, Nagpal and Kumar 108 ). These compounds include the protein NPC1L1( Reference Davis, Zhu and Hoos 109 ), and the enzymes 3-hydroxy-3-methyl-glutaryl-CoA reductase( Reference Jeun, Kim and Cho 110 ) and 7α, 27α hydrolase( Reference Park, Kim and Shin 111 ).

Agerholm-Larsen et al.( Reference Agerholm-Larsen, Raben and Haulrik 112 ) conducted a double-blind placebo-controlled study to evaluate the effect of a milk product (GAIO®; MD Foods) containing the CAUSIDO® probiotic cultures (one strain of Enterococcus faecium and two strains of S. thermophilus) and two alternative products on the risk factors for the development of CVD in overweight and obese patients. The researchers observed that after 8 weeks of consuming the product the CAUSIDO® cultures reduced LDL-cholesterol levels and increased fibrinogen rates in overweight patients. Although high levels of fibrinogen are considered a risk factor for the development of CVD, the authors observed that its levels remained within normal ranges (4·5–10 3 µmol/l). Naruszewicz et al.( Reference Naruszewicz, Johansson and Zapolska-Downar 113 ) performed an interesting study that evaluated a group of smokers who ingested daily, for 6 weeks, a drink containing the probiotic Lactobacillus plantarum 299v strain. At the end of this period, the authors observed that the probiotic group showed a reduction in the systolic blood pressure and in the concentrations of leptin and fibrinogen, when compared with the control group. Ataie-Jafari et al.( Reference Ataie-Jafari, Larijani and Majd 114 ) evaluated a group of people with hypercholesterolaemia and reported that after 6 weeks of consumption of a yogurt containing Lactobacillus acidophilus and B. lactis the blood cholesterol rates were significantly lowered, compared with the group that consumed traditional yogurt. Moreover, Jones et al.( Reference Jones, Martoni and Parent 115 ) demonstrated that the consumption of a yogurt containing microencapsulated bile salt hydrolase-active L. reuteri NCIMB 30242, taken twice during a 6-week period, was effective and safe for the reduction of LDL-cholesterol, total cholesterol and non-HDL-cholesterol in hipercholesterolaemic adults.

Bedani et al.( Reference Bedani, Rossi and Cavallini 116 ) evaluated the effect of a synbiotic soya-based product supplemented with okara soya bean by-product on the risk factors for CVD. The authors conducted a randomised, double-blind placebo-controlled trial, in which thirty-six normocholesterolaemic men were assigned into two groups: eighteen subjects consumed, on a daily basis, 100 g of the synbiotic food product fermented with L. acidophilus La-5, B. animalis Bb-12 and S. thermophilus (starter culture), whereas eighteen subjects consumed daily 100 g of unfermented soya-based product (placebo group), both for 8 weeks. Fasting blood samples and anthropometric measurements were obtained at the baseline (T0) and after 4 (T4) and 8 (T8) weeks of the food product consumption. The authors observed a significant reduction in LDL-cholesterol mean values and a significant improvement of the LDL-cholesterol:HDL-cholesterol ratio in the synbiotic group. Furthermore, a trend for a LDL-cholesterol and LDL-cholesterol:HDL-cholesterol ratio reduction (obtained as the mean differences between T8 and T0) was higher in the synbiotic group, when compared with the placebo group.

On the other hand, some studies did not find any significant effects regarding serum cholesterol reduction with the consumption of probiotic cultures( Reference de Roos, Schouten and Katan 117 , Reference Simons, Armansec and Conway 118 ).

Some researchers have investigated the effects of prebiotics on cholesterol levels, but the results were not homogeneous. Balcázar-Muñoz et al.( Reference Balcázar-Muñoz, Martínez-Abundis and Gonzáles-Ortiz 119 ) reported that the consumption of inulin (7 g/d) during 4 weeks by obese and hypercholesterolaemic subjects led to a significant reduction in total cholesterol, LDL-cholesterol, VLDL-cholesterol and TAG levels, when compared with subjects who received placebo. An interesting study published by Nichenametla et al.( Reference Nichenametla, Weidauer and Wey 120 ) evaluated individuals diagnosed with or without the metabolic syndrome (MetS) – according to the International Diabetes Federation criteria – who were enrolled in a double-blind, placebo-controlled, cluster crossover intervention. The authors studied the effects of exchanging enriched flour (containing RS type 4) for regular/control flour in different comorbidities associated with the MetS. The authors observed that the prebiotic consumption improved dyslipidaemia and body composition. On the other hand, Giacco et al.( Reference Giacco, Clemente and Luongo 121 ) reported that the daily intake of 10·6 g of short-chain FOS for 2 months by mild hypercholesterolaemic individuals had no major effects on lipid metabolism, when compared with consumption of placebo. However, Kellow et al.( Reference Kellow, Coughlan and Reid 122 ) emphasised that the conclusions of these studies were limited, as they evaluated relatively short-term prebiotic interventions periods, and longer periods would be required to draw stronger conclusions.

Other valuable studies on the role of prebiotics regarding the prevention of CVD using animal models have also been published. Rault-Nania et al.( Reference Rault-Nania, Gueux and Demougeot 123 ) demonstrated, through studies conducted with mice, that the addition of long-chain inulins in the animals’ diet was able to reduce the formation of atherosclerotic plaques. Ranganna et al.( Reference Ranganna, Yatsu and Hayes 124 ) reported that the SCFA obtained from the fermentation of dietary fibres may modulate the expression of multiple genes involved in the atherosclerosis process.

Although several studies support the hypothesis that probiotics and/or prebiotics may reduce the risk of CVD, other researchers reported controversial findings. The discrepancies found among studies could be explained by the lack of dose–response studies in order to determine the ‘minimum effective dose’ required to promote, for example, an improvement of the lipid profile in humans( Reference Ooi and Liong 106 ). In addition, although several mechanisms of probiotic and prebiotic actions have been proposed, most of them are based on in vitro tests. In this sense, new in vivo studies are needed so that these mechanisms can be properly clarified.

In recent years, evidence has pointed to the possible relationship between gut microbiota and CVD, particularly atherosclerosis( Reference Tang and Hazen 125 ). However, the understanding of this association is still limited. Keeping this in mind, a potential beneficial modulation of intestinal microbiota in a specific manner using probiotics and/or prebiotics might have a cardioprotective role in the host. In this sense, the gut microbiota could represent a new target for the treatment and prevention of CVD, and studies in this field of knowledge may be increasingly promising.

Osteoporosis

Osteoporosis is a disease characterised by bone mass insufficiency and deterioration of the structural bone tissue, resulting in an increased susceptibility to fractures. The prevention of this public health problem can bring considerable social and economic benefits( Reference Lavanda, Saad and Lobo 126 ). Among the most important nutrients to obtain a maximum bone mass during the growth phase, Ca and Mg are of great importance. Their absorption occurs preferably in the small intestine, although it takes place to a less extent in the large intestine( Reference Pedreschi, Campos and Noratto 127 , Reference Raschka and Daniel 128 ).

Although the precise mechanism for prebiotic potential effect on osteoporosis is not known, several hypotheses have been proposed to explain the effect of FOS on Ca absorption and retention. One of them is related to the bacteria effect in the colon, as they are able to ferment FOS and other non-digestible carbohydrates, increasing the production of the SCFA (such as butyrate, propionate and acetate) plus other organic acids such as lactic acid. These compounds are able to reduce the pH through the acidification of the luminal content; thus, under these conditions, insoluble compounds such as Ca phosphate are dissolved in the lumen (ionised Ca), and there is an increased absorption of these compounds by passive diffusion( Reference Kruger, Brown and Collett 129 ). In addition, the SCFA may help increase Ca absorption rates through the exchange between cellular protons and luminal Ca( Reference Coxam 130 ) and through the epithelial cell proliferation stimulation, leading to an increased Ca absorption surface( Reference Lobo, Mancini-Filho and Alvares 131 ).

According to several studies conducted with animals, inulin is the fructans that has the greatest effect on increased Ca bioavailability( Reference Lavanda, Saad and Lobo 126 ).

According to Scholz-Ahrens et al.( Reference Scholz-Ahrens, Ade and Marten 132 ), a number of studies have investigated the effect of prebiotics (inulin, oligofructose and other non-digestible carbohydrates) on mineral absorption in humans, but the outcomes were contradictory. This may be related to the experimental conditions evaluated and the physiological characteristics of the target groups included, which may vary considerably among studies.

van der Heuvel et al.( Reference van der Heuvel, Muijs and Van Dokkum 133 ) and van der Heuvel et al.( Reference van der Heuvel, Schoterman and Muijs 134 ) showed, respectively, that menopausal women who received 10 g of lactulose or 20 g of TOS daily had higher levels of absorbed Ca, when compared with their respective controls. According to Coudray et al.( Reference Coudray, Bellanger and Castiglia-Delavaud 135 ), young men who consumed 40 g of inulin daily, over a period of 26 d, showed an increase in the apparent absorption of Ca of approximately 40 %, and this effect did not negatively affect the absorption of other minerals, including Mg, Zn and Fe. However, Tahiri et al.( Reference Tahiri, Tressol and Arnaud 136 ) conducted a randomised, double-blind, placebo-controlled study with menopausal women to assess the effect of the intake of 10 g of FOS daily (for 5 weeks) on intestinal Ca absorption and did not observe a significant improved absorption by the FOS group (35·63±9·40 and 36·55±8·48 % for prebiotic and placebo groups, respectively). Among the possible explanations for this observation, we ought to mention the dose of prebiotics tested, which may have been too low.

It is believed that the effect of prebiotics on minerals bioavailability and trace elements is related to the synthesis of polyamines, which are metabolites produced by various microbial strains( Reference Noack, Dongowski and Hartmann 137 ) and are able to stimulate cell multiplication, consequently leading to an increased absorption surface( Reference Bongers and Van Den Heuvel 138 ). Furthermore, probiotics are able to produce vitamins, which are required for bone matrix formation and bone growth( Reference Scholz-Ahrens and Schrezenmeir 139 ).

However, most scientific knowledge on the effect of probiotics and prebiotics on mineral metabolism is based on animal studies, generally rats. Therefore, additional studies, particularly human trials, are required for an accurate determination of the effects and mechanisms of the action involved in lowering the risk for osteoporosis through the consumption of probiotics and prebiotics.

Female urogenital health

The vaginitis caused by yeasts belonging to the Candida genus, the bacterial vaginosis, together with the urinary tract infections, annually affect nearly a billion women around the world( Reference Reid 140 ). The traditional drug treatment for these pathologies often destroys the autochthonous microbiota, frequently selects multidrug-resistant micro-organisms and causes side effects of variable intensities( Reference Gardiner, Heinemann and Baroja 141 ). Thus, studies with harmless micro-organisms that have scientifically shown a therapeutic effect can be important tools in the re-establishment of the autochthonous microbiota, which acts as a natural barrier against a number of pathogenic and opportunistic micro-organisms.

Lactobacilli are considered the predominant members of the human vaginal microbiota of healthy women in the post-puberty period, forming a biofilm on the surface of the vaginal mucosa( Reference Antonio, Hawes and Hillier 142 ). These micro-organisms are extremely important in the protection against various infectious agents( Reference Martin, Richardson and Nyange 143 , Reference Donders, Bosmans and Dekeersmaecker 144 ). This effect is related to their ability to adhere to the vaginal epithelial cells and to synthesise several inhibitory compounds against vaginal pathogens, including hydrogen peroxide, organic acids, bacteriocins, biosurfactants, auto- and co-aggregation molecules and the arginine deaminase enzyme( Reference McLean and McGroarty 145 Reference Martin, Soberon and Vasquez 149 ).

According to Reid & Burton( Reference Reid and Burton 150 ), the main characteristics necessary for species of the Lactobacillus genus to be used as probiotic cultures in the prevention and/or treatment of urogenital infections include the capacity of adherence to the vaginal epithelial cells, inhibition of adhesion and growth of pathogenic micro-organisms, depletion of nutrients available for pathogens, changing the microenvironment and modulation of the host immune response.

It is believed that lactobacilli exert beneficial effects on the prevention and/or treatment of urogenital infections through one or more different mechanisms, namely (i) ascendance of the probiotic micro-organisms from the rectum to the vagina; (ii) reduction of pathogen transference from the rectum to the vagina; or (iii) increased intestinal mucosal immunity, which acts on the vaginal immunity by making the vaginal tract less susceptible for the colonisation of pathogenic micro-organisms( Reference Reid and Bocking 151 Reference Reid, Charbonneau and Erb 153 ).

The study of the first probiotics for urogenital health, mainly L. rhamnosus GR-1 and L. reuteri RC-14 and also L. reuteri (formerly fermentum) B-54, began in the 1980s( Reference Reid 154 ). The first clinical studies to demonstrate a significant reduction in the incidence of recurrent urinary tract infections in women after intravaginal use of L. rhamnosus GR-1 alone and combined with L. reuteri B-54 were published by Bruce & Reid( Reference Bruce and Reid 155 ) and Bruce et al.( Reference Bruce, Reid and McGroarty 156 ).

L. rhamnosus GR-1 and L. reuteri RC-14 were shown to be promising probiotic strains when co-administered with traditional antimicrobial agents for the treatment of vulvovaginal candidiasis( Reference Martinez, Franceschini and Patta 157 ) and bacterial vaginosis( Reference Anukam, Osazuwa and Ahonkhai 158 , Reference Martinez, Franceschini and Patta 159 ).

With regard to the evaluation of other probiotics, Williams et al.( Reference Williams, Yu and Tashima 160 ) conducted a double-blind, placebo-controlled study in which L. acidophilus or clotrimazole were applied intravaginally once a week and it prevented the development of vulvovaginal candidiasis in HIV-positive patients, in comparison with the group of women receiving placebo.

On the other hand, Pirotta et al.( Reference Pirotta, Gunn and Chondros 161 ) conducted a double-blind, placebo-controlled study with non-pregnant women and reported that L. rhamnosus and B. longum (administered orally) or L. rhamnosus, Lactobacillus delbrueckii, L. acidophilus and S. thermophilus (intravaginal use) were not effective in preventing the infection after the treatment with different antimicrobial drugs, in comparison with the control group.

Larsson et al.( Reference Larsson, Stray-Pedersen and Ryttig 162 ) performed a double-blind, placebo-controlled study in which they observed that the combined use of topical clindamycin and daily vaginal insertion of capsules containing Lactobacillus gasseri Lba EB01-DSM 14869 and L. rhamnosus Lbp PB01-DSM 14870, for a period of 10 d during three menstrual cycles, did not improve the effectiveness of the bacterial vaginosis therapy during the 1st month of treatment; however, the therapeutic procedure extended the time for the occurrence of relapse, evaluated after 6 months of treatment. On the other hand, Eriksson et al.( Reference Eriksson, Carlsson and Forsum 163 ) showed that the co-administration of clindamycin ovules and capsules for vaginal use with lactobacilli (L. fermentum, L. casei-rhamnosus and L. gasseri) did not raise the infection cure rate (62 %), compared with patients who used the antibiotic and capsules containing placebo (56 %).

This inconsistency of results in several clinical studies with the use of potentially probiotic LAB may be partially related to the fact that studies were limited to the effects of a specific microbial strain, to small groups of monitored patients and also to the lack of available data regarding the adequate identification and stability of the micro-organism tested, when capsules, tampons or any other vehicle were used for probiotic delivery. Other important issues when comparing studies published by different authors that assessed the effect of probiotics on urogenital infections include the fact that the micro-organisms’ dosages and the periods for both treatment and evaluation of outcomes are normally not uniform. It is also necessary to stress out that when probiotics are co-administered with antibacterial and antifungal agents, the viability of the micro-organisms ought to be previously assessed in vitro so that it can be verified that they are not adversely affected in the presence of these drugs, which could severely impair their efficacy. Thus, it is clear that further, standardised studies are required to conclusively prove the effectiveness of such approaches in the prevention and/or treatment of urogenital infections, especially at a time when few drugs are in the pipeline, and to precisely elucidate precisely the mechanisms of actions involved.

Cavities, periodontal disease and halitosis

Cavity is one of the most prevalent chronic diseases in the world( Reference Bizzini, Pizzo and Scapagnini 164 ). This pathology, bacterial in origin, has multifactorial causes and is characterised by tooth enamel demineralisation( Reference Selwitz, Ismail and Pitts 165 ). This disease occurs after changes in oral ecosystem homoeostasis, resulting in the proliferation of bacterial biofilm, predominantly composed of bacteria belonging to the Streptococcus mutans group( Reference Islam, Khan and Khan 166 ). To present beneficial effects, limiting or preventing the development of cavities, the potential probiotic micro-organism needs to attach to the dental surface and integrate the bacterial biofilm, where it must compete with cariogenic bacteria, preventing their growth. In addition, as the probiotic micro-organism metabolises sugars obtained from the diet, it is necessary that it produces low acid levels( Reference Bonifait, Chandad and Grenier 167 ).

Various clinical studies have shown that regular consumption of yogurt, milk or cheese containing probiotic cultures leads to the reduction in the number of cariogenic streptococci in saliva and dental plaque( Reference Ahola, Yli-Knnuuttila and Suomalainen 168 , Reference Nikawa, Makihira and Fukushima 169 ). An interesting study was conducted by Kang et al.( Reference Kang, Chung and Kim 170 ), in which two microbial strains of Weissella cibaria CMS1 and CMS2, isolated from the saliva of children, showed in vitro inhibitory effect against the formation of dental plaque. The researchers also evaluated seventy-two patients who used an oral rinsing solution containing 109 CFU/ml of strain CMS1, twice a day, for 1 d, and noted inhibition of dental plaque formation in 20·7 % of the subjects. Even though W. cibaria CMS1 reduced the biofilm formation, Kang et al.( Reference Kang, Chung and Kim 170 ) did not observe any antagonistic activity against S. mutans.

Periodontal diseases include gingivitis and periodontitis. The first one is characterised by inflammation confined to the gingival tissue, whereas the periodontitis is a progressive disease, which destroys the teeth support tissues, including the alveolar bone( Reference Houle and Grenier 171 ). The main micro-organisms associated with the development of gingivitis include Porphyromonas gingivalis, Treponema denticola, Tannerella forsythia and Aggregatibacter actinomycetemcomitans ( Reference Houle and Grenier 171 ). Some studies demonstrated the ability of certain lactobacilli strains to inhibit the growth of pathogens related to the development of periodontal disease( Reference Sookkhee, Chulasiri and Prachyabrued 172 , Reference Koll-Klais, Mandar and Leibur 173 ). These observations led researchers to believe that the autochthonous lactobacilli from the oral microbiota could have an important role in this microecosystem balance. Thus, Krasse et al.( Reference Krasse, Carisson and Dahl 174 ) studied the effect of a chewing gum supplemented with probiotic L. reuteri strains (LR-1 or LR-2), which was administered to volunteers with gingivitis (moderate to severe) during 14 d. The authors observed that the oral cavity of these individuals was colonised by the micro-organisms and that they even helped to reduce the dental plaque index scores. Riccia et al.( Reference Riccia, Bizzini and Perilli 175 ) studied the anti-inflammatory effects of L. brevis CD2 in a group of patients diagnosed with chronic periodontitis and observed that the micro-organism, administered over a period of 4 d, was able to improve the clinical parameters evaluated (plaque and gingival index scores and bleeding) in all volunteers.

Interestingly, according to a review paper published by Laleman & Teughels( Reference Laleman and Teughels 176 ), several studies on real periodontitis patients showed a significant reduction of gingivitis and plaque index associated with the use of probiotics, whereas this effect was not observed for experimental gingivitis patients. The authors stated that the pronounced heterogeneity in studies that assessed the efficacy of probiotics on periodontal diseases makes comparison between them difficult. This heterogeneity includes several aspects, such as a very diverse patient population studied (e.g. healthy volunteers, experimental gingivitis models, and patients with gingivitis, chronic and aggressive periodontitis) and a broad range of parameters evaluated (e.g. microbiological determinations in saliva and plaque, various plaque and gingivitis indices, bleeding on probing and probing pocket depth). Laleman & Teughels( Reference Laleman and Teughels 176 ) highlighted that the clinical studies usually evaluate probiotic use for a short period (up to 3 months), which demonstrates that real periodontal parameters, such as probing pocket depth, are not adequately characterised or are not significantly different from baseline data.

Another example of probiotic application in dental practice is in the treatment of halitosis, a disease that has several causes, including metabolic disorders, consumption of certain types of food and respiratory tract infections. However, most cases of the pathology are associated with an imbalance of the oral cavity commensal microbiota. In fact, halitosis results from the action of anaerobic bacteria that degrade proteins present in the saliva and in the food and, as a consequence, produce amino acids, which are transformed into volatile compounds, responsible for the characteristic halitosis oral malodour( Reference Scully and Greenman 177 ).

Burton et al.( Reference Burton, Chicott and Tagg 178 ) evaluated a group of patients diagnosed with halitosis who were treated with antimicrobials for 3 d and administered the probiotic organism Streptococcus salivarius K12 supplement for 2 weeks. The researchers noted that most of the volunteers exhibited reduced levels of volatile sulphur compounds (VSC) for at least 2 weeks. In addition, all patients showed increased levels of S. salivarius K12 and reduction of the bacterial populations responsible for the oral malodour, both determined in the saliva samples evaluated. Different results were reported by Keller et al.( Reference Keller, Bardow and Jensdottir 179 ), who conducted a randomised placebo-controlled study, in which a probiotic gum containing L. reuteri DSM 17938 and L. reuteri PTA 5289 was used for 14 d by patients who self-reported malodourous morning breath. The authors concluded that the probiotic gum consumption did not alter VSC levels, even though the organoleptic scores were significantly lower in the probiotic group when compared with the placebo group.

The complex oral microbiota constitutes a major challenge for the prevention and control of cavities, periodontal disease and halitosis using probiotic LAB, although several clinical studies have demonstrated such potential. According to Laleman & Teughels( Reference Laleman and Teughels 176 ), for future long-term probiotic trials, randomisation and blinding steps should be adequately followed and should include large groups of patients and assess real dental parameters (caries, plaque formation and probing pocket dept) instead of intermediate end points. Furthermore, researchers should carefully investigate strain- and dose-specific effects and evaluate the most appropriate vehicles for their delivery. As substantial scientific evidence demonstrates the usefulness of probiotics on maintenance and/or improvement of oral health, this approach can be successfully applied in dental practice.

Allergic reactions

The increasing prevalence of allergic diseases, especially in industrialised countries, has driven the research towards a better understanding of the possible causal factors involved, as well as the development of new, safer and more effective treatments( Reference Schabussova and Wiedermann 180 ).

The interest in the intestinal microbiota has increased in recent years, especially concerning the role of LAB in the development and prevention of allergic diseases( Reference Penders, Stobberingh and Van Den Brandt 181 , Reference Prescott and Bjorksten 182 ). Several studies have shown that the composition of the intestinal microbiota and, particularly, the presence of certain species of LAB, is different between healthy children and those suffering from atopic diseases( Reference Bjorksten, Sepp and Julge 183 Reference Sepp, Julge and Mikelsaar 185 ).

There is a great enthusiasm for human diet supplementation with probiotic LAB for the prevention of allergies, as these micro-organisms are well tolerated and may be appropriate for the protective immune function regulation( Reference Schabussova and Wiedermann 180 ). Certain clinical studies have demonstrated beneficial effect in the prevention of atopic disease in newborn babies, whose mothers had their diets supplemented with probiotic cultures during pregnancy( Reference Kukkonen, Savilahti and Haahtela 186 , Reference Marschan, Kuitunen and Kukkonen 187 ).

Rautava et al.( Reference Rautava, Kalliomaki and Isolauri 188 ) observed that the administration of L. rhamnosus GG to Finnish pregnant women (2×1010 CFU of the micro-organism/daily during 4 weeks before giving birth) and throughout breast-feeding (identical dose; otherwise infants received the probiotic) significantly reduced the risk of development of atopic eczema in infants during the first 2 years of life. However, Kopp et al.( Reference Kopp, Hennemuth and Heinzmann 189 ) reported different results using the same probiotic strain. These researchers observed that the administration of the micro-organism to German women during the gestation period (1×1010 CFU of L. rhamnosus GG/daily during 4–6 weeks before giving birth) and postnatally (same dose for 6 months; after 3 months, the probiotic was given only to the neonates) did not prevent the development of atopic dermatitis in newborns and also increased the risk of developing bronchitis. These results are important to emphasise the need for the careful selection of bacterial strains and their full characterisation regarding immunomodulatory properties before their use for prophylactic or therapeutic purposes. Moreover, according to Kopp et al ( Reference Kopp, Hennemuth and Heinzmann 189 ), when results obtained from clinical trials performed with diverse populations are compared (e.g. volunteers from different nationalities), it is necessary to draw attention to the fact that different genetic backgrounds may also be involved. This observation has a significant role in the final outcomes observed. Finally, the authors concluded that additional studies are necessary to determine whether there are susceptible subgroups of patients, and how they may profit from specific dietetic supplementation with probiotics.

According to a meta-analysis published by Kim et al.( Reference Kim, Ah and Yu 190 ), probiotics may be considered an option for the treatment of atopic dermatitis, namely for moderate to severe cases of the disease in children and adults, although there is not enough evidence to support their usefulness in infants. The authors also stated that, among the twenty-five randomised controlled trials included in the meta-analysis, differences in probiotic strains and their doses tested, food intake, compliance and medicaments taken simultaneously with the probiotic interventions are important constraints when making robust conclusions. Therefore, improvement of clinical trials design matching these variables (head-to-head comparison) would allow a more fair comparison between outcomes and corroborate (or not) their usefulness in clinical practice.

A study developed by Martinez-Canavate et al.( Reference Martinez-Canavate, Sierra and Lara-Villoslada 191 ) assessed the immunological effects of the consumption of a milk product containing two probiotic strains (L. gasseri CECT5714 and Lactobacillus coryniformis CECT5711) in children with respiratory allergies (asthma, pollen allergy or both). The researchers observed that the volunteers who consumed probiotic cultures showed increased levels of IgA in the mucosa, as well as higher levels of regulatory CD4+/CD25+ T cells in plasma samples. On the other hand, the researchers found a reduction in the plasma IgE levels in the group supplemented with probiotics and an increased number of natural killer (NK) cells, compared with the group that received the traditional yogurt. The authors concluded that the probiotic cultures reinforced the innate and specific immunity and improved the general status of the children’s health.

Another area of great interest in the application of probiotics is related to the prevention of foodborne allergies. It is possible that these micro-organisms are able to reduce intestinal inflammation and therefore improve the clinical features of susceptible patients. A number of studies have been conducted in the last decades to further explore this therapeutic approach( Reference Majamaa and Isolauri 192 Reference Osborn and Sinn 194 ).

Several mechanisms have been proposed to explain the interaction of probiotic LAB with the host’s immune system, including the following: stimulation of IgA secretion in the mucosal surfaces( Reference Wells and Mercenier 195 ); induction of pro-inflammatory or regulatory cytokines production( Reference Sashihara, Sueki and Ikegami 196 Reference Kekkonen, Kajasto and Miettinen 198 ); modulation of dendritic cell differential maturation( Reference Foligne, Zoumpopoulou and Dewulf 199 ); and interaction with the immune system through signalling via the ‘toll-like receptor’ (TLR)( Reference Blander and Medzhitov 200 ). In addition, it is important to highlight that the immunogenicity attributed to some probiotic LAB does not necessarily require that these micro-organisms remain viable and survive the passage throughout the GIT( Reference Fujiwara, Inoue and Wakabayashi 201 ).

In a general context, there are a limited number of clinical trials that demonstrate positive effects of probiotics on the prevention and/or treatment of allergies, whereas other studies report negative results. Until now, only a limited number of probiotic strains have shown beneficial effects, especially regarding the prevention of allergic diseases. It is important to stress out that this research area is relatively new, since the first intervention trial was reported in 1997( Reference Majamaa and Isolauri 192 ). In general, the available scientific evidence rather reflects the inherent complexity of the allergic syndromes, the characteristics and potential variables of the different probiotic strains tested and the limited understanding of the mechanism by which they can mitigate and/or neutralise different types of immune dysfunction found in allergic diseases( Reference Kalliomäki, Antoine and Herz 202 ). Therefore, a larger number of studies will be needed to ensure the effectiveness of the use of LAB probiotic strains with this purpose. According to Kalliomäki et al.( Reference Kalliomäki, Antoine and Herz 202 ), in the future, properly selected probiotic strains for allergic conditions in well-defined specific target populations may become an efficient tool to fight them. The authors also emphasise that clinical trials should use standardised criteria for both diagnosis and symptom scoring, as well as for evaluating the genetic predisposition to allergic diseases.

The effectiveness of prebiotics on allergic diseases in preterm infants was investigated by Niele et al.( Reference Niele, van Zwol and Westerbeek 203 ), who performed a randomised controlled trial in which 113 preterm infants received enteral neutral and acid oligosaccharide supplementation or placebo from day 3 to 30 of life. The authors observed that the incidence of allergic diseases during the 1st year of life was not different between both groups studied. Interestingly, a study published by Schouten et al.( Reference Schouten, Van Esch and Kormelink 204 ) demonstrated that a prebiotic mixture containing short-chain GOS and long-chain FOS was able to reduce the cumulative incidence of atopic dermatitis in infants at risk for allergy, in comparison with placebo supplementation.

Anyhow, there is no overall consensus on the effectiveness of this approach, as according to a systematic review published by Williams & Grindlay( Reference Williams and Grindlay 205 ) there is not yet any substantial evidence of benefit for the use of prebiotics in atopic dermatitis prevention. Furthermore, according to another systematic review made by Osborn & Sinn( Reference Osborn and Sinn 206 ), more controlled studies are needed before prebiotics can be routinely used for the prevention of allergy in formula-fed infants, even though some evidence exists that prebiotic addition to infant formulas might prevent eczema.

Inflammatory bowel diseases and irritable bowel syndrome

Inflammatory bowel diseases (IBD) are a group of inflammatory disorders from the GIT with multifactorial aetiology, including ulcerative colitis (UC) and Crohn’s disease (CD). UC is characterised by a continuous superficial mucosal inflammation of the colon, whereas in the case of CD inflammation is discontinuous and most often affects the ileum and the colon, although it may also affect different parts of the GIT. Even though the exact cause of the development of IBD is unknown, the hypothesis usually accepted is that the pathology begins with the loss of oral tolerance in people genetically predisposed, resulting in chronic intestinal inflammation( Reference Jonkers, Penders and Masclee 207 ).

The first evidence of intestinal microbiota in the development of IBD was obtained from a study of a colitis model, in which inflammatory reaction was observed in germ-free animals( Reference Kawada, Arihiro and Mizoguchi 208 ). Besides the normal intestinal microbiota imbalance, some patients with IBD present an overreacting immune response of commensal micro-organisms, which is believed to be an important factor in the aetiology of the disease. Although the exact mechanisms involved in the loss of oral tolerance have not yet been completely elucidated, some researchers have demonstrated an increased infiltration of activated CD4+ lymphocytes in the mucosa, dysfunction of dendritic cells and abnormal immune responses induced by macrophages( Reference Abraham and Medzhitov 209 , Reference Danese 210 ).

Besides the possibility of surgical interventions, traditional antibiotics are also used in the treatment of UC and CD. However, these may lead to important side effects, including leukopenia, abnormalities in the liver functions, nephritis and pancreatitis( Reference Dignass, Van Assche and Lindsay 211 ). This scenario stimulates the search for new strategies to be used in the appropriate management of these pathologies. Thus, the manipulation of the composition and activity of endogenous intestinal microbiota, in addition to the barrier function and the immune system, have been the main strategies assessed through intervention studies with the use of probiotic micro-organisms. Different mechanisms have been proposed to explain the beneficial effects of probiotics in patients with IBD, which include competition for nutrients and adhesion sites, production of antimicrobial substances and/or cell–cell communication( Reference Lebeer, Vanderleyden and De Keersmaecker 212 ). Probiotics may affect the immune system through the interaction of bacterial products, such as cellular components or DNA, with epithelial and immune cells associated with the intestine( Reference Oelschlaeger 213 ). In addition, some studies have also shown changes in the profile of cytokines produced, modulation on dendritic cells function, increased activity of NK cells and induction of regulatory T cells and defensins( Reference Lebeer, Vanderleyden and De Keersmaecker 212 Reference Borchers, Selmi and Meyers 214 ). Finally, probiotics may contribute towards the production of SCFA, modifying the barrier function by inducing the production of mucin, favouring tight junctions, besides reducing cell apoptosis( Reference Caballero-Franco, Keller and De Simone 215 , Reference Karczewski, Troost and Konings 216 ).

Following this line, Ng et al.( Reference Ng, Plamondon and Kamm 217 ) evaluated in vivo effects of the oral use of a commercial probiotic product VSL#3® (L. casei, L. plantarum, L. acidophilus, L. delbrueckii ssp. bulgaricus, B. longum, B. breve, B. infantis and S. thermophilus; Sigma-tau Pharmaceuticals Inc.) and steroids on colonic dendritic cells in patients with acute UC. Rectal biopsies were obtained from patients with active UC before and after treatment with VSL#3® and corticosteroids, or placebo, and from healthy controls. The authors showed that treatment of UC patients with probiotic VSL#3® and corticosteroids induced favourable intestinal dendritic cell function in vivo, increased the levels of regulatory cytokines and lowered both the levels of pro-inflammatory cytokines and TLR expression. Thus, the researchers suggested that these effects may contribute to therapeutic benefits.

Rahimi et al.( Reference Rahimi, Nikfar and Rahimi 218 ) published a meta-analysis on eight clinical trials with CD patients and concluded that probiotics were not effective in maintaining remission and in preventing clinical and endoscopic recurrence of the disease. More recently, a review article published by Shen et al.( Reference Shen, Zuo and Mao 219 ), based on a meta-analysis of twenty-three randomised controlled trials including a total of 1763 subjects, VSL#3® was the most effective treatment for the management of UC, whereas its effect on CD was much less pronounced. In fact, meta-analyses are powerful tools to demonstrate scientific evidence; however, they must be used rationally. For instance, when meta-analyses gather data from different probiotics (efficacious and non-efficacious), various conditions, and patients with diverse characteristics, the final result will merely be an average non-effect because of the heterogeneity of the benefits and the probiotics evaluated. Moreover, meta-analyses can also indicate generic activities of micro-organisms instead of the distinct functionality of a particular strain( Reference Rijkers, Bengmark and Enck 220 ).

Irritable bowel syndrome (IBS) is a highly prevalent gastrointestinal disorder, which is difficult to treat and is characterised by a set of complex symptoms( Reference Moayyedi, Ford and Talley 221 ). Patients with IBS normally have crampy abdominal pain, altered bowel habits, bloating, flatulence and disturbed defecation. These symptoms vary in intensity (mild to severe), but even though IBS is a benign disease its impact on well-being and overall quality of life is notoriously adverse. Moreover, for many subjects, IBS is a chronic condition that follows a course of relapse and remission of symptoms( Reference Spiller, Aziz and Creed 222 , Reference Clarke, Cryan and Dinan 223 ).

Several researchers have evaluated the use of probiotic micro-organisms in the treatment of IBS; however, the results are inconclusive. Some studies demonstrated an overall improvement of symptoms with probiotics, whereas others reported an absence of any beneficial outcomes. Clarke et al.( Reference Clarke, Cryan and Dinan 223 ) published a review article on clinical studies with the purpose of determining the efficacy of probiotic LAB in the treatment of IBS. The authors observed that among the forty-two studies examined thirty-four indicated positive effects in at least one of the parameters or symptoms evaluated, although a high variation in the intensity of effects and in the probiotics tested was observed. Clarke et al.( Reference Clarke, Cryan and Dinan 223 ) also pointed out several problems regarding the type of experimental design used, inadequate selection and doses of probiotic micro-organisms, unknown mode of action and scarcity of available data on the tolerance of the ingested micro-organisms during a long period of time, once the pathology is a long-lasting or chronic condition. In spite of these limitations, there seems to be a consensus among specialists in this field that some probiotic LAB are efficient in the treatment of IBS; however, the positive outcomes vary and are related to the time duration of their administration in patients( Reference Brenner, Moeller and Chey 224 Reference Horvath, Dziechciarz and Szajewska 226 ).

Ringel & Ringel-Kulka( Reference Ringel and Ringel-Kulka 227 ) also reviewed the scientific literature on clinical randomised placebo-controlled clinical studies that determined the efficiency of probiotics in the treatment of patients diagnosed with IBS. Similar to the study published by Clarke et al.( Reference Spiller, Aziz and Creed 223 ), Ringel & Ringel-Kulka( Reference Ringel and Ringel-Kulka 227 ) observed important differences regarding the experimental design, doses, probiotic strains and parameters used between the different studies reviewed. According to this meta-analysis, as a whole, probiotics were more efficient than placebo in the improvement of IBS symptoms, as well as for the reduction of the risk for persistent symptoms. Whelan( Reference Whelan 228 ) highlighted that meta-analysis is a valuable tool to gather individual small trials to improve the ability to determine the direction, size and consistency of an effect; however, it can do little to overcome the poor design of individual trials, frequently seen for the management of IBS with probiotics. The author also recommended that all future meta-analyses on probiotics ought to include a subgroup of analyses on specific combinations (species/strains).

Overall, according to reviewed and meta-analysed data, bifidobacteria, lactobacilli, Escherichia coli, E. faecalis and a mixture of bacterial strains were the most promising micro-organisms for the management of IBS. However, there is no strict consensus on the rationale use of the approach, attributed to the lack of knowledge of their exact mechanisms of action (via intestinal immune system, enteric nervous system or otherwise), complexity of the disease (variable course of symptoms and heterogeneity) and high placebo responses observed in clinical trials on IBS (up to 50 %)( Reference Haller, Antoine and Bengmark 229 ).

Two studies that addressed the positive outcomes obtained with interventions carried out using probiotics in the treatment of IBS should be mentioned. One of them was published by O’Mahony et al.( Reference O’Mahony, Mccarthy and Kelly 230 ) and evaluated the use of B. infantis 35624 or Lactobacillus salivarius UCC4331 v. placebo in the treatment of the pathology. The researchers observed a significant improvement in the symptoms (pain reduction, less bloating and reduced intestinal motility) in the group that received B. infantis 35624, in comparison with the placebo group. Whorwell et al.( Reference Whorwell, Altringer and Morel 231 ) conducted a multi-centre study and confirmed these results by the observation that patients diagnosed with IBS supplemented during 4 weeks with B. infantis 35624 exhibited significant reduced pain and an overall improvement of IBS symptoms, compared with the subjects who received placebo.

According to Bonfrate et al.( Reference Bonfrate, Tack and Grattagliano 232 ), only a limited number of clinical trials have assessed the effect of prebiotics on IBS patients. Among them, Hunter et al.( Reference Hunter, Tuffnell and Lee 233 ) performed a double-blind crossover study with patients suffering from IBS who received, three times daily, 2 g of oligofructose (Raftilose P95; Orafti) or 1 g of sucrose, both administered during 4 weeks. The researchers observed that the prebiotic tested did not alter any of the parameters studied (faecal weight and pH, whole-gut transit time, and fasting breath hydrogen concentrations) and, therefore, concluded that oligofructose given at a dose of 6 g/d presented no therapeutic value in patients diagnosed with the disease. Olesen & Gudmand-Hoyer( Reference Olesen and Gudmand-Hoyer 234 ) performed a multicentre, double-blind randomised placebo-controlled study and reported no significant differences in the severity and duration of symptoms (abdominal distension, abdominal rumbling, abnormal flatulence and abdominal pain) in IBS patients who received 20 g of FOS powder daily for 12 weeks, compared with those who had taken placebo.

On the other hand, Silk et al.( Reference Silk, Davis and Vulevic 235 ) investigated the effect of a novel prebiotic TOS on IBS patients. The authors evaluated a group of forty-four subjects diagnosed with the disease who were randomised to receive the prebiotic (3·5 or 7 g/d) or placebo (7 g/d) and observed that prebiotics increased the bifidobacteria populations in the faecal samples collected throughout the study. Silk et al.( Reference Silk, Davis and Vulevic 235 ) also reported that although the lowest dosage of the prebiotic significantly changed stool consistency, improved bloating, flatulence and subjective global assessment (SGA), the highest dosage of the prebiotic improved SGA and anxiety scores. Finally, Andriulli et al.( Reference Andriulli, Neri and Loguercio 236 ) evaluated the efficacy of a product called Flortec (Anidral Co.), containing a mixture of prebiotics and probiotic Lactobacillus paracasei B21060 in people diagnosed with IBS. The authors conducted a multicentre, randomised study in which patients received the prebiotic or Flortec, and observed that in IBS-predominant diarrhoea the latter significantly reduced bowel movements, abdominal pain and IBS scores. As shown above, different types of prebiotics have been administered at distinct doses and time periods, which make the comparison between the results of the various studies and the final decision of which one(s) is (are) the most promising to treat the disease particularly hard if not impossible.

Considered altogether, the scientific evidences on the probiotic effects are less contradictory in the case of UC, whereas for CD the available results are still scarce and not so encouraging( Reference Guandalini, Cernat and Moscoso 237 ). As for other diseases, the number of clinical studies that evaluate the effects of probiotics and prebiotics on IBD and IBS is still limited, which justifies the necessity of conducting new rigorous, long-term, well-planned, randomised clinical studies in this field of knowledge( Reference Lyra, Lahtinen and Ouwehand 86 , Reference Floch 238 ). Particularly for IBS clinical trials, substantial clinical outcomes should include the evaluation of symptom improvement with psychometrically validated SGA or validated symptom severity questionnaires( Reference Haller, Antoine and Bengmark 229 ).

Helicobacter pylori gastric infections

Helicobacter pylori infection may lead to chronic gastritis, and it constitutes the leading cause of peptic ulcer disease, besides being a risk factor for the development of gastric cancer( Reference Lesbros-Pantoflickova, Corthesy-Theulaz and Blum 239 ). Currently, there is a remarkable interest in the development of low-cost and large-scale solutions to prevent or reduce the gastric colonisation by H. pylori. In this context, probiotics become an especially interesting approach.

Several clinical studies have shown the beneficial effect of Lactobacillus johnsonii La1 on gastritis caused by H. pylori. Michetti et al.( Reference Michetti, Dorta and Wiesel 240 ) noted that the administration of L. johnsonii La1 culture supernatant suppressed the pathogen urease activity in asymptomatic volunteers. The researchers observed that the effect remained for 6 weeks after the end of the treatment, although the suppression of H. pylori urease activity was not intensified by the co-administration with omeprazole. In reality, H. pylori is known to catalyse the conversion of urea to dioxide and ammonia. The latter is then turned into ammonium hydroxide neutralising the local acidity, which favours the pathogen survival( Reference Ruggiero 241 ). Felley et al.( Reference Felley, Corthesy-Theulaz and Rivero 242 ) observed that patients diagnosed with H. pylori infection (evidenced by histological examination of the gastric biopsies), treated with clarithromycin and supplemented with preparations of milk containing L. johnsonii La-1, showed a reduction of both the infection intensity and the pathogen density in the tissue sample. The authors highlighted that such effects persisted for several weeks after the end of the probiotic intake period. Pantoflickova et al.( Reference Pantoflickova, Corthesy-Theulaz and Dorta 243 ) observed that the treatment with the strain L. johnsonii La1, administered over a period of 16 weeks and without any antibiotic treatment, reduced the intensity of gastritis associated with H. pylori, as well as the pathogen density determined in the gastric antrum. The authors reported that the positive effects related to the probiotic administration were obtained 3 weeks after the beginning of treatment and remained throughout the period of L. johnsonii La-1 intake.

Wang et al.( Reference Wang, Li and Liu 244 ) observed a positive effect regarding the reduction of the gastric infection symptoms and colonisation by H. pylori when probiotic strains L. acidophilus La-5 and B. animalis Bb-12 were tested in patients diagnosed for gastritis.

Lionetti et al.( Reference Lionetti, Miniello and Castellaneta 245 ) evaluated a H. pylori-positive paediatric population (forty subjects; mean age 12·3 years old) who received omeprazole and amoxicillin during 5 d, followed by omeprazole, clarithromycin and tinidazole for another 5 d and were then randomised to receive pills containing the probiotic strain L. reuteri ATCC 55730 or placebo. The authors reported a significant reduction of gastrointestinal symptoms in the probiotic group, when compared with children supplemented with placebo. Lionetti et al.( Reference Lionetti, Miniello and Castellaneta 245 ) concluded that the probiotic strain tested was effective in reducing the frequency and intensity of antibiotic-associated side effects during H. pylori eradication therapy.

Francavilla et al.( Reference Francavilla, Polimeno and Demichina 246 ) studied a group of 100 patients infected with H. pylori who received a combination of probiotic strains (L. reuteri DSM 17938 and L. reuteri ATCC PTA 6475) or placebo during a three-phase study (pre-eradication phase – days 0–28; eradication treatment – days 29–35; and follow-up – days 36–96). The authors observed that the probiotics tested demonstrated an inhibitory effect on the pathogen growth; however, when L. reuteri DSM 17938 and L. reuteri ATCC PTA 6475 were associated with the eradication therapy (clarithromycin-amoxicillin for 7 d), significantly increased eradication rates (about 9 %) and reduction of both gastrin-17 and antibiotic-associated side effects were found.

Several mechanisms could explain the decreased H. pylori density and the reduction in the inflammatory reaction caused by the pathogen by using probiotic cultures, administered alone or co-administered with antibiotics. Among them, the strengthening of the gastric mucosa immune defences and the increase in the specific and non-specific immune response have important roles( Reference Lesbros-Pantoflickova, Corthesy-Theulaz and Blum 239 , Reference Yang, Chuang and Yang 247 ). Moreover, some LAB (L. gasseri Chen and L. plantarum 18) were shown to inhibit H. pylori adherence to gastric epithelial cells( Reference Chen, Liu and Tian 248 ). In fact, H. pylori colonise the mucus layer in the stomach, mostly adhering to epithelial cells. As the chances for probiotic micro-organisms to arrive at this site in significant amounts are extremely reduced, it seems more feasible that, at least for therapeutic purposes, probiotics present an indirect and non-specific instead of a direct and specific anti-H. pylori activity( Reference Ruggiero 241 ).

On the other hand, Navarro-Rodriguez et al.( Reference Navarro-Rodriguez, Silva and Barbuti 249 ) reported that a probiotic compound containing L. acidophilus, L. rhamnosus, B. bifidum and Streptococcus faecium (1·25×109 CFU each) administered for 30 d to patients with peptic ulcer or functional dyspepsia because of H. pylori infection, who were previously treated with furazolidone, tetracycline and lansoprazole for 7 d, did not increase the efficacy nor improved the adverse effects of the treatment, when compared with patients supplemented with placebo. Along this line, an interesting study conducted by Szajewska et al.( Reference Szajewska, Horvath and Piwowarczyk 250 ) demonstrated that S. boulardii supplementation showed limited effect on the H. pylori eradication rate. However, the authors reported that the probiotic reduced the adverse effects related to the drug therapy.

According to a meta-analysis published by Zhang et al.( Reference Zhang, Qian and Qin 251 ), the co-administration of probiotics and antimicrobial agents was associated with an increased H. pylori eradication rate and reduction of the adverse effects; however, patient compliance was not improved, which may be related to innate personality features. The authors also highlighted the fact that not all publications include in their clinical studies important subgroups such as patients infected with antibiotic-resistant H. pylori treated with probiotics, which are of great importance for clinical practice. According to Ruggiero( Reference Ruggiero 241 ), the diverse results obtained from different clinical trials reflect the diversity of both probiotic strains and antibiotic agents tested (including respective doses and time periods of administration), as well as the variation in geographic areas that are related to distinct H. pylori strain distribution, host susceptibility and therapy efficacy. The author also emphasised that these variables make a direct comparison of the results obtained from single studies impractical; however, the global findings represent a valuable knowledge on the possible efficacy of probiotic use.

Thus, long-term placebo-controlled clinical studies, considering standardised patients traits such as age, gravity of infection and types of gastrointestinal symptoms, among others, and involving a larger number of volunteers, are still required to clarify the real benefits of the co-administration of probiotics and antimicrobials for an adequate treatment of H. pylori gastric infections. Interestingly, according to Ruggiero( Reference Ruggiero 241 ), in order to determine the efficacy of probiotics in the management of the disease, specific studies targeting H. pylori (strain-specific infectivity potential) and the host (genetic background and microbiome) are of great importance.

Conclusion

The field of probiotics and prebiotics has substantially advanced in recent years, stimulated by progress in the comprehension of the role of the human microbiota in health and diseases. Therefore, future investigations that set effective strategies to shape a healthier microbiota, which might help in our physiology and disease processes, will be increasingly stimulated and represent a potentially fruitful area of scientific research. Although there are numerous studies described in the scientific literature regarding the beneficial effects obtained from the consumption of probiotics and/or prebiotics, health claim restrictions on them are made by authorities worldwide, especially in Europe( Reference Katan 252 ). These effects are observed both in the GIT (prevention and/or treatment of gastric and intestinal infections, IBD and IBS) and in other sites (prevention and/or treatment of respiratory tract infections, CVD, osteoporosis, urogenital tract infections, oral cavity infectious diseases and atopic allergy). It is noteworthy that the effects, as well as the mechanisms of action involved, are considered strain specific. The selection of probiotic strains should be directed to the desired effects shown by the micro-organisms of interest, supported by in vitro and in vivo assays, when tested alone or incorporated into a food matrix or a pharmaceutical preparation. Thus, the probiotic strains used for production and the industrial large-scale processing should be adequately characterised and appropriate for each type of product in which they will be delivered, including high viability throughout the storage period and scientific-based evidence for specific health claims.

The improvement of probiotic and prebiotic component application to support human health relies on a better understanding of their mode of action. This clarification will enable further advances in probiotic and prebiotic research and will undoubtedly contribute to the appropriate use of these functional products. In fact, the elucidation of the mechanisms underlying the human microbiota and immune system modulation by probiotics and prebiotics will heavily depend on continuous efforts (massive financial support) to identify appropriate biomarkers of health and diseases risk factors that might improve the design of human trials required for health claim verification. Moreover, large and long-term better-aligned clinical studies in humans are required to provide more reliability and a more solid basis for the results achieved. Systematic reviews and meta-analyses are crucial means to assess the strength of scientific evidence for health effects attributed to the consumption of probiotics and prebiotics and should therefore be acceptable by the EFSA in order to substantiate health claims( Reference Glanville, King and Guarner 253 ). Although it is important to emphasise that even in case specific health claims would be accepted by EFSA, probiotic and prebiotic food manufacturers are not currently allowed to include these substantiated health claims on the food product label, as that is not the purpose of the consumption of functional foods and only the claims related to the reduction of risk for disease are accepted worldwide.

At the moment, the accumulated knowledge of the beneficial effects of probiotics indicates that the strongest evidence for their efficacy, among all the possibilities discussed in this review article, is that related to the prevention and/or management of intestinal infections, although no definite conclusions can be made. In fact, the inconsistency among studies, which may have a large number of variables, for example, the study design itself, the probiotic strain used, outcome parameters, the population studied and the still limited number of clinical studies are undoubtedly hurdles that have to be overcome( Reference Wolvers, Antoine and Myllyluoma 60 ).

Innovative, safe and cost-effective interventions, for instance, the use of probiotics as toxin scavengers, especially metals present in high levels in the bloodstream( Reference Bisanz, Enos and Mwanga 254 ), is only one example of the feasible accomplishments to be achieved in the near future in this research area. The accelerated production of scientific knowledge related to the effects of probiotics and prebiotics on consumer health tends to result in an extraordinary increase in the range of options of this class of functional foods, especially for those individuals who wish health benefits to be associated to sensory appeal. This scenario makes this niche market highly competitive and profitable.

Acknowledgements

The authors would like to thank John Milton for the final English review.

This work was supported by Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP) (S. M. I. S., financial support grant numbers 2009/51169-0 and 2013/50506-8, R. B., fellowship grant number 2009/51168-3, R. C. R. M., fellowship grant number 2010/52267-2) and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) (S. M. I. S., fellowship grant no. 305181/2013-0). FAPESP and CNPq had no role in design, analysis or writing of this article.

R. C. R. M. and R. B. carried out the scientific literature search and wrote the review. S. M. I. S. revised the paper critically for important intellectual content, included some information and opinions and suggested several improvements.

There are no conflicts of interest.

References

1. Laparra, JP & Sanz, Y (2010) Interactions of gut microbiota with functional food components and nutraceuticals. Pharmacol Res 61, 219225.CrossRefGoogle ScholarPubMed
2. Aureli, P, Capurso, L, Castellazzi, AM, et al. (2011) Probiotics and health: an evidence-based review. Pharmacol Res 63, 366376.CrossRefGoogle ScholarPubMed
3. Ley, RE, Lozupone, CA, Hamady, M, et al. (2008) Worlds within worlds: evolution of the vertebrate gut microbiota. Nat Rev Microbiol 6, 776788.CrossRefGoogle ScholarPubMed
4. Booijink, CCGM, Zoetendal, EG, Kleerebezem, M, et al. (2007) Microbial communities in the human small intestine-coupling diversity to metagenomics. Future Microbiol 2, 285295.CrossRefGoogle ScholarPubMed
5. Zoetendal, EG, Rajilic-Stojanovic, M & De Vos, WM (2008) High throughput diversity and functionality analysis of the gastrointestinal tract microbiota. Gut 57, 16051615.CrossRefGoogle ScholarPubMed
6. Tap, J, Mondot, S, Levenez, F, et al. (2009) Towards the human intestinal microbiota phylogenetic core. Environ Microbiol 11, 25742584.CrossRefGoogle ScholarPubMed
7. Phillips, ML (2009) Gut reaction: environmental effects on the human microbiota. Environ Health Perspect 117, 198205.CrossRefGoogle ScholarPubMed
8. Turnbaugh, PJ, Hamady, M, Yatsunenko, T, et al. (2009) A core gut microbiome in obese and lean twins. Nature 457, 480484.CrossRefGoogle ScholarPubMed
9. Qin, J, Li, R, Raes, J, et al. (2010) A human gut microbial gene catalogue established by metagenomic sequencing. Nature 464, 5965.CrossRefGoogle ScholarPubMed
10. Tremaroli, V & Bäckhed, F (2012) Functional interactions between the gut microbiota and host metabolism. Nature 489, 242249.CrossRefGoogle ScholarPubMed
11. Zhang, J, Guo, Z, Xue, Z, et al. (2015) A phylo-functional core of gut microbiota in healthy young Chinese cohorts across lifestyles, geography and ethnicities. ISME J 9, 19791990.CrossRefGoogle ScholarPubMed
12. Hill, MJ (1997) Intestinal flora and endogenous vitamin synthesis. Eur J Cancer Prev 6, 4345.CrossRefGoogle ScholarPubMed
13. Pryde, SE, Duncan, SH & Hold, GL (2002) The microbiology of butyrate formation in the human colon. FEMS Microbiol Lett 217, 133139.CrossRefGoogle ScholarPubMed
14. Tappenden, KA & Deutsch, AS (2007) The physiological relevance of the intestinal microbiota – contributions to human health. J Am Coll Nutr 26, 679683.CrossRefGoogle ScholarPubMed
15. Resta, SC (2009) Effects of probiotics and commensals on intestinal epithelial physiology: implications for nutrient handling. J Physiol 587, 41694174.CrossRefGoogle ScholarPubMed
16. Saulnier, DM, Kolida, S & Gibson, GR (2009) Microbiology of the human intestinal tract and approaches for its dietary modulation. Curr Pharm Des 15, 14031414.CrossRefGoogle ScholarPubMed
17. Round, JL & Mazmanian, SK (2009) The gut microbiota shapes intestinal immune responses during health and disease. Nat Rev Immunol 9, 313323.CrossRefGoogle ScholarPubMed
18. Scanlan, PD, Shanahan, F, O’Mahony, C, et al. (2006) Culture-independent analyses of temporal variation of the dominant fecal microbiota and targeted bacterial subgroups in Crohn’s disease. J Clin Microbiol 44, 39803988.CrossRefGoogle ScholarPubMed
19. Cani, PD & Delzenne, NM (2009) The role of the gut microbiota in energy metabolism and metabolic disease. Curr Pharm Des 15, 15461558.CrossRefGoogle ScholarPubMed
20. Davis, CD & Milner, JA (2009) Gastrointestinal microflora, food components and colon cancer prevention. J Nutr Biochem 20, 743752.CrossRefGoogle ScholarPubMed
21. Caesar, R, Fak, F & Bäckhed, F (2010) Effects of gut microbiota on obesisty and atherosclerosis via modulation of inflammation and lipid metabolism. J Inter Med 268, 320328.CrossRefGoogle Scholar
22. Gerritsen, J, Smidt, H, Rijkers, GT, et al. (2011) Intestinal microbiota in human health and disease: the impact of probiotics. Genes Nutr 6, 209240.CrossRefGoogle ScholarPubMed
23. Ley, RE, Peterson, DA & Gordon, JI (2006) Ecological and evolutionary forces shaping microbial diversity in the human intestine. Cell 124, 837848.CrossRefGoogle ScholarPubMed
24. Ross, RP, Mills, S, Hill, C, et al. (2010) Specific metabolite production by gut microbiota as a basis for probiotic function. Int Dairy J 20, 269276.CrossRefGoogle Scholar
25. Scott, KP, Gratz, SW, Sheridan, PO, et al. (2013) The influence of diet on the gut microbiota. Pharmacol Res 69, 5260.CrossRefGoogle ScholarPubMed
26. O’Shea, EF, Cotter, PD, Stanton, C, et al. (2012) Production of bioactive substances by intestinal bacteria as a basis for explaining probiotic mechanisms: bacteriocins and conjugated linoleic acid. Int J Food Microbiol 152, 189205.CrossRefGoogle ScholarPubMed
27. Peterson, CT, Sharma, V, Elmén, S, et al. (2015) Immune homeostasis, dysbiosis and therapeutic modulation of the gut microbiota. Clin Exp Immunol 179, 363377.CrossRefGoogle ScholarPubMed
28. McNulty, NP, Yatsunenko, T, Hsiao, A, et al. (2011) The impact of a consortium of fermented milk strains on the gut microbiome of gnotobiotic mice and monozygotic twins. Sci Transl Med 3, 106ra106.CrossRefGoogle ScholarPubMed
29. Rauch, M & Lynch, SV (2012) The potential for probiotic manipulation of the gastrointestinal microbiome. Curr Opin Biotechnol 23, 192201.CrossRefGoogle ScholarPubMed
30. Hill, C, Guarner, F, Reid, G, et al. (2014) The International Scientific Association for Probiotics and Prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat Rev Gastroenterol Hepatol 11, 506514.CrossRefGoogle ScholarPubMed
31. Stolaki, M, De Vos, W, Kleerebezem, M, et al. (2012) Lactic acid bacteria in the gut. In Lactic Acid Bacteria: Microbiological and Functional Aspects, pp. 385401 [S Lahtinen, AC Ouwehand, S Salminen and A Von Wright, editors]. Boca Raton, FL: CRC.Google Scholar
32. Tannock, GW (1998) Studies on the intestinal microflora: a prerequisite for the development of probiotics. Int Dairy J 8, 527533.CrossRefGoogle Scholar
33. Kolida, S & Gibson, GR (2011) Synbiotic in health and disease. Annu Rev Food Sci Technol 2, 373393.CrossRefGoogle ScholarPubMed
34. Champagne, CP, Ross, RP, Saarela, M, et al. (2011) Recommendations for viability assessment of probiotics as concentrated cultures and in food matrices. Int J Food Microbiol 149, 185193.CrossRefGoogle ScholarPubMed
35. Agência Nacional De Vigilância Sanitária (Brazilian Health Surveillance Agency) (2008) Alimentos com alegações de propriedades funcionais e ou de saúde, novos alimentos/ingredientes, substâncias bioativas e probióticos: lista de alegações de propriedade funcional aprovadas. Atualizado em julho, 2008. (Foods with allegation of functional and/or health properties, new foods/ingredients, bioactive compounds and probiotics: list of allegations of approved functional properties. Up-dated in 2008) http://www.anvisa.gov.br/alimentos/comissoes/tecno_lista_alega.htm (accessed August 2009).Google Scholar
36. Health Canada (2009) Accepted claims about the nature of probiotic microorganisms in foods. http://www.hc-sc.gc.ca/fn-an/label-etiquet/claims-reclam/probiotics_claims-allegations_probiotiques-eng.php (accessed October 2014).Google Scholar
37. Ministero della Salute, Commissione unica per la nutrizione e la dietetica (2013) Guidelines on probiotics and prebiotics. http://www.salute.gov.it/imgs/C_17_pubblicazioni_1016_allegato.pdf (accessed October 2014).Google Scholar
38. Lee, YK & Salminen, S (2009) Handbook of Probiotics, 2nd ed. Hoboken, NJ: John Wiley & Sons.Google Scholar
39. Gibson, GR, Probert, HM, Van Loo, JAE, et al. (2004) Dietary modulation of the human colonic microbiota: updating the concept of prebiotics. Nutr Res Rev 17, 259275.CrossRefGoogle ScholarPubMed
40. Gibson, GR, Scott, KP, Rastall, RA, et al. (2010) Dietary prebiotics: current status and new definition. Food Sci Technol Bull: Funct Foods 7, 119.Google Scholar
41. Rastall, RA (2010) Functional oligosaccharides: application and manufacture. Annu Rev Food Sci Technol 1, 305339.CrossRefGoogle ScholarPubMed
42. Roberfroid, M (2007) Prebiotic: the concept revisited. J Nutr 137, 830837.CrossRefGoogle ScholarPubMed
43. Broekaert, WF, Courtin, CM, Verbeke, K, et al. (2011) Prebiotic and other health-related effects of cereal-derived arabinoxylans, arabinoxylan-oligosaccharides, and xylooligosaccharides. Crit Rev Food Sci Nutr 51, 178194.CrossRefGoogle ScholarPubMed
44. Crittenden, RG, Morris, LF, Harvey, ML, et al. (2001) Selection of a Bifidobacterium strain to complement resistant starch in a synbiotic yoghurt. J Appl Microbiol 90, 268278.CrossRefGoogle Scholar
45. Topping, DL, Fukushima, M & Bird, AR (2003) Resistant starch as a prebiotic and synbiotic: state of the art. Proc Nutr Soc 62, 171176.CrossRefGoogle ScholarPubMed
46. Figueroa-González, I, Quijano, G, Ramírez, G, et al. (2011) Probiotics and prebiotics – perspectives and challenges. J Sci Food Agric 91, 13411348.CrossRefGoogle ScholarPubMed
47. Topping, DL & Clifton, PM (2001) Short-chain fatty acids and human colonic function: roles of resistant starch and nonstarch polysaccharides. Physiol Rev 81, 10311064.CrossRefGoogle ScholarPubMed
48. Crittenden, R & Playne, M (2009) Prebiotics. In Handbook of Probiotics and Prebiotics, pp. 535581 [YK Lee and S Salminen, editors]. Hoboken, NJ: John Wiley & Sons.Google Scholar
49. Holzapfel, WH & Schillinger, U (2002) Introduction to pre- and probiotics. Food Res Int 35, 109116.CrossRefGoogle Scholar
50. Puupponen-Pimiä, R, Aura, AM, Oksman-Caldentey, KM, et al. (2002) Development of functional ingredients for gut health. Trends Food Sci Technol 13, 311.CrossRefGoogle Scholar
51. Mattila-Sandholm, T, Myllärinen, P, Crittenden, R, et al. (2002) Technological challenges for future probiotic foods. Int Dairy J 12, 173182.CrossRefGoogle Scholar
52. Bielecka, M, Biedrzycka, E & Majkowska, A (2002) Selection of probiotics and prebiotics for synbiotics and confirmation of their in vivo effectiveness. Food Res Int 35, 125131.CrossRefGoogle Scholar
53. Sanders, ME & Marco, ML (2010) Food formats for effective delivery of probiotics. Annu Rev Food Sci Technol 1, 6585.CrossRefGoogle ScholarPubMed
54. Rolfe, RD (2000) The role of probiotic cultures in the control of gastrointestinal health. J Nutr 130, 396402.CrossRefGoogle ScholarPubMed
55. Saad, SMI (2006) Probióticos e prebióticos: o estado da arte (Probiotics and prebiotics: the state of the art). Revista Brasileira de Ciências Farmacêuticas 42, 116.CrossRefGoogle Scholar
56. de Vrese, M & Marteau, PR (2007) Probiotics and prebiotics: effects on diarrhea. J Nutr 137, 803811.CrossRefGoogle ScholarPubMed
57. Saad, N, Delattre, C, Urdaci, M, et al. (2013) An overview of the last advances in probiotic and prebiotic field. LWT – Food Sci Technol 50, 116.CrossRefGoogle Scholar
58. Floch, MH, Walker, WA, Madsen, K, et al. (2011) Recommendations for probiotic use: 2011 update. J Clin Gastroenterol 45, 168171.CrossRefGoogle ScholarPubMed
59. Guandalini, S (2011) Probiotics for prevention and treatment of diarrhea. J Clin Gastroenterol 45, 149153.CrossRefGoogle ScholarPubMed
60. Wolvers, D, Antoine, JM, Myllyluoma, E, et al. (2010) Guidance for substantiating the evidence for beneficial effects of probiotics: prevention and management of infections by probiotics. J Nutr 140, 698712.CrossRefGoogle ScholarPubMed
61. Arvola, T, Laiho, K, Torkkeli, S, et al. (1999) Prophylactic Lactobacillus GG reduces antibiotic-associated diarrhea in children with respiratory infections: a randomized study. Pediatrics 104, e64.CrossRefGoogle ScholarPubMed
62. Vanderhoof, JA, Whitney, DB, Antonson, DL, et al. (1999) Lactobacillus GG in the prevention of antibiotic-associated diarrhea in children. J Pediatr 135, 564568.CrossRefGoogle ScholarPubMed
63. Marchand, J & Vandenplas, Y (2000) Microorganisms administered in the benefit of the host: myths and facts. Eur J Gastroenterol Hepatol 12, 10771088.CrossRefGoogle ScholarPubMed
64. Dendukuri, N, Costa, V, McGregor, M, et al. (2005) Probiotic therapy for the prevention and treatment of Clostridium difficile-associated diarrhea: a systematic review. CMAJ 173, 167170.CrossRefGoogle ScholarPubMed
65. Roberfroid, M, Gibson, GR, Hoyles, L, et al. (2010) Prebiotic effects: metabolic and health benefits. Br J Nutr 104, 163.CrossRefGoogle ScholarPubMed
66. Thompson, I (2008) Clostridium difficile-associated disease: update and focus on non-antibiotic strategies. Age Ageing 37, 1418.CrossRefGoogle ScholarPubMed
67. Kale-Pradhan, PB, Jassal, HK, et al. (2010) Role of Lactobacillus in the prevention of antibiotic-associated diarrhea: a meta-analysis. Pharmacotherapy 30, 119126.CrossRefGoogle ScholarPubMed
68. Hickson, M, D’Souza, AL, Muthu, N, et al. (2007) Use of probiotic Lactobacillus preparation to prevent diarrhoea associated with antibiotics: randomised double blind placebo controlled trial. BMJ 335, 80.CrossRefGoogle ScholarPubMed
69. Pozzoni, P, Riva, A, Bellatore, AG, et al. (2012) Saccharomyces boulardii for the prevention of antibiotic-associated diarrhea in adult hospitalized patients: a single-center, randomized, double-blind, placebo-controlled trial. Am J Gastroenterol 107, 922931.CrossRefGoogle ScholarPubMed
70. Gorbach, SL (2000) Probiotics and gastrointestinal health. Am J Gastroenterol 95, 1418.CrossRefGoogle ScholarPubMed
71. Goldin, BR & Gorbach, SL (2008) Clinical indications for probiotics: an overview. Clin Infect Dis 46, 96100.CrossRefGoogle ScholarPubMed
72. Graul, T, Cain, AM & Karpa, KD (2009) Lactobacillus and bifidobacteria combinations: a strategy to reduce hospital-acquired Clostridium difficile diarrhea incidence and mortality. Med Hypotheses 73, 194198.CrossRefGoogle ScholarPubMed
73. Thomas, MR, Litin, SC, Osmon, DR, et al. (2001) Lack of effect of Lactobacillus GG on antibiotic-associated diarrhea: a randomized, placebo-controlled trial. Mayo Clin Proc 76, 883889.CrossRefGoogle ScholarPubMed
74. Balakrishnan, M & Floch, MH (2012) Prebiotics, probiotics and digestive health. Curr Opin Clin Nutr Metab Care 15, 580585.CrossRefGoogle ScholarPubMed
75. European Food Safety Authority (2010) Scientific opinion on the substantiation of a health claim related to fermented milk containing Lactobacillus casei DN-144 001 plus yoghurt symbiosis (Actimel®), and reduction of Clostridium difficile toxins in the gut of patients receiving antibiotics and reduction risk of acute diarrhoea in patients receiving antibiotics pursuant to Article 14 of Regulation (EC) No 1924/2006. EFSA J 8, 1903.Google Scholar
76. Lewis, S, Burmeister, S & Brazier, J (2005) Effect of the prebiotic oligofructose on relapse of Clostridium difficile-associated diarrhoea: a randomized, controlled study. Clin Gastroenterol Hepatol 3, 442448.CrossRefGoogle ScholarPubMed
77. Parashar, UD, Gibson, CJ, Bresse, JS, et al. (2006) Rotavirus and severe childhood diarrhea. Emerg Infect Dis 12, 304316.CrossRefGoogle ScholarPubMed
78. Salim, AF, Phillips, AD & Farthing, MJ (1990) Pathogenesis of gut virus infection. Baillières Clin Gastroenterol 4, 593607.CrossRefGoogle ScholarPubMed
79. Kaila, M, Isolauri, E, Soppi, E, et al. (1992) Enhancement of the circulating antibody secreting cell response in human diarrhea by a human Lactobacillus strain. Pediatr Res 32, 141144.CrossRefGoogle ScholarPubMed
80. Saavedra, JM, Bauman, NA, Oung, L, et al. (1994) Feeding of Bifidobacterium bifidum and Streptococcus thermophilus to infants in hospital for prevention of diarrhoea and shedding of rotavirus. Lancet 344, 10461049.CrossRefGoogle ScholarPubMed
81. Sugita, T & Togawa, M (1994) Efficacy of Lactobacillus preparation Biolactis powder in children with rotavirus enteritis. Jpn J Pediatr 47, 27552762.Google Scholar
82. Shornnikova, AV, Casas, I, Mykkanen, H, et al. (1997) Bacteriotherapy with Lactobacillus reuteri in rotavirus gastroenteritis. Pediatr Infect Dis J 16, 11031107.CrossRefGoogle Scholar
83. van Baarlen, P, Troost, F, Van der Meer, C, et al. (2011) Human mucosal in vivo transcriptome responses to three lactobacilli indicate how probiotics may modulate human cellular pathways. Proc Natl Acad Sci U S A 108, 45624569.CrossRefGoogle ScholarPubMed
84. Hilton, E, Kolakowski, P, Singer, C, et al. (1997) Efficacy of Lactobacillus GG as a diarrheal preventive in travelers. J Travel Med 4, 4143.CrossRefGoogle ScholarPubMed
85. Oksanen, PJ, Salminen, S, Saxelin, M, et al. (1990) Prevention of travellers’ diarrhoea by Lactobacillus GG. Ann Med 22, 5356.CrossRefGoogle ScholarPubMed
86. Lyra, A, Lahtinen, S & Ouwehand, AC (2012) Gastrointestinal benefits of probiotics: clinical evidence. In Lactic Acid Bacteria: Microbiological and Functional Aspects, pp. 509523 [S Lahtinen, AC Ouwehand, S Salminen and A Von Wright, editors]. Boca Raton, FL: CRC.Google Scholar
87. McFarland, LV (2007) Meta-analysis of probiotics for the prevention of traveler’s diarrhea. Travel Med Infect Dis 5, 97105.CrossRefGoogle ScholarPubMed
88. Drakoularakou, A, Tzortis, G, Rastall, RA, et al. (2010) A double-blind, placebo-controlled, randomized human study assessing the capacity of a novel galacto-oligosaccharide mixture in reducing traveller’s diarrhoea. Eur J Clin Nutr 64, 146152.CrossRefGoogle Scholar
89. Cummings, JH, Christie, S & Cole, TJ (2001) A study of fructo oligosaccharides in the prevention of traveller’s diarrhoea. Aliment Pharmacol Ther 15, 11391145.CrossRefGoogle Scholar
90. Virk, A, Mandrekar, J, Berbari, EF, et al. (2013) A randomized, double blind, placebo-controlled trial of an oral synbiotic (AKSB) for prevention of traveler’s diarrhea. J Travel Med 20, 8894.CrossRefGoogle Scholar
91. Vouloumanou, EK, Makris, GC, Karageorgopoulos, DE, et al. (2009) Probiotics for the prevention of respiratory tract infections: a systematic review. Int J Antimicrob Agents 34, 197.CrossRefGoogle ScholarPubMed
92. Thompson, M, Vodicka, TA, Blair, PS, et al. (2013) Duration of symptons of respiratory tract infections in children: systematic review. BMJ 347, f7027.CrossRefGoogle Scholar
93. Snow, V, Mottur-Pilson, C & Gonzales, R (2001) American College of Physicians-American Society of Internal Medicine, American Academy of Family Physicians, Centers for Disease Control, & Infectious Diseases Society of America. Principles of appropriate antibiotic use for treatment of nonspecific upper respiratory tract infections: background. Ann Inter Med 134, 487489.CrossRefGoogle Scholar
94. Aguilar, L, Giménez, MJ & Barberán, J (2010) Drug resistance in community-acquired respiratory tract infections: role for an emerging antibacterial. Infect Drug Resist 3, 3543.CrossRefGoogle ScholarPubMed
95. Cangemi de Gutierrez, R, Santos, V & Nader-Macias, ME (2001) Protective effect of intranasally inoculated Lactobacillus fermentum against Streptococcus pneumoniae challenge on the mouse respiratory tract. FEMS Immunol Med Microbiol 31, 187195.CrossRefGoogle ScholarPubMed
96. Yasui, H, Kiyoshima, J & Hori, T (2004) Reduction of influenza virus titer and protection against influenza virus infection in infant mice fed Lactobacillus casei Shirota. Clin Diagn Lab Immunol 11, 675679.Google ScholarPubMed
97. Racedo, S, Villena, J, Medina, M, et al. (2006) Lactobacillus casei administration reduces lung injuries in a Streptococcus pneumoniae infection in mice. Microbes Infect 8, 23592366.CrossRefGoogle Scholar
98. King, S, Glanville, J, Sanders, ME, et al. (2014) Effectiveness of probiotics on the duration of illness in healthy children and adults who develop common acute respiratory infectious conditions: a systematic review and meta-analysis. Br J Nutr 112, 4154.CrossRefGoogle ScholarPubMed
99. Arslanoglu, S, Moro, GE & Boehm, G (2007) Early supplementation of prebiotic oligosaccharides protects formula-fed infants against infections during the first 6 months of life. J Nutr 137, 24202424.CrossRefGoogle ScholarPubMed
100. Luoto, R, Ruuskanen, O, Waris, M, et al. (2014) Prebiotic and probiotic supplementation prevents rhinovirus infections in preterm infants: a randomized, placebo-controlled trial. J Allergy Clin Immunol 133, 405413.CrossRefGoogle ScholarPubMed
101. Puccio, G, Cajozzo, C, Meli, F, et al. (2007) Clinical evaluation of a new formula for infants containing live Bifidobacterium longum BL999 and prebiotics. Nutrition 23, 18.CrossRefGoogle ScholarPubMed
102. World Health Organization (2013) Cardiovascular disease fact sheet n. 317. http://www.who.int/mediacentre/factsheets/fs317/en/index.html (accessed February 2013).Google Scholar
103. Grundy, SM, Cleeman, JI, Merz, CN, et al. (2004) Implications of recent clinical trials for the National Cholesterol Education Program Adult Treatment Panel III guidelines. Circulation 110, 227239.CrossRefGoogle ScholarPubMed
104. Houston, MC, Fazio, S, Chilton, FH, et al. (2009) Nonpharmacologic treatment of dyslipidemia. Prog Cardiovasc Dis 52, 6194.CrossRefGoogle ScholarPubMed
105. Mann, GV & Spoerry, A (1974) Studies of a surfactant and cholesterolemia in the Maasai. Am J Clin Nutr 27, 464469.CrossRefGoogle Scholar
106. Ooi, LG & Liong, MT (2010) Cholesterol-lowering effects of probiotics and prebiotics: a review of in vivo and in vitro findings. Int J Mol Sci 11, 24992522.CrossRefGoogle ScholarPubMed
107. Zhuang, G, Liu, XM, Zhang, QX, et al. (2012) Research advances with regards to clinical outcome and potential mechanisms of the cholesterol-lowering effects of probiotics. Clin Lipidol 7, 501507.CrossRefGoogle Scholar
108. Kumar, M, Nagpal, R, Kumar, R, et al. (2012) Cholesterol-lowering probiotics as potential biotherapeutics for metabolic diseases. Exp Diabetes Res 2012, Article ID 902917.CrossRefGoogle ScholarPubMed
109. Davis, HR Jr, Zhu, LJ, Hoos, LM, et al. (2004) Niemann-Pick C1 Like 1 (NPC1L1) is the intestinal phytosterol and cholesterol transporter and a key modulator of whole-body cholesterol homeostasis. J Biol Chem 279, 3358633592.CrossRefGoogle Scholar
110. Jeun, J, Kim, S, Cho, SY, et al. (2010) Hypocholesterolemic effects of Lactobacillus plantarum KCTC3928 by increased bile acid excretion in C57BL/6 mice. Nutrition 26, 321330.CrossRefGoogle ScholarPubMed
111. Park, YH, Kim, JG, Shin, YW, et al. (2007) Effect of dietary inclusion of Lactobacillus acidophilus ATCC 43121 on cholesterol metabolism in rats. J Microbiol Biotechnol 17, 655662.Google ScholarPubMed
112. Agerholm-Larsen, L, Raben, A, Haulrik, N, et al. (2000) Effect of 8 week intake of probiotic milk products on risk factors for cardiovascular diseases. Eur J Clin Nutr 54, 288297.CrossRefGoogle ScholarPubMed
113. Naruszewicz, M, Johansson, ML, Zapolska-Downar, D, et al. (2002) Effect of Lactobacillus plantarum 299v on cardiovascular disease risk factors in smokers. Am J Clin Nutr 76, 12491255.CrossRefGoogle ScholarPubMed
114. Ataie-Jafari, A, Larijani, B, Majd, HA, et al. (2009) Cholesterol-lowering effects of probiotic yogurt in comparison with ordinary yogurt in mildly to moderately hypercholesterolemic subjects. Ann Nutr Metab 54, 2227.CrossRefGoogle ScholarPubMed
115. Jones, ML, Martoni, CJ, Parent, M, et al. (2012) Cholesterol-lowering efficacy of a microencapsulated bile salt hydrolase-active Lactobacillus reuteri NCIMB 30242 yoghurt formulation in hypercholesterolaemic adults. Br J Nutr 107, 15051513.CrossRefGoogle ScholarPubMed
116. Bedani, R, Rossi, EA, Cavallini, DCU, et al. (2015) Influence of daily consumption of synbiotic soy-based product supplemented with okara soybean by-product on risk factors for cardiovascular diseases. Food Res Int 73, 142148.CrossRefGoogle Scholar
117. de Roos, NM, Schouten, G & Katan, MB (1999) Yogurt enriched with Lactobacillus acidophilus does not lower blood lipids in healthy men and women with normal to borderline high serum cholesterol levels. Eur J Clin Nutr 53, 277280.CrossRefGoogle Scholar
118. Simons, LA, Armansec, SG & Conway, P (2006) Effect of Lactobacillus fermentum on serum lipids in subjects with elevated serum. Nutr Metab Cardiovasc Dis 16, 531535.CrossRefGoogle ScholarPubMed
119. Balcázar-Muñoz, BR, Martínez-Abundis, E & Gonzáles-Ortiz, M (2003) Effect of oral inulin administration on lipid profile and insulin sensitivity in subjects with obesity and dyslipidemia. Revista Médica de Chile 131, 597604.Google ScholarPubMed
120. Nichenametla, SN, Weidauer, LA, Wey, HE, et al. (2014) Resistant starch type 4-enriched diet lowered blood cholesterols and improved body composition in a double blind controlled cross-over intervention. Mol Nutr Food Res 58, 13651369.CrossRefGoogle Scholar
121. Giacco, R, Clemente, G, Luongo, D, et al. (2004) Effects of short-chain fructo-oligosaccharides on glucose and lipid metabolism in mild hypercholesterolaemic individuals. Clin Nutr 23, 331340.CrossRefGoogle ScholarPubMed
122. Kellow, NJ, Coughlan, MT & Reid, CM (2014) Metabolic benefits of dietary prebiotics in human subjects: a systematic review of randomised controlled trials. Br J Nutr 111, 11471161.CrossRefGoogle ScholarPubMed
123. Rault-Nania, MH, Gueux, E, Demougeot, C, et al. (2006) Inulin attenuates atherosclerosis in apolipoprotein E-deficient mice. Br J Nutr 96, 840844.CrossRefGoogle ScholarPubMed
124. Ranganna, K, Yatsu, FM, Hayes, BE, et al. (2000) Butyrate inhibits proliferation-induced proliferating cell nuclear antigen expression (PCNA) in rat vascular smooth muscle cells. Mol Cell Biochem 205, 149161.CrossRefGoogle ScholarPubMed
125. Tang, WH & Hazen, SL (2014) The contributory role of gut microbiota in cardiovascular disease. J Clin Invest 124, 42044211.CrossRefGoogle ScholarPubMed
126. Lavanda, I, Saad, SML, Lobo, AR, et al. (2011) Prebióticos y su efecto en la biodisponibilidad del calcio (Prebiotics and their effect on calcium bioavailability). Revista de Nutrição 24, 333344.CrossRefGoogle Scholar
127. Pedreschi, R, Campos, D, Noratto, G, et al. (2003) Andean yacon root (Smalllanthus sonchifolius Poepp. Endl) fructooligosaccharides as a potential novel source of prebiotics. J Agric Food Chem 51, 52785284.CrossRefGoogle ScholarPubMed
128. Raschka, L & Daniel, H (2005) Mechanisms underlying the effects of inulin-type fructans on calcium absorption in the large intestine of rats. Bone 37, 728735.CrossRefGoogle ScholarPubMed
129. Kruger, MC, Brown, KE, Collett, G, et al. (2003) The effect of fructooligosaccharides with various degrees of polymerization on calcium bioavailability in the growing rat. Exp Biol Med 228, 683688.CrossRefGoogle ScholarPubMed
130. Coxam, V (2005) Inulin-type fructans and bone health: state of the art and perspectives in the management of osteoporosis. Br J Nutr 93, 111123.CrossRefGoogle ScholarPubMed
131. Lobo, AR, Mancini-Filho, J, Alvares, EP, et al. (2009) Effects of dietary lipid composition and inulin-type fructans on mineral bioavailability in growing rats. Nutrition 25, 216225.CrossRefGoogle ScholarPubMed
132. Scholz-Ahrens, KE, Ade, P, Marten, B, et al. (2007) Prebiotics, probiotics, and synbiotics affect mineral absorption, bone mineral content, and bone structure. J Nutr 137, 838846.CrossRefGoogle ScholarPubMed
133. van der Heuvel, EG, Muijs, T, Van Dokkum, W, et al. (1999) Lactulose stimulates calcium absorption in postmenopausal women. J Bone Miner Res 14, 12111216.CrossRefGoogle Scholar
134. van der Heuvel, EG, Schoterman, MH & Muijs, T (2000) Transgalactooligosaccharides stimulate calcium absorption in postmenopausal women. J Nutr 130, 29382942.CrossRefGoogle Scholar
135. Coudray, C, Bellanger, J, Castiglia-Delavaud, C, et al. (1997) Effect of soluble or partly soluble dietary fibres supplementation on absorption and balance of calcium, magnesium, iron and zinc in healthy young man. Eur J Clin Nutr 51, 375380.CrossRefGoogle ScholarPubMed
136. Tahiri, M, Tressol, JC, Arnaud, J, et al. (2003) Effect of short-chain fructooligosaccharides on intestinal calcium absorption and calcium status in postmenopausal women: a stable-isotope study. Am J Clin Nutr 77, 449457.CrossRefGoogle ScholarPubMed
137. Noack, J, Dongowski, G, Hartmann, L, et al. (2000) The human gut bacteria Bacteroides thetaiotaomicron and Fusobacterium varium produce putrescine and spermidine in cecum of pectin-fed gnotobiotic rats. J Nutr 130, 12251231.CrossRefGoogle ScholarPubMed
138. Bongers, A & Van Den Heuvel, EGHM (2003) Prebiotics and the bioavailability of minerals and trace elements. Food Rev Int 19, 397422.CrossRefGoogle Scholar
139. Scholz-Ahrens, KE & Schrezenmeir, J (2007) Inulin and oligofructose and mineral metabolism: the evidence from animal trials. J Nutr 137, 25132523.CrossRefGoogle Scholar
140. Reid, G (2008) Probiotic lactobacilli for urogenital health in women. J Clin Gastroenterol 42, 234236.CrossRefGoogle ScholarPubMed
141. Gardiner, GE, Heinemann, C, Baroja, ML, et al. (2002) Oral administration of the probiotic combination Lactobacillus rhamnosus GR-1 and L. fermentum RC-14 for human intestinal applications. Int Dairy J 12, 191196.CrossRefGoogle Scholar
142. Antonio, MA, Hawes, SE & Hillier, SL (1999) The identification of vaginal Lactobacillus species and the demographic and microbiologic characteristics of women colonized by these species. J Infect Dis 180, 19501956.CrossRefGoogle ScholarPubMed
143. Martin, HLB, Richardson, BA, Nyange, PM, et al. (1999) Vaginal lactobacilli, microbial flora, and risk of human immunodeficiency virus type 1 and sexually transmitted disease acquisition. J Infect Dis 180, 18631868.CrossRefGoogle ScholarPubMed
144. Donders, GGG, Bosmans, E, Dekeersmaecker, A, et al. (2000) Pathogenesis of abnormal vaginal flora. Am J Obstet Gynecol 182, 872878.CrossRefGoogle Scholar
145. McLean, NW & McGroarty, JA (1996) Growth inhibition of metronidazole-susceptible and metronidazole-resistant strains of Gardnerella vaginalis by lactobacilli in vitro . Appl Environ Microbiol 62, 10891092.CrossRefGoogle ScholarPubMed
146. Reid, G (2001) Probiotic agents to protect the urogenital tract against infection. Am J Clin Nutr 73, 437443.CrossRefGoogle ScholarPubMed
147. Lepargneur, JP & Rousseau, V (2002) Protective role of the Doderleïn flora. J Gynecol Obstet Biol Reprod 31, 485494.Google ScholarPubMed
148. Valore, EV, Park, CH, Igreti, SL, et al. (2002) Antimicrobial components of vaginal fluid. Am J Obstet Gynecol 187, 561568.CrossRefGoogle ScholarPubMed
149. Martin, R, Soberon, N, Vasquez, F, et al. (2008) Vaginal microbiota: composition, protective role, associated pathologies, and therapeutic perspectives. Enferm Infecc Microbiol Clín 26, 160167.CrossRefGoogle ScholarPubMed
150. Reid, G & Burton, J (2002) Use of Lactobacillus to prevent infection by pathogenic bacteria. Microbes Infect 4, 319324.CrossRefGoogle ScholarPubMed
151. Reid, G & Bocking, A (2003) The potential for probiotics to prevent bacterial vaginosis and preterm labor. Am J Obstet Gynecol 189, 12021208.CrossRefGoogle ScholarPubMed
152. Reid, G & Bruce, AW (2003) Urogenital infections in women: can probiotics help? J Postgrad Med 79, 428432.CrossRefGoogle ScholarPubMed
153. Reid, G, Charbonneau, D, Erb, J, et al. (2003) Oral use of Lactobacillus rhamnosus GR-1 and L. fermentum RC-14 significantly alters vaginal flora: randomized, placebo-controlled trial in 64 healthy women. FEMS Immunol Med Microbiol 35, 131134.CrossRefGoogle ScholarPubMed
154. Reid, G (2012) Probiotic and prebiotic applications for vaginal health. J AOAC Int 95, 3134.CrossRefGoogle ScholarPubMed
155. Bruce, AW & Reid, G (1988) Intravaginal instillation of lactobacilli for prevention of recurrent urinary tract infections. Can J Microbiol 34, 339343.CrossRefGoogle ScholarPubMed
156. Bruce, AW, Reid, G, McGroarty, JA, et al. (1992) Preliminary study on the prevention of recurrent urinary tract infections in ten adult women using intravaginal lactobacilli. Int Urogynecol J 3, 2225.CrossRefGoogle Scholar
157. Martinez, RCR, Franceschini, SA, Patta, MC, et al. (2009) Improved treatment of vulvovaginal candidiasis with fluconazole plus probiotic Lactobacillus rhamnosus GR-1 and Lactobacillus reuteri RC-14. Lett Appl Microbiol 48, 269274.CrossRefGoogle ScholarPubMed
158. Anukam, K, Osazuwa, E, Ahonkhai, I, et al. (2006) Augmentation of antimicrobial metronidazole therapy of bacterial vaginosis with oral probiotic Lactobacillus rhamnosus GR-1 and Lactobacillus reuteri RC-14: randomized, double-blind, placebo controlled trial. Microbes Infect 8, 14501454.CrossRefGoogle ScholarPubMed
159. Martinez, RCR, Franceschini, SA, Patta, MC, et al. (2009) Improved cure of bacterial vaginosis with single dose of tinidazole (2g) and Lactobacillus rhamnosus GR-1 and Lactobacillus reuteri RC-14: a randomized, double-blind, placebo-controlled trial. Can J Microbiol 55, 133138.CrossRefGoogle Scholar
160. Williams, AB, Yu, C, Tashima, K, et al. (2001) Evaluation of two self-care treatments for prevention of vaginal candidiasis in women with HIV. J Gen Microbiol 12, 5157.Google ScholarPubMed
161. Pirotta, M, Gunn, J, Chondros, P, et al. (2004) Effect of Lactobacillus in preventing post-antibiotic vulvovaginal candidiasis: a randomized controlled trial. BMJ 329, 548.CrossRefGoogle Scholar
162. Larsson, PG, Stray-Pedersen, B, Ryttig, KR, et al. (2008) Human lactobacilli as supplementation of clindamycin to patients with bacterial vaginosis reduce the recurrence rate; a 6-month, double-blind, randomized, placebo-controlled study. BMC Womens Health 8, 3.CrossRefGoogle Scholar
163. Eriksson, K, Carlsson, B, Forsum, U, et al. (2005) A double-blind treatment study of bacterial vaginosis with normal vaginal lactobacilli after an open treatment with vaginal clindamycin ovules. Acta Derm Venereol 85, 4246.CrossRefGoogle ScholarPubMed
164. Bizzini, B, Pizzo, G, Scapagnini, G, et al. (2012) Probiotics and oral health. Curr Pharma Des 18, 55225531.CrossRefGoogle ScholarPubMed
165. Selwitz, RH, Ismail, AI & Pitts, NB (2007) Dental caries. Lancet 369, 5159.CrossRefGoogle ScholarPubMed
166. Islam, B, Khan, S & Khan, A (2007) Dental caries: from infection to prevention. Med Sci Monit 13, 196203.Google ScholarPubMed
167. Bonifait, L, Chandad, F & Grenier, D (2009) Probiotics for oral health: myth or reality? J Can Dent Assoc 75, 585590.Google ScholarPubMed
168. Ahola, AJ, Yli-Knnuuttila, H, Suomalainen, T, et al. (2002) Short-term consumption of probiotic-containing cheese and its effect on dental caries risk factors. Arch Oral Biol 47, 799804.CrossRefGoogle ScholarPubMed
169. Nikawa, H, Makihira, S, Fukushima, H, et al. (2004) Lactobacillus reuteri in bovine milk fermented decreases the oral carriage of mutans streptococci. Int J Food Microbiol 95, 219223.CrossRefGoogle ScholarPubMed
170. Kang, MS, Chung, J, Kim, SM, et al. (2006) Effect of Weissella cibaria isolates on the formation of Streptococcus mutans biofilm. Caries Res 40, 418425.CrossRefGoogle ScholarPubMed
171. Houle, MA & Grenier, D (2003) Maladies parodontales: connaissances actuelles (Current concepts in periodontal diseases). Méd Maladies Infect 33, 331340.CrossRefGoogle Scholar
172. Sookkhee, S, Chulasiri, M & Prachyabrued, W (2001) Lactic acid bacteria from healthy oral cavity of Thai volunteers: inhibition of oral pathogens. J Appl Microbiol 90, 172179.CrossRefGoogle ScholarPubMed
173. Koll-Klais, P, Mandar, R, Leibur, E, et al. (2005) Oral lactobacilli in chronic periodontitis and periodontal health: species composition and antimicrobial activity. Oral Microbiol Immunol 20, 354361.CrossRefGoogle ScholarPubMed
174. Krasse, P, Carisson, B, Dahl, C, et al. (2006) Decreased gum bleeding and reduced gingivitis by the probiotic Lactobacillus reuteri . Swed Dent J 30, 5560.Google ScholarPubMed
175. Riccia, DN, Bizzini, F, Perilli, MG, et al. (2007) Anti-inflammatory effects of Lactobacillus brevis (CD2) on periodontal disease. Oral Dis 13, 376385.CrossRefGoogle ScholarPubMed
176. Laleman, I & Teughels, W (2015) Probiotics in the dental practice: a review. Quintessence Int 46, 255264.Google ScholarPubMed
177. Scully, C & Greenman, J (2008) Halitosis (breath odor). Periodontology 48, 6675.CrossRefGoogle ScholarPubMed
178. Burton, JP, Chicott, CN & Tagg, JR (2005) The rationale and potential for the reduction of oral malodour using Streptococcus salivarius probiotics. Oral Dis 11, 2931.CrossRefGoogle ScholarPubMed
179. Keller, MK, Bardow, A, Jensdottir, T, et al. (2012) Effect of chewing gums containing the probiotic bacterium Lactobacillus reuteri on oral malodour. Acta Odontol Scand 70, 246250.CrossRefGoogle ScholarPubMed
180. Schabussova, I & Wiedermann, U (2008) Lactic acid bacteria as novel adjuvant systems for prevention and treatment of atopic diseases. Curr Opin Allergy Clin Immunol 8, 557564.CrossRefGoogle ScholarPubMed
181. Penders, J, Stobberingh, EE, Van Den Brandt, PA, et al. (2007) The role of the intestinal microbiota in the development of atopic disorders. Allergy 62, 12231236.CrossRefGoogle ScholarPubMed
182. Prescott, SL & Bjorksten, B (2007) Probiotics for the prevention or treatment of allergic diseases. J Allergy Clin Immunol 120, 255262.CrossRefGoogle ScholarPubMed
183. Bjorksten, B, Sepp, E, Julge, K, et al. (2001) Allergy development and the intestinal microflora during the first year of life. J Allergy Clin Immunol 108, 516520.CrossRefGoogle ScholarPubMed
184. Kalliomaki, M, Kirjavainen, P, Eerola, E, et al. (2001) Distinct patterns of neonatal gut microflora in infants in whom atopy was and was not dveloping. J Allergy Clin Immunol 107, 129134.CrossRefGoogle ScholarPubMed
185. Sepp, E, Julge, K, Mikelsaar, M, et al. (2005) Intestinal microbiota and immunoglobulin E responses in 5-year-old Estonian children. Clin Exp Allergy 35, 11411146.CrossRefGoogle ScholarPubMed
186. Kukkonen, K, Savilahti, E, Haahtela, T, et al. (2007) Probiotics and prebiotic galactooligosaccharides in the prevention of allergic diseases: a randomized, double-blind, placebo-controlled trial. J Allergy Clin Immunol 119, 192198.CrossRefGoogle Scholar
187. Marschan, E, Kuitunen, M, Kukkonen, K, et al. (2008) Probiotics in infancy induce protective immune profiles that are characteristic for chronic low-grade inflammation. Clin Exp Allergy 38, 611618.CrossRefGoogle ScholarPubMed
188. Rautava, S, Kalliomaki, M & Isolauri, E (2002) Probiotics during pregnancy and breast-feeding might confer immunomodulatory protection against topic disease in the infant. J Allergy Clin Immunol 109, 119121.CrossRefGoogle Scholar
189. Kopp, MV, Hennemuth, I, Heinzmann, A, et al. (2008) Randomized, double-blind, placebo-controlled trial of probiotics for primary prevention: no clinical effects of Lactobacillus GG supplementation. Pediatrics 121, 850856.CrossRefGoogle ScholarPubMed
190. Kim, SO, Ah, YM, Yu, YM, et al. (2014) Effects of probiotics for the treatment of atopic dermatitis: a meta-analysis of randomized controlled trials. Ann Allergy Asthma Immunol 113, 217226.CrossRefGoogle ScholarPubMed
191. Martinez-Canavate, A, Sierra, S, Lara-Villoslada, F, et al. (2009) A probiotic dairy product containing L. gasseri CECT5714 and L. coryniformis CEC5711 induces immunological changes in children suffering from allergy. Pediatr Allergy Immunol 20, 592600.CrossRefGoogle ScholarPubMed
192. Majamaa, H & Isolauri, E (1997) Probiotics: a novel approach in the management of food allergy. J Allergy Clin Immunol 99, 179185.CrossRefGoogle ScholarPubMed
193. Isolauri, E, Rautava, S, Kalliomaki, M, et al. (2002) Role of probiotics in food hypersensitivity. Curr Opini Allergy Clin Immunol 2, 263271.CrossRefGoogle ScholarPubMed
194. Osborn, DA & Sinn, JKH (2009) Probiotics in infants for prevention of allergic disease and food hypersensitivity. Cochrane Database of Systematic Reviews, issue 4, CD006475, http://onlinelibrary.wiley.com/doi/10.1002/14651858.CD006475.pub2/epdf.Google Scholar
195. Wells, JM & Mercenier, A (2008) Mucosal delivery of therapeutic and prophylactic molecules using lactic acid bacteria. Nat Rev Microbiol 6, 349362.CrossRefGoogle ScholarPubMed
196. Sashihara, T, Sueki, N & Ikegami, S (2006) An analysis of the effectiveness of heat-killed lactic acid bacteria in alleviating allergic diseases. J Dairy Sci 89, 28462855.CrossRefGoogle ScholarPubMed
197. Medina, M, Izquierdo, E, Ennahar, S, et al. (2007) Differential immunomodulatory properties of Bifidobacterium longum strains: relevance to probiotic selection and clinical applications. Clin Exp Immunol 150, 531538.CrossRefGoogle ScholarPubMed
198. Kekkonen, RA, Kajasto, E, Miettinen, M, et al. (2008) Probiotic Leuconostoc mesenteroides spp. cremoris and Streptococcus thermophilus induce IL-2 and IFN-gamma production. World J Gastroenterol 14, 11921203.CrossRefGoogle Scholar
199. Foligne, B, Zoumpopoulou, G, Dewulf, J, et al. (2007) A key role of dendritic cells in probiotic functionality. PLoS ONE 2, e313.CrossRefGoogle ScholarPubMed
200. Blander, JM & Medzhitov, R (2006) Toll-dependent selection of microbial antigens for presentation by dendritic cells. Nature 440, 808812.CrossRefGoogle ScholarPubMed
201. Fujiwara, D, Inoue, S, Wakabayashi, H, et al. (2004) The anti-allergic effects of lactic acid bacteria are strain dependent and mediated by effects on both Th1/Th2 cytokine expression and balance. Int Arch Allergy Immunol 135, 205215.CrossRefGoogle ScholarPubMed
202. Kalliomäki, M, Antoine, JM, Herz, U, et al. (2010) Guidance for substantiating the evidence for beneficial effects of probiotics: prevention and management of allergic diseases by probiotics. J Nutr 140, 713721.CrossRefGoogle ScholarPubMed
203. Niele, N, van Zwol, A, Westerbeek, EAM, et al. (2013) Effect of non-human neutral and acidic oligosaccharides on allergic and infectious diseases in preterm infants. Eur J Pediatr 172, 317323.CrossRefGoogle ScholarPubMed
204. Schouten, B, Van Esch, BC, Kormelink, TG, et al. (2011) Non-digestible oligosaccharides reduce immunoglobulin free light-chain concentrations in infants at risk for allergy. Pediatr Allergy Immunol 22, 537542.CrossRefGoogle ScholarPubMed
205. Williams, HC & Grindlay, DJ (2010) What’s new in atopic eczema? An analysis of systematic reviews published in 2007 and 2008. Part 1: definitions, causes and consequences of eczema. Clin Exp Dermatol 35, 1215.CrossRefGoogle ScholarPubMed
206. Osborn, DA & Sinn, JKH (2013) Prebiotics in infants for prevention of allergy. Cochrane Database of Systematic Reviews, issue 28, CD006474. http://onlinelibrary.wiley.com/doi/10.1002/14651858.CD006474.pub3/epdf.Google ScholarPubMed
207. Jonkers, D, Penders, J, Masclee, A, et al. (2012) Probiotics in the management of inflammatory bowel disease. Drugs 72, 803823.CrossRefGoogle ScholarPubMed
208. Kawada, M, Arihiro, A & Mizoguchi, E (2007) Insights from advances in research of chemically induced experimental models of human inflammatory bowel disease. World J Gastroenterol 13, 55815593.CrossRefGoogle Scholar
209. Abraham, C & Medzhitov, R (2011) Interaction between the host innate immune system and microbes in inflammatory bowel disease. Gastroenterology 140, 17291737.CrossRefGoogle ScholarPubMed
210. Danese, S (2011) Immune and nonimmune components orchestrate the pathogenesis of inflammatory bowel disease. Am J Physiol Gastrointest Liver Physiol 300, 716722.CrossRefGoogle ScholarPubMed
211. Dignass, A, Van Assche, G, Lindsay, JO, et al. (2010) The second European evidence-based consensus on the diagnosis and management of Crohn’s disease: current management. J Crohn’s Colitis 4, 2862.CrossRefGoogle ScholarPubMed
212. Lebeer, S, Vanderleyden, J & De Keersmaecker, SC (2008) Genes and molecules of lactobacilli supporting probiotic action. Microbiol Mol Biol Rev 72, 728764.CrossRefGoogle ScholarPubMed
213. Oelschlaeger, TA (2010) Mechanisms of probiotics actions – a review. Int J Med Microbiol 300, 5762.CrossRefGoogle ScholarPubMed
214. Borchers, AT, Selmi, V, Meyers, FJ, et al. (2009) Probiotics and immunity. J Gastroenterol 44, p. 2646.CrossRefGoogle ScholarPubMed
215. Caballero-Franco, C, Keller, K, De Simone, C, et al. (2007) The VSL#3 probiotic formula induces mucin gene expression and secretion in colonic epithelial cells. Am J Physiol Gastrointest Liver Physiol 291, 315322.CrossRefGoogle Scholar
216. Karczewski, J, Troost, FJ, Konings, I, et al. (2010) Regulation of human epithelial tight junction proteins by Lactobacillus plantarum in vivo and protective effects on the epithelial barrier. Am J Physiol Gastrointest Liver Physiol 298, 851859.CrossRefGoogle ScholarPubMed
217. Ng, SC, Plamondon, S, Kamm, MA, et al. (2010) Immunosuppressive effects via human intestinal dendritic cells of probiotic bacteria and steroids in the treatment of acute ulcerative colitis. Inflamm Bowel Dis 16, 12861298.CrossRefGoogle ScholarPubMed
218. Rahimi, R, Nikfar, S, Rahimi, F, et al. (2008) A meta-analysis on the efficacy of probiotics for maintenance of remission and prevention of clinical and endoscopic relapse in Crohn’s disease. Dig Dis Sci 53, 25242531.CrossRefGoogle ScholarPubMed
219. Shen, J, Zuo, ZX & Mao, AP (2014) Effect of probiotics on inducing remission and maintaining therapy in ulcerative colitis, Crohn’s disease, and pouchitis: meta-analysis of randomized controlled trials. Inflamm Bowel Dis 20, 2135.CrossRefGoogle ScholarPubMed
220. Rijkers, GT, Bengmark, S, Enck, P, et al. (2010) Guidance for substantiating the evidence for beneficial effects of probiotics: current status and recommendations for future research. J Nutr 140, 671676.CrossRefGoogle ScholarPubMed
221. Moayyedi, P, Ford, AC, Talley, NJ, et al. (2010) The efficacy of probiotics in the treatment of irritable bowel syndrome: a systematic review. Gut 59, 325332.CrossRefGoogle ScholarPubMed
222. Spiller, R, Aziz, Q, Creed, F, et al. (2007) Guidelines on the irritable bowel syndrome: mechanisms and practical management. Gut 56, 17701798.CrossRefGoogle ScholarPubMed
223. Clarke, G, Cryan, JF, Dinan, TG, et al. (2012) Review article: probiotics for the treatment of irritable bowel syndrome – focus on lactic acid bacteria. Aliment Pharmcol Ther 35, 403413.CrossRefGoogle ScholarPubMed
224. Brenner, DM, Moeller, MJ, Chey, WD, et al. (2009) The utility of probiotics in the treatment of irritable bowel syndrome: a systematic review. Am J Gastroenterol 104, 10331049.CrossRefGoogle ScholarPubMed
225. Parkes, GC, Sanderson, JD & Whelan, K (2010) Treating irritable bowel syndrome with probiotics: the evidence. Proc Nutr Soc 69, 187194.CrossRefGoogle ScholarPubMed
226. Horvath, A, Dziechciarz, P & Szajewska, H (2011) Meta-analysis: Lactobacillus rhamnosus GG for abdominal pain-related functional gastrointestinal disorders in childhood. Aliment Pharmcol Ther 33, 13021310.CrossRefGoogle ScholarPubMed
227. Ringel, Y & Ringel-Kulka, T (2011) The rationale and clinical effectiveness of probiotics in irritable bowel syndrome. J Clin Gastroenterol 45, 145148.CrossRefGoogle ScholarPubMed
228. Whelan, K (2014) The importance of systematic reviews and meta-analyses of probiotics and prebiotics. Am J Gastroenterol 109, 15631565.CrossRefGoogle ScholarPubMed
229. Haller, D, Antoine, JM, Bengmark, S, et al. (2010) Guidance for substantiating the evidence for beneficial effects of probiotics: probiotics in chronic inflammatory bowel disease and the functional disorder irritable bowel syndrome. J Nutr 140, 690697.CrossRefGoogle ScholarPubMed
230. O’Mahony, L, Mccarthy, J, Kelly, P, et al. (2005) Lactobacillus and Bifidobacterium in irritable bowel syndrome: symptom responses and relationship to cytokine profiles. Gastroenterology 128, 541551.CrossRefGoogle ScholarPubMed
231. Whorwell, PJ, Altringer, L, Morel, J, et al. (2006) Efficacy of an encapsulated probiotic Bifidobacterium infantis 35624 in women with irritable bowel syndrome. Am J Gastroenterol 101, 15811590.CrossRefGoogle ScholarPubMed
232. Bonfrate, L, Tack, J, Grattagliano, I, et al. (2013) Microbiota in health and irritable bowel syndrome: current knowledge, perspectives and therapeutic options. Scand J Gastroenterol 48, 9951009.CrossRefGoogle ScholarPubMed
233. Hunter, JO, Tuffnell, Q & Lee, AJ (1999) Controlled trial of oligofructose in the management of irritable bowel syndrome. J Nutr 129, 14511453.CrossRefGoogle ScholarPubMed
234. Olesen, M & Gudmand-Hoyer, E (2000) Efficacy, safety, and tolerability of fructooligosaccharides in the treatment of irritable bowel syndrome. Am J Clin Nutr 72, 15701575.CrossRefGoogle ScholarPubMed
235. Silk, DBA, Davis, A, Vulevic, J, et al. (2009) Clinical trial: the effects of a trans-galactooligosaccharide prebiotic on faecal microbiota and symptoms in irritable bowel syndrome. Aliment Pharmcol Ther 29, 508518.CrossRefGoogle ScholarPubMed
236. Andriulli, A, Neri, M, Loguercio, C, et al. (2008) Clinical trial on the efficacy of a new symbiotic formulation, Flortec, in patients with irritable bowel syndrome: a multicenter, randomized study. J Clin Gastroenterol 42, 218223.CrossRefGoogle ScholarPubMed
237. Guandalini, S, Cernat, E & Moscoso, D (2015) Prebiotics and probiotics in irritable bowel syndrome and inflammatory bowel disease in children. Benef Microbes 6, 209217.CrossRefGoogle ScholarPubMed
238. Floch, MH (2014) Recommendations for probiotic use in humans – a 2014 update. Pharmaceuticals 7, 9991007.CrossRefGoogle ScholarPubMed
239. Lesbros-Pantoflickova, D, Corthesy-Theulaz, I & Blum, AL (2007) Helicobacter pylori and probiotics. J Nutr 137, 812818.CrossRefGoogle ScholarPubMed
240. Michetti, P, Dorta, G, Wiesel, PH, et al. (1999) Effect of whey-based culture supernatant of Lactobacillus acidophilus (johnsonii) La1 on Helicobacter pylori infection in humans. Digestion 60, 203209.CrossRefGoogle ScholarPubMed
241. Ruggiero, P (2014) Use of probiotics in the fight against Helicobacter pylori . World J Gastrointest Pathophysiol 5, 384391.CrossRefGoogle ScholarPubMed
242. Felley, CP, Corthesy-Theulaz, I, Rivero, JL, et al. (2001) Favourable effect of an acidified milk (LC-1) on Helicobacter pylori gastritis in man. Eur J Gastroenterol Hepatol 13, 2529.CrossRefGoogle ScholarPubMed
243. Pantoflickova, D, Corthesy-Theulaz, I, Dorta, G, et al. (2003) Favorable effect of long-term intake of fermented milk containing Lactobacillus johnsonii on H. pylori associated gastritis. Aliment Pharmacol Ther 18, 805813.CrossRefGoogle Scholar
244. Wang, KY, Li, SN, Liu, CS, et al. (2004) Effects of ingesting Lactobacillus- and Bifidobacterium-containing yogurt in subjects with colonized Helicobacter pylori . Am J Clin Nutr 80, 737741.Google ScholarPubMed
245. Lionetti, E, Miniello, VL, Castellaneta, SP, et al. (2006) Lactobacillus reuteri therapy to reduce side-effects during anti-Helicobacter pylori treatment in children: a randomized placebo controlled trial. Aliment Pharmacol Ther 24, 14611468.CrossRefGoogle ScholarPubMed
246. Francavilla, R, Polimeno, L, Demichina, A, et al. (2014) Lactobacillus reuteri strain combination in Helicobacter pylori infection – a randomized, double-blind, placebo-controlled study. J Clin Gastroenterol 48, 407413.CrossRefGoogle ScholarPubMed
247. Yang, YJ, Chuang, CC, Yang, HB, et al. (2012) Lactobacillus acidophilus ameliorates H. pylori induced gastric inflammation by inactivating the Smad7 and NFκB pathways. BMC Microbiol 12, 38.CrossRefGoogle ScholarPubMed
248. Chen, X, Liu, XM, Tian, F, et al. (2012) Antagonistic activities of lactobacilli against Helicobacter pylori growth and infection in human gastric epithelial cells. J Food Sci 77, 914.CrossRefGoogle ScholarPubMed
249. Navarro-Rodriguez, T, Silva, FM, Barbuti, RC, et al. (2013) Association of a probiotic to a Helicobacter pylori eradication regimen does not increase efficacy or decreases the adverse effects of the treatment: a randomized, double-blind, placebo-controlled study. BMC Gastroenterol 13, 56.CrossRefGoogle ScholarPubMed
250. Szajewska, H, Horvath, A & Piwowarczyk, A (2010) Meta-analysis: the effects of Saccharomyces boulardii supplementation on Helicobacter pylori eradication rates and side effects during treatment. Aliment Pharmacol Ther 32, 10691079.CrossRefGoogle ScholarPubMed
251. Zhang, MM, Qian, W, Qin, YY, et al. (2015) Probiotics in Helicobacter pylori eradication therapy: a systematic review and meta-analysis. World J Gastroenterol 21, 43454357.CrossRefGoogle ScholarPubMed
252. Katan, MB (2012) Why the European Food Safety Authority was right to reject health claims on probiotics. Benef Microbes 3, 8589.CrossRefGoogle ScholarPubMed
253. Glanville, J, King, S, Guarner, F, et al. (2015) A review of the systematic review process and its applicability for use in evaluating evidence for health claims on probiotic foods in the European Union. J Nutr 14, 16.CrossRefGoogle ScholarPubMed
254. Bisanz, JE, Enos, MK, Mwanga, JR, et al. (2014) Randomized open-label pilot study of the influence of probiotics and the gut microbiome on toxic metal levels in Tanzanian pregnant women and school children. MBio 5, e01580-14.CrossRefGoogle ScholarPubMed
Figure 0

Table 1 Beneficial effects of probiotics and/or prebiotics on important human pathologic conditions