Next Article in Journal
Inhibition of Growth and Ammonia Production of Ruminal Hyper Ammonia-Producing Bacteria by Chinook or Galena Hops after Long-Term Storage
Next Article in Special Issue
Malting and Brewing Industries Encounter Fusarium spp. Related Problems
Previous Article in Journal
Green Biorefinery of Giant Miscanthus for Growing Microalgae and Biofuel Production
Previous Article in Special Issue
Effect of Saccharomyces, Non-Saccharomyces Yeasts and Malolactic Fermentation Strategies on Fermentation Kinetics and Flavor of Shiraz Wines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Probiotic Delivery through Fermentation: Dairy vs. Non-Dairy Beverages

by
Chaminda Senaka Ranadheera
1,*,
Janak K. Vidanarachchi
2,
Ramon Silva Rocha
3,
Adriano G. Cruz
3 and
Said Ajlouni
1
1
School of Agriculture & Food, Faculty of Veterinary & Agricultural Sciences, The University of Melbourne, Melbourne VIC 3010, Australia
2
Department of Animal Science, Faculty of Agriculture, University of Peradeniya, Peradeniya 20400, Sri Lanka
3
Department of Food, Federal Institute of Rio de Janeiro (IFRJ), Food Department, CEP 20270-021 Rio de Janeiro, RJ, Brazil
*
Author to whom correspondence should be addressed.
Fermentation 2017, 3(4), 67; https://doi.org/10.3390/fermentation3040067
Submission received: 10 November 2017 / Revised: 24 November 2017 / Accepted: 27 November 2017 / Published: 11 December 2017
(This article belongs to the Special Issue Microbiota of Fermented Beverages)

Abstract

:
Probiotics are live microorganisms that, when administered in adequate amounts, confer a health benefit on the host, mainly through the process of replacing or including beneficial bacteria in the gastrointestinal tract. Fermented dairy foods such as yogurt, fermented milk and cheese are the major vehicle in delivering probiotics, and probiotic delivery have been traditionally associated with these fermented dairy foods. Additionally, many other non-dairy probiotic products and non-food form such as capsules, pills and tablets are also available and some of these non-food forms are highly popular among the consumers. Certain non-dairy probiotic foods, especially beverages that are non-fermented products, can also play an important role in probiotic delivery. There is an increasing demand for non-dairy probiotic foods (both fermented and non-fermented) including fruit and vegetable juices, soy and certain cereal products due to vegetarianism, lactose intolerance and dairy allergies, as well as interest in low cholesterol foods. In this context, this review mainly focus on the different types of probiotic food products including beverages with special reference to their viability followed by a brief account on the applicability of using fermented and non-fermented beverage products in probiotic delivery.

Graphical Abstract

1. Probiotics: A Brief Overview

In recent years, probiotic foods have received a considerable attention among health-conscious consumers. According to the Food and Agriculture Organization of the United Nations and the World Health Organization [1], probiotics are defined as live microbial cultures of a single strain or mixture of different strains that beneficially affect the host animal, either directly or indirectly, by improving its intestinal microbial balance. Utilization of beneficial microorganisms in health promotion is not new, and in fact they have been consumed by humans, especially in the form of fermented dairy foods, for many years [2]. In the early 1990s, Noble Laureate, Elie Metchnikoff (1845–1916) observed exceptionally long healthy living among Bulgarians who regularly consumed soured/fermented dairy products, and then first documented the modern concept of probiotics in his book “The Prolongation of Life” [3]. Since then the use of probiotics in developing functional foods has gained a wide popularity in the world mainly due to the interest in gaining health benefits through consumption of probiotic fortified food products. The most common genera that have been used and possess probiotic characteristics are the lactic acid bacteria Bifidobacterium and Lactobacillus. These genera are mostly given the generally-recognized-as-safe (GRAS) status, which indicates no or less health risks to the host upon consumption [2,4,5]. Few other microorganisms mainly bacteria and some yeast have also been utilized as probiotics (Table 1). However, there are some concerns regarding the safety of some probiotic genera such as Enterococcus, since they can be pathogenic, causing illness in the host.
The human gastrointestinal tract contains trillions of microorganisms, consisting of up to 1000 or more different bacterial species, collectively known as the gut microbiota. The gut microbiota plays an important role in host health, influencing the maturation of the immune system and regulating energy metabolism [11]. In general, it is accepted that intake of probiotics contributes to the enhancement and maintenance of well-balanced intestinal microbiota. Many evidences support the use of these probiotics in prevention and treating diseases and health disorders such as high blood pressure & serum cholesterol, lactose intolerance [12] and many gastrointestinal disorders (irritable bowel syndrome, Crohn’s disease, peptic ulcers, antibiotic associated diarrhea, etc.) [13,14,15,16]. Probiotics also possess anti-carcinogenic effects [17,18,19] and enhance the immune system [20,21]. Some examples of probiotic potential for therapeutic applications have been listed in Table 2.
For probiotic bacteria in foods to be beneficial in the host, they should be able to survive gastric transit and reach the small intestine in sufficient numbers to be effective. Hence, in order to provide health benefits to the host, probiotics should maintain minimum therapeutic level/minimum viability level (106–107 cfu/mL or g of carrier food product) at the time of consumption, possess the ability to tolerate harsh gastric and intestinal conditions (including acid, bile and enzymes) and be able to attach to the gut epithelium [38,39]. A potentially successful probiotic strain is expected to have several more desirable properties (Table 3) and these characteristics may influence its potential for the commercial applications.
Probiotics can exert their effects by one or more actions, e.g., creation of a restrictive physiological environment for potentially pathogenic microorganisms. These effects are achieved by lowering the pH through production of organic acids, such as lactate and short chain volatile fatty acids, due to break-down of complex carbohydrates [44] or elaboration of antibiotic-like substances such as bacteriocin-like compounds [45]. Certain probiotic bacterial species, such as lactic acid bacteria, can adhere to the intestinal epithelium and thereby prevent invasion by pathogenic bacteria such as Escherichia coli, Salmonella and Clostridium spp. in the gut epithelium and this phenomenon is known as competitive exclusion. Feeding probiotics may also help to modulate the cellular and humoral immune system thereby enhancing the host’s resistance to enteric pathogens [38,46].

2. Dairy vs. Non-Dairy Food Matrices

Dairy products such as yogurts, fermented sour milk and cheese remain at the forefront of probiotic food development at present. Although fermented dairy foods can be considered as one of the most common as well as the traditional modes of delivering of probiotics to humans, at present, many non-dairy as well as non-traditional and convenient probiotic products, such as capsules, have been developed and commercialized in many countries [47]. Soy products, cereal based products, fruit and vegetable juices, and fermented meat and fish products can be considered as main non-dairy probiotic foods available in the market at present. There are many different types and brands of non-dairy probiotic foods as well. The diversity of probiotic food products is summarized in Table 4. Many studies have clearly indicated that the type of carrier foods could affect not only the viability of probiotics during processing and storage, but also on their functional properties, such as susceptibility to adverse conditions in the gut (acidity, bile and various enzymes), capacity to adhere to gut epithelium and immunomodulation [39,48,49]. The incorporation of probiotics into dairy foods may aid in tolerating harsh gastro-intestinal condition better than that of non-dairy carrier foods, as the buffering action of milk as well as milk fat, might protect probiotics in such conditions by reducing their direct exposure to harsh conditions [2]. Dairy foods rich in milk fat, such as ice cream, were found to be more effective in enhancing the survivability and bile acid tolerance of probiotics [50]. However, the physical structure of non-dairy probiotic carrier foods such as vegetables (for example artichokes and olives) might provide protective environment for probiotics and reduce their exposure to harsh gastrointestinal conditions as well [51]. Sausage matrix and microstructure have also shown a potential in retaining the viability of probiotics through gastrointestinal transit [52,53].
Nevertheless, both dairy and non-dairy products may contain various other ingredients such as prebiotics (ingredients which ferment in the latter part of the gastrointestinal tract and stimulate the growth and activity of beneficial gut microbes) that could interact with probiotics to alter their functional properties [105]. Presence of these substances could be specific to certain carrier foods. For example, naturally, milk does not contain inulin (plant derived polysaccharide with prebiotic properties), yet certain root vegetables/rhizomes, fruits and cereals (artichoke, oat, bananas, garlic, onions, leeks) contain high level of inulin. Nevertheless, there is also a possibility of production of probiotic dairy products by incorporating various prebiotics such as inulin or its breakdown products (fructooligosaccharides and oligofructans). Prebiotic oligosaccharides are essentially obtained by one of three processes: direct extraction of natural oligosaccharides from plants, controlled hydrolysis of natural plant polysaccharides, and enzymatic synthesis, using hydrolases and/or glycosyltransferases [106]. Apart from the direct prebiotic activities, many plant and microbial derived oligosaccharides help to deliver the probiotic organisms to the target sites. The encapsulation of probiotic organism with such compounds prevents the gastrointestinal digestion of the probiotic organism enabling them to be present at large intestine which is the target site of probiotics [107]. Prebiotics from plant sources such as Arrowroot (Maranta arundinacea) carbohydrates and Raftilose® (commercially available inulin) can be used to enhance the survivability of Lactobacillus sp. and lactic acid bacteria in bio-yoghurt during refrigerated storage [108]. These prebiotic substances may aid probiotics to survive through the gastrointestinal transit and colonize in the large intestinal epithelium. Another study [46] revealed that supplementation of broiler chicken feed with specific prebiotic compounds supported the growth of specific probiotic Lactobacillus spp. such as L. johnsonii in their ileum and caeca. Therefore, careful selection of probiotics, prebiotics and carrier food matrices (both dairy and non-dairy) when produce probiotic foods is essential in maximizing the functional efficacy of probiotics during manufacturing, storage and upon ingestion. It seems likely that in most cases the carrier food matrices possess synergistic effect on probiotic microorganisms during processing as well as in the gastrointestinal environment [2]. However, questions about such synergistic effects may arise when non-food probiotic carriers such as capsules are used.
Many probiotics with potential health benefits have be isolated from the gastrointestinal tract of healthy humans (human origin). There are some non-human sources as well. Raw cow’s milk is considered as an excellent source of probiotic bacteria [109]. Usually heat treatments such as pasteurization and sterilization destroy these beneficial microorganisms in raw milk, however better synergistic effect may be expected through re-introduction of such probiotics into milk when produce dairy products mainly due to their dairy origin. Efficacy of dairy origin probiotics when incorporated into non-dairy foods could be significantly affected and in line with this, more research is needed to discover the effect of functional properties of dairy origin probiotics when incorporate into non-dairy food matrices including beverages. However, the development of probiotic containing dairy products is not always easy and requires the overcoming of certain technological intrinsic requirements related to their processing stages. For example, selection of probiotic strains withstanding freezing is essential in production of probiotic ice cream despite its origin [50].
Although there are many benefits of having probiotics with dairy foods, non-dairy probiotic foods also play a significant role in human health. For instance, there are major drawbacks related to dairy foods such as allergy, lactose intolerance and cholesterol content, hence non-dairy probiotic foods are beneficial for the people having such health disorders. Furthermore, cultural (strict vegans) as well as specific religious believes among certain communities may also limit the consumption of dairy foods. In such situations, non-dairy probiotic carrier foods and convenient mode of deliveries such as tablets could be the only way of providing probiotics. Each probiotic strain is unique in many aspects such as optimum growth conditions and growth medium. Consequently, the selection of probiotics which perform better or equally in non-dairy foods compared with dairy foods may be useful in developing non-dairy probiotic food [110,111,112,113]. In general, most non-dairy probiotics are beverages. In this context, it could be argued that non-dairy probiotic carriers including beverages may be equally importance as dairy related carriers in terms of human health and nutrition. Some commercially available non-dairy probiotic beverages are shown in Table 5.

3. Fermented vs. Non-Fermented Beverages

Fermentation technology is one of the oldest food technology applications and fermented products are the result of the metabolic activity of a complex microbiota, consisting of the naturally occurring indigenous microorganisms, and/or selected microorganisms such as bacteria and yeasts which inoculated as starter cultures. Fermentation of food products helps their preservation due to the organic acid production as well as imparting them pleasant sensory properties and additional nutritional values [52]. In terms of probiotic beverage production, fermentation process is not compulsory (Figure 1). There are many types of probiotic fermented milk in the global market produced under various brand names. Major physicochemical properties of probiotic fermented milk products vary basically on the type of probiotic microorganism, type of milk and use of other starter cultures in the product. In addition, fermented probiotic dairy products vary in their textures ranging from liquid drinks such as acidophilus milk and kefir to semi-solid/ropy or firm products such as drinking yogurt and villi [115]. Microorganisms used in starter cultures are of great industrial significance since they play a vital role in flavour and textural development of fermented food products. These starter cultures may not necessarily possess probiotic properties. For example, the term “probiotics” may not be suitable for yogurt starter cultures (Streptococcus thermophilus and Lactobacillus delbrueckii spp. bulgaricus) due to their poor survival in the digestive tract [6]. However, some beneficial health promoting effects of yogurt starter cultures including improved lactose utilization and enhancement of immune system have also been reported [116,117]. In addition to the starter culture microorganisms, various probiotics can be added during the production of fermented food products including beverages to achieve the therapeutic benefits. Having starter cultures in probiotic products may provide benefits as well as some disadvantages. For example, starter cultures may create a suitable environment for probiotic growth during yogurt manufacturing. Yogurt starter culture bacteria, in particular S. thermophiles, are also identified as oxygen scavengers and thus may be beneficial in improving the growth and viability of anaerobic probiotics. These starter cultures were previously demonstrated to complete the fermentation of milk within 5–10 h and utilised most of the oxygen in milk [118]. In contrast, variations in the starter cultures and probiotic combinations may also influence the probiotic viability in the final product due to antagonistic or symbiotic relationships [37]. However, other options have been showing interesting, as the addition of enzymes [119] and should be considered for dairy and non-dairy processors to guarantee the viability of the probiotic strain during commercial shelf-life of the products.
The majority of the probiotic dairy beverages in the present-day market can be categorized into fermented products (Figure 1). Based on the type of microorganisms involved in the fermentation, these dairy beverages can be classified into different categories (Table 6). Many non-dairy probiotic carriers such as fruit and vegetable juices are however produced mainly without fermentation since fermentation may cause undesirable sensory properties in such products [10] due to various factors such as acidity development, changes in viscosity, texture and colour. Production of fermented and non-fermented fruit and vegetable juices can be very similar with only one additional step of fermentation when manufacture fermented juices (Figure 2). In terms of achieving desirable food characteristics, fermentation is a complex process and selecting appropriate conditions such as optimum temperature during fermentation is a critical parameter that must be considered to prevent lethal or sub-lethal damages to the probiotic cells during the processing and subsequent storage. These conditions affect the biomass yield as well [121]. Duration of fermentation (full or partial fermentation) may also affect the quality of the final product. Partial fermentation of dairy foods could results drinking dairy beverages and recently there has been high demand for such products [56]. Further, probiotics in fermented foods may demonstrate better stability in the product as fermentation time can provide them an opportunity to grow and stabilize well [50]. However, despite the variations in the production process and possible disparities in product’s physico-chemical, sensory, nutritional and therapeutic properties, modern health conscious consumers have a strong demand for both fermented and non-fermented probiotic food products, especially for probiotic beverages [10,122] and this fact can be an advantage for increase the sales of these beverage products.

4. Conclusions

Fermented dairy products remain at the forefront of probiotic delivery. However, the variety of non-dairy foods both fermented and non-fermented nature is available and these products also play a significant role in delivering probiotics to humans. Generally, most probiotic beverages, in particular those with fruit and vegetable origin are non-fermented formulations. The efficacy of probiotics when delivered through fermented vs. non-fermented status of a particular carrier food matrix with special reference to beverages has not been studied thoroughly and in order to reap the maximum benefits of probiotics, further research in this aspect is needed. Simultaneously, sensory tests should be applied to evaluate the consumer’s acceptance of these beverage products, providing more optimized formulations.

Author Contributions

Chaminda Senaka Ranadheera conceived and drafted the review. Janak K. Vidanarachchi, Ramon Silva Rocha, Adriano G. Cruz and Said Ajlouni revised the manuscript. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. FAO/WHO. Joint FAO/WHO Working Group Report on Drafting Guidelines for the Evaluation of Probiotics in Food; G. P. Putnams’s Sons: London, ON, Canada, 2002. [Google Scholar]
  2. Ranadheera, R.D.C.S.; Baines, S.K.; Adams, M.C. Importance of food in probiotic efficacy. Food Res. Int. 2010, 43, 1–7. [Google Scholar] [CrossRef]
  3. Metchnikoff, E. The Prolongation of Life; G. P. Putnams’s Sons: New York, NY, USA, 1907. [Google Scholar]
  4. Salminen, S.; Wrightb, A.; Morellic, L.; Marteaud, P.; Brassarte, B.; Vosf, W.M.; Fondeng, R.; Saxelinh, M.; Collinsi, K.; Mogensenj, G.; et al. Demonstration of safety of probiotics—A review. Int. J. Food Microbiol. 1998, 44, 93–106. [Google Scholar] [CrossRef]
  5. Penner, R.; Fedorak, R.N.; Madsen, K.L. Probiotics and nutraceuticals: non-medicinal treatments of gastrointestinal diseases. Curr. Opin. Pharmacol. 2005, 5, 596–603. [Google Scholar] [CrossRef] [PubMed]
  6. Senok, A.C.; Ismaeel, A.Y.; Botta, G.A. Probiotics: Facts and myths. Clin. Microbiol. Infect. 2005, 11, 958–966. [Google Scholar] [CrossRef] [PubMed]
  7. Shah, N.P. Functional cultures and health benifits. Int. Dairy J. 2007, 17, 1262–1277. [Google Scholar] [CrossRef]
  8. Sari, F.N.; Dizdar, E.A.; Oguz, S.; Erdeve, O.; Uras, N.; Dilmen, U. Oral probiotics: Lactobacillus sporogenes for prevention of necrotizing enterocolitis in very low-birth weight infants: A randomized, controlled trial. Eur. J. Clin. Nutr. 2011, 65, 434–439. [Google Scholar] [CrossRef] [PubMed]
  9. Caplan, M.; Frost, B. Myth: Necrotizing enterocolitis: Probiotics will end the disease, and surgical intervention improves the outcome. Semin. Fetal Neonatal Med. 2011, 16, 264–268. [Google Scholar] [CrossRef] [PubMed]
  10. Ranadheera, C.S.; Prasanna, P.H.P.; Vidanarachchi, J.K. Fruit juice as probiotic carriers. In Fruit Juices: Types, Nutritional Composition and Health Benefits; Elder, K.E., Ed.; Nova Science Publishers: New York, NY, USA, 2014; pp. 253–268. [Google Scholar]
  11. Marques, T.M.; Cryan, J.F.; Shanahan, F.; Fitzgerald, G.F.; Ross, R.P.; Dinan, T.G.; Stanton, C. Gut microbiota modulation and implications for host health: Dietary strategies to influence the gut–brain axis. Innov. Food Sci. Emerg. Technol. 2014, 22, 239–247. [Google Scholar]
  12. Rasic, J.L. Microflora of the intestine: Probiotics. In Encyclopedia of Food Sciences and Nutrition; Caballero, B., Trugo, L., Finglas, P., Eds.; Academic Press: Oxford, UK, 2003; pp. 3911–3916. [Google Scholar]
  13. Kim, H.J.; Roque, M.I.V.; Camilleri, M.; Stephens, D.; Burton, D.D.; Baxter, K.; Thomforde, G.; Zinsmeister, A.R. A randomized controlled trial of a probiotic combination VSL# 3 and placebo in irritable bowel syndrome with bloating. Neurogastroenterol. Motil. 2005, 17, 687–696. [Google Scholar] [PubMed]
  14. Hickson, M.; Souza, A.L.D.; Muthu, N.; Rogers, T.R.; Want, S.; Rajkumar, C.; Bulpitt, C.J. Use of probiotic Lactobacillus preparation to prevent diarrhoea associated with antibiotics: Randomised double blind placebo controlled trial. BMJ 2007, 335, 80. [Google Scholar] [CrossRef] [PubMed]
  15. Cabré, E.; Gassull, M.A. Probiotics for preventing relapse or recurrence in Crohn’s disease involving the ileum: Are there reasons for failure? J. Crohns Colitis 2007, 1, 47–52. [Google Scholar] [CrossRef] [PubMed]
  16. Quigley, E.M.M. Gut microbiota and the role of probiotics in therapy. Curr. Opin. Pharmacol. 2011, 11, 593–603. [Google Scholar] [CrossRef] [PubMed]
  17. Zhu, Y.; Luo, T.M.; Jobin, C.; Young, H.A. Gut microbiota and probiotics in colon tumorigenesis. Cancer Lett. 2011, 309, 119–127. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, C.F.; Pan, T.M. In Vitro Effects of Lactic acid bacteria on cancer cell viability and antioxidant activity. J. Food Drug Anal. 2010, 18, 77–86. [Google Scholar]
  19. Liu, C.T.; Chu, F.J.; Chou, C.C.; Yu, R.C. Antiproliferative and anticytotoxic effects of cell fractions and exopolysaccharides from Lactobacillus casei 01. Mutat. Res. 2011, 721, 157–162. [Google Scholar] [CrossRef] [PubMed]
  20. Lollo, P.C.B.; Moura, C.S.; Morato, P.N.; Cruz, A.G.; Castro, W.F.; Betim, C.B.; Nisishima, L.; Faria, J.A.F.; Junior, M.M.; et al. Probiotic yogurt offers higher immune-protection than probiotic whey beverage. Food Res. Int. 2013, 54, 118–124. [Google Scholar] [CrossRef]
  21. Prasanna, P.H.P.; Grandison, A.S.; Charalampopoulos, D. Bifidobacteria in milk products: An overview of physiological and biochemical properties, exopolysaccharide production, selection criteria of milk products and health benefits. Food Res. Int. 2014, 55, 247–262. [Google Scholar] [CrossRef]
  22. Kołodziej, M.; Szajewska, H. Lactobacillus reuteri DSM 17938 in the prevention of antibiotic-associated diarrhoea in children: Protocol of a randomised controlled trial. BMJ Open 2017, 7, e013928. [Google Scholar] [CrossRef] [PubMed]
  23. Oksanen, P.J.; Salminen, S.; Saxelin, M.; Hamalainen, P.; Ihanto-Vormisto, A.; Muurasniemi-Isovita, L.; Nikkari, S.; Oksanen, T.; Porsti, I.; Salminen, E.; et al. Prevention of travellers diarrhoea by Lactobacillus GG. Ann. Med. 1990, 22, 53–56. [Google Scholar] [CrossRef] [PubMed]
  24. Kajander, K.; Hatakka, K.; Poussa, T.; Farkkila, M.; Koppela, R. A probiotic mixture alleviates symptoms in irritable bowel syndrome patients: A controlled 6-month intervention. Aliment. Pharmacol. Ther. 2005, 22, 387–394. [Google Scholar] [CrossRef] [PubMed]
  25. Guyonnet, D.; Chassany, O.; Ducrotte, P.; Picard, C.; Mouret, M.; Mercier, C.H.; Matuchansky, C. Effect of a fermented milk containing Bifidobacterium animalis DN-173 010 on the health-related quality of life and symptoms in irritable bowel syndrome in adults in primary care: A multicentre, randomized, double-blind, controlled trial. Aliment. Pharmacol. Ther. 2007, 26, 475–486. [Google Scholar] [CrossRef] [PubMed]
  26. Brigidi, P.; Vitali, B.; Swennen, E.; Bazzocchi, G.; Matteuzzi, D. Effects of probiotic administration upon the composition and enzymatic activity of human fecal microbiota in patients with irritable bowel syndrome or functional diarrhea. Res. Microbiol. 2001, 152, 735–741. [Google Scholar] [CrossRef]
  27. Nobaek, S.; Johansson, M.L.; Molin, G.; Ahrne, S.; Jeppsson, B. Alteration of intestinal microflora is associated with reduction in abdominal bloating and pain in patients with irritable bowel syndrome. Am. J. Gastroenterol. 2000, 95, 1231–1238. [Google Scholar] [CrossRef] [PubMed]
  28. Marteau, P.; Cuillerier, E.; Meance, S.; Gerhardt, M.F.; Myara, A.; Bouvier, M.; Couley, C.; Tondu, F.; Bommelaer, G.; Grimaud, J.C. Bifidobacterium animalis strain DN-173 010 shortens the colonic transit time in healthy women: A double-blind, randomized, controlled study. Aliment. Pharmacol. Ther. 2002, 16, 587–593. [Google Scholar] [CrossRef] [PubMed]
  29. Campieri, M.; Rizzello, F.; Venturi, A.; Poggioli, G.; Ugolini, F.; Helwing, U.; Amadini, C.; Romboli, E.; Gionchetti, P. Combination of antibiotic and probiotic treatment is efficacious in prophylaxis of post-operative recurrence of Crohn’s disease: A randomized controlled study vs mesalamine. Gastroenterology 2000, 118, A781. [Google Scholar] [CrossRef]
  30. Kruis, W.; Fric, P.; Pokrotnieks, J.; Lukas, M.; Fixa, B.; Kamm, M.A.; Weismueller, J.; Beglinger, C.; Stolte, M.; et al. Maintaining remission of ulcerative colitis with the probiotic Escherichia coli Nissle 1917 is as effective as with standard mesalazine. Gut 2004, 53, 1617–1623. [Google Scholar] [CrossRef] [PubMed]
  31. Laake, K.O.; Bjorneklett, A.; Aamodt, G.; Aabakken, L.; Jacobsen, M.; Bakka, A.; Vatn, M.H. Outcome of four weeks’ intervention with probiotics on symptoms and endoscopic appearance after surgical reconstruction with a J-configurated ileal-pouch-anal-anastomosis in ulcerative colitis. Scand. J. Gastroenterol. 2005, 40, 43–51. [Google Scholar] [CrossRef] [PubMed]
  32. Anukam, K.; Osazuwa, E.; Ahonkhai, I.; Ngwu, M.; Osemene, G.; Bruce, A.W.; Reid, G. 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. 2006, 8, 1450–1454. [Google Scholar] [CrossRef] [PubMed]
  33. Abrahamsson, T.R.; Jakobsson, T.; Bottcher, M.F.; Fredrikson, M.; Jenmalm, M.C.; Bjorksten, B.; Oldaeus, G. Probiotics in prevention of IgE-associated eczema: A double-blind, randomized, placebo-controlled trial. J. Allergy Clin. Immunol. 2007, 119, 1174–1180. [Google Scholar] [CrossRef] [PubMed]
  34. Wickens, K.; Black, P.N.; Stanley, T.V.; Mitchell, E.; Fitzharris, P.; Tannock, G.W.; Purdie, G.; Crane, J. A differential effect of 2 probiotics in the prevention of eczema and atopy: A double-blind, randomized, placebo-controlled trial. J. Allergy Clin. Immunol. 2008, 122, 788–794. [Google Scholar] [CrossRef] [PubMed]
  35. Rosenfeldt, V.; Benfeldt, E.; Valerius, N.H.; Paerregaard, A.; Michaelsen, K.F. Effect of probiotics on gastrointestinal symptoms and small intestinal permeability in children with atopic dermatitis. J. Pediatr. 2004, 145, 612–616. [Google Scholar] [CrossRef] [PubMed]
  36. Rø, A.D.B.; Simpson, M.R.; Ro, T.B.; Storro, O.; Johnsen, R.; Videm, V.; Oien, T. Reduced Th22 cell proportion and prevention of atopic dermatitis in infants following maternal probiotic supplementation. Clin. Exp. Allergy 2017, 47, 1014–1021. [Google Scholar] [CrossRef] [PubMed]
  37. Ranadheera, C.S. Probiotic Application in the Development of Goat’s Milk Products with Special Reference to Propionibacterium Jensenii 702: Effects on Viability and Functionality. Ph.D. Thesis, The University of Newcastle Australia, Callaghan, Australia, 2012. [Google Scholar]
  38. Vasiljevic, T.; Shah, N.P. Probiotics—From Metchnikoff to bioactives. Int. Dairy J. 2008, 18, 714–728. [Google Scholar] [CrossRef]
  39. Ranadheera, C.S.; Evans, C.A.; Adams, M.C.; Baines, S.K. In vitro analysis of gastrointestinal tolerance and intestinal cell adhesion of probiotics in goat’s milk ice cream and yogurt. Food Res. Int. 2012, 49, 619–625. [Google Scholar] [CrossRef]
  40. McNaught, C.E.; MacFie, J. Probiotics in clinical practice: A critical review of the evidence. Nutr. Res. 2001, 21, 343–353. [Google Scholar] [CrossRef]
  41. Saarela, M.; Lahteenmaki, L.; Crittenden, R.; Salminen, S.; Mattila-Sandholm, T. Gut bacteria and health foods-the European perspective. Int. J. Food Microbiol. 2002, 78, 99–117. [Google Scholar] [CrossRef]
  42. Morelli, L. In vitro assessment of probiotic bacteria: From survival to functionality. Int. Dairy J. 2007, 17, 1278–1283. [Google Scholar] [CrossRef]
  43. Aureli, P.; Capurso, L.; Castellazzi, A.M.; Clerici, M.; Giovannini, M.; Morelli, L.; Poli, A.; Pregliasco, F.; Salvini, F.; Zuccotti, G.V. Probiotics and health: An evidence-based review. Pharmacol. Res. 2011, 63, 366–376. [Google Scholar] [CrossRef] [PubMed]
  44. LeBlanc, J.G.; Chain, F.; Martin, R.; Bermudez-Humaran, L.G.; Courau, S.; Langella, P. Beneficial effects on host energy metabolism of short-chain fatty acids and vitamins produced by commensal and probiotic bacteria. Microb. Cell Factories 2017, 16, 79. [Google Scholar] [CrossRef] [PubMed]
  45. Cholakov, R.; Tumbarski, Y.; Yanakieva, V.; Dobrev, I.; Salim, Y.; Denkova, Z. Antimicrobial activity of Leuconostoc lactis strain BT1&, isolated from a spontaneously fermented cereal beverage (Boza). J. Microbiol. Biotechnol. Food Sci. 2017, 7, 47–49. [Google Scholar]
  46. Vidanarachchi, J.K. Regulation of Intestinal Microflora and Productivity of Broiler Chickens by Prebiotic and Bioactive Plant Extracts. Ph.D. Thesis, The University of New England, Armidale, Australia, 2007. [Google Scholar]
  47. Prado, F.C.; Parada, J.L.; Pandey, A.; Soccol, C.R. Trends in non-dairy probiotic beverages. Food Res. Int. 2008, 41, 111–123. [Google Scholar] [CrossRef]
  48. Marco, M.L.; Tachon, S. Environmental factors influencing the efficacy of probiotic bacteria. Curr. Opin. Biotechnol. 2013, 24, 207–213. [Google Scholar] [CrossRef] [PubMed]
  49. Ranadheera, C.S.; Evans, C.A.; Adams, M.A.; Baines, S.K. Effect of dairy probiotic combinations on in vitro gastrointestinal tolerance, intestinal epithelial cell adhesion and cytokine secretion. J. Funct. Foods 2014, 8, 18–25. [Google Scholar] [CrossRef]
  50. Ranadheera, C.S.; Evans, C.A.; Adams, M.A.; Baines, S.K. Production of probiotic ice cream from goat’s milk and effect of packaging materials on product quality. Small Rumin. Res. 2013, 112, 174–180. [Google Scholar] [CrossRef]
  51. Lavermicocca, P. Highlights on new food research. Dig. Liver Dis. 2006, 38 (Suppl. S2), S295–S299. [Google Scholar] [CrossRef]
  52. Rouhi, M.; Sohrabvandi, S.; Mortazavian, A.M. Probiotic Fermented Sausage: Viability of Probiotic Microorganisms and Sensory Characteristics. Crit. Rev. Food Sci. Nutr. 2013, 53, 331–348. [Google Scholar] [CrossRef] [PubMed]
  53. Klingberg, T.D.; Budde, B.B. The survival and persistence in the human gastrointestinal tract of five potential probiotic lactobacilli consumed as freeze-dried cultures or as probiotic sausage. Int. J. Food Microbiol. 2006, 109, 157–159. [Google Scholar] [CrossRef] [PubMed]
  54. Oliveira, M.N.; Sodini, I.; Remeuf, F.; Corrieu, G. Effect of milk supplementation and culture composition on acidification, textural properties and microbiological stability of fermented milks containing probiotic bacteria. Int. Dairy J. 2001, 11, 935–942. [Google Scholar] [CrossRef]
  55. Martin-Diana, A.B.; Janer, C.; Pelaez, C.; Requena, T. Development of a fermented goat’s milk containing probiotic bacteria. Int. Dairy J. 2003, 13, 827–833. [Google Scholar] [CrossRef]
  56. 56. Ranadheera, C.S.; Evans, C.A.; Adams, M.A.; Baines, S.K. Co-culturing of probiotics influences the microbial and physico-chemical properties but not sensory quality of fermented dairy drink made from goats’ milk. Small Rumin. Res. 2016, 136, 104–108. [Google Scholar] [CrossRef]
  57. Maganha, L.C.; Rosim, R.E.; Corassin, C.H.; Cruz, A.G.; Faria, J.A.F. Viability of probiotic bacteria in fermented skim milk produced with different levels of milk powder and sugar. Int. J. Dairy Technol. 2014, 67, 89–94. [Google Scholar] [CrossRef]
  58. Shah, N.P.; Lankaputhra, W.E.V. Improving viability of Lactobacillus acidophilus and Bifidobacterium spp. in yogurt. Int. Dairy J. 1997, 7, 349–356. [Google Scholar] [CrossRef]
  59. Ekinci, F.Y.; Gurel, M. Effect of using propionic acid bacteria as an adjunct culture in yogurt production. J. Dairy Sci. 2008, 91, 892–899. [Google Scholar] [CrossRef] [PubMed]
  60. Kailasapathy, K.; Harmstorf, I.; Phillips, M. Survival of Lactobacillus acidophilus and Bifidobacterium animalis ssp. lactis in stirred fruit yogurts. LWT Food Sci. Technol. 2008, 41, 1317–1322. [Google Scholar] [CrossRef]
  61. Kumari, A.G.I.P.; Ranadheera, C.S.; Prasanna, P.H.P.; Senevirathne, N.D.; Vidanarachchi, J.K. Development of a rice incorporated synbiotic yogurt with low retrogradation properties. Int. Food Res. J. 2015, 22, 2032–2040. [Google Scholar]
  62. Prasanna, P.H.P.; Grandison, A.S.; Charalampopoulos, D. Microbiological, chemical and rheological properties of low fat set yoghurt produced with exopolysaccharide (EPS) producing Bifidobacterium strains. Food Res. Int. 2014, 51, 15–22. [Google Scholar] [CrossRef]
  63. Guler-Akin, M.B.; Akin, M.S. Effects of cysteine and different incubation temperatures on the microflora, chemical composition and sensory characteristics of bio-yogurt made from goat’s milk. Food Chem. 2007, 100, 788–793. [Google Scholar] [CrossRef]
  64. Ranadheera, C.S.; Evans, C.A.; Adams, M.A.; Baines, S.K. Probiotic viability and physico-chemical and sensory properties of plain and stirred fruit yogurts made from goat’s milk. Food Chem. 2012, 135, 1411–1418. [Google Scholar] [CrossRef] [PubMed]
  65. Guler-Akin, M.B. The effects of different incubation temperatures on the acetaldehyde content and viable bacteria counts of bio yogurt made from ewe’s milk. Int. J. Dairy Technol. 2005, 58, 174–179. [Google Scholar] [CrossRef]
  66. Alamprese, C.; Foschino, R.; Rossi, M.; Pompei, C.; Savani, L. Survival of Lactobacillus johnsonii La 1 and influence of its addition in retail-manufactured ice cream produced with different sugar and fat concentrations. Int. Dairy J. 2002, 12, 201–208. [Google Scholar] [CrossRef]
  67. Akin, M.B.; Akin, M.S.; Kirmaci, Z. Effects of inulin and sugar levels on the viability of yogurt and probiotic bacteria and the physical and sensory characteristics in probiotic ice cream. Food Chem. 2007, 104, 93–99. [Google Scholar] [CrossRef]
  68. Ferraz, J.L.; Cruz, A.G.; Cadena, R.S.; Freitas, M.Q.; Pinto, U.M.; Carvalho, C.C.; Faria, J.A.F.; Bolini, H.M.A. Sensory acceptance and survival of probiotic bacteria in ice cream produced with different overrun levels. J. Food Sci. 2012, 77, S24–S28. [Google Scholar] [CrossRef] [PubMed]
  69. Gardiner, G.E.; Bouchier, P.; O’Sullivan, E.; Kelly, J.; Collins, J.K.; Fitzgerald, G.; Ross, R.P.; Stanton, C. A spray-dried culture for probiotic Cheddar cheese manufacture. Int. Dairy J. 2002, 12, 749–756. [Google Scholar] [CrossRef]
  70. Ulpathakumbura, C.P.; Ranadheera, C.S.; Senavirathne, N.D.; Jayawardene, L.P.I.N.P.; Prasanna, P.H.P.; Vidanarachchi, J.K. Effect of biopreservatives on microbial, physico-chemical and sensory properties of Cheddar cheese. Food Biosci. 2016, 13, 21–25. [Google Scholar] [CrossRef]
  71. Buriti, F.C.A.; da Rocha, J.S.; Assis, E.G.; Saad, S.M.I. Probiotic potential of Minas fresh cheese prepared with the addition of Lactobacillus paracasei. LWT Food Sci. Technol. 2005, 38, 173–180. [Google Scholar] [CrossRef]
  72. Kasimoglu, A.; Goncuoglu, M.; Akgun, S. Probiotic white cheese with Lactobacillus acidophilus. Int. Dairy J. 2004, 14, 1067–1073. [Google Scholar] [CrossRef]
  73. Bergamini, C.V.; Hynes, E.R.; Quiberoni, A.; Suarez, V.B.; Zalazar, C.A. Probiotic bacteria as adjunct starters: Influence of the addition methodology on their survival in a semi-hard Argentinean cheese. Food Res. Int. 2005, 38, 597–604. [Google Scholar] [CrossRef]
  74. Vinderola, C.G.; Prosello, W.; Ghiberto, D.; Reinheimer, J.A. Viability of probiotic (Bifidobacterium, Lactobacillus acidophilus and Lactobacillus casei) and nonprobiotic microflora in Argentinian Fresco cheese. J. Dairy Sci. 2000, 83, 1905–1911. [Google Scholar] [CrossRef]
  75. Madureira, A.R.; Pereira, C.I.; Truszkowska, K.; Gomes, A.M.; Pintado, M.E.; Malcata, F.X. Survival of probiotic bacteria in a whey cheese vector submitted to environmental conditions prevailing in the gastrointestinal tract. Int. Dairy J. 2005, 15, 921–927. [Google Scholar] [CrossRef]
  76. Gomes, A.M.P.; Malcata, F.X. Development of probiotic cheese manufactured from goat milk: Response surface analysis via technological manipulation. J. Dairy Sci. 1998, 81, 1492–1507. [Google Scholar] [CrossRef]
  77. Gobbetti, M.; Crosetti, A.; Smacchi, E.; Zocchetti, A.; De Angelis, M. Production of crescenza cheese by incorporation of bifidobacteria. J. Dairy Sci. 1998, 81, 37–47. [Google Scholar] [CrossRef]
  78. Heenan, C.N.; Adams, M.C.; Hosken, R.W.; Fleet, G.H. Survival and sensory acceptability of probiotic microorganisms in a nonfermented frozen vegetarian dessert. LWT Food Sci. Technol. 2004, 37, 461–466. [Google Scholar] [CrossRef]
  79. Shimakava, Y.; Matsubara, S.; Yuki, N.; Ikeda, M.; Ishikawa, F. Evaluation of Bifidobacterium breve strain Yakult-fermented soymilk as a probiotic food. Int. J. Food Microbiol. 2003, 81, 131–136. [Google Scholar] [CrossRef]
  80. Ouwehand, A.C.; Kurvinen, T.; Rissanen, P. Use of a probiotic Bifidobacterium in a dry food matrix, an in vivo study. Int. J. Food Microbiol. 2004, 95, 103–106. [Google Scholar] [CrossRef] [PubMed]
  81. Helland, M.H.; Wicklund, T.; Narvhus, J.A. Growth and metabolism of selected strains of probiotic bacteria in milk- and water-based cereal puddings. Int. Dairy J. 2004, 14, 957–965. [Google Scholar] [CrossRef]
  82. Molin, G. Probiotics in foods not containing milk or milk constituents, with special reference to Lactobacillus plantarum 299v. Am. J. Clin. Nutr. 2001, 73, 380S–385S. [Google Scholar] [PubMed]
  83. Luckow, T.; Delahunty, C. Which juice is healthier? A consumer study of probiotic non-dairy juice drinks. Food Qual. Preference 2004, 15, 751–759. [Google Scholar] [CrossRef]
  84. Betoret, N.; Puente, L.; Diaz, M.J.; Pagan, M.J.; Garcia, M.J.; Gras, M.L.; Martinez-Monzo, J.; Fito, P. Development of probiotic-enriched dried fruits by vacuum impregnation. J. Food Eng. 2003, 56, 273–277. [Google Scholar] [CrossRef]
  85. Ding, W.K.; Shah, N.P. Survival of free and microencapsulated probiotic bacteria in orange and apple juices. Int. Food Res. J. 2008, 15, 219–232. [Google Scholar]
  86. Costa, M.G.M.; Fonteles, T.V.; de Jesus, A.L.T.; Rodrigues, S. Sonicated pineapple juice as substrate for L. casei cultivation for probiotic beverage development: Process optimisation and product stability. Food Chem. 2013, 139, 261–266. [Google Scholar] [CrossRef] [PubMed]
  87. AdebayoTayo, B.; Akpeji, S. Probiotic viability, physicochemical and sensory properties of probiotic pineapple juice. Fermentation 2016, 2, 20. [Google Scholar] [CrossRef]
  88. Pereira, A.L.F. Spray-Drying of Probiotic Cashew Apple Juice. Food Bioprocess Technol. 2014, 7, 2492–2499. [Google Scholar] [CrossRef]
  89. Alves, N.N. Spouted bed as an efficient processing for probiotic orange juice drying. Food Res. Int. 2017, 101, 54–60. [Google Scholar] [CrossRef] [PubMed]
  90. Barbosa, J.; Borges, S.; Amorim, M.; Pereira, M.J.; Oliveira, A.; Pintado, M.E.; Teixeira, P. Comparison of spray drying, freeze drying and convective hot air drying for the production of a probiotic orange powder. J. Funct. Foods 2015, 17, 340–351. [Google Scholar] [CrossRef]
  91. Luckow, T.; Delahunty, C. Consumer acceptance of orange juice containing functional ingredients. Food Res. Int. 2004, 37, 805–814. [Google Scholar] [CrossRef]
  92. Lavermicocca, P.; Valerio, F.; Lonigro, S.L.; Angelis, M.D.; Morelli, L.; Callegari, M.L.; Rizzello, C.G.; Visconti, A. Study of adhesion and survival of lactobacilli and bifidobacteria on table olives with the aim of formulating a new probiotic food. Appl. Environ. Microbiol. 2005, 71, 4233–4240. [Google Scholar] [CrossRef] [PubMed]
  93. Valero-Cases, E.; Frutos, M.J. Effect of Inulin on the Viability of L. plantarum during Storage and In Vitro Digestion and on Composition Parameters of Vegetable Fermented Juices. Plant Foods Hum. Nutr. 2017, 72, 161–167. [Google Scholar] [CrossRef] [PubMed]
  94. Yoon, K.Y.; Woodams, E.E.; Hang, Y.D. Fermentation of beet juice by beneficial lactic acid bacteria. LWT Food Sci. Technol. 2005, 38, 73–75. [Google Scholar] [CrossRef]
  95. Yoon, K.Y.; Woodams, E.E.; Hang, Y.D. Production of probiotic cabbage juice by lactic acid bacteria. Bioresour. Technol. 2006, 97, 1427–1430. [Google Scholar] [CrossRef] [PubMed]
  96. Valerio, F.; De Bellis, P.; Lonigro, S.L.; Morelli, L.; Visconti, A.; Lavermicocca, P. In vitro and in vivo survival and transit tolerence of potentially probiotic strains carried by Artichokes in the gastrointestinal tract. Appl. Environ. Microbiol. 2006, 72, 3042–3045. [Google Scholar] [CrossRef] [PubMed]
  97. Kim, S.Y. Production of Fermented Kale Juices with Lactobacillus Strains and Nutritional Composition. Prev. Nutr. Food Sci. 2017, 22, 231–236. [Google Scholar] [PubMed]
  98. Battistini, C.; Gullon, B.; Ichimura, E.S.; Gomes, A.M.P.; Ribeiro, E.P.; Kunigk, L.; Moreira, J.U.V.; Jurkiewicz, C.J. Development and characterization of an innovative synbiotic fermented beverage based on vegetable soybean. Braz. J. Microbiol. 2017, in press. [Google Scholar] [CrossRef] [PubMed]
  99. Fan, S.; Breidt, F.; Price, R.; Perez-Diaz, I. Survival and Growth of Probiotic Lactic Acid Bacteria in Refrigerated Pickle Products. J. Food Sci. 2017, 82, 167–173. [Google Scholar] [CrossRef] [PubMed]
  100. Erkkilä, S.; Suihko, M.L.; Eerola, S.; Petaja, E.; Mattila-Sandholm, T. Dry sausage fermented by Lactobacillus rhamnosus strains. Int. J. Food Microbiol. 2001, 64, 205–210. [Google Scholar]
  101. Speranza, B.; Racioppo, A.; Beneduce, L.; Bevilacqua, A.; Sinigaglia, M.; Corbo, M.R. Autochthonous lactic acid bacteria with probiotic aptitudes as starter cultures for fish-based products. Food Microbiol. 2017, 65, 244–253. [Google Scholar] [CrossRef] [PubMed]
  102. Ranadheera, C.S.; Evans, C.A.; Adams, M.A.; Baines, S.K. Microencapsulation of Lactobacillus acidophilus LA-5, Bifidobacterium animalis subsp. lactis BB-12 and Propionibacterium jensenii 702 by spray drying in goat’s milk. Small Rumin. Res. 2015, 123, 155–159. [Google Scholar] [CrossRef]
  103. Silva, M.P.; Tulini, F.L.; Marinho, J.F.U.; Mazzocato, M.C.; De Martinis, E.C.P.; Luccas, V.; Favaro-Trindade, C.S. Semisweet chocolate as a vehicle for the probiotics Lactobacillus acidophilus LA3 and Bifidobacterium animalis subsp. lactis BLC1: Evaluation of chocolate stability and probiotic survival under in vitro simulated gastrointestinal conditions. LWT Food Sci. Technol. 2017, 75, 640–647. [Google Scholar] [CrossRef]
  104. McMaster, L.D.; Kokott, S.A.; Reid, S.J.; Abratt, V.R. Use of traditional African fermented beverages as delivery vehicles for Bifidobacterium lactis DSM 10140. Int. J. Food Microbiol. 2005, 102, 231–237. [Google Scholar] [CrossRef] [PubMed]
  105. Mudannayake, D.C.; Silva, K.F.S.T.; Wimalasiri, K.M.S.; Ajlouni, S. In Vitro Prebiotic Properties of Partially Purified Asparagus Falcatus and Taraxacum Javanicum Inulins. J. Food Nutr. Disord. 2016, 5, 1–10. [Google Scholar] [CrossRef]
  106. Grizard, D.; Barthomeuf, C. Non-digestible oligosaccharides used as prebiotic agents: mode of production and beneficial effects on animal and human health. Reprod. Nutr. Dev. 1999, 39, 588. [Google Scholar] [CrossRef]
  107. Anal, A.K.; Singh, H. Recent advances in microencapsulation of probiotics for industrial applications and targeted delivery. Trends Food Sci. Technol. 2007, 18, 240–251. [Google Scholar] [CrossRef]
  108. Abesinghe, N.; Vidanarachchi, J.K.; Silva, S. The effect of Arrowroot (Maranta arundinacea) extract on the survival of probiotic bacteria in set yoghurt. Int. J. Sci. Res. Publ. 2012, 2, 1–4. [Google Scholar]
  109. Huang, Y.; Adams, M.C. In vitro assessment of the upper gastrointestinal tolerance of potential probiotic dairy propionibacteria. Int. J. Food Microbiol. 2004, 91, 253–260. [Google Scholar] [CrossRef] [PubMed]
  110. Cruz, A.G.; Antunes, A.E.C.; Sousa, A.L.O.P.; Faria, J.A.F.; Saad, S.M.I. Ice-cream as a probiotic food carrier. Food Res. Int. 2009, 42, 1233–1239. [Google Scholar] [CrossRef]
  111. Vinderola, G.; Binetti, A.; Burns, P.; Reinheimer, J. Cell Viability and Functionality of Prebiotic Bacteria in Dairy Products. Front. Microbiol. 2011, 2, 70. [Google Scholar] [CrossRef] [PubMed]
  112. Gawkowski, D.; Chikindas, M.L. Non-dairy probiotic beverages: The next step into human health. Benef. Microbes 2013, 4, 127–142. [Google Scholar] [CrossRef] [PubMed]
  113. Salminen, S.; Kenifel, W.; Ouwehand, A.C. Bacteria, Beneficial-Probiotics, Applications in Dairy Products. In Encyclopedia of Dairy Sciences, 2nd ed.; John, W.F., Ed.; Academic Press: San Diego, CA, USA, 2011; pp. 412–419. [Google Scholar]
  114. Patel, A.R. Probiotic fruit and vegetable juices-recent advances and future perspectives. Int. Food Res. J. 2017, 24, 1850–1857. [Google Scholar]
  115. Jayawardana, N.W.I.A.; Prasanna, P.H.P.; Ranadheera, C.S.; de Zoysa, H.K.S.; Vidanarachchi, J.K. Probiotics in functional foods. In Fermented Foods: Sources, Consumption and Health Benefits; Nova Science Publishers: New York, NY, USA, 2015; pp. 103–126. [Google Scholar]
  116. Guarner, F.; Perdigon, G.; Corthier, G.; Salminen, S. Should yoghurt cultures be considered probiotic? Br. J. Nutr. 2005, 93, 783–786. [Google Scholar] [CrossRef] [PubMed]
  117. Meydani, S.N.; Ha, W.K. Immunologic effects of yogurt. Am. J. Clin. Nutr. 2000, 71, 861–872. [Google Scholar] [PubMed]
  118. Dave, R.I.; Shah, N.P. Viability of yoghurt and probiotic bacteria in yoghurts made from commercial starter cultures. Int. Dairy J. 1997, 7, 31–41. [Google Scholar] [CrossRef]
  119. Batista, A.L.D.; Silva, R.; Cappato, L.P.; Almada, C.N.; Garcia, R.K.A.; Silva, M.C.; Raices, R.S.L.; Arellano, D.B.; Sant’Ana, A.S.; Junior, C.A.C.; et al. Quality parameters of probiotic yogurt added to glucose oxidase compared to commercial products through microbiological, physical–chemical and metabolic activity analyses. Food Res. Int. 2015, 77, 627–635. [Google Scholar] [CrossRef]
  120. Kumar, B.V.; Vijayendra, S.V.N.; Reddy, O.V.S. Trends in dairy and non-dairy probiotic products—A review. J. Food Sci. Technol. 2015, 52, 6112–6124. [Google Scholar] [CrossRef] [PubMed]
  121. Meybodi, N.M.; Mortazavian, A.M.; da Cruz, A.G.; Mohammadi, R. Probiotic Supplements and Food Products: Comparison for Different Targets. Appl. Food Biotechnol. 2017, 4, 123–132. [Google Scholar]
  122. Annunziata, A.; Vecchio, R. Consumer perception of functional foods: A conjoint analysis with probiotics. Food Qual. Preference 2013, 28, 348–355. [Google Scholar] [CrossRef]
  123. Robinson, R.K.; Tamime, A.Y. Types of fermented milks. In Fermented Milks; Blackwell Science Ltd.: Oxford, UK, 2007; pp. 1–10. [Google Scholar]
Figure 1. Classification and types of probiotic beverages. Adapted and modified from [115,120].
Figure 1. Classification and types of probiotic beverages. Adapted and modified from [115,120].
Fermentation 03 00067 g001
Figure 2. Flow chart for the production of fermented and non-fermented fruit/vegetable juices enriched with probiotics.
Figure 2. Flow chart for the production of fermented and non-fermented fruit/vegetable juices enriched with probiotics.
Fermentation 03 00067 g002
Table 1. Microorganisms used as probiotic cultures.
Table 1. Microorganisms used as probiotic cultures.
Lactobacillus spp.Bifidobacterium spp.Other spp.
L. acidophilus
L. casei
L. crispatus
L. delbrueckii subsp. bulgaricus a
L. fermentum
L. gasseri
L. johnsonii
L. paracasei
L. plantarum
L. reuteri
L. rhamnosus
L. helveticus
L. lactis
L. sporogenes
B. bifidum
B. breve
B. infantis
B. longum
B. lactis
B. animalis
B. adolescentis
B. essensis
B. laterosporus
Escherichia coli Nissle
Saccharomyces boulardii
Saccharomyces cerevisiae
Kluyveromyces lactis
Streptococcus thermophilus a
S. cremoris
S. diacetylactis
S. intermedius
S. salivarius
Enterococcus francium b
Propionibacterium freudenreichii
P. freudenreichii subsp. shermanii
P. jensenii
Pediococcus
Leuconostoc lactis subsp. cremoris
L. lactis subsp. lactis Bacillus cereus
Clostridium butyricum
a There is still debate about the probiotic activity due to poor survival during gastrointestinal transit; b Safety concerns remain because of potential pathogenicity and vancomycin resistance. Adopted and modified from [6,7,8,9,10].
Table 2. Examples of beneficial effects of therapeutic probiotic application in humans.
Table 2. Examples of beneficial effects of therapeutic probiotic application in humans.
DisorderProbiotic StrainMode of DeliveryReferences
Antibiotic-associated diarrhoea in adultsMixture of L. caseiDrinking yogurt[14]
L. bulgaricus
S. thermophilus
Antibiotic-associated diarrhoea in childrenLactobacillus reuteriDrops[22]
Traveler’s diarrhoeaSingle strain of Lactobacillus GGPowdered form dissolved in cold water[23]
Irritable bowel syndrome symptomsMixture of B. longum, B. infantis, B. breve, L. acidophilus, L. casei, L. delbrueckii, L. plantarum, S. salivariusLyophilized powdered form[13]
Mixture of L. rhamnosus, B. breve & P. freudenreichii subsp. shermaniiCapsules[24]
Mixture of B. animalis, L. bulgaricus & S. thermophilusFermented milk[25]
Mixture of B. longum, B. infantis, B. breve, L. acidophilus, L. casei, L. delbrueckii, L. plantarum, S. salivariusLyophilized powdered form[26]
Single strain of L. plantarumRose-hip drink with oat flour[27]
Single strain of B. animalisFermented semi skimmed-milk[28]
Crohn’s disease
Ulcerative colitis
Pouchitis
Mixture of B. longum, B. infantis, B. breve, L. acidophilus, L. casei, L. bulgaricus, L. plantarum, S. thermophilusLyophilized form[29]
Single strain of E. coli NissleCapsules[30]
Mixture of L. acidophilus La-5,Fermented milk[31]
Bifidobacterium Bb 12
Bacterial vaginosisMixture of L. rhamnosus & L. reuteriGelatin capsules[32]
IgE associated eczema
Atopic dermatitis
Atopic dermatitis in infants
Single strain of L. reuteriFreeze dried form in coconut or peanut oil droplets[33]
Single strain of L. rhamnosusSkim milk based freeze-dried form[34]
Mixture of L. rhamnosus & L. reuteriLyophilized powdered form[35]
Mixture of L. rhamnosus, B. animalis subsp. lactis Bb-12 (Bb-12) & L. acidophilus La-5Milk (maternal supplementation)[36]
Adapted and modified from [37].
Table 3. Key and desirable criteria for the selection of probiotics in food and nutraceutical applications.
Table 3. Key and desirable criteria for the selection of probiotics in food and nutraceutical applications.
CriteriaProperty/CharacteristicTarget and Methods to Be Assessed
SafetyOrigin
Pathogenicity and infectivity
Virulence factors-toxicity, metabolic activity and intrinsic properties, i.e., antibiotic resistance
Source or origin should be assessed: be isolated from the same species as its intended host is desirable due to higher efficacy in the same species. Probiotics of human origin may be desirable if they are intended for human use.
Pre-market clearance and post-market surveillance
Technological acceptabilityHigh viability retention during manufacturing and storage of carrier foods
Acceptable organoleptic characteristics
Ability to produce at large-scale
Phage resistance
In vitro studies and food product development
Sensory testing of model and final products and consumer studies on product formulations
FunctionalityTolerance to gastric acid and juices including acidic conditions and enzymes
Bile tolerance
Adhesion to mucosal surface and colonization
Validated and documented health effects
Model systems for gastric and bile effects (e.g., in vitro, animal and human studies)
In vitro adhesion models (e.g., intestinal segments, mucus, cell culture), animal and human studies
Health effects confirmed by clinical studies
Desirable physiological criteriaImmunomodulation
Antagonistic activity towards gastrointestinal pathogens
Antimutagenic and anticarcinogenic properties
In vitro/In vivo animal and human studies.
Adhesion and competitive exclusion of pathogens in in vitro and in vivo model systems
Adapted and modified from [37,38,40,41,42,43].
Table 4. The diversity of probiotic food products and the viability of each probiotic in different products at the end of appropriate storage conditions (either freeze or cold storage).
Table 4. The diversity of probiotic food products and the viability of each probiotic in different products at the end of appropriate storage conditions (either freeze or cold storage).
Product TypeProductProbiotic StrainViability at the End of StorageTotal Storage TimeReferences
Dairy basedFermented cow’s milksL. acidophilus
L. rhamnosus
107 cfu/g7 days[54]
Fermented goat’s milkL. acidophilus
Bifidobacterium BB-12
<106 cfu/g
106–107 cfu/g
21 days[55]
Fermented dairy drink from goat’s milkL. acidophilus
B. animalsi ssp. lactis
107 cfu/mL21 days[56]
Fermented skim milk (cow’s milk)L. acidophilus
B. animalsi ssp. lactis
106 cfu/mL21 days[57]
Cow’s milk yogurtL. acidophilus>106 cfu/g42 days[58]
B. longum
B. psedolongum
B. infantis
B. bifidum
P. jensenii105 cfu/g15 days[59]
Cow’s milk fruit yogurtL. acidophilus
B. animalsi ssp. lactis
106–107 cfu/g35 days[60]
Rice incorporated cow’s milk yogurtB. animalis subsp. lactis BB-12108 cfu/g21 days[61]
Low fat set yogurts (cow’s milk)B. infantis, B. longum subsp. infantis107 cfu/g28 days[62]
Goat’s milk yogurtL. acidophilus
B. bifidum
L. paracasei subsp. casei
107 cfu/g14 days[63,64]
Ewe’s milk yogurtL. acidophilus
B. bifidum
L. casei
107 cfu/g14 days[65]
Ice creamL. johnsonii
L. acidophilus
B. lactis
107 cfu/g
105–106 cfu/g
8 months
90 days
[66]
[67]
Ice cream (vanilla flavoured)L. acidophilus106 cfu/mL60 days[68]
Goat’s milk ice cream (chocolate flavoured)L. acidophilus, B. animalis subsp. Lactis, Propionibacterium jensenii107–108 cfu/g52 weeks[50]
Cheddar cheeseL. paracasei107 cfu/g90 days[69]
Lactococcus lactis subsp. cremoris,
L. lactis subsp. lactis, Lactobacillus helvetics, S. thermophiles,
Lactobacillus rhamnosus
108 cfu/g4 weeks[70]
Fresh Minas cheeseL. paracasei108 cfu/g21 days[71]
White Turkish cheeseL. acidophilus107 cfu/g90 days[72]
Semi hard Argentinian cheeseL. paracasei
L. acidophilus
108 cfu/g60 days[73]
Argentinian Fresco cheeseB. bifidum
B. longum
L. acidophilus
L. casei
106 cfu/g60 days[74]
Requeijao-cheese (Portuguese-whey cheese)L. animalis
L. acidophilus
L. paracasei
L. brevis
107 cfu/g28 days[75]
Semi hard goat’s cheeseL. acidophilus
B. lactis
106 cfu/g70 days[76]
Crescenza cheese (soft Italian cheese)B. bifidum
B. infantis
B. longum
105 cfu/g14 days[77]
Soya basedSoya frozen dessertL. acidophilus107 cfu/g28 weeks[78]
L. paracasei
B. lactis
L. rhamnosus
S. boulardii~105 cfu/g
Soy milkB. breve109 cfu/mL20 days[79]
Cereal basedOat barsB. lactis109 cfu/25 g bar7–14 days[80]
Milk based maize/rice puddingB. animalis
L. acidophilus
L. rhamnosus
108–109 cfu/g21 days[81]
Oat meal gruel mixed with fruit drinks (i.e., rose hip, strawberry)L. plantarum1010 cfu/mL30 days[82]
Fruit and fruit juiceBlackcurrantL. plantarumNot reported[83]
Dried apple fruitsL. casei106 cfu/g[84]
Apple juiceL. acidophilus
L. rhamnosus
L. salivarius
L. plantarum
L. paracasei
B. longum
B. lactis type Bi-04
B. lactis type Bi-07
106 cfu/mL6 weeks[85]
Pineapple juiceL. casei106 cfu/mL42 days[86]
Pediococcus pentosaceus, Lactobacillus rhamnosus, Pediococcus pentosaceus109 cfu/mL4 weeks[87]
Cashew apple juice powder (spray dried)L. casei NRRL B-442~106 cfu/g35 days[88]
Orange juice after Spouted bed dryingL. casei~106 cfu/g5 weeks[89]
Orange juice powder (spray and freeze dried)L. plantarum 299v106 cfu/g180 days[90]
OrangeLactobacillus GGNot reported[91]
Vegetable basedTable olivesL. rhamnosus
L. paracasei
B. bifidum
B. longum
106–108 cfu/g90 days[92]
Carrot blended with orange juiceL. plantarum CECT 220108–109 cfu/mL30 days[93]
Tomato juiceL. plantarum
L. acidophilus
L. casei
L. delbrueckii
104–108 cfu/g30 days[51]
Beet juiceL. plantarum
L. acidophilus
L. casei
L. delbrueckii
106–108 cfu/mL4 weeks[94]
Cabbage juiceL. plantarum
L. delbrueckii
107 cfu/mL
105 cfu/mL
4 weeks[95]
ArtichokesL. plantarum
L. paracasei
107–108 cfu/g90 days[96]
Fermented Kale juice (Brassica oleraceae)L. plantarum
L. casei
L. acidophilus
L. brevis
108 cfu/mL4 weeks[97]
Fermented vegetable soybean beverageL. acidophilus La-5,
B. animalis Bb-12
~106 cfu/mL
108 cfu/mL
28 days[98]
Vegetable pickle productsL. casei LA284104–108cfu/g70 days[99]
Meat and fish productsFermented sausage L. plantarum--[53]
Dry sausages-beef + porkL. rhamnosus108 cfu/g28 days[100]
Fermented fish sausageLactobacillus spp.satisfactory7 days[101]
MiscellaneousEncapsulated and spray dried milk powderL. acidophilus, B. animalis subsp. Lactis, Propionibacterium jensenii105–107 cfu/g24 weeks[102]
ChocolateL. acidophilus LA3, B. animalis subsp. lactis BLC1107–108 cfu/g120 days[103]
African beverages made from maize and milkB. lactis107 cfu/mL21 days[104]
Adapted and modified from [37,51].
Table 5. Commercially available non-dairy probiotic beverages.
Table 5. Commercially available non-dairy probiotic beverages.
Product LabelManufacturerMajor CharacteristicsProbiotic Strain/s
Golden Circle Healthy Life Probiotic JuiceGolden Circle, AustraliaMixture of apple juice and mango puree or orange, apple, pineapple, passionfruit with banana pureeL. paracasei 8700:2 and L. plantarum HEAL 9
PERKii Probiotic WaterPERKii, AustraliaFruit juice mixtures such as raspberry and pomegranate, lime and coconut, mango and passionfruit and strawberry and watermelonLactobacillus casai Lc431
Bravo FriscusProbi AB, SwdeenOrange apple and tropical fruit juicesL. plantarum HEAL9 and L. paracasei 8700:2
ProVivaEMEA Probi AB, SwedenFruit juice (orange, strawberry or blackcurrant)L. plantarum 299v
Bio-Live Gold & DarkBio-Live/Microbz Ltd., UKMixtures of fruit juices such as acai berry, cherry, goji, noni, pomegranate, lemon, and various herbsMixture of 13 strains including L. acidophilus, L bulgaricus, L. casei, L. fermentum, L. plantarum, Lactococcus lactis, Bacillus subtilis, B. bifidum, B. Infantis B longum; Streptococcus thermophilus, Comobcillus and Saccharomyces cerevisiae
BiolaTINE, NorwayMixture of apple, grapes and passion fruit or orange and tangerineL. rhamnosus GG
Malee ProbioticsMalee Enterprise Compny Ltd., ThailandFruit juices such as prune, grape and orangeL. paracasei
GoodBelly® Carrot Ginger Flavor Probiotics Juice Drink.Goodbelly, USACarrot Juice, ginger extract and cane sugar contains 2% or less of gluten-free oat flour,L. Plantarum 299v
KEVITAKEVITA, USAVarious fruit based mixtures such as strawberry and coconut, lime, mint and coconut, mango and coconut, pineapple and coconutBacillus coagulans GBI-30 6086, L. rhamnosus, L. plantarum, L. paracasei
Tropicana probioticsTropicana, USAFruit juice mixtures such as strawberry and banana, pineapple and mango and peach passion fruitB. lactis
Probiotic Naked JuiceNaked® Juice, USAMixture of apple, orange, pineapple juices and mango and banana puree with fructooligosacccharidesBifidobacterium
Adapted and modified from [114].
Table 6. Types of fermentation used to produce different fermented dairy beverages.
Table 6. Types of fermentation used to produce different fermented dairy beverages.
Fermentation TypeType of Microorganism InvolvedBeverage Products
Lactic FermentationMesophilic typeCultured buttermilk
Thermophilic typeBulgarian buttermilk, Drinking yogurt
TherapeuticAcidophilus milk, Yakult
Yeast-lactic FermentationYeast and lactic acid bacteriaKefir, Acidophilus yeast milk
Mould-lactic FermentationMould and lactic acid bacteriaVilli
Adapted and modified from [115,123].

Share and Cite

MDPI and ACS Style

Ranadheera, C.S.; Vidanarachchi, J.K.; Rocha, R.S.; Cruz, A.G.; Ajlouni, S. Probiotic Delivery through Fermentation: Dairy vs. Non-Dairy Beverages. Fermentation 2017, 3, 67. https://doi.org/10.3390/fermentation3040067

AMA Style

Ranadheera CS, Vidanarachchi JK, Rocha RS, Cruz AG, Ajlouni S. Probiotic Delivery through Fermentation: Dairy vs. Non-Dairy Beverages. Fermentation. 2017; 3(4):67. https://doi.org/10.3390/fermentation3040067

Chicago/Turabian Style

Ranadheera, Chaminda Senaka, Janak K. Vidanarachchi, Ramon Silva Rocha, Adriano G. Cruz, and Said Ajlouni. 2017. "Probiotic Delivery through Fermentation: Dairy vs. Non-Dairy Beverages" Fermentation 3, no. 4: 67. https://doi.org/10.3390/fermentation3040067

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop