Next Article in Journal
Daily Activity Patterns and Co-Occurrence of Duikers Revealed by an Intensive Camera Trap Survey across Central African Rainforests
Previous Article in Journal
Butyrate Permeation across the Isolated Ovine Reticulum Epithelium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

State-of-the-Art of the Nutritional Alternatives to the Use of Antibiotics in Humans and Monogastric Animals

1
Department of Veterinary Sciences, School of Agriculture and Veterinary Medicine, University of Turin, Grugliasco, Largo Braccini 2, 10095 Torino, Italy
2
Department of Agricultural, Forestry and Food Sciences, School of Agriculture and Veterinary Medicine, University of Turin, Grugliasco, Largo Braccini 2, 10095 Torino, Italy
3
Department of Surgical Sciences, Medical School, University of Turin, Corso Dogliotti 14, 10126 Torino, Italy
*
Author to whom correspondence should be addressed.
Animals 2020, 10(12), 2199; https://doi.org/10.3390/ani10122199
Submission received: 28 October 2020 / Revised: 19 November 2020 / Accepted: 20 November 2020 / Published: 24 November 2020
(This article belongs to the Section Animal Nutrition)

Abstract

:

Simple Summary

Antibiotic resistance represents a worldwide recognized issue affecting both human and veterinary medicine, with a particular focus being directed towards monogastric animals destined for human consumption. This scenario is the result of frequent utilization of the antibiotics either for therapeutic purposes (humans and animals) or as growth promoters (farmed animals). Therefore, the search for nutritional alternatives has progressively been the object of significant efforts by the scientific community. So far, probiotics, prebiotics and postbiotics are considered the most promising products, as they are capable of preventing or treating gastrointestinal diseases as well as restoring a eubiosis condition after antibiotic-induced dysbiosis development. This review provides an updated state-of-the-art of these nutritional alternatives in both humans and monogastric animals.

Abstract

In recent years, the indiscriminate use of antibiotics has been perpetrated across human medicine, animals destined for zootechnical productions and companion animals. Apart from increasing the resistance rate of numerous microorganisms and generating multi-drug resistance (MDR), the nonrational administration of antibiotics causes sudden changes in the structure of the intestinal microbiota such as dysbiotic phenomena that can have a great clinical significance for both humans and animals. The aim of this review is to describe the state-of-the-art of alternative therapies to the use of antibiotics and their effectiveness in humans and monogastric animals (poultry, pigs, fish, rabbits, dogs and cats). In particular, those molecules (probiotics, prebiotics and postbiotics) which have a direct function on the gastrointestinal health are herein critically analysed in the prevention or treatment of gastrointestinal diseases or dysbiosis induced by the consumption of antibiotics.

1. Introduction

The second half of the 20th century discovered the novel use of antibiotics as growth promoters for food in the human diet, while the 21st century saw onset and rapid increase in advanced microbial antibiotics [1]. Over the past half century, antimicrobial resistance has become a growing medical concern due to persistent, selective pressure from the widespread use of antimicrobials in humans, animals and agriculture [2,3]. In 2001, the European Union health minister recommended a more rational administration of antimicrobial agents in human medicine, with a series of specific measures aimed at curbing antimicrobial resistance spread. Furthermore, alerting this crisis, the World Health Organization (WHO) in May 2015 adopted a global action to deal with antimicrobial resistance. This plan underlined the need for an effective one health approach that requires the coordination of several sectors including human and veterinary medicine, agriculture, finance, environment and consumers [4]. Despite this, the resistance rate to different bacterial species increases each year (especially against Staphylococcus spp., Enterococcus spp., Enterobacteria spp., Pseudomonas spp. and Acinetobacter spp.) [4]. The multi-drug resistance (MDR) definitions from the European Centre for Disease Prevention and Control (ECDC), published in 2011, were mainly chosen to harmonize epidemiological surveillance data in all health contexts and countries [5]. The ECDC criteria define MDR as acquired nonsusceptibility to at least one agent in three or more antimicrobial categories [6], thus underlying a representative limitation in the use of antibiotics, especially in severe clinical settings [7]. Therefore, in recent years, the European community has invested hundreds of millions of euros in funding research in antimicrobial resistance study, thus confirming its medical and social relevance. Based on these considerations, the search for new antimicrobial remedies is of vital importance. The aim of this work is to describe the state-of-the-art regarding possible alternative therapeutic strategies to the use of antibiotics in humans and monogastric animals and their impact on intestinal microbial ecology.

2. Intestinal Microbiota: Definition, Characteristics and Factors Affecting Its Balance

2.1. Humans

The human microbiota is a diverse microbial ecosystem now identified as an integral part of the gastrointestinal tract (GI) [8,9,10]. Indeed, in the last 30 years, numerous studies have been focused on its qualitative and quantitative characterization [11,12,13,14]. The first investigation by Suau et al. and Wilson et al. showed high diversity of faecal microbiota in healthy patients, as in-depth phylogenetic analyses revealed that the vast majority of the observed rDNA diversity was attributable to hitherto unknown dominant microorganisms within the human gut [15,16]. Within the last few years, metagenomics has also offered new insights into microbial diversity of the dominant microorganisms [17,18], thus providing a clear definition of the human microbiome. In particular, focusing on quantification, human intestinal microbiota is ten times higher than the prokaryotic cells of the organism, and their genetic heritage is clearly superior to human genome [19].
Tap et al. first demonstrated the existence of a bacterial “phylogenetic core” in the faecal microbiota of healthy adults [20]. The 16S rRNA gene sequences herein obtained confirmed that the predominant human faecal microbiota belonged to five phyla: Firmicutes (79.4%), Bacteroidetes (16.9%), Actinobacteria (2.5%), Proteobacteria (1%) and Verrucomicrobia (0.1%). Nevertheless, the composition of the human microbiota can be modified by some individual factors such as age, sex, nutrition status, lifestyle and circadian rhythms. With the aim of investigating all these variables in the human microbiota, the NIH Common Fund Human Microbiome Project (HMP) was established in 2008. So far, the HMP has developed metagenomic protocols for creating, processing and interpreting distinct types of high-throughput metagenomic data available to the scientific community [19]. Thanks to next-generation sequencing (NGS) of the small subunit ribosomal RNA (16s rRNA), it has been confirmed that Bacteroidetes and Firmicutes are the most abundant taxa in the intestinal microbiota of healthy adults [21,22], with approximately 500–1000 bacterial species belonging to Bacteroides, Bifidobacterium, Eubacterium, Clostridium, Peptococcus, Peptostreptococcus, Lactobacillus and Ruminococcus [23]. Specific changes can, however, be identified depending on the different stages of life. In particular, the way a baby is delivered influences postnatal microbial exposure. Indeed, babies born by vaginal delivery are exposed to vaginal microbes (Lactobacillus and Prevotella spp.) while the microbiome of babies born via C-section is dominated by Staphylococcus, Corynebacterium and Propionibacterium spp., which resemble microbial communities of skin [24,25]. During the first years of life, due to the introduction of solid foods, the composition of the intestinal microbiome becomes more complex, with a reduction of Bifidobacteria [26], when in puberty, major changes are driven by hormones, which determine the expression of genes related to development and growth [27,28]. In adulthood, the composition of the gut microbiome reaches a certain stability by increasing microbial count and complexity, with the increase in microbes belonging to the Lachnospiraceae and Ruminococcaceae families [29]. Ageing is accompanied by significant changes in lifestyle, such as decreased locomotion, nutritional changes and chronic consumption of medication, thus being accompanied by a general expansion of Bacteroidetes and reduction of Firmicutes phyla. Microbiome-associated metabolites (vitamins B7 and B12, and creatine) and their biosynthetic pathways are also reduced in ageing, thus contributing to muscle atrophy and frailty [22,30]. Noteworthily, gut microbiota in semi-supercentenarians (105–109 years old) is characterized by increased abundance of health-associated taxa like Bifidobacterium, Christensenellaceae and Akkermansia [31].

2.2. Monogastrics Animals

The definition of “intestinal microbiota” can equally be applied to both humans and the animals, since it is defined as “the usually complex mixture of bacterial populations that colonize a given area of the gastrointestinal tract in individual human or animal hosts that have not been affected by medical or experimental intervention or disease” [32]. However, considering the profound impact of the microbiome on human health, the majority of the available studies regarding the gastrointestinal microbiota and/or microbiome functionality has mainly been focused on humans [33]. Nevertheless, characterization of the intestinal microbiota has also been a topic of great research interest in monogastric animals since the 1970s. In particular, previous studies reported that the gastrointestinal microbiota of monogastric animals is composed primarily of bacteria, especially anaerobic, Gram-positive bacteria. The main bacterial species have been estimated to range from 400 to 500, with the bacterial densities generally increasing from the proximal to the distal gut until the identification of 1010–1012 bacterial cells per each gram of colon content or faeces [34]. The recent adoption of NGS technologies instead of classical cultivation-based methods has further improved scientific knowledge about the bacterial diversity of the gastrointestinal tract, especially in monogastric livestock and fish species. The great attention given to livestock farming is a direct consequence of its primary goal, which is represented by optimization of the feed efficiency and growth performance of the animals. This scenario has created a clear separation between humans and animals in terms of microbiota research, preparing the way for the rise of the concept of “gut health”.
Kogut and Arsenault firstly introduced the “gut health” concept as “the new paradigm in food animal production”, underlining its vital importance to the performance of production animals and defining it as synonymous with animal health within animal production industries [35]. Although a clear scientific definition of “gut health” is still lacking, six major components related to gastrointestinal functionality have recently been proposed [36]. Among them—including diet, effective digestion and absorption, functional immune status, gut mucosa, and neuroendocrine and motor function of the gut—a normal and stable microbiota has a key role because a symbiotic equilibrium between the intestinal tract and the microbiota is fundamental for maintaining the welfare and performance of the animals [36]. In particular, the gut microbiota benefits the host by providing useful nutrients (i.e., short chain fatty acids (SCFAs)) from poorly digestible dietary substrates (i.e., non-starch polysaccharides) and by modulating the development and functionality of the digestive (in terms of mucosal morphology and mucin dynamics) and the immune (in terms of antimicrobial peptides and pro-inflammatory cytokines production) systems [37].
As a consequence of the development of the “gut health” concept, the gut microbiota of the main monogastric livestock (such as poultry, pig, rabbit and fish) species has been extensively characterized, also sharing similarities among each other. Indeed, Firmicutes and Bacteroidetes bacterial phyla represent the intestinal core microbiota in all monogastric livestock (poultry [38], pig [39] and rabbits [40]) and fish species [41]. Proteobacteria is another dominant phylum in poultry and pig [38,39], but it is mostly identified in fish species, along with Actinobacteria and Fusobacteria. Within the Firmicutes phylum, Lactobacillaceae, Clostridiaceae and Ruminococcaceae represent the dominant bacterial families in the monogastric livestock species, with Clostridium, Ruminococcus and Lactobacillus being also the most abundant species in poultry [38], pig [39] and rabbit [40]. On the contrary, Carnobacteriaceae and Leuconostocaceae are characteristic of fish gut microbiota [42,43] Similarly, the Bacteroidetes phylum comprises Bacteroidaceae and Bacteroides as predominant members of the intestinal microbiota of the monogastric livestock species [38,40], while the fish gut microbiota mainly displays the Flavobacteriaceae family [44]. As a final aspect to consider, within the phyla Proteobacteria, Enterobacteriaceae, Campylobacteriaceae and Helicobacteriaceae represent the dominant bacterial families in the monogastric livestock species, with Helicobacter and Actinobacillus being also the most abundant species in poultry [38] and pig [39]. On the contrary, Aeromonadaceae and Vibrionaceae are characteristic of freshwater and marine fish intestinal microbiota, respectively [45,46].
Differently from the monogastric livestock and fish species, research about the gut microbiota in domestic animals (such as dogs and cats) share more similarities with the human field, since pets are nowadays considered as genuine family members and, as a consequence, all the factors potentially promoting their longevity (including the microbiome) are topics of great interest. Both canine and feline gut microbiota are dominated by the already mentioned Firmicutes, Proteobacteria and Bacteroidetes phyla [47]. However, the poor statistical power due to small sample sizes of pet studies has led to some inconsistencies in taxonomy findings across the studies [47]. Nevertheless, Alessandri et al. recently identified Fusobacterium, Prevotella and Bacteroides as major bacterial genera in the intestinal microbiota of both dogs and cats [48].

3. Intestinal Dysbiosis in Human and Animals: The Role of Antibiotics

3.1. Humans

The term “dysbiosis”, originally established by Metchnikoff to describe altered pathogenic bacteria in the gut [49], has also been defined as “qualitative and quantitative changes in the intestinal flora, their metabolic activity and their local distribution” [50]. In addition, antibiotics administration represents a cause of major alterations in normal gastrointestinal tract (GIT) [51] by inducing long-lasting changes in intestinal microbiota correlated with disease development [52]. The potential for an antimicrobial agent to influence the gut microflora is related to its spectrum of activity [51], pharmacokinetics, dosage [53] and length of administration [53]. Regarding the spectrum of activity, an antimicrobial agent active against both the Gram-positive and -negative organisms will have greater impact on the intestinal flora [51]. In terms of pharmacokinetics, the rate of intestinal absorption plays a fundamental role. In general, oral antimicrobials that are well absorbed in the small intestine will have a minor impact on the colonic flora, whereas agents that are poorly absorbed may cause significant changes. Parenteral administration of antimicrobial agents is also not free from these consequences, as some of these agents can be secreted in their active forms in bile or saliva, or from the intestinal mucosa, thus resulting in remarkable alterations in the colonic flora [54]. Microbiota alterations induced by a particular antibiotic might also be more severe in individuals who have been subjected to multiple courses of antibiotics. So far, antibiotics have been reported to increase Proteobacteria and Tenericutes but to decrease Bacteroidetes and Firmicutes. In particular, oral vancomycin may reduce the faecal microbial diversity by decreasing Gram-positive bacteria (mainly Firmicutes) and by increasing Gram-negative bacteria (mainly Proteobacteria) [55,56]. Indeed, potential biomarkers of antibiotic administration have been reported to be Klebsiella, Parasutterella, Morganella, unclassified_f_Enterobacteriaceae, Ureaplasma and unclassified_f_Ruminococcaceae (the majority of which belong to the Proteobacteria phylum). Based on these findings, Proteobacteria may be considered a potential, useful diagnostic sign of intestinal microbiota dysbiosis [55].

3.2. Monogastric Animals

Monogastric livestock farming has frequently been characterized by the indiscriminate use of antibiotics in order to control the development of diseases and to exploit their effects as growth promoters [57]. However, similar to that previously described for human patients, antibiotic administration may cause dysbiosis, a perturbation of the number and composition of the microbiota that affects physiological microbial balance [58]. Microbiota dysfunctions are generally associated with alterations in the digestive tract and in the metabolic processes, increase in nutrient absorption and downregulation of the immune response [57], with subsequent increase in pathogenic colonization and disease susceptibility [59]. Furthermore, the antibiotic-induced changes in microbiota composition can persist for months or years after cessation of antibiotic treatment, thus representing a long-term issue [58]. Another important aspect to consider is that misuse and overuse of antibiotics in livestock farming is closely related to the growing number of antimicrobial-resistant agents, thus raising important concerns about animal and human health [60] and necessitating the search for efficacious alternatives to their use [57].

3.2.1. Poultry

Till now, several research studies have investigated the relationship between antibiotic treatment and gut dysbiosis in monogastric livestock species. Simon et al. first reported that the administration of antibiotics in the first weeks of life of laying hens may cause dysbiosis of the intestinal microbiota, thus potentially leading to alterations in immune development and, in turn, the immune competence of the birds [61]. Furthermore, zinc bacitracin administration has recently been reported to increase the abundance and the number of potential pathogens while decreasing those with beneficial functions in yellow-feathered broiler chickens [62]. These microbiota imbalances were also accompanied by lower microbial carbohydrate fermentation, higher microbial catabolism of amino acids and upregulation of pro-inflammatory cytokines after the antibiotic treatment [62]. The abovementioned research study confirms and extends the observations of Kumar et al. which identified higher abundance of Campylobacter spp. and Salmonella spp. as well as pro-inflammatory cytokines gene expression in bacitracin-fed chicks when compared to the control birds [63]. Tiamulin administration has also been reported to trigger intestinal dysbiosis in a laying hen model of gastrointestinal infection by Brachyspira pilosicoli, thus leading to higher susceptibility to pathogenic colonization and possibly causing relapse [64].

3.2.2. Pigs

A similar scenario can also be highlighted in pigs, where weaning—the most critical part of their life cycle—is frequently associated with gut dysbiosis development. Antibiotic treatment may also further decrease microbial diversity and increase opportunities for pathogenic microorganisms to colonize and trigger diseases [60]. In particular, early antibiotic exposure in suckling piglets has been reported to influence intestinal microbiota, and carbohydrate and protein metabolism [59]. Connelly et al. also observed significant changes in a porcine model of ceftriaxone-mediated gut dysbiosis, where the intestinal microbiome was characterized by a reduced abundance of commensal species, outgrowth of colitis-causing species and increased frequency of antibiotic resistance genes [58]. A more recent porcine model of carbapenem-mediated intestinal dysbiosis confirmed the abovementioned findings, as commensal bacteria were undetectable after antibiotic treatment and multi-drug-resistant species were shown, instead, to have increased [65].

3.2.3. Fish

As far as fish are concerned, the misuse of antibiotics, overdose and poor absorption after medication, as well as their high water solubility and bioactivity raise remarkable environmental issues in terms of ecological sustainability and health problems [66]. In addition to that, high concentrations of antibiotics used for treatment of bacterial diseases (florfenicol, oxytetracycline and sulfamethoxazole) have also been reported to cause severe microbiota dysfunctions in various fish species, such as Atlantic salmon [67], hybrid and Nile tilapia [66,68,69] and Zebrafish [70]. In particular, reduction [67] or increase [70] in microbial diversity and growth-depressing effects on some beneficial bacteria [66,68] were observed. Gut dysbiosis was also accompanied by upregulation of the intestinal pro-inflammatory cytokines [66,69,70] as well as altered gut histomorphology and tight junction proteins [66]. A high fat diet has also recently been reported to worsen the adverse effects of antibiotic administration on intestinal health in juvenile Nile tilapia in terms of gut microbiota disfunctions and altered intestinal histomorphology and tight junction proteins [69].

3.2.4. Pets

Canine and feline intestinal dysbiosis can be described as an intestinal microbiota alteration of the composition and count (i.e., quantity of bacterial species) that is accompanied by a reduction in SCFA production [71,72]. Historically, bacterial proliferation of the small intestine has been used to describe qualitative and quantitative gut microbiota changes based on the juices of the duodenum and fasting cultures. Indeed, the bacterial population found in the proximal small intestine of several dogs has been reported to be substantially higher than that normally found in humans, where a bacterial count higher than 105 CFU/g or ml in aspirates of the small intestine indicates bacterial overgrowth (SIBO). However, subsequent investigations have shown that healthy dogs could have bacterial loads that far exceed these established limits [73]. Therefore, SIBO is now a controversial definition and the terms antibiotic responsive diarrhoea (ARE) or small intestinal dysbiosis are preferred. More recently, the state of dysbiosis has been correlated to changes in the gut microbiota associated with gastrointestinal diseases—mostly affecting the large intestine—such as inflammatory bowel disease (IBD), granulomatous colitis, irritable bowel syndrome (IBS) [74], chronic enteropathy (CE), acute diarrhoea [75] and parasitic diseases (e.g., Giardia duodenalis) [76]. Although commensal bacterial microbiota deficiencies are linked to metabolic changes, it is not yet well investigated whether dysbiosis can be a cause or a consequence related to of gastrointestinal diseases. An overlap between these two scenarios seems reasonable, since inflammatory processes cause dysbiosis and recent functional studies have shown that dysbiosis, when present, may aggravate inflammation in genetically sensitive individuals [77]. In canine IBD, a proportional increase in bacterial genera belonging to the Proteobacteria phylum as well as a decrease in Fusobacteria, Bacteroidetes and members of Firmicutes has been observed in both the duodenum and the faeces [74,78]. Generally, there is a similarity in dysbiosis patterns between chronic and acute diarrhoea, even if some notable differences may be identified. Indeed, substantial increases in the populations of Fusobacteria have been observed in faecal samples from dogs with acute haemorrhagic diarrhoea [79], while C. perfringens overgrowth seems to characterize intestinal dysbiosis in chronic diarrhoea [74]. Unlike for those of dogs, only a few studies have focused their attention on faecal microbiota changes in cats with consolidated IBD. Janeczko et al. reported an increased duodenal count of enterobacteria in cats with IBD by fluorescence in situ hybridization, with these counts showing positive correlation with changes in mucosal architecture and cell infiltrates density [80]. An increase in Desulfovibrio spp. has also been observed [81], while Abecia et al. identified no differences between healthy cats and cats with IBD [82]. In a more recent study using 16S rRNA sequencing in cats with acute and chronic enteropathy but without a clear diagnosis, cats with chronic diarrhoea showed reduced proportions of Bacteroidetes, Fecalibacterium spp. and Turicibacter spp. and increased abundance of enterobacteria in their faecal microbiota similar to dogs with IBD [83]. However, no studies have clearly assessed whether dysbiosis patterns differ among the various forms of CE, food responsive enteropathy (FRE), ARE and IBD.
Antibiotic therapy, in association with dietary modifications, is commonly administered in dogs and cats with chronic enteropathy [84], the main antibiotic molecules represented by metronidazole and tylosin [85]. Metronidazole acts on bacteria and protozoa, while tylosin is useful dispensed during chronic tylosin-responsive diarrhoea that generally affects adults [85]. Moreover, specific enteropathies (such as boxer granulomatous colitis) usually require other antibiotic molecules belonging to the fluorquinolones class (e.g., enrofloxacin) [86]. Nevertheless, it is commonly known that antibiotic therapy may cause significant changes in the intestinal microbiota composition in either dogs or cats [77,87]. Indeed, microbiota alterations (i.e., increased abundance of organisms similar to Escherichia coli) may be still observed 14 days after tylosin withdrawal [88]. A more recent study in healthy dogs confirmed that the administration of tylosin induced dysbiosis but also demonstrated that eubiosis was not restored within 56 days after stopping tylosin, thus suggesting that the restoration of native microbiota is possible but not guaranteed [89]. Faecal bacterial diversity has also been reported to be reduced during the administration of oral metronidazole to healthy dogs for 14 days [90]. Similarly, oral administration of amoxicillin to healthy dogs for 7 days showed differences in faecal bacterial composition before and after, with several E. coli faecal isolates with greater resistance to more antibiotics during treatment [91]. In contrast, Kilpinen et al. demonstrated that dogs with tylosin-sensitive diarrhoea have shown an increase in Enterococcus spp. and other potentially probiotic bacteria (including lactic acid bacteria) in their faecal microbiota [92]. Therefore, it is thought that tylosin may have a certain probiotic action by exerting a selective increase in tylosin-resistant enterococci [92]. Nevertheless, even if these results appear interesting, there are still doubts about whether antibacterial resistance can pass horizontally from commensal or presumed probiotic bacteria to pathogenic bacteria that share the same intestinal environment [93]. Furthermore, the metabolic pathways through which antibiotics can alter intestinal homeostasis are still being investigated.

4. Nutritional Alternatives to the Antibiotics and Effects on Intestinal Health: The Role of Probiotics

4.1. Humans

In 2013, the Food and Agriculture Organization of the United Nations (FAO) and the World Health Organization (WHO) scientific committees defined probiotics as “living microorganisms which, if administered in adequate quantities, confer a benefit for the health of the host” [94]. Indeed, probiotics are constantly administered to balance the homeostasis of the gut microbiota in order to maintain human intestinal health, to prevent and treat acute and chronic gastrointestinal diseases [95] by inhibiting the infiltration and growth of pathogenic bacteria in the GIT, and to prevent infections in patients with antibiotic-induced gut dysbiosis [96]. It has been suggested that probiotics may be beneficial and safe for the prevention of antibiotic-associated diarrhoea, as demonstrated in several randomized controlled trials (RCT). In particular, probiotics have been reported to be efficacious in the prevention of C. difficile-associated diarrhoea in both adults and children [97,98]. A previous review also suggested that the protective effects of probiotics as adjunct therapy may be used in the prevention of antibiotic-associated diarrhoea in outpatients of all ages, with probiotic intervention being capable of reducing its onset by 51% without any apparent side effects [99]. Indeed, several researches revealed that dietary Lactobacillus spp. and/or Streptococcus spp. supplementation reduced the incidence of antibiotic-associated diarrhoea in seniors (12.4% vs. 23.7% [100]), adults (6.9% vs. 14.2% [101]) and children (7.3% vs. 26.3% [101]).
In the past decades, probiotics have also gained increasing attention as a potential therapy against IBD—encompassing Crohn’s disease (CD), ulcerative colitis (UC) and indeterminate colitis (IC)—which can be differentiated by localization of the inflammation in the GIT [102]. However, despite probiotics having been used to induce remission and maintenance therapy in UC [103], a recent study indicated that probiotic supplementation in patients with IBD is a promising adjuvant treatment in UC but not in CD [102]. Vanderhoof et al. first observed that the administration of Bifidobacterium longum along with a prebiotic mix was capable of downregulating the expression of TNF-α and IL-1α pro-inflammatory cytokines and of attenuating inflammation in the intestine of IBD patients [97]. Dietary Lactobacillus acidophilus and Bifidobacterium animalis supplementation has also been reported to positively modulate the gut microbiota (in terms of increased Lactobacillus and Bifidobacterium) in IBD patients [99]. Furthermore, Yang and Yu reported that CD patients administered with a mixture of Lactobacillus, Bifidobacterium and Streptococcus thermophilus probiotics were characterized by a reduction in the mucosal inflammatory cytokine levels and severe recurrence of symptoms [104]. Considering that people with UC or CD, especially young people [105], are highly predisposed to develop colorectal cancer (CRC) [100], recent studies have demonstrated the potential role of probiotics in its prevention through modulation of the intestinal microbiota (in terms of increased Lactobacillus [106]) or metabolic activities [107]. The administration of L. acidophilus, Bifidobacterium longum and Enterococcus fecalis probiotics has also been reported to reduce diarrhoea incidence in patients affected by CRC [108]. Furthermore, selected strains of Clostridia may act as live biotherapeutic products to treat certain forms of IBD, allergy and other immune-inflammatory diseases thanks to the production of SCFAs and other metabolites [109].

4.2. Monogastric Animals

Similar to humans, the high occurrence of intestinal dysbiosis and the increasing development of antibiotic-resistance issues have progressively led to the search of efficacious alternatives to the use of antibiotics in monogastric livestock species and fish. Among the different agents that have been suggested to substitute antibiotics, probiotics (such as Lactobacillus spp., Bacillus spp., Clostridium spp. and Saccharomyces spp.) are widely considered to be one of the most effective because they preserve or restore the physiological microbiota, inhibit adhesion of pathogens to the intestinal wall, prevent gut inflammation and maintain the intestinal barrier function [110]. The positive effects of probiotic administration in monogastric animals in terms of improved growth performance, selection of beneficial bacteria, reduction of potential pathogens and stimulated immune response have extensively been characterized by several research studies and reviews, but there is limited information about their potential therapeutic or preventive role during the occurrence of specific gastrointestinal diseases (especially in relation to gut health implications).

4.2.1. Poultry

Standard Lactobacillus-based probiotic administration (Lactobacillus casei 1.2435, Lactobacillus rhamnosus 621 and L. rhamnosus A4) has recently been reported to increase the relative abundance of beneficial bacteria and to decrease the relative abundance of harmful bacteria in the gut from Cherry Valley ducks infected with E. coli 77 [111]. Similar findings were also obtained when supplementing Lactobacillus plantarum K34 together with Bacillus subtilis MORI 91 and Clostridium butyricum M7 in E. coli O78-challenged broiler chickens in terms of increased total lactobacilli and total lactobacilli-enterococci populations in the intestinal microbiota and reduced isolation of E. coli 078 from liver and spleen [112]. El-Sharkawy et al. also recently reported that administration of different probiotic strains (such as L. casei ATTC334, Bifidobacterium breve JCM1192 and Bifidobacterium infantis BL2416) reduced Salmonella Typhimurium recovery from the caecal tonsils by competitive exclusion mechanism or by increased IFN-γ and TNF-α production [113]. Dietary E. faecium supplementation may also decrease the adverse effects of eimerian infection in experimentally infected broilers [114].
C. butyricum is a butyric acid-producing Gram-positive anaerobe that has been used as a probiotic to decrease clinical signs during IBD and antibiotic-associated diarrhoea in human patients [115]. In the wake of such positive outcomes, dietary supplementation of C. butyricum has been reported to improve the intestinal immune response, morphology and the digestive enzyme activities in broiler chickens challenged with E. coli K88 [116]. Zhao et al. also observed a decrease in Salmonella enteriditis colonization and reduced production of pro-inflammatory cytokines in liver, spleen and caecum from S. enteriditis-infected broilers [117]. In a more recent broiler model of necrotic enteritis, dietary administration of C. butyricum was also capable of modulating the immune response, of reducing the permeability and of decreasing C. perfringens counts in the gut [118].
Bacillus spp. represent another interesting probiotic for poultry due to their ability to produce heat-resistant spores and to their tolerance to the low pH, bile and enzymes encountered in the upper gastrointestinal tract of chickens [119]. In particular, dietary B. subtilis DSM 32315 has been reported to increase microbial alpha-diversity [120] and the relative abundance of beneficial bacteria as well as to decrease the relative abundance of C. perfringens in the gut from broiler chickens with necrotic enteritis [120,121].
Saccharomyces cerevisiae is a well-known yeast also exerting a probiotic influence in broilers by promotion of the metabolic processes of digestion and nutrient utilization. In particular, Massacci et al. has recently reported improved gut morphology and microbiota in Campylobacter jejuni challenged chickens after dietary S. cerevisiae boulardii CNCM I-1079 supplementation [122].

4.2.2. Pigs

A similar scenario can also be highlighted in pigs, where the most investigated probiotics are represented by L. rhamnosus, L. acidophilus and B. subtilis.
Li et al. firstly recommended a safe threshold for using L. rhamnosus ACTT 7469 in clinical practice, since high doses may decrease the prophylactic benefits against potential enteric pathogens. Indeed, despite both the low and the high doses of probiotic administration ameliorating diarrhoea, reducing coliform shedding in faeces, increasing intestinal Lactobacillus and Bifidobacterium counts and attenuating the increase in gut pro-inflammatory cytokines in F4 (K88)-positive E. coli-infected piglets, high doses may negate the preventative effects by disturbing the established microbial ecosystem and by interfering with mucosal immune responses against potential enteric pathogens [123]. Zhang et al. also observed an attenuation of the expansion of Prevotellaceae NK3B31; the promotion of a symbiotic synergism of Fusobacterium, Lactobacillus animalis and Propionibacterium; and a restriction in chronic inflammation in the gut of Salmonella enterica serovar 4-infected pigs after administration of L. rhamnosus GG [124]. These findings partially confirm what was previously reported by the same research group, where dietary L. rhamnosus GG supplementation ameliorated the intestinal inflammation caused by Salmonella Infantis in a pig model, also excluding Salmonella from colonization of the jejunal mucosa and increasing the relative abundance of Bifidobacterium [125]. Nordeste et al. also investigated the efficacy of bioactive molecules isolated following the growth of L. acidophilus (called proteobiotics) in piglets challenged with Escherichia coli K88, observing increased abundance of Lactobacillus and reduced colonization by Escherichia coli K88 after prebiotic administration [126].
Similar to the effects of L. rhamnosus ACTT 7469 in Escherichia coli-infected piglets, different doses of a B. licheniformisB. subtilis mixture have been reported to lead to contrasting outcomes. Indeed, pretreatment with low and moderate doses of the probiotic mixture attenuated the E. coli-induced expansion of selected bacterial species and increased the number of goblet cells in the gut of the infected piglets, while high doses disrupted colonic microbial ecology (in terms of increased abundance of Proteobacteria) and impaired goblet cell function in the ileum [127]. However, Luise et al. recently observed that dietary B. subtilis supplementation promoted the gut health of E. coli F4ac-infected piglets by reducing the Enterobacteriaceae level, favouring the upregulation of genes related to immunity and improving the amino acid metabolism and utilization [128].

4.2.3. Fish

The use of probiotics (especially Enterococcus casseliflavus, Rummeliibacillus stabekisii and Bacillus spp.) in fish farming has primarily been focused on establishing effective prevention strategies (rather than therapeutic) towards Aeromonas hydrophila and Streptococcus iniae infections.
Dietary E. casseliflavus supplementation has been reported to significantly improve the host resistance to S. iniae infection in challenged rainbow trout by increasing the intestinal lactic acid bacteria and total viable aerobic counts and by stimulating the immune response of the animals [129]. Similarly, Tan et al. observed enhanced cumulative survival in Nile tilapia fed R. stabekisii after challenge with A. hydrophila and S. iniae thanks to stimulation of the immune response of the animals and positive modulation of their gut microbiota [130].
Administration of Bacillus-based probiotics (Bacillus velezensis TPS3N, B. subtilis TPS4 and B. amyloliquefaciens TPS17) has recently been observed to reduce the mortality rates in A. hydrophila-challenged Nile tilapia by enhancing both skin mucus and intestinal immune response, by improving the antioxidant defence, by enhancing the gut morphological and by positively modulating the intestinal microbiota [131]. Ahmadifard et al. also reported the lowest cumulative mortality in ornamental fish Poecilia latipinna pretreated with B. subtilis-enriched Artemia after A. hydrophila (BCCM5/LMG3770) challenging as a consequence of increased abundance of Bacillus in the gut [132].

4.2.4. Rabbits

Unlike other monogastric animals, data regarding the potential therapeutic and/or preventive role of the probiotics against specific gastrointestinal diseases are very limited.
Ogawa et al. firstly reported that dietary L. casei strain Shirota supplementation enhanced local gut immune responses to Shiga toxin-producing E. coli (STEC) and Shiga toxins (Stx) 1 and 2 in STEC-infected infant rabbits, thus leading to the elimination of STEC, decrease in the Stx concentrations in the intestine and attenuation of the histological damage to gut mucus [133]. Administration of L. casei SABA6 strain in experimental rabbits challenged with Escherichia coli AZ1 has also recently been observed to attenuate the eruption of the intestinal epithelial cells, to reduce the incidence of diarrhoea and to stimulate the immune response of the animals [134].

4.2.5. Pets

Despite the abovementioned definition of “probiotics” by FAO and WHO being worldwide accepted, health benefits associated with the administration of probiotics in pets are not strictly related to curing a specific disease. Therefore, it appears more appropriate to define probiotics as “microorganisms with the intention of improving the health of the host” [135], especially considering that their clinical effects have recently been characterized [136]. The EU Regulation 1831/2003 classifies probiotics as zootechnical additives, which are marketed only after specific authorization. At present, only four bacterial probiotic strains have been analysed by the European Food Safety Authority (EFSA) to evaluate safety and efficacy as feed additives in companion animals: two E. faecium strains (E. faecium NCIMB 10415 E1705 and E. faecium NCIMB 10415 E1707), L. acidophilus DSM 13241 25 and B. animalis [135]. As far as E. faecium NCIMB 10415 E1707 is concerned, the EFSA concluded that sufficient information was provided to consider its use as safe in pets and humans that have connections with the administered subjects. Indeed, E. faecium NCIMB 10415 E1705 has been defined as an unlikely cause of danger in target species even if overdosed, being also incapable of promoting the growth and spread of haemolytic and non-haemolytic E. Coli strains in dogs [135]. Many commercially available probiotics for pets contain Enterococci (mainly E. faecium or E. fecalis), which are natural inhabitants of the canine gastrointestinal tract. Specific strains can exert numerous benefits on the host such as probiotics, although Enterococci can also be pathogenic and have the ability to develop rapidly, to spread resistance to antibiotics and to promote the growth of potentially harmful microbes in animals and humans [137]. In relation to L. acidophilus DSM 13241 25, the EFSA revealed no safety concerns as the strain in question is sensitive to most antibiotics (excepted ciprofloxacin). The last examined probiotic strain, in 2012, was B. animalis, albeit the effects of B. animalis on health parameters related to the GIT in dogs were considered to have a questionable biological relevance, thus not allowing the EFSA to entrust the efficacy of this product. Adding to the abovementioned strains, in Europe and United States, though limited safety and efficacy data are available, the following strains are also marketed: lactobacilli (L. acidophilus, L. casei, L. plantarum, L. paracasei, L. lactis, L. rhamnosus and L. salivarius), bifidobacteria (B. infantis, B. lactis, B. longum and B. bifidum), B. subtilis or coagulans and, in some cases, yeasts (Saccharomyces cerevisiae) or fungi (Aspergillus oryzae).
From a general point of view, probiotics compete with potential pathogens by interposing with their adhesion to the intestinal mucosa or by promoting mucin release [138]. These mechanisms are believed to be specific of each strain, with some strains having greater adherence capacities (i.e., L. rhamnosus) and other strains being capable of increasing the adherence of pathogens to the intestinal mucosa [138]. In addition, probiotic bacteria can produce various antimicrobial substances, such as fatty acids, lactic acid and acetic acid [135,139]. To remain viable, probiotics must outlive stomach acidic environment and bile acids in order to successfully colonize the intestine, where they can bring advantages [140] by improving intestinal microbiota composition in terms of increased beneficial bacteria at the expense of unwanted ones [141,142]. However, it is important to underline that the production methods, the food carrier and the growth media can widely influence the original properties of probiotics, in particular, the adhesiveness capacity to the intestinal mucosa of canine lactobacilli [143]. Although lactobacilli represent only a small part of the canine gastrointestinal microbiota, they are still widespread, and several isolated strains, including Lactobacillus spp., demonstrate antimicrobial activity in vitro [144,145], survive and dominate the microbiota of the small intestine during feeding, and are capable of modifying the intestinal micro-ecosystem [146]. In addition, L. fermentum VET9A, L. plantarum VET14A and L. rhamnosus VET16A (administered alone or together) have been reported to show good adhesion to the canine intestinal mucosa and the ability to prevent the adhesion of common enteropathogens (such as Enterococcus canis, C. perfringens and S. Typhimurium) to the canine intestinal mucosa in vitro [147]. Strompfova et al. also observed increased faecal lactobacilli and enterococci, increased total serum proteins and lipids and decreased blood glucose as well as increased production of SCFAs and reduced proliferation of clostridia and selected Gram-negative bacteria (such as coliforms, Aeromonas spp. and Pseudomonas spp.) in the follow-up in dogs supplemented with L. fermentum AD1 (a probiotic originally isolated from the faeces of healthy dogs) [148]. The administration of canine E. faecium EE3 has been reported to reduce Staphylococcus spp. and Pseudomonas-like bacteria and to increase the concentration of Lactobacillus spp. [149]. Pascher et al. also observed that the administration of L. acidophilus DSM1324 increased faecal consistency, faecal dry matter, defecation frequency, faecal lactobacilli and bifidobacteria and reduced C. perfringes and Escherichia spp in German Shorthair Pointers dogs [150]. Furthermore, some authors think that bacilli represent better probiotics than lactobacilli, since they can sporulate and, in turn, may be more resistant to environmental stress and low pH [151]. However, independently of their survival, all the probiotics must be tested for their benefits in clinical situations. Indeed, Felix et al. found that B. subtilis C-3102 improved faecal consistency in dogs due to the decrease in faecal ammonia content, but the clinical relevance of this discovery is highly questionable, thus making the use of bacilli as probiotics not recommended [151].
Probiotics also appear to be a promising tool for alleviating gastrointestinal diseases in pets. Herstad et al. observed that the administration of a probiotic cocktail including Lactobacillus farciminis, Pediococcus acidilactici and L. acidophilus MA 64/4E was capable of reducing recovery time in acute self-limiting gastroenteritis in dogs [152]. As far as the acute canine gastrointestinal infections are concerned, dietary probiotic supplementation containing four strains of Lactobacillus (L. casei, L. plantarum, L. acidophilus and L. bulgaricus), three strains of Bifidobacterium (B. longum, B. short and B. infantis) and one strain of Streptococcus salivaris (Streptococcus sulivarius subsp thermophilus) has also been reported to increase the survival rates (90% vs. 70%) and to improve the clinical score and white blood cell/lymphocyte count in puppies with confirmed parvoviral enteritis [153]. Probiotics are not always capable of colonizing the GIT as a consequence of the competition with the intestinal microbiota that is already present [154]. Indeed, Garcia-Mazcorro et al. demonstrated that the abundant growth of Enterococcus spp. [155] and Streptococcus spp. promoted by oral integration of a synbiotic product (probiotic and prebiotic) was only transitory. Another study also showed a small growth of different species after the administration of Enterococcus faecium-based probiotic [154].
As a final consideration, some of the most recent studies involving probiotic molecules and their impact on intestinal health are reported (Table 1).

5. Other Nutrients Showing a Synergic Effect with Probiotics on Intestinal Health: Prebiotics and Postbiotics

5.1. Prebiotics

5.1.1. Humans

Another useful approach to positively modulate the gut microbiota is represented by the administration of growth enhancing substrates—the so-called “prebiotics”—that can selectively be utilized by health promoting bacteria in order to stimulate their growth and the production of associated desirable metabolites. More recently, the term prebiotic was updated in December 2016 when the International Scientific Association for Probiotics and Prebiotics (ISAPP) described it as “a substrate that is selectively utilized by host microorganisms conferring a health benefit” [160]. Currently, a major prerequisite to the success of a particular prebiotic is that the target bacteria must be identifiable at a specific threshold. Indeed, if levels are below a certain threshold (as a consequence of ageing or antibiotic administration), the prebiotic may not be effective in increasing these desirable bacteria to confer the abovementioned health benefits [161]. Nevertheless, other important features need to be considered. While prebiotics can enable the growth of targeted health-promoting bacteria, more detailed microbiome analyses have, however, highlighted that these substrates are not always specific [160,162,163]. Furthermore, the benefits of prebiotics can be secondary, such as the production by health promoting bacteria of metabolic products that can inhibit the growth of enteric pathogens and/or attenuate their virulence [164,165,166].
Among the best characterized prebiotics, galactooligosaccharides (GOS), inulin and its oligomer fructooligosaccharides (FOS) appear to be the most promising ones [164,167,168]. In particular, the scientific community has mainly focused its attention on their effects on Bifidobacterium and Lactobacillus, since strains from these genera are recognized to confer health benefits to the host [169]. The GOS and FOS have overall been reported to improve the gut microbiota by increasing Bifidobacteria and by decreasing E. coli [170,171]. Nicolucci et al. also observed decreased abundance of Firmicutes and Ruminococcus as well increased abundance of Bifidobacteria in faecal samples after FOS-enriched inulin supplementation [172]. However, the effects of prebiotic therapy also depend on individual’s starting microbial profile. Indeed, in a study comparing FOS, sorghum and arabinoxylan, equally high SCFA production was observed in volunteers whose microbiota were dominant in fibre-utilizing Prevotella, while Bacteroides-dominated enterotypes showed different SCFAs levels in response to each fibre [173].
Pectin and pectic oligosaccharides (POS) have recently been identified as emerging prebiotics as they can selectively be utilized by certain members of the colonic microbiota such as E. eligens and Faecalibacterium prausnitzii [174,175]. The most common sources of pectin are citrus fruits and apple pulp, but it is also abundant in agricultural by-products such as sugar beet pulp. The POS can be obtained through depolymerization of pectin, and both the pectin and the POS are capable of escaping host digestion and, in turn, reaching the distal colon when consumed [176]. An in vitro assay of F. prausnitzii growth on different substrates firstly ascertained that growth was enhanced with apple pectin in the majority of cases [174,177]. Furthermore, lemon peel waste POS determined an increased level of Fecalibacterium and Roseburia spp. as well as lactobacilli [178].
Isomaltooligosaccharides (IMO) are also considered potential prebiotics. The IMOs are naturally found in foods such as honey or fermented foods such as miso and soy. Commercially available IMOs are composed of a mixture of α-(1,6) and α-(1,4)-linked glucosyl oligosaccharides. One of the glucose oligomers identified in IMOs is isomaltose, which is a major constituent of honey (thus giving IMOs a distinctive sweet honey taste). Breast milk is the natural, first nutritional source for newborns and provides oligosaccharides that promote the growth of desirable Bifidobacterium and Lactobacillus spp. in infant gut [179]. Indeed, the human milk oligosaccharide (HMO) and other nutrients in breast milk can act as substrates for bacteria in the gut, thus stimulating the growth of beneficial bacteria herein located. In particular, dietary HMO supplementation has recently been reported to modify the gut microbiota of adult volunteers by increasing the relative abundance of Actinobacteria (in particular Bifidobacterium) and by decreasing the relative abundance of Firmicutes and Proteobacteria [180].
Soybean oligosaccharides also seem to have a positive influence on Bifidobacterium levels and lactate production. In particular, soybean benefits are attributed to resistant starch (RS), which is a form of starch that cannot be degraded in the small intestine and, in turn, is capable of reaching the large intestine where it is fermented by the colonic bacteria [181]. Potato and maize RS have also recently been reported to modulate the gut microbiota of healthy volunteers in terms of increased Ruminococcus bromii and Bifidobacterium spp., with potato RS also leading to a greater SCFA production when compared to maize and inulin [182].
As a final consideration, synbiotics, which are a combination of prebiotics and probiotics meant to strengthen the effects of probiotics administered alone [183], also need to be mentioned, since they are capable of improving the survival of dietary live microbial supplements in the GI tract and of selectively stimulating the growth and/or activating the metabolism of health-promoting bacteria [184]. In particular, inulin has successfully been tested as a synbiotic component in the treatment of the ulcerative colitis [185,186].

5.1.2. Monogastric Animals

As in humans, the synergic utilization of prebiotics with probiotics allows manipulation of the gut microbiota of animals, thus controlling the spread of bacterial infections and, in turn, avoiding the overuse of antibiotics [187]. In particular, the scientific community has recently focused its attention on the use of prebiotics (such as inulin, yeast cell products, mannanoligosaccharide (MOS), FOS, GOS, xylooligosaccharide (XOS) and IMO) as potential therapeutic or preventive feed additives during the occurrence of specific gastrointestinal diseases (especially in relation to gut health implications).

Poultry

Dietary live yeast (S. cerevisiae) and MOS supplementation has firstly been reported to attenuate the intestinal inflammation and barrier dysfunction as well as to improve gut morphology and microbiota in Escherichia coli-challenged broiler chickens [188]. Similarly, Rahimi et al. observed enhanced intestinal morphology as well as reduced shedding of Salmonella in turkey poults challenged with Salmonella and Campylobacter after administering a mixture of S. cerevisiae wall, MOS and a direct-fed microbial [189]. Dietary whole yeast cell product alone has also been reported to decrease faecal coccidial oocyst count and intestinal coccidial oocyst count, to increase the relative abundance of Lactobacillus and to attenuate the inflammatory response in Eimeria-challenged pullet and layer chickens [190].
As far as other prebiotics are concerned, Pourabedin et al. similarly showed reduced S. enteriditis counts, increased abundance of beneficial bacteria and a significant attenuation of the increase in pro-inflammatory cytokines in MOS- and XOS-fed young broiler chickens after infection with S. enteriditis [191]. The concomitant utilization of XOS, Lactobacillus-based probiotic and fermented soybean meal has also recently been reported to decrease gut S. Typhimurium colonization, to increase intestinal lactic acid bacteria, to improve gut mucosa morphology and to reduce the heterophil to lymphocyte ratio in S. Typhimurium-challenged young broiler chickens [192]. A significant reduction in S. enteriditis [187] and Typhimurium [193,194] loads as well as positive microbiome changes [194] were also observed in Salmonella-challenged white Leghorn chickens simultaneously administered with synbiotic products (FOS and probiotics) [187] or GOS [193,194]. Tarabees et al. finally reported increased total lactobacilli and total lactobacilli-enterococci populations in the intestinal microbiota along with reduced isolation of E. coli 078 from liver and spleen in broiler chickens fed diets supplemented with IMO and a commercial probiotic mix during E. coli O78 challenging [111].

Pigs

Apart from the abovementioned prebiotics, other unconventional, prebiotic-like feed additives have also been tested in pigs. The utilization of fructan-rich chicory roots has first of all been reported to reduce the intestinal Ascaris suum worms; to increase the intestinal Trichuris suis worms; to decrease the intestinal levels of Enterobacter and Campylobacter; and to increase the intestinal levels of Lachnospiraceae and Bifidobacteria in piglets experimentally infected with Ascaris suum, Trichuris suis and E. coli O138:F8 [195]. Myhill et al. also observed a synergic activity of inulin and Trichuris suis in a porcine model of Trichuris suis infection in terms of stimulation of the immune response, reduced abundance of Proteobacteria and Firmicutes, increased abundance of Actinobacteria and Bacteroidetes, and identification of the highest Bacteroidetes–Firmicutes ratio and the lowest gut pH [196]. The same authors recently confirmed that the administration of Trichuris suis together with the consumption of inulin may positively affect the gut health of the infected pigs by increasing the relative abundance of Prevotella and by decreasing the relative abundance of Proteobacteria [197]. Interestingly, dietary human milk oligosaccharide (HMO) and GOS-FOS mixture supplementation has also been reported to shorten the duration of diarrhoea, to enhance the immune system and to increase the relative abundance of butyrate-producing Lachnospiraceae in a porcine model of rotavirus OSU infection [198].

Fish and Rabbits

Unlike other monogastric animals, data regarding the potential therapeutic and/or preventive role of the prebiotics against specific gastrointestinal diseases in fish and rabbits are quite scarce.
The administration of a plant-based diet has firstly been reported to exert a prebiotic effect in Yersinia ruckeri-challenged rainbow trout by positively modulating their gut microbiota and immune response [199]. Similarly, Wang et al. observed enhanced levels of immune enzyme activities, increased proportions of Bacillus and Lactococcus and reduced levels of Vibrio in the intestine from sea cucumber fed symbiotic-based diets (containing the probiotic Bacillus lincheniformis WS-2 and the prebiotic alginate oligosaccharide) as prevention against Vibrio infections [200].
As far as rabbits are concerned, the administration of a prebiotic mixture (composed of S. cerevisiae cell wall, MOS and dried S. cerevisiae fermentation solubles) has recently been reported to reduce the faecal oocyst count and the intestinal endogenous stage counts in rabbits experimentally infected with Eimeria spp. [201].

Pets

Prebiotics are used in both healthy and sick dogs, being particularly effective in the management of obesity and gastrointestinal diseases. Indeed, Zentek et al. firstly observed increased Bifidobacteria and decreased C. Perfringens in faecal samples from healthy dogs fed diets containing 3% of chicory (1.5% of inulin) [202]. Similarly, the administration of FOS has been reported to increase Bifidobacteria and mineral digestibility in healthy dogs [203]. Alexander et al. also recently observed accelerated excretion of bile acids, an increase of Firmicutes and a decrease of Proteobacteria in obese dogs after inulin supplementation at 0.5–1% [204]. Furthermore, the administration of chondroitin sulphate and prebiotic-mix (consisting of RS, beta-glucans and MOS) has been reported to improve the serum oxidative stress markers (such as PON 1 and TAC) and cholesterol levels in IBD dogs [205]. Synbiotics may also represent useful products, as reduced diarrhoea rates may be identified in dogs after dietary supplementation with E. faecium, arabic gum and FOS [206].
Unlike dogs, research studies on the utilization of prebiotic substances in cats are very limited, as they consume few vegetables (and, in turn, carbohydrates) in their natural diet [207]. However, dietary FOS supplementation has previously been reported to increase the relative abundance of Bifidobacteria (alone) and lactobacilli (together with pectin) as well as to decrease the relative abundance of E. coli (alone) and C. perfringens (together with pectin) in cats [208].
As a final consideration, some of the most recent studies involving prebiotic molecules and their impact on intestinal health are reported (Table 2).

5.2. Postbiotics

5.2.1. Humans

Postbiotics are bioactive compounds generally considered as secondary products of food quality microorganisms during a fermentation process, including microbial cells, cell constituents and metabolites. In detail, these functional fermentation compounds can be administered in combination with nutritional components to promote health or to increase the effectiveness of active microorganisms by transforming them into functional ingredients. Postbiotic effects derive from those microbial metabolites such as proteins, lipids, carbohydrates, vitamins, organic acids, cell wall components or other complex molecules produced in the fermented matrix [211,212]. The molecular mechanisms underlying the effects of postbiotics seem to be represented by the interaction between host and microbial products, thus activating the host’s immune system and, in turn, anti-inflammatory responses [213].

Short Chain Fatty Acids

The SCFAs (such as acetate, pronionate and butyrate) are important components of postbiotic products [214]. Primary SCFAs are well known to stimulate proliferative effects on colonocytes and to promote the absorption of water and sodium in the intestine. The most abundant SCFA detectable in human peripheral circulation is acetate, as propionate is metabolized by the liver during gluconeogenesis and butyrate is absorbed and used as a primary source of energy by colonocytes [215,216]. Furthermore, butyrate is linked to numerous beneficial health effects such as strengthening intestinal barrier function and mucosal immunity [217,218,219]. Furthermore, butyrate and, to a lesser extent, propionate are known to act as inhibitors of histone deacetylase, allowing them to exert anti-inflammatory and immune activity through the suppression of lamina propria macrophages and to cause the differentiation of dendritic cells from stem cells. bone marrow [216,220,221,222]. The SCFAs can also modulate cellular activity in the extracellular matrix via identifiable SCFA-specific G protein-coupled receptors on intestinal epithelial cells [223,224]. Another interesting aspect to underline is that the pathways for SCFA biosynthesis from the fermentation of indigestible fibres stimulate a bacterial cross-feeding complex involving several SCFA synthesis pathways to produce acetate, propionate and butyrate [216,225].

Antimicrobial Peptides (AMPs)

Postbiotic compounds might also play a crucial role in the inhibition of pathogens, since they contain the antimicrobial peptides (AMPs) [226]. The AMPs represent a wide range of molecules produced by the living organisms as a natural barrier against infections. In particular, the AMPs exhibit a wide range of activities against Gram-negative and Gram-positive bacteria, fungi, viruses and parasites. Among the AMP sources, bioactive agents derived from food proteins (mainly milk proteins) are particularly interesting because they are recognized as safe for both the humans and the animals [227,228].
Bovine and human milk contain a wide array of bioactive components such as the iron-binding glycoprotein known as lactoferrin (hLf in humans and bLf in bovine species). In both humans [229] and cows [230], the Lf concentration in milk changes according to the stage of postpartum lactation, with the highest concentrations being identified in the colostrum. This protein displays antibacterial, antifungal, antiparasitic, antiprotozoal [231,232], antiviral [233,234,235,236,237], immunomodulatory [238], anti-inflammatory, anti-catabolic and antioxidant activities [239,240,241,242], which overall contribute to the maintenance of homeostasis and the control of intestinal life-threatening diseases (especially in neonates [243]). In particular, the human Lf has been reported to show a very effective response against a wide range of bacteria, including Streptococcus, Salmonella, Shigella, Staphylococcus and Enterobacter spp. [244].

5.2.2. Monogastric Animals

Similar to humans, postbiotics mainly consist of antimicrobial metabolites such as organic acids and bacteriocins that are capable of reducing the gut pH and, in turn, inhibiting the proliferation of opportunistic pathogens in the feed and gut of animals [245]. The postbiotics that are most commonly adopted as antibiotic alternatives in monogastric animals are represented by metabolites from Lactobacillus and, in the minority of cases, Pediococcus, Enterococcus, Leuconostoc, Rhodotorula and Cetobacterium (with the latter three being exclusively used in fish).

Poultry

The administration of postbiotics obtained from L. plantarum has firstly been reported to exert beneficial effects on the gut environment and immune system of healthy laying hens [246] and broilers [247,248] in terms of reduced pH [246], increased abundance of lactic acid bacteria [246,247] and Bifidobacteria, decreased abundance of Enterobacteriaceae [246,247] and E. coli [248], increased production of propionic [245] and acetic acids [246,247], and downregulation of pro-inflammatory cytokines [248]. Humam et al. also observed improved intestinal morphology and microbiota as well as the immune system in broilers under heat stress [245]. The administration of postbiotics obtained from P. acidilactici, L. reuteri, E. faecium and L. acidophilus has also recently been reported to improve the lesion scores, C. perfringens counts and mortality rates as well as to stimulate the innate immune response and to reduce the pro-inflammatory responses in favour of a homeostatic-like response in a broiler C. perfringens challenge model [249].

Pigs

Unlike poultry, data regarding the potential utilization of postbiotics as alternative to antibiotics in the other monogastric animals are very limited. The administration of postbiotics obtained from L. plantarum has been reported to improve the gut health of postweaning piglets in terms of reduced incidence of diarrhoea, increased abundance of lactic acid bacteria, decreased abundance of Enterobacteriaceae, increased production of SCFAs, reduced pH and enhanced villus morphology [250].

Fish

A similar scenario can also be highlighted in fish, where the utilization of postbiotics derived from Lactobacillus [251,252], Leuconostoc [251], Rhodotorula and Cetobacterium [253] has led to promising results in terms of gut health. In particular, Wu et al. recently observed a positive modulation of the gut microbiota in a hybrid sturgeon fed postbiotic-based diets [253]. The administration of postbiotics has also been recently reported to improve the intestinal microbiota as well as to increase protection against Lactococcus garvieae infection in rainbow trout [251,252].

Pets

Nowadays, in companion animal medicine, there is a lack of research focused on postbiotics functional metabolism. However, it is commonly recognized that postbiotics are mainly produced by saccharolytic and putrefactive proteolytic fermentations [254]. Saccharolytic fermentation has firstly been reported to produce acetate (60%), propionate (25%)—which served as energy sources and modulated the neuroendocrine system—and butyrate (15%)—that acted as the ideal substrate for colonocyte development, and Na+ and Cl reabsorption [255,256]. Recent studies have also evaluated the functionality of postbiotic molecules produced by putrefactive protein metabolism. In particular, polyamines such as spermidine, spermine, putrescine and cadaverine from arginine and lysine catabolism are capable of slowing down the intestinal epithelial cells senescence and of maintaining the gastrointestinal barrier function through occluding production and cadherin expression [216]. The BCFAs and isobutyrate production starting from valine integration has also been demonstrated to be a compelling energy substrate for the colonocytes in an even more efficient and rapid pathway compared to the butyrate [216]. As Wernimont et al. recently suggested, postbiotics can exert positive effects on numerous diseases such as inflammatory enteropathies, kidney diseases and oral diseases [254]. In conclusion, some of the most recent studies involving probiotic compounds and their impact on intestinal health are reported (Table 3).

6. Conclusions

The increasing use and misuse of antibiotics in both human and veterinary medicine has progressively led to the spread of antibiotic resistance issues as well as the frequent onset of intestinal dysbiosis. Nutritional alternatives to the use of antibiotics herein critically summarized and discussed (such as probiotics, prebiotics and postbiotics) show a huge potential in the reduction of the use of antibiotics, since they can prevent or be useful as adjuvants in the treatment of gastrointestinal diseases (where the antibiotics are still considered the primary therapeutic choice) or in the resolution of antibiotic-induced dysbiosis.

Funding

This research received no external funding.

Acknowledgments

This research was carried out in the framework of the project “Dipartimenti di Eccellenza 2018–2022” (Ministry of University and Research). The authors are also thankful to Candioli s.r.l., (Italy) for the provided support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dodds, D.R. Antibiotic resistance: A current epilogue. Biochem. Pharmacol. 2017, 134, 139–146. [Google Scholar] [CrossRef]
  2. Breithaupt, H. Tackling resistance: Bacteria, humans, animals and the environment: An Interview with Stuart Levy, Professor at Tufts University School of Medicine and President of the Alliance for the Prudent Use of Antibiotics. EMBO Rep. 2014, 15, 127–130. [Google Scholar] [CrossRef]
  3. Ramos, S.; Chafsey, I.; Silva, N.; Hébraud, M.; Santos, H.; Capelo-Martinez, J.L.; Poeta, P.; Igrejas, G. Effect of vancomycin on the proteome of the multiresistant Enterococcus faecium SU18 strain. J. Proteom. 2015, 113, 378–387. [Google Scholar] [CrossRef]
  4. Ledingham, K.; Hinchliffe, S.; Jackson, M.; Thomas, F.; Tomson, G. Antibiotic Resistance: Using a Cultural Context of Health Approach to Address a Global Health Challenge; World Health Organization: Copenhagen, Denmark, 2019; ISBN 978-92-890-5394-5. [Google Scholar]
  5. Magiorakos, A.P.; Srinivasen, A.; Carey, R.B.; Carmeli, Y.; Falagas, M.E.; Giske, C.G.; Harbarth, S.; Hindler, J.F.; Kahlmeter, G.; Olsson-LiJequist, B.; et al. Multidrug-resistant, extensively drug-resistant and pandgrug-resistant bacteria: An international expert proposal for interim standard definitions for acquired resistance. Clin. Microbiol. Infect. 2012, 18, 268–281. [Google Scholar] [CrossRef] [Green Version]
  6. Wolfensberger, A.; Kuster, S.P.; Marchesi, M.; Zbinden, R.; Hombach, M. The effect of varying multidrug-resistence (MDR) definitions on rates of MDR gram-negative rods. Antimicrob. Resist. Infect. Control. 2019, 28, 193. [Google Scholar] [CrossRef] [Green Version]
  7. Tacconelli, E.; Cataldo, M.A.; Dancer, S.J.; De Angelis, G.; Falcone, M.; Frank, U.; Kahlmeter, G.; Pan, A.; Petrosillo, N.; Rodríguez-Baño, J.; et al. ESCMID guidelines for the management of the infection control measures to reduce transmission of multidrug resistant Gram negative bacteria in hospitalized patients. Clin. Microbiol. Infect. 2014, 20, 1–55. [Google Scholar] [CrossRef] [Green Version]
  8. Ley, R.E.; Backhed, F.; Turnbaugh, P.; Lozupone, C.A.; Knight, R.D.; Gordon, J.I. Obesity alters gut microbial ecology. Proc. Natl. Acad. Sci. USA 2005, 102, 11070–11075. [Google Scholar] [CrossRef] [Green Version]
  9. Marteau, P.; Lepage, P.; Mangin, I.; Suau, A.; Dore, J.; Pochart, P.; Seksik, P. Review article: Gut flora and inflammatory bowel disease. Aliment. Pharmacol. Ther. 2004, 20, 18–23. [Google Scholar] [CrossRef]
  10. Swidsinski, A.; Weber, J.; Loening-Baucke, V.; Hale, L.P.; Lochs, H. Spatial organization and composition of the mucosal flora in patients with inflammatory bowel disease. J. Clin. Microbiol. 2005, 43, 3380–3389. [Google Scholar] [CrossRef] [Green Version]
  11. Eckburg, P.B.; Bik, E.M.; Bernstein, C.N.; Purdom, E.; Dethlefsen, L.; Sargent, M.; Gill, S.R.; Nelson, K.E.; Relman, D.A. Diversity of the human intestinal microbial flora. Science 2005, 308, 1635–1638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Finegold, S.M.; Sutter, V.L.; Mathisen, G.E. Normal Indigenous Intestinal Flora; Academic Press: Cambridge, MA, USA, 2012; Volume 1, pp. 3–31. [Google Scholar]
  13. Frank, D.N.; Amand, A.L.S.; Feldman, R.A.; Boedeker, E.C.; Harpaz, N.; Pace, N.R. Molecular-phylogenetic characterization of microbial community imbalances in human inflammatory bowel diseases. Proc. Natl. Acad. Sci. USA 2007, 104, 13780–13785. [Google Scholar] [CrossRef] [Green Version]
  14. Savage, D.C. Microbial ecology of the gastrointestinal tract. Ann. Rev. Microbiol. 1977, 31, 107–133. [Google Scholar] [CrossRef]
  15. Suau, A.; Bonnet, R.; Sutren, M.; Godon, J.J.; Gibson, G.R.; Collins, M.D.; Dore, J. Direct analysis of genes encoding 16S rRNA from complex communities reveals many novel molecular species within the human gut. Appl. Environ. Microbiol. 1999, 65, 4799–4807. [Google Scholar] [CrossRef] [Green Version]
  16. Wilson, K.H.; Ikeda, J.S.; Blitchington, R.B. Phylogenetic placement of community members of human colonic biota. Clin. Infect. Dis. 1997, 25, S114–S116. [Google Scholar] [CrossRef] [Green Version]
  17. Gill, S.R.; Pop, M.; DeBoy, R.T.; Eckburg, P.B.; Turnbaugh, P.J.; Samuel, B.S.; Gordon, J.I.; Relman, D.A.; Fraser-Liggett, C.M.; Nelson, K.E. Metagenomic analysis of the human distal gut microbiome. Science 2006, 312, 1355–1359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Manichanh, C.; Rigottier-Gois, L.; Bonnaud, E.; Gloux, K.; Pelletier, E.; Frangeul, L.; Nalin, R.; Jarrin, C.; Chardon, P.; Marteau, P.; et al. Reduced diversity of fecal microbiota in Crohn’s disease revealed by a metagenomic approach. Gut 2006, 55, 205–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Huttenhower, C.; Gevers, D.; Knight, R.; Abubucker, S.; Badger, J.H.; Chinwalla, A.T.; Creasy, H.H.; Earl, A.M.; FitzGerald, M.G.; Fulton, R.S.; et al. Human Microbiome Project Consortium. Structure, function and diversity of the healthy human microbiome. Nature 2012, 486, 207–214. [Google Scholar] [CrossRef] [Green Version]
  20. Tap, J.; Mondot, S.; Levenez, F.; Pelletier, E.; Caron, C.; Furet, J.P.; Ugarte, E.; Muñoz-Tamayo, R.; Paslier, D.L.E.; Nanlin, R.; et al. Towards the human intestinal microbiota phylogenetic core. Environ. Microbiol. 2009, 11, 2574–2584. [Google Scholar] [CrossRef] [PubMed]
  21. Ismail, N.A.; Ragab, S.H.; ElBaky, A.A.; Shoeib, A.R.S.; Alhosary, Y.; Fekry, D. Frequency of Firmicutes and Bacteroidetes in gut microbiota in obese and normal weight Egyptian children and adults. Arch. Med. Sci. 2011, 7, 501–507. [Google Scholar] [CrossRef] [PubMed]
  22. Kundu, P.; Blacher, E.; Elinav, E.; Pettersson, S. Our Gut Microbiome: The Evolving Inner Self. Cell 2017, 171, 1481–1493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ley, R.E.; Hamady, M.; Lozupone, C.; Turnbaugh, P.J.; Ramey, R.R.; Bircher, J.S.; Schlegel, M.L.; Tucker, T.A.; Schrenzel, M.D.; Knight, R.; et al. Evolution of mammals and their gut microbes. Science 2008, 320, 1647–1651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Dominguez-Bello, M.G.; Costello, E.K.; Contreras, M.; Magris, M.; Hidalgo, G.; Fierer, N.; Knight, R. Delivery mode shapes the acquisition and structure of the initial microbiota across multiple body habitats in newborns. Proc. Natl. Acad. Sci. USA 2010, 107, 11971–11975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Penders, J.; Thijs, C.; Vink, C.; Stelma, F.F.; Snijders, B.; Kummeling, I.; van den Brandt, P.A.; Stobberingh, E.E. Factors influencing the composition of the intestinal microbiota in early infancy. Pediatrics 2006, 118, 511–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Cheng, J.; Ringel-Kulka, T.; Heikamp-de Jong, I.; Ringel, Y.; Carroll, I.; de Vos, W.M.; Salojarvi, J.; Satokari, R. Discordant temporal development of bacterial phyla and the emergence of core in the fecal microbiota of young children. ISME J. 2016, 10, 1002–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Agans, R.; Rigsbee, L.; Kenche, H.; Michail, S.; Khamis, H.J.; Paliy, O. Distal gut microbiota of adolescent children is different from that of adults. FEMS Microbiol. Ecol. 2011, 77, 404–412. [Google Scholar] [CrossRef] [PubMed]
  28. Hollister, E.B.; Riehle, K.; Luna, R.A.; Weidler, E.M.; Rubio-Gonzales, M.; Mistretta, T.A.; Raza, S.; Doddapaneni, H.V.; Metcalf, G.A.; Muzny, D.M.; et al. Structure and function of the healthy pre-adolescent pediatric gut microbiome. Microbiome 2015, 3, 36. [Google Scholar] [CrossRef] [Green Version]
  29. Yatsunenko, T.; Rey, F.E.; Manary, M.J.; Trehan, I.; Dominguez-Bello, M.G.; Contreras, M.; Magris, M.; Hidalgo, G.; Baldassano, R.N.; Anokhin, A.P.; et al. Human gut microbiome viewed across age and geography. Nature 2012, 486, 222–227. [Google Scholar] [CrossRef]
  30. O’Toole, P.W.; Jeffery, I.B. Gut microbiota and aging. Science 2015, 350, 1214–1215. [Google Scholar] [CrossRef]
  31. Biagi, E.; Franceschi, C.; Rampelli, S.; Severgnini, M.; Ostan, R.; Turroni, S.; Consolandi, C.; Quercia, S.; Scurti, M.; Monti, D.; et al. Gut Microbiota and Extreme Longevity. Curr. Biol. 2006, 26, 1480–1485. [Google Scholar] [CrossRef] [Green Version]
  32. Freter, R. Factors affecting the microecology of the gut. In Probiotics. The Scientific Basis; Fuller, R., Ed.; Chapman & Hall: London, UK, 1992; pp. 111–144. ISBN 978-94-010-5043-2. [Google Scholar]
  33. Hanning, I.; Diaz-Sanchez, S. The functionality of the gastrointestinal microbiome in non-human animals. Microbiome 2015, 3, 51. [Google Scholar] [CrossRef] [Green Version]
  34. Richards, J.D.; Gong, J.; De Lange, C.F.M. The gastrointestinal microbiota and its role in monogastric nutrition and health with an emphasis on pigs: Current understanding, possible modulations, and new technologies for ecological studies. Can. J. Anim. Sci. 2005, 85, 421–435. [Google Scholar] [CrossRef]
  35. Kogut, M.H.; Arsenault, R.J. Editorial: Gut Health: The New Paradigm in Food Animal Production. Front. Vet. Sci. 2016, 3, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Celi, P.; Cowieson, A.J.; Fru-Nji, F.; Steinert, R.E.; Kluenter, A.M.; Verlhac, V. Gastrointestinal functionality in animal nutrition and health: New opportunities for sustainable animal production. Anim. Feed Sci. Technol. 2017, 234, 88–100. [Google Scholar] [CrossRef]
  37. Pan, D.; Yu, Z. Intestinal microbiome of poultry and its interaction with host and diet. Gut Microbes 2014, 5, 108–119. [Google Scholar] [CrossRef] [PubMed]
  38. Wise, M.G.; Siragusa, G.R. Quantitative analysis of the intestinal bacterial community in one- to three-week-old commercially reared broiler chickens fed conventional or antibiotic-free vegetable-based diets. J. Appl. Microbiol. 2007, 102, 1138–1149. [Google Scholar] [CrossRef] [PubMed]
  39. Biasato, I.; Ferrocino, I.; Colombino, E.; Gai, F.; Schiavone, A.; Cocolin, L.; Vincenti, V.; Capucchio, M.T.; Gasco, L. Effects of dietary Hermetia illucens meal inclusion on cecal microbiota and small intestinal mucin dynamics and infiltration with immune cells of weaned piglets. J. Anim. Sci. Biotechnol. 2020, 11, 1–11. [Google Scholar] [CrossRef]
  40. Velasco-Galilea, M.; Piles, M.; Viñas, M.; Rafel, O.; González-Rodríguez, O.; Guivernau, M.; Sánchez, J.P. Rabbit Microbiota Changes Throughout the Intestinal Tract. Front. Microbiol. 2018, 9, 1–14. [Google Scholar] [CrossRef]
  41. Eichmiller, J.J.; Hamilton, M.J.; Staley, C.; Sadowsky, M.J.; Sorensen, P.W. Environment shapes the fecal microbiome of invasive carp species. Microbiome 2016, 4, 44. [Google Scholar] [CrossRef] [Green Version]
  42. Gajardo, K.; Rodiles, A.; Kortner, T.M.; Krogdahl, Å.; Bakke, A.M.; Merrifield, D.L.; Sørum, H. A high-resolution map of the gut microbiota in Atlantic salmon (Salmo salar): A basis for comparative gut microbial research. Sci. Rep. 2016, 6, 30893. [Google Scholar] [CrossRef] [Green Version]
  43. Lyons, P.; Turnbull, J.; Dawsonl, K.; Crumlish, M. Exploring the microbial diversity of the distal intestinal lumen and mucosa of farmed rainbow trout Oncorhynchus mykiss (Walbaum) using next generation sequencing (NGS). Aquac. Res. 2017, 48, 77–91. [Google Scholar] [CrossRef]
  44. Loch, T.P.; Faisal, M. Emerging flavobacterial infections in fish: A review. J. Adv. Res. 2015, 6, 283–300. [Google Scholar] [CrossRef] [PubMed]
  45. Sullam, K.E.; Essinger, S.D.; Lozupone, C.A.; O’Connor, M.P.; Rosen, G.; Knight, R.; Kilham, S.S.; Russell, J.A. Environmental and ecological factors that shape the gut bacterial communities of fish: A meta-analysis. Mol. Ecol. 2012, 21, 3363–3378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Tarnecki, A.M.; Burgos, F.A.; Ray, C.L.; Arias, C.R. Fish intestinal microbiome: Diversity and symbiosis unravelled by metagenomics. J. Appl. Microbiol. 2017, 123, 2–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Jha, A.R.; Shmalberg, J.; Tanprasertsuk, J.; Perry, L.; Massey, D.; Honaker, R.W. Characterization of gut microbiomes of household pets in the United States using a direct-to-consumer approach. PLoS ONE 2020, 15, e0227289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Alessandri, G.; Milani, C.; Mancabelli, L.; Longhi, G.; Anzalone, R.; Lugli, G.A.; Duranti, S.; Turroni, F.; Ossiprandi, M.C.; van Sinderen, D.; et al. Deciphering the bifidobateria populations within the canine and feline gut microbiota. Appl. Environ. Microbiol. 2020, 86, e02875-19. [Google Scholar] [CrossRef] [PubMed]
  49. Murray, M.T.; Pizzorno, J. The Encyclopedia of Natural Medicine, 4th ed.; Churchill Livingstone: London, UK, 2012; ISBN 978-14-377-2333-5. [Google Scholar]
  50. Holzapfel, W.H.; Haberer, P.; Snel, J.; Schillinger, U.; in’t Veld, J.H.H. Overview of gut flora and probiotics. Int. J. Food Microbiol. 1998, 41, 85–101. [Google Scholar] [CrossRef]
  51. Gismondo, M. Antibiotic impact on intestinal microflora. Gastroenterol. Int. 1998, 11, 29–30. [Google Scholar]
  52. Becattini, S.; Taur, Y.; Pamer, E.G. Antibiotic-Induced Changes in the Intestinal Microbiota and Disease. Trends Mol. Med. 2016, 22, 458–478. [Google Scholar] [CrossRef] [Green Version]
  53. Nord, C.E. Studies on the ecological impact of antibiotics. Eur. J. Clin. Microbiol. Infect. Dis 1990, 9, 517–518. [Google Scholar] [CrossRef]
  54. Nord, C.E.; Heimdahl, A.; Kager, L. Antimicrobial agents and the human oropharyngeal and intestinal microflora. Ann. Inst. Super. Sanita 1986, 22, 883–892. [Google Scholar]
  55. Shin, N.R.; Whon, T.W.; Bae, J.W. Proteobacteria: Microbial signature of dysbiosis in gut microbiota. Trends Biotechnol. 2015, 33, 496–503. [Google Scholar] [CrossRef] [PubMed]
  56. Vrieze, A.; Out, C.; Fuentes, S.; Jonker, L.; Reuling, I.; Kootte, R.S.; van Nood, E.; Holleman, F.; Knaapen, M.; Romijn, J.A.; et al. Impact of oral vancomycin on gut microbiota, bile acid metabolism, and insulin sensitivity. J. Hepatol. 2014, 60, 824–831. [Google Scholar] [CrossRef] [PubMed]
  57. Peralta-Sánchez, J.M.; Martín-Platero, A.M.; Ariza-Romero, J.J.; Rabelo-Ruiz, M.; Zurita-González, M.J.; Baños, A.; Rodríguez-Ruano, S.M.; Maqueda, M.; Valdivia, E.; Martínez-Bueno, M. Egg Production in Poultry Farming Is Improved by Probiotic Bacteria. Front. Microbiol. 2019, 10, 1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Connelly, S.; Bristol, J.A.; Hubert, S.; Subramanian, P.; Hasan, N.A.; Colwell, R.R.; Kaleko, M. SYN-004 (ribaxamase), an oral beta-lactamase, mitigates antibiotic-mediated dysbiosis in a porcine gut microbiome model. J. Appl. Microbiol. 2017, 123, 66–79. [Google Scholar] [CrossRef] [Green Version]
  59. Zhang, C.; Peng, Y.; Mu, C.; Zhu, W. Ileum terminal antibiotic infusion affects jejunal and colonic specific microbial population and immune status in growing pigs. J. Anim. Sci. Biotechnol. 2018, 9, 51. [Google Scholar] [CrossRef]
  60. Gresse, R.; Chaucheyras-Durand, F.; Fleury, M.A.; Van de Wiele, T.; Forano, E.; Blanquet-Diot, S. Gut Microbiota Dysbiosis in Postweaning Piglets: Understanding the Keys to Health. Trends Microbiol. 2017, 10, 851–873. [Google Scholar] [CrossRef]
  61. Simon, K.; Verwoolde, M.B.; Zhang, J.; Smidt, H.; de Vries Reilingh, G.; Kemp, B.; Lammers, A. Long-term effects of early life microbiota disturbance on adaptive immunity in laying hens. Poult. Sci. 2016, 95, 1543–1554. [Google Scholar] [CrossRef]
  62. Yu, M.; Li, Z.; Chen, W.; Wang, G.; Cui, Y.; Ma, X. Dietary Supplementation With Citrus Extract Altered the Intestinal Microbiota and Microbial Metabolite Profiles and Enhanced the Mucosal Immune Homeostasis in Yellow-Feathered Broilers. Front. Microbiol. 2019, 10, 2662. [Google Scholar] [CrossRef] [Green Version]
  63. Kumar, S.; Chen, C.; Indugu, N.; Werlang, G.O.; Singh, M.; Kim, W.K.; Thippareddi, H. Effect of antibiotic withdrawal in feed on chicken gut microbial dynamics, immunity, growth performance and prevalence of foodborne pathogens. PLoS ONE 2018, 13, e0192450. [Google Scholar] [CrossRef] [Green Version]
  64. Le Roy, C.I.; Woodward, M.J.; Ellis, R.J.; La Ragione, R.M.; Claus, S.P. Antibiotic treatment triggers gut dysbiosis and modulates metabolism in a chicken model of gastro-intestinal infection. BMC Vet. Res. 2019, 15, 37. [Google Scholar] [CrossRef]
  65. Connelly, S.; Fanelli, B.; Hasan, N.A.; Colwell, R.R.; Kaleko, M. Oral Metallo-Beta-Lactamase Protects the Gut Microbiome from Carbapenem-Mediated Damage and Reduces Propagation of Antibiotic Resistance in Pigs. Front. Microbiol. 2019, 10, 101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Limbu, S.M.; Zhou, L.; Sun, S.X.; Zhang, M.L.; Du, Z.Y. Chronic exposure to low environmental concentrations and legal aquaculture doses of antibiotics cause systemic adverse effects in Nile tilapia and provoke differential human health risk. Environ. Int. 2018, 115, 205–219. [Google Scholar] [CrossRef]
  67. Navarrete, P.; Mardones, P.; Opazo, R.; Espejo, R.; Romero, J. Oxytetracycline treatment reduces bacterial diversity of intestinal microbiota of Atlantic salmon. J. Aquat. Anim. Health 2008, 20, 177–183. [Google Scholar] [CrossRef] [PubMed]
  68. He, S.; Zhou, Z.; Liu, Y.; Cao, Y.; Meng, K.; Shi, P.; Yao, B.; Ringø, E. Do dietary betaine and the antibiotic florfenicol influence the intestinal autochthonous bacterial community in hybrid tilapia (Oreochromis niloticus ♀ × O. aureus ♂)? World J. Microbiol. Biotechnol. 2012, 28, 785–791. [Google Scholar] [CrossRef] [PubMed]
  69. Limbu, S.M.; Ma, Q.; Zhang, M.L.; Du, Z.Y. High fat diet worsens the adverse effects of antibiotic on intestinal health in juvenile Nile tilapia (Oreochromis niloticus). Sci. Total Environ. 2019, 680, 169–180. [Google Scholar] [CrossRef] [PubMed]
  70. Zhou, L.; Limbu, S.M.; Qiao, F.; Du, Z.Y.; Zhang, M. Influence of Long-Term Feeding Antibiotics on the Gut Health of Zebrafish. Zebrafish 2018, 15, 340–348. [Google Scholar] [CrossRef]
  71. Vàsquez-Baeza, Y.; Hyde, E.R.; Suchodolski, J.S.; Knight, R. Dog and human inflammatory bowel disease rely on overlapping yet distinct dysbiosis networks. Comparat. Study 2016, 3, 16177. [Google Scholar] [CrossRef]
  72. Minamoto, Y.; Minamoto, T.; Isaiah, A.; Sattasathuchana, P.; Buono, A.; Rangachari, V.R.; McNeely, I.H.; Lidbury, J.; Steiner, J.M.; Suchodolski, J.S. Fecal short-chain fatty acid concentrations and dysbiosis in dogs with chronic enteropathy. J. Vet. Intern. Med. 2019, 33, 1608–1618. [Google Scholar] [CrossRef] [Green Version]
  73. German, A.J.; Hall, E.J.; Day, M.J. Chronic intestinal inflammation and intestinal disease in dogs. J. Vet. Intern. Med. 2003, 17, 8–20. [Google Scholar] [CrossRef]
  74. Honneffer, J.B.; Minamoto, Y.; Suchodolski, J.S. Microbiota alterations in acute and chronic gastrointestinal inflammation of cats and dogs. World J. Gastroenterol. 2014, 20, 16489–16497. [Google Scholar] [CrossRef]
  75. Guard, B.C.; Barr, J.W.; Reddivari, L.; Klemashevich, C.; Jayaraman, A.; Steiner, J.M.; Vanamala, J.; Suchodolski, J.S. Characterization of microbial dysbiosis and metabolomic changes in dogs with acute diarrhea. PLoS ONE 2015, 10, e0127259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Šlapeta, J.; Dowd, S.E.; Alanazi, A.D.; Westman, M.E.; Brown, G.K. Differences in the faecal microbiome of non-diarrhoeic clinically healthy dogs and cats associated with Giardia duodenalis infection: Impact of hookworms and coccidia. Int. J. Parasitol. 2015, 45, 585–594. [Google Scholar] [CrossRef] [PubMed]
  77. Suchodolski, J.S. Diagnosis and interpretation of intestinal dysbiosis in dogs and cats. Vet. J. 2016, 215, 30–37. [Google Scholar] [CrossRef] [PubMed]
  78. Xenoulis, P.G.; Palculict, B.; Allenspach, K.; Steiner, J.M.; Van House, A.M.; Suchodolski, J.S. Molecular-phylogenetic characterization of microbial communities imbalances in the small intestine of dogs with inflammatory bowel disease. FEMS Microbiol. Ecol. 2008, 66, 579–589. [Google Scholar] [CrossRef] [Green Version]
  79. Suchodolski, J.S.; Markel, M.E.; Garcia-Mazcorro, J.F.; Unterer, S.; Heilmann, R.M.; Dowd, S.E.; Kachroo, P.; Ivanov, I.; Minamoto, Y.; Dillman, E.M.; et al. The fecal microbiome in dogs with acute diarrhea and idiopathic inflammatory bowel disease. PLoS ONE 2012, 7, e51907. [Google Scholar] [CrossRef] [Green Version]
  80. Janeczko, S.; Atwater, D.; Bogel, E.; Greiter-Wilke, A.; Gerold, A.; Baumgart, M.; Bender, H.; McDonough, P.L.; McDonough, S.P.; Goldstein, R.E.; et al. The relationship of mucosal bacteria to duodenal histopathology, cytokine mRNA, and clinical disease activity in cats with inflammatory bowel disease. Vet. Microbiol. 2008, 128, 178–193. [Google Scholar] [CrossRef]
  81. Inness, V.L.; McCartney, A.L.; Khoo, C.; Gross, K.L.; Gibson, G.R. Molecular characterisation of the gut microflora of healthy and inflammatory bowel disease cats using fluorescence in situ hybridisation with special reference to Desulfovibrio spp. J. Anim. Physiol. Anim. Nutr. (Berl.) 2007, 91, 48–53. [Google Scholar] [CrossRef]
  82. Abecia, L.; Hoyles, L.; Khoo, C.; Frantz, N.; McCartney, A.L. Effects of a novel galactooligosaccharide on the faecal microbiota of healthy and inflammatory bowel disease cats during a randomized, double-blind, cross-over feeding study. Int. J. Probiot. Prebiot. 2010, 5, 61–68. [Google Scholar]
  83. Suchodolski, J.S.; Foster, M.L.; Sohail, M.U.; Leutenegger, C.; Queen, E.V.; Steiner, J.M.; Marks, S.L. The fecal microbiome in cats with diarrhea. PLoS ONE 2015, 10, e0127378. [Google Scholar] [CrossRef] [Green Version]
  84. Makielski, K.; Cullen, J.; O’Connor, A.; Jergens, A.E. Narrative review of therapies for chronic enteropathies in dogs and cats. J. Vet. Intern. Med. 2019, 33, 11–22. [Google Scholar] [CrossRef] [Green Version]
  85. Kilpinen, S.; Spillmann, T.; Westermarck, E. Efficacy of two low-dose oral tylosin regimens in controlling the relapse of diarrhea in dogs with tylosin-responsive diarrhea: A prospective, single-blinded, two-arm parallel, clinical field trial. Acta Vet. Scand. 2014, 56, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Craven, M.; Mansfield, C.S.; Simpson, K.W. Granulomatous colitis of boxer dogs. Vet. Clin. N. Am. Small Anim. Pract. 2011, 41, 433–445. [Google Scholar] [CrossRef] [PubMed]
  87. Candellone, A.; Cerquetella, M.; Girolami, F.; Badino, P.; Odore, R. Acute Diarrhea in Dogs: Current Management and Potential Role of Dietary Polyphenols Supplementation. Antioxidants 2020, 9, 725. [Google Scholar] [CrossRef] [PubMed]
  88. Suchodolski, J.S.; Dowd, S.E.; Westermarck, E.; Steiner, J.M.; Wolcott, R.D.; Spillmann, T.; Harmoinen, J.A. The effect of the macrolide antibiotic tylosin on microbial diversity in the canine small intestine as demonstrated by massive parallel 16S rRNA gene sequencing. BMC Microbiol. 2009, 9, 210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Manchester, A.C.; Webb, C.B.; Blake, A.B.; Sarwar, F.; Lidbury, J.A.; Steiner, J.M.; Suchodolski, J.S. Long-term impact of tylosin on fecal microbiota and fecal bile acids of healthy dogs. J. Vet. Intern. Med. 2019, 33, 2605–2617. [Google Scholar] [CrossRef]
  90. Igarashi, H.; Maeda, S.; Ohno, K.; Horigome, A.; Odamaki, T.; Tsujimoto, H. Effect of oral administration of metronidazole or prednisolone on fecal microbiota in dogs. PLoS ONE 2014, 9, e107909. [Google Scholar] [CrossRef]
  91. Grønvold, A.M.; L’abée-Lund, T.M.; Sørum, H.; Skancke, E.; Yannarell, A.C.; Mackie, R.I. Changes in fecal microbiota of healthy dogs administered amoxicillin. FEMS Microbiol. Ecol. 2010, 71, 313–326. [Google Scholar] [CrossRef]
  92. Kilpinen, S.; Rantala, M.; Spillmann, T.; Björkroth, J.; Westermarck, E. Oral tylosin administration is associated with an increase of faecal enterococci and lactic acid bacteria in dogs with tylosin-responsive diarrhoea. Vet. J. 2015, 205, 369–374. [Google Scholar] [CrossRef]
  93. Wintersdorff, V.C.J.; Penders, J.; van Niekerk, J.M.; Mills, N.D.; Majumder, S.; van Alphen, L.B.; Savelkoul, P.H.; Wolffs, P.F. Dissemination of Antimicrobial Resistance in Microbial Ecosystems through Horizontal Gene Transfer. Front. Microbiol. 2016, 7, 173. [Google Scholar] [CrossRef] [Green Version]
  94. Hill, C.; Guarner, F.; Reid, G.; Gibson, G.R.; Merenstein, D.J.; Pot, B.; Morelli, L.; Canani, R.B.; Flint, H.J.; Salminen, S.; et al. Expert consensus document. The International Scientific Association for Probiotics and Prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat. Rev. Gastroenterol. Hepatol. 2014, 11, 506–514. [Google Scholar] [CrossRef] [Green Version]
  95. Floch, M.H.; Walker, W.A.; Madsen, K.; Sanders, M.E.; Macfarlane, G.T.; Flint, H.J.; Dieleman, L.A.; Ringel, Y.; Guandalini, S.; Kelly, C.P.; et al. Recommendations for probiotic use-2011 update. J. Clin. Gastroenterol. 2011, 45, 68–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ouwehand, A.C.; Kirjavainen, P.V.; Shortt, C.; Salminen, S. Probiotics: Mechanisms and established effects. Int. Dairy J. 1999, 9, 43–52. [Google Scholar] [CrossRef]
  97. Vanderhoof, J.A.; Whitney, D.B.; Antonson, D.L.; Hanner, T.L.; Lupo, J.V.; Young, R.J. Lactobacillus GG in the prevention of antibiotic-associated diarrhea in children. J. Pediatr. 1999, 135, 564–568. [Google Scholar] [CrossRef]
  98. Conway, S.; Hart, A.; Clark, A.; Harvey, I. Does eating yogurt prevent antibiotic-associated diarrhoea? A placebo-controlled randomised controlled trial in general practice. Br. J. Gen. Pract. 2007, 57, 953–959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Blaabjerg, S.; Artzi, D.M.; Aabenhus, R. Probiotics for the Prevention of Antibiotic-Associated Diarrhea in Outpatients-A Systematic Review and Meta-Analysis. Antibiotics 2017, 6, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Triantafillidis, J.K.; Nasioulas, G.; Kosmidis, P.A. Colorectal Cancer and Inflammatory Bowel Disease: Epidemiology, Risk Factors, Mechanisms of Carcinogenesis and Prevention Strategies. Anticancer Res. 2009, 29, 2727–2737. [Google Scholar] [PubMed]
  101. Hilty, M.; Burke, C.; Pedro, H.; Cardenas, P.; Bush, A.; Bossley, C.; Davies, J.; Ervine, A.; Poulter, L.; Pachter, L.; et al. Disordered Microbial Communities in Asthmatic Airways. PLoS ONE 2010, 5, e8578. [Google Scholar] [CrossRef] [Green Version]
  102. Bjarnason, I.; Sission, G.; Hayee, B. A randomised, double-blind, placebo-controlled trial of a multi-strain probiotic in patients with asymptomatic ulcerative colitis and Crohn’s disease. Inflammopharmacology 2019, 27, 465–473. [Google Scholar] [CrossRef] [Green Version]
  103. Furrie, E.; Macfarlane, S.; Kennedy, A.; Cummings, J.H.; Walsh, S.V.; O’Neil, D.A.; Macfarlane, G.T. Synbiotic therapy (Bifidobacterium longum/Synergy 1) initiates resolution of inflammation in patients with active ulcerative colitis: A randomised controlled pilot trial. Gut 2005, 54, 242–249. [Google Scholar] [CrossRef]
  104. Yang, J.; Yu, J. The association of diet, gut microbiota and colorectal cancer: What we eat may imply what we get. Protein Cell 2018, 9, 474–487. [Google Scholar] [CrossRef] [Green Version]
  105. Shadnoush, M.; Hosseini, R.S.; Khalilnezhad, A.; Navai, L.; Goudarzi, H.; Vaezjalali, M. Effects of Probiotics on Gut Microbiota in Patients with Inflammatory Bowel Disease: A Double-blind, Placebo-controlled Clinical Trial. Korean J. Gastroenterol. 2015, 65, 215–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Krebs, B. Prebiotic and Synbiotic Treatment before Colorectal Surgery—Randomised Double Blind Trial. Coll. Antropol. 2016, 40, 35–40. [Google Scholar] [PubMed]
  107. Marteau, P.; Seksik, P.; Jian, R. Probiotics and intestinal health effects: A clinical perspective. Br. J. Nutr. 2002, 88, S51–S57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Liong, M.T. Roles of probiotics and prebiotics in colon cancer prevention: Postulated mechanisms and in-vivo evidence. Int. J. Mol. Sci. 2008, 9, 854–863. [Google Scholar] [CrossRef] [Green Version]
  109. Narushima, S.; Sugiura, Y.; Oshima, K.; Atarashi, K.; Hattori, M.; Suematsu, M.; Honda, K. Characterization of the 17 strains of regulatory T cell-inducing human-derived Clostridia. Gut Microbes 2014, 5, 333–339. [Google Scholar] [CrossRef] [PubMed]
  110. Cisek, A.A.; Binek, M. Chicken intestinal microbiota function with a special emphasis on the role of probiotic bacteria. Pol. J. Vet. Sci. 2014, 17, 385–394. [Google Scholar] [CrossRef] [Green Version]
  111. Shi, S.; Qi, Z.; Jiang, W.; Quan, S.; Sheng, T.; Tu, J.; Shao, Y.; Qi, K. Effects of probiotics on cecal microbiome profile altered by duck Escherichia coli 17 infection in Cherry Valley ducks. Microb. Pathog. 2020, 138, 103849. [Google Scholar] [CrossRef]
  112. Tarabees, R.; Gafar, K.M.; El-Sayed, M.S.; Shehata, A.A.; Ahmed, M. Effects of Dietary Supplementation of Probiotic Mix and Prebiotic on Growth Performance, Cecal Microbiota Composition, and Protection Against Escherichia coli O78 in Broiler Chickens. Probiot. Antimicrob. Proteins 2019, 11, 981–989. [Google Scholar] [CrossRef]
  113. El-Sharkawy, H.; Tahoun, A.; Rizk, A.M.; Suzuki, T.; Elmonir, W.; Nassef, E.; Shukry, M.; Germoush, M.O.; Farrag, F.; Bin-Jumah, M.; et al. Evaluation of Bifidobacteria and Lactobacillus Probiotics as Alternative Therapy for Salmonella Typhimurium Infection in Broiler Chickens. Animals 2020, 10, 1023. [Google Scholar] [CrossRef]
  114. El-Sawah, A.A.; Aboelhadid, S.M.; El-Nahass, E.N.; Helal, H.E.; Korany, A.M.; El-Ashram, S. Efficacy of probiotic Enterococcus faecium in combination with diclazuril against coccidiosis in experimentally infected broilers. J. Appl. Microbiol. 2020, 129, 1020–1028. [Google Scholar] [CrossRef]
  115. Yasueda, A.; Mizushima, T.; Nezu, R.; Sumi, R.; Tanaka, M.; Nishimura, J.; Kai, Y.; Hirota, M.; Osawa, H.; Nakajima, K.; et al. The effect of Clostridium butyricum MIYAIRI on the prevention of pouchitis and alteration of the microbiota profile in patients with ulcerative colitis. Surg. Today 2016, 46, 939–949. [Google Scholar] [CrossRef]
  116. Zhang, L.; Zhang, L.; Zhan, X.; Zeng, X.; Zhou, L.; Cao, G.; Chen, A.; Yang, C. Effects of dietary supplementation of probiotic, Clostridium butyricum, on growth performance, immune response, intestinal barrier function, and digestive enzyme activity in broiler chickens challenged with Escherichia coli K88. J. Anim. Sci. Biotechnol. 2016, 7, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Zhao, X.; Yang, J.; Wang, L.; Lin, H.; Sun, S. Protection Mechanism of Clostridium butyricum against Salmonella Enteritidis Infection in Broilers. Front. Microbiol. 2017, 8, 1523. [Google Scholar] [CrossRef] [Green Version]
  118. Huang, T.; Peng, X.Y.; Gao, B.; Wei, Q.L.; Xiang, R.; Yuan, M.G.; Xu, Z.H. The Effect of Clostridium butyricum on Gut Microbiota, Immune Response and Intestinal Barrier Function During the Development of Necrotic Enteritis in Chickens. Front. Microbiol. 2019, 10, 2309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Elshaghabee, F.M.F.; Rokana, N.; Gulhane, R.D.; Sharma, C.; Panwar, H. Bacillus as potential probiotics: Status, concerns, and future perspectives. Front. Microbiol. 2017, 8, 1490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Bortoluzzi, C.; Vieira, S.B.; Dorigam, D.P.J.C.; Menconi, A.; Sokale, A.; Doranalli, K.; Applegate, T.J. Bacillus subtilis DSM 32315 Supplementation Attenuates the Effects of Clostridium perfringens Challenge on the Growth Performance and Intestinal Microbiota of Broiler Chickens. Microorganisms 2019, 7, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Whelan, R.A.; Doranalli, K.; Rinttilä, T.; Vienola, K.; Jurgens, G.; Apajalahti, J. The impact of Bacillus subtilis DSM 32315 on the pathology, performance, and intestinal microbiome of broiler chickens in a necrotic enteritis challenge. Poult. Sci. 2019, 98, 3450–3463. [Google Scholar] [CrossRef]
  122. Massacci, F.R.; Lovito, C.; Tofani, S.; Tentellini, M.; Genovese, D.A.; De Leo, A.A.P.; Papa, P.; Magistrali, C.F.; Manuali, E.; Trabalza-Marinucci, M.; et al. Dietary Saccharomyces cerevisiae boulardii CNCM I-1079 Positively Affects Performance and Intestinal Ecosystem in Broilers during a Campylobacter jejuni Infection. Microorganisms 2019, 7, 596. [Google Scholar] [CrossRef] [Green Version]
  123. Li, X.Q.; Zhu, Y.H.; Zhang, H.F.; Yue, Y.; Cai, Z.X.; Lu, Q.P.; Zhang, L.; Weng, X.G.; Zhang, F.J.; Zhou, D.; et al. Risks associated with high-dose Lactobacillus rhamnosus in an Escherichia coli model of piglet diarrhoea: Intestinal microbiota and immune imbalances. PLoS ONE 2012, 7, e40666. [Google Scholar] [CrossRef]
  124. Zhang, W.; Wu, Q.; Zhu, Y.; Yang, G.; Yu, J.; Wang, J.; Ji, H. Probiotic Lactobacillus rhamnosus GG Induces Alterations in Ileal Microbiota With Associated CD3(-)CD19(-)T-bet(+)IFNγ(+/−) Cell Subset Homeostasis in Pigs Challenged With Salmonella enterica Serovar 4,[5],12:i:-. Front. Microbiol. 2019, 10, 977. [Google Scholar] [CrossRef] [Green Version]
  125. Yang, G.Y.; Yu, J.; Su, J.H.; Jiao, L.G.; Liu, X.; Zhu, Y.H. Oral Administration of Lactobacillus rhamnosus GG Ameliorates Salmonella Infantis-Induced Inflammation in a Pig Model via Activation of the IL-22BP/IL-22/STAT3 Pathway. Front. Cell Infect. Microbiol. 2017, 7, 323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Nordeste, R.; Tessema, A.; Sharma, S.; Kovač, Z.; Wang, C.; Morales, R.; Griffiths, M.W. Molecules produced by probiotics prevent enteric colibacillosis in pigs. BMC Vet. Res. 2017, 13, 335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Zhang, W.; Zhu, Y.H.; Zhou, D.; Wu, Q.; Song, D.; Dicksved, J.; Wang, J.F. Oral Administration of a Select Mixture of Bacillus Probiotics Affects the Gut Microbiota and Goblet Cell Function following Escherichia coli Challenge in Newly Weaned Pigs of Genotype MUC4 That Are Supposed To Be Enterotoxigenic, E. coli F4ab/ac Receptor Negative. Appl. Environ. Microbiol. 2017, 83, 1–18. [Google Scholar] [CrossRef] [Green Version]
  128. Luise, D.; Bertocchi, M.; Motta, V.; Salvarani, C.; Bosi, P.; Luppi, A.; Fanelli, F.; Mazzoni, M.; Archetti, I.; Maiorano, G.; et al. Bacillus sp. probiotic supplementation diminish the Escherichia coli F4ac infection in susceptible weaned pigs by influencing the intestinal immune response, intestinal microbiota and blood metabolomics. J. Anim. Sci. Biotechnol. 2019, 10, 74. [Google Scholar] [CrossRef] [PubMed]
  129. Safari, R.; Adel, M.; Lazado, C.C.; Caipang, C.M.; Dadar, M. Host-derived probiotics Enterococcus casseliflavus improves resistance against Streptococcus iniae infection in rainbow trout (Oncorhynchus mykiss) via immunomodulation. Fish Shellfish Immunol. 2016, 52, 198–205. [Google Scholar] [CrossRef] [Green Version]
  130. Tan, H.Y.; Chen, S.W.; Hu, S.Y. Improvements in the growth performance, immunity, disease resistance, and gut microbiota by the probiotic Rummeliibacillus stabekisii in Nile tilapia (Oreochromis niloticus). Fish Shellfish Immunol. 2019, 92, 265–275. [Google Scholar] [CrossRef]
  131. Kuebutornye, F.K.A.; Wang, Z.; Lu, Y.; Abarike, E.D.; Sakyi, M.E.; Li, Y.; Xie, C.X.; Hlordzi, V. Effects of three host-associated Bacillus species on mucosal immunity and gut health of Nile tilapia, Oreochromis niloticus and its resistance against Aeromonas hydrophila infection. Fish Shellfish Immunol. 2020, 97, 83–95. [Google Scholar] [CrossRef]
  132. Ahmadifard, N.; Aminlooi, R.V.; Tukmechi, A.; Agh, N. Evaluation of the Impacts of Long-Term Enriched Artemia with Bacillus subtilis on Growth Performance, Reproduction, Intestinal Microflora, and Resistance to Aeromonas hydrophila of Ornamental Fish Poecilia latipinna. Probiot. Antimicrob. Proteins 2019, 11, 957–965. [Google Scholar] [CrossRef]
  133. Ogawa, M.; Shimizu, K.; Nomoto, K.; Takahashi, M.; Watanuki, M.; Tanaka, R.; Tanaka, T.; Hamabata, T.; Yamasaki, S.; Takeda, Y. Protective effect of Lactobacillus casei strain Shirota on Shiga toxin-producing Escherichia coli O157:H7 infection in infant rabbits. Infect. Immun. 2001, 69, 1101–1108. [Google Scholar] [CrossRef] [Green Version]
  134. Fayyaz, I.; Zahoor, M.A.; Shahid, M.; Rasool, M.H.; Nawaz, Z. Effect of Lactobacillus casei on serum interleukins following enteropathogenic E. coli infection in experimental rabbits. Pak. J. Pharm. Sci. 2018, 31, 2131–2136. [Google Scholar]
  135. Schmitz, S.; Suchodolski, J. Understanding the Canine Intestinal Microbiota and Its Modification by Pro-, Pre- And Synbiotics—What Is the Evidence? Vet. Med. Sci. 2016, 2, 71–94. [Google Scholar] [CrossRef] [PubMed]
  136. Grzeskowiak, L.; Endo, A.; Beasley, S.; Salminen, S. Microbiota and Probiotics in Canine and Feline Welfare. Anaerobe 2015, 34, 14–23. [Google Scholar] [CrossRef] [PubMed]
  137. Hammerum, A.M. Enterococci of animal origin and their significance for public health. Clin. Microbiol. Infect. 2012, 18, 619–625. [Google Scholar] [CrossRef] [PubMed]
  138. Collado, M.C.; Grzeskowiak, L.; Salminen, S. Probiotic strains and their combination inhibit in vitro adhesion of pathogens to pig intestinal mucosa. Curr. Microbiol. 2007, 55, 260–265. [Google Scholar] [CrossRef]
  139. Saarela, M.; Mogensen, G.; Fondén, R.; Matto, J.; Mattila-Sandholm, T. Probiotic bacteria: Safety, functional and technological properties. J. Biotechnol. 2000, 84, 197–215. [Google Scholar] [CrossRef]
  140. Ceapa, C.; Wopereis, H.; Rezaiki, L.; Kleerebezem, M.; Knol, J.; Oozeer, R. Influence of fermented milk products, prebiotics and probiotics on microbiota composition and health. Best Pract. Res. Clin. Gastroenterol. 2013, 27, 139–155. [Google Scholar] [CrossRef]
  141. Baillon, M.L.A.; Marshall-Jones, Z.V.; Butterwick, R.F. Effects of probiotic Lactobacillus acidophilus strain DSM13241 in healthy adult dogs. Am. J. Vet. Res. 2004, 65, 338–343. [Google Scholar] [CrossRef]
  142. Biagi, G.; Cipollini, I.; Pompei, A.; Zaghini, G.; Matteuzzi, D. Effect of a Lactobacillus animalis strain on composition and metabolism of the intestinal microflora in adult dogs. Vet. Microbiol. 2007, 124, 160–165. [Google Scholar] [CrossRef]
  143. Grzeskowiak, L.; Endo, A.; Collado, M.C.; Pelliniemi, L.J.; Beasley, S.; Salminen, S. The effect of growth media and physical treatments on the adhesion properties of canine probiotics. J. Appl. Microbiol. 2013, 115, 539–545. [Google Scholar] [CrossRef]
  144. Beasley, S.S.; Manninen, T.J.K.; Saris, P.E.J. Lactic acid bacteria isolated from canine faeces. J. Appl. Microbiol. 2006, 101, 131–138. [Google Scholar] [CrossRef]
  145. Lin, C.F.; Lin, M.Y.; Lin, C.N.; Chiou, M.T.; Chen, J.W.; Yang, K.C.; Wu, M.C. Potential probiotic of Lactobacillus strains isolated from the intestinal tracts of pigs and feces of dogs with antibacterial activity against multidrug-resistant pathogenic bacteria. Arch. Microbiol. 2020, 202, 1849–1860. [Google Scholar] [CrossRef] [PubMed]
  146. Manninen, T.J.K.; Rinkinen, M.L.; Beasley, S.S.; Saris, P.E.J. Alteration of the canine small intestinal lactic acid bacterium microbiota by feeding of potential probiotics. Appl. Environ. Microbiol. 2006, 72, 6539–6543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Grzeskowiak, L.; Collado, M.C.; Beasley, S.; Salminen, S. Pathogen exclusion properties of canine probiotics are influenced by the growth media and physical treatments simulating industrial processes. J. Appl. Microbiol. 2014, 116, 1308–1314. [Google Scholar] [CrossRef] [PubMed]
  148. Strompfova, V.; Placha, I.; Cobanova, K.; Gancarcıkova, S.; Mudronova, D.; Laukova, A. Experimental addition of Eleutherococcus senticosus and probiotic to the canine diet. Cent. Eur. J. Biol. 2012, 7, 436–447. [Google Scholar] [CrossRef] [Green Version]
  149. Marcináková, M.; Simonová, M.; Strompfová, V.; Lauková, A. Oral application of Enterococcus faecium strain EE3 in healthy dogs. Folia Microbiol. (Praha) 2006, 51, 239–242. [Google Scholar] [CrossRef]
  150. Pascher, M.; Hellweg, P.; Khol-Parisini, A.; Zentek, J. Effects of a probiotic Lactobacillus acidophilus strain on feed tolerance in dogs with non-specific dietary sensitivity. Arch. Anim. Nutr. 2008, 62, 107–116. [Google Scholar] [CrossRef] [PubMed]
  151. Felix, A.P.; Netto, M.V.T.; Murakami, F.Y.; De Brito, C.B.M.; De Oliveira, S.G.; Maiorka, A. Digestibility and fecal characteristics of dogs fed with Bacillus subtilis in diet. Cienc. Rural 2010, 40, 2169–2173. [Google Scholar] [CrossRef]
  152. Herstad, H.K.; Nesheim, B.B.; L’Abee-Lund, T.; Larsen, S.; Skancke, E. Effects of a probiotic intervention in acute canine gastroenteritis—A controlled clinical trial. J. Small Anim. Pract. 2010, 51, 34–38. [Google Scholar] [CrossRef]
  153. Arslan, H.H.; Aksu, D.S.; Terzi, G.; Nisbet, C. Therapeutic effects of probiotic bacteria in parvoviral enteritis in dogs. Rev. Med. Vet. 2012, 163, 55–59. [Google Scholar]
  154. Pilla, R.; Suchodolski, J.S. The role of the canine gut microbiome and metabolome in health and gastrointestinal disease. Front. Vet. Sci. 2020, 6, 498. [Google Scholar] [CrossRef] [Green Version]
  155. Garcia-Mazcorro, J.F.; Lanerie, D.J.; Dowd, S.E.; Paddock, C.G.; Gruetzner, N.; Steiner, J.M.; Ivanek, R.; Suchodolski, J.S. Effect of a multi-species synbiotic formulation on fecal bacterial microbiota of healthy cats and dogs as evaluated by pyrosequencing. FEMS Microbiol. Ecol. 2011, 78, 542–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Allen, S.J.; Martinez, E.G.; Gregorio, G.V.; Dans, L.F. Probiotics for treating acute infectious diarrhoea. Cochrane Database Syst. Rev. 2010, 2010, Cd003048. [Google Scholar] [CrossRef] [PubMed]
  157. Shimizu, K.; Yamada, T.; Ogura, H.; Mohri, T.; Kiguchi, T.; Fujimi, S.; Asahara, T.; Ojima, M.; Ikeda, M.; Shimazu, T. Synbiotics modulate gut microbiota and reduce enteritis and ventilator-associated pneumonia in patients with sepsis: A randomized controlled trial. Crit. Care 2018, 22, 239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Bond, D.M.; Morris, J.M.; Nassar, N. Study protocol: Evaluation of the probiotic Lactobacillus Fermentum CECT5716 for the prevention of mastitis in breastfeeding women: A randomised controlled trial. BMC Pregnancy Childbirth 2017, 17, 148. [Google Scholar] [CrossRef] [Green Version]
  159. Rossi, G.; Pengo, G.; Caldin, M.; Piccionello, A.P.; Steiner, J.M.; Cohen, N.D.; Jergens, A.E.; Suchodolski, J.S. Comparison of Microbiological, Histological, and Immunomodulatory Parameters in Response to Treatment with Either Combination Therapy with Prednisone and Metronidazole or Probiotic VSL#3 Strains in Dogs with Idiopathic Inflammatory Bowel Disease. PLoS ONE 2014, 9, e94699. [Google Scholar] [CrossRef]
  160. Gibson, G.R.; Hutkins, R.; Sanders, M.E.; Prescott, S.L.; Reimer, R.A.; Salminen, S.J.; Scott, K.; Stanton, C.; Swanson, K.S.; Cani, P.D.; et al. Expert consensus document: The International Scientific Association for Probiotics and Prebiotics (ISAPP) consensus statement on the definition and scope of prebiotics. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 491–502. [Google Scholar] [CrossRef] [Green Version]
  161. Venkataraman, A.; Sieber, J.R.; Schmidt, A.W.; Waldron, C.; Theis, K.R.; Schmidt, T.M. Variable responses of human microbiomes to dietary supplementation with resistant starch. Microbiome 2016, 4, 33. [Google Scholar] [CrossRef] [Green Version]
  162. Gibson, G.R.; Probert, H.M.; Van Loo, J.; Rastall, R.A.; Roberfroid, M.B. Dietary modulation of the human colonic microbiota: Updating the concept of prebiotics. Nutr. Res. Rev. 2004, 17, 259–275. [Google Scholar] [CrossRef] [Green Version]
  163. Bindels, L.B.; Neyrinck, A.M.; Salazar, N.; Taminiau, B.; Druart, C.; Muccioli, G.G.; François, E.; Blecker, C.; Richel, A.; Daube, G. Non Digestible Oligosaccharides Modulate the Gut Microbiota to Control the Development of Leukemia and Associated Cachexia in Mice. PLoS ONE 2015, 10, e0131009. [Google Scholar] [CrossRef] [Green Version]
  164. Macfarlane, G.T.; Steed, H.; Macfarlane, S. Bacterial metabolism and health-related effects of galacto-oligosaccharides and other prebiotics. J. Appl. Microbiol. 2008, 104, 305–344. [Google Scholar] [CrossRef]
  165. Sun, Y.; O’Riordan, M.X.D. Regulation of Bacterial Pathogenesis by Intestinal Short-Chain Fatty Acids. Adv. Appl. Microbiol. 2013, 85, 93–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Dewulf, E.M.; Cani, P.D.; Claus, S.P.; Fuentes, S.; Puylaert, P.G.B.; Neyrinck, A.M.; Bindels, L.B.; de Vos, W.M.; Gibson, G.R.; Thissen, J.P.; et al. Insight into the prebiotic concept: Lessons from an exploratory, double blind intervention study with inulin-type fructans in obese women. Gut 2013, 62, 1112–1121. [Google Scholar] [CrossRef] [PubMed]
  167. Kaplan, H.; Hutkins, R.W. Fermentation of fructooligosaccharides by lactic acid bacteria and bifidobacteria. Appl. Environ. Microbiol. 2000, 66, 2682–2684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Macfarlane, S.; Macfarlane, G.T.; Cummings, J.H. Review article: Prebiotics in the gastrointestinal tract. Aliment. Pharmacol. Ther. 2006, 24, 701–714. [Google Scholar] [CrossRef] [PubMed]
  169. Davis, L.M.G.; Martinez, I.; Walter, J.; Hutkins, R. A dose dependent impact of prebiotic galactooligosaccharides on the intestinal microbiota of healthy adults. Int. J. Food Microbiol. 2010, 144, 285–292. [Google Scholar] [CrossRef] [PubMed]
  170. Musilova, S.; Rada, V.; Marounek, M.; Nevoral, J.; Duskova, D.; Bunesova, V.; Vlkova, E.; Zelenka, R. Prebiotic Effects of a Novel Combination of Galactooligosaccharides and Maltodextrins. J. Med. Food 2015, 18, 685–689. [Google Scholar] [CrossRef]
  171. Markowiak, P.; Slizewska, K. Effects of Probiotics, Prebiotics, and Synbiotics on Human Health. Nutrients 2017, 9, 1021. [Google Scholar] [CrossRef]
  172. Nicolucci, A.C.; Hume, M.P.; Martinez, I.; Mayengbam, S.; Walter, J.; Reimer, R.A. Prebiotics Reduce Body Fat and Alter Intestinal Microbiota in Children Who Are Overweight or With Obesity. Gastroenterology 2017, 153, 711–722. [Google Scholar] [CrossRef] [Green Version]
  173. Chen, T.T.; Long, W.M.; Zhang, C.H.; Liu, S.; Zhao, L.P.; Hamaker, B.R. Fiber-utilizing capacity varies in Prevotella-versus Bacteroides-dominated gut microbiota. Sci. Rep. 2017, 7, 2594. [Google Scholar] [CrossRef]
  174. Lopez-Siles, M.; Khan, T.M.; Duncan, S.H.; Harmsen, H.J.M.; Garcia-Gil, L.J.; Flint, H.J. Cultured Representatives of Two Major Phylogroups of Human Colonic Fecalibacterium prausnitzii Can Utilize Pectin, Uronic Acids, and Host-Derived Substrates for Growth. Appl. Environ. Microbiol. 2012, 78, 420–428. [Google Scholar] [CrossRef] [Green Version]
  175. Chung, W.S.F.; Meijerink, M.; Zeuner, B.; Holck, J.; Louis, P.; Meyer, A.S.; Wells, J.M.; Flint, H.J.; Duncan, S.H. Prebiotic potential of pectin and pectic oligosaccharides to promote anti-inflammatory commensal bacteria in the human colon. FEMS Microbiol. Ecol. 2017, 93, fix127. [Google Scholar] [CrossRef] [PubMed]
  176. Gullon, B.; Gomez, B.; Martinez-Sabajanes, M.; Yanez, R.; Parajo, J.C.; Alonso, J.L. Pectic oligosaccharides: Manufacture and functional properties. Trends Food Sci. Technol. 2013, 30, 153–161. [Google Scholar] [CrossRef]
  177. Koutsos, A.; Lima, M.; Conterno, L.; Gasperotti, M.; Bianchi, M.; Fava, F.; Vrhovsek, U.; Lovegrove, J.A.; Tuohy, K.M. Effects of Commercial Apple Varieties on Human Gut Microbiota Composition and Metabolic Output Using an In Vitro Colonic Model. Nutrients 2017, 9, 533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Gomez, B.; Gullon, B.; Yanez, R.; Schols, H.; Alonso, J.L. Prebiotic potential of pectins and pectic oligosaccharides derived from lemon peel wastes and sugar beet pulp: A comparative evaluation. J. Funct. Foods 2016, 20, 108–121. [Google Scholar] [CrossRef]
  179. Geurts, L.; Neyrinck, A.M.; Delzenne, N.M.; Knauf, C.; Cani, P.D. Gut microbiota controls adipose tissue expansion, gut barrier and glucose metabolism: Novel insights into molecular targets and interventions using prebiotics. Benef. Microbes 2014, 5, 3–17. [Google Scholar] [CrossRef] [Green Version]
  180. Elison, E.; Vigsnaes, L.K.; Krogsgaard, L.R.; Rasmussen, J.; Sorensen, N.; McConnell, B.; Hennet, T.; Sommer, M.O.A.; Bytzer, P. Oral supplementation of healthy adults with 2-O-fucosyllactose and lacto-N-neotetraose is well tolerated and shifts the intestinal microbiota. Br. J. Nutr. 2016, 116, 1356–1368. [Google Scholar] [CrossRef] [Green Version]
  181. Rycroft, C.E.; Jones, M.R.; Gibson, G.R.; Rastall, R.A. A comparative in vitro evaluation of the fermentation properties of prebiotic oligosaccharides. J. Appl. Microbiol. 2001, 91, 878–887. [Google Scholar] [CrossRef] [Green Version]
  182. Baxter, N.T.; Schmidt, A.W.; Venkataraman, A.; Kim, K.S.; Waldron, C.; Schmidt, T.M. Dynamics of Human Gut Microbiota and Short-Chain Fatty Acids in Response to Dietary Interventions with Three Fermentable Fibers. mBio 2019, 10, 1–13. [Google Scholar] [CrossRef] [Green Version]
  183. Frei, R.; Akdis, M.; O’Mahony, L. Prebiotics, probiotics, synbiotics, and the immune system: Experimental data and clinical evidence. Curr. Opin. Gastroenterol. 2015, 31, 153–158. [Google Scholar] [CrossRef] [Green Version]
  184. Pandey, K.R.; Naik, S.R.; Vakil, B.V. Probiotics, prebiotics and synbiotics—A review. J. Food Sci. Technol. 2015, 52, 7577–7587. [Google Scholar] [CrossRef]
  185. Leenen, C.H.M.; Dieleman, L.A. Inulin and oligofructose in chronic inflammatory bowel disease. J. Nutr. 2007, 137, 2572S–2575S. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Plaza-Diaz, J.; Ruiz-Ojeda, F.J.; Vilchez-Padial, L.M.; Gil, A. Evidence of the Anti-Inflammatory Effects of Probiotics and Synbiotics in Intestinal Chronic Diseases. Nutrients 2017, 9, 555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Luoma, A.; Markazi, A.; Shanmugasundaram, R.; Murugesan, G.R.; Mohnl, M.; Selvaraj, R. Effect of synbiotic supplementation on layer production and cecal Salmonella load during a Salmonella challenge. Poult. Sci. 2017, 96, 4208–4216. [Google Scholar] [CrossRef] [PubMed]
  188. Wang, W.; Li, Z.; Han, Q.; Guo, Y.; Zhang, B.; D’inca, R. Dietary live yeast and mannan-oligosaccharide supplementation attenuate intestinal inflammation and barrier dysfunction induced by Escherichia coli in broilers. Br. J. Nutr. 2016, 116, 1878–1888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Rahimi, S.; Kathariou, S.; Fletcher, O.; Grimes, J.L. Effect of a direct-fed microbial and prebiotic on performance and intestinal histomorophology of turkey poults challenged with Salmonella and Campylobacter. Poult. Sci. 2019, 98, 6572–6578. [Google Scholar] [CrossRef]
  190. Markazi, A.D.; Perez, V.; Sifri, M.; Shanmugasundaram, R.; Selvaraj, R.K. Effect of whole yeast cell product supplementation (CitriStim®) on immune responses and cecal microflora species in pullet and layer chickens during an experimental coccidial challenge. Poult. Sci. 2017, 96, 2049–2056. [Google Scholar] [CrossRef]
  191. Pourabedin, M.; Chen, Q.; Yang, M.; Zhao, X. Mannan— Mannan- and xylooligosaccharides modulate caecal microbiota and expression of inflammatory-related cytokines and reduce caecal Salmonella Enteritidis colonisation in young chickens. FEMS Microbiol. Ecol. 2017, 93, fiw226. [Google Scholar] [CrossRef] [Green Version]
  192. Jazi, V.; Mohebodini, H.; Ashayerizadeh, A.; Shabani, A.; Barekatain, R. Fermented soybean meal ameliorates Salmonella Typhimurium infection in young broiler chickens. Poult. Sci. 2019, 98, 5648–5660. [Google Scholar] [CrossRef]
  193. Hughes, R.A.; Ali, R.A.; Mendoza, M.A.; Hassan, H.M.; Koci, M.D. Impact of Dietary Galacto-Oligosaccharide (GOS) on Chicken’s Gut Microbiota, Mucosal Gene Expression, and Salmonella Colonization. Front. Vet. Sci. 2017, 4, 192. [Google Scholar] [CrossRef] [Green Version]
  194. Azcarate-Peril, M.A.; Butz, N.; Cadenas, M.B.; Koci, M.; Ballou, A.; Mendoza, M.; Ali, R.; Hassan, H. An Attenuated Salmonella enterica Serovar Typhimurium Strain and Galacto-Oligosaccharides Accelerate Clearance of Salmonella Infections in Poultry through Modifications to the Gut Microbiome. Appl. Environ. Microbiol. 2018, 84, 1–16. [Google Scholar] [CrossRef] [Green Version]
  195. Jensen, A.N.; Mejer, H.; Mølbak, L.; Langkjær, M.; Jensen, T.K.; Angen, Ø.; Martinussen, T.; Klitgaard, K.; Baggesen, D.L.; Thamsborg, S.M.; et al. The effect of a diet with fructan-rich chicory roots on intestinal helminths and microbiota with special focus on Bifidobacteria and Campylobacter in piglets around weaning. Animal 2011, 5, 851–860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Myhill, L.J.; Stolzenbach, S.; Hansen, T.V.A.; Skovgaard, K.; Stensvold, C.R.; Andersen, L.O.; Nejsum, P.; Mejer, H.; Thamsborg, S.M.; Williams, A.R. Mucosal Barrier and Th2 Immune Responses Are Enhanced by Dietary Inulin in Pigs Infected With Trichuris suis. Front. Immunol. 2018, 9, 2557. [Google Scholar] [CrossRef] [PubMed]
  197. Stolzenbach, S.; Myhill, L.J.; Andersen, L.O.; Krych, L.; Mejer, H.; Williams, A.R.; Nejsum, P.; Stensvold, C.R.; Nielsen, D.S.; Thamsborg, S.M. Dietary Inulin and Trichuris suis Infection Promote Beneficial Bacteria Throughout the Porcine Gut. Front. Microbiol. 2020, 11, 312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Li, M.; Monaco, M.H.; Wang, M.; Comstock, S.S.; Kuhlenschmidt, T.B.; Fahey, G.C., Jr.; Miller, M.J.; Kuhlenschmidt, M.S.; Donovan, S.M. Human milk oligosaccharides shorten rotavirus-induced diarrhea and modulate piglet mucosal immunity and colonic microbiota. ISME J. 2014, 8, 1609–1620. [Google Scholar] [CrossRef] [Green Version]
  199. Ingerslev, H.C.; Strube, M.L.; Jørgensen, L.; Dalsgaard, I.; Boye, M.; Madsen, L. Diet type dictates the gut microbiota and the immune response against Yersinia ruckeri in rainbow trout (Oncorhynchus mykiss). Fish Shellfish Immunol. 2014, 40, 624–633. [Google Scholar] [CrossRef]
  200. Wang, X.; Sun, Y.; Wang, L.; Li, X.; Qu, K.; Xu, Y. Synbiotic dietary supplement affects growth, immune responses and intestinal microbiota of Apostichopus japonicus. Fish Shellfish Immunol. 2017, 68, 232–242. [Google Scholar] [CrossRef]
  201. El-Ashram, S.A.; Aboelhadid, S.M.; Abdel-Kafy, E.M.; Hashem, S.A.; Mahrous, L.N.; Farghly, E.M.; Moawad, U.K.; Kamel, A.A. Prophylactic and Therapeutic Efficacy of Prebiotic Supplementation against Intestinal Coccidiosis in Rabbits. Animals 2019, 9, 965. [Google Scholar] [CrossRef] [Green Version]
  202. Zentek, J.; Marquart, B.; Pietrzak, T.; Ballèvre, O.; Rochat, F. Dietary effects on bifidobacterial and Clostridium perfringens in the canine intestinal tract. J. Anim. Physiol. Anim. Nutr. (Berl.) 2003, 87, 397–407. [Google Scholar] [CrossRef] [PubMed]
  203. Pinna, C.; Vecchiato, C.G.; Bouldan, C.; Grandi, M.; Stefanelli, C.; Windisch, W.; Zaghini, G.; Biagi, G. Influence of dietary protein and fructooligosaccharides on fecal fermentative end -products, fecal bacterial populations and apparent total tract digestibility in dogs. BMC Vet. Res. 2018, 14, 106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Alexander, C.; Cross, T.-W.L.; Devendran, S.; Neumer, F.; Theis, S.; Suchodolski, J.S.; Godoy, M.R.C.; Swanson, K.S. Effects of prebiotic inulin-type fructans on blood metabolite and hormone concentrations and fecal microbiota and metabolites in overweight dogs. Br. J. Nutr. 2018, 120, 711–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Segarra, S.; Subiela, S.M.; Cerdà-Cuellar, M.; Martinez-Puig, D.; Munoz-Prieto, A.; Rodriguez-Franco, F.; Rodriguez-Bertos, A.; Allenspach, K.; Velasco, K.; Ceron, J. Oral chondroitin sulfate and prebiotics for the treatment of canine inflammatory bowel disease: A randomized controlled clinical trial. BMC Vet. Res. 2016, 12, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Rose, L.; Rose, S.; Gosling, S.; Holmes, M. Efficacy of a probiotic supplement on incidence of diarrhea in a dog shelter: A randomized, double-blind, placebo-controlled trial. J. Vet. Intern. Med. 2017, 31, 377–382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Lubbs, D.C.; Vester Boler, B.M.; Kerr, K.R.; Dowd, S.E.; Swanson, K.S. Dietary protein concentration affects intestinal microbiota of adult cats: A study using DGGE and qPCR to evaluate differences in microbial population in the feline gastrointestinal tract. J. Anim. Physiol. Anim. Nutr. (Berl.) 2009, 93, 113–121. [Google Scholar] [CrossRef] [PubMed]
  208. Barry, K.A.; Wojcicki, B.J.; Middelbos, I.S.; Vester, B.M.; Swanson, K.S.; Fahey, G.C., Jr. Dietary cellulose, fructooligosaccharides, and pectin modify fecal protein catabolites and microbial populations in adult cats. J. Anim. Sci. 2010, 88, 2978–2987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Vandeputte, D.; Kathagen, G.; D’Hoe, K.; Vieira-Silva, S.; Valles-Colomer, M.; Sabino, J.; Wang, J.; Tito, R.Y.; De Commer, L.; Darzi, Y.; et al. Quantitative microbiome profiling links gut community variation to microbial load. Nature 2017, 551, 507–511. [Google Scholar] [CrossRef] [PubMed]
  210. Rafter, J.; Bennett, M.; Caderni, G.; Clune, Y.; Hughes, R.; Karlsson, P.C.; Klinder, A.; O’Riordan, M.; O’Sullivan, G.C.; Pool-Zobel, B.; et al. Dietary synbiotics reduce cancer risk factors in polypectomized and colon cancer patients. Am. J. Clin. Nutr. 2007, 85, 488–496. [Google Scholar] [CrossRef]
  211. Aguilar-Toala, J.E.; Garcia-Varela, R.; Garcia, H.S.; Mata-Haro, V.; Gonzalez-Cordova, A.F.; Vallejo-Cordoba, B.; Hernandez-Mendoza, A. Postbiotics: An evolving term within the functional foods field. Trends Food Sci. Technol. 2018, 75, 105–114. [Google Scholar] [CrossRef]
  212. Konstantinov, S.R.; Kuipers, E.J.; Peppelenbosch, M.P. Functional genomic analyses of the gut microbiota for CRC screening. Nat. Rev. Gastroenterol. Hepatol. 2013, 10, 741–745. [Google Scholar] [CrossRef]
  213. Gosalbez, L.; Ramon, D. Probiotics in transition: Novel strategies. Trends Biotechnol. 2015, 33, 195–196. [Google Scholar] [CrossRef]
  214. Scheppach, W. Effects of short-chain fatty-acids on gut morphology and function. Gut 1994, 35, S35–S38. [Google Scholar] [CrossRef] [Green Version]
  215. Den Besten, G.; Lange, K.; Havinga, R.; van Dijk, T.H.; Gerding, A.; van Eunen, K.; Muller, M.; Groen, A.K.; Hooiveld, G.J.; Bakker, B.M.; et al. Gut-derived short-chain fatty acids are vividly assimilated into host carbohydrates and lipids. Am. J. Physiol. Gastrointest. Liver Physiol. 2013, 305, G900–G910. [Google Scholar] [CrossRef] [PubMed]
  216. Koh, A.; De Vadder, F.; Kovatcheva-Datchary, P.; Backhed, F. From Dietary Fiber to Host Physiology: Short-Chain Fatty Acids as Key Bacterial Metabolites. Cell 2016, 165, 1332–1345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Bedford, A.; Gong, J. Implications of butyrate and its derivatives for gut health and animal production. Anim. Nutr. 2018, 4, 151–159. [Google Scholar] [CrossRef]
  218. Liu, H.; Wang, J.; He, T.; Becker, S.; Zhang, G.L.; Li, D.F.; Ma, X. Butyrate: A Double-Edged Sword for Health? Adv. Nutr. 2018, 9, 21–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Wu, X.Q.; Wu, Y.B.; He, L.M.; Wu, L.H.; Wang, X.C.; Liu, Z.P. Effects of the intestinal microbial metabolite butyrate on the development of colorectal cancer. J. Cancer 2018, 9, 2510–2517. [Google Scholar] [CrossRef]
  220. Johnstone, R.W. Histone-deacetylase inhibitors: Novel drugs for the treatment of cancer. Nat. Rev. Drug Discov. 2002, 1, 287–299. [Google Scholar] [CrossRef]
  221. Singh, N.; Thangaraju, M.; Prasad, P.D.; Martin, P.M.; Lambert, N.A.; Boettger, T.; Offermanns, S.; Ganapathy, V. Blockade of Dendritic Cell Development by Bacterial Fermentation Products Butyrate and Propionate through a Transporter (Slc5a8)-dependent Inhibition of Histone Deacetylases. J. Biol. Chem. 2010, 285, 27601–27608. [Google Scholar] [CrossRef] [Green Version]
  222. Lukovac, S.; Belzer, C.; Pellis, L.; Keijser, B.J.; de Vos, W.M.; Montijn, R.C.; Roeselers, G. Differential Modulation by Akkermansia muciniphila and Fecalibacterium prausnitzii of Host Peripheral Lipid Metabolism and Histone Acetylation in Mouse Gut Organoids. mBio 2014, 5, 1–10. [Google Scholar] [CrossRef] [Green Version]
  223. Morrison, D.J.; Preston, T. Formation of short chain fatty acids by the gut microbiota and their impact on human metabolism. Gut Microbes 2016, 7, 189–200. [Google Scholar] [CrossRef] [Green Version]
  224. Gill, P.A.; van Zelm, M.C.; Muir, J.G.; Gibson, P.R. Short chain fatty acids as potential therapeutic agents in human gastrointestinal and inflammatory disorders. Aliment. Pharmacol. Ther. 2018, 48, 15–34. [Google Scholar] [CrossRef] [Green Version]
  225. O’Grady, J.; O’Connor, E.M.; Shanahan, F. Review article: Dietary fibre in the era of microbiome science. Aliment. Pharmacol. Ther. 2019, 49, 506–515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Dobson, A.; Cotter, P.D.; Ross, R.P.; Hill, C. Bacteriocin production: A probiotic trait? Appl. Environ. Microbiol. 2012, 78, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Korhonen, H.; Pihlanto, A. Food-derived bioactive peptides—Opportunities for designing future foods. Curr. Pharm. Des. 2003, 9, 1297–1308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Atanasova, J.; Moncheva, P.; Ivanova, I. Proteolytic and antimicrobial activity of lactic acid bacteria grown in goat milk. Biotechnol. Biotechnol. Equip. 2014, 28, 1073–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Lonnerdal, B.; Erdmann, P.; Thakkar, S.K.; Sauser, J.; Destaillats, F. Longitudinal evolution of true protein, amino acids and bioactive proteins in breast milk: A developmental perspective. J. Nutr. Biochem. 2017, 41, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Alkudsi, N.H.; Hussein, D.A. The concentration of lactoferrin and its relationship with minerals and amino acids in cows milk. Sci. Pap. Ser. Anim. Sci. 2016, 59, 157–163. [Google Scholar]
  231. Leboffe, L.; Giansanti, F.; Antonini, G. Antifungal and antiparasitic activities of lactoferrin. Anti Infect. Agents Med. Chem. 2009, 8, 114–127. [Google Scholar] [CrossRef]
  232. Embleton, N.D.; Berrington, J.E.; McGuire, W.; Stewart, C.J.; Cummings, S.P. Lactoferrin: Antimicrobial activity and therapeutic potential. Semin. Fetal Neonat Med. 2013, 18, 143–149. [Google Scholar] [CrossRef]
  233. Andersen, J.H.; Jenssen, H.; Gutteberg, T.J. Lactoferrin and lactoferricin inhibit Herpes simplex 1 and 2 infection and exhibit synergy when combined with acyclovir. Antivir. Res. 2003, 58, 209–215. [Google Scholar] [CrossRef]
  234. Mistry, N.; Drobni, P.; Nasland, J.; Sunkari, V.G.; Jenssen, H.; Evander, M. The anti-papillomavirus activity of human and bovine lactoferricin. Antivir. Res. 2007, 75, 258–265. [Google Scholar] [CrossRef]
  235. Berlutti, F.; Pantanella, F.; Natalizi, T.; Frioni, A.; Paesano, R.; Polimeni, A.; Valenti, P. Antiviral Properties of Lactoferrin-A Natural Immunity Molecule. Molecules 2011, 16, 6992–7018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Shestakov, A.; Jenssen, H.; Nordstrom, I.; Eriksson, K. Lactoferricin but not lactoferrin inhibit herpes simplex virus type 2 infection in mice. Antivir. Res. 2012, 93, 340–345. [Google Scholar] [CrossRef] [PubMed]
  237. Ng, T.; Cheung, R.; Wong, J.; Wang, Y.; Ip, D.; Wan, D.; Xia, J. Antiviral activities of whey proteins. Appl. Microbiol. Biotechnol. 2015, 99, 6997–7008. [Google Scholar] [CrossRef] [PubMed]
  238. Siqueiros-Cendon, T.; Arevalo-Gallegos, S.; Iglesias-Figueroa, B.F.; Garcia-Montoya, I.A.; Salazar-Martinez, J.; Rascon-Cruz, Q. Immunomodulatory effects of lactoferrin. Acta Pharmacol. Sin. 2014, 35, 557–566. [Google Scholar] [CrossRef]
  239. Brock, J.H. Lactoferrin—50 years on. Biochem. Cell Biol. 2012, 90, 245–251. [Google Scholar] [CrossRef]
  240. Kim, J.S.; Ellman, M.B.; Yan, D.Y.; An, H.S.; Ranjan, K.C.; Li, X.; Chen, D.; Xiao, G.Z.; Cs-Szabo, G.; Hoskin, D.W.; et al. Lactoferricin mediates anti-inflammatory and anti-catabolic effects via inhibition of IL-1 and LPS activity in the intervertebral disc. J. Cell Physiol. 2013, 228, 1884–1896. [Google Scholar] [CrossRef] [Green Version]
  241. Zhang, Y.L.; Lima, C.F.; Rodrigues, L.R. Anticancer effects of lactoferrin: Underlying mechanisms and future trends in cancer therapy. Nutr. Rev. 2014, 72, 763–773. [Google Scholar] [CrossRef] [Green Version]
  242. Kanwar, J.R.; Roy, K.; Patel, Y.; Zhou, S.F.; Singh, M.R.; Singh, D.; Nasir, M.; Sehgal, R.; Sehgal, A.; Singh, R.S.; et al. Multifunctional iron bound lactoferrin and nanomedicinal approaches to enhance its bioactive functions. Molecules 2015, 20, 9703–9731. [Google Scholar] [CrossRef]
  243. Giansanti, F.; Panella, G.; Leboffe, L.; Antonini, G. Lactoferrin from Milk: Nutraceutical and Pharmacological Properties. Pharmaceuticals 2016, 9, 61. [Google Scholar] [CrossRef] [Green Version]
  244. Arnold, R.R.; Brewer, M.; Gauthier, J.J. Bactericidal activity of human lactoferrin: Sensitivity of a variety of microorganisms. Infect. Immun. 1980, 28, 893–898. [Google Scholar]
  245. Humam, A.M.; Loh, T.C.; Foo, H.L.; Samsudin, A.A.; Mustapha, N.M.; Zulkifli, I.; Izuddin, W.I. Effects of Feeding Different Postbiotics Produced by Lactobacillus plantarum on Growth Performance, Carcass Yield, Intestinal Morphology, Gut Microbiota Composition, Immune Status, and Growth Gene Expression in Broilers under Heat Stress. Animals 2019, 9, 644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Loh, T.C.; Choe, D.W.; Foo, H.L.; Sazili, A.Q.; Bejo, M.H. Effects of feeding different postbiotic metabolite combinations produced by Lactobacillus plantarum strains on egg quality and production performance, fecal parameters and plasma cholesterol in laying hens. BMC Vet. Res. 2014, 10, 149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Kareem, K.Y.; Loh, T.C.; Foo, H.L.; Akit, H.; Samsudin, A.A. Effects of dietary postbiotic and inulin on growth performance, IGF1 and GHR mRNA expression, fecal microbiota and volatile fatty acids in broilers. BMC Vet. Res. 2016, 12, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Kareem, K.Y.; Loh, T.C.; Foo, H.L.; Asmara, S.A.; Akit, H. Influence of postbiotic RG14 and inulin combination on cecal microbiota, organic acid concentration, and cytokine expression in broiler chickens. Poult. Sci. 2017, 96, 966–975. [Google Scholar] [CrossRef]
  249. Johnson, C.N.; Kogut, M.H.; Genovese, K.; He, H.; Kazemi, S.; Arsenault, R.J. Administration of a Postbiotic Causes Immunomodulatory Responses in Broiler Gut and Reduces Disease Pathogenesis Following Challenge. Microorganisms 2019, 7, 268. [Google Scholar] [CrossRef] [Green Version]
  250. Thu, T.V.; Loh, T.C.; Foo, H.L.; Yaakub, H.; Bejo, M.H. Effects of liquid metabolite combinations produced by Lactobacillus plantarum on growth performance, faeces characteristics, intestinal morphology and diarrhoea incidence in postweaning piglets. Trop. Anim. Health Prod. 2011, 43, 69–75. [Google Scholar] [CrossRef] [Green Version]
  251. Pérez-Sánchez, T.; Mora-Sánchez, B.; Vargas, A.; Balcázar, J.L. Changes in intestinal microbiota and disease resistance following dietary postbiotic supplementation in rainbow trout (Oncorhynchus mykiss). Microb. Pathog. 2020, 142, 104060. [Google Scholar] [CrossRef]
  252. Mora-Sánchez, B.; Balcázar, J.L.; Pérez-Sánchez, T. Effect of a novel postbiotic containing lactic acid bacteria on the intestinal microbiota and disease resistance of rainbow trout (Oncorhynchus mykiss). Biotechnol. Lett. 2020, 42, 1957–1962. [Google Scholar] [CrossRef]
  253. Wu, X.; Teame, T.; Hao, Q.; Ding, Q.; Liu, H.; Ran, C.; Yang, Y.; Zhang, Y.; Zhou, Z.; Duan, M.; et al. Use of a paraprobiotic and postbiotic feed supplement (HWF™) improves the growth performance, composition and function of gut microbiota in hybrid sturgeon (Acipenser baerii x Acipenser schrenckii). Fish Shellfish Immunol. 2020, 104, 36–45. [Google Scholar] [CrossRef]
  254. Wernimont, S.M.; Radosevich, J.; Jackson, M.I.; Ephraim, E.; Badri, D.V.; MacLeay, J.M.; Sucholdoski, J.S. The effects of nutrition on the gastrointestinal microbiome of cats and dogs: Impact on health and disease. Front. Microbiol. 2020, 11, 1266. [Google Scholar] [CrossRef]
  255. Sawada, D.; Sugawara, T.; Ishida, Y.; Aihara, K.; Aoki, Y.; Takehara, I.; Takano, K.; Fujiwara, S. Effect of continuous ingestion of a beverage prepared with Lactobacillus gasseri CP2305 inactivated by heat treatment on the regulation of intestinal function. Food Res. Int. 2016, 79, 33–39. [Google Scholar] [CrossRef] [Green Version]
  256. Wakabayashi, H.; Oda, H.; Yamauchi, K.; Abe, F. Lactoferrin for prevention of common viral infections. J. Infect. Chemother. 2014, 20, 666–671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Luceri, C.; Femia, A.P.; Fazi, M.; Di Martino, C.; Zolfanelli, F.; Dolara, P.; Tonelli, F. Effect of butyrate enemas on gene expression profiles and endoscopic/histopathological scores of diverted colorectal mucosa: A randomized trial. Dig. Liver Dis. 2016, 48, 27–33. [Google Scholar] [CrossRef] [PubMed]
  258. Jackson, M.I.; Jewell, D.E. Balance of saccharolysis and proteolysis underpins improvements in stool quality induced by adding a fiber bundle containing bound polyphenols to either hydrolyzed meat or grain-rich foods. Gut Microbes 2019, 10, 298–320. [Google Scholar] [CrossRef] [Green Version]
Table 1. Impact of probiotics on gastrointestinal health: main experimental studies.
Table 1. Impact of probiotics on gastrointestinal health: main experimental studies.
Animal SpeciesChallengingProbioticDoseImpact on Gut HealthReferences
HumansCampylobacterEnterococcus faecium LAB SF68108 CFU three times dailyTreatment of acute diarrhoea in adultsAllen et al. (2010) [156]
HumansUlcerative colitisLactobacillus rhamnosus NCIMB 30174, Lactobacillus plantarum NCIMB 30173, Lactobacil-lus acidophilus NCIMB 30175 and Enterococcus faecium NCIMB 30176 Suspension of barley containing about 10 billion live bacteria ↓Intestinal inflammation
↓Faecal calprotectin levels in the UC patients receiving the probiotic.
Bjarnason et al. (2019) [102]
HumansEnterococcus,
Pseudomonas,
Enteritis in patients with sepsi
B. breve,
L. casei
1 × 108 CFU/g,
1 × 108 CFU/g
↑Beneficial bacteria and SCFAs
↓Incidence of enteritis and VAP in patients with sepsis
Shimizu et al. (2018) [157]
HumansStaphylococcus aureus and Streptocococcus epidermidis Lactobacillus fermentum CECT5716 1 × 1010 CFU/g ↑Breastfeeding duration
↓Antibiotic usage
Bond et al. (2017) [158]
Broiler chickensEscherichia coli O78Lactobacillus plantarum K34, Bacillus subtilis MORI 91 and Clostridium butyricum M72 × 108 CFU/g, 2 × 108 CFU/g, 2.06 × 108 CFU/g↑total lactobacilli and total lactobacilli-enterococciTarabees et al. (2019) [112]
Broiler chickens Salmonella TyphimuriumLactobacillus casei ATTC334, Bifidobacterium breve JCM1192 and Bifidobacterium infantis BL2416108 CFU/mLSalmonella Typhimurium recovery from caecal tonsils
↑IFN-γ and TNF-α
El-Sharkawy et al. (2020) [113]
Broiler chickensEimeria spp.Enterococcus faecium2 × 1012 CFU/kg↓histopathological scores
↓oocyst counts in faecal samples
El-Sawah et al. (2020) [114]
Broiler chickensEscherichia coli K88Clostridium butyricum2 × 107 CFU/kg of feed↑TNF-α and IL-4
↑Vh and ↓Cd
↑amylase, protease and lipase
Zhang et al. (2016) [116]
Broiler chickensSalmonella enteriditisClostridium butyricum106 CFU/0.2 mL↓INF-γ, IL-1β, IL-8 and TNF-αZhao et al. (2017) [117]
Broiler chickensClostridium perfringensClostridium butyricum1 × 109 CFU/kg↓IL-17A
↑Claudin-1 and IL-10
Clostridium perfringens
Huang et al. (2019) [118]
Broiler chickensClostridium perfringensBacillus subtilis DSM 323151 × 106 CFU/g of feed↑microbial alpha-diversity ↑Ruminococcus, Ruminococcaceae and Bacteroides
Clostridium perfringens
Bortoluzzi et al. (2019) [120]
Broiler chickensClostridium perfringensBacillus subtilis DSM 323152 × 109 CFU/gBacillus spp., Lactobacillaceae, Lactobacillus salivarius and Lactobacillus johnsonii
Clostridium perfringens
Whelan et al. (2018) [121]
Broiler chickensCampylobacter jejuniSaccharomyces cerevisiae boulardii CNCM I-10791 × 109 CFU/kg↑Vh
Lactobacillus reuteri and Fecalibacterium prausnitzii
Campylobacter spp.
Massacci et al. (2019) [122]
PigletsEscherichia coli F4 (K88)Lactobacillus rhamnosus ACTT 74691010 CFU/day or 1012 CFU/day↓diarrhoea frequency
↓faecal coliforms
Lactobacillus and Bifidobacterium
↓TNF-α, IL-8 and pBD2
Li et al. (2012) [123]
PigsSalmonella enterica serovar 4Lactobacillus rhamnosus GG1 × 109 CFU/mL (10 mL/day)↓Prevotellaceae NK3B31 ↑Fusobacterium, Lactobacillus animalis and Propionibacterium
↓CD3CD19cell subsets
Zhang et al. (2019) [124]
PigsSalmonella InfantisLactobacillus rhamnosus GG1.0 × 1010 CFU/day↑T-bet
Activation of STAT3
↓CCL20
SalmonellaBifidobacterium
Yang et al. (2017) [125]
PigsEscherichia coli K88Lactobacillus acidophilus0, 0.5, 1.0 and 2.0 mL/kg of feedLactobacillus
Escherichia coli K88
Nordeste et al. (2017) [126]
PigletsEscherichia coliBacillus licheniformis and Bacillus subtilis3.9 × 108 CFU/day (low dose), 7.8 × 108 CFU/day (moderate dose), 3.9 × 109 CFU/day (high dose)Low/moderate doses:
Bacteroides uniformis, Eubacterium eligens, Acetanaerobacterium and Sporobacter
Clostridium, Turicibacter, Lactobacillus gasseri and Lactobacillus salivarius
↑GC number
High doses:
↑Proteobacteria
↓GC function
Zhang et al. (2017) [127]
PigletsEscherichia coli F4acBacillus amyloliquefaciens DSM25840 and Bacillus subtilis DSM258411.28 × 106 CFU/g feed↓Enterobacteriaceae
↑immunity genes
↑amino acid metabolism and utilization
Luise et al. (2019) [128]
Rainbow troutStreptococcus iniaeEnterococcus casseliflavus107 CFU/g of feed, 108 CFU/g of feed, 109 CFU/g of feed↑lactic acid bacteria and total viable aerobic counts
↑total serum protein and albumin, IgM, C3 complement and lysozyme
↑respiratory burst activity of blood leukocytes
Safari et al. (2016) [129]
Nile tilapiaAeromonas hydrophila and Streptococcus iniaeRummeliibacillus stabekisii106 CFU/g, 107 CFU/g↑leukocytes and lysozyme
↑IL-1β, TNF-α, TGF-β and heat shock protein 70
Bacillus and LactobacillusStreptococcus and Staphylococcus
Tan et al. (2019) [130]
Nile tilapiaAeromonas hydrophilaBacillus velezensis TPS3N, Bacillus subtilis TPS4 and Bacillus amyloliquefaciens TPS171.0 × 108 CFU/ml↑nitric oxide, IgM and AKP
↑CAT and SOD
↑Vh, Vw, GC count and muscle thickness
Staphylococcus and Aeromonas
Kuebutornye et al. (2019) [131]
Ornamental fish (Poecilia latipinna)Aeromonas hydrophila (BCCM5/LMG3770)Bacillus subtilis-enriched Artemia1 × 105 CFU/mLBacillusAhmadifard et al. (2019) [132]
Infant rabbitsShiga toxin-producing Escherichia coli (STEC)Lactobacillus casei strain Shirota2 × 103 CFU/mL↑IgA
↓STEC
↓Stx
↓histological damage to the gut mucus
Ogawa et al. (2001) [133]
RabbitsEscherichia coli AZ1Lactobacillus casei SABA61 × 108 CFU/mL↓Eruption of the intestinal epithelial cells
↓Diarrhoea incidence
↑IL-6 and IL-10
Fayyaz et al. (2018) [134]
DogsAereomonas, psedudomonas, coliformsLactobacillus fermentum AD13 × 109 CFU/mL↑Faecal lattobacilli
↓Proteobacteria
Strompfova et al. (2012) [148]
DogsIBDL. casei, L. plantarum. L. acidophilus, L. delbrueckii subsp. Bulgaricus, B. longum, B. breve, and B. infantis and Streptococcus sulivarius subsp thermophilus112 to 225 × 109 CFU × 10 kg↑Increases in Faecalibacterium spp. and Lactobacillus spp. Rossi et al. (2014) [159]
↑ = increase; ↓ = decrease.
Table 2. Impact of prebiotics on gastrointestinal health: main experimental studies.
Table 2. Impact of prebiotics on gastrointestinal health: main experimental studies.
Animal SpeciesChallengingPrebioticDoseImpact on Gut HealthReferences
HumansChanges in the faecal microbiota2′-O-fucosyllactose (2′FL) and/or lacto-N-neotetraose (LNnT) 5, 10 or 20 g per day↑Bifidobacteria abundance,
= abundance beneficial taxa such as Faecali bacterium, Roseburia, Akkermansia or Lactobacillus
Elison S. et al. (2016) [180]
HumansNoneInulin12 g per daygrowth of a limited number of colon bacteria,
↓Bilophila growth
Vandeputte et al. (2017) [209]
HumansObesity and associated dysbiosisInulin/oligofructose8 g twice a dayChanges in the gut microbiota correlated in ↓ fat mass, ↓ serum LPS levels and ↓ metabolism (hippurate, lactate and PC) Dewulf et al. (2013) [166]
HumansObesity in childrenoligofructose-enriched inulin8 g per day↑Bifidobacterium and
↓Bacteroides vulgatus
↓body weight, body fat, percent trunk fat
Nicolucci et al. (2017) [172]
HumansClostridium perfringensoligofructose- enriched inulin, Lactobacillus rhamnosus and Bifidobacterium lactis Bb12 Lactobacillus rhamnosus and Bifidobacterium lactis Bb12 > log10 CFU/g product
oligofructose/inulin 8 g per day
↑bifidobacteria and lactobacilli,
Clostridium perfringens,
Rafter et al. (2007) [210]
Broiler chickens Escherichia coliMOS (together with Saccharomyces cerevisiae)0.5%↓TLR4, NF-κB p65 and IL- 1β
↑IL-10 and occludin
↑Vh and Vh/Cd
↑bacterial diversity
Wang et al. (2016) [188]
Turkey poultsSalmonella and CampylobacterMOS and Saccharomyces cerevisiae wall (together with a direct-fed microbial)0.05%↑villi surface area and Vh/Cd
Salmonella shedding
Rahimi et al. (2019) [189]
Pullet and layer chickens Eimeria spp.Whole yeast cell product0, 0.1 or 0.2%↓faecal and gut coccidial oocyst counts
Lactobacillus
↓CD8+ T cell in the cecal tonsils
↓IL-10
Markazi et al. (2017) [190]
Young broiler chickensSalmonella enteriditisMOS and XOS1% and 2%Salmonella enteriditis counts
Coprococcus, Ruminococcus, Enterococcus, Clostridium, Lactobacillus and Roseburia
↓TNF-α and IFN-γ
Pourabedin et al. (2016) [191]
Broiler chickensSalmonella TyphimuriumXOS (together with Lactobacillus-based probiotic and fermented soybean meal)2%Salmonella Typhimurium colonization
↑intestinal lactic acid bacteria
↑mucosal morphology
Jazi et al. (2019) [192]
White Leghorn chickensSalmonella spp.FOS and probiotics0.5 or 1%Salmonella enteriditis Luoma et al. (2017) [187]
White Leghorn chickensSalmonella spp.GOS1%Salmonella TyphimuriumHughes et al. (2017) [193]
White Leghorn chickensSalmonella spp.GOS1%Salmonella Typhimurium
Alistipes, Lactobacillus reuteri and Christensenellaceae
Azcarate-Peril et al. (2018) [194]
Broiler chickensEscherichia coli O78IMO (together with a probiotic mix)0.5%↑total lactobacilli and total lactobacilli-enterococciTarabees et al. (2019) [112]
PigletsAscaris suum, Trichuris suis and Escherichia coli O138:F8Fructan-rich chicory roots30%Ascaris suum worms
Trichuris suis worms
Enterobacter and Campylobacter
↑Lachnospiraceae and Bifidobacteria
Jensen et al. (2011) [195]
PigsTrichuris suisInulin10%↑IL-13, IL-5 and TFF3
↓IFN-γ, IL-1α, IL-8 and CXCL9
↓Proteobacteria and Firmicutes
↑Actinobacteria, Bacteroidetes and Bacteroidetes:Firmicutes ratio
↓pH
Myhill et al. (2018) [196]
PigsTrichuris suisInulin10%↑Prevotella
↓Proteobacteria
Stolzenbach et al. (2020) [197]
Pigs Rotavirus OSUHMO and GOS-FOS mixture4 g/L and 3.6 g/L–0.4 g/L↓diarrhoea duration
↑IFN-γ and IL-10
↑IgM
↑Lachnospiraceae
Li et al. (2014) [198]
Rainbow troutYersinia ruckeriPlant-based dietReplacement of 10% of fishmealStreptococcus, Leuconostoc and Weissella counts
Yersinia ruckeri counts
↓IL-1β and MBL-2
Ingerslev et al. (2014) [199]
Sea cucumberVibrio spp.AOS and Bacillus licheniformis WS-2 10%↑immune enzyme activities
Bacillus and Lactococcus
Vibrio
Wang et al. (2017) [200]
RabbitsEimeria spp.Saccharomyces cerevisiae cell wall, MOS and dried Saccharomyces cerevisiae fermentation solubles2 g/L↓faecal oocyst and endogenous stage counts El-Ashram et al. (2019) [201]
Dogsfaecal fermentative end- products FOS0.5%↑Growth of Bifidobacteria Pinna et al. (2018) [203]
CatEscherichia coliFOS and cellulose 0.05%↑Bifidobacteria
↓Escherichia coli
Barry et al. (2010) [208]
Cat Clostridium perfringensDiet with the supplement of FOS and pectin 1% and 2%↓Clostridium perfringens
↑lactobacilli
Barry et al. (2010) [208]
↑ = increase; ↓ = decrease.
Table 3. Impact of postbiotics on gastrointestinal health: main experimental studies.
Table 3. Impact of postbiotics on gastrointestinal health: main experimental studies.
Animal SpeciesChallengingPostbioticDoseImpact on Gut HealthReferences
HumansRegulation of intestinal functionLactobacillus gasseri CP2305 1 × 1010 bacterial cells per container ↑ F. prausnitzii,
↑ SCFAs
Sawada et al. (2016) [255]
HumansViral gastroenteritis (Rotavirus)Lactoferrin Apo-bLF 50 mg/mLInhibition of cytopathic effect In human colon adenocarcinoma (HT-29) cellsWakabayashi et al. (2014) [256]
HumansDiverted colorectal mucosaSodium butyrate30 mL of sodium butyrate 600 mmol/LenemaEndoscopic scores were significantly improved
↑Upregulation of genes associated with mucosal repair
Luceri et al. (2016) [257]
Laying hens NoneLactobacillus plantarum TL1, RS5, RG14, RG11 and RI110.6%↑Lactic acid bacteria
↓Enterobacteriace
↑Propionic acid
Loh et al. (2014) [246]
Broiler chickensNoneLactobacillus plantarum RI11 and RG140.3%↓pH
↑Lactic acid bacteria
↓Enterobacteriaceae
↑Acetic acid
Kareem et al. (2016) [247]
Broiler chickensNoneLactobacillus plantarum RG140.15, 0.30, 0.45 and 0.6%↑Bifidobacteria
Escherichia coli
↓IFN-γ and LITAF
↑IL-6
Kareem et al. (2017) [248]
Broiler chickensHeat stressLactobacillus plantarum RI11, RS5 and UL40.3%↑Vh, ↓Cd and ↑Vh/Cd
Lactobacillus and Bifidobacteria
↓Enterobacteriaceae, Salmonella and Escherichia coli
Humam et al. (2019) [245]
Broiler chickensClostridium perfringensPediococcus acidilactici B-67717, Lactobacillus reuteri B-67718, Enterococcus faecium B-67720 and Lactobacillus acidophilus B-677011 ounce/gallon↓Lesion scores
Clostridium perfringens counts
Johnson et al. (2019) [249]
PigletsNoneLactobacillus plantarum TL1, RG11, RI11, RG14 and RS50.3%↓diarrhoea incidence
↑lactic acid bacteria
↓Enterobacteriaceae
↑SCFAs
↓pH
↑Vh
Thu et al. (2011) [250]
Hybrid sturgeonNoneRhodotorula and Cetobacterium5%↑Firmicutes and Clostridium
↓Proteobacteria, Actinobacteria, Chlamydiae and Lactococcus
Wu et al. (2020) [253]
Rainbow troutLactococcus garvieaeLactobacillus and Leuconostoc3%↑bacterial diversity and richness ↑Firmicutes, Lactobacillus, Enhydrobacter, Paracoccus and Pseudomonas
↓Proteobacteria and Fusobacteria
Pérez-Sánchez et al. (2020) [251]
Rainbow troutNoneLactobacillus3%↑Tenericutes, Bacteroidetes and Spirochaetes
↓Fusobacteria
Mora-Sánchez et al. (2020) [252]
DogsGut healthbSCFA; isobutyrate (2-methylpropionate) Substrate from Protein; valine Source of energy for colonocytes; refeed starved colonocytes with greater efficiency and rapidity than butyrate Koh et al. (2016) [216]
DogsGut healthSCFA; butyrate Substrate from Carbohydrate; indigestible polysaccharides ↑Energy substrate for colonocytes; epigenetic modulation; colon electrolyte balance, motility, blood flow Jackson et al. (2019) [258]
↑ = increase; ↓ = decrease.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Saettone, V.; Biasato, I.; Radice, E.; Schiavone, A.; Bergero, D.; Meineri, G. State-of-the-Art of the Nutritional Alternatives to the Use of Antibiotics in Humans and Monogastric Animals. Animals 2020, 10, 2199. https://doi.org/10.3390/ani10122199

AMA Style

Saettone V, Biasato I, Radice E, Schiavone A, Bergero D, Meineri G. State-of-the-Art of the Nutritional Alternatives to the Use of Antibiotics in Humans and Monogastric Animals. Animals. 2020; 10(12):2199. https://doi.org/10.3390/ani10122199

Chicago/Turabian Style

Saettone, Vittorio, Ilaria Biasato, Elisabetta Radice, Achille Schiavone, Domenico Bergero, and Giorgia Meineri. 2020. "State-of-the-Art of the Nutritional Alternatives to the Use of Antibiotics in Humans and Monogastric Animals" Animals 10, no. 12: 2199. https://doi.org/10.3390/ani10122199

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