SciELO - Scientific Electronic Library Online

 
vol.9Evaluation of the anti-Newcastle disease vaccine serum on dogs with canine distemperPharmacological effect of T-NilPlus® and Synertox®, commercial anti-mycotoxins, on broiler chickens in Egypt índice de autoresíndice de materiabúsqueda de artículos
Home Pagelista alfabética de revistas  

Servicios Personalizados

Revista

Articulo

Indicadores

Links relacionados

  • No hay artículos similaresSimilares en SciELO

Compartir


Veterinaria México OA

versión On-line ISSN 2448-6760

Veterinaria México OA vol.9  Ciudad de México  2022  Epub 16-Mayo-2023

https://doi.org/10.22201/fmvz.24486760e.2022.985 

Original research articles

High levels of dietary digestible protein transiently promote beneficial bacteria in adult dog feces

Karina Elizabeth Cosío-Carpintero1 
http://orcid.org/0000-0002-3908-0234

Carlos Gutiérrez Olvera1 
http://orcid.org/0000-0002-7088-7159

Claudia Cecilia Márquez-Mota1 
http://orcid.org/0000-0002-4311-381X

María Esther Ortega-Cerrilla2 
http://orcid.org/0000-0001-9971-5169

María Guadalupe Sánchez-González3 
http://orcid.org/0000-0003-0253-1369

Lilia Gutiérrez Olvera4 
http://orcid.org/0000-0002-4823-0388

1Universidad Nacional Autónoma de México. Facultad de Medicina Veterinaria y Zootecnia. Departamento de Bioquímica y Nutrición Animal. Ciudad de México, México

2Instituto de Recursos Genéticos y Productividad. Colegio de Postgraduados. Km. 36.5, México 136 5, Montecillo, 56230.

3Universidad Nacional Autónoma de México. Facultad de Medicina Veterinaria y Zootecnia. Departamento de Genética y Bioestadística. Ciudad de México, México.

4Universidad Nacional Autónoma de México. Facultad de Medicina Veterinaria y Zootecnia. Departamento de Fisiología y Farmacología. Ciudad de México, México.


Abstract

Nowadays, many commercial dog diets select the ingredients not only to meet the needs of the individual and maintain a balance between nutrition/ microbiome/health; fiber and protein quality is known to be one of the most important factors in this balance. Diets high in low digestible protein tend to favor microorganisms that when fermenting amino acids generate substances that induce inflammation, while high digestible protein and fermentable fibers have been related to beneficial bacteria. The study’s objective was to identify changes in the relative abundance of specific microorganisms (Clostridium perfringens, Enterococcus faecium, Lactobacillus salivarius, Bacteroides fragilis and Fusobacterium varium) by PCR (polymerase chain reaction), associated with two diets of different quality and digestibility. Twenty adult dogs were used, divided into two groups, the first one fed with a high digestibility diet (HD) (n = 10), the second one with a low digestibility diet (LD) (n = 10). After 3 days of adaptation to the diet, fecal samples were taken at days 15 and 30. The results showed that the high-quality diet promotes a transient increase (15 days) in the relative abundance of F. varium and E. faecium, as well as a persistent increase in that of L. salivarius and B. fragilis until day 30. Apparently, however, healthy adult dogs eventually balance their fecal microbiota, regardless of the dietary protein level and digestibility. Therefore, it is difficult to identify clear patterns of the ideal dietary profile in this species.

Keywords: dog; microbiota; diet; quality; protein-digestibility

Study contribution

Intestinal microbiota is an important topic in gastroenteric human and veterinary health. This study shows that high levels of digestible protein transiently promote beneficial bacteria in dog feces. It also demonstrated that, at least in healthy dogs, a low digestibility diet does not increase the abundance of pathogenic bacteria. Therefore, healthy dogs on either high or low protein content and digestibility diets can standardize and balance their microbiota.

Introduction

Dog and cat nutrition is an area that has gained interest recently. Nowadays, processed animal feed is designed not only to nourish but also to offer health benefits through the ingredients used in their formulation.1,2 Macronutrients, such as carbohydrates, fats, and protein, have a great impact on the intestinal microbiome in cats and dogs.3-4 Hence, intense research has focused on studying the interactions between microorganisms, food, and host, both in healthy and diseased animals.5

Part of the main associations found, include the fermentation of non-digestible carbohydrates, such as fiber. This dietary component favors greater production of volatile fatty acids, specifically butyrate, which is a nutritional source for enterocytes and has antineoplastic properties, thus being a fundamental part of gastrointestinal health.6,3 Conversely, proteins and amino acids used by proteolytic bacteria tend to increase fecal pH and stimulate the production of harmful metabolites for the intestine, such as ammonia, indoles and phenols.3,7

Key microorganisms that actively participate in fermentative processes, such as Eubacterium, Bacteroides, Clostridium, Peptococcus, Bifidobacterium, and Lactobacillus have also been identified.8 Some of these microorganisms interact with each other modify the microenvironment, prevent the colonization of pathogens, intervene in the digestion of some nutrients, and stimulate the immune system.9,10 Others, such as Clostridium perfringens, increase the concentration of putrefaction compounds,10,8 as well as biogenic amines that generate possible inflammatory effects, associated with chronic diseases.9 These predominant microorganisms can maintain stability, which is undoubtedly altered positively or negatively depending mainly on the diet consumed by the animal.11

Studies on how protein, fiber, carbohydrates, and other nutrients influence the microbiota have been increasing and have focused on comparing raw diets (BARF = Biologically Appropriate Raw Food) with commercial diets. It has been observed that food with high protein and fat, obtained from natural sources, reduce the proportion of genera such as Lactobacillus spp., Paralactobacillus spp., and Prevotella spp.,12 as well as Proteobacteria.5 Other studies that have evaluated the effect of protein and its sources (by-products) have determined that high concentrations of this macronutrient favor Fusobacteria.13 Other studies on functional fibers and prebiotics, such as inulin and fructans, have reported a reduction in the concentration of Fusobacteria as well as an increase in Firmicutes, although results may vary across individuals.14

The digestibility of the diet also plays an essential role in the microbiome. Everything that is not digested and absorbed in the small intestine by the animal represents a substrate available for bacteria in the colon.4 This study aimed to evaluate the effect of protein digestibility and the quality of two processed diets on the relative abundance of specific microorganisms.

Materials and methods

Ethical statement

This study was evaluated and approved by the Institutional Subcommittee for the Care and Use of Experimental Animals (SICUAE), Faculty of Veterinary Medicine, National Autonomous University of Mexico, with protocol number DC-2018/2-10.

Animals

Twenty clinically healthy, non-sterilized adult dogs, 15 females and 5 males, different breeds between 2 and 6 years old, with an average weight of 7 kg, were used in a completely randomized design. Two groups were randomly formed; the food selection was made considering the parameters established by AFFCO (2011) for dogs in maintenance. The first group was fed with high protein kibble (>22 % CP in DM), high energy density (> 3 500 kcal/100 g) and protein digestibility greater than 75 % (high quality-digestibility (HD); the second group was fed a minimum level of protein (20 % CP in DM), low energy density (< 3 200 kcal/100 g) and protein digestibility less than 75 % low quality-low digestibility (LD).

A 3-day adaptation period to the new food was contemplated to avoid digestive problems. To determine the amounts to be offered, the energy requirement of the animals was calculated individually and based on the following formula:15

Maintenance energy requirement (REM) = (live weight)0.75 × 95) kcal

All animals were kept in house conditions, without cage confinement, with access to the yard to reduce stress, without additional food and evaluated for the presence of undesirable behaviors such as coprophagia. At the beginning, middle and end of the experimental period, the body condition, muscle mass and weight of the animals were evaluated.

Food

The food was weighed for individual administration to each experimental subject, offered in a daily intake in which total consumption was guaranteed. The offered food was subjected to a bromatological study to determine moisture, crude protein (CP), crude fat, crude fiber, ashes, carbohydrates, as well as true protein (TP), indigestible protein (IP), in vitro protein digestibility (IVPD), neutral detergent fiber (NDF) and acid detergent fiber (ADF). All these analyses were performed according to the reference methods of AOAC International,16 except for NDF and ADF17 (Table 1).

Table 1 Means and standard deviations of proximate chemical composition (n = 3) of the experimental diets (wet basis) 

Analyte (%) Diet1
HD LD
Moisture 5.0 ± 0.87 7.5 ± 0.42
Crude protein 33.8 ± 0.28 20.1 ± 2.1
Crude fat 14.9 ± 0.01 7.1 ± 0.01
Crude fiber 1.5 ± 0.2 4.1 ± 0.01
Ashes 5.7 ± 0.01 9.0 ± 0.01
Carbohydrates 42.0 ± 1.25 53.0 ± 1.8
True protein 33.6 ± 0.07 19.1 ± 0.01
Indigestible protein 9.6 ± 0.01 -
Protein digestibility 71.8 ± 0.09 12.3 ± 0.28
Neutral detergent fiber 10.3 ± 0.32 8.5 ± 0.2
Acid detergent fiber 28.6 ± 0.36 31.5 ± 0.15

1 HD: high quality and digestibility; LD: low quality and digestibility

Stool samples

Samples were collected on the day the transition was completed (0, 15, and 30) of the experimental period. Excreta were collected immediately after defecation, ensuring that they did not touch the ground. Hermetic bags were used to preserve them and kept in deep-freezing conditions (-70 °C).

DNA extraction

For DNA extraction, the bacterial pellet was first obtained using 2 g of samples and 5 mL of 1× PBS (137-mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, 2 mM KH2PO4). A modified centrifugation was performed at 200 × g for 5 min and 1 mL of the supernatant was recovered in a 2 mL Eppendorf tube. To obtain a cleaner pellet, two more centrifugation steps (200 × g for 5 min) were performed, decanting the liquid and recovering the pellet at each step, ending with a centrifugation at 10 000 rpm for 15 min and recovering the pellet in 1 mL of 1 × PBS.

In the final pellet, bacterial genomic (BG) DNA was extracted using a modified cetyl trimethyl ammonium bromide (CTAB, 10%) technique16,17 that includes precipitation in isopropanol leaving the sample under refrigeration for at least 12 h. We used the following modifications of the technique: A) Re-suspend the pellet obtained from washing in 570 µL of TE buffer + 50 µL of lysozyme (10 mg/mL) incubating at 37 °C for 30 min. B) Incubate at 56 °C for 1 h after the addition of 30 µL sodium dodecyl sulfate + 4 µL proteinase K.

The integrity of the DNA was confirmed by agarose gel electrophoresis (0.8 % agarose) run at 70 V and 300 mA for 45 min. The DNA was visualized using a gel photodocumentation system (Figure 1). Subsequently, quantifications were performed in a Nanodrop® spectrophotometer.

Figure 1 Example of a visualized gel showing bacterial genomic DNA integrity 

Real-time PCR

All samples were standardized at 4 ng/mL of DNA concentration, from which qPCR was performed using the KAPA SYBR® FAST for light cycler® 480 kit. The reaction volume was 12 µL (6µL KAPA SYBR® FAST + 3.4 µL nuclease-free molecular biology grade water + 0.6 µL forward and reverse primers + 2 µL sample). The following primers were used for detecting microorganisms targeting a fragment of the 16S gene. (Table 2) The samples were carried out in duplicate on the Rotor Gene® with an initial denaturation step (95° C/3 min), followed by 35 cycles of denaturation (95 °C/30 s), primer annealing (specific for each microorganism, see Table 2), amplification (72 °C/30 s), and final extension (72 °C/3 min).

Table 2 Forward (F) and reverse (R) primers used to amplify a fragment of the 16S gene 

Microorganism Sequence Amplicon size (bp) Annealing
temperature (°C)
Bacteroides fragilis F: CAGTTCGCCATACAA
R: GGATTCTCTTTCCGCTTTGAC
131 57
Clostridium perfringens F: TGA AAC TGG GAG ACT TGA GTG C
R: CTT AGG TAA GGT TCT TCG CGT TGC
100 55
Enterococcus faecium F: GCATAGCCCGCACCTG
R: GTTACTCTCATCCTTGTTCTTCTC
160 54
Fusobacterium varium F: GGGATGTCAAACGCTGG
R: GGCGCTGAGGTTCGAG
143 57
Lactobacillus salivarius F: GTTCTCCTACGGCTACCTTGTTACG
R: TTCTCAGTTCGGATTGTAGGCTG
225 57
Universal F: ACTCCTACGGGAGGCAGCAG
R: ATTACCGCGGCTGG
136 60

The results were analyzed with the Qrex® program. The obtained cycle time (Ct) threshold values were considered acceptable in replicates that differed by a maximum of one cycle from each other. Relative abundance was determined (Microsoft® Excel) by the ΔΔCt method.18 The universal primer was used as reference Ct and the day zero sampling Ct as controls. We used the following formula to calculate the ΔΔCt:

ΔCt A = Ct of the target microorganism - Average reference Ct

ΔCt B = Day 0 Ct (control) of the target microorganism - Average reference Ct

ΔΔCt = Average ΔCt B - ΔCt A

The ΔΔCt  value was then used to calculate the relative abundance (RA):

RA = 2-ΔΔCt

Statistical analysis

The experiment had a completely randomized design with two treatments, each one with ten replicates. The results of relative abundance were analyzed with the IBM® SPSS® Statistics program, version 23. Normality (Shapiro-Wilk test) and homogeneity of variances (Levene’s test) were evaluated19; by not complying with the assumptions for parametric tests, non-parametric tests were used, working with ranges for the results.

We compared the bacterial abundance between diets over time by using the Mann Whitney U test (rank-sum). The Kruskall Wallis test was used to evaluate the differences between the relative abundance of each microorganism. Finally, to determine wether there was a difference in the relative bacterial abundance of the diets over time, the Friedman test was used as an alternative to ANOVA for paired samples. A P-value (P < 0.05) was considered a statistically significant difference. All the original values (relative abundance) were transformed into ranks for analysis.

Results

Main rank of relative abundance between diets

Table 3 shows the difference in the main rank of relative abundance of each microorganism between diets over time. We observed a higher abundance of the microorganisms Fusobacterium varium and Enterococcus faecium in the first 15 days of consumption and Lactobacillus salivarius and Bifidobacterium animalis after 30 days of consumption of the HD diet.

Table 3 Main rank of relative abundance of microorganisms in the feces of dogs* 

Microorganism Day 15 P-value Day 30 P-value
1Diet 1Diet
HD LD HD LD
C. perfringens 12.3
(4-20)
8.7
(2-16)
0.1900 10.3
(6-18)
10.7
(2-20)
0.9120
F. varium 14.7a
(10-20)
6.3b
(3-14)
0.0010 13.1
(8-20)
7.9
(2-16)
0.0520
L. salivarius 15.1a
(10-20)
5.9b
(2-12)
< 0.0001 13.4a
(8-20)
7.6b
(2-16)
0.0280
E. faecium 14.7a
(10-20)
6.3b
(3-14)
0.0010 11.5
(4-20)
9.5
(2-18)
0.4810
B. fragilis 13.9a
(8-20)
7.1b
(4-16)
0.0090 13.5a
(10-18)
7.5b
(2-20)
0.0230

* Diets at 15 and 30 days post administration (n = 10)

1 HD: high quality and digestibility; LD: low quality and digestibility

a,b Different superscripts within the same row denote a statistically significant difference (P < 0.05) according to a Mann Whitney U test.

Main rank of relative abundance of populations

The main rank of relative abundances of the populations of microorganisms did not change significantly during the study period per diet offered (Figure 2).

Figure 2 Main rank (range) of relative abundance of bacterial species under study in the feces of dogs (n = 10) fed either high (HD) or low (LD) protein and digestibility diets at 15 and 30 days post-administration. 

Main rank of relative abundance between days 15 and 30

Fecal samples from animals on the HD diet had higher overall bacterial relative abundance (P = 0.025) on day 15 than on day 30 post administration: main rank 2 (range 1-2) and 1 (range 1-2), respectively. Conversely, overall bacterial relative abundance was similar (P = 0.180) at 15 and 30 days post administration in the feces of dogs fed the LD diet: main ranks 1.2 (range 1-2) and 1.8 (range 1-2) respectively.

Discussion

Current nutrition in dogs seeks to meet the physiological requirements of the animals and generate greater intestinal integrity and health, which is reflected in the health of the individual, ranging from the immune system to possible repercussions on mental health.20 It is becoming increasingly evident how diet affects the intestinal microbiota, not only within an individual, since it differs between individuals.3

Given these facts, it is evident that in many cases it has not been possible to establish clear patterns for an ideal dietary profile. Still, it allows us to think of tailor-made or individual diet profiles that lead to this “healthy gut” balance. Several studies on microbiome, have established that the intestinal anatomophysiology differs between species or breeds as there is an evident individual variability.21 This situation is usually reflected in the results that show that even with changes in the diet, the microbiome tends to return to a balance.3

This “dietary” balance depends on multiple factors such as the use of antibiotics, stress, infectious processes, and chronic diseases. When the individual homeostasis is broken, generating alterations such as colitis, malabsorption syndromes, or persistent diarrhea can lead to the individual´s death.10 The microorganism associated with this type of digestive disorder is Clostridium perfringens. Experimental evidence6,12,13 has shown that environments rich in protein and with high amounts of amino acids promote the growth of both Clostridium difficile and Clostridium perfringens, microorganisms related to inflammatory processes.22 For instance, a study reported an increase in the concentration of C. perfringens (from 3.3 to > 8 logCFU/g feces) in dogs under high-protein diets (> 40 % CP) with low quality and digestibility.22

Despite expecting a similar change, in this study, it was impossible to associate the protein digestibility of the diet with an increase in the relative abundance of C. perfringens. Thus, it appears that even if the diet has low digestibility and allows a colonic environment rich in amino acids, the microbiome can stabilize the growth of possible pathogens in response to the presence of other nutrients or intrinsic factors of a healthy dog.23 In our study, the stool consistency was not evaluated, nor was the diet evaluated for more than 30 days.

The results of Fusobacterium varium obtained in our study are similar to those reported by Mori,24 who compared the microbiome of healthy dogs fed with four prescription diets. In that study, high relative abundance of Fusobacterium was found when the animals consumed a diet with the highest level of crude protein (30 %) (abundances: 7.6 vs. 2.9, 1.4, and 0). Here, the higher crude protein diet (33.8 vs. 20.1 %) resulted in a higher relative abundance of Fusobacterium (14.7 vs. 6.3). However this effect was only significant for the first 15 days.

Moinard also reported an increase in Fusobacterium when the diet was high in protein (29.3 % CP).25Fusobacterium spp. is considered an amino acid fermenter, which is sometimes related to the presentation of digestive disorders. Hang13 observed a high relative abundance of this type when a very high protein diet (60 % CP) was administered. In that study, diarrhea and changes in the consistency of the feces were also reported. Nevertheless, the tendency of Fusobacterium spp. to increase its relative abundance in high protein diets could also have a beneficial effect as this microorganism produces butyrate.6 This volatile fatty acid is used as an energy source for the intestine and is associated with antineoplastic properties.

The genus Enterococcus presents great ecological diversity. While Enterococcus faecium is a microorganism widely distributed in nature, some strains are considered causative of nosocomial diseases in humans and zoonotic.26 Conversely, other strains have a probiotic potential effect.27,28 The identification and quantification of this microorganism has been documented in both healthy and diseased dogs.29,30

In this study it was possible to identify a higher relative abundance of E. faecium in the high protein and high-quality diets (14.7 vs 6.3), the associations between the type of diet or substrates with the growth of this microorganism are few. However, according to previous studies, it has a high capacity to use carbohydrates.28 In the case of the HD diet, it presented a lower concentration of carbohydrates than the LD (42 vs 53 %), which could explain the reported abundance of this microorganism. It is essential to consider that the primers used for the molecular identification of E. faecium¸ did not contemplate any particular strain. Therefore, it would be interesting to conduct further studies to determine wether the abundance associated with the diet belongs to potential zoonotic or innocuous strains.

Regarding Lactobacillus salivarius, its relative abundance remained high in the feces of dogs on HD diet during the 30 days of experimentation. These findings are consistent with observations of Middelboss,31 who found a higher relative abundance of this microorganism in the feces of dogs fed a diet supplemented with prebiotic fiber compared with controls fed unsupplemented feed (12.2 vs. 10). This effect is associated with the fact that Lactobacillus genera are microorganisms associated with the presence of functional fibers such as inulin, fructooligosaccharides, mannanoligosaccharides and beet pulp.

High quality and highly digestible feeds contain inulin among its ingredients, which could explain the behavior of this microorganism. In turn, Bermingham et al., 2017 dietinduced changes in faecal microbiota observed in humans and rodents have been extrapolated to pets in spite of their very different dietary and metabolic requirements. This lack of direct evidence means that the mechanisms by which microbiota influences health in dogs are poorly understood. We hypothesised that changes in faecal microbiota correlate with physiological parameters including apparent macronutrient digestibility. Methods. Fifteen adult dogs were assigned to two diet groups, exclusively fed either a premium kibbled diet (kibble; K; n = 8, also reported a correlation between the level of digestibility of the diet, and the growth of Lactobacillary, similar to the results of this.32

Bacteroides is a gender in which important discrepancies have been detected. While some authors have reported high concentrations in patients with diarrhea,33,34 others have reported it in lower concentrations when the patient presents with inflammatory bowel disease.35 These possible differences are associated with issues such as the extraction technique used, the type of sample, species (dog or cat), as well as the type of diet.(34) Regarding this last factor, it has been possible to associate Bacteroides with diets high in protein and fat.24,36 Deusch et al.37 found that cats fed a high-protein, low-carbohydrate diet (110 g crude protein and 49 g crude fat/1 000 kcal ME) had a higher relative abundance of this microorganism compared to those fed moderate protein and carbohydrate diets (78 g crude protein and 44 g crude fat/1 000 kcal). This result is similar to that obtained in this experimental, considering that the high-quality diet has a profile of 91.2 g crude protein and 40.15 g crude fat/1 000 kcal. Swanson34 also found a relationship between the type of fibers contained in the feed and growth of species belonging to the genre Bacteroides. Similar to what was mentioned with L. salivarius, the high quality-digestibility diet contains inulin among its ingredients, a component that could favor these two microorganisms during the 30 days of the experimental period.

Finally, it is possible to observe that the general relative abundance of microorganisms was modified over time in the HD diet. This effect may be associated with the ability of the intestinal microbiome to adapt quickly and regulate itself to transient changes intrinsically.23 They offered two intercalated diets to Labrador dogs, and observed that after some time, the microorganisms recovered their initial stability or profile, with significant changes in certain bacteria, such as Fusobacterium, Bacteroides and Bifidobacterium. This effect has also been reported in humans,38 where evaluating the microbiome of 98 individuals fed with two types of diets (with different levels of fat and fiber), significant changes were detected during the first 24 h, and a stabilization of the population after 10 days.

Conclusions

This study demonstrated that the diet’s protein content, energy density, and digestibility significantly modify the relative abundance of some microorganisms commonly present in the dog’s fecal microbiome, except for the pathogen C. perfringens. Apparently, this effect does not persist over time for some of the bacterial species under study.

Data availability

The data sets used and analyzed in this experiment are available from the corresponding author on reasonable request.

Acknowledgments

The authors thank the support of Animal Nutrition and Biochemistry Department, Faculty of Veterinary Medicine, National Autonomous University of Mexico (UNAM).

Conflicts of interest

The authors have declared no conflict of interest.

Author contributions

Conceptualization: L Gutiérrez, C Gutiérrez

Data curation: KE Cosío

Formal analysis: MG Sánchez

Funding acquisition: L Gutiérrez, C Gutiérrez

Investigation: KE Cosío, CC Márquez

Methodology: CC Márquez

Writing original draft: KE Cosío

Writing-review and editing: KE Cosío, L Gutiérrez, C Gutiérrez, CC Márquez, ME Ortega

Funding statement

This study was funded by UNAM’s Support Program for Research and Technological Innovation Projects (PAPIIT) through award number IT200222. Mexico’s National Council of Science and Technology (CONACYT) also awarded fellowship 282603 to author Karina Cosío for conducting her doctorate at UNAM.

References

1. Arboleda M. Estudio de la oferta y la demanda de productos de mascotas relacionados con la alimentación y los accesorios en la ciudad de Medellín para determinar oportunidades de innovación y desarrollo de nuevos productos (Tesis de maestría). Medellín, Colombia. Universidad EAFIT; 2017. [ Links ]

2. Michel KE. Unconventional diets for dogs and cats. Veterinary Clinics of North America: Small Animal Practice. 2006;36(6):1269-81. doi: 10.1016/j.cvsm.2006.08.003. [ Links ]

3. Herstad, KMV, Gajardo, K, Bakke, AM, Moe, L, Ludvigsen, J, Rudi, K, et al. A diet change from dry food to beef induces reversible changes on the faecal microbiota in healthy, adult client-owned dogs. BMC Vet Res. 2017;13(1):1-13. doi: 10.1186/s12917-017-1073-9. [ Links ]

4. Wernimont, SM, Radosevich, J, Jackson, MI, Ephraim, E, Badri, DV, Macleay, JM, et al. The effects of nutrition on the gastrointestinal microbiome of cats and dogs: impact on health and disease. Frontiers in Microbiology. 2020;11(June):1-24. doi: 10.3389/fmicb.2020.01266. [ Links ]

5. Moon, CD, Young, W, Maclean, PH, Cookson, AL, Bermingham, EN. Metagenomic insights into the roles of Proteobacteria in the gastrointestinal microbiomes of healthy dogs and cats. Microbiology Open. 2018;7(5):1-20. doi: 10.1002/mbo3.677. [ Links ]

6. Vital, M, Gao, J, Rizzo, M, Harrison, T, Tiedje, JM. Diet is a major factor governing the fecal butyrate-producing community structure across Mammalia , Aves and Reptilia. The ISME Journal. 2015;832-43. doi: 10.1038/ismej.2014.179. [ Links ]

7. Xu, J, Verbrugghe, A, Lourenço, M, Cools, A, Liu, DJX, Wiele, T Van De, et al. The response of canine faecal microbiota to increased dietary protein is influenced by body condition. BMC Veterinary Research. 2017;1-11. doi: 10.1186/ s12917-017-1276-0. [ Links ]

8. Hooda, S, Minamoto, Y, Suchodolski, JS, Swanson, KS. Current state of knowledge : the canine gastrointestinal microbiome. Animal Health Research Reviews 2012;13(1):78-88. doi: 10.1017/S1466252312000059. [ Links ]

9. Lagkouvardos, Illias, Overmann, Jöng, Clavel, T. Cultured microbes represent a substantial fraction of the human and mouse gut microbiota. Gut Microbes. 2017;8(5):493-503. doi: 10.1080/19490976.2017.1320468. [ Links ]

10. Suchodolski JS. Companion animals symposium: Microbes and gastrointestinal health of dogs and cats. J Anim Sci. 2011;89(5):1520-30. doi: 10.2527/ jas.2010-3377. [ Links ]

11. Vanhoutte, T, Huys, G, De Brandt, E, Fahey, GCJ, Swings, J. Molecular monitoring and characterization of the faecal microbiota of healthy dogs during fructan supplementation. FEMS Microbiol Lett. 2005 Aug;249(1):65-71. doi: 10.1016/j.femsle.2005.06.003. [ Links ]

12. Sandri, M, Dal Monego, S, Conte, G, Sgorlon, S, Stefanon, B. Raw meat based diet influences faecal microbiome and end products of fermentation in healthy dogs. BMC Vet Res. 2017;13(1):1-11. doi: 10.1016/j.femsle.2005.06.003. [ Links ]

13. Hang, I, Rinttila, T, Zentek, J, Kettunen, A, Alaja, S, Apajalahti, J, et al. Effect of high contents of dietary animal-derived protein or carbohydrates on canine faecal microbiota. BMC Veterinary Research . 2012; http://www.biomedcentral.com/1746-6148/8/90Links ]

14. Middelbos, IS, Boler, BMV, Qu, A, White, BA, Swanson, KS, Fahey, GC. Phylogenetic characterization of fecal microbial communities of dogs fed diets with or without supplemental dietary fiber using 454 pyrosequencing. PLoS One. 2010;5(3). doi: 10.1371/journal.pone.0009768. [ Links ]

15. Case, LP, Leighann, D, Hayek, MG, Raasch, M. Canine and Feline Nutrition: A Resource for Companion Animal Professionals. 3rd. St. Louis Missouri. USA: Elesevier; 2010. 576 p. (Elsevier Health Sciences). [ Links ]

16. Wilson K. Preparation of genomic DNA from bacteria. Curr Protoc Mol Biol [internet]. 2001;56(1):2.4.1-2.4.5. doi: 10.1002/0471142727.mb0204s56. [ Links ]

17. García A. Implementación de técnicas de extracción de ADN a partir de líquido ruminal para ladetección e identificación de microorganismos mediante PCR. México: UNAM; 2019. [ Links ]

18. Green MR. Molecular Cloning: A Laboratory Manual. 4th ed. New York: Cold Spring Harbor Laboratory Press; 2012. 2 028 p. [ Links ]

19. Daniel W. Bioestadística: base para el análisis de las ciencias de la salud. 4th. ed. Buenos Aires: Limusa-Wiley; 2010. 634 p. [ Links ]

20. Pilla, R, Suchodolski, JS. The role of the canine gut microbiome and metabolome in health and gastrointestinal disease. Front Vet Sci. 2020;6(January):1-12. doi: 10.3389/fvets.2019.00498. [ Links ]

21. Suchodolski JS. Assessment of the qualitative variation in bacterial microflora among compartments of the intestinal tract of dogs by use of a molecular fingerprinting technique. Am J Vet Res. 2005;66:1555-62. [ Links ]

22. Zentek, J, Marquart, B, Pietrzak, T, Ballèvre, O, Rochat, F. Dietary effects on bifidobacteria and Clostridium perfringens in the canine intestinal tract. J Anim Physiol Anim Nutr (Berl). 2003;87(11-12):397-407. doi: 10.1046/j.0931-2439.2003.00451.x. [ Links ]

23. Allaway, D, Haydock, R, Lonsdale, ZN, Deusch, OD, O’Flynn, C, Hughes, KR. Rapid reconstitution of the fecal microbiome after extended diet-induced changes indicates a stable gut microbiome in healthy adult dogs. Appl Environ Microbiol. 2020;86(13):1-13. doi: 10.1128/AEM.00562-20. [ Links ]

24. Mori, A, Goto, A, Kibe, R, Oda, H, Kataoka, Y. Comparison of the effects of four commercially available prescription diet regimens on the fecal microbiome in healthy dogs. J. Vet. Med. Sci. 2019;1783-1790. doi: 10.1292/jvms.19-0055. [ Links ]

25. Moinard, A, Payen, C, Ouguerram, K, André, A, Hernandez, J, Nguyen, P, et al. Effects of high-fat diet at two energetic levels on fecal microbiota, colonic barrier and, metabolic parameters in dogs. Frontiers in Veterinary Science 2020;7(September):1-15. doi: 10.3389/fvets.2020.566282. [ Links ]

26. Santaniello, A, Sansone, M, Fioretti, A, Menna, LF. Systematic review and meta-analysis of the occurrence of ESKAPE bacteria group in dogs, and the related zoonotic risk in animal-assisted therapy, and in animal-assisted activity in the health context. International Journal of Environmental Research and Public Health. Review 2020. doi: 10.3390/ijerph17093278. [ Links ]

27. Lam, MMC, Seemann, T, Bulach, DM, Gladman, SL, Chen, H, Haring, V, et al. Comparative analysis of the first complete Enterococcus faecium. Journal of Bacteriology. 2012;2334-41. doi: 10.1128/JB.00259-12. [ Links ]

28. Ghattargi, VC, Gaikwad, MA, Meti, BS, Nimonkar, YS, Dixit, K, Prakash, O, et al. Comparative genome analysis reveals key genetic factors associated with probiotic property in Enterococcus faecium strains. BMC Genomics 2018;1-16. doi: 10.1186/s12864-018-5043-9. [ Links ]

29. Zishiri O. Prevalence of virulence genes in Enterococcus species isolated from companion animals and livestock. Onderstepoort Journal of Veterinary Research 2018;1-8. doi: 10.4102/ojvr.v85i1.1583. [ Links ]

30. Jha, AR, Shmalberg, J, Tanprasertsuk, J, Perry, LA, Massey, D, Honaker, RW. Characterization of gut microbiomes of household pets in the United States using a direct-to-consumer approach. PLoS One. 2020;15(2):1-20. doi: 10.1371/journal.pone.0227289. [ Links ]

31. Middelbos, IS, Fastinger, ND, Fahey, GC. Evaluation of fermentable oligosaccharides in diets fed to dogs in comparison to fiber standards. J Anim Sci. 2007;85(11):3033-3044. doi: 10.2527/jas.2007-0080. [ Links ]

32. Bermingham, EN, Maclean, P, Thomas, DG, Cave, NJ, Young, W. Key bacterial families (Clostridiaceae, Erysipelotrichaceae and Bacteroidaceae) are related to the digestion of protein and energy in dogs. PeerJ. 2017;2017(3). doi: 10.7717/peerj.3019. [ Links ]

33. Jia, J, Frantz, N, Khoo, C, Gibson, GR, Rastall, RA, Mccartney, AL. Investigation of the faecal microbiota associated with canine chronic diarrhoea. FEMS. 2010;71:30412. doi: 10.1111/j.1574-6941.2009.00812.x. [ Links ]

34. Deng, P, Swanson, KS. Gut microbiota of humans , dogs and cats : current knowledge and future opportunities and challenges. British Journal of Nutrition. 2015; doi: 10.1017/S0007114514002943. [ Links ]

35. Inness, VL, Mccartney, AL, Khoo, C, Gross, KL, Gibson, GR. Molecular characterisation of the gut microflora of healthy and inflammatory bowel disease cats using fluorescence in situ hybridisation with special reference to Desulfovibrio spp. Journal of Animal Physiology and Animal Nutrition. 2007;91:48-53. doi: 10.1111/j.1439-0396.2006.00640.x. [ Links ]

36. Bermingham, EN, Young, W, Kittelmann, S, Kerr, KR, Swanson, KS, Roy, NC, et al. Dietary format alters fecal bacterial populations in the domestic cat (Felis catus). Microbiology Open. 2013;173-81. doi: 10.1002/mbo3.60. [ Links ]

37. Deusch, O, Flynn, CO, Colyer, A, Morris, P, Allaway, D, Jones, PG, et al. Deep Illumina-based shotgun sequencing reveals dietary effects on the structure and function of the fecal microbiome of growing kittens. Plos One. 2014;9(7). doi: 10.1371/journal.pone.0101021. [ Links ]

38. Wu, GD, Chen, J, Hoffmann, C, Bittinger, K, Chen, YY, Keilbaugh, SA, et al. Linking long-term dietary patterns with gut microbial enterotypes. Science (80). 2011;334(6052):105-8. doi: 10.1126/science.1208344. [ Links ]

Cite this as:

Cosío-Carpintero KE, Gutiérrez Olvera C, Márquez Mota CC, Ortega-Cerrilla ME, Sánchez-González MG, Gutiérrez Olvera L. High levels of dietary digestible protein transiently promote beneficial bacteria in adult dog feces. Veterinaria México OA. 2022;9. doi: 10.22201/fmvz.24486760e.2022.985.

Received: August 16, 2021; Accepted: June 09, 2022; Published: December 01, 2022

*Corresponding author: Email address: liliago@unam.mx

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License