Oxalobacter formigenes

Last updated

Oxalobacter formigenes
Scientific classification OOjs UI icon edit-ltr.svg
Domain: Bacteria
Phylum: Pseudomonadota
Class: Betaproteobacteria
Order: Burkholderiales
Family: Oxalobacteraceae
Genus: Oxalobacter
Species:
O. formigenes
Binomial name
Oxalobacter formigenes
Allison et al, 1985 [1]
Type strain
Oxalobacter formigenes OxBT

Oxalobacter formigenes is a Gram negative oxalate-degrading anaerobic bacterium that was first isolated from the gastrointestinal tract of a sheep in 1985. [1] To date, the bacterium has been found to colonize the large intestines of numerous vertebrates, including humans, and has even been isolated from freshwater sediment. [2] It processes oxalate by decarboxylation into formate (oxalyl-CoA decarboxylase), producing energy for itself in the process. [3]

Contents

The broad-spectrum quinolone antibiotics kill O. formigenes.[ citation needed ] If a person's gastrointestinal (GI) tract lacks this bacterium, and therefore lacks the primary source of the oxalyl-CoA decarboxylase enzyme, then the GI tract cannot degrade dietary oxalates; after some vitamin B6-modulated partial metabolic degradation in the body, the oxalates are excreted in the kidney, where they precipitates to form calcium oxalate kidney stones. [4] [5] [6] [7] Oxalobacter formigenes can protect against kidney stones by degrading oxalate. [7]

The role and presence of O. formigenes in the human gut is an area of active research.

Genome

Genomic information
Genome size 2.41-2.47 Mb [8] [9] [10]

The genome of O. formigenes has been sequenced by at least three different researchers. It has a G+C content of 49.6%. [9] [11]

Taxonomy

Based on fatty acid profile, 16S ribosomal RNA sequencing, and DNA probes specific to the oxc (oxalyl-CoA decarboxylase) gene and frc (formyl-CoA transferase), O. formigenes has been divided into two groups. [1] [12] [13] [14] Group 1 has less diversity and better growth compared to group 2. To date, most research has focused on group 1 strains due to their ease of growth.

Interestingly, analysis with the DNA probes showed that group 2 may be further divided into two subgroups. [13] Whole genome sequencing has revealed that the original O. formigenes taxon can be divided into three additional species: Oxalobacter aliiformigenes , Oxalobacter paeniformigenes , and Oxalobacter paraformigenes . [11]

Metabolism

O. formigenes uses oxalate as its primary carbon source. [1] Oxalate is absorbed through an oxalate:formate antiporter (OxlT) in a 1:1 proportion. [15] Imported oxalate is then converted to oxalyl-CoA via formyl-CoA transferase (frc). Oxalyl-CoA is decarboxylated using and H+ via oxalyl-CoA decarboxylase (oxc), releasing CO2, and generating formyl-CoA, which is used for the frc reaction. In total, approximately 1 mol of formate and CO2 are produced per mol of oxalate consumed. [16] 3H+ are imported via an ATPase to provide H+ for the decarboxylation reaction. [17]

Cell biomass generation

Biomass in O. formigenes is primarily generated by oxalate consumption through the metabolism of oxalyl-CoA in the glycerate pathway. [18] [19] Acetate and carbonate are also used for cell biomass, but to a lesser extent than oxalate. [18]

Growth in culture

O. formigenes was isolated in oxalate containing anerobic media. [1] Currently, O. formigenes is grown in anaerobic Hungate tubes using a CO2-bicarbonate buffered oxalate media. [2] Optimal growth is achieved at a pH between 6 and 7. Oxalate is used at 20 mM for freezer recovery and general maintenance but concentrations can be increased to 100 mM for increased cell density. While oxalate is the main carbon source, small amounts of acetate and yeast extract are supportive of growth. [2] [16] O. formigenes can reach stationary phase in approximately 24 – 48 hours but is sometimes delayed to 72 hours.

Enriched anaerobic complex media (e.g. Brain heart infusion) fail to support the growth of O. formigenes unless supplemented with oxalate. Therefore, these media can be used to assess the purity of O. formigenes cultures.

Antibiotic resistance and susceptibility

Given the fastidious nature of O. formigenes, traditional methods for antibiotic susceptibility testing are not sufficient. Instead, bacteria are cultured in the presence of antibiotics and screened for viability using opaque anaerobic oxalate agar. [2] [20] [21] This method demonstrated that O. formigenes is resistant to nalidixic acid, ampicillin, amoxicillin, streptomycin, and vancomycin. [20] [21] O. formigenes was also found to be susceptible to ciprofloxacin, clarithromycin, clindamycin, doxycycline, gentamicin, levofloxacin, metronidazole, and tetracycline. [20] [21]

Prevalence in the mammalian gut

O. formigenes is found in the mammalian gastrointestinal tract and often isolated from feces. In addition to culture-based methods, O. formigenes is presence is detected using molecular methods such as qPCR and next generation sequencing.

Humans

Humans are not typically born with O. formigenes and only become colonized when they begin crawling around in their environment. [22] In adulthood, the frequency of O. formigenes in the gut microbiota varies across different populations. In North India, O. formigenes is prevalent in approximately 65% of the population. [23] In South Korea and Japan, O. formigenes is present in about 75% of individuals. [24] [25] In the United States of America, O. formigenes is only detected in about 30% of the human population. [26] [27] Populations who do not practice modern medicine or life in a Western lifestyle typically have an increased prevalence of O. formigenes, which could imply that these practices affect O. formigenes colonization. [28] [29]

Ruminants

The idea that ruminants are colonized by oxalate-degrading bacteria came from the observation that sheep grazing on oxalate-rich plants (e.g. Halogeton glomeratus ) consumed large quantities of this plant and died of renal intoxication from oxalate. [2] However, by slowly acclimatizing sheep to high-oxalate intake, they would survive the consumption of large quantities of oxalate-rich plants. [30] This led to the proposal that resident oxalate-degrading bacteria were enriched by the gradual introduction to a oxalate-rich diet, which protected the sheep from oxalate-induced renal damage. [31] [32] In 1980, the first oxalate-degrading bacteria were isolated from the rumen of sheep, and it was later named Oxalobacter formigenes. [1] [16]

Potential role in kidney stone disease

O. formigenes has been investigated for its role in mitigating calcium oxalate kidney stone disease because it metabolizes oxalate as its primary carbon source.

Oxalate degradation

In vitro experiments find that O. formigenes is a specialist oxalate consuming bacteria that can degrade oxalate more efficiently than other generalist oxalate consuming bacteria. [33] Initial research pointed to the loss of oxalate-degrading bacteria, such as O. formigenes, following antibiotic usage as primary contributor to calcium oxalate kidney stone disease. [34] [35] Colonization with O. formigenes has been observed to results in a decrease in urinary oxalate [35] [4] and reduced frequency of kidney stones [4] [7] [36]

In a small study, oral supplementation with O. formigenes HC-1 along with a loading dose of oxalate resulted in reduced oxalate excretion during the 6 h immediately following ingestion. [20] However, clinical trials have been unsuccessful in establishing O. formigenes and reducing urinary and plasma concentrations of oxalate. [37] [38] [39]

Recent work using next-generation sequencing has found that O. formigenes colonizes both calcium oxalate kidney stone formers and non-stone forming controls. [40] [41] This observation has led to the notion that O. formigenes alone may not be responsible for regulating oxalate degradation in the gut microbiota, but instead it may be part of a network of co-occurring bacterial taxa that modulate oxalate degradation together. [42] [43] [44]

Secretagogues to promote intestinal oxalate dumping

It has been proposed that O. formigenes produces secretagogues that can stimulate oxalate transport in epithelial cells. While epithelial oxalate secretion has been shown in human cell lines and rodent models, [45] [46] it has not been confirmed in humans. Candidate bioactive molecules have been identified and tested in animal models. [45] [47]


Related Research Articles

<span class="mw-page-title-main">Kidney stone disease</span> Formation of mineral stones in the urinary tract

Kidney stone disease, also known as renal calculus disease, nephrolithiasis or urolithiasis, is a crystallopathy where a solid piece of material develops in the urinary tract. Renal calculi typically form in the kidney and leave the body in the urine stream. A small calculus may pass without causing symptoms. If a stone grows to more than 5 millimeters, it can cause blockage of the ureter, resulting in sharp and severe pain in the lower back that often radiates downward to the groin. A calculus may also result in blood in the urine, vomiting, or painful urination. About half of people who have had a renal calculus are likely to have another within ten years.

<span class="mw-page-title-main">Oxalic acid</span> Simplest dicarboxylic acid

Oxalic acid is an organic acid with the systematic name ethanedioic acid and chemical formula HO−C(=O)−C(=O)−OH, also written as (COOH)2 or (CO2H)2 or H2C2O4. It is the simplest dicarboxylic acid. It is a white crystalline solid that forms a colorless solution in water. Its name comes from the fact that early investigators isolated oxalic acid from flowering plants of the genus Oxalis, commonly known as wood-sorrels. It occurs naturally in many foods. Excessive ingestion of oxalic acid or prolonged skin contact can be dangerous.

<span class="mw-page-title-main">Oxalate</span> Any derivative of oxalic acid; chemical compound containing oxalate moiety

Oxalate is an anion with the chemical formula formula C2O2−4. This dianion is colorless. It occurs naturally, including in some foods. It forms a variety of salts, for example sodium oxalate, and several esters such as dimethyl oxalate. It is a conjugate base of oxalic acid. At neutral pH in aqueous solution, oxalic acid converts completely to oxalate.

Lactiplantibacillus plantarum is a widespread member of the genus Lactiplantibacillus and commonly found in many fermented food products as well as anaerobic plant matter. L. plantarum was first isolated from saliva. Based on its ability to temporarily persist in plants, the insect intestine and in the intestinal tract of vertebrate animals, it was designated as a nomadic organism. L. plantarum is Gram positive, bacilli shaped bacterium. L. plantarum cells are rods with rounded ends, straight, generally 0.9–1.2 μm wide and 3–8 μm long, occurring singly, in pairs or in short chains. L. plantarum has one of the largest genomes known among the lactic acid bacteria and is a very flexible and versatile species. It is estimated to grow between pH 3.4 and 8.8. Lactiplantibacillus plantarum can grow in the temperature range 12 °C to 40 °C. The viable counts of the "L. plantarum" stored at refrigerated condition (4 °C) remained high, while a considerable reduction in the counts was observed stored at room temperature.

<i>Proteus mirabilis</i> Species of bacterium

Proteus mirabilis is a Gram-negative, facultatively anaerobic, rod-shaped bacterium. It shows swarming motility and urease activity. P. mirabilis causes 90% of all Proteus infections in humans. It is widely distributed in soil and water. Proteus mirabilis can migrate across the surface of solid media or devices using a type of cooperative group motility called swarming. Proteus mirabilis is most frequently associated with infections of the urinary tract, especially in complicated or catheter-associated urinary tract infections.

Ferroglobus is a genus of the Archaeoglobaceae.

In biology, syntrophy, syntrophism, or cross-feeding is the cooperative interaction between at least two microbial species to degrade a single substrate. This type of biological interaction typically involves the transfer of one or more metabolic intermediates between two or more metabolically diverse microbial species living in close proximity to each other. Thus, syntrophy can be considered an obligatory interdependency and a mutualistic metabolism between different microbial species, wherein the growth of one partner depends on the nutrients, growth factors, or substrates provided by the other(s).

<i>Aeromonas hydrophila</i> Species of heterotrophic, Gram-negative, bacterium

Aeromonas hydrophila is a heterotrophic, Gram-negative, rod-shaped bacterium mainly found in areas with a warm climate. This bacterium can be found in fresh or brackish water. It can survive in aerobic and anaerobic environments, and can digest materials such as gelatin and hemoglobin. A. hydrophila was isolated from humans and animals in the 1950s. It is the best known of the species of Aeromonas. It is resistant to most common antibiotics and cold temperatures and is oxidase- and indole-positive. Aeromonas hydrophila also has a symbiotic relationship as gut flora inside of certain leeches, such as Hirudo medicinalis.

Dehalococcoides is a genus of bacteria within class Dehalococcoidia that obtain energy via the oxidation of hydrogen and subsequent reductive dehalogenation of halogenated organic compounds in a mode of anaerobic respiration called organohalide respiration. They are well known for their great potential to remediate halogenated ethenes and aromatics. They are the only bacteria known to transform highly chlorinated dioxins, PCBs. In addition, they are the only known bacteria to transform tetrachloroethene to ethene.

<i>Pseudomonas stutzeri</i> Species of bacterium

Pseudomonas stutzeri is a Gram-negative soil bacterium that is motile, has a single polar flagellum, and is classified as bacillus, or rod-shaped. While this bacterium was first isolated from human spinal fluid, it has since been found in many different environments due to its various characteristics and metabolic capabilities. P. stutzeri is an opportunistic pathogen in clinical settings, although infections are rare. Based on 16S rRNA analysis, this bacterium has been placed in the P. stutzeri group, to which it lends its name.

In enzymology, a formyl-CoA transferase is an enzyme that catalyzes the chemical reaction

<span class="mw-page-title-main">Oxalyl-CoA decarboxylase</span>

The enzyme oxalyl-CoA decarboxylase (OXC) (EC 4.1.1.8), primarily produced by the gastrointestinal bacterium Oxalobacter formigenes, catalyzes the chemical reaction

Faecalibacterium is a genus of bacteria. The genus contains several species including Faecalibacterium prausnitzii, Faecalibacterium butyricigenerans, Faecalibacterium longum, Faecalibacterium duncaniae, Faecalibacterium hattorii, and Faecalibacterium gallinarum. Its first known species, Faecalibacterium prausnitzii is gram-positive, mesophilic, rod-shaped, and anaerobic, and is one of the most abundant and important commensal bacteria of the human gut microbiota. It is non-spore forming and non-motile. These bacteria produce butyrate and other short-chain fatty acids through the fermentation of dietary fiber. The production of butyrate makes them an important member of the gut microbiota, fighting against inflammation.

<span class="mw-page-title-main">Zetaproteobacteria</span> Class of bacteria

The class Zetaproteobacteria is the sixth and most recently described class of the Pseudomonadota. Zetaproteobacteria can also refer to the group of organisms assigned to this class. The Zetaproteobacteria were originally represented by a single described species, Mariprofundus ferrooxydans, which is an iron-oxidizing neutrophilic chemolithoautotroph originally isolated from Kamaʻehuakanaloa Seamount in 1996 (post-eruption). Molecular cloning techniques focusing on the small subunit ribosomal RNA gene have also been used to identify a more diverse majority of the Zetaproteobacteria that have as yet been unculturable.

Desulfitobacterium hafniense is a species of gram positive bacteria, its type strain is DCB-2T..

<i>Akkermansia muciniphila</i> Species of bacterium

Akkermansia muciniphila is a human intestinal symbiont, isolated from human feces. It is a mucin-degrading bacterium belonging to the genus, Akkermansia, discovered in 2004 by Muriel Derrien and Willem de Vos at Wageningen University of the Netherlands. It belongs to the phylum Verrucomicrobiota and its type strain is MucT. It is under preliminary research for its potential association with metabolic disorders.

Symbiobacterium thermophilum is a symbiotic thermophile that depends on co-culture with a Bacillus strain for growth. It is Gram-negative and tryptophanase-positive, with type strain T(T). It is the type species of its genus. Symbiobacterium is related to the Gram-positive Bacillota and Actinomycetota, but belongs to a lineage that is distinct from both.S. thermophilum has a bacillus shaped cell structure with no flagella. This bacterium is located throughout the environment in soils and fertilizers.

Oxalobacter aliiformigenes is a Gram negative, non-spore-forming, oxalate-degrading anaerobic bacterium that was first isolated from human fecal samples. O. aliiformigenes consumes oxalate as its main carbon source but is negative for indole production and negative for sulfate and nitrate reduction. Cells appear rod shaped, though occasionally present as curved, and do not possess flagella.

Oxalobacter paraformigenes is a Gram negative, non-spore-forming, oxalate-degrading anaerobic bacterium that was first isolated from human fecal samples. O. paraformigenes may have a role in calcium oxalate kidney stone disease because of its unique ability to utilize oxalate as its primary carbon source.

Oxalobacter paeniformigenes is a Gram negative, non-spore-forming, oxalate-degrading anaerobic bacterium that was first isolated from human fecal samples. Similar to other species in the Oxalobacter genus, O. paeniformigenes uses oxalate as its primary carbon source. O. paeniformigenes is negative for indole production and negative for sulfate and nitrate reduction. Cells appear rod shaped, though occasionally present as curved, and do not possess flagella.

References

  1. 1 2 3 4 5 6 Allison MJ, Dawson KA, Mayberry WR, Foss JG (February 1985). "Oxalobacter formigenes gen. nov., sp. nov.: oxalate-degrading anaerobes that inhabit the gastrointestinal tract". Archives of Microbiology. 141 (1): 1–7. Bibcode:1985ArMic.141....1A. doi:10.1007/BF00446731. PMID   3994481. S2CID   10709172.
  2. 1 2 3 4 5 Daniel SL, Moradi L, Paiste H, Wood KD, Assimos DG, Holmes RP, et al. (August 2021). Pettinari JM (ed.). "Forty Years of Oxalobacter formigenes, a Gutsy Oxalate-Degrading Specialist". Applied and Environmental Microbiology. 87 (18): e0054421. Bibcode:2021ApEnM..87E.544D. doi:10.1128/AEM.00544-21. PMC   8388816 . PMID   34190610.
  3. Unden G (2013). "Energy Transduction in Anaerobic Bacteria". Encyclopedia of Biological Chemistry. pp. 204–209. doi:10.1016/B978-0-12-378630-2.00282-6. ISBN   978-0-12-378631-9.
  4. 1 2 3 Troxel SA, Sidhu H, Kaul P, Low RK (April 2003). "Intestinal Oxalobacter formigenes colonization in calcium oxalate stone formers and its relation to urinary oxalate". Journal of Endourology. 17 (3): 173–176. doi:10.1089/089277903321618743. PMID   12803990.
  5. Tunuguntla HS (2001). "Can the recurrence of oxalate stones be prevented? Role of Oxalobacter formigenes in stone recurrence". Journal of Urology. 165: S246.
  6. Pearle MS, Goldfarb DS, Assimos DG, Curhan G, Denu-Ciocca CJ, Matlaga BR, et al. (August 2014). "Medical management of kidney stones: AUA guideline". The Journal of Urology. 192 (2): 316–324. doi:10.1016/j.juro.2014.05.006. PMID   24857648. S2CID   206623478.
  7. 1 2 3 Siener R, Bangen U, Sidhu H, Hönow R, von Unruh G, Hesse A (June 2013). "The role of Oxalobacter formigenes colonization in calcium oxalate stone disease". Kidney International. 83 (6): 1144–1149. doi: 10.1038/ki.2013.104 . PMID   23536130.
  8. Sun NY, Gao Y, Yu HJ (October 2019). Stewart FJ (ed.). "Genome Sequence of Oxalobacter formigenes Strain SSYG-15". Microbiology Resource Announcements. 8 (42): e01059–19. doi:10.1128/MRA.01059-19. PMC   6797538 . PMID   31624173.
  9. 1 2 Hatch M, Allison MJ, Yu F, Farmerie W (July 2017). "Genome Sequence of Oxalobacter formigenes Strain OXCC13". Genome Announcements. 5 (28): e00534–17. doi:10.1128/genomeA.00534-17. PMC   5511905 . PMID   28705966.
  10. Hatch M, Allison MJ, Yu F, Farmerie W (July 2017). "Genome Sequence of Oxalobacter formigenes Strain HC-1". Genome Announcements. 5 (27): e00533–17. doi:10.1128/genomeA.00533-17. PMC   5502849 . PMID   28684568.
  11. 1 2 Chmiel JA, Carr C, Stuivenberg GA, Venema R, Chanyi RM, Al KF, et al. (2022-12-21). "New perspectives on an old grouping: The genomic and phenotypic variability of Oxalobacter formigenes and the implications for calcium oxalate stone prevention". Frontiers in Microbiology. 13: 1011102. doi: 10.3389/fmicb.2022.1011102 . PMC   9812493 . PMID   36620050.
  12. Jensen NS, Allison MJ (1994). Studies on the diversity among anaerobic oxalate-degrading bacteria now in the species Oxalobacter formigenes, abstr. I-12. 94th General Meeting of the American Society for Microbiology 1994. Washington, D.C., USA: American Society for Microbiology. p. 255.
  13. 1 2 Sidhu H, Allison M, Peck AB (February 1997). "Identification and classification of Oxalobacter formigenes strains by using oligonucleotide probes and primers". Journal of Clinical Microbiology. 35 (2): 350–353. doi:10.1128/jcm.35.2.350-353.1997. PMC   229578 . PMID   9003594.
  14. Garrity GM, Bell JA, Lilburn T (2005). "Class II. Betaproteobacteria class. nov.". In Brenner DJ, Krieg NR, Staley JT (eds.). Bergey's Manual® of Systematic Bacteriology. Boston, MA: Springer US. pp. 575–922. doi:10.1007/978-0-387-29298-4_2. ISBN   978-0-387-24145-6 . Retrieved 2022-11-10.
  15. Anantharam, V; Allison, M J; Maloney, P C (1989). "Oxalate:formate exchange". Journal of Biological Chemistry. 264 (13): 7244–7250. doi: 10.1016/S0021-9258(18)83227-6 .
  16. 1 2 3 Dawson KA, Allison MJ, Hartman PA (October 1980). "Isolation and some characteristics of anaerobic oxalate-degrading bacteria from the rumen". Applied and Environmental Microbiology. 40 (4): 833–839. Bibcode:1980ApEnM..40..833D. doi:10.1128/aem.40.4.833-839.1980. PMC   291667 . PMID   7425628.
  17. Kuhner, C H; Hartman, P A; Allison, M J (1996). "Generation of a proton motive force by the anaerobic oxalate-degrading bacterium Oxalobacter formigenes". Applied and Environmental Microbiology. 62 (7): 2494–2500. Bibcode:1996ApEnM..62.2494K. doi:10.1128/aem.62.7.2494-2500.1996. ISSN   0099-2240. PMC   168031 . PMID   8779588.
  18. 1 2 Cornick, N. A.; Allison, M. J. (1996). "Assimilation of oxalate, acetate, and CO 2 by Oxalobacter formigenes". Canadian Journal of Microbiology. 42 (11): 1081–1086. doi:10.1139/m96-138. ISSN   0008-4166. PMID   8941983.
  19. Cornick, N A; Allison, M J (1996). "Anabolic Incorporation of Oxalate by Oxalobacter formigenes". Applied and Environmental Microbiology. 62 (8): 3011–3013. Bibcode:1996ApEnM..62.3011C. doi:10.1128/aem.62.8.3011-3013.1996. ISSN   0099-2240. PMC   1388924 . PMID   16535386.
  20. 1 2 3 4 Duncan SH, Richardson AJ, Kaul P, Holmes RP, Allison MJ, Stewart CS (August 2002). "Oxalobacter formigenes and its potential role in human health". Applied and Environmental Microbiology. 68 (8): 3841–3847. Bibcode:2002ApEnM..68.3841D. doi:10.1128/AEM.68.8.3841-3847.2002. PMC   124017 . PMID   12147479.
  21. 1 2 3 Lange JN, Wood KD, Wong H, Otto R, Mufarrij PW, Knight J, et al. (June 2012). "Sensitivity of human strains of Oxalobacter formigenes to commonly prescribed antibiotics". Urology. 79 (6): 1286–1289. doi:10.1016/j.urology.2011.11.017. PMC   3569510 . PMID   22656407.
  22. Sidhu H, Enatska L, Ogden S, Williams WN, Allison MJ, Peck AB (June 1997). "Evaluating Children in the Ukraine for Colonization With the Intestinal Bacterium Oxalobacter formigenes, Using a Polymerase Chain Reaction-based Detection System". Molecular Diagnosis. 2 (2): 89–97. doi:10.1016/S1084-8592(97)80015-X. PMID   10462596.
  23. Kumar R, Mukherjee M, Bhandari M, Kumar A, Sidhu H, Mittal RD (March 2002). "Role of Oxalobacter formigenes in calcium oxalate stone disease: a study from North India". European Urology. 41 (3): 318–322. doi:10.1016/S0302-2838(02)00040-4. PMID   12180235.
  24. Kwak C, Jeong BC, Kim HK, Kim EC, Chox MS, Kim HH (May 2003). "Molecular epidemiology of fecal Oxalobacter formigenes in healthy adults living in Seoul, Korea". Journal of Endourology. 17 (4): 239–243. doi:10.1089/089277903765444384. PMID   12816588.
  25. Kodama T, Mikami K, Akakura K, Takei K, Naya Y, Ueda T, Ito H (July 2003). "[Detection of Oxalobacter formigenes in human feces and study of related genes in a new oxalate-degrading bacterium]". Hinyokika Kiyo. Acta Urologica Japonica. 49 (7): 371–376. PMID   12968475.
  26. Barnett C, Nazzal L, Goldfarb DS, Blaser MJ (February 2016). "The Presence of Oxalobacter formigenes in the Microbiome of Healthy Young Adults". The Journal of Urology. 195 (2): 499–506. doi:10.1016/j.juro.2015.08.070. PMC   4747808 . PMID   26292041.
  27. Kelly JP, Curhan GC, Cave DR, Anderson TE, Kaufman DW (April 2011). "Factors related to colonization with Oxalobacter formigenes in U.S. adults". Journal of Endourology. 25 (4): 673–679. doi:10.1089/end.2010.0462. PMC   3071521 . PMID   21381959.
  28. PeBenito A, Nazzal L, Wang C, Li H, Jay M, Noya-Alarcon O, et al. (January 2019). "Comparative prevalence of Oxalobacter formigenes in three human populations". Scientific Reports. 9 (1): 574. Bibcode:2019NatSR...9..574P. doi:10.1038/s41598-018-36670-z. PMC   6346043 . PMID   30679485.
  29. Clemente JC, Pehrsson EC, Blaser MJ, Sandhu K, Gao Z, Wang B, et al. (April 2015). "The microbiome of uncontacted Amerindians". Science Advances. 1 (3). Bibcode:2015SciA....1E0183C. doi:10.1126/sciadv.1500183. PMC   4517851 . PMID   26229982.
  30. James LF, Cronin EH (November 1974). "Management practices to minimize death losses of sheep grazing Halogeton-infested range". Journal of Range Management. 27 (6): 424–426. doi:10.2307/3896714. hdl: 10150/647155 . JSTOR   3896714.
  31. Allison MJ, Littledike ET, James LF (November 1977). "Changes in ruminal oxalate degradation rates associated with adaptation to oxalate ingestion". Journal of Animal Science. 45 (5): 1173–1179. doi:10.2527/jas1977.4551173x. PMID   599103.
  32. Daniel SL, Cook HM, Hartman PA, Allison MJ (August 1989). "Enumeration of anaerobic oxalate-degrading bacteria in the ruminal contents of sheep". FEMS Microbiology Letters. 62 (5): 329–334. doi: 10.1111/j.1574-6968.1989.tb03387.x .
  33. Federici F, Vitali B, Gotti R, Pasca MR, Gobbi S, Peck AB, Brigidi P (September 2004). "Characterization and heterologous expression of the oxalyl coenzyme A decarboxylase gene from Bifidobacterium lactis". Applied and Environmental Microbiology. 70 (9): 5066–5073. Bibcode:2004ApEnM..70.5066F. doi:10.1128/AEM.70.9.5066-5073.2004. PMC   520889 . PMID   15345383.
  34. Sidhu H, Hoppe B, Hesse A, Tenbrock K, Brömme S, Rietschel E, Peck AB (September 1998). "Absence of Oxalobacter formigenes in cystic fibrosis patients: a risk factor for hyperoxaluria". Lancet. 352 (9133): 1026–1029. doi:10.1016/S0140-6736(98)03038-4. PMID   9759746. S2CID   25936201.
  35. 1 2 Mittal RD, Kumar R, Bid HK, Mittal B (January 2005). "Effect of antibiotics on Oxalobacter formigenes colonization of human gastrointestinal tract". Journal of Endourology. 19 (1): 102–106. doi:10.1089/end.2005.19.102. PMID   15735393.
  36. Kaufman DW, Kelly JP, Curhan GC, Anderson TE, Dretler SP, Preminger GM, Cave DR (June 2008). "Oxalobacter formigenes may reduce the risk of calcium oxalate kidney stones". Journal of the American Society of Nephrology. 19 (6): 1197–1203. doi:10.1681/ASN.2007101058. PMC   2396938 . PMID   18322162.
  37. Hoppe B, Groothoff JW, Hulton SA, Cochat P, Niaudet P, Kemper MJ, et al. (November 2011). "Efficacy and safety of Oxalobacter formigenes to reduce urinary oxalate in primary hyperoxaluria". Nephrology, Dialysis, Transplantation. 26 (11): 3609–3615. doi: 10.1093/ndt/gfr107 . PMID   21460356.
  38. Hoppe B, Niaudet P, Salomon R, Harambat J, Hulton SA, Van't Hoff W, et al. (May 2017). "A randomised Phase I/II trial to evaluate the efficacy and safety of orally administered Oxalobacter formigenes to treat primary hyperoxaluria". Pediatric Nephrology. 32 (5): 781–790. doi:10.1007/s00467-016-3553-8. PMID   27924398. S2CID   52271121.
  39. Milliner D, Hoppe B, Groothoff J (August 2018). "A randomised Phase II/III study to evaluate the efficacy and safety of orally administered Oxalobacter formigenes to treat primary hyperoxaluria". Urolithiasis. 46 (4): 313–323. doi:10.1007/s00240-017-0998-6. PMC   6061479 . PMID   28718073.
  40. Tang R, Jiang Y, Tan A, Ye J, Xian X, Xie Y, et al. (November 2018). "16S rRNA gene sequencing reveals altered composition of gut microbiota in individuals with kidney stones". Urolithiasis. 46 (6): 503–514. doi:10.1007/s00240-018-1037-y. PMID   29353409. S2CID   11340007.
  41. Ticinesi A, Milani C, Guerra A, Allegri F, Lauretani F, Nouvenne A, et al. (December 2018). "Understanding the gut-kidney axis in nephrolithiasis: an analysis of the gut microbiota composition and functionality of stone formers". Gut. 67 (12): 2097–2106. doi:10.1136/gutjnl-2017-315734. PMID   29705728. S2CID   14055215.
  42. Ticinesi A, Nouvenne A, Meschi T (July 2019). "Gut microbiome and kidney stone disease: not just an Oxalobacter story". Kidney International. 96 (1): 25–27. doi: 10.1016/j.kint.2019.03.020 . PMID   31229040. S2CID   195327195.
  43. Miller AW, Choy D, Penniston KL, Lange D (July 2019). "Inhibition of urinary stone disease by a multi-species bacterial network ensures healthy oxalate homeostasis". Kidney International. 96 (1): 180–188. doi:10.1016/j.kint.2019.02.012. PMC   6826259 . PMID   31130222.
  44. Liu M, Koh H, Kurtz ZD, Battaglia T, PeBenito A, Li H, et al. (August 2017). "Oxalobacter formigenes-associated host features and microbial community structures examined using the American Gut Project". Microbiome. 5 (1): 108. doi: 10.1186/s40168-017-0316-0 . PMC   5571629 . PMID   28841836.
  45. 1 2 Arvans D, Jung YC, Antonopoulos D, Koval J, Granja I, Bashir M, et al. (March 2017). "Oxalobacter formigenes-Derived Bioactive Factors Stimulate Oxalate Transport by Intestinal Epithelial Cells". Journal of the American Society of Nephrology. 28 (3): 876–887. doi:10.1681/ASN.2016020132. PMC   5328155 . PMID   27738124.
  46. Hatch M, Cornelius J, Allison M, Sidhu H, Peck A, Freel RW (February 2006). "Oxalobacter sp. reduces urinary oxalate excretion by promoting enteric oxalate secretion". Kidney International. 69 (4): 691–698. doi: 10.1038/sj.ki.5000162 . PMID   16518326.
  47. Arvans D, Chang C, Alshaikh A, Tesar C, Babnigg G, Wolfgeher D, et al. (July 2023). "Sel1-like proteins and peptides are the major Oxalobacter formigenes-derived factors stimulating oxalate transport by human intestinal epithelial cells". American Journal of Physiology. Cell Physiology. 325 (1): C344–C361. doi:10.1152/ajpcell.00466.2021. PMC  10393326. PMID   37125773.