Bacteria

Cyanobacterium

Figure 1. A filament of cyanobacterial cells from the family Oscillatoriaceae. The short filament shown (called a hormogonium) was formed from fragmentation of a larger filament as a means of reproductive dispersal. The hormogonium is capable of moving by gliding over surfaces. This particular taxon is named for its oscillatory movements.

Some General Characteristics

Bacteria are the most abundant and widespread cellular life forms on Earth. They are generally microscopic single-celled organisms that in some cases form relatively simple multicellular entities (Figure 1). Almost all bacteria have a cell wall. The characteristic component of that cell wall is a polymer called peptidoglycan. General characteristics of bacteria are described below.

Habitats

Bacteria have amassed an extraordinary number of adaptations and metabolic abilities, which have enabled them to survive in practically every habitat on earth. They are found in a diversity of aquatic and terrestrial environments— in waters, soils, rocks, and in the cells and/or bodies of other organisms.

They also occur in extreme environments such as high in the atmosphere (Smith et al. 2018); in the deepest oceanic trenches (Glud et al. 2013); in the deepest layer of ocean crust (Mason et al. 2010); in the driest deserts (Schulze-Makuch et al. 2018); in water bodies that are very acidic (Quatrini and Johnson 2018) or alkaline (Padan et al. 2005), and at extreme temperatures. Some bacteria remain biologically active at temperatures as low as -20°C, which is believed to be the lowest temperature limit for sustaining life (D'Amico et al. 2006, Rivkina et al. 2000); other bacteria are capable of growing at temperatures as high as 100°C (Kashefi et al. 2002).

Although some bacteria are able to survive long periods under very dry conditions, their cells ultimately require water for metabolic activity— this is true for all cells regardless of the type of organism. Thus, all bacterial cells (and all cells in general) are in essence aquatic. The so-called "terrestrial" bacteria actually live within thin films of water found on the surfaces of soils, plants or other objects, or within water-filled pores of rocks or soils (Fenchel 1994).

Cell Size

Bacteria typically have cell volumes ranging from around 0.4 to 3 µm3 (Levin and Angert 2015) and cell diameters that are almost always larger than 0.2 µm—which is the filter pore size used for "sterile" filtration. However, some bacteria are smaller, and others much larger, than those typical sizes. Within a given species, cell size is known to vary greatly with nutrient availability, with starved cells tending to be much smaller than cells growing under nutrient-replete conditions (Chien et al. 2013).

The smallest described free-living bacterial cells include a group of marine Actinobacteria with an average cell volume of approximately 0.013 µm3 (Ghai et al. 2013), and diverse bacterial taxa in groundwater with a median cell volume of 0.009±0.002 µm3 (Luef et al. 2015). At the other extreme are bacterial cells that are large enough to be seen with the naked eye, such as Thiomargarita namibiensis, which is spherical and has a cell diameter of 750 µm (Schulz et al. 1999), and Epulopiscium fishelsoni, with a cell size of approximately 80 by 600 µm (Angert et al. 1993). The largest known bacterium has not yet been grown in culture. It is Candidatus Thiomargarita magnifica and has an average cell length greater than 9,000 µm (Volland et al. 2022).

Cell Shape

Bacteria shapes

Figure 2. Some shapes of bacterial cells include (A) cocci, (B) bacilli, (C) coccobacilli, (D) vibrios, (E) spiral-shaped (spirilla and spirochaetes), (F) filamentous, (G) prosthecate, and (H) pleomorphic.

Bacterial cells come in many different shapes (Figure 2). Many bacteria have scientific names that describe their shapes. Some cell shapes include:

  • spheres— also known as cocci (singular: coccus), e.g., Streptococcus pyogenes, Staphylococcus aureus, and Veillonella parvula.

  • rods—also referred to as bacilli (singular: bacillus), e.g., Lactobacillus acidophilus, Bacillus subtilis, and Clostridium botulinum.

  • short rods or ovals— known as coccobacilli (singular: coccobacillus), e.g., Bordetella pertussis, Haemophilus influenzae, and Pasteurella multocida

  • comma shaped (curved rods or incomplete spirals)— called vibrios, e.g., Vibrio cholerae, Butyrivibrio fibrisolvens, and Desulfovibrio desulfuricans.

  • rigid spirals— called spirillas, e.g., Rhodospirillum rubrum, Helicobacter pylori, and Campylobacter jejuni.

  • flexible spirals— called spirochaetes, e.g., Treponema pallidum, Borreliella (Borrelia) burgdorferi, and Leptospira interrogans.

  • filamentous— with long narrow filament-like structures (modified cells) called hyphae. The hyphae grow by extension of the tip and by formation of new branches in a manner similar to the hyphae of fungi. Like fungi, some filamentous bacteria may form masses of branching hyphae known as mycelia. The Phylum Actinobacteriota (Actinobacteria) contains taxa that form hyphae as part of their normal life cycle, e.g., Streptomyces spp., Actinomadura spp., and Microbispora spp.

  • prosthecate— having stalk-like cellular appendages called prosthecae (singular: prostheca). These appendages are narrowed portions of the cell. Like other parts of the cell they contain cytoplasm and are surrounded by cell membrane and cell wall. Some bacteria have a holdfast — gel-like adhesive material— at the end of their prostheca, which allows them to attach to surfaces. Examples of prosthecate bacteria include Hyphomonas neptunium, Caulobacter crescentus, and Asticcacaulis excentricus.

  • variable shaped (with no characteristic shape)— also known as pleomorphic. e.g., Mycoplasma pneumoniae, Phytoplasma mali, and L-forms of Mycobacterium tuberculosis.

Cell Components and Structure

The cell of a bacterium has two main parts: (1) the cytoplasm and (2) the various layers that surround the cytoplasm, known as the cell envelope.

Cytoplasm

The bulk of a bacterial cell consists of cytoplasm. It includes the watery matrix of the cell, called the cytosol (sometimes referred to as the cytoplasmic matrix), and all cellular components found within the cytosol. These components include ribosomes— tiny complex structures that carry out protein synthesis; and the nucleoid— the region within the cell where the bacterial chromosome (or chromosomes) are located. Small extrachromosomal DNA molecules may also be found within the cytosol. In addition, bacterial cells may contain various inclusions (also called inclusion bodies)— materials stored directly within the cytosol or within special structures that are bounded by protein or lipid membrane. Examples of inclusions include:

  • cyanophycin granules— these are inclusions that apparently lack any kind of surrounding membrane or proteinaceous shell (Allen and Weathers 1980, Seufferheld et al. 2003). They consist of a nitrogen-rich polymer called multi-L-arginyl poly-(L-aspartic acid), which contains equimolar amounts of arginine and aspartate (Allen and Weathers 1980, Klemke et al. 2016). Cyanophycin granules are found in many cyanobacteria and in some heterotrophic bacteria, and they apparently serve as a storage reserve for nitrogen, carbon, and energy (Füser and Steinbüchel, 2007).

  • protein-bounded gas vesicles, which enable some bacteria to control their buoyancy and position within a water column. These vesicles are impermeable to water but highly permeable to gases (Walsby 1994). Gas vesicles are found mostly in planktonic bacteria and archaea, including many cyanobacteria, purple sulfur bacteria, green sulfur bacteria, and green nonsulfur bacteria (Walsby 1994);

  • carboxysomes— protein-bounded structures found in cyanobacteria, many chemoautotrophs, and a few purple sulfur bacteria (Turmo et al. 2017). These structures contain concentrated CO2 and enzymes needed for "fixing" CO2 — i.e., for converting CO2 into organic compounds (Turmo et al. 2017, Rae et al. 2013).

  • lipid droplets— which contain lipids and are surrounded by a lipid membrane. The membrane consists of a monolayer of phospholipids with many proteins (Yang et al. 2012). Lipid droplets are found in the cells of bacteria, archaea, and eukaryotes— including humans (Murphy 2012), and may have been present in an ancient ancestor of these groups (Zhang and Liu 2017). They were originally believed to function only as storage reservoirs for lipids; however, they are now recognized as dynamic organelles with multiple functions, including the storage, transport, and metabolism of lipids, and the binding and protection of nucleic acids (Zhang and Liu 2017, Yang et al. 2012). Additional functions have been identified in eukaryotes, but have not yet been well studied in bacteria (Zhang and Liu 2017). Lipid droplets are known to occur in most bacteria (Docampo 2016).

  • magnetosomes— organelles that consist of a magnetic iron-containing crystal, usually magnetite (Fe3O4) or greigite (Fe3S4), surrounded by a lipid bilayer membrane that contains proteins (Docampo 2016). These organelles are typically arranged in one or more chains within a cell. Bacteria containing magnetosomes— the so-called magnetotactic bacteria— use these organelles to align themselves with and swim along geomagnetic field lines (Yan et al. 2012). Magnetotactic bacteria have a gram-negative cell envelope, they are found in environments with low concentrations of molecular oxygen (i.e., they are typically anaerobes or microaerophiles), and they belong to at least six different phyla of bacteria— Desulfobacterota, Proteobacteria, Nitrospirota, and Planctomycetota; and the uncultured (candidate) phyla Omnitrophota (Omnitrophica) and Latescibacterota (Latescibacteria) (Lin et al. 2018, Yan et al. 2012, Genome Taxonomy Database 2018). The magnetosome genes in these different lineages appear to have a common origin (Lin et al. 2018).

  • acidocalcisomes— acidic membrane-bound organelles containing phosphorous compounds (including polyphosphate, pyrophosphate, and orthophosphate) that are complexed with calcium and other cations (Docampo et al. 2010, Docampo 2016). These organelles also contain enzymes for the synthesis and degradation of pyrophosphate and polyphosphate (Docampo and Moreno 2011). The surrounding membrane consists of a lipid bilayer and contains proteins, including proton pumps, which maintain the acidity of the organelle (Docampo 2016). Acidocalcisomes are found in bacteria, eukaryotes (including humans), and probably archaea (Seufferheld et al 2011). They were previously referred to as volutin granules, metachromatic granules, or polyphosphate granules, vacuoles or bodies (Docampo et al. 2005, 2010; Seufferheld et al. 2011). Given their presence in the different cellular domains, these organelles, like lipid droplets, may have been present in an ancient ancestor, or they could be examples of convergent evolution (Seufferheld et al. 2011, Docampo 2016). Acidocalcisomes have been identified in the bacteria Agrobacterium tumefaciens and Rhodospirillum rubrum (Docampo 2016). They are important sites for the storage of cations and phosphorus (Docampo and Moreno 2011, Docampo et al. 2010). In eukaryotes, they are also known to be involved in calcium/cation homeostasis, regulation of intracellular pH, and osmoregulation (Docampo et al. 2010).

  • anammoxosomes— membrane-bound organelles found in anaerobic ammonium-oxiding (anammox) bacteria from the phylum Planctomycetes (Order Brocadiales). The membrane surrounding this organelle consists of a lipid bilayer (Neumann et al. 2014). This bilayer is made up of unusual lipids called ladderanes and also contain proteins (Docampo 2016), including ATP synthase (de Almeida et al. 2015)— a molecular machine used to produce adenosine triphosphate (ATP). The anammoxosome is a dedicated compartment where the ammonium-oxidizing catabolic reactions take place and where usable energy— in the form of ATP— is produced (de Almeida et al. 2015). In the anammox reactions, ammonium (NH4+) and nitrite (NO2-) are converted to nitrogen gas (N2). In the process high energy electrons are captured, and fed into an electron transport chain. Energy released from the chemical reactions of the electron transport chain is used to drive proton pumps in the anammoxosome membrane, resulting in a highly energized membrane (due to a difference in proton concentration and charge across the membrane). The energy stored in the membrane is used to produce ATP. Thus the anammoxome, like the mitochondrion in eukaryotes, is believed to be an energy-generating organelle (Neumann et al. 2014, van Niftrik and Jetten 2012). Unlike the mitochondrion, the anammoxosome is believed to have originated from an invagination of the cytoplasmic membrane— because all membranes within the cell resemble one another in structure and lipid composition (Neumann et al. 2014). The anammoxome organelle has been successfully isolated from cells of anammox bacteria; thus it appears likely that the membrane surrounding this organelle is completely separate from the other membranes in the cell (Neumann et al. 2014).

Another feature that may be found in the cytoplasm of bacteria— at least within certain groups of bacteria— is intracytoplasmic membrane. Intracytoplasmic membranes are essentially invaginations of the cytoplasmic membrane. They can take the form of tubules, interconnected vesicles, and plate-like layers of membrane referred to as lamellae (Niederman 2006). These membranes may provide additional surface area for the production of ATP and they create compartments where metabolic processes can take place (Niederman 2006, LaSarre et al. 2018). The thylakoid membranes of cyanobacteria are examples of intracytoplasmic membrane. Bacteria that typically contain intracytoplasmic membranes include:

  • phototrophic bacteria— which harvest energy from sunlight; examples include the cyanobacterium Synechococcus and the purple nonsulfur bacterium Rhodobacter;
  • nitrifying bacteria— autotrophic bacteria that oxidize ammonia or nitrite to obtain energy; examples include the ammonia-oxidizing Nitrosomonas and the nitrite-oxidizing Nitrobacter.
  • methane-utilizing bacteria— which generate energy by oxidizing methane; for example, Methylomicrobium album.
  • bacteria from the phyla Planctomycetota (Planctomycetes) and Verrucomicrobiota (Verrucomicrobia)— these bacteria tend to have cells that are compartmentalized by internal membranes. For example, Gemmata obscuriglobus.

Note: there is some overlap when describing intracytoplasmic membrane and some of the membrane-bound organelles. For example, the membranes of magnetosomes have been found to be invaginations of the cytoplasmic membrane (Komeili et al. 2006) and thus could be described as intracytoplasmic membrane. Often it can be difficult to tell whether a membrane surrounding an internal compartment is completely separate from the cytoplasmic membrane.

Some bacteria may form intracytoplasmic membranes only under certain growth conditions. For example, in one study, cells of Methylobacterium organophilum produced intracytoplasmic membranes when they were grown in medium containing methane, but they failed to form intracytoplasmic membranes when they were given methanol or glucose as their sole carbon and energy source (Patt and Hanson 1978). Environmental conditions may also affect the amount of intracytoplasmic membrane that a cell produces. Photosynthetic bacteria may produce a greater amount of intracytoplasmic membrane under low light conditions than under high light conditions (Adamas and Hunter 2012). This is because under low light, there are fewer photons available per unit area and time, thus cells need to produce more pigments to effectively harvest light—and more intracytoplasmic membrane is necessary to contain those pigments (Adams and Hunter 2012). In laboratory studies, scientists have been able to induce the formation of intracytoplasmic membranes in Escherichia coli by making the cells overproduce certain membrane proteins (Jamin et al. 2018, Arechaga et al. 2000).

Ribosomes

Scattered within the cytosol of all bacteria are thousands of ribosomes. Ribosomes may also be found loosely attached to the cytoplasmic membrane (Prescott et al. 1993). These structures are responsible for the synthesis of proteins in a cell. In essence, they are molecular machines that take segments of an organism's genetic code (specified by RNA nucleotides) and "translate" them into sequences of amino acids to form polypeptides (chains of amino acids). Once a polypeptide is completed and dissociates from the ribosome, it spontaneously folds up to form a protein (with the shape of the protein determined by the sequence of amino acids in the polypeptide chain).

Ribosomes are found in all cellular domains— Bacteria, Eukarya, and Archaea. They consist of two subunits— a large subunit and a small subunit. Each of the subunits is made up of a complex of RNA and proteins. Over 50 different ribosomal proteins have been identified in bacteria; 34 of these proteins are also found in the ribosomes of archaea and eukaryotes (Yutin et al. 2012).

Ribosomes and their components are described in terms of their sedimentation values or "S values" (also known as sedimentation coefficients), which refer to their rate of sedimentation in an ultracentrifuge. S values are measured in Svedberg units (abbreviated as "S"). When particles are suspended in a liquid medium and a centrifugal force is applied to them, the rate at which they separate out to form sediment depends on their mass, volume, and shape; with heavier, more compact particles tending to move/sink to the bottom of the centrifuge tube at a faster rate (i.e., have a larger S value) than lighter, less compact particles.

Bacteria and archaea have ribosomes with a sedimentation value of 70S. The large subunit has a sedimentation value of 50S and is made up to two ribosomal RNA (rRNA) molecules (a 5S rRNA and a 23S rRNA) and approximately 34 ribosomal proteins, while the small subunit has a sedimentation value of 30S and is made up of a single rRNA molecule (a 16S rRNA) and approximately 21 ribosomal proteins (Prescott et al. 1993). In comparison, the eukaryotic ribosome is larger— 80S, and it is made up of 40S and 60S subunits (Prescott et al. 1993).

Protein synthesis is of vital importance to cells. Virtually all metabolic reactions that take place in cells involve proteins. Proteins are also involved in the transport of substances across membranes, cell signaling, the segregation of chromosomes, cell division, and the maintenance of cell shape, among other functions. The bulk of a cell's total energy budget goes towards making proteins (Harold 1986, Russel and Cook, 1995).

When bacterial cells are stressed, e.g., when they are starved of nutrients, they are able to deactivate many of their ribosomes to conserve energy—so they don't waste energy trying to make proteins under unfavorable conditions. With a substantial reduction in the number of proteins being produced, other metabolic processes within the cell (which are dependent on proteins) are likely scaled back proportionately, further conserving energy (McKay and Portnoy 2015).

The deactivation of ribosomes occurs as follows: When nutrient-limited or otherwise stressed, bacteria produce proteins known as dimerizing factors. These proteins interact with (bind to) 70S ribosomes and cause two 70S ribosomes to bind together, forming what is known as a 100S ribosome. The 100S ribosomes are nonfunctional—they are unable to produce proteins and thus are called "hibernating ribosomes." Once environmental conditions become favorable again (e.g., once starved cells are provided with fresh, nutrient-rich media) the 100S ribosomes dissociate into 70S ribosomes, which once again become active and capable of making proteins.

A bacterium's ability to produce 100S ribosomes increases its survival rate enormously— reportedly by 2 to 3 orders of magnitude (Gohara and Yap 2018). Without this ability, the ribosomes in stressed bacteria become rapidly degraded, and the cells tend to die prematurely (Basu and Yap 2017). Virtually all bacteria appear to be capable of producing 100S ribosomes, although the mechanisms involved vary depending on the type of bacterium (McKay and Portnoy 2015).

Genetic Material

The genetic material, transcription, and translation

Most bacteria have their DNA in the form of a single circular chromosome (diCenzo and Finan 2017). The DNA is double-stranded, highly condensed— i.e., linearly compacted over 1,000-fold (Holmes and Cozzarelli 2000, Trun and Marko 1998), and occurs in association with various proteins or protein complexes (Badrinarayan et al. 2015, Trojanowski et al. 2018). The associated proteins are involved in the organization, condensation, and segregation of the chromosomes and may have other functions related to the cell cycle (Badrinarayan et al. 2015, Trojanowski et al. 2018).

Bacterial DNA is densely packed with genes— it contains very few intravening sequences that don't code for proteins. Thus in bacteria, the size of the genome (the complete set of genetic material for the organism) is strongly proportional to the number of genes found in that organism (Koonin 2011). This is in contrast to eukaryotes, where the relationship between genome size and gene size is decoupled— eukaryotes tend to contain relatively large amounts of noncoding DNA, and the quantity of noncoding DNA may vary hugely depending on the species (Koonin 2011).

Unlike in eukaryotes, bacterial chromosomes are not enclosed in a special compartment within the cell. Instead, the genetic material is suspended within the cytoplasmic matrix. The area of the cell containing the chromosome(s) is referred to as the nucleoid. (The outline of the nucleoid is coincident with the outline of the chromosomal material within the cell.) The nucleoid can be thought of as an organized tangle of chromosomal material that takes the form of a mesh, which has a certain pore size (Castellana et al. 2016). Molecules smaller than the pores can freely pass into the interior of the mesh, while molecules larger than the pores cannot.

The nucleoid is typically located near the middle of the cell (Bakshi et al. 2012). In the nucleoid, RNA is produced from DNA in a process known as transcription. The types of RNA that are produced include ribosomal RNA (rRNA)— the primary component of ribosomes, messenger RNA (mRNA)— which is used as the template from which to make proteins, and transfer RNA (tRNA)— which carries amino acids to the mRNA based on the sequence of nucleotides encoded in the mRNA. Whereas rRNA and tRNA are stable forms of RNA, mRNA has a lifespan of around 6 to 7 minutes (Bernstein et al. 2002).

Unlike the case of eukaryotes, where transcription (production of RNA from DNA) and translation (synthesis of proteins) occur at separate locations and times (transcription in the nucleus and translation in ribosome-rich areas of the cytoplasm and endoplasmic reticulum), in bacteria the two processes can be coupled (in a phenomenon known as co-transcriptional translation) (Miller et al. 1970, Bakshi et al. 2015). Although co-transcriptional translation occurs within the nucleoid, the nucleoid has been found to be largely devoid of ribosomes— in part because bound ribosomes are larger than the average pore size of the nucleoid (Castellana et al. 2016). At least in the case of Escherichia coli, the ribosomes tend to be strongly localized to the cell poles and segregated from the nucleoid (Bakshi et al. 2012). However, it has been shown that free ribosomal subunits (30S and 50S subunits) can readily enter the nucleoid (Sanamrad et al. 2014). Thus the following scenario is believed to occur: as the mRNAs are formed, the ribosomal subunits bind to them, i.e., a 30S subunit binds to an initiation site on the mRNA, and then a 50S subunit binds to the 30S unit, forming a functional 70S ribosome. Multiple ribosomes bind to a single mRNA in this manner, forming what is known as a 70S polysome. Then, through some phenomenon— perhaps due at least in part to excluded volume effects— the 70S polysomes are excluded from the nucleoid and migrate towards the poles of the cell (Castellana et al. 2016, Bakshi et al. 2015). Protein synthesis continues outside of the nucleoid; the mRNAs can be translated into proteins repeatedly before finally degrading (Bakshi et al. 2015).

At least with regards to Escherichia coli, there is a cycling of ribosomes within the cell: ribosomal subunits flow into the nucleoid where they bind to mRNA to form 70S polysomes; the polysomes flow out of the nucleoid towards the cell poles where proteins synthesis is continued; then at the poles the mRNAs degrade and the ribosomes dissociate into subunits, which flow into the nucleoid (Bashi et al. 2015, Castellano et al. 2016). Based on studies with Escherichia coli, most protein synthesis occurs post-transcriptionally, outside of the nucleoid (Bashi et al. 2015, Castellano et al. 2016), while around 10 to 15 percent of protein synthesis appears to occur co-transcriptionally, within the nucleoid (Bashi et al. 2015).

DNA replication

Replication of circular chromosome Replication of linear chromosome

Figure 3. Replication of the Bacterial Chromosome. The original DNA is shown in blue, the newly synthesized DNA is shown in pink. The origin of replication (oriC) is shown in yellow. OriC is a locus on the chromosome where replication is initiated. Replication is shown for (A) a circular chromosome and (B) a linear chromosome. In both cases, a replication "bubble" forms at oriC and the "bubble" expands until the entire chromosome is copied.

The bacterial chromosome consists of a single replicon—it has a single origin of replication (replication is initiated from just one point on the chromosome) and the whole chromosome is replicated as a single unit. From the origin of replication (known as oriC), replication of the DNA proceeds in two directions around the chromosome. Starting at the origin, the two strands of the original DNA begin to separate, forming a "bubble" and new DNA is synthesized along each of the original strands. As replication proceeds, the "bubble" expands in both directions until the entire chromosome is copied (Figure 3).

In contrast to bacteria, eukaryotes have linear chromosomes with multiple origins of replication— they can potentially have hundreds of replication origins per chromosome (Kuzminov 2014). On eukaryotic chromosomes, origins of replication occur every 10 to 100 μm and each of the associated replicons can be copied simultaneously, so that a large DNA molecule may be quickly replicated (Prescott et al. 1993). Because bacterial chromosomes have only a single origin of replication, replication time increases with the size of the chromosome. Thus, it is advantageous for bacteria to have small chromosomes so that they can maximize their rate of reproduction. [1]

Although most bacteria have circular chromosomes, linear chromosomes are found in some taxa, such as Borreliella (Borrelia) species, which cause lyme disease (Ferdows and Barbour 1989, Casjens et al. 1995), and several species of Streptomyces (Yang et al. 2002), which are the source of many of our antibiotics (Chater 2006). Linear chromosomes in bacteria are apparently evolved from circular bacterial chromosomes (Volff and Altenbuchner 2000). Thus, in contrast to eukaryotic chromosomes, linear chromosomes in bacteria have only one origin of replication (Trojanowski et al 2018). The origin is typically located near the middle of the chromosome (Volff and Altenbuchner 2000, Picardeau et al. 1999, Hopwood 2006).

The cell cycle

The cell cycle in bacteria and other organisms involves, among other things, the replication of the chromosome(s), the segregation of the daughter chromosomes to different ends of the cell, and cell division— the splitting apart of the cell into daughter cells (with each daughter cell receiving a copy of the genetic material). Whereas eukaryotes have cell cycles with discrete, non-overlapping stages, bacterial cell cycles are quite different.

Multifork Replication

Figure 4. Multifork Replication. When certain species of bacteria such as Escherichia coli grow at fast growth rates, multiple generations of chromosomes are replicated in parallel. In this figure, three generations of chromosomes are being replicated. The replication bubbles for the daughter chromosomes, granddaughter chromosomes, and great-granddaughter chromosomes are labeled A, B, and C, respectively. The origins of replication (oriC) are shown in yellow. The parental DNA is shown in dark blue and the three rounds of newly synthesized DNA are shown in pink, light blue and orange.

Compared to eukaryotes, bacteria are unusual in that they have cell cycles in which DNA replication and chromosome segregation occur simultaneously (Kuzminov 2013, Badrinarayanan et al. 2015). Shortly after a given locus on a chromosome is replicated, the sister loci are segregated; so the segregation of the various loci along the chromosome occurs sequentially rather than simultaneously, and the overall rate of segregation equals the rate of replication (Kuzminov 2013). In addition, in at least some taxa of bacteria multiple rounds of chromosome replication can occur in parallel within a single cell cycle in a phenomenon known as multifork replication (Yoshikawa et al. 1964, Kuzminov 2013, Bremer and Dennis 1996, Trojanowski et al. 2017). In other words, shortly after a cell has begun replication to create daughter chromosomes (and before the daughter chromosomes are fully copied and fully segregated into daughter nucleoids) replication of the daughter chromosomes may commence (to produce granddaughter chromosomes). And shortly after replication of the granddaughter chromosomes has started (and before the daughter and granddaughter chromosomes have been fully replicated and segregated), the cell can begin replication of the granddaughter chromosomes (to produce great-granddaughter chromosomes), etc. In Escherichia coli, as many as four overlapping rounds of chromosome replication can occur (Morigen et al. 2009). Multifork replication tends to occur in bacteria that are growing rapidly under ideal, nutrient-rich conditions (Wang and Levin 2009). In slowly growing cells of Escherichia coli, the chromosome is only replicated once prior to cell division (Cooper and Helmstetter 1968, Skarstad et al. 1983). However, multifork replication has been observed in a fraction of slow-growing Mycobacterium smegmatus cells, including cells grown under suboptimal conditions (Trojanowski et al. 2017). So the interplay of factors causing multifork replication is not fully known. Whether or not multifork replication occurs may depend on the growth rate (Wang and Levin 2009), the taxa of bacteria (Fossum et al. 2007), and other factors.

The phenomenon of multifork replication has some interesting consequences: (1) The number of copies (copy number) of the chromosome (ploidy) can change from one generation to the next— depending on how fast the cells are growing and what stage of growth they are in (Pecoraro et al. 2011, Maldonado et al. 1994); (2) Cells can inherit incomplete copies of chromosomes (i.e., copy numbers that are not whole numbers). For example, when growing under optimal conditions, Escherichia coli is merodiploid (having one full copy plus one or more partial copies of its chromosome) to mero-oligoploid (having a few full copies and one or more partial copies of its chromosome) (Pecoraro et al. 2010); and (3) In stark contrast to eukaryotic cells —where the time required to complete the cell cycle is equal to the sum of the amounts of time required to complete each of the discrete cell cycle stages— many bacteria that undergo multifork replication can divide faster than the amount of time it takes to replicate their chromosome. For example, Escherichia coli can divide as quickly as once every 24 minutes whereas it takes 42 minutes to over an hour to replicate its chromosome (Bremer and Dennis 1996).

The bacterial cell cycle varies somewhat depending on the taxa of bacteria. Bacteria with nontypical cell cycles include species with more than one chromosome/essential replicon, those that undergo cell differentiation, and species with complicated life cycles (Trojanowski et al. 2018). The cell cycle in bacteria is still being elucidated. It is described by some as appearing to consist of "a set of coordinated but independent events"—as opposed to one unified process (Wang and Levin 2009)— and by others as a single event, with the different cell cycle processes occurring together, in the same location, probably by physically interacting entities (Kuzminov 2013).

Chromosomes and other replicons within the cell

Chromosome (Primary Chromosome, Primary Essential Replicon).

The bacterial chromosome is sometimes referred to as the primary chromosome or the primary essential replicon, because other types of replicons—secondary replicons— may be present in a cell. (All DNA molecules found within bacterial cells [i.e., chromosomes and secondary replicons] consist of single replicons [i.e., having one origin of replication]— this is in contrast to eukaryotic chromosomes, which are made up of many replicons.)

The chromosome is most easily distinguished from secondary replicons based on size: the chromosome is always the largest replicon in the cell (Harrison et al. 2010, diCenzo and Finan 2017). It contains all or most of the essential genes needed by the bacterium for survival (Harroson et al. 2010, diCenzo and Finan 2017, Ochman 2002, Egan et al. 2005), and its replication is tied to the cell cycle (Egan et al. 2005).

Secondary replicons

Secondary replicons include plasmids, megaplasmids, and chromids (diCenzo and Finan 2017). One or more of these types of secondary replicons may be found within a cell. Based on available genome data, approximately 40 percent of bacterial species contain secondary replicons (diCenzo and Finan 2017). Like chromosomes, secondary replicons consist of double-stranded DNA molecules, which are typically in the form of a closed circle, although some are linear (Ochman 2002, Hopwood 2006, Chaconas and Kobryn 2010). Secondary replicons have a replication system that is distinct from that of the chromosome; they have a plasmid-like replication system with a characteristic plasmid-like origin of replication (Fournes et al. 2018). Secondary replicons that are greater than 25 kilobases in size tend to carry genes for partition systems (Planchenault et al. 2020). Partition systems segregate copies of secondary replicons, so that when a cell divides, each daughter cell will receive a copy (or copies) of the replicon. Partition systems are generally absent from secondary replicons less than 25 kilobases (Planchenault et al. 2020); with no active partition system, these smaller replicons are randomly segregated into daughter cells. (When random segregation occurs, there is no guarantee that both daughter cells will receive a copy of the replicon.) Secondary replicons are described below:

  • plasmids — are small DNA molecules that are considered to be dispensible to the cell; they don't carry any core bacterial genes essential for the survival/viability of their host (Egan et al. 2005). A plasmid is an example of an acellular parasite. It is an independent agent that utilizes the host cell's machinery to replicate its DNA and, in many cases, to spread itself to other host cells. Plasmids are replicated independently of the host cell's chromosome and independently of the cell cycle (del Solar et al. 1998 referenced in Egan et al. 2005).[2] A plasmid's DNA typically includes core plasmid genes— for replication of the plasmid and transmission of the plasmid to other bacteria, and accessory genes— other miscellaneous genes, which in some cases may prove beneficial to the bacterial host, for example by providing antibiotic resistance or resistance to heavy metals (Maclean and San Millan 2015). Plasmids that lack partition systems (i.e., plasmids < 25 kilobases) are normally found along the outer edge (just outside) of the nucleoid (they appear to be excluded from the nucleoid), while plasmids that have partition systems are found within the volume of the nucleoid (but separate from the bacterial chromosome) (Planchenault et al. 2020).[3]

  • megaplasmids— like plasmids, megaplasmids don't carry any genes needed by the bacterium for survival/viability. As is true for larger sized plasmids, megaplasmids are found within the volume of the nucleoid and they contain active partition systems. Megaplasmids are distinguished from plasmids solely based on size; however, there is no standardly accepted size cutoff between megaplasmids and plasmids.[4] Although they are essentially defined as large plasmids, megaplasmids appear to have genomic features and other characteristics that are intermediate between those of plasmids and chromids (diCenzo and Finan 2017). In contrast to plasmids, megaplasmids may replicate and segregate in coordination with the bacterial cell cycle (diCenzo and Finan 2017).

  • chromids (Secondary Chromosomes, Secondary Essential Replicons)—are so named because they contain characteristics of both plasmids and chromosomes. Unlike plasmids and megaplasmids, chromids contain one or more core bacterial genes essential for the survival/viability of the bacterium; however, chromids contain fewer of these essential genes than the chromosome (Egan et al. 2005). Chromids are generally larger (consist of more base pairs) than megaplasmids, although their sizes tend to overlap (diCenzo and Finan 2017). Like megaplasmids, chromids appear to replicate and segregate in coordination with the host's cell cycle (diCenzo and Finan 2017) and they have active partition systems (Planchenault et al. 2020). And, apparently because they possess a partition system, they are able to reside within the volume of the nucleoid (Planchenault et al. 2020), where they may take advantage of entropic flow to segregate copies of themselves into daughter cells (Planchenault et al. 2020). Often chromids have genes for specialized functions that may be beneficial to bacteria under certain environmental conditions (Egan et al. 2005).

Evolution of secondary replicons

It is commonly believed that these three types of secondary replicons represent an evolutionary continuum— from recently acquired DNA (in the form of a plasmid), to DNA that's stably or nearly stably incorporated into the host's genome (in the form of a chromid).

Plasmids are replicons that were apparently obtained fairly recently, from another bacterium or other organism by means of horizontal gene transfer. (Horizontal gene transfer includes a number of processes by which genes are transmitted from one individual to another by some means other than by inheritance from a parent.) Through horizontal gene transfer, bacteria have the potential to rapidly adapt to new or changing environmental conditions—they can instantly obtain genes needed to survive in a new environment from some other organism that already has those genes, rather than wait for the needed genes to evolve over time and over many, many generations through mutations.

A plasmid that contains no genes that benefit a bacterium, will impose a fitness cost on its bacterial host (MacLean and San Millan 2015). A bacterium carrying such a plasmid needs to expend energy and resources to replicate, transcribe and translate the plasmid's genes. Organisms generally don't have unlimited energy and resources. Therefore, a bacterium carrying a parasitic plasmid, will have less energy and resources available to devote to its own growth and reproduction. Consequently, an infected bacterium will not be able to effectively compete with similar (plasmid-free) bacteria, and ultimately the plasmid-carrying bacterium will produce fewer offspring. Thus there is a strong evolutionary tendency for nonbeneficial plasmids to be lost from bacteria over time (unless the plasmids can spread themselves to other bacteria at a rate faster than the bacteria can lose the plasmids).

If on the other hand, a plasmid contains one or more genes that benefit the bacterium— the plasmid functions as a symbiont rather than a parasite—there will be a tendency for the plasmid (or at least the plasmid's beneficial genes) to be retained over time. The beneficial genes may enable the bacterium to outcompete other bacteria or to colonize a novel environment or niche where few or no competitors exist.

When a foreign DNA molecule such as a plasmid is taken into a cell, initially it has a very different nucleotide composition than that of the host cell's chromosome. However, over long periods of time, the foreign DNA becomes increasingly "domesticated" ("ameliorated") so that its nucleotide composition (genetic signature) begins to closely resemble that of the host cell's chromosome (Lawrence and Ochman 1997). For example, the percentage of the bases guanine and cytosine (G+C) making up the DNA, and the dinucleotide relative abundance become more similar to that of the host chromosome (diCenzo and Finan 2017). These changes in the foreign DNA are believed to occur because the host cell has a very different internal environment compared to the environment in which the foreign DNA formerly occurred. The unique environment of the host cell exerts certain directional mutation pressures—the DNA in the cell is affected by mutational biases— so that over time the DNA takes on a characteristic genomic signature (Sueoka 1988, Lawrence and Ochman 1997, Suzuki et al. 2008). Differences in genomic signatures (e.g. percentage of G+C) can be used to estimate how long foreign DNA has resided in a cell (Lawrence and Ochman 1997) or to determine whether bacteria are closely related to one another (Sueoka 1962).

Most scientists believe that chromids evolved from plasmids (i.e., from megaplasmids). Some of the lines of evidence in support of this theory include the plasmid-like replication machinery of secondary replicons (including chromids) [Fournes et al. 2018], as well as evidence from genomic signature data. Chromids have genomic signatures (GC content and dinucleotide relative abundances) that are very similar to those of their respective host chromosomes (e.g., GC content differing by less than 1% [diCenzo and Finan 2017]). Compared to chromids, the genomic signatures of megaplasmids are less similar to those of the chromosome, and the genomic signatures of plasmids are even more different from the chromosome (diCenzo and Finan 2017). Thus the data suggest that chromids have resided in cells for a relatively long evolutionary time period, that megaplasmids have been present in cells for a somewhat shorter period of time, and that plasmids were acquired fairly recently.

For a plasmid to evolve into a chromid, not only does "genomic amelioration" need to occur, but the plasmid must also acquire one or more essential (core) bacterial genes. Although the phenomenon of genomic domestication/amelioration is well known (Suzuki et al. 2008), scientists don't have a clear understanding of how plasmids/chromids acquire essential bacterial genes.

Plasmids and chromosomes may carry other acellular parasites— for example genetic elements known as transposons, which have the ability to "jump" from one DNA molecule to another. (The "jumping" involves excision of the DNA sequence from the original site [or duplication of the DNA sequence] followed by insertion into the new location). The beneficial genes carried by plasmids are often found in transposons or other mobile genetic elements (MacLean and San Millan 2015). Thus, it is possible for genes to move— via mobile genetic elements— from plasmid to chromosome or vice-versa.

It is also known that one or more secondary replicons may fuse with the host chromosome, creating a single replicon from the cointegrated DNA molecules (Guo et al. 2003, Val et al. 2014).[5] The fused DNA molecule (called a cointegrate) can split back into its original replicons, as long as the splits occur at the same sites where the molecules were fused; this appears to be what typically occurs when cointegrates break apart (Guo et al. 2003). However, if the splits occur at other sites— sites within the original replicons rather than between them— new replicons may be created with novel combinations of genes (Guo et al. 2003). Thus if a chromosome and a plasmid fused together and then split apart, it may be possible that one or more core genes from the chromosome could end up on the plasmid (and one or more plasmid genes could end up on the chromosome (Poirion and Lafay 2018).

Another way that an essential core bacterial gene, formerly located on a chromosome, could end up on a plasmid is if the cell acquired a new copy of the gene on the plasmid/chromid (in addition to the existing copy on the chromosome) (diCenzo and Finan 2017). Then if the copy of the gene was lost from the chromosome (e.g., due to a detrimental mutation), the cell could still survive, relying on the copy on the chromid. Gene redundancy could potentially occur if a core gene on the chromosome was duplicated and a copy ended up on a plasmid, or if the cell acquired— through horizontal gene transfer— a plasmid with a copy of the core gene (or a functionally equivalent gene) (diCenzo and Finan 2017).

Multipartite genomes

Bacteria containing two or more large DNA molecules—i.e., a chromosome and one or more megaplasmids and/or chromids—are said to have a "multipartite" genome. Based on available genome data, around 10 percent of bacterial species have multipartite genomes (Harrison et al. 2010, diCenzo and Finan 2017). Bacteria with multipartite genomes include species that interact with eukaryotes— as pathogens or symbionts, species found in polluted environments that are capable of catabolizing pollutants, and species that occur in extreme environments (diCenzo and Finan 2017). However, not all bacteria occupying these niches have multipartite genomes (diCenzo and Finan 2017). Nonetheless, those that do apparently acquired genes from plasmids enabling these bacteria to thrive in new environments. Bacteria found in the same environments that don't have multipartite genomes may have also acquired their niche-specific genes via horizontal gene transfer, but from a different acellular parasite.

Scientists have recently begun to recognize the prevalence of genetic elements known as integrative and conjugative elements (ICEs), which, like plasmids, appear to be important in enabling bacteria to quickly adapt to new niches. Like plasmids, ICEs are capable of spreading copies of themselves from one host cell to another. ICEs spread to new host cells in the same manner as plasmids— via a process known as conjugation. Also like plasmids, ICEs often carry genes that may enhance the fitness of their bacterial hosts. ICEs differ from plasmids in that they normally integrate themselves into the host cell's chromosome (rather than remaining separate from it). However, the distinction between these two types of genetic elements is not always sharp and over time one type may give rise to another (Guglielmini et al. 2011). Scientists have tended to overlook ICEs because they are difficult to identify and delimit from sequenced bacterial genomes (Ambroset et al. 2015, Cury et al. 2017). However, given their apparent importance in microbiology, ICEs are increasingly being studied.

Cell Envelope

The bacterial cell envelope consists of cell membrane, cell wall, and other materials that surround the cytoplasm. The actual layers making up the cell envelope vary depending on the species of bacterium. All cells, at a minimum, have a cytoplasmic membrane. There are two common cell envelope types— gram positive and gram negative. These envelope types have traditionally been determined using a staining technique known as gram staining, which is still used today.

The gram staining procedure was developed over 130 years ago by the Danish doctor Hans Christian Gram (Gram 1884). Dr. Gram noticed that different types of bacteria vary in their ability to retain dyes. He tested whether bacteria retained the dye crystal violet by

  1. Applying the crystal violet dye to a heat-fixed smear of bacteria on a glass slide;
  2. Washing off the excess dye using water;
  3. Applying iodine in order to set or "fix" the dye;
  4. Rinsing with alcohol to wash off the dye; and
  5. Rinsing with water to stop the action of the alcohol.

Dr. Gram found that certain bacteria retained the violet-colored dye after this procedure (the gram-positive bacteria), while other bacteria did not retain the dye (the gram-negative bacteria). [6]

Although Dr. Gram recognized that bacteria could be categorized into two types based on their ability to retain the crystal violet stain, he was unaware of the significance of his finding and did not have knowledge of the physical differences between the two types of bacteria (Sandle 2004).

Today we know that gram-positive- and gram-negative-staining bacteria have very different cell envelopes.

Gram-positive Cell Envelope

Gram Positive

Figure 5. A cell with a classical gram-positive envelope.

A classical gram-positive cell envelope consists of the following layers (listed from interior to exterior):

  1. A cytoplasmic membrane (also referred to as the cell membrane or plasma membrane), consisting of a symmetric bilayer of phospholipids with embedded proteins. This membrane is around 5 to 10 nanometers thick (Prescott et al. 1993). It encloses the cytoplasm and controls the movement of substances into and out of the cytoplasmic compartment. It is also the site of various biochemical processes. The electron transport chains involved in respiration and photosynthesis are located here.

    Bacterial cells maintain an electrochemical gradient across the cytoplasmic membrane. (There is a lower concentration of protons inside the cytoplasmic compartment than outside of it.) A multi-subunit protein complex called ATP synthase is embedded in the cytoplasmic membrane. ATP synthase uses the potential energy of the transmembrane gradient to make ATP (a ready-to-use source of energy for the cell).

    bilayer of cytoplasmic membrane

    Figure 6. The lipid bilayer of the cytoplasmic membrane. The bilayer is symmetrical because both leaflets of the membrane are made up of phospholipids. Hydrophilic glycerol phosphate heads are at the outside edges of the membrane and hydrophobic fatty acid tails are in the interior of the membrane.

  2. Outside of the cytoplasmic membrane is a relatively thick cell wall, containing multiple layers of peptidoglycan, with a total thickness ranging from 30 to 100 nm (Silhavy et al. 2010). In gram positive bacteria the peptidoglycan layer is usually complexed with polymers of teichoic acids, which may constitute a substantial portion of the cell wall's mass (Brown et al. 2013, Silhavy et al. 2010). Two types of teichoic acids are present— wall teichoic acids, which are attached to the peptidoglycan and extend beyond the peptidoglycan layer (to the outermost reaches of the cell wall), and lipoteichoic acids, which are attached to the cytoplasmic membrane and extend into the peptidoglycan layer (Brown et al. 2013). Teichoic acids are negatively charged and cause the gram-positive cell wall to have an overall net negative charge (Prescott et al. 1993, Jordan et al. 2007). These polymers have many functions in the cell, which may vary depending on the species of bacteria (Silhavy et al. 2010). Wall teichoic acids appear to play a role in cell morphology and division, autolytic activity, ion homeostasis, tolerance of high temperatures and osmotic stress, cell adhesion and biofilm formation, and defense against antibiotics (Brown et al. 2013).

    Recently, the gram-positive cell wall has been found to comprise two zones: a high density outer wall zone and a low density inner wall zone (Matias et al. 2005). The outer wall zone corresponds to the peptidoglycan layer of the cell wall, while the inner wall zone consists of a space located between the cytoplasmic membrane and the outer wall zone (Zuber et al. 2006, Matias and Beveridge 2005). The inner wall zone is sometimes referred to as the periplasmic space, although that term is more commonly used to refer to the space/compartment between two membranes (Forster and Marquis 2012).

Gram-negative Cell Envelope

In contrast to the classical gram positive cell, the classical gram negative cell has (1) a relatively thin cell wall, (2) an additional membrane—called the outer membrane— located outside of the cell wall, and (3) a fluid-filled compartment called the periplasmic space or periplasm, which is bounded by the two membranes of the cell.

The layers found in a classical gram-negative cell envelope are listed and described below:

Gram Negative

Figure 7. A cell with a classical gram-negative envelope.

  1. A cytoplasmic membrane (also referred to as the inner membrane, cell membrane, or plasma membrane). It has the same structure, composition, and functions as the cytoplasmic membrane found in gram-positive bacteria.

  2. A relatively thin cell wall, containing one to a few layers of peptidoglycan, with a thickness of 2 to 6 nm (Silhavy et al. 2010, Hoiczyk and Hansel 2000). Unlike the cell walls of gram-positive bacteria, gram-negative cell walls lack teichoic acids (Cheng and Costerton 1977, Prescott et al. 1993). The cell wall is attached to both the cytoplasmic membrane and the outer membrane by membrane-anchored proteins (Kovacs-Simon et al. 2011), with some protein complexes extending across both membranes (Nikaido 1996).

  3. An outer membrane, around 7.5 to 10 nm thick (Salton and Kim 1996), consisting of an asymmetric lipid bilayer. This membrane screens out chemicals that are harmful to the cell and adds structural stability. The inner leaflet of the membrane is made up of phospholipid molecules, while the outer leaflet normally consists of glycolipids— typically lipopolysaccharide (LPS) molecules (Galdiero et al. 2012). Approximately 50 percent of the outer membrane is made up of proteins (Koebnik et al. 2000). Two main types of proteins are found in outer membranes: lipoproteins and transmembrane beta-barrel proteins.

    the outer membrane

    Figure 8. The outer membrane. This membrane is highly asymmetrical, with phospholipids in the inner leaflet (facing the periplasm) and lipopolysaccharide molecules in the outer leaflet (facing the outside environment). Also shown is a porin— a trimeric beta-barrel protein (in yellow) and a lipoprotein (in orange).

    Lipoproteins are found in both gram-negative and gram-positive bacteria and occur in both cytoplasmic and outer membranes (Kovacs-Simon et al. 2011). They may be found in the outer leaflet of the cytoplasmic membrane, the inner leaflet of the outer membrane, and/or the outer leaflet of the outer membrane (Kovacs-Simon et al. 2010). They do not span both layers of the membrane. These proteins have a lipid component, which apparently aligns with and anchors them (via hydrophobic interactions) to the fatty acid tails of the phosopholipid and LPS molecules in the membrane (Kovacs-Simon et al. 2011). Lipoproteins in the outer membrane are involved in cell wall synthesis, secretion systems, and antibiotic efflux pumps (Grabowicz and Silhavy 2017). One type of lipoprotein, called Braun's lipoprotein (murein lipoprotein or Lpp), attaches the peptidoglycan wall to the outer membrane. Additional proteins are apparently involved in a network that links together the cell wall and the outer membrane; together the Braun's lipoprotein and the other proteins in the network appear to be of critical importance in maintaining the integrity of the gram-negative cell envelope (Dramsi et al. 2008).

    Beta-barrel transmembrane proteins are characteristic components of the gram-negative outer membrane. These proteins (sometimes referred to as outer membrane proteins) are not found in the cytoplasmic membrane or in any other cellular membrane (except for the outer membranes of mitochondria and chloroplasts— two eukaryotic organelles that apparently evolved from gram-negative bacteria) (Koebnik et al. 2000, Fairman et al. 2011).

    A common type of beta-barrel protein found in outer membranes are porins. They are so-named because they form pores or channels in the membrane. The pores are filled with water and allow small hydrophilic molecules to diffuse across the membrane (Galdiero et al. 2012). General/nonspecific porins are abundant in outer membranes and allow many different types of ions and polar molecules to pass through (as long as the diffusing entities are small enough to fit through the channel), with some selectivity for either cations or anions depending on the porin (Galdiero et al. 2012). Less common are substrate-specific porins (i.e., porins that only allow a specific molecule to pass through— for example sucrose-specific porins). Substrate-specific porins are expressed at increased levels when needed (their synthesis is induced). These porins have channel interiors with features that make it easier for particular molecules to diffuse into the cell (Forst et al. 1998). Besides their role in nutrient exchange across the outer membrane, porins may also be involved in pathogenicity (Galdiero et al. 2012).

    In addition to porins, other types of beta-barrel proteins found in outer membranes include proteins involved in the active transport of large substrates into the cell (some of these proteins are also known to play a part in signal transduction[7]), proteins involved in the active transport of molecules out of the cell; those involved in the secretion of proteins (primarily virulence factors); ones that function as enzymes; and proteins with structural functions (Kim et al. 2012, Koebnik et al. 2010).

    Unlike the cytoplasmic membrane, the outer membrane is not an energized membrane. No gram-negative bacterium is known to have an electrochemical gradient across its outer membrane and ATP synthase has never been found in bacterial outer membranes (Küper et al. 2010). However, active transport does occur across the outer membrane; the energy required for this transport is ultimately obtained from the electrochemical potential of the cytoplasmic membrane (Braun et al. 1996).

  4. A periplasmic space or periplasm—a compartment located between the cytoplasmic membrane and the outer membrane. (The cell wall extends through this space.) Like the cytoplasm, the periplasm is a viscous fluid-filled compartment. It may constitute 7 to 40 percent of the total cell volume (Kulp and Koehn 2010). Found within this compartment are binding and trafficking proteins, various enzymes (including degradative and detoxifying enzymes), secreted materials, and components for synthesis of the cell wall and outer membrane (Beveridge 1999). The periplasm is devoid of ATP and other molecular energy sources (Kulp and Kuehn 2010).

Although both gram-negative and gram-positive species can be highly pathogenic (depending on the species), gram-negative bacteria have some special characteristics associated with their outer membrane that may enhance their pathogenicity, as summarized below:

  • The outer membrane provides gram-negative bacteria with an extra layer of protection— an additional selective barrier— against antibiotics, toxins, and other harmful substances. In general, gram-negative species are more resistant to antibiotics than gram-positive species, at least in part because of their extra membrane (Silhavy et al. 2010, Nikaido 1996). The outer membrane is impermeable to large charged molecules (Galdiero et al. 2012), and due to the presence of LPS, provides an effective barrier against hydrophobic molecules (Silvahy et al. 2010). Hydrophobic molecules diffuse across the outer membrane about two orders of magnitude slower than across a typical cytoplasmic membrane (Nikaido 2003). This is a benefit because most clinically important antibiotics have some degree of hydrophobicity (Cohen 2004). Although both gram-positive and gram-negative bacteria have efflux pumps that pump antibiotics and other toxins out of the cell, these pumps tend to be more effective in gram-negative bacteria because their cell envelopes are less permeable (Fernández and Hancock 2012)— they can pump out the toxins faster than the toxins can enter the cell.

    lipopolysaccharide

    Figure 9. A lipopolysaccharide (LPS). The Lipid A component contains diglucosamine "heads" (shown in red), which typically have a phosphate group attached at each end (shown in green). It also includes fatty acid "tails" (shown in black). Attached to the outer side of Lipid A is a core polysaccharide (shown in blue) and an O-antigen (shown in pink).

  • Outer membranes contain LPS molecules, which are known endotoxins. The toxin is released when the bacterial cells die and break open. Small amounts of LPS are also released from live cells, as a by-product of the normal, on-going maintenance and renewal of the outer membrane (Wassenaar and Zimmermann 2018) and during cell division (Matsuura 2013). Virtually all gram-negative species contain LPS in their outer membranes.[8] In mammals, LPS stimulates the immune system and can induce fever and inflammation (Wassenaar and Zimmermann 2018). If a gram-negative bacterial infection spreads to the bloodstream, septic shock may occur (due to an overzealous inflammatory response), potentially resulting in organ failure and death (Wassenaar and Zimmermann 2018).[9]

    LPS molecules contain three main components: (1) Lipid A— a phosphorylated diglucosamine molecule with fatty acid tails, which anchors the LPS into the membrane; (2) a "core" oligosaccharide, which is attached to Lipid A; and (3) a relatively long O-antigen polysaccharide, which is connected to the core oligosaccharide and extends out from the cell exterior into the surrounding medium (Raetz and Whitfield 2002, Nikaido 2003, Clifton et al. 2013). The toxic part of LPS is the Lipid A component (Raetz and Whitfield 2002, Wassenaar and Zimmermann 2018, Stewart et al. 2016).[10] Lipid A is the portion of the molecule that is recognized by the innate immune system; thus the presence of lipid A alone in the bloodstream can lead to septic shock (Ramachandran 2014, Raetz and Whitfield 2002). However, the adaptive immune system— which launches its attacks at cells containing specific antigens— usually directs its response based on the composition of the O-antigen (Bryant et al. 2010).

  • The o-antigen is the most variable part of the LPS molecule (Lerouge and Vanderleyden 2001, Matsuura 2013, Stewart et al. 2006). Variability in the o-antigen may enable some bacteria to evade the host's immune response (Raetz and Whitfield 2002, Nikaido 2003), resulting in persistent infections. The immune systems of mammals and other vertebrates are able to distinguish their own cells from those of foreign invaders. They recognize the presence of a foreign invader based on a macromolecule (e.g. a protein or a polysaccharide) that occurs in association with that invader, typically on the surface of the invader's cell. Such a macromolecule—which is recognized by the immune system and elicits an immune response—is called an antigen. The first time a host is exposed to a foreign antigen it can take several days for the immune system to launch a response that specifically targets the pathogen containing that antigen (Alberts et al. 1994). (A nonspecific immune response happens right away but is generally less effective.) However, the immune system retains a memory of the antigen after the infection has passed, and upon subsequent infection with the same pathogen/antigen the immune system is able to quickly launch a strong targeted attack. When an antigen, such as the o-antigen, is highly variable and changes its structure and/or composition frequently it becomes much harder for the immune system to clear the infection. Once the antigen changes its form, the T-cells, B-cells, antibodies, and memory cells that were once specific to that pathogen are no longer effective in fighting the infection. Each time the antigen changes, a new adaptive immune response needs to be launched from scratch.

    In some cases a single population of bacteria may contain cells that express different o-antigens (Lerouge and Vanderleyden 2001). Thus to control a single infection, a host may need to produce immune cells and antibodies specific to each type of o-antigen present.

    Certain pathogenic bacteria may evade the host's immune response by producing o-antigens (or other types of antigens) that closely resemble or "mimic" macromolecules found on the surface of the host's own cells (Matsuura 2013, Lerouge and Vanderleyden 2001). For example, Helicobacter pylori, a species commonly found in the human stomach, may express o-antigens that mimic blood group structures found on the surface of cells in the gastric mucosa (Appelmelk et al. 1996). This phenomenon, known as "molecular mimicry," may play a role in autoimmune disorders (Cusick et al. 2012).

    membrane vesicle formation

    Figure 10. Development and release of a membrane vesicle from an outer membrane.

    membrane vesicle formation

    Figure 10. Development and release of a membrane vesicle from an outer membrane.

    membrane vesicle formation

    Figure 10. Development and release of a membrane vesicle from an outer membrane.

  • All or virtually all gram-negative bacteria are known to produce membrane vesicles from their outer membrane (Beveridge 1999, Kulp and Kuehn 2010)— they are said to "bleb" membrane vesicles off their outer membrane. These vesicles, which are released to the outside environment, may contribute to the pathogenicity of their associated bacteria.[11] Outer membrane vesicles are generally spherical and have a diameter of around 10 to 500 nm (Domingues and Nielsen 2017). They start off as small pockets that form in the outer membrane. The pockets then pinch off to form vesicles (Figure 10). In some cases huge numbers of vesicles may bleb off an outer membrane at once (Beveridge 1999). An outer membrane vesicle contains periplasm (and whatever constituents are present in the periplasm) and is bounded by a lipid bilayer membrane. The membrane surrounding a membrane vesicle closely resembles the bacterium's outer membrane— it contains an inner leaflet of phospholipids, an outer leaflet of LPS molecules, as well as outer membrane proteins (Beveridge 1999). Outer membrane vesicles are enriched in certain proteins and lipids relative to the outer membrane from which they were formed; thus it is likely that they are produced by some specific cellular mechanism rather than from random blebbing of the membrane (Kulp and Kuehn 2010). Researchers have found that greater numbers of outer membrane vesicles are produced in response to increased stresses on the cell envelope (McBroom and Kuehn 2007).

    Outer membrane vesicles may serve as vessels to carry substances to other cells. These substances may include enzymes that break down cell walls, toxins (including the endotoxin LPS), enzymes that break down antibiotics, or even plasmids that contain genes for antibiotic resistance (Beveridge 1999, Kadurugamuwa and Beveridge 1996, Ciofu et al. 2000, Dorward et al. 1989). Consequently, membrane vesicles are capable of attacking (dissolving the cell walls of) bacterial competitors (presumably to release nutrients under poor growth conditions), they may transport and release toxins that weaken host defenses, deliver antibiotic-degrading enzymes to bacterial cells that may lack such enzymes (or deliver these enzymes to host tissues to protect infecting bacteria), and they may spread genes for antibiotic resistance (Beveridge 1999).

    The means by which outer membrane vesicles deliver substances to other cells depends on the type of recipient cell. Outer membrane vesicles attack gram-positive cells by binding to the cell wall. The membrane vesicles then break open and release enzymes that degrade the recipient's cell wall (Kadurugamuwa and Beveridge 1996). When they contact the outer membrane of gram-negative bacteria, membrane vesicles fuse to the membrane, releasing the contents of the vesicles into the recipient cell's periplasm (Beveridge 1999, Kadurugamuwa and Beveridge 1996). Outer membrane vesicles deliver their contents to eukaryotic cells either by fusing to the cell's membrane (and releasing the contents into the cell's cytoplasm) (Jäger et al. 2015) or they may be taken into a cell by endocytosis (Vanaja et al. 2016, Sharp et al. 2011, Furuta et al. 2009).

    In addition to transporting and delivering substances, outer membrane vesicles may absorb or adsorb toxins or other harmful agents, protecting bacterial cells from chemical or viral attack (Manning and Kuehn 2011, Ciofu et al. 2000). They may also play a role in biofilm formation and thus help bacteria colonize surfaces and survive in harsh environments (Schooling and Beveridge 2006, Yonezawa et al. 2009). Additionally, membrane vesicles may be used to package and remove harmful substances from the cell envelope, including misfolded proteins (McBroom and Kuehn 2007).

Monoderm vs. Diderm Bacteria

Bacteria that have one membrane (a cytoplasmic membrane only) are said to be monoderm (having "one skin"), while bacteria with two membranes (a cytoplasmic and an outer membrane) are described as being diderm (having "two skins"). The gram-staining technique is a quick and easy way to determine whether a bacterium is monoderm (which typically stain gram positive) or diderm (which typically stain gram negative), but the technique is not foolproof. Not all bacteria have a classicial gram-positive or a classical gram-negative cell envelope. Some monoderm bacteria have atypically thin cell walls and stain gram negative, and some diderm bacteria have unusually thick cell walls and stain gram positive. Culture conditions may also affect the gram stain result. For example, monoderm bacteria obtained from cultures that are old or are otherwise stressed have a tendency to stain gram negative (Beveridge 1990, Benson 1990). Some bacteria— even under ideal conditions— stain in a variable manner. They may stain gram positive or gram negative depending on their phase of growth (Beveridge 1990). Thus, in order to determine with certainty whether a bacterium is monoderm or diderm, an alternative method should be used (such as one involving transmission electron microscopy).

Which came first — monoderm or diderm bacteria?

An interesting question in biology is whether the ancestral bacterium was monoderm or diderm. In short, we don't know. Reconstructing phylogenies for ancient events is notoriously difficult (Gribaldo et al. 2010) due to increased noise, decreased evolutionary signal, and consequently a greater number of artifacts confounding the analysis (Gouy et al. 2015). Scientists cannot determine with certainty how the tree of life is rooted— for example, whether Bacteria evolved first and gave rise to Archaea or whether an ancestral cell of some other type (which no longer exists) gave rise to both Bacteria and Archaea. Scientists also can't say with confidence how Bacteria are rooted or identify the lineages that form its earliest branches (Antunes et al. 2016).

Most bacteria alive today are diderms. The known exceptions include bacteria from three phyla: Actinobacteriota, Chloroflexota, and Firmicutes. The Actinobacteriota and the Firmicutes are unusual among the bacterial phyla in that they contain both monoderm and diderm species. Most of the taxa making up these phyla are monoderm. However, one suborder of the Actinobacteriota (the Corynebacterineae) consists of bacteria with an outer membrane of mycolic acid, and at least two classes of Firmicutes contain bacteria with outer membranes of LPS— the Negativicutes and the Halanaerobiia (Halanaerobiales).

Results of a study conducted by Antunes et al. (2016) strongly suggest that, at least in the case of the Firmicutes, the ancestral bacterium was diderm (with an outer membrane containing LPS), and in multiple instances, monoderm bacteria evolved from diderm ancestors by losing the outer membrane. The authors found that both the Negativicutes and the Halanerobiia are more closely related to monoderm relatives than they are related to each other. Yet representatives from both of these diderm classes contain a large cluster of genes associated with the outer membrane that is highly conserved among the two groups— thus it appears that this large gene cluster was inherited from a common (diderm) ancestor.[12]

It is very unlikely that the Negativicutes and/or the Halanerobiia obtained the outer membrane gene cluster via horizontal gene transfer. The outer membrane is a complex structure, and given the size of the gene cluster in these taxa, it would be quite difficult for a cell to successfully incorporate all of the genes at once, losing other pre-existing genes as necessary, to form a functional outer membrane (Antunes et al. 2016). It would be much easier for a cell to lose an outer membrane (or other complex structure in general) than to gain one. For example, this could possibly occur if a bacterium had a mutation that resulted in a thicker cell wall, which in turn could affect the anchoring of the outer membrane to the cell wall (Cavalier-Smith 2006), or alternatively, mutations could occur in a gene for an anchoring protein, resulting in a defective protein and the destabilization and detachment of the outer membrane (Antunes et al. 2016).

To summarize: over the course of the evolution of bacteria, it is apparent that in multiple instances monoderm cells evolved from diderm ancestors (at least in the case of the Firmicutes). But we still don't know with any certainty whether the ancestor shared by all bacteria was diderm or monoderm.

In all likelihood, the first cells to exist on Earth were monoderm. Monoderms have a simpler cell envelope, and evolution, over the long term, tends to build more complex structures from simpler pre-existing designs. As noted by Canadian researcher Radhey Gupta (Gupta 2011), it is difficult to imagine any simple model where a cell with two membranes could have evolved without the initial development of a cell with a single membrane.

How could a monoderm cell become diderm?

Endospore formation

Figure 11. Endospore formation in a gram-negative bacterium. The outer membrane of the original cell is shown in orange. (Areas containing peptidoglycan are shaded in light orange.) The cytoplasmic membrane, and the membranes derived from the cytoplasmic membrane, are shown in blue. (Areas containing or formerly containing cytoplasm are shaded in light blue.) Endospore formation begins with a cell replicating and segregating its DNA. The cytoplasmic membrane then begins to invaginate near one end of the cell (A), forming a septum (B), which separates the cell into two unequal parts, each containing a copy of the genetic material. Instead of splitting apart into two daughter cells, the cytoplasmic membrane of the larger part grows around the smaller part (C, D), surrounding it entirely (E), forming a double-membraned "forespore" within the mother cell. Upon maturation of the spore, the mother cell breaks open, releasing the spore (F).

Thus, it is likely that at some time in the past diderm cells evolved from a monoderm ancestor. This transition is not as simple as losing an outer membrane. Besides acquiring all the genes needed for an outer membrane, the cell would need to undergo some process to form that membrane. Cell membranes aren't synthesized from scratch— they're inherited (Beisson 2008). So it doesn't matter if all the genes are present to build a membrane. There has to be some pre-existing membrane structure, which then can be modified and adapted as necessary from the gene products that are produced by the cell.

So, as far as we know, the outer membrane must have been formed from some other membrane. But how? Thanks to research performed by Elitza Tocheva and colleagues (Tocheva et al. 2011) we know that outer membranes can be formed from cytoplasmic membranes as the result of a process known as sporulation (or more specifically endospore formation).

Some bacteria are able to produce spores when growth conditions become unfavorable. Spores are durable dormant cells that are capable of surviving harsh conditions that would kill normal, biologically active cells. Once environmental conditions become favorable again, a spore "germinates" and a biologically active cell grows out from it.

Endospores are the type of spores produced by bacteria from the phylum Firmicutes. They are the most durable cells known on earth— being capable of surviving heat, UV and gamma radiation, desiccation, and oxidizing agents (Nicholson et al. 2000). Endospores can successfully germinate after remaining dormant for extremely long periods of time— sometimes thousands of years or longer (Gest and Mandelstam 1987, Kennedy et al. 1994, Potts 1994).

Spore formation occurs primarily in monoderm bacteria, although some diderm species are known to form spores. Endospores of monoderm and diderm bacteria have the same structure and they are formed in the same manner (Tocheva et al. 2016). Sporulation appears to have evolved from a cell replication event that went awry (Tocheva et al. 2011). It starts off like cell replication, with the bacterium replicating and segregating its DNA, and forming a septum between the two halves of the cell. However, instead of dividing into two daughter cells, the membrane of one half (referred to as the mother cell) begins growing around the other half (the soon to be spore), eventually surrounding it entirely. The maturing spore, located within the mother cell, is enclosed by two membranes, both of which were derived from the cytoplasmic membrane of the original mother cell. Upon maturation of the spore, the mother cell breaks open, releasing the spore (Figure 11).

When the spore of a monoderm bacterium germinates, the outer membrane of the spore is lost and the inner membrane of the spore becomes the cytoplasmic membrane of the outgrowing cell. However, in a diderm bacterium, both of the spore's membranes are retained and they become the cytoplasmic and outer membranes of the emerging cell (Tocheva et al. 2011).

Tocheva et al. 2011 examined sporulation in the diderm Acetonema longum. Through electron cryotomography imaging, they observed that the spore of this bacterium was surrounded by two membranes that were derived from the cytoplasmic membrane of the mother cell. They also found that following germination, both membranes of the spore were retained to become the cytoplasmic and outer membranes of the outgrowing cell.[13] Acetonema longum has a classical gram-negative cell envelope, with an outer membrane containing LPS and typical outer membrane proteins. Therefore, at some point in the sporulation/germination process, the outer cytoplasmic membrane of the spore was transformed into a bona fide outer membrane.

Since the spores of both monoderm and diderm bacteria are surrounded by two membranes and have the same overall structure, it's easy to image a scenario where an ancient sporulating monoderm could have given rise to a diderm. There would simply need to be some mutation(s) causing the outer membrane of the spore to be retained, so that the outgrowing cell would have an outer membrane. Likewise a monoderm species could possibly evolve from a sporulating diderm if a mutation occurred that resulted in the loss of the outer spore membrane.

The first diderms to evolve likely had an outer membrane that was very similar in composition to the cytoplasmic membrane. But over time, diderm bacteria would have acquired genes that enhanced the function of the outer membrane, leading to the present day outer membranes (Tocheva et al. 2011). There are at least two unique types of outer membranes among the known bacteria— the LPS outer membrane (characteristic of most diderm bacteria), and the mycolic acid-based outer membrane (found in some bacteria from the Phylum Actinobacteriota). Thus it is possible that cells with outer membranes evolved from monoderm progenitors on at least two occasions (Sutcliffe and Dover 2016).

The Cell Wall and Osmosis

A bacterium's cell wall is a porous protective barrier that helps to maintain the shape of the cell and prevent the cell from bursting (lysing) in hypotonic environments. The cell wall surrounds the semi-permeable cytoplasmic membrane, which encloses the contents of the cell. Water flows freely across the membrane, but the membrane does not allow the free passage of many dissolved particles. Because the solute (particle) concentration inside a cell is typically higher than that of the surrounding medium, there is a tendency for water to flow (via osmosis) from the environment into the cell. Water will continue to flow into the cell until the solute concentration is the same on both sides of the membrane or until the force or pressure of water flowing into the cell is counterbalanced by the force or pressure inside the cell. Without a cell wall or other reinforcement of the cell membrane, the influx of water into a cell typically causes the cell to swell, stretching and stressing the membrane until it ruptures and the cell lyses. However, when a membrane is reinforced by a cell wall, the cell wall prevents the membrane from expanding when water enters the cell. Instead, pressure builds up inside the cell from the inflow of water, but the cell wall is strong enough to resist the pressure and the cell is able to hold its shape. Because of the need to counterbalance the force of osmosis, the pressure inside some gram-positive bacterial cells may be as high as 20 atmospheres or 300 pounds per square inch (Prescott et al. 1993, Mitchell and Moyle 1956). Gram-negative bacteria, because of their thinner cell walls, are believed to support much lower internal pressures (Koch and Doyle 1986) and thus lower internal solute concentrations. Besides tolerating a range of internal pressures, bacteria can also manage osmotic stresses by accumulating or releasing solute particles (Wood 2015).

Groups of Bacteria

For asexual organisms, such as bacteria, the classical definition of a species— a group of individuals that interbreed and produce fertile offspring— does not apply. In order for a scientist to determine whether a bacterium belongs to a particular species (or any other taxonomic group), he or she has to utilize arbitrary criteria to establish cut-off points between the different groups.

Currently there is no standard definition for what constitutes a species of bacterium. The assignment of bacterial species to higher taxonomic ranks (e.g., genus, family, etc.) is also problematic and lacks a consistent, standardized approach. Bacteria have been classified based on phenotypic characteristics, genomic characteristics, or a combination of the two. Some of the methods that have been used to classify bacteria are listed below.

  • Variation in DNA base composition (mol% G+C content) (Fournier et al. 2006).
  • DNA-DNA hybridization. (Wayne et al. 1987, Konstantinidis et al. 2006).
  • Gene sequence similarity using select DNA markers (including 16S rRNA) (Chan et al. 2012, Varghese et al. 2015)
  • Whole-genome sequencing (Church et al. 2020)
  • Presence of certain metabolites, morphological features, growth requirements, or other biochemical or enzymological characters (Varghese et al. 2015, Chan et al. 2012).

Compared to macroscopic organisms, bacteria are very challenging to study and classify. They're generally invisible to the unaided eye, they're found in all sorts of habitats, and the vast majority of them can't be successfully isolated and cultured. To obtain an idea of the total biodiversity of bacteria on Earth, researchers need to sample all imaginable habitats and all types of environmental media (e.g., soils, waters, tissues, etc.). Fortunately, with recent advances in molecular biology, scientists can now identify genomes of bacteria from environmental samples— without having to culture the organisms. So now, new taxa of bacteria are being identified at a rapid rate.

Bacterial taxa that have no cultured representatives (i.e., no member taxa that have been isolated and grown in pure culture) are referred to as "candidate" taxa and are written with the Latin word candidatus preceeding the taxon name (e.g., candidate phylum [or division] Zixibacteria, also referred to as Candidatus Zixibacteria, or abbreviated Ca. Zixibacteria). Note when Candidatus is used with a species name, the word Candidatus is italicized but the species name is not, e.g., Candidatus Scalindua brodae. Currently there are more candidate phyla of bacteria than there are bacterial phyla with cultured representatives.

The table below provides information on various phyla of bacteria. The list shown is a subset of the phyla identified in the Genome Taxonomy Database (GTDB 2018). Whereas the GTDB contains all identified phyla, both cultured and uncultured, the list below includes only those phyla that have at least one cultured representative. One candidate phylum— Entotheonellota— is included in the list below. Although Entotheonellota contains a species that has been cultured, that species (a symbiont) has not been grown in pure culture (only in mixed culture with its host), and therefore is considered to have candidate status.

The classification of bacteria is a work in progress. Hopefully over time, classification methodologies will become more standardized and universally applied. It is anticipated that many, many more taxa of bacteria will be discovered and described in the future, and the taxonomic placement of known bacteria will change somewhat as more data become available and as phylogenetic methods are improved.


Phyla of Bacteria
Phylum Representative Classes Envelope Type Examples Notes
Acidobacteriota (Acidobacteria) Acidobacteriae, Aminicenantia, Blastocatellia, Holophagae, Luteitaleia, Thermoanaerobaculia Gram-negative; contain genes for LPS (Antunes et al. 2016) Acidobacteria capsulatum— first isolate to be cultured from this phylum. It was obtained from an acid mine drainage and thus was named Acidobacteria to reflect its preference for acidic environments; Candidatus Chloroacidobacterium thermophilum— an uncultured phototrophic bacterium, which possesses a Type I (iron-sulfur type) photochemical reaction center. It was identified from an alkaline hot spring. Members of this group occur in diverse habitats and are especially common and abundant in soils and sediments. They are very difficult to grow in culture— most have not been cultured. Despite their name, only some members of this phylum appear to occur in and be adapted to acid environments.
Actinobacteriota (Actinobacteria) Actinobacteria, Acidimicrobiia, Coriobacteriia, Rubrobacteria, Thermoleophilia Gram-positive (Monoderm). Except bacteria from suborder Corynebacterineae (under class Actinobacteria, order Corynebacteriales) have an outer membrane of mycolic acids, which is attached to the peptidoglycan wall by the polymer arabinogalactan. Species from this phylum that have been sequenced lack genes for LPS (Antunes et al. 2016) Streptomyces griseus— source of the antibiotic Streptomycin; Frankia alni— nitrogen-fixing symbiont in the roots of alders; Mycobacterium tuberculosus— causes tuberculosis; Corynebacterium diphtheriae— causes diptheria; some other species of Corynebacterium are commonly found in arm pits and are known to cause body odor. Many are filamentous. Some reproduce by sporulation. Includes free-living taxa in soils and water, pathogens, symbionts of plants, and commensal members of the gut microbiota. They are the source of most of our naturally-derived antibiotics. The Actinobacteriota are often referred to as the high G+C gram-positive bacteria (i.e., their DNA is typically made up of a high proportion of guanine and cytosine nucleotides). DNA with high G+C content is considered to be energy demanding and nitrogen demanding (Liu et al. 2016).
Aquificota (Aquificae) Aquificae, Desulfurobacteriia Gram-negative; contain genes for LPS (Antunes et al. 2016) Aquifex aeolicus— a chemolithoautotroph and one of the most thermophilic bacteria known— grows at 95°C (Deckert et al. 1998). Thermophilic and hyperthermophilic taxa. Non-spore forming.
Armatimonadota (Armatimonadetes) Chthonomonadetes, Fimbriimonadia Gram-negative Chthonomonas calidirosea— an uncommon thermophilic soil bacterium. It contains genes for LPS (Antunes et al. 2016). This phylum has not yet been well characterized.
Bacteroidota (Bacteroidetes) Bacteroidia, Chlorobia, Ignavibacteria, Kapabacteria, Kryptonia, Rhodothermia Gram-negative; many (but not all) taxa contain genes for LPS (Antunes et al. 2016). Chlorobaculum tepidum—a green sulfur bacterium, and Bacteroides fragilis— a normal inhabitat of the human gut, which can become pathogenic if it escapes the gut. Found in diverse habitats. Includes the anoxygenic photosynthetic green sulfur bacteria, which have a Type I (iron-sulfur type) photochemical reaction center; thermophilic taxa; and taxa found in the gastrointestinal tract of animals. Along with the Firmicutes, the Bacteroidota dominate the bacterial flora in our gut. They are known to be particularly adept at degrading complex polysaccharides, which otherwise can't be broken down by the host organism.
Bdellovibrionota Bacteriovoracia, Bdellovibrionia, Oligoflexia Gram-negative Bdellovibrio bacteriovorus— a predatory bacterium known to occur in terrestrial and aquatic habitats and in the human gut. Includes predatory bacteria, which attack and feed on larger, typically gram-negative bacteria, invading their periplasm. Also includes some filamentous taxa found in aquatic and terrestrial ecosystems, and a pleomorphic species found in marine coral.
Caldisericota (Caldiserica) Caldisericia Gram-negative Caldisericum exile— an anaerobic, thermophilic, thiosulfate-reducing bacterium. It lacks genes for LPS (Antunes et al. 2016). This phylum has not yet been well characterized.
Calditrichota (Calditrichaeota) Calditrichia Gram-negative Caldithrix abyssi— the first cultivated representative of the phylum. It is anaerobic and thermophilic. This phylum has not yet been well characterized.
Campylobacterota Campylobacteria, Desulfurellia Gram-negative Helicobacter pylori— found in the lining of the stomach in more than half of the world's people; Campylobacter spp.— one of the most common causes of food poisoning and diarrhea in humans. This is a new phylum that combines the former phyla of Epsilonproteobacteria and Desulfurellia. It consists of species found in the gastrointestinal tract of animals, as well as many thermophilic taxa, including thermophilic anaerobic, sulfur-reducing bacteria.
Chloroflexota (Chloroflexi) Chloroflexia, Anaerolineae, Dehalococcoidia, Ktedonobacteria Atypical monoderm bacteria; many stain gram negative. Cell wall lacks peptidoglycan or has atypical peptidoglycan. Species that have been sequenced lack genes for LPS (Antunes et al. 2016). Chloroflexus aurantiacus— an anoxygenic photosynthetic green non-sulfur bacterium, which posseses a Type II (quinone-type) reaction center; Dehalococcoides mccartyi (along with other members of Class Dehalococcoidia)— considered very important for groundwater remediation because of their ability to break down toxic chlorinated organic compounds. Mostly gliding, filamentous bacteria. This phylum includes phototrophs, many thermophiles, and bacteria that respire chloro-organic compounds.
Chrysiogenetota (Chrysiogenetes) Chrysiogenetes Gram-negative; contain genes for LPS (Antunes et al. 2016). Chrysiogenes arsenatis— an anaerobic arsenate-respiring bacterium; and Desulfurispirillum indicum— an anaerobic selenate- and selenite-respiring bacterium This phylum has not yet been well characterized.
Coprothermobacterota Coprothermobacteria Gram-negative Coprothermobacter platensis and C. proteolyticus— both anaerobic thermophilic protein-digesting bacteria isolated from agroindustrial waste treatment systems. This phylum has not yet been well characterized.
Cyanobacteriota Cyanobacteriia, Sericytochromatia, Vampirovibrionia Atypical diderm— the Cyanobacteriia have a peptidoglycan layer that is much thicker than most gram-negative bacteria, ranging from 10 to 700 nm depending on the species; the LPS in their outer membrane often lacks ketodeoxyoctonate, and their outer membrane contains components not usually found in gram-negative bacteria (Hoiczyk and Hansel 2000). The Sericytochromatia and the Vampirovibrionia are diderms but sufficient information is not yet available to know whether they are atypical diderms. Nostoc punctiforme— a photosynthetic, nitrogen-fixing filamentous species that may be free-living or occur in symbiosis with certain plants, including the liverwort Blasia pusilla, hornworts such as Anthoceros punctatus, the angiosperm Gunnera, and cycads such as Macrozamia; Microcystis aeruginosa— a photosynthetic species known to form harmful blooms, typically in nutrient-enriched fresh water. It often produces the toxin microcystin, which is known to cause liver damage in mammals and may also cause damage to the kidneys, heart, reproductive system, and lungs; Vampirovibrio chlorellavorus— a nonphotosynthetic bacterium that is an obligate predator of the alga Chlorella vulgaris. This phylum includes one group of photosynthetic bacteria—class Cyanobacteriia (also referred to as Oxyphotobacteria or crown-group Cyanobacteria), and two groups of nonphotosynthetic bacteria—class Sericytochromatia and class Vampirovibrionia (also referred to as Melainabacteria). The Cyanobacteriota have both a Type I (iron-sulfur type) and a Type II (quinone-type) photochemical reaction center, known as Photosystem I and Photosystem II, respectively. They are the only bacteria known to perform oxygenic photosynthesis. Many also fix nitrogen. An entrapped cyanobacterium gave rise to the chloroplast organelle found in green plants and eukaryotic algae. While many members of the Cyanobacteriia are well known, currently there are no cultured representatives of the Sericytochromatia and only one cultured representative of the Vampirovibrionia (Vampirovibrio chlorellavorus). Sericytochromatia genomes have been identified from samples collected from a coal bed methane well, a bioreactor, and subsurface groundwater; Vampirovibrionia genomes have been recovered from the human gut, koala gut, wastewater treatment plants, subsurface groundwater, lake water, and the gut of termites (Soo et al. 2017, Soo et al. 2014, Utami et al. 2018). The genomes examined from Sericytochromatia and Vampirovibrionia lack genes for phototrophy and carbon fixation (Soo et al. 2017).
Deferribacterota (Deferribacteres) Deferribacteres Gram-negative; contain genes for LPS (Antunes et al. 2016). Flexistipes sinusarabici— a thermophilic halophile isolated from the Atlantis Deep brines of the Red Sea—an extreme environment characterized by high heat, high salinity and high concentrations of heavy metals; and Mucispirillum schaedleri—lives in the gut of rodents and other animals. Generally anaerobic, non-spore-forming, heterotrophic bacteria. Phylum includes mesophiles, thermophiles and thermohalophiles.
Deferrisomatota Deferrisomatia Gram-negative Deferrisoma camini— an anaerobic thermophile isolated from a deep-sea hydrothermal vent. This phylum has not yet been well characterized.
Deinococcota (Deinococcus-Thermus) Deinococci Diderm. Some species contain genes for LPS, others do not (Antunes et al. 2016). Have an S-layer. Deinococcus spp. stain gram positive due to a thick peptiodoglycan layer. Other taxa in this phylum stain gram negative. Deinococcus radiodurans— a species that is highly resistant to ionizing radiation, desiccation, UV radiation and oxidative stress. Phylum includes taxa highly resistant to radiation and desiccation (Deinococcus spp.) and thermophiles (e.g., Thermus spp.)
Dependentiae Babeliae Gram-negative Babela massiliensis—an obligate intracellular parasite of the amoeba Acanthamoeba castellanii; and Chromulinavorax destructans— a virus-like bacterium that infects the predatory flagellate Spumella elongata. C. destructans appears to be inactive until taken up by a host cell, is apparently dependent on its host for replication, and appears to contain no complete metabolic pathways (Deeg et al. 2019). Widespread parasites of eukaryotic microbes.
Desulfobacterota (Thermodesulfobacteria) Desulfarculia, Desulfobaccia, Desulfobacteria, Desulfobulbia, Desulfofervidia, Desulfomonila, Desulfovibrionia, Dissulfuribacteria, Syntrophia, Syntrophobacteria, Thermodesulfobacteria Gram-negative; contain genes for LPS (Antunes et al. 2016). Desulfamplus magnetovallimortis— an anaerobic, sulfate-reducing magnetotactic bacterium; and Desulfotalea psychrophila— a sulfate-reducing species from Arctic sediments capable of growing at temperatures below 0°C. This phylum includes some taxa that were formerly classified under the former phylum of Deltaproteobacteria. It contains many anaerobic, sulfate-reducing bacteria and encompasses a broad range of organisms, including taxa that are mesophilic, thermophilic, psychrophilic, and halophilic, species found in the guts of animals, heterotrophs and chemoautotrophs.
Desulfobacterota_B Syntrophorhabdia Gram-negative (Diderm) Syntrophorhabdus aromaticivorans— a fermenting anaerobe that degrades phenol to acetate with the help of a hydrogen- (and/or formate)-scavenging methanogen. (A methanogen is a methane-producing archaeon.) This phylum has not yet been well characterized. The name of the representative class (and the genus name for the species given as an example) is derived from the word syntrophy, meaning "cross feeding"; it refers to the phenomenon of one species living off of the products of another. In the example given, S. aromaticivorans depends on the products of a methanogen in order to be able to metabolize phenol.
Desulfuromonadota Desulfuromonadia Gram-negative (Diderm) Geobacter sulfurreducens—An anaerobe that obtains energy by oxidizing organic substances in sediment. It uses a metal oxide in the outside environment (Fe[III] oxide) as its terminal electron acceptor in respiration, and transfers electrons to the metal oxide via "nanowires." Since G. sulfurreducens will also transfer electrons to an electrode (an anode) buried in the sediment, this species (and others like it) can be used in microbial fuel cells to generate electricity from aquatic sediments. This phylum includes some taxa that were formerly classified under the former phylum of Deltaproteobacteria. It generally consists of anaerobes that respire sulfur and/or metals (i.e., sulfur- and metal-reducers) and fermenters.
Dictyoglomota (Dictyoglomi) Dictyoglomia Gram-negative (Diderm); contain genes for LPS (Antunes et al. 2016). Dictyoglomus thermophilum and Dictyoglomus turgidum— both are anaerobic, extremely thermophilic bacteria isolated from hot springs. This phylum has not yet been well characterized
Elusimicrobiota (Elusimicrobia) Elusimicrobia, Endomicrobia Diderm. Contain genes for LPS (Antunes et al. 2016). Elusimicrobium minutum— an anaerobic fermenting ultramicrobacterium isolated from the hindgut of a scarab beetle larva; and Endomicrobium proavitum— a free-living, anaerobic ultramicrobacterium isolated from the gut of a termite. Includes intracellular symbionts of eukaryotic flagellates found in the guts of termites, and free-living bacteria found in the guts of termites and other insects.
Entotheonellota (Candidatus Tectomicrobia) Entotheonellia Gram-negative Candidatus Entotheonella palauensis Entotheonella are filamentous symbionts of the marine sponge Theonella swinhoei. Based on their genomes, they apparently have the potential to form endospores.
Fibrobacterota (Fibrobacteres) Chitinivibrionia, Fibrobacteria Gram-negative (Diderm); contain genes for LPS (Antunes et al. 2016). Chitinispirillum alkaliphilum— a haloalkaliphilic chitin-utilizing anaerobe isolated from sediments from hypersaline alkaline lakes; and Fibrobacter succinogenes— an important member of the microbial flora in the rumen of large herbivorous mammals; it degrades cellulose, which is otherwise indigestible by the host, and, along with other microbes in the rumen, contributes to the production of volatile fatty acids—its host's primary energy source. [Note: a ruminant obtains 100% of its protein from digesting its gut microorganisms— it receives no protein from the plant material it ingests, since all of those proteins get broken down and fermented in the rumen (Fenchel and Finlay 1995).] This phylum contains polymer-degrading bacteria, some of which are found in the guts of animals.
Firmicutes Alicyclobacillia, Bacilli, Bacilli_A, Desulfuribacillia Monoderm (generally Gram-positive). Bacteria from the orders Acholeplasmatales and Mycoplasmatales (Class Bacilli) are monoderm but lack a cell wall. Kyrpidia tusciae—a thermoacidophilic, hydrogen-oxidizing facultative chemolithoautotroph isolated from ponds in a volcanically-active area; Lactobacillus acidophilus— found in the human gut, it produces lactic acid from the fermentation of carbohydrates and it is commonly added to yogurt as a probiotic; Streptococcus pyogenes— a human pathogen known to cause sore throat, skin infections, scarlet fever, and streptococcal toxic shock syndrome; and Desulfuribacillus alkaliarsenatis— an anaerobic, sulfur and arsenate-reducing haloalkaliphilic bacterium isolated from sediments from alkaline lakes. This phylum includes free-living taxa as well as pathogenic and commensal species. Some are spore-forming. They are known as the low G+C gram-positive bacteria.
Firmicutes_A Clostridia, Mahellia, Thermoanaerobacteria, Thermovenabulia Many are gram-positive, but with several exceptions. Exceptions include Caldicoprobacter faecalis and Thermosediminibacter oceani, which stain gram-negative, and Thermovenabulum gondwanense, which stains gram-positive but appears to have a gram-negative cell envelope. Clostridium botulinum— an anaerobic spore-forming, neurotoxin-producing species that causes botulism; Caldicoprobacter faecalis— an anaerobic thermophilic fermenter isolated from sewage sludge; Mahella australiensis— an anaerobic spore-forming thermophilic species isolated from an oil well; and Thermovenabulum gondwanense— an anaerobic thermophile isolated from microbial mats from thermal waters. This phylum includes some taxa formerly classified under the former phylum of Deltaproteobacteria. They are generally anaerobic. Most form spores. Many are thermophilic.
Firmicutes_B Dehalobacteriia, Desulfitobacteriia, Desulfotomaculia, Moorellia, Peptococcia, Syntrophomonadia, Thermincolia Generally stain gram-positive. Syntrophobotulus glycolicus stains gram-negative but has a gram-positive-type cell envelope (monoderm). Heliobacterium modesticaldum— an anaerobic, thermophilic photoheterotrophic nitrogen-fixing bacterium with a Type I (iron-sulfur type) photochemical reaction center; it is missing only one gene needed for autotrophic carbon fixation (Tang et al. 2010); Candidatus Desulforudis audaxviator— a thermophilic, nitrogen-fixing chemoautotroph whose genome was identified from fracture water collected from a gold mine. This species was found to be the sole member of a deep subsurface ecosystem (2.8 kilometers below the earth's surface); and Carboxydothermus hydrogenoformans— an anaerobic thermophilic, chemoautotroph isolated from a hydrothermal swamp. It uses carbon monoxide as its carbon and energy source and it uses water as an electron acceptor, converting water to hydrogen gas. Anaerobic bacteria, many of which produce spores. Includes chemoheterotrophs that respire and/or ferment, photoheterotrophs, and chemoautotrophs.
Firmicutes_C Negativicutes Diderm. Contains genes for LPS (Antunes et al. 2016). Anaerosporomusa subterranea— an anaerobic spore-forming fermenter isolated from a subsurface sediment core containing weathered rock; and Allisonella histaminiformans— a bacterium that ferments the amino acid histidine, producing the inflammatory agent histimine as a waste product. This species grows in the rumen of cows and horses that are given certain types of feed (e.g., alfalfa but not timothy hay), and may be associated with rumen disorders and laminitis (inflammation of hooves). Members of this phylum are anaerobic or microaerophilic fermenters. Some form spores but many do not. Several species have been isolated from humans and other animals; some are known to cause infections.
Firmicutes_D Dethiobacteria, Natranaerobiia, Proteinivoracia Monoderm bacteria. Stain gram positive except for Anaerobranca spp., which stain gram-negative due to an atypically thin cell wall. Dethiobacter alkaliphilus— an anaerobic, haloalkaliphilic, sulfide-producing facultative chemolithoautotroph isolated from sediments of a hypersaline soda lake. This species has a Na+-coupled vacuolar (VoV1-type) ATP synthase; Natranaerobius thermophilus— an anaerobic, halophilic, alkaliphilic, thermophilic fermenter isolated from sediment of alkaline, hypersaline lakes. Generally anaerobic, alkaliphilic and halophilic. Some are thermophilic. Some produce endospores.
Firmicutes_E Sulfobacillia, Symbiobacteriia, Thermaerobacteria Described as gram-positive or as staining gram-negative but having a gram-positive-type (monoderm) cell envelope. Symbiobacterium thermophilum— a syntrophic thermophilic bacterium isolated from compost. It is dependent on the metabolic products of another bacterium (Geobacillus sp.); and Thermaerobacter marianensis— an extremely thermophilic aerobic heterotroph, isolated from the Mariana Trench at an ocean depth of around 11,000 meters below sea level. Includes thermophiles and acidophiles. They are heterotrophic or mixotrophic. Some are strict aerobes, others are facultative anaerobes or microaerophiles. Some produce spores.
Firmicutes_F Halanaerobiia Gram-negative (diderm); contain genes for LPS (Antunes et al. 2016). Haloarsenatibacter silvermanii— a facultative chemoautotrophic arsenate-respiring extremophile isolated from a salt-saturated, arsenic-rich, alkaline lake; and Halanaerobium hydrogeniformans— an extremophile isolated from a highly haloalkaline lake that produces hydrogen gas as an end product of fermentation. It is a species of potential interest in the biofuel industry, because its ability to produce hydrogen under high pH/high salinity conditions could potentially eliminate some steps in the production process (Begemann et al. 2012). Generally anaerobic halophiles. Some are also alkaliphiles and/or thermophiles. Metabolism types found in this phylum include fermentation, chemoautotrophy, methylotrophy, and anaerobic respiration. Some species form spores.
Firmicutes_G Limnochordia Gram-negative Limnochorda pilosa— a thermophilic, facultatively anaerobic species, with cells consisting of pleomorphic filaments. Produces endospore-like structures. Isolated from sediment of a brackish lake. This phylum has not yet been well characterized.
Fusobacteriota (Fusobacteria) Fusobacteriia Diderm; contains genes for LPS (Antunes et al. 2016). Fusobacterium nucleatum— an anaerobic bacterium commonly occurring as a commensal in the human mouth but sometimes becoming pathogenic; it is known to play a role in a large number of disorders and diseases; and Streptobacillus moniliformis— a commensal species found in the upper respiratory tract of rats; humans bitten by rats carrying this bacterium may develop rat bite fever, which has a mortality rate of around 10% if left untreated (Elliott 2007). Members of this phylum are generally anaerobic, facultatively anaerobic or microaerophilic heterotrophs. Many have been isolated from humans and other animals and occur as commensals and opportunistic pathogens. Some taxa have been isolated from marine sediments.
Gemmatimonadota (Gemmatimonadetes) Gemmatimonadetes, Glassbacteria Gram-negative (Diderm); contains genes for LPS (Antunes et al. 2016). Gemmatimonas phototrophica— a facultative photoheterotroph with a Type II (quinone-type) photochemical reaction center. Commonly found in soils and sediments; also known to occur in seawater, freshwater, wastewater, and as a sponge symbiont. Few have been cultivated.
Myxococcota Myxococcia, Polyangia Gram-negative (Diderm) Myxococcus xanthus— a soil bacterium that preys on both gram-positive and gram-negative bacteria. It hunts individually or in coordinated multicellular groups known as swarms, gliding across surfaces and killing prey cells by lysing them. When nutrients become depleted, this species forms multicellular, spore-filled fruiting bodies; Pajaroellobacter abortibovis— a pathogen transmitted by the Pajaroello tick that causes epizootic bovine abortion—the fetuses of pregnant cows die due to an immunological response to the infection (although the infection does not cause an inflammatory response in adult cows); and Sorangium cellulosum— a species that has the ability to degrade cellulose. It has a complex social lifestyle, an extensive network of regulatory genes (Schneiker et al. 2007), and the largest known genome of any bacterium (Han et al. 2013). This phylum includes some taxa formerly classified under the former phylum of Deltaproteobacteria. In general, these bacteria are chemoheterotrophic, rod-shaped, found mostly in soils, move by gliding, and are aerobic (except for Anaeromyxobacter, which is facultatively anaerobic). Many form cellular aggregations known as fruiting bodies, which contain dormant, stress-resistant spores called myxospores. Some marine species are known. Members of this phylum produce a number of bioactive molecules— including antibacterial, antifungal, anti-cancer, and cytotoxic compounds, which are of potential interest for drug development.
Nitrospinota (Nitrospinae) Nitrospinia Gram-negative; contain genes for LPS (Antunes et al. 2016). Nitrospina gracilis— an aerobic nitrite-oxidizing chemoautotroph isolated from ocean surface waters. This phylum has not yet been well characterized.
Nitrospirota (Nitrospirae) Nitrospiria, Thermodesulfovibrionia Gram-negative; contain genes for LPS (Antunes et al. 2016). Nitrospira inopinata—the first cultured complete ammonia oxidizer (comammox). It is an aerobic chemolithoautotroph capable of carrying out both steps of nitrification— i.e., oxidizing ammonia to nitrite, and nitrite to nitrate; Magnetobacterium casensis— an uncultured, autotrophic, magnetotactic bacterium; and Thermodesulfovibrio yellowstonii— an anaerobic, thermophilic, heterotrophic, sulfate-reducing bacterium. This phylum includes nitrite oxidizers, complete ammonia oxidizers, magnetotactic bacteria, and sulfate reducers.
Nitrospirota_A Leptospirillia Gram-negative Leptospirillum ferrooxidans and Leptospirillum ferriphilum— these are aerobic, acidophilic, chemolithoautotrophic iron-oxidizing bacteria. Leptospirillum bacteria are important in biomining operations. Their metabolic actions cause the leaching of metals from mineral ores. The metal-rich leach solution may then be processed to recover the desired metal(s).
Planctomycetota (Planctomycetes) Brocadiae, Phycisphaerae, Planctomycetes Atypical Diderm. Recent studies indicate the Planctomycetota have a gram-negative-type cell plan with an outer membrane and an inner cytoplasmic membrane. There is still some debate over whether the innermost membrane consists of invaginations of the cytoplasmic membrane, or if there are compartments in the cell surrounded by distinct membranes that are separate from the cytoplasmic membrane. Peptidoglycan and LPS have been detected in at least some members of this group (van Teeseling et al. 2015, Jeske et al. 2015, Mahat et al. 2016). Gemmata obscuriglobus— a bacterium described as having membrane-bound cellular compartments, including a "nuclear body" containing DNA and ribosomes, surrounded by a double membrane; some internal membranes in the cell have been found to contain pores similar to eukaryotic nuclear pores (Sagulenko et al. 2017); Kuenenia stuttgartiensis— an anaerobic ammonium-oxidizing (anammox) bacterium. The anammox bacteria convert ammonia to N2 gas using nitrite as the electron acceptor. These bacteria are found in a wide variety of aquatic and terrestrial environments and have complex internal membrane systems. They include many heterotrophic taxa as well as the chemolithoautotrophic anammox bacteria. The Planctomycetota are closely related to the Verrucomicrobia and Chlamydiae bacteria. The Planctomycetota lack the bacterial division protein FtsZ. Most divide by budding, with an exception being the anammox bacteria, which divide by binary fission. Membrane coat-like proteins have been found in these bacteria, and at least some members of this phylum are capable of uptaking macromolecules from the environment using an active/energy-dependent process. This uptake has been described as occurring via an endocytosis-like process (Lonhienne et al. 2010) or by another, unknown process that possibly involves outer membrane complexes called crateriform structures (Boedeker et al. 2017). The anammox bacteria are believed to produce a major portion of the nitrogen gas in our atmosphere (Francis et al. 2007). They are of great interest for use in sewage treatment, as they may provide an efficient, cost-effective and environmentally friendly means for removing ammonium from wastewater (Kartal et al. 2010, Stauffer and Spuhler 2018).
Proteobacteria Alphaproteobacteria, Gammaproteobacteria, Magnetococcia, Zetaproteobacteria Gram-negative; many (but not all) contain genes for LPS (Antunes et al. 2016). Roseobacter denitrificans— a photoheterotrophic purple bacterium; Escherichia coli— commonly found in the intestines of people and other mammals, and in contaminated food, water, and soils; Wolbachia species— endosymbionts in the cells of insects and other invertebrates; they are generally considered to be parasites of insects, but some nematode worms require these endosymbionts for normal development; and Yersinia pestis— the pathogen that causes plague. Class Gammaproteobacteria includes Order Betaproteobacteriales (formerly Class Betaproteobacteria). The Proteobacteria are a diverse group that encompasses free-living taxa as well as intracellular mutualists and pathogens. They include the anoxygenic phototrophic purple bacteria (which have a Type II [quinone-type] photochemical reaction center), magnetotactic bacteria, chemolithoautotrophic iron-oxidizing bacteria, chemolithoautotrophic sulfur-oxidizing bacteria, nitrifying bacteria, nitrogen-fixing bacteria (including nitrogen-fixing symbionts found in nodules in the roots of plants), methane-utilizing bacteria, and chitin-degrading bacteria among other types of bacteria. The mitochondrion in eukaryotes is believed to have evolved from an alphaproteobacterium.
Spirochaetota (Spirochaetes) Brachyspirae, Brevinematia, Leptospirae, Spirochaetia Gram-negative (Diderm); some taxa contain genes for LPS, some do not (Antunes et al. 2016). Treponema pallidum— a human pathogen that causes venereal syphilis (subspecies pallidum), endemic syphilis (subspecies endemicum), and yaws (subspecies pertenue); Borrelia burgdorferi— carried by ticks and causes Lyme disease in humans; its outer membrane lacks LPS and instead contains immunoreactive glycolipids. Most known members of the phylum are spiral-shaped and move by means of internal flagella (endoflagella)— located within the periplasm of the cell. The movement of the flagella causes the whole cell to move in a rotating and/or undulating manner. This phylum includes many pathogens of vertebrates; commensal and symbiotic species, including species occurring in the guts of arthropods (including termites and wood-eating cockroaches); and free-living taxa found in waters, soils, and aquatic and marine and sediments. The Spirochaetota include both aerobic and anaerobic taxa. Some are known to fix nitrogen.
Synergistota (Synergistetes) Synergistia Gram-negative (Diderm); contain genes for LPS (Antunes et al. 2016). Acetomicrobium hydrogeniformans— an anaerobic, thermophilic, sodium chloride-requiring fermenter isolated from oil production water. It is apparently capable of anaerobic respiration using sulfur compounds via an as yet undescribed pathway. Its genome lacks genes for the FoF1-type ATP synthase, which is normally found in bacteria. Instead it possesses genes for either archaeal-like or vacuolar-like ATP synthases; and Synergistes jonesii— a species that may be found in the rumen of some livestock animals. This bacterium degrades a toxic amino acid that is found in the legume Leucaena leucocephala, which is frequently used to feed livestock in the tropics. Inoculation of ruminant livestock with Synergistes jonesii helps protect them against Leucaena toxicity. Includes many anaerobic fermenters. They are generally able to degrade amino acids and/or peptides and many are incapable of degrading carbohydrates. Some are thermophilic. They are not known to produce spores. Some of the habitats they have been isolated from include petroleum reservoirs, sludge from sewage and wastewater treatment operations, saline springs, and the bodies of mammals, including humans.
Thermodesulfobiota Thermodesulfobiia Gram-negative (diderm) Thermodesulfobium narugense— a chemoautotrophic, sulfur-reducing thermophile isolated from a hot spring. This phylum has not yet been well characterized.
Thermosulfidibacterota Thermosulfidibacteria Diderm Thermosulfidibacter takaii— an anaerobic, thermophilic, hydrogen-oxidizing, sulfur-reducing, chemolithoautotroph obtained from a deep-sea hydrothermal field. This phylum has not yet been well characterized.
Thermotogota (Thermotogae) Thermotogae Atypical diderm. In place of an outer membrane has an outer layer called a "toga" that lacks LPS and consists mostly of proteins rather than lipids. Mesoaciditoga lauensis— a thermophilic heterotroph isolated from a deep-sea hydrothermal vent deposit; and Thermotoga maritima— a hyperthermophilic anaerobic fermenter isolated from geothermally-heated sea floors, able to grow at temperatures as high as 90°C. Around one quarter of the DNA sequences making up this species' genome may have been obtained from archaea, through horizontal gene transfer (Nelson et al. 1999). Many of the cultured species are anaerobic, thermophilic, heterotrophs that were isolated from hydrothermal sediments or oil or gas reservoirs.
Verrucomicrobiota (Verrucomicrobia) Chlamydiia, Kiritimatiellae, Lentisphaeria, Verrucomicrobiae Diderm; contain genes for LPS (Antunes et al. 2016). Chlamydia trachomatis— an intracellular pathogen that causes a sexually-transmitted disease in humans; Akkermansia muciniphila— a generally common, abundant, and beneficial inhabitant of the human gut; and Methylacidiphilum fumariolicum— a free-living thermoacidophile isolated from a volcanic mudpot that consumes methane but is also able to grow chemoautotrophically on hydrogen and carbon dioxide. Class Chlamydiia includes intracellular pathogens of humans and other animals, and intracellular pathogens and endosymbionts of amoebas (some of which also infect animals). The phylum also includes free-living aerobic and anaerobic species found in terrestrial, freshwater and marine environments, as well as commensal or symbiotic species found in the guts or other tissues of animals. Some taxa in this phylum have highly compartmentalized cells and have been found to possess membrane coat-like proteins. Membrane coat-like proteins are only known to occur in eukaryotes and in the bacterial phyla Verrucomicrobiota and Planctomycetota.

Notes:
Names of phyla and associated classes and species are from the Genome Taxonomy Database (GTDB). Release 03-RS86 (19th August 2018). Phyla names in parentheses are from the National Center for Biotechnology Information (NCBI)'s taxonomy database (Accessed December 2018). The GTDB phylogeny is inferred from the concatenation of 120 ubiquitous single-copy proteins; higher taxonomic ranks are normalized based on relative evolutionary divergence (Parks et al. 2018).

An envelope type of "Gram-negative" indicates that the taxon stained gram negative and is presumed to have a gram-negative (diderm) cell envelope. Where the word "diderm" is specified, the taxon is known to have both a cytoplasmic membrane and an outer membrane.

An envelope type of "Gram-positive" indicates that the taxon stained gram positive and is presumed to have a gram-positive (monoderm) cell envelope. Where the word "monoderm" is specified, the taxon is known to lack an outer membrane (i.e., possessing only a cytoplasmic membrane).

[1] Although bacteria are unable to replicate large amounts of DNA efficiently like eukaryotes can (because their chromosome consists of a single replicon) their DNA replication rate is faster than that of eukaryotes. Bacteria have a faster replication rate in part because the bacterial version of the replication enzyme is speedier and because bacterial DNA is more "naked" (it lacks the histone-based packaging characteristic of eukaryotic DNA) and so less time and energy is needed for processing/unwinding the DNA for replication (and transcription) to take place (Kuzminov 2014). So for organisms with small genomes, the bacterial mode of life is more efficient, whereas organisms with large genomes are better suited being eukaryotes. [Return to text]

[2] A given copy of a plasmid can be replicated multiple times over the course of a cell division cycle. This is in contrast to a given copy of a chromosome, which can be replicated no more than once per cell division cycle (Egan et al. 2005).[Return to text]

[3] The small plasmids that lack partition systems have a high copy number per cell, which makes it much more likely that these plasmids will get randomly distributed to both daughter cells upon cell division (Lili et al. 2007). The larger plasmids (and other large secondary replicons), which have partition systems tend to have a low copy number (Lili et al. 2007)— lots of copies aren't needed when there is an active system to ensure that the replicons get distributed to both daughter cells.[Return to text]

[4]Some authors suggest that megaplasmids should be defined as having a minimum size of 100 kilobases (Anton et al. 1995, Barton et al. 1995), while others recommend a minimum size of 350 kilobases (diCenzo and Finan 2017).[Return to text]

[5]Replicons can fuse if each has repeated DNA sequences that are very similar to those on the other molecule and of sufficient length, and thus can undergo homologous recombination. According to Guo et al. (2003), ideally the matching sequences on the two replicons should have more than 95% sequence similarity and be larger than 1 kilobase. [Return to text]

[6]Carl Weigert later modified the procedure, adding the counterstain safranin so that gram negative cells would stain red (Sandle 2004).[Return to text]

[7]Signal transduction involves a complex set of events that results in the expression of certain genes.[Return to text]

[8]Gram-negative species known to lack LPS include Sorangium cellulosum (phylum Myxococcota); Treponema pallidum, Borrelia (Borreliella) burgdorferi, and B. hispanica (phylum Spirochaetota); Sphingomonas capsulate and S. paucimobilis (phylum Proteobacteria); Deinococcus radiodurans , Thermus thermophilus, and Meiothermus taiwanensis (phylum Deinococcota); and Thermotoga maritima (phylum Thermotogota) (Keck et al. 2011, Raetz and Whitfield 2002). In addition, some species have LPS-like glycolipids that lack the O-antigen; thus these species are described by some as having lipooligosaccharides (LOS) rather than lipopolysaccharides (LPS) (Stewart et al. 2006, Nikaido 2003). Species with LOS glycolipids in their outer membrane include proteobacteria of the genera Neisseria, Haemophilus, Bordetella, and Moraxella (Branhamella) (Preston et al. 1996).[Return to text]

[9]Many gram-positive species also produce toxins that can result in infections that cause sepsis and death (e.g., the superantigens of Staphylococcus aureus, exotoxin A of Pseudomonas aeruginosa, and the secreted toxins of Bacillus anthracis (Ramachandran 2014).[Return to text]

[10]The toxicity of LPS/lipid A can vary considerably among different species of bacteria, ranging from nontoxic (e.g., in the purple nonsulfur bacteria Rhodobacter sphaeroides and R. capsulatus) to potentially highly toxic (e.g., in proteobacteria from the family Enterobacteriaceae) (Stewart et al. 2006). [Return to text]

[11]Recently, gram-positive bacteria from the phyla Firmicutes and Actinobacteriota have been found to release extracellular vesicles; it's not yet known how the vesicles get past the cell wall (Liu et al. 2018). [Return to text]

[12]Outside of the Firmicutes, it is unusual for a bacterium to have many of its outer membrane-associated genes grouped together in one large cluster in the genome (Antunes et al. 2016). This makes it harder to convincingly show in general that all diderm bacteria with an LPS outer membrane are descended from a common diderm ancestor. However, it appears that across bacterial phyla, LPS-containing diderm bacteria often contain a cluster in their genomes with at least four genes involved in LPS biosynthesis, with the positions of the genes within the cluster highly conserved (Opiyo et al. 2010). Based on this, the homology of outer membrane proteins among diderm phyla (Tocheva et al. 2016), and results from phylogenetic analyses, many scientists believe it is likely that all bacteria with an LPS outer membrane are descended from a common diderm ancestor. If true, the ancestral diderm bacteria likely possessed a primordial form of Lipid A in their outer membrane, since only four LPS biosynthesis enzymes are conserved across bacterial phyla (Opiyo et al. 2010). [Return to text]

[13]Refer to Tocheva et al. 2011 supplementary materials (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3176627/) for excellent videos showing the processes of endospore formation, germination and outgrowth in Acetonema longum. [Return to text]

Sources

Adams, P.G. and C.N. Hunter. 2012. Adaptation of intracytoplasmic membranes to altered light intensity in Rhodobacter sphaeroides. Biochimica et Biophysica Acta (BBA) - Bioenergetics. 1817(9):1616-1627.

Alberts, B., D. Bray, J. Lewis, M. Raff, K. Roberts, and J.D. Watson. 1994. Molecular Biology of the Cell. Third Edition. Garland Publishing, Inc. New York. 1294 pp.

Allen, J.P. and J.C. Williams. 1998. Photosynthetic reaction centers. FEBS Letters. 438:5-9.

Allen, M.M. and P.J. Weathers. 1980. Structure and composition of cyanophycin granules in the cyanobacterium Aphanocapsa 6308. Journal of Bacteriology. 141(2):959-962.

Allison, M.J., W.R. Mayberry, C.S. Mcsweeney, and D.A. Stahl. 1992. Synergistes jonesii, gen. nov., sp.nov.: A Rumen Bacterium That Degrades Toxic Pyridinediols. Systematic and Applied Microbiology. 15(4):522-529.

Ambroset C., C. Coluzzi, G. Guédon, M.-D. Devignes, V. Loux, T. Lacroix, S. Payot, and N. Leblond-Bourget. 2016. New insights into the classification and integration specificity of Streptococcus integrative conjugative elements through extensive genome exploration. Frontiers in Microbiology. 6:1483. doi: 10.3389/fmicb.2015.01483.

Angert, E.R., K.D. Clements, and N.R. Pace. 1993. The largest bacterium. Nature. 362:239-241.

Anton J., R. Amilsz, C.L. Smith, and P. Lopez-Garcia. 1995. Comparative restriction maps of the archaeal megaplasmid pHM300 in different Haloferax mediterranei strains. Systematic and Applied Microbiology. 18(3):439-447.

Antunes L.C.S., D. Poppleton, A. Klingl, A. Criscuolo, B. Dupuy, C. Brochier-Armanet, C. Beloin, and S. Gribaldo. 2016. Phylogenomic analysis supports the ancestral presence of LPS-outer membranes in the Firmicutes. eLife. 5:e14589. doi: 10.7554/eLife.14589.

Appelmelk B.J., I. Simoons-Smit, R. Negrini, A.P. Moran, G.O. Aspinall, J.G. Forte, T. De Vries, H. Quan, T. Verboom, J.J. Maaskant, P. Ghiara, E.J. Kuipers, E. Bloemena, T.M. Tadema, R.R. Townsend, K. Tyagarajan, J.M. Crothers Jr., M.A. Monteiro, A. Savio and J. De Graaff. 1996. Potential role of molecular mimicry between Helicobacter pylori lipopolysaccharide and host lewis blood group antigens in autoimmunity. Infection and Immunity. 64(6):2031-2040.

Arechaga, I., B. Miroux, S. Karrasch, R. Huijbregts, B. de Kruijff, M.J. Runswick, and J.E. Walker. 2000. Characterisation of new intracellular membranes in Escherichia coli accompanying large scale over-production of the b subunit of F1FO ATP synthase. FEBS Letters. 482(3):215-219.

Badrinarayanan, A., T.B.K. Le, and M.T. Laub. 2015. Bacterial chromosome organization and segregation. Annu. Rev. Cell Dev. Biol. 31:171-199. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4706359/.

Bailey, P. 2016. Bizarre Bacteria Causing Major Cattle Disease Named by UC Davis Researchers. https://www.ucdavis.edu/news/bizarre-bacteria-causing-major-cattle-disease-named-uc-davis-researchers/. Published on July 26 in Food & Agriculture.

Bakshi S., H. Choi, and J.C. Weisshaar. 2015. The spatial biology of transcription and translation in rapidly growing Escherichia coli. Frontiers in Microbiology. 6:636. doi: 10.3389/fmicb.2015.00636.

Bakshi S., A. Siryaporn, M. Goulian, and J.C. Weisshaar. 2012. Superresolution imaging of ribosomes and RNA polymerase in live Escherichia coli cells. Mol Microbiol 85(1):21-38.

Barbour A.G. and C.F. Garon. 1987. Linear plasmids of the bacterium Borrelia burgdorferi have covalently closed ends. Science 237: 409–411.

Barka, E.A., P. Vatsa, L. Sanchez, N. Gaveau-Vaillant, C. Jacquard, H.-P. Klenk, C. Clément, Y. Ouhdouch, and G.P. van Wezel. 2016. Taxonomy, Physiology, and Natural Products of Actinobacteria. Microbiol Mol Rev. 80(1):1-43.

Barton B.M., G.P. Harding, and A.J. Zuccarelli. 1995. A general method for detecting and sizing large plasmids. Analytical Biochemistry. 226(2):235-240.

Begemann, M.B., M.R. Mormile, O.C. Sitton, J.D. Wall, and D.A. Elias. 2012. A Streamlined Strategy for Biohydrogen Production with Halanaerobium hydrogeniformans, an Alkaliphilic Bacterium. Front. Microbiol. 3:93. doi:10.3389/fmicb.2012.00093.

Beisson, J. 2008. Preformed cell structure and cell heredity. Prion. 2(1):1-8.

Benson, H.J. 1990. Microbiological Applications. A Laboratory Manual in General Microbiology. Fifth Edition. Wm. C. Brown Publishers. Dubuque, IA.

Bernstein J.A., A.B. Khodursky, P.-H. Lin, S. Lin-Chao, and S.N. Cohen. 2002. Global analysis of mRNA decay and abundance in Escherichia coli at single-gene resolution using two-color fluorescent DNA microarrays. PNAS. 99(15):9697-9702.

Beveridge T.J. 1990. Mechanism of gram variability in select bacteria. Journal of Bacteriology. 172(3):1609-1620.

Beveridge T.J. 1999. Structures of gram-negative cell walls and their derived membrane vesicles. Journal of Bacteriology. 181(16):4725-4733.

Blum, J.W., S. Han, B. Lanoil, C. Saltikov, B. Witte, F.R. Tabita, S. Langley, T.J. Beveridge, L. Jahnke, and R.S. Oremland. 2009. Ecophysiology of “Halarsenatibacter silvermanii” Strain SLAS-1T, gen. nov., sp. nov., a Facultative Chemoautotrophic Arsenate Respirer from Salt-Saturated Searles Lake, California. Applied and Environmental Microbiology. 75(7):1950-1960.

Boedeker, C., M. Schüler, G. Reintjes, O. Jeske, M.C.F. van Teeseling, M. Jogler, P. Rast, D. Borchert, D.P. Devos, M. Kucklick, M. Schaffer, R. Kolter, L. van Niftrik, S. Engelmann, R. Amann, M. Rohde, H. Engelhardt, and C. Jogler. 2017. Determining the bacterial cell biology of Planctomycetes. Nature Communications. 8:14853. doi:10.1038/ncomms14853.

Bond, D.R., and D.R. Lovley. 2003. Electricity Production by Geobacter sulfurreducens Attached to Electrodes. Applied and Environmental Microbiology. 69(3):1548-1555.

Braun V., S. Gaisser, C. Herrmann, K. Kampfenkel, H. Killmann, and I. Traub. 1996. Energy-coupled transport across the outer membrane of Escherichia coli: ExbB binds ExdB and TonB in vitro, and Leucine 132 in the periplasmic region and Aspartate 25 in the transmembrane region are important for ExbD activity. Journal of Bacteriology. 178(10):2836-2845.

Bremer, H. and P.P. Dennis. 1996. Modulation of chemical composition and other parameters of the cell by growth rate. Escherichia coli and Salmonella, ed Neidhardt, F.C. (Am Soc Microbiol, Washington, DC), 2nd Ed, pp 1553–1569.

Briffotaux J. and K. Kobryn. 2010. Preventing broken Borrelia telomeres. ResT couples dual hairpin telomere formation with product release. The Journal of Biological Chemistry 285(52):41010-41018.

Brooks, R.S., M.T. Blanchard, K.A. Clothier, S. Fish, M.L. Anderson, and J.L. Stott. 2016. Characterization of Pajaroellobacter abortibovis, the etiologic agent of epizootic bovine abortion. Veterinary Microbiology. 192:73-80.

Brumm, P.J., K. Gowda, F.T. Robb, and D.A. Mead. 2016. The Complete Genome Sequence of Hyperthermophile Dictyoglomus turgidum DSM 6724™ Reveals a Specialized Carbohydrate Fermentor. Front microbiol. 7:1979. doi:10.3389/fmicb.2016.01979.

Bryant, D.A., A.M. Costas, J.A. Maresca, A.G. Chew, C.G. Klatt, M.M. Bateson, L.J. Tallon, J. Hostetler, W.C. Nelson, J.F. Heidelberg, and D.M. Ward. 2007. Candidatus Chloracidobacterium thermophilum: an aerobic phototrophic Acidobacterium. Science. 317(5837):523-526.

Bryant C.E., D.R. Spring, M. Gangloff, and N.J. Gay. 2010. The molecular basis of the host response to lipopolysaccharide. Nature Reviews Microbiology. 8:8-14.

Bull, M. S. Plummer, J. marchesi, E. Mahenthiralingam. 2013. The life history of Lactobacillus acidophilus as a probiotic: a tale of revisionary taxonomy, misidentification and commercial success. FEMS Microbiology Letters. 349(2):77-87.

Cardona, T. 2015. A fresh look at the evolution and diversification of photochemical reaction centers. Photosyn Res. 126:111-134.

Casjens S., M. Delange, H.L. Ley III, P. Rosa, and W.M. Huang. 1995. Linear chromosomes of lyme disease agent spirochetes: genetic diversity and conservation of gene order. Journal of Bacteriology. 177(10):2769-2780.

Castellana M., S. H.-J. Li, and N.S. Wingreen. 2016. Spatial organization of bacterial transcription and translation. PNAS. 11(33):9286-9291.

Castillo F., A. Benmohamed, and G. Szatmari. 2017. Xer site specific recombination: double and single recombinase systems. Frontiers in Microbiology. 8(453):1-18. doi:10.3389/fmicb.2017.00453.

Cavalier-Smith T. 2006. Rooting the tree of life by transition analyses. Biology Direct. 1:19. doi: 10.1186/1745-6150-1-19.

Chaconas G. and K. Kobryn. 2010. Structure, function, and evolution of linear replicons in Borrelia. Annual Review of Microbiology. 64:185-202. doi:10.1146/annurev.micro.112408.134037.

Chan, J.Z.-M., M.R. Halachev, N.J. Loman, C. Constantinidou, and M.J. Pallen. 2012. Defining bacterial species in the genomic era: insights from the genus Acinetobacter. BMC Microbiology. 12:302. https:/doi.org/10.1186/1471-2180-12-302.

Chater, K.F. 2006. Streptomyces inside-out: a new perspective on the bacteria that provide us with antibiotics. Phil. Trans. R. Soc. B. 361:761-768.

Chen, L., D.A. Bazylinski, and B.H. Lower. 2010. Bacteria that synthesize nano-sized compasses to navigate using earth's geomagnetic field. Nature Education Knowledge. 3(10):30.

Cheng K.-J. and J.W. Costerton. 1977. Ultrastructure of Butyrivibrio fibrisolvens: a gram-positive bacterium? Journal of Bacteriology. 129(3):1506-1512.

Chien, A-C., N.S. Hill, and P.A. Levin. 2012. Cell size control in bacteria. Current Biology. 22(9):R340-R349. https://doi.org/10.1016/j.cub.2012.02.032.

Chivian, D., E.L. Brodie, E.J. Alm, D.E. Culley, P.S. Dehal, T.Z. DeSantis, T.M. Gihring, A. Lapidus, L.-H. Lin, S.R. Lowry, D.P. Moser, P.M. Richardson, G. Southam, G. Wanger, L.M. Pratt, G.L. Andersen, T.C. Hazen, F.J. Brockman, A.P. Arkin, T.C. Onstott. 2008. Environmental Genomics Reveals a Single-Species Ecosystem Deep Within Earth. Science. 322(5899):275-278.

Church, DL, Cerutti L, Gürtler AGriener T, Zelazny A, Emler S 2020. Performance and Application of 16S rRNA Gene Cycle Sequencing for Routine Identification of Bacteria in the Clinical Microbiology Laboratory. Clin Microbiol Rev 33:10.1128/cmr.00053-19. https://doi.org/10.1128/cmr.00053-19

Ciofu O., T.J. Beveridge, J. Kadurugamuwa, J. Walther-Rasmussen, and N. Høiby. 2000. Chromosomal β-lactamase is packaged into membrane vesicles and secreted from Pseudomonas aeruginosa. Journal of Antimicrobial Chemotherapy. 45(1):9-13. https://doi.org/10.1093/jac/45.1.9.

Clifton L.A., M.W.A. Skoda, E.L. Daulton, A.V. Hughes, A.P. Le Brun, J.H. Lakey, and S.A. Holt. 2013. Asymmetric phosopholipid: lipopolysaccharide bilayers; a gram-negative bacterial outer membrane mimic. Journal of the Royal Society Interface. 10:20130810. http://dx.doi.org/10.1098/rsif.2013.0810.

Cohen G.N. 2004. The outer membrane of gram-negative bacteria and the cytoplasmic membrane. In: Microbial Biochemistry. Chapter 2. Springer, Dordrecht.

Cook, L.E., S.S. Gang, A. Ihlan, M. Maune, R.S. Tanner, M.J. McInerney, G. Weinstock, E.A. Lobos, and R.P. Gunsalus. 2018. Genome Sequence of Acetomicrobium hydrogeniformans OS1. Genome Announc. 6(26):e00581-18. doi:10.1128/genomeA.00581-18.

Cooper S. and C.E. Helmstetter. 1968. Chromosome replication and the division cycle of Escherichia coli B/r. J Mol Biol. 1968;31:519–540.

Cortez D., S. Quevillon-Cheruel, S. Gribaldo, N. Desnoues, G. Sezonov, P. Forterre, M.-C. Serre. 2010. Evidence for a Xer/dif system for chromosome resolution in Archaea. PLoS Genetics. 6(10):e1001166. https://doi.org/101371/journal.pgen.1001166.

Cunningham, M.W. 2000. Pathogenesis of Group A Streptococcal Infections. Clinical Microbiology Reviews. 13(3):470-511.

Cury J., M. Touchon, and E.P.C. Rocha. Integrative and conjugative elements and their hosts: composition, distribution and organization. Nucleic Acids Research. 45(15):8943-8956. doi: 10.1093/nar/gkx607.

Cusick M.F., J.E. Libbey, and R.S. Fujinami. 2012. Molecular mimicry as a mechanism of autoimmune disease. Clinic Rev Allerg Immunol. 42:102-111. doi: 10.1007/s12016-011-8294-7.

D'Amico, S., T. Collins, J.-C. Marx, G. Feller, and C. Gerday. 2006. Psychrophilic microorganisms: challenges for life. EMBO reports. 7(4):385-389.

Daims, H., E.v. Lebedeva, P. Pjevac, P. Han, C. Herbold, M. Albertsen, N. Jehmlich, M. Palatinszky, J. Vierheilig, A. bulaev, R.H. Kirkegaard, M. von Bergen, T. Rattei, B. Bendinger, P.H. Nielsen, and M. Wagner. 2015. Complete nitrification by Nitrospira bacteria. Nature. 528(7583):504-509.

de Almeida, N.M., S. Neumann, R.J. mesman, C. Ferousi, J.T. Keltjens, M.S.M. Jetten, B. Kartal, and L. van Niftrik. 2015. Immunogold localization of key metabolic enzymes in the anammoxosome and on the tubule-like structures of Kuenenia stuttgartiensis. Journal of Bacteriology. 197(14):2432-2441.

Deckert, G., P.V. Warren, T. Gaasterland, W.G. Young, A.L. Lenox, D.E. Graham, R. Overbeek, M.A. Snead, M. Keller, M. Aujay, R. Huber, R.A. Feldman, J.M. Short, G.J. Olsen, and R.V. Swanson. 1998. The complete genome of the hyperthermophilic bacterium Aquifex aeolicus. Nature, 392:353-358.

Deeg C.M., M.M. Zimmer, E.E. George, F. Husnik, P.J. Keeling, and C.A. Suttle. 2019. Chromulinavorax destructans, a pathogen of microzooplankton that provides a window into the enigmatic candidate phylum Dependentiae. PLOS. May 31. https://journals.plos.org/plospathogens/article?id=10.1371/journal.ppat.1007801.

Descamps, E.C.T., C.L. Monteil, N. Menguy, N. Ginet, D. Pignol, D.A. Bazylinski and C.T. Lefèvre. 2017. Desulfamplus magnetovallimortis gen. nov., sp. nov., a magnetotactic bacterium from a brackish desert spring able to biomineralize greigite and magnetite, that represents a novel lineage in the Desulfobacteraceae. Syst Appl Microbiol. 40(5):280-289.

DiCenzo, G.C. and T.M. Finan. 2017. The divided bacterial genome: structure, function, and evolution. Microbiology and Molecular Biology Reviews. 81(3): e0019-17.

Docampo, R. 2016. The origin and evolution of the acidocalcisome and its interactions with other organelles. Mol Biochem Parasitol. 209(1-2):3-9. doi:10.1016/j.molbiopara.2015.10.003.

Docampo, R., W. de Souza, K. Miranda, P. Rohloff, and S.N.J. Moreno. 2005. Nature Reviews Microbiology. 3:251-261.

Docampo, R. and S.N.J. Moreno. 2011. Acidocalcisomes. Cell Calcium. 50(2):113-119. doi:10.1016/j.ceca.2011.05.012.

Docampo, R., P. Ulrich, and S.N.J. Moreno. 2010. Evolution of acidocalcisomes and their role in polyphosphate storage and osmoregulation in eukaryotic microbes. Phil. Trans. R. Soc. B. 365:775-784. doi:10.1098/rstb.2009.0179.

Domingues S. and K.M. Nielsen. 2017. Membrane vesicles and horizontal gene transfer in prokaryotes. Current Opinion in Microbiology. 38:16-21. http://dx.doi.org/10.1016/j.mib.2017.03.012.

Dorward D.W., C.F. Garon, and R.C. Judd. 1989. Export and intercellular transfer of DNA via membrane blebs of Neisseria gonorrhoeae. Journal of Bacteriology. 171(5):2499-2505.

Dramsi S., S. Magnet, S. Davison, and M. Arthur. 2008. Covalent attachment of proteins to peptidoglycan. FEMS Microbiology Reviews. 32(2):307-320. https://doi.org/10.1111/j.1574-6976.2008.00102.x.

Elliott, S.P. 2007. Rat Bite Fever and Streptobacillus moniliformis. Clin Microbiol Rev. 20(1):13-22.

Errington, J. 2013. L-form bacteria, cell walls and the origin of life. Open Biol. 3(1):120143.

Etchebehere, C., M.E. Pavan, J. Zorzópulos, M. Soubes, and L. Muxí. 1998. Coprothermobacter platensis sp. nov., a new anaerobic proteolytic thermophilic bacterium isolated from an anaerobic mesophilic sludge. International Journal of Systematic Bacteriology. 48:1297-1304.

Fenchel, T. 1994. Microbial ecology on land and sea. Phil. Trans. R. Soc. Lond. B. 343:51-56.

Fenchel, T. and B.J. Finlay. 1995. Ecology and Evolution in Anoxic Worlds. Oxford University Press: New York. 276 pages.

Ferdows M.S. and A.G. Barbour. 1989. Megabase-sized linear DNA in the bacterium Borrelia burgdorferi, the Lyme disease agent. Proc. Natl. Acad. Sci. USA 86:5969-5973.

Fernández L. and R.E.W. Hancock. 2012. Adaptive and mutational resistance: role of porins and efflux pumps in drug resistance. Clinical Microbiology Reviews. 25(4):661-681.

Flärdh, K. and M. Buttner. 2009. Streptomyces morphogenetics: dissecting differentiation in a filamentous bacterium. Nature Reviews Microbiology. 7:36-49.

Flot, J.-F., G. Wörheide, and S. Dattagupta. 2010. Unsuspected diversity of Niphargus amphipods in the chemoautotrophic cave ecosystem of Frasassi, central Italy. BMC Evolutionary Biology. 10:171. https://doi.org/10.1186/1471-2148-10-171.

Forst D., W. Welte, T. Wacker, and K. Diederichs. 1998. Structure of the sucrose-specific porin ScrY from Salmonella typhimurium and its complex with sucrose. Nature Structural Biology. 5:37-46. http://dx.doi.org/10.1038/nsb0198-37.

Fournes F., M.-E. Val, O. Skovgaard, and D. Mazel. 2018. Replicate once per cell cycle: replication control of secondary chromosomes. Frontiers in Microbiology. 9:1833. doi: 10.3389/fmicb.2018.01833.

Fournier, P-E., K. Suhre, G. Fournous, and D. Raoult. 2006. Estimation of prokaryotic DNA G+C content by sequencing universally conserved genes. International Journal of Systematic and Evolutionary Microbiology. 56:1025-1029. DOI: 10.1099/ijs.0.63903-0.

Fossum S., E. Crooke, and K. Skarstad. 2007. Organization of sister origins and replisomes during multifork DNA replication in Escherichia coli. EMBO J. 26(21):4514-4522. doi:10.1038/sj.emboj.7601871.

Francis, C.A., J.M. Beman, and M.M.M. Kuypers. 2007. New processes and players in the nitrogen cycle: the microbial ecology of anaerobic and archaeal ammonia oxidation. The ISME Journal. 1:19-27.

Frigaard, N.U., A.G. Chew, J.A. Maresca, and D.A. Bryant. 2003. Chlorobium tepidum: insights into the structure, physiology, and metabolism of a green sulfur bacterium derived from teh complete genome sequence. Photosynth Res. 78(2):93-117.

Fuerst, J.A. and R.I. Webb. 1991. Membrane-bounded nucleoid in the eubacterium Gemmata obscuriglobus. Proc. Natl. Acad. Sci. USA. 88:8184-8188.

Furuta N., K. Tsuda, H. Omori, T. Yoshimori, F. Yoshimura, and A. Amano. 2009. Porphyromonas gingivalis Outer Membrane Vesicles Enter Human Epithelial Cells via an Endocytic Pathway and Are Sorted to Lysosomal Compartments. Infection and Immunity. 77(10):4187-4196.

Füser, G. and A. Steinbüchel. 2007. Analysis of Genome Sequences for Genes of Cyanophycin Metabolism: Identifying Putative Cyanophycin Metabolizing Prokaryotes. Macromolecular Bioscience. 7(3):278-296.

Galdiero S., A. Falanga, M. Cantisani, R. Tarallo, M.E. Della Pepa, V. D'Oriano, and M. Galdiero. 2012. Microbe-host interactions: structure and role of gram-negative bacterial porins. Current Protein and Peptide Science. 13:843-854.

Geissinger, O., D.P.R. Herlemann, E. Mörschel, U.G. Maier, and A. Brune. 2009. The Ultramicrobacterium “Elusimicrobium minutum” gen. nov., sp. nov., the First Cultivated Representative of the Termite Group 1 Phylum. Appl Environ Microbiol. 75(9):2831-2840.

Genome Taxonomy Database (GTDB). Release 03-RS86 (19th August 2018). http://gtdb.ecogenomic.org/.

Gest H. and J. Mandelstam. 1987. Longevity of microorganisms in natural environments. Microbiol. Sci. 4:69-71.

Ghai, R., C.M. Mizuno, A. Picazo, A. Camacho, and F. Rodriguez-Valera. 2013. Metagenomics uncovers a new group of low GC and ultra-small marine Actinobacteria. Scientific Reports. 3:2471. DOI: 10.1038/srep02471.

Glud, R.N., F. Wenzhöfer, M. Middelboe, K. Oguri, R. Turnewitsch, D.E. Canfield, and H. Kitazato. 2013. Nature Geoscience. 6:284-288.

Gohara, D.W. and Yap, M.-N.F. 2018. Survival of the drowsiest: the hibernating 100S ribosome in bacterial stress management. Current Genetics. 64:753-760. https://doi.org/10.1007/s00294-017-0796-2.

Gorlenko, V., A. Tsapin, Z. Namsaraev, T. Teal, T. Tourova, D. Engler, R. Mielke, and K. Nealson. 2004. Anaerobranca californiensis sp. nov., an anaerobic, alkalithermophilic, fermentative bacterium isolated from a hot spring on Mono Lake.Int. J. Syst. Evol. Microbiol. 54:739-743.

Goux, H.J., D. Chavan, M. Crum, K. Kourentzi, and R.C. Wilson. 2018. Akkermansia muciniphila as a Model Case for the Development of an Improved Quantitative RPA Microbiome Assay. Front Cell Infect Microbiol. 8:237. doi:10.3389/fcimb.2018.00237.

Gouy, R., D. Baurain, and H. Philippe. 2015. Rooting the tree of life: the phylogenetic jury is still out. Phil. Trans. R. Soc. B. 370:20140329. http://dx.doi.org/10.1098/rstb.2014.0329.

Grabowicz M. and T.J. Silhavy. 2017. Redefining the essential trafficking pathway for outer membrane lipoproteins. PNAS. 114(18):4769-4774. https://doi.org/10.1073/pnas.1702248114.

Gralnick, J.A. and D.K. Newman. 2007. Extracellular respiration. Mol. Microbiol. 65(1):1-11.

Gram H. C. 1884. Über die isolierte Färbung der Schizomyceten in Schnitt- und Trockenpräparaten. Fortschritte der Medizin (in German). 2:185–189.

Greub, G. 2009. Parachlamydia acanthamoebae, an emerging agent of pneumonia. Clin Microbiol Infect. 15(1):18-28. doi: 10.1111/j.1469-0691.2008.02633.x.

Gribaldo, S., A.M. Poole, V. Daubin, P. Forterre, and C. Brochier-Armanet. 2010. The origin of eukaryotes and their relationship with the Archaea: are we at a phylogenomic impasse? Nature Rev. Microbiol. 8:743-752.

Griffiths, E. and R.S. Gupta. 2006. Molecular signatures in protein sequences that are characteristics of the phylum Aquificae. International Journal of Systematic and Evolutionary Microbiology. 56:99-107.

Guglielmini J., L. Quintais, M.P. Garcillán-Barcia, F. de la Cruz, and E.P.C. Rocha. 2011. The repertoire of ICE in prokaryotes underscores the unity, diversity, and ubiquity of conjugation. PLoS Genetics. 7(8):e1002222. doi: 10.1371/journal.pgen.1002222.

Guo X., M. Flores, P. Mavingui, S.I. Fuentes, G. Hernández, G. Dávila, and R. Palacios. 2003. Natural genomic design in Sinorhizobium meliloti: Novel genomic architectures. Genome Research. 13:1810-1817. doi:10.1101/gr.1260903.

Gupta R.S. 2011. Origin of diderm (gram-negative) bacteria: antibiotic selection pressure rather than endosymbiosis likely led to the evolution of bacterial cells with two membranes. Antonie van Leeuwenhoek. 100:171-182. doi:10.1007/s10482-011-9616-8.

Gupta, R.S. 2014. The Phylum Aquificae. In: Rosenberg E., DeLong E.F., Lory S., Stackebrandt E., Thompson F. (eds) The Prokaryotes. Springer, Berlin, Heidelberg

Gupta, R.S., S. Mahmood, and M. Adeolu. 2013. A phylogenomic and molecular signature based approach for characterization of the phylum Spirochaetes and its major clades: proposal for a taxonomic revision of the phylum. Front Microbiol. 4:217. doi:10.3389/fmicb.2013.00217.

Hahn, M.W., J. Schmidt, U. Koll, M. Rohde, S. Verbarg, A. Pitt, R. Nakai, T. Naganuma, and E. Lang. 2017. Silvanigrella aquatica gen. nov., sp. nov., isolated from a freshwater lake, description of Silvanigrellaceae fam. nov. and Silvanigrellales ord. nov., reclassification of the order Bdellovibrionales in the class Oligoflexia, reclassification of the families Bacteriovoracaceae and Halobacteriovoraceae in the new order Bacteriovoracales ord. nov., and reclassification of the family Pseudobacteriovoracaceae in the order Oligoflexales. International journal of systematic and evolutionary microbiology. 67(8):2555-2568.

Halliday, M.J., J. Padmanabha, C.S. McSweeney, G. Kerven and H.M. Shelton. 2013. Leucaena toxicity: a new perspective on the most widely used forage tree legume. Tropical Grasslands - Forrajes Tropicales. 1:1-11.

Han, C., R. Mwirichia, Ol Chertkov, B. Held, A. Lapidus, M. Nolan, S. Lucas, N. Hammon, S. Deshpande, J.-F. Cheng, R. Tapia, L. Goodwin, S. Pitluck, M. Huntemann, K. Liolios, N. Ivanova, I. Pagani, K. Mavromatis, G. Ovchinikova, A. Pati, A. Chen, K. Palaniappan, M. Land, L. Hauser, E.-M. Brambilla, M. Rohde, S. Spring, J. Sikorski, M. Göker, T. Woyke, J. Bristow, J.A. Eisen, V. Markowitz, P. Hugenholtz, N.C. Kyrpides, H.-P. Klenk, and J.C. Detter. 2011. Complete genome sequence of Syntrophobotulus glycolicus type strain (FlGlyR). Stand. Genomic Sci. 4(3):371-380.

Han, K., Z.-F. Li, R. Peng, L.-P. Zhu, T. Zhou, L.-G. Wang, S.-G. Li, X.-B. Zhang, W. Hu, Z.-H. Wu, N. Qin, and Y.-Z. Li. 2013. Extraordinary expansion of a Sorangium cellulosum genome from an alkaline milieu. Scientific Reports. 3:2101. doi:10.1038/srep02101. 2013.

Han, Y.W. 2015. Fusobacterium nucleatum: a commensal-turned pathogen. Curr. Opin. Microbiol. 0:141-147. doi:10.1016/j.mib.2014.11.013.

Harold, F.M. 1986. The Vital Force: A Study of Bioenergetics. Freeman, New York.

Harrison P.W., R.P. Lower, N.K. Kim, and J.P. Young. 2010. Introducing the bacterial 'chromid': not a chromosome, not a plasmid. Trends Microbiol. 18(4):141-148. doi: 10.1016/j.tim.2009.12.010.

Hayflick L. 1965. The limited in vitro lifetime of human diploid cell strains. Experimental Cell Research. 37:614-636.

Henry, E.A., R. Devereux, J.S. Maki, C.C. Gilmour, C.R. Woese, L. Mandelco, R. Schauder, C.C. Remsen, and R. Mitchell. 1994. Characterization of a new thermophilic sulfate-reducing bacterium Thermodesulfovibrio yellowstonii, gen. nov. and sp. nov.: its phylogenetic relationship to Thermodesulfobacterium commune and their origins deep within the bacterial domain. Arch. Microbiol. 161(1):62-69.

Hoiczyk, E. and A. Hansel. 2000. Cyanobacterial cell walls: news from an unusual prokaryotic envelope. Journal of Bacteriology. 182:1191-1199.

Holmes, V.F. and N.R. Cozzarelli. 2000. Closing the ring: links between SMC proteins and chromosome partitioning, condensation, and supercoiling. Proc Natl Acad Sci U S A. 97:1322–1324.

Hooi, J.K.Y., W.Y. Lai, W.K. Ng, M.M.Y. Suen, F.E. Underwood, D. Tanyingoh, P. Malfertheiner, D.Y. Graham, V.W.S. Wong, J.C.Y. Wu, F.K.L. Chan, J.J.Y. Sung, G.G. Kaplan, and S.C. Ng. 2017. Global Prevalence of Helicobacter pylori Infection: Systematic Review and Meta-Analysis. Gastroenterology. 153(2):420-429.

Hopwood D.A. 2006. Soil to genomics: the Streptomyces chromosome. Annu. Rev. Genet. 40:1-23. doi:10.1146/annurev.genet.40.110405.090639.

Hornsby P.J. 2007. Telomerase and the aging process. Exp Gerontol. 42(7):575-581. doi:10.1016/j.exger.2007.03.007.

Huang W.M, J. DaGloria, H. Fox, Q. Ruan, J. Tillou, K. Shi, H. Aihara, J. Aron, and S. Casjens. 2012. Linear chromosome-generating system of Agrobacterium tumefaciens C58. The Journal of Biological Chemistry. 287(30):25551-25563.

Huber, R., T.A. Langworthy, H. König, M. Thomm, C.R. Woese, U.B. Sleytr, and K.O. Stetter. 1986. Thermotoga maritima sp. nov. represents a new genus of unique extremely thermophilic eubacteria growing up to 90°C. Archives of Microbiology. 144(4):324-333.

Hugenholtz, P. and E. Stackebrandt. 2004. Reclassification of Sphaerobacter thermophilus from the subclass Sphaerobacteridae in the phylum Actinobacteria to the class Thermomicrobia (emended description) in the phylum Chloroflexi (emended description). International Journal of Systematic and Evolutionary Microbiology. 54:2049-2051.

Iebba, V., Santangelo, F., Totino, V., Nicoletti, M., Gagliardi, A., De Biase, R. V., Cucchiara, S., Nencioni, L., Conte, M. P., and Schippa, S. (2013). Higher prevalence and abundance of Bdellovibrio bacteriovorus in the human gut of healthy subjects. PloS one, 8(4), e61608. http://doi.org/10.1371/journal.pone.0095166. doi:10.1371/journal.pone.0061608

Iebba, V., Totino, V., Santangelo, F., Gagliardi, A., Ciotoli, L., Virga, A., Ambrosi, C., Pompili, M., De Biase, R. V., Selan, L., Artini, M., Pantanella, F., Mura, F., Passariello, C., Nicoletti, M., Nencioni, L., Trancassini, M., Quattrucci, S., … Schippa, S. (2014). Bdellovibrio bacteriovorus directly attacks Pseudomonas aeruginosa and Staphylococcus aureus Cystic fibrosis isolates. Frontiers in microbiology, 5, 280. doi:10.3389/fmicb.2014.00280

Imhoff, J.F., T. Rahn, S. Künzel, and S.C. Neulinger. 2017. Photosynthesis Is Widely Distributed among Proteobacteria as Demonstrated by the Phylogeny of PufLM Reaction Center Proteins. Front Microbiol. 8:2679. doi:10.3389/fmicb.2017.02679.

Ishida, K., T. Sekizuka, K. Hayashida, J. matsuo, F. Takeuchi, M. Kuroda, S. Nakamura, T. Yamazaki, M. Yoshida, K. Takahashi,H. Nagai, C. Sugimoto, and H. Yamaguchi. 2014. Amoebal Endosymbiont Neochlamydia Genome Sequence Illuminates the Bacterial Role in the Defense of the Host Amoebae against Legionella pneumophila. PLOS One. 9(4):e95166.

Issotta, F., P.A. Galleguillos, A. Moya-Beltrán, C.S. Davis-Belmar, G. Rautenbach, P.C. Covarrubias, M. Acosta, F.J. Ossandon, Y. Contador, D.S. Holmes, S. Marin-Eliantonio, R. Quatrini, and C. Demergasso. 2016. Draft genome sequence of chloride-tolerant Leptospirillum ferriphilum Sp-Cl from industrial bioleaching operations in northern Chile. Stand. Genomic Sci. 11:19. doi:10.1186/s40793-016-0142-1.

Jäger J., S. Keese, M. Roessle, M. Steinert, and A.B. Schromm. 2015. Fusion of Legionella pneumophila outer membrane vesicles with eukaryotic membrane systems is a mechanism to deliver pathogen factors to host cell membranes. Cellular Microbiology. 17(5):607-620. doi:10.1111/cmi.12392.

Jamin, N. M. Garrigos, C. Jaxel, A. Frelet-Barrand, and S. Orlowski. 2018. Ectopic neo-formed intracellular membranes in Escherichia coli: a response to membrane protein-induced stress involving membrane curvature and domains. Biomolecules. 8:88. doi:10.3390/biom8030088.

Jeske, O., M. Schüler, P. Schumann, A. Schneider, C. Boedeker, M. Jogler, D. Bollschweiler, M. rohde, C. Mayer, H. Engelhardt, S. Spring, and C. Jogler. 2015. Planctomycetes do possess a peptidoglycan cell wall. Nature Communications. 6:7116. doi: 10.1038/ncomms8116.

Johnson, E.L., S.L. Heaver, W.A. Walters, and R.E. Ley. 2017. Microbiome and metabolic disease: revisiting the bacterial phylum Bacteroidetes. J Mol Med (Berl). 95(1):1-8.

Kadurugamuwa J.L. and T.J. Beveridge. 1996. Bacteriolytic effect of membrane vesicles from Pseudomonas aeruginosa on other bacteria including pathogens: conceptually new antibiotics. J. Bacteriol. 178:2767-2774.

Kant, R., M.W.J. van Passel, P. Sangwan, A. Palva, S. Lucas, A. Copeland, A. Lapidus, T. Glavina del Rio, E. Dalin, H. Tice, D. Bruce, L. Goodwin, S. Pitluck, O. Chertkov, F.W. Larimer, M.L. Land, L. Hauser, T.S. Brettin, J.C. Detter, S. Han, W.M. de Vos, P.H. Janssen, and H. Smidt. 2011. Genome sequence of "Pedosphaera parvula" Ellin514, an aerobic Verrucomicrobial isolate from pasture soil. Journal of Bacteriology. 193(11):2900-2901.

Kartal, B., J.G. Kuenen, and M.C.M. van Loosdrecht. 2010. Sewage treatment with anammox. Science. 328(5979):702-703.

Kashefi, K., D.E. Holmes, A.-L. Reysenbach, and D.R. Lovley. 2002. Use of Fe(III) as an electron acceptor to recover previously uncultured hyperthemophiles: isolation and characterization of Geothermobacterium ferrireducens gen. nov., sp. nov. Appl Environ Microbiol. 68(4):1735-1742.

Keane, R. and J. Berleman. 2016. The predatory life cycle of Myxococcus xanthus. Microbiology. 162:1-11. doi:10.1099/mic.0.000208.

Keck M., N. Gisch, H. Moll, F.-J. Vorhölter, K. Gerth, U. Kahmann, M. Lissel, B. Lindner, K. Niehaus, and O. Holst. 2011. Unusual outer membrane lipid composition of the gram-negative, lipopolysaccharide-lacking myxobacterium Sorangium cellulosum So ce56. The Journal of Biological Chemistry. 286(15):12850-12859.

Kennedy M.J., S.L. Reader, and L.M. Swierczynski. 1994. Preservation records of micro-organisms: evidence of the tenacity of life. Microbiology. 140:2513-2519.

Kersters, I., G.M. Maestrojuan, U. Torck, M. Vancanneyt, K. Kersters, and W. Verstraete. 1994. Isolation of Coprothermobacter proteolyticus from an Anaerobic Digest and Further Characterization of the Species. Systematic and Applied Microbiology. 17(2):289-295.

Kielak, A.M., C.C. Barreto, G.A. Kowalchuk, J.A. van Veen, and E.E. Kuramae. The ecology of Acidobacteria: moving beyond genes and genomes. Front. Microbiol. 7:744. doi: 10.3389/fmicb.2016.00744.

Kim K.H., S. Aulakh, and M. Paetzel. 2012. The bacterial outer membrane

β-barrel assembly machinery. Protein Science. 21(6):751-768.

Kits, K.D., C.J. Sedlacek, E.V. Lebedeva, P. Han, A. bulaev, P. Pjevac, A. Daebeler, S. Romano, M. Albertsen, L.Y. Stein, H. Daims, and M. Wagner. 2017. Kinetic analysis of a complete nitrifier reveals an oligotrophic lifestyle. Nature. 549(7671):269-272.

Klemke, F., D.J. Nürnberg, K. Ziegler, G. Beyer, U. Kahmann, W. Lockau, and T. Volkmer. 2016. CphA2 is a novel type of cyanophycin synthetase in N2-fixing cyanobacteria. Microbiology. 162:526-536. doi:10.1099/mic.0.000241.

Klenk, H.-P., A. Lapidus, O. Chertkov, A. Copeland, T.G. Del Rio, M. Nolan, S. Lucas, F. Chen, H. Tice, J.-F. Cheng, C. Han, D. Bruce, L. Goodwin, S. Pitluck, A. Pati, N. Ivanova, K. Mavromatis, C. Daum, A. Chen, K. Palaniappan, Y.-J. Chan, M. Land, L. Hauser, C.D. Jeffries, J.C. Detter, M. Rohde, B. Abt, R. Pukall, M. Göker, J. Bristow, V. Markowitz, P. Hugenholtz, and J.A. Eisen. 2011. Complete genome sequence of the thermophilic, hydrogen-oxidizing Bacillus tusciae type strain (T2T) and reclassification in the new genus, Kyrpidia gen. nov. as Kyrpidia tusciae comb. nov. and emendation of the family Alicyclobacillaceae da Costa and Rainey, 2010. 2011. Stand. Genomic Sci. 5(1):121-134.

Koch, A.L. and R.J. Doyle 1986. The growth strategy of the gram-positive rod. FEMS Microbiology Reviews. 32:247-254.

Koebnik R., K.P. Locher, and P. Van Gelder. 2000. Structure and function of bacterial outer membrane proteins: barrels in a nutshell. Molecular Microbiology. 37(2):239-253.

Komeili, A., Z. Li, D.K. Newman, and G.J. Jensen. 2006. Magnetosomes are cell membrane invaginations organized by the actin-like protein MamK. Science. 311:242-245.

Konstandinidis, K., A. Ramette, and J.M. Tiedje. 2006. The bacterial species definition in the genomic era. Phil. Trans. R. Soc. B. 361:1929-1940.

Koonin, E.V. 2011. The logic of chance: the nature and origin of biological evolution. FT Press Science. Upper Saddle River, NJ.

Kovacs-Simon A., R.W. Titball, and S.L. Michell. 2011. Lipoproteins of bacterial pathogens. Infection and Immunity. 79(2):548-561.

Kublanov, I.V., O.M. Sigalova, S.N Gavrilov, A.V. Lebedinsky, C. Rinke, O. Kovaleva, N.A. Chernyh, N. Ivanova, C. Daum, T.B.K. Reddy, H.-P. Klenk, S. Spring, M. Göker, O.N. Reva, M.L. Miroshnichenko, N.C. Kyrpides, T. Woyke, M.S. Gelfand, and E.A. Bonch-Osmolovskaya. 2017. Genomic analysis of Caldithrix abyssi, the thermophilic anaerobic bacterium of the novel bacterial phylum Calditrichaeota. Frontiers in Microbiology. Vol. 8, Article 195.

Kulp A. and M.J. Kuehn. 2010. Biological functions and biogenesis of secreted bacterial outer membranes. Annu Rev Microbiol. 64:163-184. doi:10.11146/annurev.micro.091208.073413.

Kumaresan, D., J. Stephenson, A.C. Doxey, H. Bandukwala, E. Brooks, A. Hillebrand-Voiculescu, A.S. Whiteley and J.C. Murrell. 2018. Aerobic proteobacterial methylotrophs in Movile Cave: genomic and metagenomic analyses. Microbiome. 6:1. https://doi.org/10.1186/s40168-017-0383-2.

Küper U., C. Meyer, V. Müller, R. Rachel, and H. Huber. 2010. Energized outer membrane and spatial separation of metalic processes in the hyperthermophilic archaeon Ignicoccus hospitalis. PNAS. 107(7):3152-3156.

Kuzminov, A. 2013. The chromosome cycle of prokaryotes. Mol Microbiol. 90(2):214-227. doi:10.1111/mmi.12372.

Lapidus, A., O Chertkov, M. Nolan, S. Lucas, N. Hammon, S. Deshpande, J.-F. Cheng, R. Tapia, C. Han, L. Goodwin, S. Pitluck, K. Liolios, I. Pagani, N. Ivanova, M. Huntemann, K. Mavromatis, N. Mikhailova, A. Pati, A. Chen, K. Palaniappan, M. Land, L. Hauser, E.-M. Brambilla, M. Rohde, B. Abt, S. Spring, M. Göker, J. Bristow, J.A. Eisen, V. Markowitz, P. Hugenholtz, N.C. Kyrpides, H.-P. Klenk, and T. Woyke. 2011. Genome sequence of the moderately thermophilic halophile Flexistipes sinusarabici strain (MAS10T). Stand Genomic Sci. 5(1):86-96.

LaSarre, B. D.T. Kysela, B.D. Stein, A. ducret, Y.V. Brun, and J.B. McKinlay. 2018. Restricted Localization of Photosynthetic Intracytoplasmic Membranes (ICMs) in Multiple Genera of Purple Nonsulfur Bacteria. mBio. 9(4):e00780-18. doi:10.1128/mBio.00780-18.

Lau, M.C.Y., T.L. Kieft, O. Kuloyo, B. Linage-Alvarez, E. van Heerden, M.R. Lindsay, C. Magnabosco, W. Wang, J.B. Wiggins, L. Guo, D.H. Perlman, S. Kyin, H.H. Shwe, R.L. Harris, Y. Oh, M.J. Yi, R. Purtschert, G.F. Slater, S. Ono, S. Wei, L. Li, B.S. Lollar, and T.C. Onstott. 2016. An oligotrophic deep-subsurface community dependent on syntrophy is dominated by sulfur-driven autotrophic denitrifiers. PNAS. 113(49):E7927-E7936. https://doi.org/10.1073/pnas.1612244113.

Lawrence J.G. and H. Ochman. 1997. Amelioration of bacterial genomes: rates of change and exchange. J. Mol Evol. 44:383-397.

Le Bourgeois P. and F. Cornet. 2013. Chromosome dimer resolution by site-specific recombination. In: Stanley Maloy and Kelly Hughes, editors. Brenner's Encyclopedia of Genetics 2nd edition, Vol 1. San Diego: Academic Press. pp. 551-554.

Lee H.-H., C.-C. Hsu, Y.-L. Lin, and C.W. Chen. 2011. Linear plasmids mobilize linear but not circular chromosomes in Streptomyces: support for the 'end first' model of conjugal transfer. Microbiology. 157:2556-2568. doi: 10.1099/mic.0.051441-0.

Lee, K.C., M.B. Stott, P.F. Dunfield, C. Huttenhower, I.R. McDonald, X.C. Morgan. 2016. The Chthonomonas calidirosea genome is highly conserved across geographic locations and distinct chemical and microbial environments in New Zealand's Taupō Volcanic Zone. Applied and Environmental Microbiology. 82(12):3572-3581. Envir

Lerouge I. and J. Vanderleyden. 2001. O-antigen structural variation: mechanisms and possible roles in animal/plant-microbe interactions. FEMS Microbiology Reviews. 26:17-47.

Levin, P.A. and E.R. Angert. 2015. Small but mighty: cell size and bacteria. Cold Spring Harb Perspect Biol. 7(7):a019216. doi: 10.1101/cshperspect.a019216.

Liaimer, A., E.J.N. Helfrich, K. Hinrichs, A. Guljamow, K. Ishida, C. Hertweck and E. Dittmann. 2015. Nostopeptolide plays a governing role during cellular differentiation of the symbiotic cyanobacterium Nostoc punctiforme. Proc natl Acad Sci USA. 112(6):1862-1867. doi:10.1073/pnas.1419543112.

Lilburn, T.G., K.S. Kim, N.E. Ostrom, K.R. Byzek, J.R. Leadbetter, and J.A. Breznak. 2001. Nitrogen fixation by symbiotic and free-living spirochetes. Science. 292(5526):2495-2498.

Lin, W., A. Deng, Z. Wang, Y. Li, T. Wen, L.-F. Wu, M. Wu, and Y. Pan. 2014. Genomic insights into the uncultured genus 'Candidatus Magnetobacterium' in the phylum Nitrospirae. ISME J. 8(12):2463-2477. doi:10.1038/ismej.2014.94.

Lin, W., W. Zhang, X. Zhao, A.P. Roberts, G.A. Paterson, D.A. Bazylinski, and Y. Pan. 2018. Genomic expansion of magnetotactic bacteria reveals an early common origin of magnetotaxis with lineage-specific evolution. The ISME Journal. 12:1508-1519. https://doi.org/10.1038/s41396-018-0098-9.

Liu, F., J. Li, G. Feng, and Z. Li. 2016. New genomic insights into "Entotheonella" symbionts in Theonella swinhoei: mixotrophy, anaerobic adapation, resilience, and interaction. Front. Microbiol. 7:1333. doi: 10.3389/fmicb.2016.01333.

Liu Y., K.A.Y. Defourny, E.J. Smid, and T. Abee. 2018. Gram-positive bacterial extracellular vesicles and their impact on health and disease. Frontiers in Microbiology. 9:1502. doi: 10.3389/fmicb.2018.01502.

Lonhienne, T.G.A., E. Sagulenko, R.I. Webb, K. Lee, J. Franke, D.P. Devos, A. Nouwens, B.J. Carroll, and J.A. Fuerst. 2010. Endocytosis-like protein uptake in the bacterium Gemmata obscuriglobus. PNAS, 107(29):12883-12888.

Loy, A.,C. Pfann, M. Steinberger, B. Hanson, S. Herp, S. Brugiroux, J.C. Gomes Neto, M.V. Boekschoten, C. Schwab, T. Urich, A.E. Ramer-Tait, T. Rattei, B. Stecher, D. Berry. 2017. Lifestyle and Horizontal Gene Transfer-Mediated Evolution of Mucispirillum schaedleri, a Core Member of the Murine Gut Microbiota. mSystems. 2(1):e00171-16. doi: 10.1128/mSystems.00171-16.

Löffler, F.E., J. Yan, K.M. Ritalhti, L. Adrian, E.A. Edwards, K.T. Konstantinidis, J.A. Müller, H. Fullerton, S.H. Zinder, and A.M. Spormann. 2013. Dehalococcoides mccartyi gen. nov., sp. nov., obligately organohalide-respiring anaerobic bacteria relevant to halogen cycling and bioremediation, belong to a novel bacterial class, Dehalococcoidia classis nov., order Dehalococcoidales ord. nov. and family Dehalococcoidaceae fam. nov., within the phylum Chloroflexi. International Journal of Systematic and Evolutionary Microbiology. 63:625-635.

Lue N.F. and S. Jiang. 2004. Reverse transcriptase at bacterial telomeres. PNAS, 101(40):14307-14308.

Luef, B., K.R. Frischkorn, K.C. Wrighton, H-Y.N. Holman, G. Birarda, B.C. Thomas, A. Singh, K.H. Williams, C.E. Siegerist, S.G. Tringe, K.H. Downing, L.R. Comolli, and J.F. Banfield. 2015. Diverse uncultivated ultra-small bacterial cells in groundwater. Nature Communications. 6:6372. DOI: 10.1038/ncomms7372.

Lücker, S., B. Nowka, T. Rattei, E. Spieck, and H. Daims. 2013. The Genome of Nitrospina gracilis Illuminates the Metabolism and Evolution of the Major Marine Nitrite Oxidizer. Front. Microbiol. 4:27. doi:10.3389/fmicb.2013.00027.

MacLean R.C. and A. San Millan. 2015. Microbial evolution: towards resolving the plasmid paradox. Current Biology. 25:R764-R767. doi: 10.1016/j.cub.2015.07.006.

Macy, J.M., K. Nunan, K.D. Hagen, D.R. Dixon, P.J. Harbour, M. Cahill, and L.I. Sly. 1996. Chrysiogenes arsenatis gen. nov., sp. nov., a new arsenate-respiring bacterium isolated from gold mine wastewater. Int J Syst Bacteriol. 46(4):1153-1157.

Mahat, R., C. Seebart, F. Basile, and N.L. Ward. 2016. Global and targeted lipid analysis of Gemmata obscuriglobus reveals the presence of lipopolysaccharide, a signature of the classical gram-negative outer membrane. Journal of Bacteriology. 198(2):221-236.

Majekodunmi, S.O. 2015. A review on centrifugation in the pharmaceutical industry. American Journal of Biomedical Engineering. 5(2):67-78. doi:10.5923/j.ajbe.20150502.03.

Maldonado R., J. Jiménez, and J. Casadesús. 1994. Changes of ploidy during the Azotobacter vinelandii growth cycle. Journal of Bacteriology. 176(13):3911-3919.

Manning A.J. and M.J. Kuehn. 2011. Contribution of bacterial outer membrane vesicles to innate bacterial defense. BMC Microbiology. 11:258. http://www.biomedcentral.com/1471-2180/11/258.

Marchandin, H. C. Teyssier, J. Campos, H. Jean-Pierre, F. Roger, B. Gay, J.-P. Carlier, and E. Jumas-Bilak. 2010. Negativicoccus succinicivorans gen. nov., sp. nov., isolated from human clinical samples, emended description of the family Veillonellaceae and description of Negativicutes classis nov., Selenomonadales ord. nov. and Acidaminococcaceae fam. nov. in the bacterial phylum Firmicutes. International Journal of Systematic and Evolutionary Microbiology. 60:1271-1279.

Mardanov A.V. and N.V. Ravin. 2012. The impact of genomics on research in diversity and evolution of archaea. Biochemistry (Mosc) 77:799-812. http://dx.doi.org/10.1134/S0006297912080019.

Mason, O.U., T. Nakagawa, M. Rosner, J.D. van Nostrand, J. Zhou, A. Maruyama, M.R. Fisk, and S.J. Giovannoni. 2010. First investigation of the microbiology of the deepest layer of ocean crust. PLoS ONE. 5(11):e15399. doi:10.1371/journal.pone.0015399.

Matsuura M. 2013. Structural modifications of bacterial lipopolysaccharide that facilitate gram-negative bacteria evasion of host innate immunity. Frontiers in immunology. 4:109. doi:10.3389/fimmu.2013.00109.

Maune, M.W. and R.S. Tanner. 2012. Description of Anaerobaculum hydrogeniformans sp. nov., an anaerobe that produces hydrogen from glucose, and emended description of the genus Anaerobaculum. Int. J. Syst. Evol. Microbiol. 62:832-838. doi:10.1099/ijs.0.024349-0.

McBroom A.J. and M.J. Kuehn. 2007. Release of outer membrane vesicles by Gram-negative bacteria is a novel envelope stress response. Mol. Microbiol. 63(2):545-558. doi: 10.1111/j.1365-2958.2006.05522.x.

McCauley, E.P., B. Haltli, and R.G. Kerr. 2015. Description of Pseudobacteriovorax antillogorgiicola gen. nov., sp. nov., a bacterium isolated from the gorgonian octocoral Antillogorgia elisabethae, belonging to the family Pseudobacteriovoracaceae fam. nov., within the order Bdellovibrionales. International Journal of Systematic and Evolutionary Microbiology. 65:522-530.

McKay, S.L. and Portnoy, D.A. 2015. Ribosome hibernation facilitates tolerance of stationary-phase bacteria to aminoglycosides. antimicrob Agents Chemother. 59:6992-6999. doi:10.1128/AAC.01532-15.

McLellan, N.L. and R.A. Manderville. 2017. Toxic mechanisms of microcystins in mammals. Toxicol Res (Camb). 6(4):391-405.

Melton, E.D., D.Y. Sorokin, L. Overmars, A.L. Lapidus, M. Pillay, N. Ivanova, T.G. del Rio, N.C. Kyrpides, T. Woyke, and G. Muyzer. 2017. Draft genome sequence of Dethiobacter alkaliphilus strain AHT1T, a gram-positive sulfidogenic polyextremophile. Stand. Genomic Sc. 12:57. doi:10.1186/s40793-017-0268-9.

Merialdi, G., M. Ramini, G. Parolari, S. Barbuti, M.A. Frustoli, R. Taddei, S. Pongolini, P. Ardigò, and P. Cozzolino. 2016. Study on Potential Clostridium Botulinum Growth and Toxin Production in Parma Ham. Italian Journal of Food Safety. 5(2):5564. doi:10.4081/ijfs.2016.5564.

Meriläinen, L., A. Herranen, A. Schwarzbach, and L. Gilbert. 2015. Morphological and biochemical features of Borrelia burgdorferi pleomorphic forms. Microbiology. 161(Pt3):516-527. doi:10.1099/mic.0.000027.

Mesbah, N.M., G.M. Cook, and J. Wiegel. 2009. The halophilic alkalithermophile Natranaerobius thermophilus adapts to multiple environmental extremes using a large repertoire of Na+(K+)/H+ antiporters. Molecular Microbiology. 74(2):270-281.

Mesbah, N.M., D.B. Hedrick, A.D. Peacock, M. Rohde, and J. Wiegel. 2007. Natranaerobius thermophilus gen. nov., sp. nov., a halophilic, alkalithermophilic bacterium from soda lakes of the Wadi An Natrun, Egypt, and proposal of Natranaerobiaceae fam. nov. and Natranaerobiales ord. nov. International Journal of Systematic and Evolutionary Microbiology. 57:2507-2512. doi:10.1099/ijs.0.65068-0.

Miller, O.L., B.A. Hamkalo, and C.A. Thomas. 1970. Visualization of bacterial genes in action. Science. 169:392-395.

Miroshnichenko, M.L., G.A Gongadze, A.M. Lysenko, and E.A. Bonch-Osmolovskaya. 1994. Desulfurella multipotens sp. nov., a new sulfur-respiring thermophilic eubacterium from Raoul Island (Kermadec archipelago, New Zealand). Archives of Microbiology. 161:88-93. 10.1007/BF00248898.

Miroshnichenko, M.L., N.A. Kostrikina, N.A. Chernyh, N.V. Pimenov, T.P. Tourova, A.N. Antipov, S. Spring, E. Stackebrandt, and E.A. Bonch-Osmolovskaya. 2003. Caldithrix abyssi gen. nov., sp. novl, a nitrate-reducing, thermophilic, anaerobic bacterium isolated from a Mid-Atlantic ridge hydrothermal vent, represents a novel bacterial lineage. Int J Syst Evol Microbiol. 53(Pt 1):323-329.

Mitchell, P. and J. Moyle. 1956. Osmotic function and structure in bacteria. Symp. Soc. Gen. Microbiol. 6:150-180.

Mohammadi, S. A. Pol, T.A. van Alen, M.S.M. Jetten, and H.J.M. Op den Camp, H.J.M. 2017. Methylacidiphilum fumariolicum SolV, a thermoacidophilic ‘Knallgas’ methanotroph with both an oxygen-sensitive and -insensitive hydrogenase. The ISME Journal. 11:945-958.

Mori, K. , K. Hongik, T. Kakegawa, and S. Hanada. 2003. A novel lineage of sulfate-reducing microorganisms: Thermodesulfobiaceae fam. nov., Thermodesulfobium narugense, gen. nov., sp. nov., a new thermophilic isolate from a hot spring. Extremophiles.7:283-290.

Mori, K., K. Yamaguchi, Y. Sakiyama, T. Urabe, and K.-I. Suzuki. 2009. Caldisericum exile gen. nov., sp. nov., and anaerobic, thermophilic, filamentous bacterium of a novel bacterial phylum, Caldiserica phyl. nov., originally called the candidate phylum OP5, and description of Caldisericaceae fam. nov., Caldisericales ord. nov. and Caldisericia classis nov. International Journal of Systematic and Evolutionary Microbiology. 59:2894-2898.

Morigen M., I. Odsbu, and K. Skarstad. 2009. Growth rate dependent numbers of SeqA structures organize the multiple replication forks in rapidly growing Escherichia coli. Genes to Cells. 14:643-657.

Muller, H. J. 1964. The relation of recombination to mutational advance. Mutat. Res. 106:2–9.

Murphy, D.J. 2012. The dynamic roles of intracellular lipid droplets: from archaea to mammals. Protoplasma. 249(3):541-585. doi:10.1007/s00709-011-0329-7.

Nagata, R., Y. Takaki, A. Tame, T. Nunoura, H. Muto, S. Mino, S. Sawayama, K. Takai, and S. Nakagawa. 2017. Lebetimonas natsushimae sp. nov., a novel strictly anaerobic, moderately thermophilic chemoautotroph isolated from a deep-sea hydrothermal vent polychaete nest in the Mid-Okinawa Trough. Syst Appl Microbiol. 40(6):352-356.

Nakagawa, S. and K. Takai. 2008. Deep-sea vent chemoautotrophs: diversity, biochemistry and ecological significance. FEMS Microbiology Ecology. 65(1):1-14. https://doi.org/10.1111/j.1574-6941.2008.00502.x

Nakai, R., Nishijima, M., Tazato, N., Handa, Y., Karray, F., Sayadi, S., Isoda, H., and Naganuma, T. (2014). Oligoflexus tunisiensis gen. nov., sp. nov., a Gram-negative, aerobic, filamentous bacterium of a novel proteobacterial lineage, and description of Oligoflexaceae fam. nov., Oligoflexales ord. nov. and Oligoflexia classis nov. International journal of systematic and evolutionary microbiology, 64(Pt 10), 3353-9.

National Center for Biotechnology Information (NCBI). Taxonomy Database. http://https://www.ncbi.nlm.nih.gov/guide/taxonomy/. Accessed December 2018.

Nelson K., R. Clayton, S. Gill, et al. 1999. Evidence for lateral gene transfer between Archaea and Bacteria from genome sequence of Thermotoga maritima. Nature. 399:323–329. https://doi.org/10.1038/20601

Newmann, S., H.J.C.T. Wessels, W.I.C. Rijpstra, J.S. Sinninghe Damsté, B. Kartal, M.S.M. Jetten, and L. van Niftrik. 2014. Isolation and characterization of a prokaryotic cell organelle from the anammox bacterium Kuenenia stuttgartiensis. Molecular Microbiology. 94(4):794-802.

Nicholson W.L., N. Munakata, G. Horneck, H.J. Melosh, and P. Setlow. 2000. Resistance of Bacillus endospores to extreme terrestrial and extraterrestrial environments. Microbiology and Molecular Biology Reviews. 64(3):548-572.

Niederman, R.A. 2006. Structure, function and formation of bacterial intracytoplasmic membranes. In Shively, J.M. (Ed.), Complex Intracellular Structures in Prokaryotes. Springer-Verlag Berlin Heidelberg.

Nikaido H. 2003. Molecular basis of bacteria outer membrane permeability revisited. Microbiology and Molecular Biology Reviews. 67(4):593-656.

Nunoura, T., H. Oida, M. Miyazaki, and Y. Suzuki. 2008. Thermosulfidibacter takaii gen. nov., sp. nov., a thermophilic, hydrogen-oxidizing, sulfur-reducing chemolithoautotroph isolated from a deep-sea hydrothermal field in the Southern Okinawa Trough. Int. J. Syst. Evol. Microbiol. 58(Pt 3):659-665.

Ochman H. 2002. Bacterial evolution: chromosome arithmetic and geometry. Current Biology. 12:R427-R428.

Ogg, C.D., A.C. Greene, and B.K.C. Patel. 2010. Thermovenabulum gondwanense sp. nov., a thermophilic anaerobic Fe(III)-reducing bacterium isolated from microbial mats thriving in a Great Artesian Basin bore runoff channel. International Journal of Systematic and Evolutionary Microbiology. 60:1079-1084.

Ohno, M., H. Shiratori, M.-J. Park, Y. Saitoh, Y. Kumon, N. Yamashita, A. Hirata, H. Nishida, K. Ueda, and T. Beppu. 2000. Symbiobacterium thermophilum gen. nov., sp. nov., a symbiotic thermophile that depends on co-culture with a Bacillus strain for growth. International Journal of Systematic and Evolutionary Microbiology. 50:1828-1832.

Olabarria,G. J.L. Carrascosa, M.A. De Pedro, and J. Berenguer. 1996. A conserved motif in S-layer proteins is involved in peptidoglycan binding in Thermus thermophilus. Journal of Bacteriology. 178(16):4765-4772.

Opiyo S.O., R.L. Pardy, H. Moriyama, and E.N. Moriyama. 2010. Evolution of Kdo2-lipid A biosynthsis in bacteria. BMC Evolutionary Biology. 10:362. http://www.biomedcentral.com/1471-2148/10/362.

Oshima, K., K. Ueda, T. Beppu, and H. Nishida. 2011. Unique Evolution of Symbiobacterium thermophilum Suggested from Gene Content and Orthologous Protein Sequence Comparisons. International Journal of Evolutionary Biology. 2011:Article ID 376831. http://dx.doi.org/10.4061/2011/376831.

Padan, E., E. Bibi, M. Ito, and T.A. Krulwich. 2005. Alkaline pH homeostasis in bacteria: new insights. Biochim Biophys Acta. 1717(2):67-88.

Pagnier, I., N. Yutin, O. Croce, K.S. Makarova, Y.I. Wolf, S. Benamar, D. Raoult, E.V. Koonin, and B. La Scola. 2015. Babela massiliensis, a representative of a widespread bacterial phylum with unusual adaptations to parasitism in amoebae. Biol Direct. 10:13.

Parks, D.H., M. Chuvochina, D.W. Waite, C. Rinke, A. Skarshewski, P.-A. Chaumeil, and P. Hugenholtz. 2018. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nature Biotechnology. doi:10.1038/nbt.4229.

Patt, T.E. and R.S. Hanson. 1978. Intracytoplasmic membrane, phospholipid, and sterol content of Methylobacterium organophilum cells grown under different conditions. Journal of Bacteriology. 134(2):636-644.

Pecoraro V., K. Zerulla, C. Lange, and J. Soppa. 2011. Quantification of ploidy in proteobacteria revealed the existence of monoploid, (mero-)oligoploid and polyploid species. PLoS ONE 6(1):e16392. doi:10.1371/journal.pone.0016392.

Peter B.J., C. Ullsperger, H. Hiasa, K.J. Marians, and N.R. Cozzarelli. 1998. The structure of supercoiled intermediates in DNA replication. Cell. 94(6):819-827.

Picardeau M., J.R. Lobry, and B.J. Hinnebusch. 1999. Physical mapping of an origin of bidirectional replication at the centre of the Borrelia burgdorferi linear chromosome. Molecular Microbiology. 32(2):437-445.

Planchenault C., M. Pons, C. Schiavon, P. Siguier, J. Rech, C. Guynet, J. Dauverd-Girault, J. Cury, E. Rocha, I. Junier, F. Cornet, and O. Espéli. 2020. Intracellular positioning systems limit the entropic eviction of secondary replicons toward the nucleoid edges in bacterial cells. Journal of Molecular Biology. 432(3):745-761. https://doi.org/10.1016/j.jmb.2019.11.027.

Por, F.D., Ch. Dimentman, A. Frumkin, I. Naaman. 2013. Animal life in the chemoautotrophic ecosystem of the hypogenic groundwater cave of Ayyalon (Israel): A summing up. Natural Science. 5(4A):article ID:30808, 7 pages. doi:10.4236/ns.2013.54A002.

Poirion O.B. and B. Lafay. 2018. Neo-formation of chromosomes in bacteria. bioRxiv. Preprint server for biology. Posted August 27, 2018. doi: https://doi.org/10.1101/264945.

Potts M. 1994. Desiccation tolerance of prokaryotes. Microbiol. Rev. 58:755-805.

Prescott, L.M., J.P. Harley, and D.A. Klein. Microbiology. Second Edition. 1993. Wm. C. Brown Publishers. Dubuque, Iowa.

Preston A., R.E. Mandrell, B.W. Gibson, and M.A. Apicella. The lipooligosaccharides of pathogenic gram-negative bacteria. Crit Rev Microbiol. 22(3):139-180.

Prowe, S.G. and G. Antranikian. 2001. Anaerobranca gottschalkii sp. nov., a novel thermoalkaliphilic bacterium that grows anaerobically at high pH and temperature. Int. J. Syst. Evol. Microbiol. 51:457-465.

Qiu, Y.-L., S. Hanada, A. Ohashi, H. Harada, Y. Kamagata, and Y. Sekiguchi. 2008. Syntrophorhabdus aromaticivorans gen. nov., sp. nov., the first cultured anaerobe capable of degrading phenol to acetate in obligate syntrophic associations with a hydrogenotrophic methanogen. Appl environ Microbiol. 74(7):2051-2058.

Quatrini, R. and D.B. Johnson. 2018. Microbiomes in extremely acidic environments: functionalities and interactions that allow survival and growth of prokaryotes at low pH. Current Opinion in Microbiology. 43:139-147.

Rabus, R., A. Ruepp, T. Frickey, T. Rattei, B. Fartmann, M. Stark, M. Bauer, A. Zibat, T. Lombardot, I. Becker, J. Armann, K. Gellner, H. Teeling, W.D. Leuschner, F.O. Glöckner, A.N. Lupas, R. Amann, and H.P. Klenk. 2004. Environ Microbiol. 6(9):887-902.

Radolf, J.D., R.K. Deka, A. Anand, D. Šmajs, M.V. Norgard, and X.F. Yang. 2016. Treponema pallidum, the syphilis spirochete: making a living as a stealth pathogen. Nat Rev microbiol. 14(12):744-759.

Raetz C.R.H. and C. Whitfield. 2002. Lipopolysaccharide endotoxins. Annu Rev Biochem. 71:635-700. doi:10.1146/annurev.biochem.71.110601.135414.

Rahman, N.A., D.H. Parks, I. Vanwonterghem, M. Morrison, G.W. Tyson and P. Hugenholtz. 2015. A phylogenomic analysis of the bacterial phylum Fibrobacteres. Front Microbiol. 6:1469. doi:10.3389/fmicb.2015.01469.

Ramachandran G. 2014. Gram-positive and gram-negative bacterial toxins in sepsis. A brief review. Virulence. 5(1):213-218.

Rauschenbach, I., P. Narasingarao, and M.M. Häggblom. 2011. Desulfurispirillum indicum sp. nov., a selenate- and selenite-respiring bacterium isolated from an estuarine canal. International Journal of Systematic and Evolutionary Microbiology. 61:654-658.

Reyes-Lamothe R., E. Nicolas, and D.J. Sherratt. 2012. Chromosome replication and segregation in bacteria. Annu. Rev. Genet. 46:121-143. doi:10.1146/annurev-genet-110711-155421.

Reysenbach, A.L., Y. Liu, A.R. Lindgren, I.D. Wagner, C.D. Sislak, A. Mets, S. Schouten. 2013. Mesoaciditoga lauensis gen. nov., sp. nov., a moderately thermoacidophilic member of the order Thermotogales from a deep-sea hydrothermal vent. Int. J. Syst. Evol. Microbiol. 63(Pt 120:4724-4729.

Rivkina, E.M., E.I. Friedmann, C.P. MacKay, and D.A. Gilichinsky. 2000. Metabolic activity of permafrost bacteria below the freezing point. Appl Environ Microbiol. 66(8):3230-3233.

Rosenberg, E., E.F. Delong, S. Lory, E. Stackebrandt, and F. Thompson (eds.). 2014. The Prokaryotes. Firmicutes and Tenericutes. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-30120-9.

Russell, J.B. and G.M. Cook. 1995. Energetics of bacterial growth: Balance of anabolic and catabolic reactions. Microbiol Rev 59:48–62.

Russell, J.B. and M.R. Garner. 2007. Allisonella histaminiformans, a novel histamine-producing bacterium that may play a role in rumen disorders and bovine laminitis. Engormix. 3/1/2007. https://en.engormix.com/dairy-cattle/articles/allisonella-histaminiformans-novel-histamine-t33595.htm.

Russell, J.B. and M.R. Garner. 2002. Allisonella histaminiformans gen. nov., sp. nov. A novel bacterium that produces histamine, utilizes histidine as its sole energy source, and could play a role in bovine and equine laminitis. Syst. Appl. Microbiol. 25(4):498-506.

Sagulenko, E., A. Nouwens, R.I. Webb, K. Green, B. Yee, G. Morgan, A. Leis, K.-C-. Lee, M.K. Butler, N. Chia, U. T. Phuong Pham, Stinus Lindgreen, R. Catchpole, A.M. Poole, and J.A. Fuerst. 2017. Nuclear pore-like structures in a compartmentalized bacterium. PLoS ONE. 12(2):e0169432. doi:10.1371/journal.pone.0169432.

Saiki, T., Y. Kobayashi, K. Kawagoe,and T. Beppu. 1985. Dictyoglomus thermophilum gen. nov. , sp. nov. , a Chemoorganotrophic, Anaerobic, Thermophilic Bacterium. International Journal of Systematic Bacteriology. 35(3):253-259.

Saiyari, D.M., H.-P. Chuang, D.B. Senoro, T.-F. Lin, L.-M. Whang, Y.-T. Chiu, and Y.-H. Chen. 2018. A review in the current developments of genus Dehalococcoides, its consortia and kinetics for bioremediation options of contaminated groundwater. Sustainable Environment Research. 28(4):149-157.

Salinas, M.B., M.L. Fardeau, P. Thomas, J.L. Cayol, B.K. Patel, and B. Ollivier. 2004. Mahella australiensis gen. nov., sp. nov., a moderately thermophilic anaerobic bacterium isolated from an Australian oil well. Int. J. Syst. Evol. Microbiol. 54(Pt6):2169-2173.

Salton M.R.J. and K.-S. Kim. 1996. 'Structure.' In Baron, S. (ed.) Medical Microbiology. 4th edition. Galveston (TX): University of Texas Medical Branch at Galveston. Chapter 2. https:////www.ncbi.nlm.nih.gov/books/NBK8477/.

Sanamrad A., F. Persson, E.G. Lundius, D. Fange, A.H. Gynnå, and J. Elf. 2014. Single-particle tracking reveals that free ribosomal subunits are not excluded from the Escherichia coli nucleoid. PNAS. 111(31):11413-11418.

Sand, W., K. Rohde, B. Sobotke, and C. Zenneck. 1992. Evaluation of Leptospirillum ferrooxidans for leaching. Appl. Environ. Microbiol. 58(1):85-92.

Sandle, T. 2004. Gram's Stain: History and explanation of the fundamental technique of determinative bacteriology, IST Science and Technology Journal. 54:3-4.

Santarella-Mellwig, R., S. Pruggnaller, N. Roos, I.W. Mattaj, and D.P. Devos. 2013. Three-dimensional reconstruction of bacteria with a complex endomembrane system. PLoS Biol. 11(5):e1001565. doi: 10.1371/journal.pbio.1001565.

Sattley, W.M., M.T. Madigan, W.D. Swingley, P.C. Cheung, K.M. Clocksin, A.L. Conrad, L.C. Dejesa, B.M. Honchak, D.O. Jung, L.E. Karbach, A. Kurdoglu, S. Lahiri, S.D. Mastrian, L.E. Page, H.L. Taylor, Z.T. Wang, J. Raymond, M. Chen, R.E. Blankenship, and J.W. Touchman. 2008. The genome of Heliobacterium modesticaldum, a phototrophic representative of the Firmicutes containing the simplest photosynthetic apparatus. J. Bacteriol. 190(13):4687-4696.

Schneiker, S., O. Perlova, O. Kaiser, K. Gerth, A. Alici, M.O. Altmeyer, D. Bartels, T. Bekel, S. Beyer, E. Bode, et al. 2007. Complete genome sequence of the myxobacterium Sorangium cellulosum. Nature Biotechnology. 25(11):1281-1289

Schooling S.R. and T.J. Beveridge. 2006. Membrane vesicles: an overlooked component of the matrices of biofilms. J. Bacteriol. 188(16):5945-5957.

Schulze-Makuch, D., D. Wagner, S.P. Kounaves, K. Mangelsdorf, K.G. Devine, J.-P. de Vera, P. Schmitt-Kopplin, H.-P. Grossart, V. Parro, M. Kaupenjohann, A. Galy, B. Schneider, A. Airo, J. Frösler, A.F. Davila, F.L. Arens, L. Cáceres, F.S. Cornejo, D. Carrizo, L. Dartnell, J. DiRuggiero, M. Flury, L. Ganzert, M.O. Gessner, P. Grathwohl, L. Guan, J. Heinz, M. Hess, F. Keppler, D. Maus, C.P. McKay, R.U. Meckenstock, W. Montgomery, E.A. Oberlin, A.J. Probst, J.S. Sáenz, T. Sattler, J. Schirmack, M.A. Sephton, M. Schloter, J. Uhl, B. Valenzuela, G. Vestergaard, L. Wörmer, and P. Zamorano. 2018. Transitory microbial habitat in the hyperarid Atacama Desert. PNAS. 115(11):2670-2675. https://doi.org/10.1073/pnas.1714341115.

Schulz, H.N., T. Brinkhoff, T.G. Ferdelman, M.H. Mariné, A. Teske and B.B. Jørgensen. 1999. Dense populations of a giant sulfur bacterium in Namibian shelf sediments. Science. 284(5413):493-495.

Seufferheld, M.J., K. Mo Kim, J. Whitfield, A. Valerio, and G. Caetano-Anollés. 2011. Evolution of vacuolar proton pyrophosphatase domains and volutin granules: clues into the early evolutionary origin of the acidocalcisome. Biology Direct. 6:50. http://www.biology-direct.com/content/6/1/50.

Seufferheld, M., M.C.F. Vieira, F.A. Ruiz, C.O. Rodrigues, S.N.J. Moreno, and R. Docampo. 2003. Identification of organelles in bacteria similar to acidocalcisomes of unicellular eukaryotes. The Journal of Biological Chemistry. 278(32):29971-29978.

Sharpe S.W., M.J. Kuehn, and K.M. Mason. 2011. Elicitation of epithelial cell-derived immune effectors by outer membrane vesicles of nontypeable Haemophilus influenzae. Infect Immun 79:4361–4369. doi:10.1128/IAI.05332-11.

Siddiqui, M.H., A. Kumar, K.K. Kesari, and J.M. Arif. 2009. Biomining— a useful approach toward metal extraction. American-Eurasian Journal of Agronomy. 2(2):84-88.

Silhavy T.j., D. Kahne, and S. Walker. 2010. The bacterial cell envelope. Cold Spring Harb Perspect Biol 2:a000414. doi: 10.1101/cshperspect.a000414.

Skarstad K., H.B. Steen, and E. Boye. 1983. Cell cycle parameters of slowly growing Escherichia coli B/r studied by flow cytometery. J Bacteriol 154(2):656-662.

Slatko, B.E., A.N. Luck, S.L. Dobson, and J.M. Foster. 2014. Wolbachia endosymbionts and human disease control. Molecular and Biochemical Parasitology. 195(2):88-95.

Slobodkina, G.B., A.L. Reysenbach, A.N. Panteleeva, N.A. Kostrikina, I.D. Wagner, E.A. Bonch-Osmolovskaya, and A.I. Slobodkin. 2012. Deferrisoma camini gen. nov., sp. nov., a moderately thermophilic, dissimilatory iron(III)-reducing bacterium from a deep-sea hydrothermal vent that forms a distinct phylogenetic branch in the Deltaproteobacteria. Int J Syst Evol Microbiol. 62 (Pt10):2463-2468.

Smith, D.J., J.D. Ravichandar, S. Jain, D.W. Griffin, H. Yu, Q. Tan, J. Thissen, T. Lusby, P. Nicoll, S. Shedler, P. Martinez, A. Osorio, J. Lechniak, S. choi, K. Sabino, K. Iverson, L. Chan, C. Jaing, and J. McGrath. 2018. Airborne bacteria in Earth's lower stratosphere resemble taxa detected in the troposphere: results from a new NASA aircraft bioaerosol collector (ABC). Frontiers in Microbiology. 9:1752. doi: 10.3389/fmicb.2018.01752.

Smith, J.L., B.J. Campbell, T.E. Hanson, C.L. Zhang, and S.C. Cary. 2008. Nautilia profundicola sp. nov., a thermophilic, sulfur-reducing epsilonproteobacterium from deep-sea hydrothermal vents. International Journal of Systematic and Evolutionary Microbiology. 58: 1598-1602.

Soo, R.M., J. Hemp, D.H. Parks, W.W. Fischer, and P. Hugenholtz. 2017. On the origins of oxygenic photosynthesis and aerobic respiration in Cyanobacteria. Science. 355(6332):1436-1440.

Soo, R.M., C.T. Skennerton, Y. Sekiguchi, M. Imelfort, S.J. Paech, P.G. Dennis, J.A. Steen, D.H. Parks, G.W. Tyson, and P. Hugenholtz. 2014. An expanded genomic representation of the phylum Cyanobacteria. Genome Biol. Evol. 6(5):1031-1045.

Soo, R.M., B.J. Woodcroft, D.H. Parks, G.W. Tyson, and P. Hugenholtz. 2015. Back from the dead; the curious tale of the predatory cyanobacterium Vampirovibrio chlorellavorus. PeerJ. 3:e968. doi:10.7717/peerj.968.

Sorokin, D.Y., T.P. Tourova, M.V. Sukhacheva, and G. Muyzer. 2012. Desulfuribacillus alkaliarsenatis gen. nov. sp. nov., a deep-lineage, obligately anaerobic, dissimilatory sulfur and arsenate-reducing, haloalkaliphilic representative of the order Bacillales from soda lakes. Extremophiles. 16(4):597-605.

Sorokin, D.Y., A.L. Rakitin, V.M. Gumerov, A.V. Beletsky, J.S. Sinninghe Damsté, A.V. Mardanov, and N.V. Ravin. 2016. Phenotypic and Genomic Properties of Chitinispirillum alkaliphilum gen. nov., sp. nov., A Haloalkaliphilic Anaerobic Chitinolytic Bacterium Representing a Novel Class in the Phylum Fibrobacteres. Front Microbiol. 7:407. doi: 10.3389/fmicb.2016.00407.

Spickler, A.R. 2017. Chlamydiosis. Retrieved from http://www.cfsph.iastate.edu/DiseaseInfo/factsheets.php.

Stauffer, B. and Spuhler, D. 2018. Anammox Factsheet. Sustainable Sanitation and Water Management Toolbox. 27 April. https://sswm.info/water-nutrient-cycle/wastewater-treatment/hardwares/semi-centralised-wastewater-treatments/anammox.

Stewart I., P.J. Schluter, and G.R. Shaw. 2006. Cyanobacterial lipopolysaccharides and human health - a review. Environmental Health: A Global Access Science Source. 5:7. http://www.ehjournal.net/content/5/1/7.

Suen, G., P.J. Weimer, D.M. Stevenson, F.O. Aylward, J. Boyum, J. Deneke, C. Drinkwater, N.N. Ivanova, N. Mikhailova, O. Chertkov, L.A. Goodwin, C.R. Currie, D. Mead, and P.J. Brumm. 2011. The Complete Genome Sequence of Fibrobacter succinogenes S85 Reveals a Cellulolytic and Metabolic Specialist. PLoS One. 6(4):e18814. doi:10.1371/journal.pone.0018814.

Sueoka N. 1962. On the genetic basis of variation and heterogeneity of DNA base composition. Proc. N.A.S. 48:582-592.

Sueoka N. 1988. Directional mutation pressure and neutral molecular evolution. Proc. Natl. Acad. Sci. USA. 85:2653-2657.

Sutcliffe, I.C. 2011. Cell envelope architecture in the Chloroflexi: a shifting frontline in a phylogenetic turf war. Environmental Microbiology. 13(2):279-282.

Sutcliffe I.C. and L.G. Dover. 2016. Comment on Tocheva et al. "Sporulation, bacterial cell envelopes and the origin of life." Nature Reviews Microbiology. 14(9):600. doi:10.1038/nrmicro.2016.113.

Suzuki H., M. Sota, C.J. Brown, and E.M. Top. 2008. Using Mahalanobis distance to compare genomic signatures between bacterial plasmids and chromosomes. Nucleic Acids Research. 36(22):e147. doi:10.1093/nar/gkn753.

Svetlichny, V.A., T.G. Sokolova, M. Gerhardt, M. Ringpfeil, N.A. Kostrikina, and G.A. Zavarzin. 1991. Carboxydothermus hydrogenoformans gen. nov., sp. nov., a CO-utilizing Thermophilic Anaerobic Bacterium from Hydrothermal Environments of Kunashir Island. Systematic and Applied Microbiology. 14(3):254-260.

Swingley, W.D., S. Sadekar, S.D. Mastrian, H.J. Matthies, J. Hao, H. Ramos, C.R. Acharya, A.L. Conrad, H.L. Taylor, L.C. Dejesa, M.K. Shah, M.E. O'Huallachain, M.T. Lince, R.E. Blankenship, J.T. Beatty, and J.W. Touchman. 2007. The Complete Genome Sequence of Roseobacter denitrificans Reveals a Mixotrophic Rather than Photosynthetic Metabolism. Journal of Bacteriology. 189(3):683-690.

Takeuchi, N., K. Kaneko, and E.V. Koonin.2014. Horizontal gene transfer can rescue prokaryotes fromMuller’s ratchet: benefit of DNA from dead cells and population subdivision. G3 (Bethesda) 4:325–339.

Takai, K., A. Inoue, and K. Horikoshi. 1999. Thermaerobacter marianensis gen. nov., sp. nov., an aerobic extremely thermophilic marine bacterium from the 11,000 m deep Mariana Trench. Int. J. Syst. Bacteriol. 49:619-628.

Tamarozzi, F., A. Halliday, K. Gentil, A. Hoerauf, E. Pearlman, and M.J. Taylor. 2011. Onchocerciasis: the Role of Wolbachia Bacterial Endosymbionts in Parasite Biology, Disease Pathogenesis, and Treatment. Clin. Microbiol. Rev. 24(3):459-468.

Tanabe, Y., Y. Hodoki, T. Sano, K. Tada, and M.W. Watanabe. 2018. Adaptation of the Freshwater Bloom-Forming Cyanobacterium Microcystis aeruginosa to Brackish Water Is Driven by Recent Horizontal Transfer of Sucrose Genes. Front. Microbiol. 05 June 2018. https://doi.org/10.3389/fmicb.2018.01150.

Tang, K.-H., K. Barry, O. Chertkov, E. Dalin, C.S. Han, L.J. Hauser, B.M. Honchak, L.E. Karbach, M.L. Land, A. Lapidus, F.W. Larimer, N. Mikhailova, S. Pitluck, B.K. Pierson, and R.E. Blankenship. 2011. Complete genome sequence of the filamentous anoxygenic phototrophic bacterium Chloroflexus aurantiacus. BMC Genomics. 12:334. doi: 10.1186/1471-2164-12-334.

Tang, K.-H., H. Yue, and R.E. Blankenship. 2010. Energy metabolism of Heliobacterium modesticaldum during phototrophic and chemotrophic growth. BMC Microbiology. 10:150. https://doi.org/10.1186/1471-2180-10-150.

Tocheva E.I., E.G. Matson, D.M. Morris, F. Moussavi, J.R. Leadbetter, and G.J. Jensen. 2011. Peptidoglycan remodeling and conversion of an inner membrane into an outer membrane during sporulation. Cell. 146(5):799-812. doi:10.1016/j.cell.2011.07.029.

Tocheva E.I., D.R. Ortega, and G.J. Jensen. 2016. Sporulation, bacterial cell envelopes and the origin of life. Nat Rev Microbiol. 14(8):535-542. doi:10.1038/nrmicro.2016.85.

Toh, E., H.D. Kurtz Jr., Y.V. Brun. Characterization of the Caulobacter crescentus holdfast polysaccharide biosynthesis pathway reveals significant redundancy in the initiating glycosyltransferase and polymerase steps. 2008. Journal of Bacteriology. 190(21):7219-7231

Trojanowski, D., J. Hołówka, K. Ginda, D. Jakimowicz, and J. Zakrzewska-Czerwinska. 2017. Multifork chromosome replication in slow-growing bacteria. Scientific Reports. 7:43836. doi: 10.1038/srep43836.

Trojanowski, D., J. Hołówka, and J. Zakrzewska-Czerwinska. 2018. Where and when bacterial chromosome replication starts: a single cell perspective. Front. Microbiol. 9:2819. doi: 10.3389/fmicb.2018.02819. https://www.frontiersin.org/articles/10.3389/fmicb.2018.02819/full

Trun, N.J. and J.F. Marko. 1998. Architecture of a bacterial chromosome. American Society for Microbiology News. 64(5):276-283.

Utami, Y.D., H. Kuwahara, T. Murakami, T. Morikawa, K. Sugaya, K. Kihara, M. Yuki, N. Lo, P. Deevong, S. Hasin, W. Boonriam, T. Inoue, A. Yamada, M. Ohkuma, and Y. Hongoh. 2018. Phylogenetic Diversity and Single-Cell Genome Analysis of "Melainabacteria", a Non-Photosynthetic Cyanobacterial Group, in the Termite Gut. Microbes Environ. 33(1):50-57.

Val M.E., S.P. Kennedy, A.J. Soler-Bistué, V. Barbe, C. Bouchier, M. Ducos-Galand, O. Skovgaard, and D. Mazel. 2014. Fuse or die: how to survive the loss of Dam in Vibrio cholerae. Molecular Microbiology. 91(4):665-678. doi: 10.1111/mmi.12483.

van Elsas, J.D., A.V. Semenov, R. Costa, and J.T. Trevors. 2011. Survival of Escherichia coli in the environment: fundamental and public health aspects. ISME J. 5(2):173-183.

van Niftrik, L. and M.S.M. Jetten. 2012. Anaerobic ammonium-oxidizing bacteria: unique microorganisms with exceptional properties. Microbiology and Molecular Biology Reviews. 76(3):585-596.

van Teeseling, M.C.F., R.J. Mesman, E. Kuru, A. Espaillat, F. Cava, Y.V. Brun, M.S. VanNieuwenhze, B. Kartal, and L. van Niftrik. 2015. Anammox Planctomycetes have a peptidoglycan cell wall. Nature Communications. 6:6878. doi: 10.1038/ncomms7878.

Vanaja S.K., A.J. Russo, B. Behl, I. Banerjee, M. Yankova, S.D. Deshmukh, and V.A. Rathinam. 2016. Bacterial outer membrane vesicles mediate cytosolic localization of LPS and caspase-11 activation. Cell 165:1106–1119. doi:10.1016/j.cell.2016.04.015.

Varghese, N.J., S. Mukherjee, N.Ivanova, K.T. Konstantinidis, K. Mavrommatis, N.C. Kyrpides, A. Pati. 2015. Microbial species delineation using whole genome sequences, Nucleic Acids Research, 43(14):6761–6771, https://doi.org/10.1093/nar/gkv657

Volff J.-N. and J. Altenbuchner. 2000. A new beginning with new ends: linearisation of circular chromosomes during bacterial evolution. FEMS Microbiology Letters. 186:143-150.

Volland J.-E., S. Gonzalex-Rizzo, O. Gros, et al. 2022. A centimeter-long bacterium with DNA contained in metabolically active, membrane-bound organelles. Science. 376:1453-1458.

von Morgen P. and J. Maciejowski. 2018. The ins and outs of telomere crisis in cancer. Genome Med. 10:89. doi:10.1186/s13073-018-0596-4.

Voordeckers, J.W., V. Starovoytov, and C. Vetriani. 2005. Caminibacter mediatlanticus sp. nov., a thermophilic, chemolithoautotrophic, nitrate-ammonifying bacterium isolated from a deep-sea hydrothermal vent on the Mid-Atlantic Ridge. Int J Syst Evol Microbiol. 55(Pt 2):773-779.

Walsby, A.E. 1994. Gas vesicles. Microbiol Rev. 58(1):94-144.

Wang J.D. and P.A. Levin. 2009. Metabolism, cell growth and the bacterial cell cycle. Nat Rev Microbiol. 7(11):822-827. doi:10.1038/nrmicro2202.

Watanabe, M., H. Kojima, and M. Fukui. 2015. Limnochorda pilosa gen. nov., sp. nov., a moderately thermophilic, facultatively anaerobic, pleomorphic bacterium and proposal of Limnochordaceae fam. nov., Limnochordales ord. nov. and Limnochordia classis nov. in the phylum Firmicutes. International Journal of Systematic and Evolutionary Microbiology. 65:2378-2384. doi: 10.1099/ijs.0.0000267.

Ward, L.M., J. Hemp, P.M. Shih, S.E. McGlynn, and W.W. Fischer. 2018. Evolution of phototrophy in the Chloroflexi phylum driven by horizontal gene transfer. Frontiers in Microbiology. Vol. 9, Article 260. doi: 10.3389/fmicb.2018.00260.

Ward, N.L., J.F. Challacombe, P.H. Janssen, B. Henrissat, P.M. Coutinho, M. Wu, G. Xi, D.H. Haft, M. Sait, J. Badger, R.D. Barabote, B. Bradley, T.S. Brettin, L.M. Brinkac, D. Bruce, T. Creasy, S.C. Daugherty, T.M. Davidsen, R.T. Deboy, J.C. Detter, R.C. Dodson, A.S. Durkin, A. Ganapathy, M. Gwin-Giglio, C.S. Han, H. Khouri, H. Kiss, S.P. Kothari, R. Madupu, K.E. Nelson, W.C. Nelson, I. Paulsen, K. Penn, Q. Ren, M.J. Rosovitz, J.D. Selengut, S. Shrivastava, S.A. Sullivan, R. Tapia, L.S. Thompson, K.L. Watkins, Q. Yang, C. Yu, N. Zafar, L. Zhou, and C.R. Kuske. 2009. Three Genomes from the phylum Acidobacteria provide insight into the lifestyles of these microorganisms in soils. Appl. Environ Microbiol. 75(7):2046-2056.

Wassenaar T.M. and K. Zimmermann. 2018. Lipopolysaccharides in food, food supplements, and probiotics: should we be worried? European Journal of Microbiology and Immunology. 8(3):63-69. doi: 10.1556/1886.2018.00017.

Wayne, L.G., D.J. Brenner, R.R. Colwell, P.A.D. Grimont, O. Kandler, M.I. Krichevsky, L.H. Moore, W.E.C. Moore, R.G.E. Murray, E. Stackebrandt, M.P. Starr, and H.G. Trüper. 1987. Report of the ad hoc committee on reconciliation of approaches to bacterial systematics. Int. J. Syst. Bacteriol. 37(4):463–464.

Wexler, H.M. 2007. Bacteroides: the good, the bad, and the nitty-gritty. Clin Microbiol Rev. 20(4):593-621.

Winter, J., E. Braun and H.P. Zabel. 1987. Acetomicrobium faecalis spec, nov. a strictly anaerobic bacterium from sewage sludge, producing ethanol from pentoses. Syst Appl Microbiol 9:71–76. [note: species has since been renamed/reclassified as Caldicoprobacter faecalis]

Wolgemuth, C.W. 2015. Flagellar motility of the pathogenic spirochetes. Semin. Cell Dev. Biol. 46:104-112.

World Health Organization. 2018. Campylobacter fact sheet. https://www.who.int/news-room/fact-sheets/detail/campylobacter. 23 January 2018.

Wu, M., Q. Ren, A.S. Durkin, S.C. Daugherty, L.M. Brinkac, R.J. Dodson, R. Madupu, S.A. Sullivan, J.F. Kolonay, W.C. Nelson, L.J. Tallon, K.M. Jones, L.E. Ulrich, J.M. Gonzalez, I.B. Zhulin, F.T. Robb, and J.A. Eisen. 2005. Life in Hot Carbon Monoxide: The Complete Genome Sequence of Carboxydothermus hydrogenoformans Z-2901. PLoS Genet. 1(5);e65. doi:10.1371/journal.pgen.0010065.

Yan, L., S. Zhang, P. Chen, H. Liu, H. Yin, and H. Li. 2012. Magnetotactic bacteria, magnetosomes and their application. Microbiological Research. 167(9):507-519.

Yang, L., Y. Ding, Y. Chen, S. Zhang, C. Huo, Y. Wang, J. Yu, P. Zhang, H. Na, H. Zhang, Y. Ma, and P. Liu. 2012. The proteomics of lipid droplets: structure, dynamics, and functions of the organelle conserved from bacteria to humans. J Lipid Res. 53(7):1245-1253. doi:10.1194/jlr.R024117.

Yang C.C., C.H. Huang, C.Y. Li, Y.G. Tsay, S.C. Lee, and C.W. Chen. 2002. The terminal proteins of linear Streptomyces chromosomes and plasmids: a novel class of replication priming proteins. Mol Microbiol. 43(2):297-305.

Yonezawa H., T. Osaki, S. Kurata, M. Fukuda, H. Kawakami, K. Ochiai, T. Hanawa, and S. Kamiya. 2009. Outer membrane vesicles of Helicobacter pylori TK1402 are involved in biofilm formation. BMC Microbiology. 9:197. https://doi.org/10.1186/1471-2180-9-197.

Yoshikawa H., A. O'Sullivan, and N. Sueoka. 1964. Sequential replication of the Bacillus subtilis chromosome, III. Regulation of initiation. Proc Natl Acad Sci USA. 52:973-980.

Young, K.D. 2010. Bacterial shape: two-dimensional questions and possibilities. Annu Rev Microbiol. 64:223-240. doi:10.1146/annurev.micro.112408.134102.

Yurkov, V.V., and J.T. Beatty. 1998. Aerobic anoxygenic phototrophic bacteria. Microbiol. Mol. Biol. Rev. 62:695-724.

Yutin, N. P. Puigbò, E.V. Koonin, and Y.I. Wolf. 2012. Phylogenomics of prokaryotic ribosomal proteins. PLoS One. 7(5):e36972. https://doi.org/10.1371/journal.pone.0036972.

Zaburannyi, N., B. Bunk, J. Maier, J. Overmann, and R. Müller. 2016. Genome Analysis of the Fruiting Body-Forming Myxobacterium Chondromyces crocatus Reveals High Potential for Natural Product Biosynthesis. Applied and Environmental Microbiology. 82(6):1945-1957.

Zhang, C. and P. Liu. 2017. The lipid droplet: a conserved cellular organelle. Protein Cell. 8(11):796-800. doi:10.1007/s13238-017-0467-6.

Zheng, H. and A. Brune. 2015. Complete Genome Sequence of Endomicrobium proavitum, a Free-Living Relative of the Intracellular Symbionts of Termite Gut Flagellates (Phylum Elusimicrobia). Genome Announc. 3(4):e00679-15. doi:10.1128/genomeA.00679-15.

Zeng, Y., V. Selyanin, M. Lukeš, J. Dean, D. Kaftan, F. Feng, and M. Koblížek. 2015. Characterization of the microaerophilic, bacteriochlorophyll a-containing bacterium Gemmatimonas phototrophica sp. nov., and emended descriptions of the genus Gemmatimonas and Gemmatimonas aurantiaca. Int. J. Syst. Evol. Microbiol. 65(8):2410-2419. doi: 10.1099/ijs.0.000272.

Zeng, Y., F. Feng, H. Medová, J. Dean, and M. Koblížek. 2014. Functional type 2 photosynthetic reaction centers found in the rare bacterial phylum Gemmatimonadetes. PNAS. 111(21):7795-7800. https://doi.org/10.1073/pnas.1400295111.

Last edited: 24 Jan. 2019

How to cite:

Note: this is an incomplete draft. I will be adding additional sections at a later date.

Hegarty, S. 2019. Bacteria. January 24, 2019 DRAFT. Mass Nature, https://massnature.com/bacteria.