Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Appl Environ Microbiol. 2006 Mar; 72(3): 1719–1728.
PMCID: PMC1393246
PMID: 16517615

Identifying the Dominant Soil Bacterial Taxa in Libraries of 16S rRNA and 16S rRNA Genes

From near to far,

from here to there,

funny things are everywhere.

—Dr. Seuss, One Fish Two Fish Red Fish Blue Fish

HISTORICAL ASPECTS

In 1909, H. Joel Conn (25) expressed the hope that methods would soon be at hand by which the significance of the different bacteria present in any soil could be determined. However, by 1918 he was pointing out that the methods available to him, which relied on cultivation of bacteria on artificial media, resulted in the formation of colonies by only 1.5 to 10% of the bacterial cells in soil (26). Fifty years later, Vagn Jensen concluded a review of cultivation-based methodologies by stating his suspicion that those cells that were forming colonies were unrepresentative of the total bacterial community (59). This was confirmed when cultivation-independent methods began to be used to study soil bacteria (see below). Even so, in the absence of better methods, the pure cultures derived from the colonies that did form were extensively and successfully studied throughout the 20th century. Much of our basic knowledge of soil bacteria, as well as the discovery of many important antibiotics, came from investigations of pure cultures (2, 86, 104). Cultured isolates are still very important in developing our understanding of bacterial physiology, genetics, and ecology (85, 122).

Beginning in the 1990s, the application of molecular ecological methods, especially those based on surveys of genes after PCR amplification, has allowed cultivation-independent investigations of the microbial communities of soils to be made. The power of these methods has largely rendered obsolete the plate count approach to detecting and enumerating subsets of soil bacteria, and a range of diagnostic and quantitative methods that target functional genes, phylogenetically informative genes, or RNAs has been developed (49, 69). In particular, 16S rRNA and its gene have proven to be useful and powerful markers for the presence of bacteria in samples (36, 56, 88). The utility of these markers is facilitated by the availability of primers that allow amplification of almost the complete gene or its RNA product (66) and by the phylogenetic inferences that can be made from the resultant nucleotide sequences, permitting placement of the host organism within a phylogenetic framework even if closely related cultured organisms are lacking (36, 56, 74). Since the initial pioneering studies to survey soil bacterial communities using molecular ecological surveys (13, 68, 79, 91, 111), a number of libraries of 16S rRNA and 16S rRNA genes derived from soils have become available (Table (Table11).

TABLE 1.

Libraries of 16S rRNA or 16S rRNA genes prepared from RNA and DNA extracted from soilsa

LibrarySourcebNo. of clonesNo. of sequencesReference(s)
NMS8Canada, forest, DNA91675
NMW3Canada, forest, DNA90615
NOS7Canada, forest, DNA92735
NOW2Canada, forest, DNA93765
SMS9Canada, forest, DNA94835
SMW4Canada, forest, DNA92565
FVMBrazil, forest, DNA484812
FVPBrazil, pasture, DNA494912
C0United States, arid woodland, DNA16216232, 33, 65
C1United States, arid woodland, DNA10910932, 33, 65
S0United States, arid woodland, DNA16316332, 33, 65
S1United States, arid woodland, DNA10210232, 33, 65
a13United Kingdom, DNA*393835
HBUnited States, cropping rotation, DNA959545
TcSwitzerland, RNA + DNA*14114150
KolmAustria, forest, DNA454551
RothAustria, forest, DNA515151
STAustria, forest, DNA414151
SturtAustralia, arid landscape, DNA12212254
WMARS00United States, DNA*25825871
WMARS97United States, DNA*13813871
NiwotUnited States, alpine meadow, DNA18015972
LBSSwitzerland, pasture, DNA272776
TBSSwitzerland, pasture, DNA262676
safUnited Kingdom, pasture, DNA13713777
s1United Kingdom, pasture, DNA13713777
WittGermany, moorland, RNA + DNA39639682, 83
C6Germany, forest, DNA313192
MTPUnited States, wheat, DNA9393105
ABSUnited States, grassland, DNA5959121
DSUnited States, grassland, DNA6262121
EBAustralia, pasture, DNA135135c
Totals3,3983,240
aAll libraries were used to ascertain the genus and family affiliations of soil bacteria, but only libraries with ≥90 clones were used to determine phylum and class level abundances.
b*, details of vegetation not given.

It is important to realize that libraries of PCR-amplified 16S rRNA and 16S rRNA genes may not represent a complete or accurate picture of the bacterial community. Firstly, the species diversity is so great (28, 46, 109) that libraries of <400 cloned sequences must represent only an incomplete sampling. Even all of the currently published sequences combined would seem to constitute an incomplete census of all of the 16S rRNA genes on earth (98). In addition, there may be biases in the contributions of the various bacterial groups to libraries. The efficiencies of nucleic acid extraction may be different for different bacteria, the number of copies of 16S rRNA or 16S rRNA genes per cell varies, and there may be preferential amplification of some sequence types relative to others by PCR (36, 43, 113). Some sequences may arise from contaminating DNA and may not represent bacteria actually present in the sample being studied (108). Assigning physiologies and functions to the hosts of 16S rRNA gene sequences is complicated in many cases by the lack of characterized close relatives (e.g., see references 31, 57, 81, and 88) and by the diversity of phenotypes among close relatives in some groups (1, 95). Some, but not all, of these biases may be overcome as metagenomic data sets accumulate (71, 87, 110). In the meantime, the available libraries of 16S rRNA and 16S rRNA genes permit an initial survey of the global soil bacterial community structure.

SEQUENCE DATA SETS

Thirty-two libraries of 16S rRNA and 16S rRNA genes of members of the domain Bacteria, prepared from a variety of soils, were analyzed to gain an understanding of the general composition of soil bacterial communities (Table (Table1).1). Libraries or sequences from rhizosphere samples were not included in this synthesis. Libraries consisting predominantly of sequences of <300 nucleotides were also excluded, as phylogenetic assignment from very short sequences can be unreliable (36, 74), and sequences of <300 nucleotides were removed from the libraries that were included. Some published libraries were generated with primers that could not be expected to sample most known bacteria or were screened in such a way that the total number of clones belonging to each group could not be deduced from the published data. These were not included in this survey. Two potentially interesting libraries containing unexpectedly large numbers of sequences assignable to the genus Escherichia were excluded because these may not have originated from DNA from the soil being investigated (30). As a consequence of the exclusions, the final data set is not as geographically comprehensive as it might have been, but a number of different vegetation types are included (Table (Table1).1). Some libraries consist of data from multiple reports in which the sample site or the sample itself appeared to be the same. In a few cases, multiple libraries that originated from highly similar replicate samples were pooled to increase the library size. A total of 3,240 sequences from the 32 libraries were assigned to genus level groupings by the “Classifier” program of Ribosomal Database Project II (24) and then weighted for multiple clone assignments to one sequence type, and this pooled data set of 3,398 clones was treated as one global set.

The contribution of phylum level groupings to soil bacterial communities was calculated only from the 21 libraries with ≥90 clones. Smaller libraries (<90 clones) contained representatives of very few phyla. It was felt that the smaller libraries might skew the outcomes, since the contributions were normalized to compensate for library size and then analyzed further. For some of the better-characterized dominant phyla (Acidobacteria, Actinobacteria, Bacteroidetes, Firmicutes, Proteobacteria, and Verrucomicrobia), the clones were also assigned to subphylum groups (class, subclass, or subdivision). Some of these subphylum groups are organizational rather than evolutionarily equivalent lineages, especially the classes of the phylum Proteobacteria (74), but are useful for the purposes of surveying the global data set. These assignments were based on results obtained with Classifier (17), published phylogenetic trees and tables in publications or their supporting material, BLAST (3) in GenBank databases (8), Ribosomal Database Project II databases (24), and phylogenetic analyses of sequences against references of known affiliation (e.g., see references 60, 93, and 96).

HISTORICALLY IMPORTANT SOIL BACTERIA

In his landmark book, the second edition of Introduction to Soil Microbiology, Martin Alexander (2) listed what were then considered to be important genera of soil bacteria, based on cultivation studies. He suggested that members of nine genera were significant in soils: Agrobacterium, Alcaligenes, Arthrobacter, Bacillus, Flavobacterium, Micromonospora, Nocardia, Pseudomonas, and Streptomyces (Table (Table2).2). Since 1977, two things that have changed the validity of this list have occurred. First, many of the genera have undergone taxonomic revision, and some of their species have been reclassified into new or other genera. This is especially true for the genera Flavobacterium and Pseudomonas (4, 9). Second, surveying of 16S rRNA genes in soils has permitted a more direct census of soil bacteria, without the limitations inherent in cultivation-based studies. These surveys suggest that members of Alexander's nine genera, as they are currently defined, together make up only 2.5 to 3.2% of soil bacteria (Table (Table2).2). Of the nine genera, Pseudomonas spp. are the most abundant in soil bacterial communities, contributing 1.6% of the cloned sequences from soils (Table (Table22).

TABLE 2.

Abundance of members of well-known genera of soil bacteria in libraries of 16S rRNA and 16S rRNA genes compared with their historical significance as colony-forming soil bacteria and their contribution to soil isolates held in the ATCC

GenusContribution to libraries (%) at different confidence levelsa
Range of abundance among colonies (%)bContribution to soil isolates in ATCC (%)c
100%≥80%
Actinomadura00d1.5
Actinoplanes00.061.5
Agrobacterium000-13
Alcaligenes001-8
Arthrobacter0.330.533-401.3
Bacillus0.180.625-457.6
Clostridium0.030.091.6
Flavobacterium0.350.381-7
Flexibacter001.2
Hyphomicrobium0.030.031.2
Micromonospora000-52.1
Mycobacterium0.330.502.6
Nocardia003-10
Paenibacillus0.120.181.4
Pseudomonas1.601.602-106.0
Ralstonia001.0
Rhodococcus001.4
Streptomyces0.060.0623-3025.2
aConfidence levels calculated by the program Classifier of the Ribosomal Database Project (24).
bCalculated from the data of Alexander (2), assuming a mean of one-third of colonies being actinomycetes (filamentous members of the subclass Actinobacteridae, including Micromonospora, Nocardia, and Streptomycetes) and the remaining two-thirds being other non-actinomycete bacteria (2).
cData from Floyd et al. (41).
d—, no data.

Recently, Floyd et al. (41) presented a breakdown of cultures of prokaryotic organisms in the American Type Culture Collection (ATCC). The 14 genera of soil bacteria with the most deposited cultures, together encompassing over half of all soil-derived isolates in the ATCC, are also not common among the clones detected in libraries (Table (Table2).2). Together, members of these genera make up only 2.7 to 3.7% of soil bacteria. This shows that the cultured part of soil bacterial diversity is not representative of the total diversity, as suggested by Jensen (59) and many others since.

GENUS LEVEL DIVERSITY

The Classifier algorithm (24) returns a confidence value with which a 16S rRNA gene sequence can be assigned to a taxon (genus and higher) that is represented by a set of sequences, based on the number of times, out of 100 trials, that random subsets of the query sequence match sequences assigned to that taxon. The algorithm also returns the name of the taxon to which the sequence was most often assigned in those 100 trials. Using Classifier, as many as 17% of the sequences could be assigned to a known genus with 100% confidence; this increased to 32% with a confidence level of 80% or greater. However, these outcomes were greatly influenced by sequences falling into poorly defined groups with very few described species. Sequences affiliated with these classes or phyla tended to be identified as members of the one genus or few genera in them, because there were no other genera to draw the sequences during bootstrap analysis. Nearly all of these assignments were spurious, as the sequence similarities to the few named species were <96%. Everett et al. suggest that 16S rRNA gene sequence similarities of <96% are indicative of the hosts of the genes belonging to different genera (37). When the genera Acidobacterium (phylum Acidobacteria), Alterococcus, Verrucomicrobium, Xiphinematobacter (phylum Verrucomicrobia), Gemmatimonas (phylum Gemmatimonadetes), and Conexibacter and Rubrobacter (subclass Rubrobacteridae of the phylum Actinobacteria) were removed, the number of sequences able to be assigned to known genera decreased. In their absence, only 11% of sequences could be assigned to known genera with 100% confidence and 21% at ≥80% confidence. These figures are still perhaps surprisingly high, given that soil bacterial diversity is high and our ability to culture these bacteria is generally considered to be poor (28, 46, 88, 98, 109, 122). It seems that our ability to culture representatives of the phylogenetic diversity of soil bacteria, at least as judged at the genus level, is better than the 1% often quoted, even if culturability as a function of cell numbers is low.

The majority (79 to 89%) of 16S rRNA gene sequences are from bacteria that are not affiliated with known genera. Some of these are associated with well-studied lineages of bacteria, such as Actinobacteridae, Flavobacteria, Sphingobacteria, Bacilli, Clostridia, Alphaproteobacteria, Betaproteobacteria, Gammaproteobacteria, and Deltaproteobacteria. In some of these groups, the number of bacteria affiliated with known genera is high. For example, up to 60% of sequences from members of the class Sphingobacteria and up to 76% of sequences from members of Flavobacteria may come from described and named genera (Table (Table3).3). Among sequences assigned to Actinobacteridae this is lower, with less than half belonging to described genera, while in the Deltaproteobacteria, the number of sequences assignable to known genera is even lower (Table (Table3).3). Since members of the latter two groups have been the source of many chemically novel bioactive compounds (104), this indicates considerable scope for more discovery. In some groups well represented by cultured isolates, such as in the classes Alphaproteobacteria, Betaproteobacteria, and Gammaproteobacteria, less than half of all sequences could be assigned to these described and named genera (Table (Table3).3). Overall, only 20 to 40% of sequences affiliated with well-characterized groups of bacteria can be assigned to known genera. Of the genera in the well-defined groups, the three most abundant at 100% assignment confidence are Burkholderia (class Betaproteobacteria), Pseudomonas (class Gammaproteobacteria), and Chitinophaga (class Sphingobacteria), which constitute 2.7, 1.6, and 1.0% of all the sequences, respectively.

TABLE 3.

Assignment of cloned 16S rRNA and 16S rRNA genes affiliated with well-characterized subphylum groupings (class or subclass) to described genera

SubphylumProportion of all 3,398 clones (%)Proportion of clones assignable to genera (%) at different confidence levelsa
Total no. of described genera in groupb
100%≥80%
Actinobacteridae4.251.12 (26)2.01 (47)158
Flavobacteria0.500.35 (70)0.38 (76)25
Sphingobacteria4.341.15 (26)2.60 (60)28
Bacilli1.890.38 (20)1.03 (54)79
Clostridia0.590.06 (10)0.27 (46)135
Alphaproteobacteria18.552.10 (11)6.38 (34)160
Betaproteobacteria10.873.31 (30)4.82 (44)93
Deltaproteobacteria2.840.03 (1)0.59 (21)70
Gammaproteobacteria7.772.10 (27)2.60 (33)194
Total51.6010.6020.68942
aConfidence levels calculated by the program Classifier of the Ribosomal Database Project (24). The percentage of sequences affiliated with each class that were able to be assigned to a genus is given in parentheses.
bData from Garrity et al. (48).

Although recognized as the essential basis of bacterial systematics (115), the genus rank has not been well defined. In essence, though, genera tend to consist of species that share major phenotypic characteristics that differentiate them from species of related genera. The consequence of being able to assign only 10 to 21% of sequences to known genera is that the broad characteristics of most of the soil bacteria are not known.

ABUNDANCE OF DIFFERENT PHYLA

16S rRNA genes from soil bacteria are affiliated with at least 32 phylum-level groups. The contributions that members of different phyla make to the different soil bacterial communities vary (Fig. (Fig.1).1). The dominant phyla in the libraries are Proteobacteria, Acidobacteria, Actinobacteria, Verrucomicrobia, Bacteroidetes, Chloroflexi, Planctomycetes, Gemmatimonadetes, and Firmicutes (Table (Table4).4). Members of these nine phyla make up an average of 92% of soil libraries (normalized for the size of the individual libraries). Although there are at least 52 bacterial phyla (88), and 24 are recognized by Bergey's Manual (48), soils seem to be dominated by only a small number of these.

An external file that holds a picture, illustration, etc.
Object name is zam0030665400001.jpg

Contributions of 16S rRNA and 16S rRNA genes from members of different phyla in libraries prepared from soil bacterial communities (2,920 clones in 21 libraries). The horizontal line in the middle of each block indicates the mean, the block represents 1 standard deviation on either side of the mean, and the vertical lines extending above and below each block indicate the minimum and maximum contributions of each phylum.

TABLE 4.

Contribution of 16S rRNA and 16S rRNA genes from members of different phyla and subphylum groups (class, subdivision, or subclass) to soil bacterial communitiesa

PhylumSubphylumMean contribution (%)Range (%)
AcidobacteriaSubdivision 13.30-14.4
Subdivision 20.50-3.3
Subdivision 31.80-4.9
Subdivision 47.70-35.0
Subdivision 50.40-2.2
Subdivision 64.50-12.8
Subdivision 71.50-7.4
ActinobacteriaAcidimicrobidae2.40-8.9
Actinobacteridae4.70-18.3
Rubrobacteridae5.60-24.8
ProteobacteriaAlphaproteobacteria18.81.8-43.1
Betaproteobacteria10.02.1-31.1
Gammaproteobacteria8.11.1-34.1
Deltaproteobacteria2.30-10.1
Epsilonproteobacteria0.040-0.8
VerrucomicrobiaVerrucomicrobiae0.030-0.7
Spartobacteria6.30-21.1
Subdivision 30.50-4.7
Subdivision 40.20-1.1
BacteroidetesFlavobacteria0.40-3.2
Sphingobacteria4.60-15.9
FirmicutesBacilli1.60-7.0
Clostridia0.20-1.4
Chloroflexi3.20-15.8
Planctomycetes2.00-7.8
Gemmatimonadetes2.00-3.7
Other groups5.22.2-9.9
Unknown2.40-12.6
aOnly the 21 libraries with ≥90 clones were included in this survey (total, 2,920 clones).

Members of the phyla Proteobacteria and Acidobacteria are the most abundant soil bacteria, as judged by the occurrence of 16S rRNA and 16S rRNA genes that are assignable to these groups (Table (Table4).4). All of the libraries surveyed contained some sequences assignable to these two phyla. The other dominant phyla are not found in all libraries, but this is likely to be a consequence of library size. Given that there is variation in the contribution of different phyla and classes, abundances at the lower end of any range may mean that no sequences are detected if the library is too small. In addition to representatives of the nine major phyla, members of a number of other phylum-level lineages, such as Chlamydiae, Chlorobi, Cyanobacteria, Deinococcus-Thermus, Fibrobacteres, Nitrospirae, BRC1, NKB19, OP10, OP11, OS-K, SC3, SC4, termite group I, TM6, TM7, WS2, and WS3, are present in the global data set. Some members of a few of these phyla are quite well studied, but in general very little is known about the soil-inhabiting members of these groups. Most of these phyla are virtually unstudied and have few or no known pure culture representatives from soils (74).

A number of studies using methods other than analysis of libraries have estimated the contribution of members of different bacterial groups to the microbial population of soils (Table (Table5).5). These studies support the general trends observed in clone libraries (Fig. (Fig.1,1, Table Table4)4) that Alphaproteobacteria, Acidobacteria, and Actinobacteria are often abundant in soils and that members of Bacteroidetes, Firmicutes, and Planctomycetes are generally less abundant (Table (Table5).5). They also support the observation that the estimated abundance of the major phyla varies between different soils (or samples). It is not possible to state to what degree the variations are method based. Fluorescence in situ hybridization (FISH) and other hybridization methods may detect bacteria other than the intended target group, or the phylogenetic coverage of oligonucleotide probes may not be comprehensive. The same applies to oligonucleotides designed for quantitative PCR approaches. Detection of extracted rRNA is affected by ribosome levels in bacteria, while clone library compositions are influenced by PCR steps and by rrn copy number. The results obtained with all the methods are affected by the physical nature of bacterial cells, which may vary between groups and under different conditions, affecting oligonucleotide probe permeability and successful nucleic acid extraction.

TABLE 5.

Estimates of abundance of the members of different bacterial groups made by cultivation-independent methods other than clone library analysis

Source (site designation)aMethodbNo. of members of indicated groupc
Reference(s)
ACIACTALFBETGAMDELVERBACFIRPLA
Cropland, ItalyFISH42015586518
Organic soil, NorwayFISH<12-5<1-5<15-8<13-720
Mineral soil, GermanyFISH<17-10<1-4<14-7<13-720
Tundra, RussiaFISH411238664
Cropland, GermanyFISH310252102
Forest, GermanyFISH<17<113<17118
Tilled cropland, United States (CT)rRNA<1-95-2218-412-6<1-3<15-715, 16
Tilled cropland, United States (NI)rRNA7264315
No-till cropland, United States (AF)rRNA7264315
No-till cropland, United States (NT)rRNA11315315
Abandoned field, United States (HCS)rRNA1-39-277-331-51-3<13-915, 16
Abandoned field, United States (LS)rRNA4103812<11316
Tilled grassland, United States (HCST)rRNA2144<1<1<1216
Meadow, United States (NCS)rRNA<1-39-1812-411-941-31-24-1215, 16
Tree plantation, United States (PL)rRNA17161<1<1316
Meadow, The NetherlandsrRNA19224839
Desert, United StatesqPCR195741440
Forest, United StatesqPCR1451451340
Prairie, United StatesqPCR236981640
aThe site designations are those used by authors to identify particular sources within studies with multiple soil samples.
bFISH, counting of cells in soil samples with group-specific oligonucleotide probes; rRNA, estimation of abundance of rRNA in total rRNA by hybridization with group-specific oligonucleotide probes; qPCR, quantitative PCR estimate of 16S rRNA genes using group-specific assays, relative to estimates of total bacteria using Bacteria-specific assays.
cACI, phylum Acidobacteria; ACT, phylum Actinobacteria; ALF, class Alphaproteobacteria; BET, class Betaproteobacteria; GAM, class Gammaproteobacteria; DEL, class Deltaproteobacteria; VER, phylum Verrucomicrobia; BAC, phylum Bacteroidetes; FIR, phylum Firmicutes; PLA, phylum Planctomycetes.

PHYLUM LEVEL DIVERSITY

Members of the phylum Proteobacteria make up an average of 39% (range, 10 to 77%) of libraries derived from soil bacterial communities (Fig. (Fig.1).1). Most soil-dwelling members of the phylum Proteobacteria can be classified within the classes Alphaproteobacteria, Betaproteobacteria, Gammaproteobacteria, and Deltaproteobacteria (Table (Table4).4). The phylum Proteobacteria currently contains some 528 named and described genera (48), but the number of proteobacterial sequences that can be confidently assigned to known genera is relatively low. Depending on the level of confidence used, only 19 to 36% of the proteobacterial sequences can be assigned to a known genus (Table (Table3),3), indicating that many proteobacterial groups still remain to be described and named. Libraries from soils reveal the existence of lineages not affiliated with known isolates (e.g., see references 5, 37, and 77). Given the extent of physiological diversity within the phylum, no guesses as to their general metabolism can be made. At present, research interest seems to be directed toward the as-yet-uncultured groups within phyla such as Acidobacteria and Verrucomicrobia, which constitute smaller parts of the soil bacterial community; the uncultured proteobacteria have been largely ignored. Members of some cosmopolitan family or order level groups of the Alphaproteobacteria and Gammaproteobacteria without named and described genera have been isolated but are not yet formally named (29, 60, 94). 16S rRNA genes affiliated with these isolates have been found in most of the libraries listed in Table Table11.

Members of the phylum Acidobacteria make up an average of 20% (range, 5 to 46%) of soil bacterial communities (Fig. (Fig.1).1). The phylum Acidobacteria is divided into at least eight subdivisions that may have class level rank (56). Three of these are particularly abundant in soils: subdivisions 1, 4, and 6 (Table (Table4).4). There are, however, only three formally described genera in the phylum (48). Members of subdivision 1 have been shown to be readily culturable. Kishimoto and Tano (62) originally cultured eight isolates, one of which was named Acidobacterium capsulatum (63). Since then, at least 99 further isolates that span the phylogenetic breadth of the subdivision have been reported (29, 52, 58, 60, 93, 94, 103). A few isolates of members of subdivisions 2, 3, and 4 have been obtained in pure culture (29, 60, 78, 92, 94). All isolates appear to be aerobic heterotrophs, but this may be a result of the cultivation strategies employed to date. The phylogenetic breadth of subdivisions 3 and 4 in particular is much greater than that covered by the few isolates (e.g., see references 5, 7, 50, 56, 71, 87, and 121). To date, no isolates of subdivisions 6 and 7 have been reported, even though members of these subdivisions are common in soils (Table (Table4).4). Members of subdivision 6 may be aerobes, since they were not detected in a permanently anoxic soil system (116) and colonies have recently been found to form on plates incubated under air (K. E. R. Davis and P. H. Janssen, unpublished data). The phylogenetic depth of the phylum Acidobacteria is nearly as great as in the phylum Proteobacteria, and the different class rank subdivisions may contain bacteria with very different physiologies (31, 56, 75). For example, the only known members of subdivision 8 are obligate anaerobes, with two different basic physiologies, which contrast with the known aerobes of subdivisions 1 to 4 (23, 63, 70, 93).

Members of the phylum Actinobacteria make up an average of 13% (range, 0 to 34%) of soil bacterial communities (Fig. (Fig.1).1). The phylum Actinobacteria contains three subclasses that are common in soil: Actinobacteridae, Acidimicrobidae, and Rubrobacteridae (Table (Table4).4). For the purposes of this synthesis, the subclasses are used as subphylum groupings. The majority of described genera of the phylum Actinobacteria are within the subclass Actinobacteridae. This group consists of some 158 genera, many of which are well known and well studied (48). Although members of the subclass Actinobacteridae have been extensively investigated, only 26 to 47% of sequences could be assigned to described genera and there remains considerable scope for the isolation of novel members of this group, especially new rare genera that may yield novel bioactive compounds (67).

In addition to members of the subclass Actinobacteridae, soils contain many members of the less-studied subclasses Rubrobacteridae and Acidimicrobidae (Table (Table4).4). To date, there are no validly named and described members of the subclass Acidimicrobidae from soil, and only five isolates from soil been reported (29, 60). The only named and characterized members of this subclass are Acidimicrobium ferrooxidans and Ferromicrobium acidophilus, which are ferrous-iron-oxidizing acidophiles, and Microthrix parvicella, an as-yet-uncultured filamentous bacterium found in activated sewage sludge (6, 11, 22). Only Acidimicrobium ferrooxidans is currently recognized by Bergey's Manual (48).

Two genera of aerobic heterotrophs from soil in the subclass Rubrobacteridae have been described. These are Solirubrobacter and Conexibacter, each represented by one species with one strain each (80, 100). Twenty further isolates, some phylogenetically distant from these two genera, have been obtained from soil (29, 58, 60, 94, 96). These are all aerobic heterotrophs. The few other members of this subclass are aquatic thermophiles of the genera Rubrobacter and Thermoleophilum (19, 21, 106, 117). Overall, there are many lineages without cultured representatives in all three subclasses of soil-inhabiting actinobacteria, especially in the subclasses Rubrobacteridae and Acidimicrobidae, but also some in the subclass Actinobacteridae (e.g., see references 5, 50, 54, 73, 79, 89, and 90). Recently, members of some of the previously uncultured lineages of the Actinobacteridae have been shown to be culturable (60). The phylogenetic depth of the phylum Actinobacteria appears to be lower than that of other major phyla, but the degree of phenotypic diversity in this phylum is high (31, 47). The as-yet-uncultured actinobacterids can be expected to be aerobic heterotrophs, but the subclasses Rubrobacteridae and Acidimicrobidae may yet contain other metabolic types.

Members of the phylum Verrucomicrobia make up an average of 7% (range, 0 to 21%) of soil bacterial communities (Fig. (Fig.1).1). The phylum has been divided into five major class level subdivisions (56). The major group of Verrucomicrobia found in soil is the class Spartobacteria (Table (Table4),4), which is the name proposed for subdivision 2 of Verrucomicrobia (97). Chthoniobacter flavus is the first named cultured isolate of the class Spartobacteria (58, 97). C. flavus and a further nine isolates belong to at least two genera within the family Chthoniobacteraceae in the class Spartobacteria (96). These are all aerobic heterotrophs. The class Spartobacteria also contains bacterial symbionts of nematodes in the family Xiphinematobacteraceae (112), and the cloned sequences detected in soils may therefore have come from free-living or symbiotic bacteria. The 10 isolates of the class Spartobacteria do not cover the phylogenetic breadth of the class (96), and many unrepresented lineages (e.g., see references 5 and 56), indicative of novel genera and families, remain to be cultured. There is no indication of what the physiologies of members of those families could be. Only six isolates of subdivision 3 of the phylum Verrucomicrobia have been obtained (60, 96). These are aerobic heterotrophs, but they also do not cover the full phylogenetic breadth of the subdivision and so do not yet give a complete picture of the phenotypes of members of this group (96).

Members of the phylum Bacteroidetes make up an average of 5% (range, 0 to 18%) of soil bacterial communities (Fig. (Fig.1).1). There is some evidence suggesting that members of this phylum may be underrepresented in libraries of PCR-amplified 16S rRNA genes (27, 34). Even so, members of the class Sphingobacteria of the phylum Bacteroidetes are common in soils (Table (Table4).4). Some members of this group are aerobes, while others are anaerobes or facultative anaerobes, and so the species composition of members of this class within a soil may depend on oxygen levels or the amount of variation in oxygen availability. Members of the class Flavobacteria are less common (Table (Table4),4), and members of the third class, Bacteroidetes, seem to be absent from soils. Of all the sequences affiliated with the phylum Bacteroidetes, 34 to 62% could be assigned to known genera, including Chitinophaga, Flavobacterium, Hymenobacter, and Pedobacter, suggesting that this is one of the few groups of dominant soil bacteria that is readily culturable. However, given the high diversity of soil bacteria, we should not be surprised if many novel genera, albeit in low abundance, exist in soil. Lineages without cultured representatives have been detected (e.g., see references 72 and 77).

Members of the phylum Chloroflexi make up an average of 3% (range, 0 to 16%) of soil bacterial communities (Fig. (Fig.1).1). The phylum Chloroflexi consists of perhaps eight candidate classes, and the phylogenetic depth is comparable with the phylum Proteobacteria (31, 88). Only eight described genera are known, and these are not evenly distributed within the classes (48). Rappé and Giovannoni (88) and Hugenholtz et al. (56) recognize the Thermomicrobia as a class of the phylum Chloroflexi, with one described genus, but Garrity et al. (48) accord the Thermomicrobia separate phylum status. The diversity of phenotypes in this phylum is high, even among the relatively few isolates that have been cultured to date (88). 16S rRNA genes from uncultured soil bacteria belonging to the phylum Chloroflexi are affiliated with a number of the candidate classes and so may display quite different physiologies (88). Only one isolate from soil has been reported (29). It is a filamentous aerobic heterotroph, but no conclusions can be drawn yet about the general properties of soil chloroflexi. There is evidence that other members of this group are culturable (Davis and Janssen, unpublished data).

Members of the phylum Planctomycetes make up an average of 2% (range, 0 to 8%) of soil bacterial communities (Fig. (Fig.1).1). The planctomycetes are a group of budding bacteria that lack peptidoglycan and possess membrane-bound intracellular compartments (44). The phylogenetic depth within the group is sufficient to suggest that the phylum could be composed of at least three classes (31, 88). One of these could consist of bacteria such as those of the candidates Brocadia and Kuenenia, which are involved in anaerobic nitrification (88). The other two major groups are defined by the relatively well studied genera classified in the class Planctomycetacia (48) and by sequences affiliated with the WPS-1 lineage (82). Soil bacteria are affiliated with all three major groups, and there are many lineages without any cultured representatives (e.g., see references 50, 68, and 88). Most isolates of this phylum are from aquatic sources, and it is not clear whether these are physiologically and genetically good models for soil planctomycetes. Isolates from soil, including a few members of the WPS-1 lineage, have been reported (29, 60, 114), but these do not represent the full phylogenetic breadth suggested by the sequences detected in soils.

Members of the phylum Gemmatimonadetes make up an average of 2% (range, 0 to 4%) of soil bacterial communities (Fig. (Fig.1).1). The phylum Gemmatimonadetes contains only one named and described species, Gemmatimonas aurantiacus (120). This bacterium is a gram-negative aerobic heterotroph isolated from an anaerobic-aerobic sequential batch reactor and belongs to subdivision 1, also known as the class Gemmatimonadetes (48, 120). Four isolates from soil have been obtained (29, 60). They too belong to subdivision 1 and display an aerobic, heterotrophic phenotype. Soil-inhabiting representatives of the phylum are found through most of the phylogenetic breadth of the group, which may contain a number of discrete class rank taxa (81, 88, 120). The diversity of general physiologies of this group remains to be ascertained.

Members of the genera Bacillus and Clostridium have long been considered to be common members of the soil bacterial community, but the classes Bacilli and Clostridia of the phylum Firmicutes, together comprising of some 214 genera, including Bacillus and Clostridium (48), contribute only a mean of 2% (range, 0 to 8%) to the libraries (Fig. (Fig.1).1). It is possible that members of this group are underrepresented in libraries because the cells or spores may be difficult to lyse and so are not detected in PCR-based analyses that rely on DNA extraction from soil. Until evidence for such a bias is available, members of this group must be considered to be relatively minor components of soil bacterial communities. They may, however, be locally abundant, such as in a grassland soil in The Netherlands (38, 39). Of the sequences affiliated with the phylum Firmicutes in the 32 libraries analyzed, 17 to 52% could be assigned to known genera, suggesting that a number of new genera remain to be isolated, named, and described.

This review deals with members of the domain Bacteria, but members of the domain Archaea have also been detected in soils (10, 12, 17, 61, 111), although their abundance is generally low (17, 84, 99). These studies also reveal the presence of high-rank taxa of the domain Archaea with no cultured representatives. Some of these soil archaea may prove to have unexpected physiologies (42), and they appear to be culturable (99).

CHALLENGES AND GOALS

There is considerable variability in the abundance of members of different phyla and classes in different soils, judged by the abundance of 16S rRNA or 16S rRNA genes in libraries. It is not yet clear to what extent the variations are systematic, in response to conditions in the soil environment, and to what degree method-induced artifacts impact the data. The number of different biological, chemical, and physical factors that may influence the abundance of different bacterial groups could be very large. It has been suggested that the abundance of verrucomicrobia is influenced by soil moisture (14), and the abundance of members of subdivision 1 of the acidobacteria appears to be controlled by soil pH (93). It is not known whether the abundance of members of other high-rank taxa is controlled by single, readily identifiable factors. The degree of phenotypic variation within some of the groups must mean that the total abundance of a particular group may not change as much as the representation of species within that group, and so the abundance of such phenotypically diverse groups cannot be expected to be controlled by single variables.

Regardless of whether one is interested in functional or phylogenetic groupings, it is clear that the physiologies and characteristics of the poorly studied groups of soil bacteria must be of interest to soil microbiologists (49, 55, 85, 88, 122). Those interested in functions will want to identify all the major contributors to that function and will not want to disregard the possibility that bacteria among the poorly characterized part of the community are involved. The complexity of soil microbial communities means that metagenomic approaches to studying soil bacteria and assembling genomes of uncultured bacteria to understand their physiologies seem impractical at present (110). Although the assignment of functions and associated genes to phylogenetic markers is possible in the absence of cultures (49, 69), it can be achieved more easily with cultured bacteria.

Culturing of the total diversity of species, estimated at 104 to 106 per 10-g soil sample (28, 46, 109), currently seems an unreachable goal, given that such large numbers of isolates are not routinely cultured and identified and that the rarer species can be expected to be difficult to find among the colonies that do form. An initial objective should be to obtain a range of isolates from representatives of members of the different phyla and classes and to determine to what extent there is functional and genetic diversity within these groups. This will give us an initial overview of the potential roles of different soil bacterial groups and will also enable us to learn the tricks required to culture their more recalcitrant relatives. This strategy has been successfully applied to verrucomicrobia and acidobacteria (93, 96). At the same time, genome sequences of selected isolates will help fill in the bacterial genome tree (55).

Some advances in culturing soil bacteria have been made in the last few years (29, 58, 60, 93, 94, 96, 103, 107, 119). These recent advances mean that it is probably incorrect to speak of the majority of bacterial species in soil as being unculturable. Instead, we should be aware that isolating them will take patience and careful selection of appropriate strategies. Many of the isolates of rarely isolated groups are very slow growing and are difficult to maintain in the laboratory (29). The formation of visible colonies requires weeks or months rather than hours or days. Although the growth rates are low, they are still much higher than their likely growth rates in soil, where cells may divide only a few times per year (53). It is likely that the few isolates of Acidobacteria, Verrucomicrobia, Planctomycetes, Gemmatimonadetes, Chloroflexi, Acidimicrobidae, and Rubrobacteridae, as well as many of the isolates of the better-studied Proteobacteria, Bacteroidetes, and Actinobacteridae, are actually the more readily cultured representatives of these groups. To isolate type strains and a representative collection of related strains, and to deposit them in culture collections as required for the valid description of new species (101), will require great dedication and a high level of commitment and expertise from soil microbiologists and from culture collections. Successful approaches to culturing these organisms require patience, but the outcomes are immensely satisfying to microbiologists who enjoy the challenge and savor the reward of observing the colony on the plate, seeing the cells of the pure culture under the microscope, elucidating the bacterium's physiology, or releasing its genome sequence into public databases.

Acknowledgments

The results obtained in the author's laboratory have been due to the intellectual and experimental efforts of Sally Cairnduff, Kathyrn Davis, Bronwyn Grinton, Shayne Joseph, Suzana Kovac, Matthew O'Neill, Catherine Osborne, Michelle Sait, Parveen Sangwan, Liesbeth Schoenborn, and Penelope Yates.

Research has been supported by grants from the Australian Research Council.

REFERENCES

1. Achenbach, L. A., and J. D. Coates. 2000. Disparity between bacterial phylogeny and physiology. ASM News 66:714-716. [Google Scholar]
2. Alexander, M. 1977. Introduction to soil microbiology, 2nd ed. John Wiley and Sons, New York, N.Y.
3. Altschul, S. F., W. Gish, W. Miller, E. W. Myers, and D. J. Lipman. 1990. Basic local alignment search tool. J. Mol. Biol. 215:403-410. [PubMed] [Google Scholar]
4. Anzai, Y., H. Kim, J. Y. Park, H. Wakabayashi, and H. Oyaizu. 2000. Phylogenetic affiliation of the pseudomonads based on 16S rRNA sequence. Int. J. Syst. Evol. Microbiol. 50:1563-1589. [PubMed] [Google Scholar]
5. Axelrood, P. E., M. L. Chow, C. C. Radomski, J. M. McDermott, and J. Davies. 2002. Molecular characterization of bacterial diversity from British Columbia forest soils subjected to disturbance. Can. J. Microbiol. 48:655-674. [PubMed] [Google Scholar]
6. Bacelar-Nicolau, P., and D. B. Johnson. 1999. Leaching of pyrite by acidophilic heterotrophic iron-oxidizing bacteria in pure and mixed cultures. Appl. Environ. Microbiol. 65:585-590. [PMC free article] [PubMed] [Google Scholar]
7. Barns, S. M., S. L. Takala, and C. R. Kuske. 1999. Wide distribution and diversity of members of the bacterial kingdom Acidobacterium in the environment. Appl. Environ. Microbiol. 65:1731-1737. [PMC free article] [PubMed] [Google Scholar]
8. Benson, D. A., I. Karsch-Mizrachi, D. J. Lipman, J. Ostell, and D. L. Wheeler. 2005. GenBank. Nucleic Acids Res. 33:D34-D38. [PMC free article] [PubMed] [Google Scholar]
9. Bernardet, J. F., Y. Nakagawa, and B. Holmes. 2002. Proposed minimal standards for describing new taxa of the family Flavobacteriaceae and emended description of the family. Int. J. Syst. Evol. Microbiol. 52:1049-1070. [PubMed] [Google Scholar]
10. Bintrim, S. B., T. J. Donohue, J. Handelsman, G. P. Roberts, and R. M. Goodman. 1997. Molecular phylogeny of Archaea from soil. Proc. Natl. Acad. Sci. USA 94:277-282. [PMC free article] [PubMed] [Google Scholar]
11. Blackall, L. L., E. M. Seviour, M. A. Cunningham, R. J. Seviour, and P. Hugenholtz. 1994. “Microthrix parvicella” is a novel, deep branching member of the actinomycetes subphylum. Syst. Appl. Microbiol. 17:513-518. [Google Scholar]
12. Borneman, J., and E. W. Triplett. 1997. Molecular microbial diversity in soils from eastern Amazonia: evidence for unusual microorganisms and microbial population shifts associated with deforestation. Appl. Environ. Microbiol. 63:2647-2653. [PMC free article] [PubMed] [Google Scholar]
13. Borneman, J., P. W. Skroch, K. M. O'Sullivan, J. A. Palus, N. G. Rumjanek, J. L. Jansen, J. Nienhuis, and E. W. Triplett. 1996. Molecular microbial diversity of an agricultural soil in Wisconsin. Appl. Environ. Microbiol. 62:1935-1943. [PMC free article] [PubMed] [Google Scholar]
14. Buckley, D. H., and T. M. Schmidt. 2001. Environmental factors influencing the distribution of rRNA from Verrucomicrobia in soil. FEMS Microbiol. Ecol. 35:105-112. [PubMed] [Google Scholar]
15. Buckley, D. H., and T. M. Schmidt. 2001. The structure of microbial communities in soil and the lasting impact of cultivation. Microb. Ecol. 42:11-21. [PubMed] [Google Scholar]
16. Buckley, D. H., and T. M. Schmidt. 2003. Diversity and dynamics of microbial communities in soils from agro-ecosystems. Environ. Microbiol. 5:441-452. [PubMed] [Google Scholar]
17. Buckley, D. H., J. R. Graber, and T. M. Schmidt. 1998. Phylogenetic analysis of nonthermophilic members of the kingdom Crenarchaeota and their diversity and abundance in soils. Appl. Environ. Microbiol. 64:4333-4339. [PMC free article] [PubMed] [Google Scholar]
18. Caracciolo, A. B., P. Grenni, R. Ciccoli, G. Di Landa, and C. Cremisini. 2005. Simazine biodegradation in soil: analysis of bacterial community structure by in situ hybridization. Pest Manag. Sci. 61:863-869. [PubMed] [Google Scholar]
19. Carreto, L., E. Moore, M. F. Nobre, R. Wait, P. W. Riley, R. J. Sharp, and M. S. da Costa. 1996. Rubrobacter xylanophilus sp. nov., a new thermophilic species isolated from a thermally polluted effluent. Int. J. Syst. Bacteriol. 46:460-465. [Google Scholar]
20. Chatzinotas, A., R. A. Sandaa, W. Schönhuber, R. Amann, F. L. Daae, V. Torsvik, J. Zeyer, and D. Hahn. 1998. Analysis of broad-scale differences in microbial community composition of two pristine forest soils. Syst. Appl. Microbiol. 21:579-587. [PubMed] [Google Scholar]
21. Chen, M.-Y., S.-H. Wu, G.-H. Lin, C.-P. Lu, Y.-T. Lin, W.-C. Chang, and S.-S. Tsay. 2004. Rubrobacter taiwanensis sp. nov., a novel thermophilic, radiation-resistant species isolated from hot springs. Int. J. Syst. Evol. Microbiol. 54:1849-1855. [PubMed] [Google Scholar]
22. Clark, D. A., and P. R. Norris. 1996. Acidimicrobium ferrooxidans gen. nov., sp. nov.: mixed-culture ferrous iron oxidation with Sulfobacillus species. Microbiology 142:785-790. [PubMed] [Google Scholar]
23. Coates, J. D., D. J. Ellis, C. V. Gaw, and D. R. Lovley. 1999. Geothrix fermentans gen. nov., sp. nov., a novel Fe(III)-reducing bacterium from a hydrocarbon-contaminated aquifer. Int. J. Syst. Bacteriol. 49:1615-1622. [PubMed] [Google Scholar]
24. Cole, J. R., B. Chai, R. J. Farris, Q. Wang, S. A. Kulam, D. M. McGarrell, G. M. Garrity, and J. M. Tiedje. 2005. The Ribosomal Database Project (RDP-II): sequences and tools for high-throughput rRNA analysis. Nucleic Acids Res. 33:D294-D296. [PMC free article] [PubMed] [Google Scholar]
25. Conn, H. J. 1909. Future methods of soil bacteriological investigations. Zentbl. Bakteriol. Parasitenkd. Infektkrankh. Abt. 2 25:454-457. [Google Scholar]
26. Conn, H. J. 1918. The microscopic study of bacteria and fungi in soil. N. Y. Agric. Exp. Sta. Tech. Bull. 64:3-20. [Google Scholar]
27. Cottrell, M. T., and D. L. Kirchman. 2000. Community composition of marine bacterioplankton determined by 16S rRNA gene clone libraries and fluorescence in situ hybridization. Appl. Environ. Microbiol. 66:5116-5122. [PMC free article] [PubMed] [Google Scholar]
28. Curtis, T. P., W. T. Sloan, and J. C. Scannell. 2002. Estimating prokaryotic diversity and its limits. Proc. Natl. Acad. Sci. USA 99:10494-10499. [PMC free article] [PubMed] [Google Scholar]
29. Davis, K. E. R., S. J. Joseph, and P. H. Janssen. 2005. Effects of growth medium, inoculum size, and incubation time on the culturability and isolation of soil bacteria. Appl. Environ. Microbiol. 71:826-834. [PMC free article] [PubMed] [Google Scholar]
30. Demba Diallo, M., M. Martens, N. Vloemans, S. Cousin, T. T. M. Vanderkerckhove, M. Neyra, P. de Lajudie, A. Willems, M. Gillis, W. Vyverman, and K. Van der Gucht. 2004. Phylogenetic analysis of partial bacterial 16S rDNA sequences of tropical grass pasture soil under Acacia tortilis subsp. raddiana in Senegal. Syst. Appl. Microbiol. 27:238-252. [PubMed] [Google Scholar]
31. Dojka, M. A., J. K. Harris, and N. R. Pace. 2000. Expanding the known diversity and environmental distribution of an uncultured phylogenetic division of bacteria. Appl. Environ. Microbiol. 66:1617-1621. [PMC free article] [PubMed] [Google Scholar]
32. Dunbar, J., S. M. Barns, L. O. Ticknor, and C. R. Kuske. 2002. Empirical and theoretical bacterial diversity in four Arizona soils. Appl. Environ. Microbiol. 68:3035-3045. [PMC free article] [PubMed] [Google Scholar]
33. Dunbar, J., S. Takala, S. M. Barns, J. A. Davis, and C. R. Kuske. 1999. Levels of bacterial community diversity in four arid soils compared by cultivation and 16S rRNA gene cloning. Appl. Environ. Microbiol. 65:1662-1669. [PMC free article] [PubMed] [Google Scholar]
34. Eilers, H., J. Pernthaler, F. O. Glöckner, and R. Amann. 2000. Culturability and in situ abundance of pelagic bacteria from the North Sea. Appl. Environ. Microbiol. 66:3044-3051. [PMC free article] [PubMed] [Google Scholar]
35. Ellis, R. J., P. Morgan, A. J. Weightman, and J. C. Fry. 2003. Cultivation-dependent and -independent approaches for determining bacterial diversity in heavy-metal-contaminated soil. Appl. Environ. Microbiol. 69:3223-3230. [PMC free article] [PubMed] [Google Scholar]
36. Embley, T. M., and E. Stackebrandt. 1997. Species in practice: exploring uncultured prokaryote diversity in natural samples, p. 61-81. In M. F. Claridge, H. A. Dawah, and M. R. Wilson (ed.), Species: the units of biodiversity. Chapman and Hall, London, United Kingdom.
37. Everett, K. D., R. M. Bush, and A. A. Andersen. 1999. Emended description of the order Chlamydiales, proposal of Parachlamydiaceae fam. nov. and Simkaniaceae fam. nov., each containing one monotypic genus, revised taxonomy of the family Chlamydiaceae, including a new genus and five new species, and standards for the identification of organisms. Int. J. Syst. Bacteriol. 49:415-440. [PubMed] [Google Scholar]
38. Felske, A., A. Wolterink, R. van Lis, and A. D. L. Akkermans. 1998. Phylogeny of the main bacterial 16S rRNA sequences in Drentse A grassland soils (The Netherlands). Appl. Environ. Microbiol. 64:871-879. [PMC free article] [PubMed] [Google Scholar]
39. Felske, A., A. Wolterink, R. van Lis, W. M. de Vos, and A. D. Akkermans. 2000. Response of a soil bacterial community to grassland succession as monitored by 16S rRNA levels of the predominant ribotypes. Appl. Environ. Microbiol. 66:3998-4003. [PMC free article] [PubMed] [Google Scholar]
40. Fierer, N., J. A. Jackson, R. Vilgalys, and R. B. Jackson. 2005. Assessment of soil microbial community structure by use of taxon-specific quantitative PCR assays. Appl. Environ. Microbiol. 71:4117-4120. [PMC free article] [PubMed] [Google Scholar]
41. Floyd, M. M., J. Tang, M. Kane, and D. Emerson. 2005. Captured diversity in a culture collection: case study of the geographic and habitat distributions of environmental isolates held at the American Type Culture Collection. Appl. Environ. Microbiol. 71:2813-2823. [PMC free article] [PubMed] [Google Scholar]
42. Francis, C. A., K. J. Roberts, J. M. Beman, A. E. Santoro, and B. B. Oakley. 2005. Ubiquity and diversity of ammonia-oxidizing archaea in water columns and sediments of the ocean. Proc. Natl. Acad. Sci. USA 102:14683-14688. [PMC free article] [PubMed] [Google Scholar]
43. Frostegård, Å., S. Courtois, V. Ramisse, S. Clerc, D. Bernillon, F. Le Gall, P. Jeannin, X. Nesme, and P. Simonet. 1999. Quantification of bias related to the extraction of DNA directly from soils. Appl. Environ. Microbiol. 65:5409-5420. [PMC free article] [PubMed] [Google Scholar]
44. Fuerst, J. A. 2005. Intracellular compartmentation in planctomycetes. Annu. Rev. Microbiol. 59:299-328. [PubMed] [Google Scholar]
45. Furlong, M. A., D. R. Singleton, D. C. Coleman, and W. B. Whitman. 2002. Molecular and culture-based analyses of prokaryotic communities from an agricultural soil and the burrows and casts of the earthworm Lumbricus rubellus. Appl. Environ. Microbiol. 68:1265-1279. [PMC free article] [PubMed] [Google Scholar]
46. Gans, J., M. Wolinsky, and J. Dunbar. 2005. Computational improvements reveal great bacterial diversity and high metal toxicity in soil. Science 309:1387-1390. [PubMed] [Google Scholar]
47. Garrity, G. M., and J. G. Holt. 2001. The road map to the Manual, p. 119-166. In D. R. Boone and R. W. Castenholz (ed.), Bergey's manual of systematic bacteriology, vol. 1: the Archaea and the deeply branching and phototrophic Bacteria. Springer-Verlag, New York, N.Y. [Google Scholar]
48. Garrity, G. M., J. A. Bell, and T. G. Lilburn. 2004. Taxonomic outline of the prokaryotes, release 5.0. Springer-Verlag, New York, N.Y.
49. Gray, N. D., and I. M. Head. 2001. Linking genetic identity and function in communities of uncultured bacteria. Environ. Microbiol. 3:481-492. [PubMed] [Google Scholar]
50. Gremion, F., A. Chatzinotas, and H. Harms. 2003. Comparative 16S rDNA and 16S rRNA sequence analysis indicates that Actinobacteria might be a dominant part of the metabolically active bacteria in heavy metal contaminated bulk and rhizosphere soil. Environ. Microbiol. 5:896-907. [PubMed] [Google Scholar]
51. Hackl, E., S. Zechmeister-Boltenstern, L. Bodrossy, and A. Sessitsch. 2004. Comparison of diversities and compositions of bacterial populations inhabiting natural forest soils. Appl. Environ. Microbiol. 70:5057-5065. [PMC free article] [PubMed] [Google Scholar]
52. Hallberg, K. B., and D. B. Johnson. 2003. Novel acidophiles isolated from moderately acidic mine drainage waters. Hydrometallurgy 71:139-148. [Google Scholar]
53. Harris, D., and E. A. Paul. 1994. Measurements of bacterial growth rates in soil. Appl. Soil Ecol. 1:277-290. [Google Scholar]
54. Holmes, A. J., J. Bowyer, M. P. Holley, M. O'Donoghue, M. Montgomery, and M. R. Gillings. 2002. Diverse, yet-to-be-cultured members of the Rubrobacter subdivision of the Actinobacteria are widespread in Australian arid soils. FEMS Microbiol. Ecol. 33:111-120. [PubMed] [Google Scholar]
55. Hugenholtz, P. 2002. Exploring prokaryotic diversity in the genomic era. Genome Biol. 3:Reviews0003.1-0003.8. [Online.] http://genomebiology.com/2002/3/2/reviews/0003. [PMC free article] [PubMed] [Google Scholar]
56. Hugenholtz, P., B. M. Goebel, and N. R. Pace. 1998. Impact of culture-independent studies on the emerging phylogenetic view of bacterial diversity. J. Bacteriol. 180:4765-4774. [PMC free article] [PubMed] [Google Scholar]
57. Hugenholtz, P., G. W. Tyson, R. I. Webb, A. M. Wagner, and L. L. Blackall. 2001. Investigation of candidate division TM7, a recently recognized major lineage of the domain Bacteria with no known pure-culture representatives. Appl. Environ. Microbiol. 67:411-419. [PMC free article] [PubMed] [Google Scholar]
58. Janssen, P. H., P. S. Yates, B. E. Grinton, P. M. Taylor, and M. Sait. 2002. Improved culturability of soil bacteria and isolation in pure culture of novel members of the divisions Acidobacteria, Actinobacteria, Proteobacteria, and Verrucomicrobia. Appl. Environ. Microbiol. 68:2391-2396. [PMC free article] [PubMed] [Google Scholar]
59. Jensen, V. 1968. The plate count technique, p. 158-170. In T. R. G. Gray and D. Parkinson (ed.), The ecology of soil bacteria. Liverpool University Press, Liverpool, United Kingdom.
60. Joseph, S. J., P. Hugenholtz, P. Sangwan, C. A. Osborne, and P. H. Janssen. 2003. Laboratory cultivation of widespread and previously uncultured soil bacteria. Appl. Environ. Microbiol. 69:7210-7215. [PMC free article] [PubMed] [Google Scholar]
61. Jurgens, G., K. Lindstrom, and A. Saano. 1997. Novel group within the kingdom Crenarchaeota from boreal forest soil. Appl. Environ. Microbiol. 63:803-805. [PMC free article] [PubMed] [Google Scholar]
62. Kishimoto, N., and T. Tano. 1987. Acidophilic heterotrophic bacteria isolated from acidic mine drainage, sewage and soils. J. Gen. Appl. Microbiol. 33:11-25. [Google Scholar]
63. Kishimoto, N., Y. Kosako, and T. Tano. 1991. Acidobacterium capsulatum gen. nov., sp. nov.: an acidophilic chemoorganotrophic bacterium containing menaquinone from acidic mineral environment. Curr. Microbiol. 22:1-7. [PubMed] [Google Scholar]
64. Kobabe, S., D. Wagner, and E.-M. Pfeiffer. 2004. Characterisation of microbial community composition of a Siberian tundra soil by fluorescence in situ hybridisation. FEMS Microbiol. Ecol. 50:13-23. [PubMed] [Google Scholar]
65. Kuske, C. R., S. M. Barns, and J. D. Busch. 1997. Diverse uncultivated bacterial groups from soils of the arid southwestern United States that are present in many geographic regions. Appl. Environ. Microbiol. 63:3614-3621. [PMC free article] [PubMed] [Google Scholar]
66. Lane, D. J. 1991. 16S/23S rRNA sequencing, p. 115-175. In E. Stackebrandt and M. Goodfellow (ed.), Nucleic acid techniques in bacterial systematics. John Wiley & Sons, Chichester, United Kingdom.
67. Lazzarini, A., L. Cavaletti, G. Troppo, and F. Marinelli. 2001. Rare genera of actinomycetes as potential sources of new antibiotics. Antonie Leeuwenhoek 79:399-405. [PubMed] [Google Scholar]
68. Liesack, W., and E. Stackebrandt. 1992. Occurrence of novel groups of the domain Bacteria as revealed by analysis of genetic material isolated from an Australian terrestrial environment. J. Bacteriol. 174:5072-5078. [PMC free article] [PubMed] [Google Scholar]
69. Liesack, W., and P. F. Dunfield. 2002. Biodiversity in soils: use of molecular methods for its characterization, p. 528-544. In G. Bitton (ed.), Encyclopedia of environmental microbiology. John Wiley and Sons, New York, N.Y.
70. Liesack, W., F. Bak, J.-U. Kreft, and E. Stackebrandt. 1994. Holophaga foetida gen. nov., sp. nov., a new, homoacetogenic bacterium degrading methoxylated aromatic compounds. Arch. Microbiol. 162:85-90. [PubMed] [Google Scholar]
71. Liles, M. R., B. F. Manske, S. B. Bintrim, J. Handelsman, and R. M. Goodman. 2002. A census of rRNA genes and linked genomic sequences within a soil metagenomic library. Appl. Environ. Microbiol. 69:2684-2691. [PMC free article] [PubMed] [Google Scholar]
72. Lipson, D. A., and S. K. Schmidt. 2004. Seasonal changes in an alpine soil bacterial community in the Colorado Rocky Mountains. Appl. Environ. Microbiol. 70:2867-2879. [PMC free article] [PubMed] [Google Scholar]
73. Lüdemann, H., and R. Conrad. 2000. Molecular retrieval of large 16S rRNA gene fragments from an Italian rice paddy soil affiliated with the class Actinobacteria. Syst. Appl. Microbiol. 23:582-584. [PubMed] [Google Scholar]
74. Ludwig, W., and H.-P. Klenk. 2001. Overview: a phylogenetic backbone and taxonomic framework for procaryotic systematics, p. 49-65. In D. R. Boone and R. W. Castenholz (ed.), Bergey's manual of systematic bacteriology, vol. 1: the Archaea and the deeply branching and phototrophic Bacteria. Springer-Verlag, New York, N.Y. [Google Scholar]
75. Ludwig, W., S. H. Bauer, M. Bauer, I. Held, G. Kirchhof, R. Schulze, I. Huber, S. Spring, A. Hartman, and K. H. Schleifer. 1997. Detection and in situ identification of a widely distributed new bacterial phylum. FEMS Microbiol. Lett. 153:181-190. [PubMed] [Google Scholar]
76. Marilley, L., and M. Aragno. 1999. Phylogenetic diversity of bacterial communities differing in degree of proximity of Lolium perene and Trifolium repens roots. Appl. Soil Ecol. 13:127-136. [Google Scholar]
77. McCaig, A. E., L. A. Glover, and J. I. Prosser. 1999. Molecular analysis of bacterial community structure and diversity in unimproved and improved upland grass pastures. Appl. Environ. Microbiol. 65:1721-1730. [PMC free article] [PubMed] [Google Scholar]
78. McCaig, A. E., S. J. Grayston, J. I. Prosser, and L. A. Glover. 2001. Impact of cultivation on characterisation of species composition of soil bacterial communities. FEMS Microbiol. Ecol. 35:37-48. [PubMed] [Google Scholar]
79. McVeigh, H. P., J. Munro, and T. M. Embley. 1996. Molecular evidence for the presence of novel actinomycete lineages in a temperate forest soil. J. Ind. Microbiol. 17:197-204. [Google Scholar]
80. Monciardini, P., L. Cavaletti, P. Schumann, M. Rohde, and S. Donadio. 2003. Conexibacter woesii gen. nov., sp. nov., a novel representative of a deep evolutionary line of descent within the class Actinobacteria. Int. J. Syst. Evol. Microbiol. 53:569-576. [PubMed] [Google Scholar]
81. Mummey, D. L., and P. D. Stahl. 2003. Candidate division BD: phylogeny, distribution and abundance in soil ecosystems. Syst. Appl. Microbiol. 26:228-235. [PubMed] [Google Scholar]
82. Nogales, B., E. R. B. Moore, E. Llobet-Brossa, R. Rossello-Mora, R. Amann, and K. N. Timmis. 2001. Combined use of 16S ribosomal DNA and 16S rRNA to study the bacterial community of polychlorinated biphenyl-polluted soil. Appl. Environ. Microbiol. 67:1874-1884. [PMC free article] [PubMed] [Google Scholar]
83. Nogales, B., E. R. B. Moore, W. R. Abraham, and K. N. Timmis. 1999. Identification of the metabolically active members of a bacterial community in a polychlorinated biphenyl-polluted moorland soil. Environ. Microbiol. 1:199-212. [PubMed] [Google Scholar]
84. Ochsenreiter, T., D. Selezi, A. Quaiser, L. Bonch-Osmolovskaya, and C. Schleper. 2003. Diversity and abundance of Crenarchaeota in terrestrial habitats studied by 16S RNA surveys and real time PCR. Environ. Microbiol. 5:787-797. [PubMed] [Google Scholar]
85. Palleroni, N. J. 1997. Prokaryotic diversity and the importance of culturing. Antonie Leeuwenhoek 72:3-19. [PubMed] [Google Scholar]
86. Paul, E. A., and F. E. Clark. 1996. Soil microbiology and biochemistry, 2nd ed. Academic Press, San Diego, Calif.
87. Quaiser, A., T. Ochsenreiter, C. Lanz, S. C. Schuster, A. H. Treusch, J. Eck, and C. Schleper. 2003. Acidobacteria form a coherent but highly diverse group within the bacterial domain: evidence from environmental genomics. Mol. Microbiol. 50:563-575. [PubMed] [Google Scholar]
88. Rappé, M. S., and S. J. Giovannoni. 2003. The uncultured microbial majority. Annu. Rev. Microbiol. 57:369-394. [PubMed] [Google Scholar]
89. Rheims, H., and E. Stackebrandt. 1999. Application of nested polymerase chain reaction for the detection of as yet uncultured organisms of the class Actinobacteria in environmental samples. Environ. Microbiol. 1:137-143. [PubMed] [Google Scholar]
90. Rheims, H., A. Felske, S. Seufert, and E. Stackebrandt. 1999. Molecular monitoring of an uncultured group of the class Actinobacteria in two terrestrial environments. J. Microbiol. Methods 36:65-75. [PubMed] [Google Scholar]
91. Rheims, H., F. A. Rainey, and E. Stackebrandt. 1996. A molecular approach to search for diversity among bacteria in the environment. J. Ind. Microbiol. 17:159-169. [Google Scholar]
92. Rösch, C., A. Mergel, and H. Bothe. 2002. Biodiversity of denitrifying and dinitrogen-fixing bacteria in an acid forest soil. Appl. Environ. Microbiol. 68:3818-3829. [PMC free article] [PubMed] [Google Scholar]
93. Sait, M., K. E. R. Davis, and P. H. Janssen. 2006. Effect of pH on isolation and distribution of members of subdivision 1 of the phylum Acidobacteria occurring in soil. Appl. Environ. Microbiol. 72:1852-1857. [PMC free article] [PubMed] [Google Scholar]
94. Sait, M., P. Hugenholtz, and P. H. Janssen. 2002. Cultivation of globally-distributed soil bacteria from phylogenetic lineages previously only detected in cultivation-independent surveys. Environ. Microbiol. 4:654-666. [PubMed] [Google Scholar]
95. Saito, A., H. Mitsui, R. Hattori, K. Minamisawa, and T. Hattori. 1998. Slow-growing and oligotrophic bacteria phylogenetically close to Bradyrhizobium japonicum. FEMS Microbiol. Ecol. 25:277-286. [Google Scholar]
96. Sangwan, P., S. Kovac, K. E. R. Davis, M. Sait, and P. H. Janssen. 2005. Detection and cultivation of soil verrucomicrobia. Appl. Environ. Microbiol. 71:8402-8410. [PMC free article] [PubMed] [Google Scholar]
97. Sangwan, P., X. Chen, P. Hugenholtz, and P. H. Janssen. 2004. Chthoniobacter flavus gen. nov., sp. nov., the first pure-culture representative of subdivision two, Spartobacteria classis nov., of the phylum Verrucomicrobia. Appl. Environ. Microbiol. 70:5875-5881. [PMC free article] [PubMed] [Google Scholar]
98. Schloss, P. D., and J. Handelsman. 2004. Status of the microbial census. Microbiol. Mol. Biol. Rev. 68:686-691. [PMC free article] [PubMed] [Google Scholar]
99. Simon, H. M., C. E. Jahn, L. T. Bergerud, M. K. Sliwinski, P. J. Weimer, D. K. Willis, and R. M. Goodman. 2005. Cultivation of mesophilic soil crenarchaeotes in enrichment cultures from plant roots. Appl. Environ. Microbiol. 71:4751-4760. [PMC free article] [PubMed] [Google Scholar]
100. Singleton, D. R., M. A. Furlong, A. D. Peacock, D. C. White, D. C. Coleman, and W. B. Whitman. 2003. Solirubrobacter pauli gen. nov., sp. nov., a mesophilic bacterium within the Rubrobacteridae related to common soil clones. Int. J. Syst. Evol. Microbiol. 53:485-490. [PubMed] [Google Scholar]
101. Stackebrandt, E., W. Frederiksen, G. M. Garrity, P. A. Grimont, P. Kämpfer, M. C. Maiden, X. Nesme, R. Rosselló-Mora, J. Swings, H. G. Trüper, L. Vauterin, A. C. Ward, and W. B. Whitman. 2002. Report of the ad hoc committee for the re-evaluation of the species definition in bacteriology. Int. J. Syst. Evol. Microbiol. 52:1043-1047. [PubMed] [Google Scholar]
102. Stein, S., D. Selesi, R. Schilling, I. Pattis, M. Schmid, and A. Hartmann. 2005. Microbial activity and bacterial composition of H2-treated soils with net CO2 fixation. Soil Biol. Biochem. 37:1938-1945. [Google Scholar]
103. Stevenson, B. S., S. A. Eichorst, J. T. Wertz, T. M. Schmidt, and J. A. Breznak. 2004. New strategies for cultivation and detection of previously uncultured microbes. Appl. Environ. Microbiol. 70:4748-4755. [PMC free article] [PubMed] [Google Scholar]
104. Strohl, W. R. 2004. Antimicrobials, p. 336-355. In A. T. Bull (ed.), Microbial diversity and bioprospecting. ASM Press, Washington, D.C.
105. Sun, H. Y., S. P. Deng, and W. R. Raun. 2004. Bacterial community structure and diversity in a century-old manure-treated agroecosystem. Appl. Environ. Microbiol. 70:5868-5874. [PMC free article] [PubMed] [Google Scholar]
106. Suzuki, K., M. D. Collins, E. Iijima, and K. Komagata. 1988. Chemotaxonomic characterization of a radiotolerant bacterium, Arthrobacter radiotolerans: description of Rubrobacter radiotolerans gen. nov., comb. nov. FEMS Microbiol. Lett. 52:33-40. [Google Scholar]
107. Svenning, M. M., I. Wartiainen, A. G. Hestnes, and S. J. Binnerup. 2003. Isolation of methane oxidising bacteria from soil by use of a soil substrate membrane system. FEMS Microbiol. Ecol. 44:347-354. [PubMed] [Google Scholar]
108. Tanner, M. A., B. M. Goebel, M. A. Dojka, and N. R. Pace. 1998. Specific ribosomal DNA sequences from diverse environmental settings correlate with experimental contaminants. Appl. Environ. Microbiol. 64:3110-3113. [PMC free article] [PubMed] [Google Scholar]
109. Torsvik, V., J. Goksøyr, and F. L. Daae. 1990. High diversity in DNA of soil bacteria. Appl. Environ. Microbiol. 56:782-787. [PMC free article] [PubMed] [Google Scholar]
110. Tringe, S. G., C. von Mering, A. Kobayashi, A. A. Salamov, K. Chen, H. W. Chang, M. Podar, J. M. Short, E. J. Mathur, J. C. Detter, P. Bork, P. Hugenholtz, and E. M. Rubin. 2005. Comparative metagenomics of microbial communities. Science 308:554-557. [PubMed] [Google Scholar]
111. Ueda, T., Y. Suga, and T. Matsuguchi. 1996. Molecular phylogenetic analysis of a soil microbial community in a soybean field. Eur. J. Soil Sci. 46:415-421. [Google Scholar]
112. Vandekerkhove, T. T. M., A. Willems, M. Gillis, and A. Coomans. 2000. Occurrence of novel verrucomicrobial species, endosymbiotic and associated with parthenogenesis in Xiphinema americanum-group species (Nematoda, Longidoridae). Int. J. Syst. Evol. Microbiol. 50:2197-2205. [PubMed] [Google Scholar]
113. von Wintzingerode, F., U. B. Göbel, and E. Stackebrandt. 1997. Determination of microbial diversity in environmental samples: pitfalls of PCR-based rRNA analysis. FEMS Microbiol. Rev. 21:213-229. [PubMed] [Google Scholar]
114. Wang, J., C. Jenkins, R. I. Webb, and J. A. Fuerst. 2002. Isolation of Gemmata-like and Isosphaera-like planctomycete bacteria from soil and freshwater. Appl. Environ. Microbiol. 68:417-422. [PMC free article] [PubMed] [Google Scholar]
115. 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. 37:463-464. [Google Scholar]
116. Weber, S., S. Stubner, and R. Conrad. 2001. Bacterial populations colonizing and degrading rice straw in anoxic paddy soil. Appl. Environ. Microbiol. 67:1318-1327. [PMC free article] [PubMed] [Google Scholar]
117. Yakimov, M. M., H. Lünsdorf, and P. N. Golyshin. 2003. Thermoleophilum album and Thermoleophilum minutum are culturable representatives of group 2 of the Rubrobacteridae (Actinobacteria). Int. J. Syst. Evol. Microbiol. 53:377-380. [PubMed] [Google Scholar]
118. Zarda, B., D. Hahn, A. Chatzinotas, W. Schönhuber, A. Neef, R. I. Amann, and J. Zeyer. 1997. Analysis of bacterial community structure in bulk soil by in situ hybridization. Arch. Microbiol. 168:185-192. [Google Scholar]
119. Zengler, K., G. Toledo, M. Rappé, J. Elkins, E. J. Mathur, J. M. Short, and M. Keller. 2002. Cultivating the uncultured. Proc. Natl. Acad. Sci. USA 99:15684-15686. [PMC free article] [PubMed] [Google Scholar]
120. Zhang, H., Y. Sekiguchi, S. Hanada, P. Hugenholtz, H. Kim, Y. Kamagata, and K. Nakamura. 2003. Gemmatimonas aurantiaca gen. nov., sp. nov., a gram-negative, aerobic, polyphosphate-accumulating micro-organism, the first cultured representative of the new bacterial phylum Gemmatimonadetes phyl. nov. Int. J. Syst. Evol. Microbiol. 53:1155-1163. [PubMed] [Google Scholar]
121. Zhou, J., B. Xia, H. Huang, D. S. Treves, L. J. Hauser, R. J. Mural, A. V. Palumbo, and J. M. Tiedje. 2003. Bacterial phylogenetic diversity and a novel candidate division of two humid region, sandy surface soils. Soil Biol. Biochem. 35:915-924. [Google Scholar]
122. Zinder, S. H., and A. A. Salyers. 2001. Microbial ecology—new directions, new importance, p. 101-109. In D. R. Boone and R. W. Castenholz (ed.), Bergey's manual of systematic bacteriology, vol. 1: the Archaea and the deeply branching and phototrophic Bacteria. Springer-Verlag, New York, N.Y. [Google Scholar]

Articles from Applied and Environmental Microbiology are provided here courtesy of American Society for Microbiology (ASM)

-