Next Article in Journal
A Plea for a New Synthesis: From Twentieth-Century Paleobiology to Twenty-First-Century Paleontology and Back Again
Next Article in Special Issue
Nested Real-Time PCR Assessment of Vertical Transmission of Sandalwood Spike Phytoplasma (‘Ca. Phytoplasma asteris’)
Previous Article in Journal
Quantitative Proteomics and Network Analysis of Differentially Expressed Proteins in Proteomes of Icefish Muscle Mitochondria Compared with Closely Related Red-Blooded Species
Previous Article in Special Issue
Impact of the “Flavescence Dorée” Phytoplasma on Xylem Growth and Anatomical Characteristics in Trunks of ‘Chardonnay’ Grapevines (Vitis vinifera)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phytoplasma Taxonomy: Nomenclature, Classification, and Identification

Molecular Plant Pathology Laboratory, Beltsville Agricultural Research Center, Agricultural Research Service, United States Department of Agriculture, Beltsville, MD 20705, USA
*
Author to whom correspondence should be addressed.
Biology 2022, 11(8), 1119; https://doi.org/10.3390/biology11081119
Submission received: 30 June 2022 / Revised: 22 July 2022 / Accepted: 25 July 2022 / Published: 26 July 2022

Abstract

:

Simple Summary

Phytoplasmas are vector-borne and graft-transmissible bacteria that cause various plant diseases, leading to severe economic losses. Since phytoplasmas cannot be cultured in cell-free media, their identification and taxonomy rely on molecular techniques and gene sequences. In this article, we summarize the recent advances in phytoplasma taxonomy from three different aspects, including (i) nomenclature (naming Candidatus Phytoplasma species); (ii) classification (group and subgroup assignment based on 16S rRNA gene sequences); and (iii) identification (fine differentiation of phytoplasma strains). In addition, some important issues, especially those related to recognizing new ‘Candidatus Phytoplasma’ species, are discussed. This information will be helpful for rapid diagnosis of phytoplasma diseases and accurate taxonomic identification of both emerging and known phytoplasma strains.

Abstract

Phytoplasmas are pleomorphic, wall-less intracellular bacteria that can cause devastating diseases in a wide variety of plant species. Rapid diagnosis and precise identification of phytoplasmas responsible for emerging plant diseases are crucial to preventing further spread of the diseases and reducing economic losses. Phytoplasma taxonomy (identification, nomenclature, and classification) has lagged in comparison to culturable bacteria, largely due to lack of axenic phytoplasma culture and consequent inaccessibility of phenotypic characteristics. However, the rapid expansion of molecular techniques and the advent of high throughput genome sequencing have tremendously enhanced the nucleotide sequence-based phytoplasma taxonomy. In this article, the key events and milestones that shaped the current phytoplasma taxonomy are highlighted. In addition, the distinctions and relatedness of two parallel systems of ‘Candidatus phytoplasma’ species/nomenclature system and group/subgroup classification system are clarified. Both systems are indispensable as they serve different purposes. Furthermore, some hot button issues in phytoplasma nomenclature are also discussed, especially those pertinent to the implementation of newly revised guidelines for ‘Candidatus Phytoplasma’ species description. To conclude, the challenges and future perspectives of phytoplasma taxonomy are briefly outlined.

1. Introduction

Plant diseases characterized by flower abnormality, yellowing, and witches’-broom were long thought to be caused by viruses until 1967 [1,2], when Doi et al. discovered small bacteria with pleomorphism and lack of cell walls in ultrathin electron microscopic sections of infected phloem tissues [3]. The bacteria were named mycoplasma-like organisms (MLOs) because of their morphological resemblance to the mycoplasmas that infect humans and animals [3]. The first 16S rRNA gene sequence of MLO (Oenothera MLO 86-7, accession number M30790) was reported in 1989, which was distinct from mycoplasmas in phylogeny [4]. In 1993, according to the proposal of International Committee on Systematic Bacteriology (ICSB) Subcommittee on the Taxonomy of Mollicutes, the new trivial name, phytoplasma, was used to replace MLO [5]. Phytoplasma is derived from two Greek words, phyto (plant) and plasma (a thing), emphasizing its plant origin. Phytoplasmas have undergone reductive evolution, losing many of the genes involved in metabolic pathways that are essential for free-living organisms [6,7,8]. This not only makes phytoplasmas highly depend on host nutrition, but also is the main reason why axenic culture of phytoplasma has not been achieved yet despite various efforts (Figure 1) [1,2,3,4,5,6,9,10,11,12,13,14,15,16,17,18,19,20,21,22].
Taxonomy (from Greek taxis (arrangement) and nomos (law)) is a broad biological science concerned with nomenclature, classification and identification, which are three related but distinct aspects [23]. Nomenclature is the naming of organisms. Scientific names established according to the binomial nomenclature help scientists around the world better communicate and study the same organism(s). Classification is the orderly arrangement of organisms into groups or taxa based on their similarity. Identification is to recognize a known or unknown organism and to assign it to an existing or a new taxon [23]. Nomenclature is more academic, but classification can be designed according to practical needs [24]. In recent years, significant progress has been made in the bacterial taxonomy; for example, transition from 16S rRNA gene toward whole genome sequence in culturable bacteria [20]. For unculturable phytoplasma, the taxonomy has also been advanced in order to adapt and align better with culturable bacteria. In this article, recent advancements in phytoplasma taxonomy are explored from nomenclature, classification and identification.

2. Phytoplasma Nomenclature: Delineation of Candidatus Phytoplasma Species

Traditional polyphasic approach, which integrates phenotypic and genotypic data and reflects the ecological nature of the bacteria, is considered as the gold standard for bacterial taxonomy [25]. The phenotypic markers mainly include morphological, physiological, and biochemical characteristics of cultivatable bacteria [26]; however, inability to culture phytoplasma in vitro impeded the accessibility of the above-mentioned phenotypic characteristics to differentiate phytoplasmas. Several decades ago, scientists attempted to distinguish phytoplasmas by using symptoms induced by phytoplasmas, plant host range, insect vector specificity and serological correlations as markers, but were ultimately unsuccessful due to lack of consistency [27,28,29,30,31]. The subsequent development of culture-independent modern genotypic approach based on heredity information has rapidly and considerably enhanced the entire bacterial systematics, providing high levels of resolution and differentiation. In particular, the advent of DNA sequencing technology and exploitation of 16S rRNA gene sequences have tremendously facilitated taxonomy, tree of life, evolution, and diversity studies of unculturable bacteria [32,33,34]. Based on 16S rRNA gene sequences, many bacteria have been reclassified and renamed [35,36].
As with many other unculturable bacteria, the higher rank taxa of phytoplasmas (Mycoplasmatota (originally named Tenericutes_/Mollicutes/Acholeplasmatales/incertae sedis—Family II]) were named in the absence of type genus and species [37,38,39]. While the Candidatus status was used to reserve the putative lower rank taxa (Genus and Species [10]). The term Candidatus was first introduced in 1994 to nonculturable bacteria, granting appropriate status of potential taxa based on 16S rRNA gene sequences ([10]; Figure 1). Candidatus is not a rank, nor is it governed by Prokaryotic Code [40]. Currently, all phytoplasma strains are accommodated within the provisional Candidatus Phytoplasma genus. The main function of the phytoplasma taxonomic nomenclature system is naming ‘Candidatus Phytoplasma’ species as species is the most basic taxon of bacteria [12].

2.1. Transition from 16S rRNA Gene to Whole Genome-Based Nomenclature of Candidatus Phytoplasma Species?

The first ‘Candidatus Phytoplasma’ species (Ca. Phytoplasma. aurantifolia) was named based on 16S rRNA gene sequence in 1995 [41]. In 2004, the detailed guidelines (rules a through g) for naming Candidatus Phytoplasma were proposed by the IRPCM Phytoplasma/Spiroplasma Working Team—Phytoplasma taxonomy group [12]. According to the guidelines, Ca. Phytoplasma species may be delineated based on the identity of their 16S rRNA gene sequences greater than 1200 bp (rule a). A new ‘Candidatus Phytoplasma’ species can be recognized if the phytoplasma shares lower than 97.5% sequence identity in 16S rRNA gene with previously established ‘Candidatus Phytoplasma’ species (rule b). Such sequence identity threshold value was adopted because it corresponded to DNA-DNA hybridization (DDH) reassociation value (70%) suitable for demarcating bacterial species. If a phytoplasma shares higher than 97.5% sequence identity in 16S rRNA gene with existing species but clearly represents “an ecologically separated population”, the phytoplasma also qualifies for a new ‘Candidatus species’ (rule c).
In genotypic characterization, 16S rRNA gene serves as a backbone for bacterial taxonomy [35,39,40,41,42]. In many bacterial genera, the 16S rRNA gene alone is not sufficient to differentiate species, and, in some cases, multi-locus sequence analysis (MLSA) of alternative housekeeping genes is needed for phylogenetic studies [43,44,45,46]. In addition, with the advancement of genome sequencing technology, whole genome sequence-based genotypic characterization becomes possible. In recent years, the whole-genome average nucleotide identity (ANI) has emerged as a robust method for assessing species boundaries and estimating the genetic relatedness between two genomes. Ample data have shown whole genome ANI is correlated with the traditional microbiological concept of DNA–DNA hybridization relatedness for defining species [47,48]. In bacteria, an ANI value of 95% to 96% has been generally accepted for circumscribing species [47,49]. In 2019, Bergey’s Manual suggested the beginning of the transition from 16S rRNA gene sequence-based to whole genome-based taxonomy in bacteria (Figure 1) [20].
The first complete genome of phytoplasma (onion yellows phytoplasma mild strain (OY-M) was published in 2004 [6] (data deposited into GenBank in 2003). So far, 47 phytoplasma genomes (35 draft and 12 complete) have been sequenced involving 13 groups and 29 subgroups (Figure 2 and Table 1, [6,7,14,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81]). The size of the complete ge-nomes ranges from 576 to 960 Kb. As shown in Figure 2, 12 phytoplasma genomes were published in year 2021 alone. Even so, genomes of only a small proportion of phytoplasmas have been sequenced compared to nearly one thousand known phytoplasma strains (covering 37 groups and more than 150 subgroups, see Section 3). No doubt, the ever-increasing phytoplasma genome sequence data will serve as an excellent and more comprehensive frame-work for phytoplasma taxonomy.
The 2004 guidelines have served nearly 20 years. To date, approximately 50 ‘Candidatus phytoplasma’ species have been formally named [15]. While most of species were delineated based on 16S rRNA gene identity scores (rules a and b), several species were recognized as they each represent an ecologically distinct population (rule c). In contrast, the delineation of ‘Ca. Phytoplasma tritici’ and ‘Ca. Phytoplasma sacchari’ exploited whole genome information in addition to unique ecological properties [14,82]. Phytoplasma species naming based on whole genome information goes beyond the 2004 guidelines. On the other hand, in recent years, after pairwise comparison of the 16S rRNA gene sequence identity score and the corresponding whole genome ANI score, bacterial taxonomists revised twice the 16S rRNA gene sequence identity threshold value for delineating new bacterial species: changing from 97% to 98.7% and then to the current 98.65% [49,83,84]. To embrace these new developments and to incorporate “whole genome” concept to phytoplasma taxonomy, Bertaccini et al. recently revised guidelines for naming Ca. Phytoplasma species (referred to as “2022 guidelines” thereafter) [15].

2.2. The Newly Revised 2022 Guidelines and Proposed Amendments

The major revisions in the 2022 guidelines for naming a new Candidatus Phytoplasma species include (1) the length of 16S rRNA gene sequence was extended from >1200 bp to >1500 bp (full length or nearly full length of 16S rRNA gene); (2) the threshold of 16S rRNA gene identity was changed from 97.5% to 98.65%; (3) a whole genome ANI criterium was proposed; and (4) if a strain shares >98.65% identity in 16S rRNA gene sequence and >95% genome ANI with previously established species, a MLSA approach can be used to demarcate a new species. Criteria for five housekeeping genes were proposed. However, in our opinion, some provisions in the 2022 guidelines lack clarity and precision. Below, we compare the newly revised 2022 guidelines with the original 2004 guidelines and discuss the issues that require clarification and amendment (Table 2).
In rule (a) of 2004 IRPCM guidelines, the ‘related strain’ was clearly defined. That is, the strain from which this sequence was obtained should be named the ‘reference strain’ and not the ‘type strain’. Strains in which even minimal differences in the 16S rRNA gene sequence from the reference strain are detected do not ‘belong’ to the Candidatus species but are ‘related’ to it. However, in the following statement of the 2022 revised guidelines, “Strains sharing >98.65% sequence identity when compared with the reference strain are considered members of the respective ‘Ca. Phytoplasma’ species. Strains showing identity <98.65% to the reference strain, but >98.65% with other strains of the same ‘Ca. Phytoplasma’ species should be considered as related to this ‘Ca. Phytoplasma’ species.”.
The term ‘member strain’ was not very well conceived and could lead to erroneous assignment of a single given strain to more than one species. For example, alder yellows phytoplasma strain ALY (AY197646) shares 99.7%, 99.35%, and 98.98% identity with that of ‘Ca. Phytoplasma ulmi’ reference strain EY1 (AY197655), ‘Ca. Phytoplasma rubi’ reference strain RuS (AY197648), and ‘Ca. Phytoplasma ziziphi’ reference strain JWB-G1 (AB052876), respectively, in their 16S rRNA gene sequences. According to the 2022 guidelines, ALY would be a member strain of ‘Ca. Phytoplasma ulmi’, ‘Ca. Phytoplasma rubi’, and ‘Ca. Phytoplasma ziziphi’ simultaneously. Likewise, plum leptonecrosis phytoplasma strain LNp (JQ868450) shares 99.93%, 98.86%, and 98.66% identity with ‘Ca. Phytoplasma prunorum’ reference strain ESFY-G1 (AJ542544), ‘Ca. Phytoplasma pyri’ reference strain PD1 (AJ542543), and ‘Ca. Phytoplasma mali’ reference strain AP15 (AJ542541), respectively, in their 16S rRNA gene sequences. Therefore, according to the 2022 guidelines, LNp would be a member strain of ‘Ca. Phytoplasma prunorum’, ‘Ca. Phytoplasma pyri’, and ‘Ca. Phytoplasma mali’ at the same time.
Conceptually and practically, a strain can be related to more than one species, but cannot be a member of more than one species simultaneously. Therefore, the term “member strain” should be abolished and the term “related strain” as coined in the 2004 original guidelines should be restored.
Rule (c) of the 2004 IRPCM guidelines emphasized the importance of ‘ecological population’; it reflects the ecological nature of bacteria. No matter how we “modernize” our standards or framework, this provision should be retained. In other words, in addition to 16S rRNA sequence identity- and whole ANI-based criteria for demarcating Candidatus Phytoplasma species, ecological feature/property-based delineation criteria should be in place as well.
In the 2022 revised guidelines, a proposal was made to allow naming new phytoplasma species based on two out of five housekeeping genes (groEL, tuf, rp, secA and secY) with individual criteria if a strain shares >98.65% identity in 16S rRNA gene sequence and >95% genome ANI with previously reported species. So far, about 20,000 phytoplasma-related nucleotide sequences have been found in NCBI database, including approximately 8000 16S rRNA gene sequences, 1300 ribosomal protein-encoding gene sequences (rps3, rpl15, and rpl22, etc), 880 secY gene sequences, 570 tuf gene sequences, and sequences of other genes that are often used for differentiation of closely related phytoplasma strains such as vmp, Cpn60, amp, map, and SecA, etc. For a particular gene, such as secY gene, the sequence length of different phytoplasma strains deposited in GenBank varies considerably, ranging from 150 to 1400 bp. Excessively short sequences bear little value for comparative analysis. In addition, so far, there is no single gene other than the 16S rRNA gene has universal primers that are capable of amplifying all phytoplasma strains. Even widely used generic rp and secY primers can only amplify phytoplasmas that belong to certain 16Sr groups. Without universal primers and sufficient sequence data for a thorough comparative analysis of these five housekeeping genes, it is still difficult to establish objective criteria for MLSA-based species delineation. However, if the strain clearly represents ecologically separated populations, MLSA could be used to demonstrate significant molecular diversity in addition to fulfilling the unique vectorship and host specificity requirement. In such case, the MLSA assay should not be limited to these five genes proposed in 2022 guidelines. Evidence of significant molecular diversity from other housekeeping genes should be accepted as well.
Based on the above reasoning, we propose the following amendments to the 2022 revised guidelines. For clarity and consistency, the amendments are structured in the same fashion as the original 2004 IRCPM guidelines:
(a)
The ‘Ca. Phytoplasma’ species description should refer to a single, unique 16S rRNA gene sequence (full length or nearly full length, >1500 bp) or whole genome sequence with at least 60% coverage. The strain from which this sequence was obtained should be named the ‘reference strain’ and not the ‘type strain’. Strains in which even minimal differences in the 16S rRNA gene sequence from the reference strain are detected are referred as ‘related’ to the Candidatus species.
(b)
In general, a strain can be described as a novel ‘Ca. Phytoplasma’ species if its 16S rRNA gene sequence shares <98.65% identity or its whole genome shares an ANI score <95–96% to that of any previously described ‘Ca. Phytoplasma’ species.
(c)
There are, however, cases of phytoplasmas that share >98.65% identity of their 16S rRNA gene sequences or >95–96% ANI of their genomes, but clearly represent ecologically separated populations and, therefore, may deserve description as separate species. For such cases, description of two different species is recommended only when all three of the following conditions apply:
(i)
the two phytoplasmas are transmitted by different vectors;
(ii)
the two phytoplasmas have a different natural plant host (or, at least, their behavior is significantly different in the same plant host);
(iii)
there is evidence of significant molecular diversity, achieved by either hybridization to cloned DNA probes, serological reaction or multilocus sequence analysis (MLSA).
(d)
The rank of subspecies should not be used.
(e)
Due to strict international regulations, collection of micropropagation is no longer feasible or realistic; the gene clones of the reference strain should be deposited to the authorized scientific organizations in America, Europe, and Asia, etc. (to be determined by phytoplasma scientists).
(f)
Manuscripts that describe a novel ‘Ca. Phytoplasma’ species should preferably be submitted to the Int. J. Syst. Evol. Microbiol (IJSEM).
(g)
The abbreviation for Candidatus is Ca. (e.g., ‘Ca. Phytoplasma japonicum’ stands for ‘Candidatus Phytoplasma japonicum’).
In summary, a phytoplasma may be recognized as a novel ‘Ca. Phytoplasma’ species if it meets one of the following three criteria: sharing <98.65% 16S rRNA gene sequence identity, or sharing <95–96% genome-wide ANI or representing an ecologically separated population. Fulfillment of the rule (c) shall be demonstrated by vector specificity, unique host or host behavior, and molecular divergence.

3. Phytoplasma Classification: 16Sr Group/Subgroup Classification System Based on Collective RFLP Profiles

Classification is the systematic and orderly arrangement of organisms into groups or categories according to established criteria. Different from taxonomic nomenclature system, a classification scheme is often designed to meet practical needs, emphasizing less academic significance. Therefore, different scientists may classify the same organism differently [24]. Phytoplasma classification also has followed this principle. Phenotypic approaches such as symptomology, vectorship, and serology were employed to classify phytoplasmas in early days, but this has proved not suitable or practical [85,86] as in many cases the same phytoplasma strain may induce different symptoms in different hosts, and different phytoplasma strains may share a common vector or cause diseases exhibiting similar symptoms [87]. Until the 1990s, the 16Sr group/subgroup classification scheme was established based on RFLP profiles of PCR amplified F2nR2 fragment of the 16S rRNA gene [11,35,88,89]. This classification system is most widely adopted by phytoplasma researchers so far [90,91,92,93].
The RFLP-based phytoplasma classification scheme exploits a high-resolution subset of the 16S rRNA gene characteristics, namely, the recognition sites of 17 restriction enzymes, to differentiate diverse phytoplasmas [11,87]. The 16Sr groups delineated with this RFLP classification scheme are consistent with the 16S rRNA gene phylogenetic clades. More advantageously, by distinguishing subtle pattern differences, this RFLP analysis-based scheme is able to identify and distinguish different subgroup lineages within any given group [13,88,94,95]. Operationally, traditional RFLP analysis requires actual enzymatic gel electrophoresis and visual comparisons of various banded patterns. It is inconvenient, and few people are willing to do that anymore. The current virtual RFLP analysis approach is operated based on DNA sequences but retains the principles and criteria of the original phytoplasma classification scheme. Using accurate sequence data, the virtual gel patterns generated by computer simulated RFLP analysis can faithfully duplicate the classical and authoritative patterns established by conventional RFLP analysis. The new pattern types derived from virtual RFLP analysis have also been confirmed by actual enzymatic gel electrophoresis [94]. Furthermore, based on the virtual RFLP analysis approach, the interactive online tool iPhyClassifier was constructed, enabling and facilitating database-guided phytoplasma classification and identification [13].
Some scientists might think that the RFLP approach is obsolete. The truth is RFLP analysis still plays an important role in the classification and differentiation of many unculturable and fastidious bacteria, and fungi [96,97,98]. Examples include classifications of genus Basidiobolus [97] and genus Vibrio [98]. In the past five years (2017 to present), around 15,000 papers have been published on the classification and differentiation of bacteria and fungi based on RFLP analysis, including nearly 1600 articles on phytoplasma classification and identification. Computer-simulated virtual RFLP analysis undoubtedly enhanced the applicability of the RFLP analysis-based classification.
Importantly, the 16Sr group/subgroup classification system complements ‘Candidatus Phytoplasma’ species affiliation assignment. A striking example is the aster yellows (AY) phytoplasma group, which contains hundreds of known strains around the globe. The current taxonomic system assigns all the AY strains as ‘Ca. Phytoplasma asteris’-related strains, which grossly masks the differences among the strains. On the other hand, the existing 16Sr group classification scheme can differentiate the AY strains into more than two dozen subgroups, each of which has its own unique RFLP profile. In addition, some subgroups are only (or predominantly) present in certain geological regions and associated with different ecological niches [93,98].
In addition, in certain cases, the current phytoplasma taxonomic system may even have difficulty to assign certain strains to the existing ‘Ca. Phytoplasma’ species. For example, a strain (KJ452548) in the elm yellows phytoplasma group shares 99.1–99.3% identities with ‘Ca. Phytoplasma ulmi’- and ‘Ca. Phytoplasma ziziphi’-related strains in their 16S rRNA gene sequences. So, what species should this strain be affiliated with, ‘Ca. Phytoplasma ulmi’ or ‘Ca. Phytoplasma ziziphi’? Well, the RFLP-based group/subgroup classification system can at least provide distinguishing RFLP markers to separate them and classify the strain into a new subgroup other than 16SrV-A and 16SrV-B. This example strongly demonstrates that the group/subgroup classification system effectively avoids the ambiguity caused by the term, ‘Candidatus Phytoplasma sp.’-related strain, and helps diagnosticians and regulatory agencies distinguish closely-related phytoplasma strains.
In 2007, based on the virtual RFLP analysis of all 16S rRNA gene sequences available at the time (F2nR2 fragment of about 1250 bp), the number of phytoplasma classification groups was expanded from 19 to 28 (16SrXIX-16SrXXVIII), and some potentially new species were proposed with suggested reference strains (Table 3). In the present review, groups/subgroups corresponding to Candidatus Phytoplasma species, especially the newly named species are updated (Table 3 [99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137]). Two new groups (16SrXXXVIII and 16SrXXXIX) are established based on the criterium which requires the collective F2nR2 RFLP pattern of any new group representative has a similarity coefficient of <0.85 with that of all previously recognized 16Sr groups [94] (Supplementary Table S1). The reference strains of ‘Ca. Phytoplasma noviguineense’ and ‘Ca. Phytoplasma dypsidis’ were designated as representative strain of 16SrXXXVIII-A (LC228755) and 16SrXXXIX-A (MT536195), respectively.
Currently, there are a total of 37 groups and 48 named Candidates phytoplasma species (Table 3). Each group should contain at least one Candidatus species [138]. As shown in Table 3, nearly ten novel groups have been identified since 2007 (16SrXXIX-16SrXXXIX). However, it is noteworthy that no new phytoplasmas have been identified in groups 16SrXXIII-16SrXXVIII during the past 15 years. This suggests that the phytoplasmas belonging to these groups may be rare or the sequences representing these groups contain errors. In addition, we also noted that several pairs of strains share high sequence identity, but very low RFLP similarity coefficients. Such discrepancy might be caused by indels or sequencing errors that occurred within restriction enzyme recognition sites.

4. Phytoplasma Identification: Detection, Diagnostics and Characterization

The early identification and diagnosis of phytoplasmas and phytoplasmal diseases are vital for the formulation and implementation of rapid control measures. This not only thwarts the further spread of disease and reduce direct economic losses from plant death/damage, but also prevents delays and restrictions on the import and export of plant materials. Plants infected by phytoplasmas often exhibit remarkable symptoms. These symptoms include virescence (flower petals turning green), phyllody (leafy flowers), cauliflower-like inflorescence (repetitive initiation of inflorescence meristems), and witches’-broom (excessive shoot proliferation) [139,140]. In addition to these characteristic symptoms, phytoplasma infection can also induce some general symptoms seen in diseases caused by various other plant pathogens. Such general symptoms include leaf discoloration (such as purple leaves and leaf yellowing), little leaf, stem fasciation, and stunting [139,140,141]. Furthermore, asymptomatic phytoplasma infections were reported as well [142].
As phytoplasmas cannot be cultured in vitro, the routine culture-dependent metrics and characteristics for bacterial identification (morphological observation, biochemical assay, serotyping and antibiotic inhibition/resistance pattern assessment) cannot be employed. Phytoplasma detection and characterization heavily rely on the molecular diagnostic techniques. With the rapid development of molecular diagnostic techniques, a variety of fast, sensitive, and cost effective phytoplasma detection methods have emerged, ranging from PCR, nested PCR, real time PCR, droplet digital PCR (ddPCR), and loop-mediated isothermal amplification (LAMP) to CRISPR-based detection methods. These methods are devised based on highly conserved gene sequences of phytoplasmas, namely 16S rRNA gene, rp gene, SecY gene and tuf gene, etc. [11,143,144,145,146,147,148,149].
Currently, the most widely adopted procedure for the phytoplasma identification and further classification includes the following steps: (i) PCR or nested PCR amplification of phytoplasma DNA using universal primers of 16S rRNA gene, for example, P1, P7, P1A, P7A, 16S-SR, 16RF2n, and R16R2 [103,144,150,151]; (ii) Sequencing of PCR amplicons (direct sequencing or sequencing after amplicon cloning); and (iii) Sequence analysis using iPhyClassifier, classifying the phytoplasma strain under study to existing 16Sr group/subgroup and assigning (relating) the strain to previously named Candidatus Phytoplasma species. Results from the last step also offer opportunities for establishing new groups/subgroups and discovering novel Candidatus Phytoplasma species.
MLSA-based classification schemes have been established in many bacteria, but not yet implemented in non-culturable phytoplasmas (see Section 2.2 for reasons). However, this does not affect MLSA as a very effective method for phytoplasma diversity studies and fine differentiation of closely related phytoplasmas. For example, MLSA-based approach revealed the genetic diversity of apple proliferation phytoplasmas [152]; in addition, 16S rRNA, rp, and secY genes based MLSA characterization also indicated azalea little leaf phytoplasmas represented a distinct lineage within 16SrI group [153].

5. Challenges and Perspectives

Phytoplasma genome sequence information is essential to further advancing phytoplasma taxonomy. Since axenic phytoplasma culture is unattainable, DNA samples for phytoplasma genome sequencing are usually prepared from infected plants. As host DNA accounts for an overwhelming majority in the genomic DNA preparations, there is a risk of host DNA contamination in the process of genome assembly and mapping. A careful assessment of genome coverage statistics is vital as genome information-based delineation of new Candidatus Phytoplasma species solely relies on the accuracy of the genome sequences. Genome coverage statistics can indicate not only the contamination, but also the completeness of the genome. For those incomplete (draft) genomes, a minimum of 60% coverage has been suggested for microbial species delineation [154].
In addition, according to rule (e) in the 2004 Guidelines, the reference strain (maintained in micro-propagation if available) should be sent to and deposited to the scientific community or authorized organizations by the authors of the Candidatus species description paper. Implementation of this rule has become increasingly difficult or even no longer feasible due to strict international regulations. A good alternative might be to submit the gene clones of the reference strain to the scientific committee or the authorized organizations in different countries in America, Europe, and Asia, etc. (to be determined by phytoplasma scientists).
Currently, Groups 16SrV, 16SrX, and 16SrXII each contains four or five Candidatus Phytoplasma species (Table 3). There are occasions where an unknown strain shares an identical (or nearly identical) 16S rDNA sequence identity score with the reference strains of more than one Candidatus species within a given 16Sr group. Such a scenario makes it difficult to determine with which Candidatus species the unknown strain should be affiliated. The decision-making process is even tough if any involved Candidatus species has quarantine implications. The 2022 Guidelines revised the 16S rRNA gene sequence identity threshold value for demarcating phytoplasma species from 97.5% to 98.65%. The new threshold will likely result in an increase in the number of new Candidatus species within certain 16Sr groups, which will make the matters worse. Subgroup classification and multilocus strain typing may help alleviate the problem.

6. Conclusions

This article reviews the latest progress in phytoplasma taxonomy from three aspects: nomenclature, classification, and identification. ‘Candidatus Phytoplasma’ species/nomenclature system and group/subgroup classification system are two parallel systems and serve different purposes. The nomenclature system focuses more on naming new species based on one of the three criteria: 16S rRNA gene sequence identity (<98.65%), whole genome ANI (<95–96%), or representing ecologically separated populations. Currently, 48 Candidatus Phytoplasma species have been named. The group/subgroup classification system is based on collective RFLP profiles of the F2nR2 region of 16S rRNA gene. The genetically diverse phytoplasmas have been classified into 37 groups and more than 150 subgroups.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/biology11081119/s1, Table S1: Identification of two new 16Sr groups/subgroups (16SrXXXVIII-A and 16SrXXXIX-A) based on similarity coefficients derived from virtual RFLP analysis of 16S rRNA genes.

Author Contributions

Conceptualization, W.W. and Y.Z.; writing—original draft preparation, W.W. and Y.Z.; writing—review and editing; funding acquisition, W.W. and Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the US Department of Agriculture, Agricultural Research Service (Project number 8042-22000-320-00D).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Severin, H.H. Infection of perennial delphiniums by California-aster-yellows virus. Hilgardia 1942, 14, 411–440. [Google Scholar] [CrossRef] [Green Version]
  2. Black, L.M. Transmission of plant viruses by cicadellids. In Advances in Virus Research; Academic Press: Cambridge, MA, USA, 1953; Volume 1, pp. 69–89. [Google Scholar]
  3. Doi, Y.; Teranaka, M.; Yora, K.; Asuyama, H. Mycoplasma- or PLT group-like microorganisms found in the phloem elements of plants infected with mulberry dwarf, potato witches’ broom, aster yellows or paulownia witches’ broom. Ann. Phytopathol. Soc. Jpn. 1967, 33, 259–266. [Google Scholar] [CrossRef]
  4. Lim, P.O.; Sears, B.B. 16S rRNA sequence indicates that plant-pathogenic mycoplasmalike organisms are evolutionarily distinct from animal mycoplasmas. J. Bacteriol. 1989, 171, 5901–5906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. International Committee on Systematic Bacteriology Subcommittee on the Taxonomy of Mollicutes. Minutes of the Interim Meetings, 1 and 2 August, 1992, Ames, Iowa. Int. J. Syst. Bacteriol. 1993, 43, 394–397. [Google Scholar] [CrossRef] [Green Version]
  6. Oshima, K.; Kakizawa, S.; Nishigawa, H.; Jung, H.Y.; Wei, W.; Suzuki, S.; Arashida, R.; Nakata, D.; Miyata, S.; Ugaki, M.; et al. Reductive evolution suggested from the complete genome sequence of a plant-pathogenic phytoplasma. Nat. Genet. 2004, 36, 27–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Bai, X.; Zhang, J.; Ewing, A.; Miller, S.A.; Jancso Radek, A.; Shevchenko, D.V.; Tsukerman, K.; Walunas, T.; Lapidus, A.; Campbell, J.W.; et al. Living with genome instability: The adaptation of phytoplasmas to diverse environments of their insect and plant hosts. J. Bacteriol. 2006, 188, 3682–3696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Tan, Y.; Li, Q.; Zhao, Y.; Wei, H.; Wang, J.; Baker, C.J.; Liu, Q.; Wei, W. Integration of metabolomics and existing omics data reveals new insights into phytoplasma-induced metabolic reprogramming in host plants. PLoS ONE 2021, 16, e0246203. [Google Scholar] [CrossRef] [PubMed]
  9. Kirkpatrick, B.C.; Stenger, D.C.; Morris, T.J.; Purcell, A.H. Cloning and detection of DNA from a nonculturable plant pathogenic mycoplasma-like organism. Science 1987, 238, 197–200. [Google Scholar] [CrossRef] [PubMed]
  10. Murray, R.G.E.; Schleifer, K.H. Taxonomic notes: A proposal for recording the properties of putative taxa of procaryotes. Int. J. Syst. Evol. Microbiol. 1994, 44, 174–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Lee, I.M.; Bertaccini, A.; Vibio, M.; Gundersen, D.E. Detection of multiple phytoplasmas in perennial fruit trees with decline symptoms in Italy. Phytopathology 1995, 85, 728–735. [Google Scholar] [CrossRef]
  12. IRPCM Phytoplasma/Spiroplasma Working Team–Phytoplasma Taxonomy Group. ‘Candidatus Phytoplasma’, a taxon for the wall-less, non-helical prokaryotes that colonize plant phloem and insects. Int. J. Syst. Evol. Microbiol. 2004, 54, 1243–1255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zhao, Y.; Wei, W.; Lee, M.; Shao, J.; Suo, X.; Davis, R.E. Construction of an interactive online phytoplasma classification tool, iPhyClassifier, and its application in analysis of the peach X-disease phytoplasma group (16SrIII). Int. J. Syst. Evol. Microbiol. 2009, 59 Pt 10, 2582. [Google Scholar] [CrossRef] [PubMed]
  14. Kirdat, K.; Tiwarekar, B.; Thorat, V.; Sathe, S.; Shouche, Y.; Yadav, A. ‘Candidatus Phytoplasma sacchari’, a novel taxon-associated with Sugarcane Grassy Shoot (SCGS) disease. Int. J. Syst. Evol. Microbiol. 2021, 71, 004591. [Google Scholar] [CrossRef] [PubMed]
  15. Bertaccini, A.; Arocha-Rosete, Y.; Contaldo, N.; Duduk, B.; Fiore, N.; Montano, H.G.; Kube, M.; Kuo, C.H.; Martini, M.; Oshima, K.; et al. Revision of the ‘Candidatus Phytoplasma’species description guidelines. Int. J. Syst. Evol. Microbiol. 2022, 72, 005353. [Google Scholar] [CrossRef]
  16. Schildkraut, C.L.; Marmur, J.; Doty, P. The formation of hybrid DNA molecules and their use in studies of DNA homologies. J. Mol. Biol. 1961, 3, 595–617, IN15–IN16. [Google Scholar] [CrossRef]
  17. Woese, C.R.; Fox, G.E. Phylogenetic structure of the prokaryotic domain: The primary kingdoms. Proc. Natl. Acad. Sci. USA 1977, 74, 5088–5090. [Google Scholar] [CrossRef] [Green Version]
  18. Wayne, L.G.; Brenner, D.J.; Colwell, R.R.; Grimont, P.A.D.; Kandler, O.; Krichevsky, M.I.; Moore, L.H.; Moore, W.E.C.; Murray, R.G.E.; Stackebrandt, E.S.M.P.; et al. Report of the ad hoc committee on reconciliation of approaches to bacterial systematics. Int. J. Syst. Bacteriol. 1987, 37, 463–464. [Google Scholar] [CrossRef] [Green Version]
  19. Hillis, D.M.; Dixon, M.T. Ribosomal DNA: Molecular evolution and phylogenetic inference. Q Rev. Biol. 1991, 66, 411–453. [Google Scholar] [CrossRef] [PubMed]
  20. Hugenholtz, P.; Chuvochina, M.; Oren, A.; Parks, D.H.; Soo, R.M. Prokaryotic taxonomy and nomenclature in the age of big sequence data. ISME J. 2021, 15, 1879–1892. [Google Scholar] [CrossRef] [PubMed]
  21. Stackebrandt, E.; Frederiksen, W.; Garrity, G.M.; Grimont, P.A.; Kämpfer, P.; Maiden, M.C.; Nesme, X.; Rosselló-Mora, R.; Swings, J.; Trüper, H.G.; et al. Report of the ad hoc committee for the re-evaluation of the species definition in bacteriology. Int. J. Syst. Evol. Microbiol. 2002, 52, 1043–1047. [Google Scholar]
  22. Auch, A.F.; von Jan, M.; Klenk, H.P.; Göker, M. Digital DNA-DNA hybridization for microbial species delineation by means of genome-to-genome sequence comparison. Stand. Genomic. Sci. 2010, 2, 117–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Johansson, K.E.; Pettersson, B. Taxonomy of mollicutes. In Molecular Biology and Pathogenicity of Mycoplasmas (1–29); Springer: Boston, MA, USA, 2002. [Google Scholar]
  24. Baron, E.J. Classification. In Medical Microbiology, 4th ed.; Baron, S., Ed.; University of Texas Medical Branch at Galveston: Galveston, TX, USA, 1996; Chapter 3. Available online: https://www.ncbi.nlm.nih.gov/books/NBK8406/ (accessed on 6 June 2022).
  25. Kämpfer, P.; Glaeser, S.P. Prokaryotic taxonomy in the sequencing era–the polyphasic approach revisited. Environ. Microbiol. 2012, 14, 291–317. [Google Scholar] [CrossRef] [PubMed]
  26. Rosselló-Mora, R.; Amann, R. The species concept for prokaryotes. FEMS Microbiol. Rev. 2001, 25, 39–67. [Google Scholar] [CrossRef]
  27. Chiykowski, L.N. Clover phyllody virus in Canada and its transmission. Can. J. Bot. 1962, 40, 397–404. [Google Scholar] [CrossRef]
  28. Freitag, J.H. Interaction and mutual suppression among three strains of aster yellows virus. Virology 1964, 24, 401–413. [Google Scholar] [CrossRef]
  29. Granados, R.R.; Chapman, R.K. Identification of some new aster yellows virus strains and their transmission by aster leafhopper Macrosteles fascifrons. Phytopathology 1968, 58, 1685. [Google Scholar]
  30. Chiykowski, L.N.; Sinha, R.C. Differentiation of MLO diseases by means of symptomatology and vector transmission. In Recent advances in mycoplasmology. In Proceedings of the 7th congress of the International Organization for Mycoplasmology, Baden near Vienna, Austria, 2–9 June 1988; Gustav Fischer Verlag: Baden near Vienna, Austria, 1990; pp. 280–287. [Google Scholar]
  31. McCoy, R.E.; Caudwell, A.; Chang, C.G.; Chen, T.A.; Chiykowski, L.N.; Cousin, M.T.; De Leeuw, G.D.; Golino, D.A.; Hacke, K.J.; Kirkpatrick, B.C.; et al. Mycoplasmalike organisms. Mycoplasmas 1989, 5, 545–568. [Google Scholar]
  32. Stackebrandt, E.; GOEBEL, B.M. Taxonomic note: A place for DNA-DNA reassociation and 16S rRNA sequence analysis in the present species definition in bacteriology. Int. J. Syst. Evol. Microbiol. 1994, 44, 846–849. [Google Scholar] [CrossRef] [Green Version]
  33. Stephen, J.R.; McCaig, A.E.; Smith, Z.; Prosser, J.I.; Embley, T.M. Molecular diversity of soil and marine 16S rRNA gene sequences related to beta-subgroup ammonia-oxidizing bacteria. Appl. Environ. Microbiol. 1996, 62, 4147–4154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Drancourt, M.; Bollet, C.; Carlioz, A.; Martelin, R.; Gayral, J.P.; Raoult, D. 16S ribosomal DNA sequence analysis of a large collection of environmental and clinical unidentifiable bacterial isolates. J. Clin. Microbiol. 2000, 38, 3623–3630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Moore, E.R.; Krüger, A.S.; Hauben, L.; Seal, S.E.; De Baere, R.; De Wachter, R.; Timmis, K.N.; Swings, J. 16S rRNA gene sequence analyses and inter-and intrageneric relationships of Xanthomonas species and Stenotrophomonas maltophilia. FEMS Microbiol. Lett. 1997, 151, 145–153. [Google Scholar] [CrossRef] [PubMed]
  36. Woo, P.C.; Lau, S.K.; Teng, J.L.; Tse, H.; Yuen, K.Y. Then and now: Use of 16S rDNA gene sequencing for bacterial identification and discovery of novel bacteria in clinical microbiology laboratories. Clin. Microbiol. Infect. 2008, 14, 908–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Oren, A.; Garrity, G.M. Valid publication of the names of forty-two phyla of prokaryotes. Int. J. Syst. Evol. Microbiol. 2021, 71, 005056. [Google Scholar] [CrossRef]
  38. Murray, R.G.E.; Sneath, P.H.A.; Mair, N.S.; Sharpe, M.E. Kingdom Procaryotae. Bergey’s Man. Syst. Bacteriol. 1984, 1, 34–36. [Google Scholar]
  39. Gasparich, G.E.; Bertaccini, A.; Zhao, Y. Candidatus Phytoplasma. In Bergey’s Manual of Systematics of Archaea and Bacteria; Trujillo, M.E., Dedysh, S., DeVos, P., Hedlund, B., Kämpfer, P., Rainey, F.A., Whitman, W.B., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2020. [Google Scholar] [CrossRef]
  40. Oren, A. A plea for linguistic accuracy–also for Candidatus taxa. Int. J. Syst. Evol. Microbiol. 2017, 67, 1085–1094. [Google Scholar] [CrossRef] [PubMed]
  41. Zreik, L.; Carle, P.; Bove, J.M.; Garnier, M. Characterization of the mycoplasmalike organism associated with witches’-broom disease of lime and proposition of a “Candidatus” taxon for the organism, “Candidatus Phytoplasma aurantifolia”. Int. J. Syst. Bacteriol. 1995, 45, 449–453. [Google Scholar] [CrossRef]
  42. Yarza, P.; Yilmaz, P.; Pruesse, E.; Glöckner, F.O.; Ludwig, W.; Schleifer, K.H.; Whitman, W.B.; Euzéby, J.; Amann, R.; Rosselló-Móra, R. Uniting the classification of cultured and uncultured bacteria and archaea using 16S rRNA gene sequences. Nat. Rev. Microbiol. 2014, 12, 635–645. [Google Scholar] [CrossRef] [PubMed]
  43. Fox, G.E.; Wisotzkey, J.D.; Jurtshuk, P., Jr. How close is close: 16S rRNA sequence identity may not be sufficient to guarantee species identity. Int. J. Syst. Evol. Microbiol. 1992, 42, 166–170. [Google Scholar] [CrossRef] [Green Version]
  44. Lu, Z.; Zhang, W. Comparative phylogenies of ribosomal proteins and the 16S rRNA gene at higher ranks of the class Actinobacteria. Curr. Microbiol. 2012, 65, 1–6. [Google Scholar] [CrossRef]
  45. Maiden, M.C. Multilocus sequence typing of bacteria. Annu. Rev. Microbiol. 2006, 60, 561–588. [Google Scholar] [CrossRef]
  46. Dingle, K.E.; Colles, F.M.; Wareing, D.R.A.; Ure, R.; Fox, A.J.; Bolton, F.E.; Bootsma, H.J.; Willems, R.J.L.; Urwin, R.; Maiden, M.C.J. Multilocus sequence typing system for Campylobacter jejuni. J. Clin. Microbiol. 2001, 39, 14–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Richter, M.; Rosselló-Móra, R. Shifting the genomic gold standard for the prokaryotic species definition. Proc. Natl. Acad. Sci. USA 2009, 106, 19126–19131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Jain, C.; Rodriguez-R, L.M.; Phillippy, A.M.; Konstantinidis, K.T.; Aluru, S. High throughput ANI analysis of 90K prokaryotic genomes reveals clear species boundaries. Nat. Commun. 2018, 9, 5114. [Google Scholar] [CrossRef] [Green Version]
  49. Kim, M.; Oh, H.S.; Park, S.C.; Chun, J. Towards a taxonomic coherence between average nucleotide identity and 16S rRNA gene sequence similarity for species demarcation of prokaryotes. Int. J. Syst. Evol. Microbiol. 2014, 64 Pt 2, 346–351. [Google Scholar] [CrossRef] [PubMed]
  50. Kakizawa, S.; Makino, A.; Ishii, Y.; Tamaki, H.; Kamagata, Y. Draft genome sequence of “Candidatus Phytoplasma asteris” strain OY-V, an unculturable plant-pathogenic bacterium. Genome Announc. 2014, 2, e00944-14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Kirdat, K.; Tiwarekar, B.; Thorat, V.; Narawade, N.; Dhotre, D.; Sathe, S.; Shouche, Y.; Yadav, A. Draft genome sequences of two phytoplasma strains associated with sugarcane grassy shoot (SCGS) and bermuda grass white leaf (BGWL) diseases. Mol. Plant-Microbe Interact 2020, 33, 715–717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Chang, S.H.; Cho, S.T.; Chen, C.L.; Yang, J.Y.; Kuo, C.H. Draft genome sequence of a 16SrII-A subgroup phytoplasma associated with purple coneflower (Echinacea purpurea) witches’ broom disease in Taiwan. Genome Announc. 2015, 3, e01398-15. [Google Scholar] [CrossRef] [Green Version]
  53. Cui, W.; Quiroga, N.; Curkovic, S.T.; Zamorano, A.; Fiore, N. Detection and identification of 16SrXIII-F and a novel 16SrXIII phytoplasma subgroups associated with strawberry phyllody in Chile. Eur. J. Plant Pathol. 2019, 155, 1039–1046. [Google Scholar] [CrossRef]
  54. Huang, C.T.; Cho, S.T.; Lin, Y.C.; Tan, C.M.; Chiu, Y.C.; Yang, J.Y.; Kuo, C.H. Comparative Genome Analysis of ‘Candidatus Phytoplasma luffae’Reveals the Influential Roles of Potential Mobile Units in Phytoplasma Evolution. Front. Microbiol. 2022, 13, 773608. [Google Scholar] [CrossRef]
  55. Kube, M.; Schneider, B.; Kuhl, H.; Dandekar, T.; Heitmann, K.; Migdoll, A.M.; Reinhardt, R.; Seemüller, E. The linear chromosome of the plant-pathogenic mycoplasma ‘Candidatus Phytoplasma mali’. BMC Genet. 2008, 9, 306. [Google Scholar] [CrossRef] [Green Version]
  56. Fernández, F.D.; Conci, L.R. Genome characterization of ‘Candidatus Phytoplasma meliae’ (isolate ChTYXIII). bioRxiv 2021. [Google Scholar] [CrossRef]
  57. Fischer, A.; Santana-Cruz, I.; Wambua, L.; Olds, C.; Midega, C.; Dickinson, M.; Kawicha, P.; Khan, Z.; Masiga, D.; Jores, J.; et al. Draft genome sequence of “Candidatus Phytoplasma oryzae” strain Mbita1, the causative agent of Napier grass stunt disease in Kenya. Genome Announc. 2016, 4, e00297-16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Quaglino, F.; Kube, M.; Jawhari, M.; Abou-Jawdah, Y.; Siewert, C.; Choueiri, E.; Sobh, H.; Casati, P.; Tedeschi, R.; Lova, M.M.; et al. ‘Candidatus Phytoplasma phoenicium’associated with almond witches’-broom disease: From draft genome to genetic diversity among strain populations. BMC Microbiol. 2015, 15, 148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Cai, W.; Shao, J.; Zhao, Y.; Davis, R.E.; Costanzo, S. Draft genome sequence of ‘Candidatus Phytoplasma pini’-related strain MDPP: A resource for comparative genomics of gymnosperm-infecting phytoplasmas. Plant Dis. 2020, 104, 1009–1010. [Google Scholar] [CrossRef] [Green Version]
  60. Fernández, F.D.; Zübert, C.; Huettel, B.; Kube, M.; Conci, L.R. Draft Genome Sequence of “Candidatus Phytoplasma pruni” (X-disease group, subgroup 16SrIII-B) strain ChTDIII from Argentina. Microbiol. Resour. Announc. 2020, 9, e00792-20. [Google Scholar] [CrossRef] [PubMed]
  61. Lee, I.M.; Shao, J.; Bottner-Parker, K.D.; Gundersen-Rindal, D.E.; Zhao, Y.; Davis, R.E. Draft genome sequence of “Candidatus Phytoplasma pruni” strain CX, a plant-pathogenic bacterium. Genome Announc. 2015, 3, e01117-15. [Google Scholar] [CrossRef] [Green Version]
  62. Music, M.S.; Samarzija, I.; Hogenhout, S.A.; Haryono, M.; Cho, S.T.; Kuo, C.H. The genome of ‘Candidatus Phytoplasma solani’strain SA-1 is highly dynamic and prone to adopting foreign sequences. Syst. Appl. Microbiol. 2019, 42, 117–127. [Google Scholar] [CrossRef]
  63. Mitrović, J.; Siewert, C.; Duduk, B.; Hecht, J.; Mölling, K.; Broecker, F.; Beyerlein, P.; Büttner, C.; Bertaccini, A.; Kube, M. Generation and Analysis of Draft Sequences of ‘Stolbur’Phytoplasma from Multiple Displacement Amplification Templates. Microb. Physiol. 2014, 24, 1–11. [Google Scholar] [CrossRef]
  64. Cai, W.; Nunziata, S.O.; Srivastava, S.K.; Wilson, T.; Chambers, N.; Rivera, Y.; Nakhla, M.; Costanzo, S. Draft Genome Sequence Resource of AldY-WA1, a Phytoplasma Strain Associated with Alder Yellows of Alnus rubra in Washington, U.S.A. Plant Dis. 2022, 106, 1971–1973. [Google Scholar] [CrossRef]
  65. Chen, W.; Li, Y.; Wang, Q.; Wang, N.; Wu, Y. Comparative genome analysis of wheat blue dwarf phytoplasma, an obligate pathogen that causes wheat blue dwarf disease in China. PLoS ONE 2014, 9, e96436. [Google Scholar] [CrossRef]
  66. Wang, J.; Song, L.; Jiao, Q.; Yang, S.; Gao, R.; Lu, X.; Zhou, G. Comparative genome analysis of jujube witches’-broom Phytoplasma, an obligate pathogen that causes jujube witches’-broom disease. BMC Genom. 2018, 19, 689. [Google Scholar] [CrossRef] [Green Version]
  67. Pacifico, D.; Galetto, L.; Rashidi, M.; Abbà, S.; Palmano, S.; Firrao, G.; Bosco, D.; Marzachì, C. Decreasing global transcript levels over time suggest that phytoplasma cells enter stationary phase during plant and insect colonization. Appl. Environ. Microbiol. 2015, 81, 2591–2602. [Google Scholar] [CrossRef] [Green Version]
  68. Nijo, T.; Iwabuchi, N.; Tokuda, R.; Suzuki, T.; Matsumoto, O.; Miyazaki, A.; Maejima, K.; Oshima, K.; Namba, S.; Yamaji, Y. Enrichment of phytoplasma genome DNA through a methyl-CpG binding domain-mediated method for efficient genome sequencing. J. Gen. Plant Pathol. 2021, 87, 154–163. [Google Scholar] [CrossRef]
  69. Saccardo, F.; Martini, M.; Palmano, S.; Ermacora, P.; Scortichini, M.; Loi, N.; Firrao, G. Genome drafts of four phytoplasma strains of the ribosomal group 16SrIII. Microbiology 2012, 158, 2805–2814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Orlovskis, Z.; Canale, M.C.; Haryono, M.; Lopes, J.R.S.; Kuo, C.H.; Hogenhout, S.A. A few sequence polymorphisms among isolates of Maize bushy stunt phytoplasma associate with organ proliferation symptoms of infected maize plants. Ann. Bot. 2017, 119, 869–884. [Google Scholar] [CrossRef] [Green Version]
  71. Luo, L.; Zhang, X.; Meng, F.; Wang, Y.; Zhou, Y.; Liu, J. Draft Genome Sequences Resources of Mulberry Dwarf Phytoplasma Strain MDGZ-01 Associated with Mulberry Yellow Dwarf (MYD) Diseases. Plant Dis. 2022; online ahead of print. [Google Scholar] [CrossRef]
  72. Sparks, M.E.; Bottner-Parker, K.D.; Gundersen-Rindal, D.E.; Lee, I.M. Draft genome sequence of the New Jersey aster yellows strain of ‘Candidatus Phytoplasma asteris’. PLoS ONE 2018, 13, e0192379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Cho, S.T.; Lin, C.P.; Kuo, C.H. Genomic characterization of the periwinkle leaf yellowing (PLY) phytoplasmas in Tawan. Front. Microbiol. 2019, 10, 2194. [Google Scholar] [CrossRef] [PubMed]
  74. Town, J.R.; Wist, T.; Perez-Lopez, E.; Olivier, C.Y.; Dumonceaux, T.J. Genome sequence of a plant-pathogenic bacterium, “Candidatus Phytoplasma asteris” strain TW1. Microbiol. Resour. Announc. 2018, 7, e01109-18. [Google Scholar] [CrossRef] [Green Version]
  75. Cao, Y.; Sun, G.; Zhai, X.; Xu, P.; Ma, L.; Deng, M.; Zhao, Z.; Yang, H.; Dong, Y.; Shang, Z.; et al. Genomic insights into the fast growth of paulownias and the formation of Paulownia witches’ broom. Mol. Plant 2021, 14, 1668–1682. [Google Scholar] [CrossRef]
  76. Chung, W.C.; Chen, L.L.; Lo, W.S.; Lin, C.P.; Kuo, C.H. Comparative analysis of the peanut witches’-broom phytoplasma genome reveals horizontal transfer of potential mobile units and effectors. PLoS ONE 2013, 8, e62770. [Google Scholar] [CrossRef] [Green Version]
  77. Zhu, Y.; He, Y.; Zheng, Z.; Chen, J.; Wang, Z.; Zhou, G. Draft genome sequence of rice orange leaf phytoplasma from Guangdong, China. Genome Announc. 2017, 5, e00430-17. [Google Scholar] [CrossRef] [Green Version]
  78. Ranebennur, H.; Kirdat, K.; Tiwarekar, B.; Rawat, K.; Chalam, V.C.; Solanke, A.U.; Yadav, R.; Singh, K.; Sathe, S.; Yadav, A.; et al. Draft genome sequence of ‘Candidatus Phytoplasma australasia’, strain SS02 associated with sesame phyllody disease. 3 Biotech 2022, 12, 107. [Google Scholar] [CrossRef]
  79. Tran-Nguyen, L.T.; Kube, M.; Schneider, B.; Reinhardt, R.; Gibb, K.S. Comparative genome analysis of “Candidatus Phytoplasma australiense” (subgroup tuf-Australia I; rp-A) and “Ca. Phytoplasma asteris” strains OY-M and AY-WB. J. Bacteriol. 2008, 190, 3979–3991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Andersen, M.T.; Liefting, L.W.; Havukkala, I.; Beever, R.E. Comparison of the complete genome sequence of two closely related isolates of ‘Candidatus Phytoplasma australiense’reveals genome plasticity. BMC Genom. 2013, 14, 529. [Google Scholar] [CrossRef] [Green Version]
  81. Debonneville, C.; Mandelli, L.; Brodard, J.; Groux, R.; Roquis, D.; Schumpp, O. The Complete Genome of the “Flavescence Dorée” Phytoplasma Reveals Characteristics of Low Genome Plasticity. Biology 2022, 11, 953. [Google Scholar] [CrossRef]
  82. Zhao, Y.; Wei, W.; Davis, R.E.; Lee, M.; Bottner-Parker, K.D. The agent associated with blue dwarf disease in wheat represents a new phytoplasma taxon, ‘Candidatus Phytoplasma tritici’. Int. J. Syst. Evol. Microbiol. 2021, 71, 004604. [Google Scholar] [CrossRef]
  83. Stackebrandt, E. Taxonomic parameters revisited: Tarnished gold standards. Microbiol. Today 2006, 33, 152–155. [Google Scholar]
  84. Chun, J.; Oren, A.; Ventosa, A.; Christensen, H.; Arahal, D.R.; da Costa, M.S.; Rooney, A.P.; Yi, H.; Xu, X.W.; De Meyer, S.; et al. Proposed minimal standards for the use of genome data for the taxonomy of prokaryotes. Int. J. Syst. Evol. Microbiol. 2018, 68, 461–466. [Google Scholar] [CrossRef]
  85. Kirkpatrick, B.C. Strategies for characterizing plant pathogenic mycoplasma-like organisms and their effects on plants. Plant-Microbe Interact. USA 1989. [Google Scholar]
  86. Davis, R.E.; Lee, I.; Dally, E.L.; Dewitt, N.; Douglas, S.M. Cloned nucleic acid hybridization probes in detection and classification of mycoplasmalike organisms (MLOs). In Proceedings of the VII International Symposium on Virus Diseases of Ornamental Plants, Sanremo, Italy, 29 May–2 June 1988; Volume 234, pp. 115–122. [Google Scholar]
  87. Zhao, Y.; Wei, W.; Davis, R.E.; Lee, I.-M. Recent advances in 16S rRNA gene-based phytoplasma differ-entiation, classification and taxonomy. In Phytoplasmas: Genomes, Plant Hosts and Vector; Weintraub, P., Jones, P., Eds.; CABI Publishing: Wallingford, UK, 2010; pp. 64–92. [Google Scholar]
  88. Lee, I.M.; Gundersen-Rindal, D.E.; Davis, R.E.; Bartoszyk, I.M. Revised classification scheme of phytoplasmas based on RFLP analyses of 16S rRNA and ribosomal protein gene sequences. Int. J. Syst. Evol. Microbiol. 2010, 48, 1153–1169. [Google Scholar] [CrossRef] [Green Version]
  89. Marcone, C.; Ragozzino, A.; Seemüller, E. Detection of Bermuda grass white leaf disease in Italy and characterization of the associated phytoplasma by RFLP analysis. Plant Dis. 1997, 81, 862–866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Pérez-López, E.; Luna-Rodríguez, M.; Olivier, C.Y.; Dumonceaux, T.J. The underestimated diversity of phytoplasmas in Latin America. Int. J. Syst. Evol. Microbiol. 2016, 66, 492–513. [Google Scholar] [CrossRef] [PubMed]
  91. Fernandez, F.D.; Meneguzzi, N.G.; Guzman, F.A.; Kirschbaum, D.S.; Conci, V.C.; Nome, C.F.; Conci, L.R. Detection and identification of a novel 16SrXIII subgroup phytoplasma associated with strawberry red leaf disease in Argentina. Int. J. Syst. Evol. Microbiol. 2015, 65 Pt 8, 2741–2747. [Google Scholar] [CrossRef] [Green Version]
  92. Nejat, N.; Vadamalai, G. Diagnostic techniques for detection of phytoplasma diseases: Past and present. J. Plant Dis. Prot. 2013, 120, 16–25. [Google Scholar] [CrossRef]
  93. Lee, I.-M.; Davis, R.E.; Gundersen-Rindal, D.E. Phytoplasma: Phytopathogenic mollicutes. Annu. Rev. Microbiol. 2000, 54, 221–255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Wei, W.; Davis, R.E.; Lee, M.; Zhao, Y. Computer-simulated RFLP analysis of 16S rRNA genes: Identification of ten new phytoplasma groups. Int. J. Syst. Evol. Microbiol. 2007, 57, 1855–1867. [Google Scholar] [CrossRef] [PubMed]
  95. Wei, W.; Lee, M.; Davis, R.E.; Suo, X.; Zhao, Y. Automated RFLP pattern comparison and similarity coefficient calculation for rapid delineation of new and distinct phytoplasma 16Sr subgroup lineages. Int. J. Syst. Evol. Microbiol. 2008, 58, 2368–2377. [Google Scholar] [CrossRef] [PubMed]
  96. Edel, V. Use of PCR and RFLP in fungal systematics. In Chemical Fungal Taxonomy; CRC Press: Boca Raton, FL, USA, 2020; pp. 51–76. [Google Scholar]
  97. Claussen, M.; Schmidt, S. Differentiation of Basidiobolus s isolates: RFLP of a diagnostic PCR amplicon matches sequence-based classification and growth temperature preferences. J. Fungi 2021, 7, 110. [Google Scholar] [CrossRef]
  98. Silvester, R.; Alexander, D.; Antony, A.C.; Hatha, M. GroEL PCR-RFLP–an efficient tool to discriminate closely related pathogenic Vibrio species. Microb. Pathog. 2017, 105, 196–200. [Google Scholar] [CrossRef] [PubMed]
  99. Lee, I.M.; Gundersen-Rindal, D.E.; Davis, R.E.; Bottner, K.D.; Marcone, C.; Seemüller, E. ‘Candidatus Phytoplasma asteris’, a novel phytoplasma taxon associated with aster yellows and related diseases. Int. J. Syst. Evol. Microbiol. 2004, 54, 1037–1048. [Google Scholar] [CrossRef] [Green Version]
  100. Arocha, Y.; Antesana, O.; Montellano, E.; Franco, P.; Plata, G.; Jones, P. ‘Candidatus Phytoplasma lycopersici’, a phytoplasma associated with ‘hoja de perejil’ disease in Bolivia. Int. J. Syst. Evol. Microbiol. 2007, 57, 1704–1710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. White, D.T.; Blackall, L.L.; Scott, P.T.; Walsh, K.B. Phylogenetic positions of phytoplasmas associated with dieback, yellow crinkle and mosaic diseases of papaya, and their proposed inclusion in ‘Candidatus Phytoplasma australiense’ and a new taxon, ‘Candidatus Phytoplasma australasia’. Int. J. Syst. Evol. Microbiol. 1998, 48, 941–951. [Google Scholar] [CrossRef]
  102. Davis, R.E.; Zhao, Y.; Dally, E.L.; Lee, M.; Jomantiene, R.; Douglas, S.M. ‘Candidatus Phytoplasma pruni’, a novel taxon associated with X-disease of stone fruits, Prunus s: Multilocus characterization based on 16S rRNA, secY, and ribosomal protein genes. Int. J. Syst. Evol. Microbiol. 2013, 63 Pt 2, 766–776. [Google Scholar] [CrossRef] [PubMed]
  103. Lee, M.; Martini, M.; Marcone, C.; Zhu, S.F. Classification of phytoplasma strains in the elm yellows group (16SrV) and proposal of ‘Candidatus Phytoplasma ulmi’ for the phytoplasma associated with elm yellows. Int. J. Syst. Evol. Microbiol. 2004, 54, 337–347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Jung, H.Y.; Sawayanagi, T.; Kakizawa, S.; Nishigawa, H.; Wei, W.; Oshima, K.; Miyata, S.I.; Ugaki, M.; Hibi, T.; Namba, S. ‘Candidatus Phytoplasma ziziphi’, a novel phytoplasma taxon associated with jujube witches’-broom disease. Int. J. Syst. Evol. Microbiol. 2003, 53, 1037–1041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Malembic-Maher, S.; Salar, P.; Filippin, L.; Carle, P.; Angelini, E.; Foissac, X. Genetic diversity of European phytoplasmas of the 16SrV taxonomic group and proposal of ‘Candidatus Phytoplasma rubi’. Int. J. Syst. Evol. Microbiol. 2011, 61, 2129–2134. [Google Scholar] [CrossRef] [PubMed]
  106. Win, N.K.K.; Lee, S.Y.; Bertaccini, A.; Namba, S.; Jung, H.Y. ‘Candidatus Phytoplasma balanitae’ associated with witches’ broom disease of Balanites triflora. Int. J. Syst. Evol. Microbiol. 2013, 63 Pt 2, 636–640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Hiruki, C.; Wang, K. Clover proliferation phytoplasma: ‘Candidatus Phytoplasma trifolii’. Int. J. Syst. Evol. Microbiol. 2004, 54, 1349–1353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Davis, R.E.; Zhao, Y.; Dally, E.L.; Jomantiene, R.; Lee, M.; Wei, W.; Kitajima, E.W. ‘Candidatus Phytoplasma sudamericanum’, a novel taxon, and strain PassWB-Br4, a new subgroup 16SrIII-V phytoplasma, from diseased passion fruit (Passiflora edulis f. flavicarpa Deg.). Int. J. Syst. Evol. Microbiol. 2012, 62 Pt 4, 984–989. [Google Scholar] [CrossRef] [PubMed]
  109. Griffiths, H.M.; Sinclair, W.A.; Smart, C.D.; Davis, R.E. The phytoplasma associated with ash yellows and lilac witches’-broom: ‘Candidatus Phytoplasma fraxini’. Int. J. Syst. Evol. Microbiol. 1999, 49, 1605–1614. [Google Scholar] [CrossRef]
  110. Davis, R.E.; Zhao, Y.; Wei, W.; Dally, E.L.; Lee, M. ‘Candidatus Phytoplasma luffae’, a novel taxon associated with witches’ broom disease of loofah, Luffa aegyptica Mill. Int. J. Syst. Evol. Microbiol. 1999, 67, 3127–3133. [Google Scholar] [CrossRef] [PubMed]
  111. Verdin, E.; Salar, P.; Danet, J.L.; Choueiri, E.; Jreijiri, F.; El Zammar, S.; Gelie, B.; Bove, J.M.; Garnier, M. ‘Candidatus Phytoplasma phoenicium’sp. nov., a novel phytoplasma associated with an emerging lethal disease of almond trees in Lebanon and Iran. Int. J. Syst. Evol. Microbiol. 2003, 53, 833–838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Seemüller, E.; Schneider, B. ‘Candidatus Phytoplasma mali’, ‘Candidatus Phytoplasma pyri’and ‘Candidatus Phytoplasma prunorum’, the causal agents of apple proliferation, pear decline and European stone fruit yellows, respectively. Int. J. Syst. Evol. Microbiol. 2004, 54, 1217–1226. [Google Scholar] [CrossRef] [PubMed]
  113. Marcone, C.; Gibb, K.S.; Streten, C.; Schneider, B. ‘Candidatus Phytoplasma spartii’, ‘Candidatus Phytoplasma rhamni’ and ‘Candidatus Phytoplasma allocasuarinae’, respectively associated with spartium witches’-broom, buckthorn witches’-broom and allocasuarina yellows diseases. Int. J. Syst. Evol. Microbiol. 2004, 54, 1025–1029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Jung, H.Y.; Sawayanagi, T.; Wongkaew, P.; Kakizawa, S.; Nishigawa, H.; Wei, W.; Oshima, K.; Miyata, S.I.; Ugaki, M.; Hibi, T.; et al. ‘Candidatus Phytoplasma oryzae’, a novel phytoplasma taxon associated with wheat yellow dwarf disease. Int. J. Syst. Evol. Microbiol. 2003, 53, 1925–1929. [Google Scholar] [CrossRef]
  115. Šafárŏvá, D.; Zemanek, T.; Valova, P.; Navratil, M. ‘Candidatus Phytoplasma cirsii’, a novel taxon from creeping thistle [Cirsium arvense (L.) Scop.]. Int. J. Syst. Evol. Microbiol. 2016, 66, 1745–1753. [Google Scholar] [CrossRef] [PubMed]
  116. Davis, R.E.; Dally, E.L.; Gundersen, D.E.; Lee, I.-M.; Habili, N. “Candidatus Phytoplasma australiense,” a new phytoplasma taxon associated with Australian grapevine yellows. Int. J. Syst. Bacteriol. 1997, 47, 262–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Sawayanagi, T.; Horikoshi, N.; Kanehira, T.; Shinohara, M.; Bertaccini, A.; Cousin, M.T.; Hiruki, C.; Namba, S. ‘Candidatus Phytoplasma japonicum’, a new phytoplasma taxon associated with Japanese Hydrangea phyllody. Int. J. Syst. Evol. Microbiol. 1999, 49, 1275–1285. [Google Scholar] [CrossRef] [Green Version]
  118. Valiunas, D.; Staniulis, J.; Davis, R.E. ‘Candidatus Phytoplasma fragariae’, a novel phytoplasma taxon discovered in yellows diseased strawberry, Fragaria× ananassa. Int. J. Syst. Evol. Microbiol. 1999, 56, 277–281. [Google Scholar] [CrossRef] [Green Version]
  119. Quaglino, F.; Zhao, Y.; Casati, P.; Bulgari, D.; Bianco, P.A.; Wei, W.; Davis, R.E. ‘Candidatus Phytoplasma solani’, a novel taxon associated with stolbur-and bois noir-related diseases of plants. Int. J. Syst. Evol. Microbiol. 2013, 63 Pt 8, 2879–2894. [Google Scholar] [CrossRef] [Green Version]
  120. Martini, M.; Marcone, C.; Mitrović, J.; Maixner, M.; Delić, D.; Myrta, A.; Ermacora, P.; Bertaccini, A.; Duduk, B. ‘Candidatus Phytoplasma convolvuli’, a new phytoplasma taxon associated with bindweed yellows in four European countries. Int. J. Syst. Evol. Microbiol. 2012, 62 Pt 12, 2910–2915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Davis, R.E.; Harrison, N.A.; Zhao, Y.; Wei, W.; Dally, E.L. ‘Candidatus Phytoplasma hispanicum’, a novel taxon associated with Mexican periwinkle virescence disease of Catharanthus roseus. Int. J. Syst. Evol. Microbiol. 2016, 66, 3463–3467. [Google Scholar] [CrossRef]
  122. Fernández, F.D.; Galdeano, E.; Kornowski, M.V.; Arneodo, J.D.; Conci, L.R. Description of ‘Candidatus Phytoplasma meliae’, a phytoplasma associated with Chinaberry (Melia azedarach L.) yellowing in South America. Int. J. Syst. Evol. Microbiol. 2016, 66, 5244–5251. [Google Scholar] [CrossRef] [PubMed]
  123. Marcone, C.; Schneider, B.; Seemüller, E. ‘Candidatus Phytoplasma cynodontis’, the phytoplasma associated with Bermuda grass white leaf disease. Int. J. Syst. Evol. Microbiol. 2004, 54, 1077–1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Montano, H.G.; Davis, R.E.; Dally, E.L.; Hogenhout, S.; Pimentel, J.P.; Brioso, P.S. ‘Candidatus Phytoplasma brasiliense’, a new phytoplasma taxon associated with hibiscus witches’ broom disease. Int. J. Syst. Evol. Microbiol. 2001, 51, 1109–1118. [Google Scholar] [CrossRef] [Green Version]
  125. Arocha, Y.; Lopez, M.; Pinol, B.; Fernandez, M.; Picornell, B.; Almeida, R.; Palenzuela, I.; Wilson, M.R.; Jones, P. ‘Candidatus Phytoplasma graminis’ and ‘Candidatus Phytoplasma caricae’, two novel phytoplasmas associated with diseases of sugarcane, weeds and papaya in Cuba. Int. J. Syst. Evol. Microbiol. 2005, 55, 2451–2463. [Google Scholar] [CrossRef] [Green Version]
  126. Lee, M.; Bottner, K.D.; Secor, G.; Rivera-Varas, V. ‘Candidatus Phytoplasma americanum’, a phytoplasma associated with a potato purple top wilt disease complex. Int. J. Syst. Evol. Microbiol. 2005, 56, 1593–1597. [Google Scholar] [CrossRef] [Green Version]
  127. Jung, H.Y.; Sawayanagi, T.; Kakizawa, S.; Nishigawa, H.; Miyata, S.I.; Oshima, K.; Ugaki, M.; Lee, J.T.; Hibi, T.; Namba, S. ‘Candidatus Phytoplasma castaneae’, a novel phytoplasma taxon associated with chestnut witches’ broom disease. Int. J. Syst. Evol. Microbiol. 2002, 52, 1543–1549. [Google Scholar]
  128. Schneider, B.; Torres, E.; Martín, M.P.; Schröder, M.; Behnke, H.D.; Seemüller, E. ‘Candidatus Phytoplasma pini’, a novel taxon from Pinus silvestris and Pinus halepensis. Int. J. Syst. Evol. Microbiol. 2005, 55, 303–307. [Google Scholar] [CrossRef]
  129. Harrison, N.A.; Davis, R.E.; Oropeza, C.; Helmick, E.E.; Narvaez, M.; Eden-Green, S.; Dollet, M.; Dickinson, M. ‘Candidatus Phytoplasma palmicola’, associated with a lethal yellowing-type disease of coconut (Cocos nucifera L.) in Mozambique. Int. J. Syst. Evol. Microbiol. 2014, 64 Pt 6, 1890–1899. [Google Scholar] [CrossRef] [Green Version]
  130. Al-Saady, N.A.; Khan, A.J.; Calari, A.; Al-Subhi, A.M.; Bertaccini, A. ‘Candidatus Phytoplasma omanense’, associated with witches’-broom of Cassia italica (Mill.) Spreng. in Oman. Int. J. Syst. Evol. Microbiol. 2008, 58, 461–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Zhao, Y.; Sun, Q.; Wei, W.; Davis, R.E.; Wu, W.; Liu, Q. ‘Candidatus Phytoplasma tamaricis’, a novel taxon discovered in witches’-broom-diseased salt cedar (Tamarix chinensis Lour.). Int. J. Syst. Evol. Microbiol. 2009, 59, 2496–2504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Lee, I.M.; Bottner-Parker, K.D.; Zhao, Y.; Villalobos, W.; Moreira, L. ‘Candidatus Phytoplasma costaricanum’, a novel phytoplasma associated with an emerging disease in soybean (Glycine max). Int. J. Syst. Evol. Microbiol. 2011, 61, 2822–2826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Nejat, N.; Vadamalai, G.; Davis, R.E.; Harrison, N.A.; Sijam, K.; Dickinson, M.; Abdullah, S.N.A.; Zhao, Y. ‘Candidatus Phytoplasma malaysianum’, a novel taxon associated with virescence and phyllody of Madagascar periwinkle (Catharanthus roseus). Int. J. Syst. Evol. Microbiol. 2013, 63 Pt 2, 540–548. [Google Scholar] [CrossRef] [PubMed]
  134. Naderali, N.; Nejat, N.; Vadamalai, G.; Davis, R.E.; Wei, W.; Harrison, N.A.; Kong, L.; Kadir, J.; Tan, Y.H.; Zhao, Y. ‘Candidatus Phytoplasma wodyetiae’, a new taxon associated with yellow decline disease of foxtail palm (Wodyetia bifurcata) in Malaysia. Int. J. Syst. Evol. Microbiol. 2017, 67, 3765–3772. [Google Scholar] [CrossRef]
  135. Jardim, B.R.; Kinoti, W.M.; Tran-Nguyen, L.T.; Gambley, C.; Rodoni, B.; Constable, F.E. ‘Candidatus Phytoplasma stylosanthis’, a novel taxon with a diverse host range in Australia, characterised using multilocus sequence analysis of 16S rRNA, secA, tuf, and rp genes. Int. J. Syst. Evol. Microbiol. 2021, 71, ijsem004589. [Google Scholar]
  136. Miyazaki, A.; Shigaki, T.; Koinuma, H.; Iwabuchi, N.; Rauka, G.B.; Kembu, A.; Saul, J.; Watanabe, K.; Nijo, T.; Maejima, K.; et al. ‘Candidatus Phytoplasma noviguineense’, a novel taxon associated with Bogia coconut syndrome and banana wilt disease on the island of New Guinea. Int. J. Syst. Evol. Microbiol. 2018, 68, 170–175. [Google Scholar] [CrossRef] [PubMed]
  137. Jones, L.M.; Pease, B.; Perkins, S.L.; Constable, F.E.; Kinoti, W.M.; Warmington, D.; Allgood, B.; Powell, S.; Taylor, P.; Pearce, C.; et al. ‘Candidatus Phytoplasma dypsidis’, a novel taxon associated with a lethal wilt disease of palms in Australia. Int. J. Syst. Evol. Microbiol. 2021, 71, 004818. [Google Scholar] [CrossRef]
  138. Zhao, Y.; Davis, R.E. Criteria for phytoplasma 16Sr group/subgroup delineation and the need of a platform for proper registration of new groups and subgroups. Int. J. Syst. Evol. Microbiol. 2021, 66, 2121–2123. [Google Scholar] [CrossRef]
  139. MacLean, A.M.; Sugio, A.; Makarova, O.V.; Findlay, K.C.; Grieve, V.M.; Tóth, R.; Nicolaisen, M.; Hogenhout, S.A. Phytoplasma effector SAP54 induces indeterminate leaf-like flower development in Arabidopsis plants. Plant Physiol. 2011, 157, 831–841. [Google Scholar] [CrossRef] [Green Version]
  140. Wei, W.; Davis, R.E.; Nuss, D.L.; Zhao, Y. Phytoplasmal infection derails genetically preprogrammed meristem fate and alters plant architecture. Proc. Natl. Acad. Sci. USA 2013, 110, 19149–19154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Wu, W.; Cai, H.; Wei, W.; Davis, R.E.; Lee, I.M.; Chen, H.; Zhao, Y. Identification of two new phylogenetically distant phytoplasmas from S enna surattensis plants exhibiting stem fasciation and shoot proliferation symptoms. Ann. Appl. Biol. 2012, 160, 25–34. [Google Scholar] [CrossRef]
  142. Zwolińska, A.; Krawczyk, K.; Borodynko-Filas, N.; Pospieszny, H. Non-crop sources of Rapeseed Phyllody phytoplasma (‘Candidatus Phytoplasma asteris’: 16SrI-B and 16SrI-(B/L)L), and closely related strains. Crop Prot. 2019, 119, 59–68. [Google Scholar] [CrossRef]
  143. Lee, I.M.; Hammond, R.W.; Davis, R.E.; Gundersen, D.E. Universal amplification and analysis of pathogen 16S rDNA for classification and identification of mycoplasmalike organisms. Phytopathology 1993, 83, 834–842. [Google Scholar] [CrossRef]
  144. Gundersen, D.E.; Lee, I.-M. Ultrasensitive detection of phytoplasmas by nested-PCR assays using two universal primer pairs. Phytopathol. Mediterr. 1996, 35, 114–151. [Google Scholar]
  145. Christensen, N.M.; Nicolaisen, M.; Hansen, M.; Schulz, A. Distribution of phytoplasmas in infected plants as revealed by real-time PCR and bioimaging. Mol. Plant-Microbe Interact. 1993, 17, 1175–1184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Wei, W.; Kakizawa, S.; Suzuki, S.; Jung, H.Y.; Nishigawa, H.; Miyata, S.; Oshima, K.; Ugaki, M.; Hibi, T.; Namba, S. In planta dynamic analysis of onion yellows phytoplasma using localized inoculation by insect transmission. Phytopathology 2004, 94, 244–250. [Google Scholar] [CrossRef] [Green Version]
  147. Mehle, N.; Dreo, T.; Ravnikar, M. Quantitative analysis of “flavescence doreé” phytoplasma with droplet digital PCR. Phytopathogenic Mollicutes 2014, 4, 9–15. [Google Scholar] [CrossRef]
  148. Dickinson, M. Loop-mediated isothermal amplification (LAMP) for detection of phytoplasmas in the field. In Plant Pathology; Humana Press: New York, NY, USA, 2015; pp. 99–111. [Google Scholar]
  149. Wheatley, M.S.; Wang, Q.; Wei, W.; Bottner-Parker, K.D.; Zhao, Y.; Yang, Y. Cas12a-based diagnostics for potato purple top disease complex associated with infection by ‘Candidatus Phytoplasma trifolii’-related strains. Plant Dis. 2022, PDIS09212119RE. [Google Scholar] [CrossRef]
  150. Deng, S.J.; Hiruki, C. Amplification of 16S ribosomal-RNA genes from culturable and nonculturable Mollicutes. J. Microbiol. Methods 1991, 14, 53–61. [Google Scholar] [CrossRef]
  151. Schneider, B.; Seemüller, E.; Smart, C.D.; Kirkpatrick, B.C. Phylogenic classification of plant pathogenic mycoplasmalike organisms or phytoplasmas. In Molecular and Diagnostic Procedures in Mycoplasmology; Razin, I.R., Tully, J.G., Eds.; Academic Press: San Diego, CA, USA, 1995; pp. 369–380. [Google Scholar] [CrossRef]
  152. Casati, P.; Quaglino, F.; Stern, A.R.; Tedeschi, R.; Alma, A.; Bianco, P.A. Multiple gene analyses reveal extensive genetic diversity among ‘Candidatus Phytoplasma mali’populations. Ann. Appl. Biol. 2011, 158, 257–266. [Google Scholar] [CrossRef]
  153. Wei, W.; Cai, H.; Jiang, Y.; Lee, I.M.; Davis, R.E.; Ding, Y.; Yuan, E.; Chen, H.; Zhao, Y. A new phytoplasma associated with little leaf disease in azalea: Multilocus sequence characterization reveals a distinct lineage within the aster yellows phytoplasma group. Ann. Appl. Biol. 2011, 158, 318–330. [Google Scholar] [CrossRef]
  154. Varghese, N.J.; Mukherjee, S.; Ivanova, N.; Konstantinidis, K.T.; Mavrommatis, K.; Kyrpides, N.C.; Pati, A. Microbial species delineation using whole genome sequences. Nucleic Acids Res. 2015, 43, 6761–6771. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Key events of phytoplasmas and milestones of phytoplasma taxonomy. Details please see references [1,2,3,4,5,6,9,10,11,12,13,14,15,16,17,18,19,20,21,22]. [1] Severin, 1942; [2] Black, 1953; [3] Doi et al., 1967; [4] Lim and Sears, 1989; [5] ICSB, 1993; [6] Oshima et al., 2004; [9] Kirkpatrick et al., 1987; [10] Murray and Schleifer, 1994; [11] Lee et al., 1995; [12] IRPCM, 2004; [13] Zhao et al., 2009; [14] Kirdat et al., 2021; [15] Bertaccini et al., 2022; [16] Schildkraut et al., 1961; [17] Woese and Fox, 1977; [18] Wayne et al., 1987; [19] Hills and Dixon, 1991; [20] Hugenholtz et al., 2021; [21] Stackebrandt et al., 2002; [22] Auch et al., 2010.
Figure 1. Key events of phytoplasmas and milestones of phytoplasma taxonomy. Details please see references [1,2,3,4,5,6,9,10,11,12,13,14,15,16,17,18,19,20,21,22]. [1] Severin, 1942; [2] Black, 1953; [3] Doi et al., 1967; [4] Lim and Sears, 1989; [5] ICSB, 1993; [6] Oshima et al., 2004; [9] Kirkpatrick et al., 1987; [10] Murray and Schleifer, 1994; [11] Lee et al., 1995; [12] IRPCM, 2004; [13] Zhao et al., 2009; [14] Kirdat et al., 2021; [15] Bertaccini et al., 2022; [16] Schildkraut et al., 1961; [17] Woese and Fox, 1977; [18] Wayne et al., 1987; [19] Hills and Dixon, 1991; [20] Hugenholtz et al., 2021; [21] Stackebrandt et al., 2002; [22] Auch et al., 2010.
Biology 11 01119 g001
Figure 2. Number of phytoplasma genomes sequenced per year since 2003.
Figure 2. Number of phytoplasma genomes sequenced per year since 2003.
Biology 11 01119 g002
Table 1. A list of complete or draft phytoplasma genomes.
Table 1. A list of complete or draft phytoplasma genomes.
Organism NameOrganism Infraspecific Names Strain16Sr Group ClassificationHost SymptomsCountryReferences or GenBank DepositionAssembly AccessionAssembly Stats Total Sequence LengthAssembly LevelAssembly Submission Date
‘Catharanthus roseus’ aster yellows phytoplasmaDe VillaI-BMaize Bushy Stunt-likeSouth AfricaCoetzee et al. depositedGCF_004214875.1603,949Complete Genome20 February 2019
‘Chrysanthemum coronarium’ phytoplasmaOY-VI-Bonion yellowsJapan[50]GCF_000744065.1739,592Contig14 August 2014
‘Cynodon dactylon’ phytoplasmaLW01XIV-ABermuda grass white leafIndia[51]GCF_009268075.1483,935Scaffold22 October 2019
‘Echinacea purpurea’ witches’-broom phytoplasmaNCHU2014II-Apurple coneflower witches’ broomTaiwan[52]GCF_001307505.1545,427Contig7 October 2015
‘Fragaria x ananassa’ phyllody phytoplasmaStrPh-ClXIII-Fstrawberry phyllodyChile[53]GCF_018274325.1627,584Contig4 May 2021
‘Parthenium hysterophorus’ phyllody phytoplasmaPR34II-new subgroupSanta-Maria phyllodyIndiaKirdat depositedGCF_015100165.1740,170Contig29 October 2020
‘Parthenium sp.’ Phyllody phytoplasmaPR08II-DSanta-Maria phyllodyIndiaKirdat depositedGCF_015239935.1586,816Contig10 May 2021
‘Santalum album’ aster yellows phytoplasmaSW86I-BSandalwood SpikeIndiaTiwarekar depositedGCF_018283495.1554,025Contig5 May 2021
Aster yellows witches’-broom phytoplasma AYWBAYWBI-AAster yellows witches’-broom in lettuceUSA[7]GCF_000012225.1723,970Complete Genome1 November 2006
Ca. Phytoplasma aurantifoliaWBDLII-CLime witches’ broom phytoplasma in periwinkleOmanFoissac and Carle depositedGCF_002009625.1474,669Contig2 March 2017
Ca. Phytoplasma luffaeNCHU2019VIII-ALoofah witches’ broomTaiwan[54]GCF_018024475.1769,143Complete Genome16 April 2021
Ca. Phytoplasma maliATX-Aapple proliferationNA[55]GCF_000026205.1601,943Complete Genome4 July 2008
Ca. PhytoplasmaChTYXIII-MoXIII-GChinaberry yellowingArgentina[56]GCF_016876135.2751,949Contig14 April 2021
Ca. Phytoplasma oryzaeNGS-S10XI-ANapier Grass StuntKenya[57]GCF_003263355.1484,488Contig25 June 2018
Ca. Phytoplasma phoeniciumSA213XI-Dalmond witches’-broomLebanon[58]GCF_001189415.1345,965Contig30 July 2015
Ca. Phytoplasma piniMDPPXXI-Bpine phytoplasmaUSA[59]GCF_007821455.1474,136Contig1 August 2019
Ca. Phytoplasma pruniChTDIIIIII-BChina-tree declineArgentina[60]GCF_013391955.1790,517Contig8 July 2020
Ca. Phytoplasma pruniCXIII-AStone fruit tree declineNA[61]GCF_001277135.1598,511Contig1 September 2015
Ca. Phytoplasma sacchariSCGSXI-BSugarcane Grassy ShootIndia[51]GCF_009268105.1505,173Contig4 November 2019
Ca. Phytoplasma solaniSA-1XII-ABois noir in PeriwinkleNA[62]GCF_003698095.1821,322Contig30 October 2018
Ca. Phytoplasma solani284/09XII-AStolbur phytoplasma (in tobacco and parsley)NA[63]GCF_000970375.1570,238Chromosome22 October 2013
Ca. Phytoplasma sp. AldY-WA1AldY-WA1V-AAlder yellowsUSA[64]GCF_020312115.1457,625Scaffold6 October 2021
Ca. Phytoplasma triticiWBDI-C00420042pe blue dwarfChina[65]GCF_000495255.1611,462Contig1 November 2013
Ca. Phytoplasma ziziphiJwb-nkyV-Bjujube witches’ broomChina[66]GCF_003640545.1750,803Complete Genome12 October 2018
Chrysanthemum yellows phytoplasmaCYPI-BChrysanthemum yellowsItaly[67]GCF_000803325.1659,699Contig18 December 2014
Hydrangea phyllody phytoplasmaHPI-DHydrangea phyllodyJapan[68]GCF_018327665.1597,775Contig28 April 2021
Italian clover phyllody phytoplasma str. MA1MA1III-BItalian clover phyllody (in periwinkle)Italy[69]GCF_000300695.1597,245Contig1 October 2012
Maize bushy stunt phytoplasmaM3I-BMaize bushy stuntBrazil[70]GCF_001712875.1576,118Complete Genome25 August 2016
Milkweed yellows phytoplasma str. MW1MW1III-FMilkweed yellows (in periwinkle)Italy[69]GCF_000309485.1583,806Contig1 October 2012
Mulberry dwarf phytoplasmaMDGZ-01I-BMulberry dwarfChina[71]GCF_020714625.1622,358Complete Genome2 November 2021
New Jersey aster yellows phytoplasmaNJAYI-ANew Jersey aster yellows (in periwinkle)USA[72]GCA_002554195.1652,092Contig16 October 2017
Periwinkle leaf yellowing phytoplasmaDY2014I-BPeriwinkle leaf yellowingTaiwan[73]GCA_005093185.1824,596Contig2 May 2019
‘Brassica napus’ phytoplasmaTW1I-new subgroupRapeseed stunting and virescenceCanada[74]GCA_003181115.1743,598Contig31 May 2018
‘Elaeagnus angustifolia’ witches’-broom phytoplasmaTBZ1I-new subgroupRussian olive tree witches’-broomIranAzizpour et al. depositedGCA_018598675.1833,199Contig30 May 2021
Onion yellows phytoplasmaOYI-BOnion yellows (in chrysanthemum)Japan[6]GCA_000009845.1853,092Complete Genome9 December 2003
Paulownia witches’-broom phytoplasmaZhengzhouI-DPaulownia witches’-broomChina[75]GCF_019396865.1891,641Complete Genome29 July 2021
Peanut witches’-broom phytoplasma NTU2011NTU2011II-APeanut witches’-broom (in periwinkle)Taiwan[76]GCF_000364425.1566,694Contig26 March 2013
Poinsettia branch-inducing phytoplasma str. JR1JR1III-HPoinsettia branch-inducing (in periwinkle)Italy[69]GCF_000309465.1631,440Contig1 October 2012
Rice orange leaf phytoplasmaLD1IX-ARice orange leafChina[77]GCF_001866375.1599,264Contig4 November 2016
Sesame phyllody phytoplasmaSS02II-A or II-DSesame phyllodyIndia[78]GCF_018390775.1536,153Contig17 May 2021
Ca. Phytoplasma australiense XII-BMaintained in periwinkleAustralia[79]GCA_000069925.1879,959Complete Genome2 April 2008
Strawberry lethal yellows phytoplasma (CPA)NZSb11XII-B variantStrawberry lethal yellowsAustralia and New Zealand[80]GCF_000397185.1959,779Complete Genome16 May 2013
Texas Phoenix palm phytoplasmaFlo-TPPDIV-DTexas Phoenix Palm declineUSABao et al. depositedGCF_005774685.1744,506Contig23 May 2019
Vaccinium witches’-broom phytoplasma str. VACVACIII-FVaccinium witches’-broom (in periwinkle)Italy[69]GCF_000309405.1647,754Contig1 October 2012
Ca. Phytoplasma sp.Tabriz.2I-BElaeagnus sp. (symptoms not described)IranZirak et al. depositedGCA_019841745.1762,261Contig24 August 2021
Ca. Phytoplasma trifolii-relatedCBPPT1VI-APotato purple top (in periwinkle)USAWei et al. depositedPRJNA839414514,536Contig18 May 2022
Florescence dorée (FD) phytoplasmaCHV-AFlorescence dorée (in insect vector Scaphoideus titanusSwitzerland[81]PRJNA838420654,223Complete Genome27 June 2022
Table 2. A comparison of the 2022 revised guidelines and the 2004 original guidelines: issues require clarification and amendment.
Table 2. A comparison of the 2022 revised guidelines and the 2004 original guidelines: issues require clarification and amendment.
2004 Guidelines (IRPCM [12])2022 Revised Guidelines [15]Suggested Clarification and Amendments to the 2022 Revised Guidelines
(a) The ‘Ca. Phytoplasma’ species description should refer to a single, unique 16S rRNA gene sequence (>1200 bp). The strain from which this sequence was obtained should be named the ‘reference strain’ and not the ‘type strain’. Strains in which even minimal differences in the 16S rRNA gene sequence from the reference strain are detected do not ‘belong’ to the Candidatus species, but are ‘related’ to it.Extended the required length of 16S rRNA gene sequence from >1200 bp to full length or nearly full length.
Introduced the term “member strains”.
Comment: The new term may lead to erroneous assignment of a single given strain to more than one species. See Section 2.2 for details)
(a) The ‘Ca. Phytoplasma’ species description should refer to a single, unique 16S rRNA gene sequence (full length or nearly full length, >1500 bp) or whole genome sequence with at least 60% coverage (see Section 4 and Section 5). The strain from which this sequence was obtained should be named the ‘reference strain’ and not the ‘type strain’. Strains in which even minimal differences in the 16S rRNA gene sequence from the reference strain are detected are referred as ‘related’ to the Candidatus species.
(b) In general, a strain can be described as a novel ‘Ca. Phytoplasma’ species if its 16S rRNA gene sequence has <97.5% similarity to that of any previously described ‘Ca. Phytoplasma’ species.Revised the threshold value for 16S rRNA gene sequence identity-based Ca. Phytoplasma species delineation to 98.65%.
Proposed whole genome ANI-based criterum (95%) for Ca. Phytoplasma species delineation.
(b) In general, a strain can be described as a novel ‘Ca. Phytoplasma’ species if its 16S rRNA gene sequence shares <98.65% identity or its whole genome shares an ANI score <95–96% to that of any previously described ‘Ca. Phytoplasma’ species.
(c) There are, however, cases of phytoplasmas that share >97.5% of their 16S rRNA gene sequence, but clearly represent ecologically separated populations and, therefore, may deserve description as separate species. For such cases, description of two different species is recommended only when all three of the following conditions apply:
(i) the two phytoplasmas are transmitted by different vectors;
(ii) the two phytoplasmas have a different natural plant host (or, at least, their behaviour is significantly different in the same plant host);
(iii) there is evidence of significant molecular diversity, achieved by either hybridization to cloned DNA probes, serological reaction or PCR-based assay.
If a strain shares >98.65% similarity in 16S rRNA gene sequence and >95% genome ANI with previously established species, two out of five housekeeping genes (groEL, tuf, rp, secA and secY) with suggested criteria can be used for delineating new species.
Comment: All taxonomic frameworks attempt to reflect the ecological nature of organisms. Therefore, Rule (c) of the 2004 guidelines should be retained with necessary modifications. Due to the lack of universal primers and sufficient comparative analysis of housekeeping gene sequence data, it is still difficult to establish objective criteria for species delineation based on MLSA. But if the strain under study clearly represents ecologically separated populations, MLSA could be used to demonstrate significant molecular diversity in addition to fulfilling the unique vectorship and host specificity requirement (see Section 2.2 for details).
(c) There are, however, cases of phytoplasmas that share >98.65% identity in their 16S rRNA gene sequences or >95–96% ANI in their genomes, but clearly represent ecologically separated populations and, therefore, may deserve description as separate species. For such cases, description of two different species is recommended only when all three of the following conditions apply:
(i) the two phytoplasmas are transmitted by different vectors;
(ii) the two phytoplasmas have a different natural plant host (or, at least, their behaviour is significantly different in the same plant host);
(iii) there is evidence of significant molecular diversity, achieved by either hybridization to cloned DNA probes, serological reaction or MLSA assay on at least two housekeeping genes.
(d) The rank of subspecies should not be used. (d) The rank of subspecies should not be used.
(e) The reference strain should be made available to the scientific community from the authors of the Candidatus species description paper and it should be deposited (unless in vitro micropropagation proves impossible) in the micropropagated collection of Dr Assunta Bertaccini, DiSTA, Patologia Vegetale, Universita‘ di Bologna, Italy. (e) Due to strict international regulations, the rule e may no longer feasible; The gene clones of the reference strain should be deposited to the scientific committee or the authorized organizations in different countries in America, Europe, and Asia, etc (to be determined by phytoplasma scientists).
(f) Manuscripts that describe a novel ‘Ca. Phytoplasma’ species should preferably be submitted to the Int. J. Syst. Evol. Microbiol (IJSEM). (f) Manuscripts that describe a novel ‘Ca. Phytoplasma’ species should preferably be submitted to the Int. J. Syst. Evol. Microbiol (IJSEM).
(g) The abbreviation for Candidatus is Ca. (e.g., ‘Ca. Phytoplasma japonicum’ stands for ‘Candidatus Phytoplasma japonicum’). (g) The abbreviation for Candidatus is Ca. (e.g., ‘Ca. Phytoplasma japonicum’ stands for ‘Candidatus Phytoplasma japonicum’).
Table 3. An updated list of 16Sr groups/subgroups corresponding to named Candidatus Phytoplasma species.
Table 3. An updated list of 16Sr groups/subgroups corresponding to named Candidatus Phytoplasma species.
GroupNumber of ‘Ca. Phytoplasma’ SpeciesAccession Number of Reference StrainSubgroupReference
16SrI: Aster yellows group3Ca. Phytoplasma asteris’M3079016SI-B[99]
Ca. Phytoplasma lycopersici’EF19954916SrI-Y[100]
Ca. Phytoplasma tritici’NZ AVAO0100000316SrI-C[82]
16SrII: Peanut witches’ broom group1Ca. Phytoplasma aurantifolia’U1544216SrII-B[41]
* AbolishedCa. Phytoplasma australasia’Y1009616SrII-D[101]
16SrIII: X-disease group1Ca. Phytoplasma pruni’JQ04439316SrIII-A[102]
16SrIV: Coconut lethal yellows group2Ca. Phytoplasma palmae’U1874716SrIV-A[12,15]
Ca. Phytoplasma cocostanzaniae’X8011716SrIV-C[12,15]
16SrV: Elm yellows group4Ca. Phytoplasma ulmi’AY19765516SrV-A[103]
Ca. Phytoplasma ziziphi’AB05287616SrV-B[104]
Ca. Phytoplasma rubi’AY19764816SrV-E[105]
Ca. Phytoplasma balanitae’AB68967816SrV-new subgroup[106]
16SrVI: Clover proliferation group2Ca. Phytoplasma trifolii’AY39026116SrVI-A[107]
Ca. Phytoplasma sudamericanum’GU29208116SrVI-I[108]
16SrVII: Ash yellows group1Ca. Phytoplasma fraxini’AF09220916SrVII-A[109]
16SrVIII: Loofah witches’ broom group1Ca. Phytoplasma luffae’AF24895616SrVIII-A[110]
16SrIX: Pigeon pea witches’ broom group1Ca. Phytoplasma phoenicium’AF24895616SrIX-D[111]
16SrX: Apple proliferation group4Ca. Phytoplasma mali’AJ54254116SrX-A[112]
Ca. Phytoplasma pyri’AJ54254316SrX-C[112]
Ca. Phytoplasma prunorum’AJ54254416SrX-F[112]
Ca. Phytoplasma spartii’X9286916SrX-D[113]
16SrXI: Rice yellow dwarf group3Ca. Phytoplasma oryzae’AB05287316SrXI-A[114]
Ca. Phytoplasma cirsii’KR86914616SrXI-D[115]
Ca. Phytoplasma sacchari’VWXM0000000016SrXI-B[14]
16SrXII: Stolbur group5Ca. Phytoplasma australiense’L7686516SrXII-B[116]
Ca. Phytoplasma japonicum’AB01042516SrXII-D[117]
Ca. Phytoplasma fragariae’DQ08642316SrXII-E[118]
Ca. Phytoplasma solani’AF24895916SrXII-A[119]
Ca. Phytoplasma convolvuli’JN83370516SrXII-H[120]
16SrXIII: Mexican periwinkle virescence group2Ca. Phytoplasma hispanicum’AF24896016SrXIII-A[121]
Ca. Phytcoplasma meliaeKU85094016SrXIII-G[122]
16SrXIV: Bermudagrass white leaf group1Ca. Phytoplasma cynodontis’AJ55098416SrXIV-A[123]
16SrXV: Hibiscus witches’ broom group1Ca. Phytoplasma brasiliense’AF14770816SrXV-A[124]
16SrXVI: Sugar cane yellow leaf syndrome group1Ca. Phytoplasma graminis’AY72522816SrXVI-A[125]
16SrXVII: Papaya bunchy top group1Ca. Phytoplasma caricae’AY72523416SrXVII-A[125]
16SrXVIII: American potato purple top wilt group1Ca. Phytoplasma americanum’DQ17412216SrXVIII-A[126]
16SrXIX: Japanese chestnut witches’ broom group1Ca. Phytoplasma castaneae’AB05498616SrXIX-A[127]
16SrXX: Buckthorn witches’ broom group1Ca. Phytoplasma rhamni’X7643116SrXX-A[113]
16SrXXI: Pine shoot proliferation group1Ca. Phytoplasma pini’AJ63215516SrXXI-A[128]
16SrXXII: Nigerian coconut lethal decline group1Ca. Phytoplasma palmicola’KF75138716SrXXII-A[129]
16SrXXIII: Buckland Valley grapevine yellows group1 unnamed species identifiedAY08360516SrXXIII-A[94]
16SrXXIV: Sorghum bunchy shoot group 1 unnamed new species identifiedAF50932216SrXXIV-A
16SrXXV: Weeping tea tree witches’ broom group 1 unnamed new species identifiedAF52167216SrXXV-A
16SrXXVI: Mauritius sugar cane yellows D3T1 group1 unnamed new species identifiedAJ53917916SrXXVI-A
16SrXXVII: Mauritius sugar cane yellows D3T2 group1 unnamed new species identifiedAJ53918016SrXXVII-A
16SrXXVIII: Havana derbid group 1 unnamed new species identifiedAY74494516SrXXVII-A
16SrXXIX: Cassia witches’ broom group1Ca. Phytoplasma omanense’EF66605116SrXXIX-A[130]
16SrXXX: Salt cedar witches’ broom group1Ca. Phytoplasma tamaricis’FJ43266416SrXXX-A[131]
16SrXXXI: Soybean stunt phytoplasma group1Ca. Phytoplasma costaricanum’HQ22563016SrXXXI-A[132]
16SrXXXII: Malaysian periwinkle virescence group1Ca. Phytoplasma malaysianum’EU37193416SrXXXII-A[133]
16SrXXXIII: Allocasuarina group1Ca. Phytoplasma allocasuarinae’AY13552316SrXXXIII-A[12]
16SrXXXIV: grapevine yellows No new species identified, abolishedDQ232752
16SrXXXV: Pepper witches’-broom No new species identified, abolishedEU125184
16SrXXXVI: foxtail palm yellow decline group1Ca. Phytoplasma wodyetiae’KC84487916SrXXXVI-A[134]
16SrXXXVII: Stylosanthes little leaf group1Ca. Phytoplasma stylosanthis’MT43155016SrXXXVII-A[135]
16SrXXXVIII: Bogia coconut syndrome group1Ca. Phytoplasma noviguineense’LC22875516SrXXXVIII-A[136]
16SrXXXIX: Palm lethal wilt group1Ca. Phytoplasma dypsidis’MT53619516SrXXXIX-A[137]
* Abolished: ‘Ca. Phytoplasma australasia’ was originally described by White et al. [101]. It was later removed from the ‘Ca. Phytoplasma’ species list by the IRPCM as its 16S rRNA gene sequence shares 99.5% sequence identity with that of ‘Ca. Phytoplasma aurantifolia’ and there is no evidence that it represents an ecologically separated population [12]. ‘Ca. Phytoplasma australasia’ was erroneously included in 2022 guidelines [15] and should be removed.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wei, W.; Zhao, Y. Phytoplasma Taxonomy: Nomenclature, Classification, and Identification. Biology 2022, 11, 1119. https://doi.org/10.3390/biology11081119

AMA Style

Wei W, Zhao Y. Phytoplasma Taxonomy: Nomenclature, Classification, and Identification. Biology. 2022; 11(8):1119. https://doi.org/10.3390/biology11081119

Chicago/Turabian Style

Wei, Wei, and Yan Zhao. 2022. "Phytoplasma Taxonomy: Nomenclature, Classification, and Identification" Biology 11, no. 8: 1119. https://doi.org/10.3390/biology11081119

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop