Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Microbiological Research 239 (2020) 126522

Contents lists available at ScienceDirect

Microbiological Research
journal homepage: www.elsevier.com/locate/micres

Plant growth-promoting bacteria isolated from wild legume nodules and T


nodules of Phaseolus vulgaris L. trap plants in central and southern Mexico
Erika Yanet Tapia-García, Verónica Hernández-Trejo, Joseph Guevara-Luna,
Fernando Uriel Rojas-Rojas, Ivan Arroyo-Herrera, Georgina Meza-Radilla,
María Soledad Vásquez-Murrieta, Paulina Estrada-de los Santos*
Instituto Politécnico Nacional, Escuela Nacional de Ciencias Biológicas, Prol. Carpio y Plan de Ayala s/n, Col. Santo Tomás, C.P. 11340, Mexico City, Mexico

A R T I C LE I N FO A B S T R A C T

Keywords: Central southern Mexico contains highly diverse legumes. In this study, nodule-associated bacteria (NAB) were
Nodule associated bacteria isolated from wild legume nodules and from nodules on Phaseolus vulgaris plants used as a plant-trap in soils from
Nodulation the same areas as the wild legumes. The bacteria were identified through the 16S rRNA gene sequence analysis,
Biological nitrogen fixation tested for plant growth-promoting (PGP) activities and the production of antimicrobial compounds, and ana-
Mimosa
lyzed for potential nodulation by amplifying the nodC gene. Several genera with PGP activity were isolated from
Rhizosphere
legume nodules, including Achromobacter, Acinetobacter, Bacillus, Brevibacillus, Brevibacterium, Burkholderia,
Cupriavidus, Dyella, Ensifer, Enterobacter, Herbaspirillum, Kosakonia, Labrys, Microbacterium, Moraxella,
Paraburkholderia, Pseudomonas, Rhizobium, Stenotrophomonas; and Aeromonas, Marinococcus Pseudarthrobacter
and Pseudoxanthomonas were found in plant legume nodules for the first time. Pseudomonas was the most
common bacteria, and Mimosa pudica was colonized by the largest number of genera (6 different genera). A
Burkholderia strain from the Burkholderia cepacia complex and a firmicutes strain harbor the nodC gene, iden-
tifying them as potential novel nodulating bacteria and showing that most of the strains isolated in this study
were NAB. The most frequent PGP activity identified among the strains isolated from wild legumes was IAA
synthesis. Two bacteria, Stenotrophomonas sp. and Rhizobium sp., synthesized more than 250 μg/ml, which is
more than the level of synthesis reported in this study for Azospirillum brasilense Sp7 (59.77 μg/ml). Nitrogen
fixation and antimicrobial compound production were not common, but the production of siderophores was
frequently found among all the strains. This study shows that diverse NAB with PGP activity are very common in
the legume nodules from central southern Mexico.

1. Introduction legumes remain mostly unexplored (Selvakumar et al., 2008; Sánchez-


Cruz et al., 2019). In any case, the role of NAB within the nodule has
Some endophytes are plant growth-promoting (PGP) bacteria that been understudied, and whether they are important to plant develop-
are important in plant development in challenging environmental ment is unknown (Martínez-Hidalgo and Hirsch, 2017). Understanding
conditions (Muresu et al., 2019). These microorganisms can coexist the strategies used by legumes to select the best partners for improving
with Alpha- or Beta-rhizobia in the legume nodules (Ibáñez et al., 2017; BNF is important for sustainable agriculture (Clua et al., 2018). This
Muresu et al., 2008) but are themselves unable to nodulate. Some of knowledge would be useful in formulating biofertilizers based on single
them can fix nitrogen (BNF, biological nitrogen fixation) or carry out strains or combining NAB and rhizobia, thereby diminishing the use of
other PGP activities, and are therefore called nodule-associated bacteria chemical fertilizers (Martínez-Hidalgo and Hirsch, 2017).
(NAB) or non-rhizobial bacteria (Ibáñez et al., 2017; Clua et al., 2018; In Mexico the study of Alpha-rhizobia, in Phaseolus vulgaris L., has
Martínez-Hidalgo and Hirsch, 2017). been extensive, and several bacterial species have been described
Wild legumes can associate with noncultivable rhizobia and some within the genera Rhizobium, such as Rhizobium acidisoli, Rhizobium
cultivable endophytes; in contrary, agricultural legumes contain several endophyticum, Rhizobium esperanzae, Rhizobium etli, Rhizobium hi-
cultivable rhizobia but fewer endophytes (Muresu et al., 2019). Gen- dalgoense, Rhizobium mesoamericanum and Rhizobium tropici (Martínez-
erally, the study of NAB is derived from agricultural legumes, but wild Romero et al., 1991; Segovia et al., 1993; López-López et al., 2010,


Corresponding author.
E-mail address: pestradadelossantos@gmail.com (P. Estrada-de los Santos).

https://doi.org/10.1016/j.micres.2020.126522
Received 28 February 2020; Received in revised form 24 May 2020; Accepted 30 May 2020
Available online 13 June 2020
0944-5013/ © 2020 Elsevier GmbH. All rights reserved.
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

2012; Cordeiro et al., 2017; Yan et al., 2017; Román-Ponce et al., chloride to keep them dry and the soils were placed in plastic bags. The
2016). R. etli was found in nodules of Acacia farnesiana and Acacia plant specimens were preserved for identification, and the soil was used
cochliacantha used as trap plants with soil from Morelos state in Mexico for plant-trap experiments and physicochemical tests (see below).
(Miranda-Sánchez et al., 2016). Acaciella angustissima collected from
the Sumidero Canyon National Park in Chiapas, Mexico was found 2.2. Bacterial isolation from legume nodules
nodulated by Sinorhizobium chiapanecum, Ensifer mexicanum, Rhizobium
tropici, Mesorhizobium plurifarium and Agrobacterium tumefaciens (now Five legume nodules were randomly selected from each wild plant
Agrobacterium radiobacter) (Rincón-Rosales et al., 2009). In Desmodium and washed with sterile water 3 times. The legume nodule surface was
nodules Bradyrhizobium was found in Northern of Mexico (Parker, sterilized by soaking in ethanol for 30 s, immersed in 10 % commercial
2002). In Mimosa, R. etli, R. mesoamericanum, Burkholderia sp., Ensifer chloride for 10 min and rinsed 5 times with sterile water. The water
sp., and Rhizobium sp. have been isolated in Mexico (Wang et al., 1999; from the final rinse was used to verify the sterilization. Finally, the
López-López et al., 2012; Bontemps et al., 2016). Also, in Mexico Rhi- legume nodules were crushed with a plastic pestle with 40 μl of water.
zobium grahamii was isolated from nodule of Leucaena leucocephala The nodule suspension was inoculated (15 μl) onto plates with both
(López-López et al., 2012). However, P. vulgaris L. is the second most yeast extract mannitol (YM) medium containing 5 g of mannitol and
important crop in the world, after maize (Avelar et al., 2012; Cordeiro peptone yeast (PY) medium. The plates were incubated at 28 °C for 3–5
et al., 2017). Previous works showed that P. vulgaris L. is a promiscuous days. Isolates with different colony morphologies were selected and
legume that can be nodulated by different rhizobia (Martínez-Romero, stored in 30 % glycerol at −70 °C.
2003), however, none these studies focused on the isolation of NAB.
In a recent study the NAB Enterobacter sp. NOD1 isolated from 2.3. Plant-trap experiments
Mimosa pudica L. nodules in Chiapas (Mexico) was characterized
(Sánchez-Cruz et al., 2019). This strain was unable to form nodules in P. The experiments were carried out by mixing soil with vermiculite
vulgaris L., but synthesizes indoleacetic acid (IAA) and siderophores, (3:1) and adding the mixture to a 325 mL pot. P. vulgaris var. Negro
and solubilizes phosphate. Other strains from the same study identified Jamapa seeds were disinfected with 10 % commercial chloride for 10
as Serratia sp. were antagonists to the phytopathogens Alternaria solani min. The seeds were placed in 15 % agar-water plates and incubated at
and Phytophthora capsici. 30 °C for 72 h. The germinated seeds were planted in the pots and
In order to understand more about NAB from wild legumes in incubated for 45 days in a greenhouse at 30 °C, 14 h light – 10 h dark.
Mexico, the objective of this work was to analyze the PGP features of The legume nodules were recovered and treated as described above for
cultivable bacteria isolated from legume nodules of wild plants in bacterial isolation. For each soil, two replicates were included.
central and southeastern Mexico and bacteria obtained from P. vulgaris
nodules used as a trap plant grown in soil from the same areas. 2.4. Bacterial identification

2. Materials and methods 2.4.1. Bacterial DNA isolation


The strains were grown in YM liquid medium for 2 days at 30 °C.
2.1. Plants, legume nodules and soil sampling After the incubation period, cell pellets were obtained by centrifugation
(12 300 X g for 5 min). The DNA was isolated according to Ausubel
Soil, plants and legume nodules were collected from Chiapas, et al. (1987).
México, Guerrero, Morelos, Veracruz and Tabasco states in México
during April and May 2013 (Fig. 1). Chiapas and Veracruz are the states 2.4.2. 16S rRNA gene amplification
with the greatest biological diversity in México, and Mexico State, The 16S rRNA fragment of selected isolates was amplified using
Guerrero and Tabasco are states where the Mimosa genus is frequently primers 27F/1492R (Lane, 1991). The PCR conditions consisted of an
found. The legumes collected were identified as Acacia sp. W., Acaciella initial denaturation cycle (95 °C, 5 min), 30 amplification cycles (95 °C,
sp. B. & R., Arachis sp. L., Crotalaria pumila Ort., Desmodium sp. Desv., 1 min; 57 °C, 1 min; 72 °C, 1.5 min), and then a final elongation cycle
Mimosa affinis B. L. Rob., Mimosa benthamii J. F. Macbr., Mimosa galeottii (72 °C, 10 min). The PCR fragments were sequenced at Macrogen Inc.
Benth., Mimosa pigra L., Mimosa pudica L., Mimosa quadrivalvis L., Leu- (https://dna.macrogen.com/). Sequences were edited with ChromasPro
caena sp. Benth., and Trifolium sp. L. The collected plants belonged to 2.1.5 (Technelysium Pty Ltd). The identification of the isolates was
the families Mimosoideae and Papilionoideae. Thirty-four legume no- carried out with the tool BLASTn from NCBI (https://blast.ncbi.nlm.
dules were recovered from wild plants and 46 rhizospheric soils sam- nih.gov/Blast.cgi) using the database 16S ribosomal RNA sequences for
ples were taken from the collection sites (Table S1). The legume no- Bacteria and Archaea and selecting the option Sequences from type
dules were transported to the laboratory in tubes containing calcium material. All sequences were deposited in GenBank (Table S2).

Fig. 1. Location sites (indicated with green circles) for collecting legume plants and soil analyzed in this study.

2
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

2.4.3. Phylogenetic analysis 2.7. Inhibition of bacterial plant pathogens


Nucleotide 16S rRNA alignments were constructed using the soft-
ware MUSCLE (https://www.ebi.ac.uk/Tools/msa/muscle/). The phy- The ability of the strains to control bacterial phytopathogens was
logenetic tree was constructed using the Bayesian inference method tested using the double-layer agar technique (Rojas-Rojas et al., 2018;
with the software BEAST v2.6.0 (Bouckaert et al., 2019). jModeltest Chávez-Ramírez et al., 2020). Briefly, bacterial strains (producer
2.1.10 software (Darriba et al., 2012) was used to calculate the ap- strains) were spotted (2 μl from a bacterial culture adjusted to 0.1
propriate evolution model for the data according to the AIC (Akaike OD600) on PDA (potato dextrose agar) medium and incubated at 30 °C
information criterion) (Akaike, 1974). The GTR + G+I model, with α for 72 h. Next, plates were exposed to chloroform vapor and overlaid
= 0.6330 for the gamma distribution and p-inv = 0.3430, was used. with soft agar medium seeded with the phytopathogen bacteria (in-
dicator or sensitive strain). Antimicrobial activity was reported as an
2.5. Plant growth-promoting activities inhibition zone around the producer strain.

2.5.1. Free-living nitrogen fixation 2.8. Soil analysis


Free-living nitrogen fixation was tested by inoculating the bacterial
suspension (100 μl, OD600 = 0.1) in 10 ml vials containing 5 ml of The collected soil was tested for pH and water retention capacity. A
BMGM medium (Estrada-de los Santos et al., 2001). The cultures were mixture of soil and water (1:2/5 w/v) was prepared to determine pH
incubated at 29 °C for 72 h. Ten percent of the culture atmosphere was (Thomas, 1996), resulting slightly acidic, with values of 5.8–6.7. The
replaced with acetylene and incubated for 24 h. The bacterial nitrogen water retention capacity was quantified gravimetrically (Gardner,
fixation was tested with an acetylene reduction activity (ARA) assay 1986), showing that the soils used in the trap experiments were
using a Clarus 580 gas chromatographer (PerkinElmer). Each isolate 132–209 %. The analyses were performed with three replicates.
was tested in triplicate.
2.9. Data analysis
2.5.2. Phosphate solubilization
The isolates were grown in YM medium for 24–72 h at 29 °C. The Venn diagrams were generated using the Venn tool for
cultures were adjusted to 0.1 (OD600), and 2 μl was inoculated on plates Bioinformatics and Evolutive Genomics, using the server at the
containing NBRIP medium with tricalcium phosphate [Ca3(PO4)2] University of Ghent (http://bioinformatics.psb.ugent.be/webtools/
(Nautiyal, 1999). The plates were incubated at 29 °C for 3–5 days. The Venn/). For the analysis of PGP features (IAA, siderophore biosynth-
solubilization halos were measured, and the result was gathered by esis and phosphate solubilization), media data were compared using the
subtracting the size of the colony from the size of the halo. Each isolate post hoc Tukey test p < 0.05 with the Minitab v17.1.0 software.
was tested in triplicate.
3. Results
2.5.3. Siderophore production
The universal chemical assay of chromoazurol S (CAS) agar plates 3.1. Bacterial isolation from wild plants and trap plants nodules
(Caballero-Mellado et al., 2007) was modified by omitting the deferri-
zation process, and 3 % glucose was added as a carbon source in MM9 There were 257 bacteria isolated from wild legume nodules. These
medium (Schwyn and Neilands, 1987). The bacterial strains were pre- isolates were derived from the plant genera Mimosa, Desmodium,
viously grown in YM medium for 24–72 h at 29 °C. Then, the culture Arachis, Leucaena, Acacia and Acaciella. Out of the 46 soils used in the
was adjusted to 0.1 (OD600), and 2 μl was inoculated in CAS-MM9 trap experiments, 37 resulted in the development of bean root nodules.
medium in triplicate. The yellow halos were measured, and the colony Among these, more than 20 root nodules were formed with 13 soils.
was subtracted to obtain the result. The number of microorganisms isolated from legume nodules in trap
experiments was 171 and the microorganisms were originated from
2.5.4. Indoleacetic acid production soils where Mimosa, Leucaena, Crotalaria, Acacia and P. vulgaris plants
All isolates were grown in 5 ml YM medium at 28 °C for 24–72 h at were growing.
120 rpm. The cultures were adjusted to 0.1 (OD600). An aliquot of 100
μl was inoculated into 5 ml of JP medium broth (Jain and Patriquin, 3.2. Bacterial identification and phylogenetic inference
1985), in triplicate and incubated at 29 °C for 72 h. The cultures were
centrifuged at 12,000 rpm, and the supernatant was used to quantify The identification of the isolates was carried out by sequencing and
the production of IAA with Salkowski solution (1:1) (Glickmann and analyzing the 16S rRNA fragment. Several genera were identified,
Dessaux, 1995). After 30 min of incubation at room temperature in the among which the phyla Proteobacteria was the most common and,
dark, the absorbance was measured at OD535. Each culture was com- Actinobacteria and Firmicutes were the least common (Table 1, Fig. 2).
pared with a standard curve of pure IAA (Sigma-Aldrich). Within the Proteobacteria, we identified the genera Ensifer, Labrys and
Rhizobium from the Alpha-proteobacteria, the genera Achromobacter,
2.6. nodC gene amplification and phylogenetic inference Burkholderia, Cupriavidus, Herbaspirillum and Paraburkholderia from the
Beta-proteobacteria and the genera Acinetobacter, Aeromonas, Dyella,
The nodC gene was amplified by PCR as an indirect nodulation test. Enterobacter, Kosakonia, Moraxella, Pseudomonas, Stenotrophomonas and
For Alpha-proteobacteria, a fragment of 560 bp was amplified with the Pseudoxhanthomonas from the Gamma-proteobacteria. Within the Ac-
primers NodCfor540 TGATYGAYATGGARTAYTGGCT and tinobacteria, the genus identified were Brevibacterium, Microbacterium
NodCrev1160 (Sarita et al., 2005). A 635 bp fragment for Beta-pro- and Pseudarthrobacter and within the Firmicutes Bacillus, Brevibacillus
teobacteria nodC was amplified using the primers nodCfor ACTSATA- and Marinococcus. The genus Pseudomonas was the most common
CTYAACGTMGAYTC and nodCrev GMRAAYCCRAGAAATCGAAG among the microorganisms isolated from wild legume and trap plant
(Bontemps et al., 2010). The PCR conditions consisted of an initial nodules, and Rhizobium, Cupriavidus and Paraburkholderia were the
denaturation cycle (95 °C, 5 min), 30 amplification cycles (95 °C, 1 min; most common among the genera recovered from trap plant nodules
57 °C, 1 min; 72 °C, 45 s), and then a final elongation cycle (72 °C, 10 after Pseudomonas (Figs. 2 and 3). The phylogenetic analysis from
min). The nodC sequences were compared to those of several genera Alpha-proteobacteria (Fig. S1) showed a strain belonging to the genus
that are known to nodulate legumes. The alignment and phylogenetic Labrys close to Labrys okinawensis/Labrys monachus (98.35 % similarity)
analyses were performed similarly to the 16S rRNA sequences analysis. and a strain from the genus Ensifer close to Ensifer adherens (99.77 %

3
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

Table 1
Bacterial species isolated from wild legume nodules in Mexico and nodules from Phaseolus vulgaris used as a trap plant.

(continued on next page)

4
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

Table 1 (continued)

IAA, indoleacetic acid. For IAA, phosphate solubilization and siderophore production, the asterisk means significative difference compared to the control (above the
control) (P < 0.05) (n = 3). The statistical analysis was performed independently for each plant growth promotion feature tested. In grey are the values that
represent the highest activity.
–, production or activity absent. +, production or activity present. ND, not determined. NI, not identified. The geographical reference for each plant or soil sample is
indicated as: 1N 15°6'4.68” W 92°23'53.447”. 2N 14°57'46.44” W 92°9'20.375”. 3N 14°58'19.308” W 92°9'26.315”. 4N 14°52'30.18” W 92°21'24.768”. 5N 15°5'38.004”
W 92°23'31.239”. 6N 14°52'30.18’’ W 92°21'24.768”. 7N 18°54'20.999” W 99°1'28.999”. 8N 19°0'49” W 98°50'35.998”. 9N 14°58'19.308” W 92°9'26.315”. 10N
15°5'38.004” W 92°23'31.239”. 11N 17°54'51.408” W 99°13'0.407”. 12N 17°36'12.096” W 99°31'25.571”. 13N 17°34'12.576” W 99°35'33.504”. 14N 19°0'49” W 98° 50'
35.998”. 15N 18°0'24.264” W 93°19'8.939”. 16N 18°0'24.264” W 93°19'8.939”. 17N 18°0'27.54” W 93°19'5.664”. 18N 18°5'17.16” W 93°15'10.368”. 19N 18°13'24.816”
W 93°25'0.803”. 20N 18°29'39.516” W 95°2'35.015”.
£
Inhibition of any of the following phytopathogens: Ralstonia solanacearum LMG 2299T, R. solanacearum LMG 2300, Xanthomonas axonopodis pv phaseoli XHFR-02
and Agrobacterium tumefaciens AGJI-04, see Fig. S7 for more details.

similarity). Moreover, a number of strains identified as Rhizobium were (99.47–99.55 % similarity), Aeromonas sanarellii (99.62–99.85 % simi-
intermingled with several Rhizobium species. Within the Beta-prote- larity), Moraxella osloensis (99.0–99.13 % similarity) and Pseudox-
bacteria (Fig. S2), a strain was close to Herbaspirillum robiniae, 99.42 % anthomonas indica (99.11 % similarity). Some strains were identified as
similarity) and a strain to Achromobacter deleyi (99.49 % similarity). Acinetobacter lactucae (99.78 % similarity), Acinetobacter lwoffii (99.85
The strains from the genus Burkholderia belong to the Burkholderia ce- % similarity) and Acinetobacter johnsonii (99.43 % similarity) and other
pacia complex and were close to Burkholderia contaminans (99.63 % strains were related to Dyella marensis (99.93–100 % similarity). The
similarity). The Paraburkholderia genus was also present, and the strains strains identified as Stenotrophomonas were close to Stenotrophomonas
were close to Paraburkholderia caballeronis (100 % similarity), Para- maltophilia (99.28–99.57 % similarity). Pseudomonas was the genus
burkholderia unamae (99.35 % similarity), Paraburkholderia mimosarum with the highest number of isolated strains (Fig. S4) which were close to
(99.71 % similarity) and Paraburkholderia phymatum (96-97-98.40 % Pseudomonas oryzihabitans, Pseudomonas plecoglossicida, Pseudomonas
similarities). The strains identified as Cupriavidus were close to Cu- aeruginosa, Pseudomonas entomophila, Pseudomonas reidholzensis, Pseu-
priavidus oxalaticus (98.52–99.46 similarity), Cupriavidus taiwnanensis/ domonas japonica and Pseudomonas ntritireducens, with similarity values
Cupriavidus alkaliphilus (98.71–98.56 % similarity) and Cupriavidus higher than 99 %. Within the Firmicutes (Fig. S5), one strain was close
metallidurans (99.07–99.10 % similarity). Within the Gamma-proteo- to Marinococcus tarijensis (99.50 % similarity). Some strains were
bacteria (Fig. S3), a strain was close to Enterobacter cloacae (99.78 % identified as Bacillus aryabhattai (99.93 % similarity), Bacillus altitu-
similarity) and several strains were close to Kosakonia cowanii dinis/Bacillus aerius (99.13 % similarity) and Bacillus drentensis (99.42),

5
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

Fig. 2. Frequency of bacterial genera identified from legume nodules. WLN, nodules from wild legumes. TPN, nodules from Phaseolus vulgaris used as trap plant.

and two strains were similar to Brevibacillus nitrificans (98.71–99.93 %). (Caballero-Mellado et al., 2007), a result that was confirmed in this
Within the Actinobacteria (Fig. S6), two strains were identified as study.
Pseudarthrobacter oxydans (99.33–99.77 % similarity). The strain
CpTa8-2b was related to Microbacterium azadirachtae (99.93 % simi- 3.4. Nodulation potential in the strains
larity).
Two sets of primers were used to amplify nodC, one targeting the
nodC from Alpha-proteobacteria and the other targeting nodC from
3.3. Characterization of bacterial plant growth promotion activity
Beta-proteobacteria. Among the genera found to have the potential to
form legume nodules were Paraburkholderia, Cupriavidus, Labrys,
All strains were analyzed for the production of metabolites involved
Brevibacillus, Burkholderia and Rhizobium (Table 2). Among these, the
directly in plant growth promotion (PGP), such as IAA, and nitrogen
Beta-proteobacteria Burkholderia and the Firmicutes Brevibacillus are
fixation, and for their indirect PGP activity, such as phosphate solubi-
not known as bacteria that are able to nodulate. The phylogenetic
lization, siderophore synthesis or inhibition of bacterial phytopatho-
analysis of the nodC sequences showed that Burkholderia and Breviba-
gens. Forty-four bacteria isolated from wild legumes had at least one
cillus clustered with alpha-rhizobia, which may have horizontally
PGP feature, and 35 bacteria with at least one PGP feature were isolated
transferred to these two genera (Fig. 5).
from trap-plants, for a total of 79 strains in total (Fig. 4). None of the
strains harbored all the analyzed features. However, among the strains
from wild legume nodules, the most common PGP feature was IAA 3.5. Anti-phytopathogenic bacterial activity
production (38 strains) (Table 1, Fig. 4). Most of the strains synthesized
IAA, in amounts from 1.84 to 45.09 μg/ml. However, Rhizobium sp. All strains were tested against phytopathogenic bacteria such as
LIz42a and Stenotrophomonas sp. LXo13 produced more than 250 μg/ml Agrobacterium tumefaciens AGJI-04, Ralstonia solanacearum LMG 2299T
of IAA, approximately 4 times the quantity synthesized by Azospirillum and LMG 2300, and Xanthomonas axonopodis pv. phaseoli XHFR-02.
brasilense Sp7 (59.77 μg/ml), which is known as a good IAA producer Nine strains from the genera Dyella, Kosakonia and Stenotrophomonas
and was used as a positive control in this analysis (Table 1). Phosphate were able to inhibit the four phytopathogens used in the analysis (Fig.
solubilization was less common among the strains; only 9 bacteria ob- S7). The strains LXo13, LEh18 and LTz4a from the genus
tained from wild legume nodules and 17 from trap-plant nodules ex- Stenotrophomonas were able to inhibit only R. solanacearum LMG 2300
hibited phosphate solubilization (Table 1). The ability to fix nitrogen (Fig. S7).
was found in only 13 strains, one isolated from wild legume nodules
and 12 from trap-plant nodules. The synthesis of siderophores was 4. Discussion
present in many strains (45), among which Kosakonia sp. LIz40b41 and
Pseudomonas sp. LIz40b42 (collected from the plant Desmodium sp., Many plants have developed in and become specialized for the
from Chiapas) exhibited a halo of 3.8 ± 0.01 cm. This halo was the different climates in Mexico, including the tropical-wet climate in the
largest among the isolates from this study and even larger than that southeastern area of the country (Rzedowski, 1991). This part of
produced by B. cenocepacia TAtl-371, which was 2.2 cm in diameter Mexico contains the highest number of plant species, especially in the

6
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

Fig. 3. Phylogenetic tree based in the analysis of the 16S rRNA sequence from several taxonomic groups which are close to the nodule associated bacteria isolated in
this study. The analysis was performed with the Bayesian inference applying the model GTR + I+G. The bar represents.

Fig. 4. Venn diagrams showing the isolates with plant growth


promotion (PGP) activity. A) Isolates from wild legume nodules
with any PGP activity (N = 43). B) Isolates from Phaseolus vulgaris
nodules used as a trap plant with PGP activity (N = 26). IAA,
indoleacetic acid production. PS, phosphate solubilization ac-
tivity. NF, nitrogen fixation activity. SP, siderophore synthesis.

7
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

Table 2 Few studies have analyzed NAB from wild legume nodules growing in
Amplification of nodC in bacteria isolated from legume nodules. Mexico. An exploration of M. pudica nodules collected in Chiapas
Species Strain nodC amplification showed a number of strains from the Enterobacter and Serratia genera
with several PGP activities (Sánchez-Cruz et al., 2019). In addition to
Alpha Beta this study, no other reports are known to exist. Therefore, in the present
proteobacteria proteobacteria study, we explored NAB thriving in the root nodules of legume species
Bacteria isolated from wild legume nodules growing in wild areas in the southern Mexico, where a tropical climate
Paraburkholderia sp. LIz1*1 – + prevails. Nodules were collected from wild legumes belonging to the
Cupriavidus sp. LEh21 – + taxa Acaciella sp., Arachis sp., Desmodium sp., M. pudica, M. quadrivalvis,
Cupriavidus sp. LEh25 – + M. galeottii and Leucaena sp. in Chiapas, Tabasco and Veracruz, states in
Paraburkholderia sp. LEh15*2 – +
Mexico known for their richness in legume species (Rzedowski, 1991),
Paraburkholderia sp. LEh19 – +
Labrys sp. LIt4 – + and from Mexico State, Morelos and Guerrero, where many species
from the genus Mimosa exist (Grether et al., 1996; Bontemps et al.,
Bacteria isolated from Phaseolus vulgaris used as trap plant 2016).
Brevibacillus sp. DeCh30-1a*3 + –
Most of the soils (37 out of 48) used in the plant-trap experiments
Burkholderia sp. CpTa8-5 + –
Cupriavidus sp. AcVe19-1a*4 – +
with P. vulgaris resulted in the formation of nodules. P. vulgaris is known
Cupravidus sp. AcVe19-6a*4 + – as a promiscuous plant that can form nodules with several bacterial
Rhizobium sp. AcVe20-2b*4 – + species, namely, Rhizobium, Ensifer and Paraburkholderia, which belong
Paraburkholderia mimosarum PAS44T – + to different groups in Proteobacteria (Martínez-Romero, 2003;
Paraburkholderia phymatum STM815T + +
Dall’Agnol et al., 2016). The genera isolated in this study with PGP
1, strain isolated from Mimosa sp. nodules (Chiapas, N 14° 58' 19.308" W 92° 9' activity were diverse, and Pseudomonas was frequently found in legume
26.315"). 2, Strain isolated from Leucaena sp. nodules (Chiapas, N 14° 52' 30.18" nodules from both wild and trap plants. This is consistent with other
W 92° 21' 24. 768"). 3, Strain isolated from P. vulgaris grown in soil collected studies in which Pseudomonas was abundant as NAB in legumes (Peix
where Desmodium sp. was growing (Chiapas, N 14º 57’ 46.44” W 92º 9’ et al., 2015). The identification of the isolates was performed with the
20.375”). 4, Strain isolated from P. vulgaris grown in soil collected where Acacia 16S rRNA gene, which in many cases and depending on the genus
sp. was growing (Veracruz, N 18º 29’ 39.516 W 95º 2’ 36.015”). analyzed it lacks resolution for the identification at the species level,
Nd, not determined. therefore, other techniques such as MLSA (multilocus sequence ana-
* The strain didn’t show any plant growth promotion activity, therefore are lysis) or whole genome sequencing is used to identify species or to
not listed in Table 1.
describe novel species (Rosselló-Mora and Amann, 2015; Chun et al.,
2018). In this study, the isolated bacteria were identified at the genus
states of Chiapas, Oaxaca and Veracruz (Rzedowski, 1991). The Faba-
level and when possible at species level. The specific genera found in
ceae family (Leguminoseae) is a taxonomic group that is present
the wild legume nodules were Achromobacter, Brevibacillus, Ensifer,
throughout Mexico. This family is divided into three subfamilies a)
Stenotrophomonas, Pseudoxanthomonas, Kosakonia, Dyella and Labrys,
Caesalpinioideae, b) Mimosoideae and c) Papilionoideae (Kenicer,
and the wild plant containing the largest number of bacterial genera
2005). The genus Mimosa from the subfamily Mimosoideae is found
was M. pudica (6 different genera). M. pudica forms nodules with several
extensively in Mexico, which is the second largest radiation center for
bacterial genera, such as Bradyrhizobium, Paraburkholderia, Cupriavidus
this genus (c.a. 100 species) (Grether et al., 1996; Simon et al., 2011).

Fig. 5. Phylogenetic analysis based in the nodC gene sequence of several species from Paraburkholderia (P), Cupriavidus (C), Rhizobium (R), Ensifer (E), Bradyrhizobium
(B), Mesorhizobium (M), and Azorhizobium caulinodans as an out group. The tree was calculated using the Bayesian method with the evolutive model HKY + G. The
bar represents the nucleotide differences between two sequences. Accession numbers of nodC described in this study are in Table S2.

8
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

and Rhizobium (Andrews and Andrews, 2017), but no information has been isolated from P. vulgaris and Pithecellobium sp. nodules, respec-
been reported about NAB in this plant. These associations might help tively, but both were uncapable to nodulate siratro and common bean
the plant adapt to different soils, given that M. pudica is widely dis- (Da Silva et al., 2016). Burkholderia sp. CpTa8-5, isolated in this study,
tributed in the world (https://www.gbif.org/species/2969284). In no- was positive for the amplification of nodC; it belongs to the Bcc, and the
dules from trap-plants, the specific genera isolated were Burkholderia, closest species are Burkholderia seminalis/Burkholderia territorii (99.78
Aeromonas, Moraxella, Herbaspirillum, Enterobacter and Marinococcus. It %), B. cepacia (99.71 %), and many other Burkholderia species. It is
has been reported that due to plant domestication (in this case, that of common to find high 16S rRNA similarity among the species of this
P. vulgaris, used as a trap-plant), interaction with endophytes di- group. A closer phylogenetic analysis comparing the 16S rRNA se-
minishes (Muresu et al., 2019), but this was not the case in the present quence and the atpD gene of Bcc species, the strains CpTa8-5 and
study; both wild legumes and trap plants contained several NAB. This AcTa6-5 showed that both are not related to any Burkholderia species
result may be due to the use of soil from wild plants rather than soil (data not shown). In the near future, we will analyze the genome se-
from agricultural fields for the trap experiments. Furthermore, the soil quence and test the strain CpTa8-5 in nodulation experiments to de-
from C. pumila used with bean as the trap-plant produced nodules with termine whether it is a novel Burkholderia species with the ability to
the largest number of bacterial genera (7). Crotalaria belongs to the nodulate.
Papilionoideae subfamily and is a native legume of Mexico, where it is Not much is known about gram-positive bacteria with the ability to
called “quelite” or more traditionally “chepil”, “chipil” or “chipilin”. It nodulate plant legumes (Ampomah and Huss-Danell, 2011), although
is an edible plant that is part of traditional agroecosystems but is cur- many have been found to be NAB. The Firmicutes Brevibacillus sp.
rently underused. Some studies have established that this plant controls DeCh30-1a tested positive for the amplification of nodC genes with
Helicobacter pylori-associated diseases (Gomez-Chang et al., 2018) or Alpha-proteobacteria primers. This is interesting, but more information
has antioxidant effects (Villa-Ruano et al., 2013). However, there is needs to be collected to learn about the possible nodulation activity of
little information on bacteria associated with C. pumila, although Rhi- this microorganism, such as genome sequencing to identify the genes
zobium has been isolated from Crotolaria pumila and Crotolaria mollicula involved in nodulation and performing inoculation experiments in
root nodules in France but not in Mexico (Martínez et al., 1985). No- beans and with plants where the soil was collected for the trap plant
tably, this plant species has been understudied but seems to be associate experiments, such as Desmodium sp., to determine their ability to no-
with a number of bacterial genera. Additionally, other genera in addi- dulate.
tion to Pseudomonas were found both in wild legumes and trap plant Among the species found in this study, some were interesting to us,
nodules, such as Cupriavidus, Acinetobacter, Bacillus, Paraburkholderia, such as Labrys sp. LIt4, this strain was close to L. neptuniae and was
Pseudarthrobacter and Rhizobium. Many of the genera isolated in this isolated from the root nodules of Acaciella sp. This strain was positive
study have been found previously in legume nodules (Abd-Alla et al., for the amplification of nodC, and it produces IAA and fixes nitrogen as
2018; Benjelloun et al., 2019; Busby et al., 2016; Cardoso et al., 2018; a free-living bacterium. Labrys is not widely recognized as a plant no-
De Meyer et al., 2015, 2018; Ndlovu et al., 2013; Noori et al., 2018; dulating bacteria; a strain identified as Labrys monachus performed ef-
Pandya et al., 2013; Rangel et al., 2017; Torche et al., 2016); however, fective symbiosis with Crotalaria spectabilis, but the genes involved in
to the best of our knowledge, this is the first time that Pseudox- this activity were not analyzed (Rangel et al., 2017). Therefore, it will
anthomonas, Aeromonas, Pseudarthrobacter and Marinococcus were found be interesting to more deeply study this bacterium in relation to its
in legume nodules. This adds information to this study of the bacterial formation of legume nodules in order to show that this genus might be
genera found in legume nodules. However, although in this study many further involved in symbiosis.
genera were found residing in the legume nodule, the isolation of Another species found in this study was P. caballeronis, described as
bacteria from this structure might be limited by the culture media used a nitrogen-fixing bacterium isolated from the rhizosphere of tomato
in any analysis, therefore it would be interesting to study the nodule cultivated in Mexico State and with the ability to effectively nodulate P.
microbiome from these understudied legumes in Mexico using high vulgaris (Martínez-Aguilar et al., 2013). However, when the genome
throughput sequencing methods (Hartman et al., 2017). sequence of the type strain was analyzed, no nodulation genes were
The potential capacity of the strains to nodulate was tested by found and the ability to nodulate was lost (Rojas-Rojas et al., 2017). In
amplifying the nodC gene. Only a few bacteria were positive for the this study, several P. caballeronis strains were isolated from the nodules
amplification of this gene. This suggests that most of the strains isolated on bean cultivated in soil from Leucaena sp. (Chiapas and Mexico State),
in this study are NAB and not typical symbionts; symbionts can be M. benthamii (Guerrero), C. pumila (Tabasco) and Acacia sp. (Veracruz).
uncultured, as has been observed when metagenomic analyses are None of these strains harbored the nodC gene; the genome of strain
performed in legume nodules (Martínez-Hidalgo and Hirsch, 2017). MbGu45-6, also isolated in this study and available from the Joint
Nonetheless, it was not surprising to find genera that contain this gene Genome Institute (JGI) as NZu45-6, lacks nodulation genes, like the
since, since genera such as Paraburkholderia, Cupriavidus, Rhizobium and other P. caballeronis strains available at JGI that we provided for the
Pseudomonas have been reported to be symbionts (Andrus et al., 2012; GEBA Project (Genomic Encyclopedia of Bacteria and Archaea, JGI-
Dall’Agnol et al., 2016; González et al., 2019). It was not expected to DOE-USA). It seems that this species may have had nodulation genes
find as potential symbionts some bacteria that are typically not known that were lost due to subcultivation in rich culture media or for another
to exhibit symbiont activity, namely the Beta-proteobacteria Bur- unidentified reason. The role of P. caballeronis as a NAB is interesting to
kholderia and the Firmicutes Brevibacillus. It is important to clarify the study because this species was isolated from the soils used in the trap
difference between Burkholderia sensu lato and Burkholderia sensu stricto; experiments from several regions of Mexico but not in nodules from
the former has been recently divided into the genera Burkholderia, wild legumes collected in the same locations. Apparently, this species
Paraburkholderia, Caballeronia, Robbsia, Mycetohabitans, Trinickia and may not be very competent at forming legume nodules in plants other
Pararobbsia (Estrada-de los Santos et al., 2018; Lin et al., 2019), and the than P. vulgaris. Currently, this species has been isolated in Mexico but
latter contains the Burkholderia cepacia complex (Bcc), the Burkholderia not in any other country.
pseudomallei group and a few plant pathogenic species (Rojas-Rojas Cupriavidus, another noteworthy genus, was found in this study. We
et al., 2018). The ability to nodulate has been identified in a number of previously reported that C. alkaliphilus, C. plantarum and other
species of Paraburkholderia and Trinickia (Estrada-de los Santos et al., Cupriavidus sp. from the northern Mexico (Tamaulipas state) are asso-
2018), but not specifically in Burkholderia. The species Burkholderia ciated with the rhizosphere of several non-legume plants (Estrada-de los
vietnamiensis, Burkholderia contaminans and Burkholderia lata are free- Santos et al., 2011, 2012, 2014), and from soil and sediment collected
living nitrogen fixers (Gillis et al., 1995; Da Silva et al., 2012) but are in two contaminated sites in Michoacán state (unpublished data).
not recognized as nodulating legumes; B. contaminans and B. lata have However, there are no reports of Cupriavidus species found in legumes

9
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

in Mexico. In this study, Cupriavidus sp. was isolated from wild legume activities that are important for promoting the plant growth.
nodules from Chiapas and a strain from trap-plant cultivated with soil
from Acacia sp. collected in Veracruz. The strains were close to C. al- Funding
kaliphilus (AcVe19-1a, AcVe19-6a, AcVe19-6b and AcVe19-1b), C. ox-
alaticus (LEh21 and LEh25) and C. metallidurans (LEh25b, LXo12c and This work was partially support by Instituto Politécnico Nacional,
LXo12a). From these strains, LEh21, LEh25, AcVe19-1a and AcVe19-6a Secretaría de Investigación y Posgrado (México) with the projects:
were positive for the amplification of nodC. However, it remains to be SIP2016-0077, SIP2017-0492, SIP2018-0117 and SIP2019-6674
confirmed whether the strains are the previously described Cupriavidus awarded to PES.
species or whether they represent novel species and symbionts. Nodu-
lation in Cupriavidus has not been shown before in C. alkaliphilus or in C. CRediT authorship contribution statement
oxalaticus. A group of C. alkaliphilus-like strains were found as sym-
bionts from Mimosa spp. growing in alkaline soils on Christmas Island, Erika Yanet Tapia-García: Conceptualization, Methodology,
Western Australia, but the taxonomy of these strains needs to be re- Validation, Formal analysis, Investigation, Data curation, Writing -
viewed in order to confirm whether they belong to C. alkaliphilus or to a original draft, Visualization. Verónica Hernández-Trejo:
novel Cupriavidus species (De Meyer et al., 2018). Methodology, Validation, Investigation. Joseph Guevara-Luna:
The phylogenetic analysis of nodC sequences showed that the strains Formal analysis. Fernando Uriel Rojas-Rojas: Methodology. Ivan
identified as Paraburkholderia, Cupriavidus and Labrys clustered with the Arroyo-Herrera: Methodology. Georgina Meza-Radilla:
nodC of these genera. However, Burkholderia and Brevibacillus nodC Methodology, Formal analysis. María Soledad Vásquez-Murrieta:
were placed among Rhizobium species. Since Burkholderia and Funding acquisition, Writing - review & editing. Paulina Estrada-de
Brevibacillus are not known as nodulating bacteria, it is possible that los Santos: Conceptualization, Validation, Formal analysis,
these strains may have acquired the symbiotic characteristic through Investigation, Resources, Data curation, Writing - original draft, Writing
the lateral transfer of symbiotic genes from alpha-rhizobia. - review & editing, Visualization, Supervision, Project administration,
All bacteria recovered from legume nodules were tested for PGP Funding acquisition.
activity. Rhizobia and NAB might act together as a community in the
nodule to promote plant growth, especially under environmental stress Declaration of Competing Interest
(Martínez-Hidalgo and Hirsch, 2017). Among the activities tested, the
synthesis of IAA was common among the isolates, with Steno- The authors declare no conflict of interest.
trophomonas sp. LXo13 and Rhizobium sp. Liz42a as the greatest pro-
ducers, synthesizing even more than A. brasilense, a typical IAA pro- Acknowledgements
ducer. The role of IAA in NAB may be related to improved nodulation
and plant growth when an NAB IAA producer strain is combined with a We thank Dr. Carlos Hugo Avendaño Arrazate (INIFAP-Rosario
symbiotic bacterium (Kumawat et al., 2019). It is possible that the IAA Izapa) for the logistical support in collecting legumes in Chiapas. We
functions as a signal in bacteria-bacteria communication (quorum are grateful to MSc Angélica Martínez Bernal (UAM-Iztapalapa) for
sensing) and between plant-bacteria communication (Spaepen et al., taxonomic identification of the legumes. We thank MSc. Tania Zacaria
2007). The second most common activity was the production of side- Vital for helping to collect plant samples in Guerrero. We are grateful to
rophores. Kosakonia sp. LIz40b41 and Pseudomonas sp. LIz40b2, were Angel Adalberto Cruz Alonzo for technical support. EYTG, JGL, FURR,
the greatest siderophore producers. Siderophores are molecules that IAH, and GMR were recipients of a Consejo Nacional de Ciencia y
bind Fe for leghemoglobin and help to increase the enzymatic activity, Tecnología. fellowship. EYTG, VHT, FURR and GMR were awarded a
and the water and mineral uptake of the plant (Ferchichi et al., 2019). BEIFI-IPN fellowship. MSVM and PES thank EDI, COFAA and SNI for
Moreover, siderophores can protect plants from pathogens (Martínez- support.
Hidalgo and Hirsch, 2017). It is possible that some of the siderophore
producer strains isolated in this study exhibit this activity. Another PGP Appendix A. Supplementary data
feature tested was phosphate solubilization. Kosakonia sp. LIz40b41
produced large solubilization halos, twice as larger as those produced Supplementary material related to this article can be found, in the
by P. tropica Ppe8, a species known for this activity (Caballero-Mellado online version, at doi:https://doi.org/10.1016/j.micres.2020.126522.
et al., 2007). It has been observed that co-inoculation of phosphate-
solubilizing bacteria with rhizobia increases nodulation, nitrogen fixa- References
tion and plant growth (Yasmeen and Bano, 2014). Phosphorus is im-
portant for nitrogen fixation because this activity requires large Abd-Alla, M.H., Bashandy, S.R., Nafady, N.A., Hassan, A.A., 2018. Enhancement of exo-
amounts of energy in the form of ATP (Pérez‐Fernández et al., 2017; polysaccharide production by Stenotrophomonas maltophilia and Brevibacillus para-
Stevens et al., 2019). The diversity of NAB found in this study and the brevis isolated from root nodules of Cicer arietinum L. and Vigna unguiculata L. (Walp.)
plants. Rend. Fis. Acc. Lincei 29, 117–129.
PGP activities performed by these bacteria may represent possible in- Akaike, H., 1974. A New Look at the Statistical Model of Identification. Selected Papers of
teractions occurring in the nodule among bacteria and with rhizobia. Hirotugu Akaike Springer, New York, NY, pp. 215–222.
Ampomah, O.Y., Huss-Danell, K., 2011. Genetic diversity of root nodule bacteria nodu-
lating Lotus corniculatus and Anthyllis vulneraria in Sweden. Syst. Appl. Microbiol. 34,
5. Concluding remarks 267–275.
Andrews, M., Andrews, M.E., 2017. Specificity in legume-rhizobia symbioses. Int. J. Mol.
This study revealed the presence of NAB belonging to Alpha-, Beta- Sci. 18, 705.
Andrus, A.D., Andam, C., Parker, M.A., 2012. American origin of Cupriavidus bacteria
and Gamma proteobacteria, Firmicutes and Actinobacteria in un- associated with invasive Mimosa legumes in the Philippines. FEMS Microbiol. Ecol.
explored legumes in Mexico. The majority of the strains were capable of 80, 747–750.
performing at least one PGP activity that could assist nodule develop- Ausubel, F.M., Brent, R., Kingston, R.E., More, D.D., Seidman, J.G., Smith, J.A., Struhl, K.,
1987. Current Protocols in Molecular Biology. John Wiley & Sons, Inc., New
ment by symbiotic bacteria. New bacterial genera were reported as
York, N.Y.
NAB, such as Pseudoxanthomonas, Aeromonas, Pseudarthrobacter and Avelar, F.P.A., Bomfeti, C.A., Soares, B.L., de Souza, M.F.M., 2012. Efficient nitrogen-
Marinococcus, and potential new symbionts such as Burkholderia from fixing Rhizobium strains isolated from Amazonian soils are highly tolerant to acidity
the B. cepacia complex and strains belonging to the Firmicutes and aluminum. World J. Microbiol. Biotechnol. 28, 1947–1959.
Benjelloun, I., Thami-Alami, I., Douira, A., Udupa, S.M., 2019. Phenotypic and genotypic
Brevibacillus were observed. This shows that many bacteria remain diversity among symbiotic and non-symbiotic bacteria present in chickpea nodules in
undiscovered in unexplored plants and soils, and they may perform Morocco. Front. Microbiol. 10, 1885.

10
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

Bontemps, C., Elliott, G.N., Simon, M.F., Dos Reis Junior, F.B., Gross, E., Lawton, Nicolau, Gillis, M., Van Van, T., Bardin, R., Goor, M., Hebbar, P., Willems, A., Segers, P., Kersters,
E.N., Loureiro, M.D.F., De Faria, S.M., Sprent, J.I., James, E.K., Young, P.W., 2010. K., Heulin, T., Fernandez, M.P., 1995. Polyphasic taxonomy in the genus Burkholderia
Burkholderia species are ancient symbionts of legumes. Mol. Ecol. 19, 44–52. leading to an emended description of the genus and proposition of Burkholderia
Bontemps, C., Rogel, M.A., Wiechmann, A., Mussabekova, A., Moody, S., Simon, M.F., vietnamiensis sp. nov. for N2-fixing isolates from rice in Vietnam. Int. J. Syst. Evol.
Moulin, L., Elliot, G.N., Lacercat-Didier, L., Dasilva, C., Grether, R., Camargo-Ricalde, Microbiol. 45, 274–289.
S.L., Chen, W., Sprent, J.I., Martínez-Romero, E., Young, J.P.W., James, E.K., 2016. Glickmann, E., Dessaux, Y., 1995. A critical examination of the specificity of the
Endemic Mimosa species from Mexico prefer alphaproteobacterial rhizobial sym- Salkowski reagent for indolic compounds produced by phytopathogenic bacteria.
bionts. New Phytol. 209, 319–333. Appl. Environ. Microbiol. 61, 793–796.
Bouckaert, R., Vaughan, T.G., Barido-Sottani, J., Duchêne, S., Fourment, M., Gomez-Chang, E., Uribe-Estanislao, G.V., Martínez-Martínez, M., Gálvez-Mariscal, A.,
Gavryushkina, A., Heled, J., Jones, G., Kühnet, D., De Maio, N., Matschiner, M., Romero, I., 2018. Anti-Helicobacter pylori potential of three edible plants known as
Mendes, F.K., Müller, N.F., Ogilvie, H.A., du Plessis, L., Popinga, A., Rambaut, A., Quelites in Mexico. J. Med. Food 21, 1150–1157.
Rasmussen, D., Siver, I., Suchard, M.A., Wu, C., Xie, D., Zhang, C., Stadler, T., González, A.H., Morales Londoño, D., Pille da Silva, E., Nascimento, F.X.I., de Souza, L.F.,
Drummond, A.J., 2019. BEAST 2.5: an advanced software platform for Bayesian da Silva, B.G., Canei, A.D., de Armas, R.D., Giachini, A.J., Soares, C.R.F.S., 2019.
evolutionary analysis. PLoS Comput. Biol. 15, 1–28. Bradyrhizobium and Pseudomonas strains obtained from coal‐mining areas nodulate
Busby, R.R., Rodriguez, G., Gebhart, D.L., Yannarell, A.C., 2016. Native Lespedeza species and promote the growth of Calopogonium muconoides plants used in the reclamation
harbor greater non-rhizobial bacterial diversity in root nodules compared to the of degraded areas. J. Appl. Microbiol. 126, 523–533.
coexisting invader, L. cuneate. Plant Soil 401, 427–436. Grether, R., Camargo, S.L., Martínez, A., 1996. Especies del género Mimosa
Caballero-Mellado, J., Onofre-Lemus, J., Estrada-de los Santos, P., Martínez-Aguilar, L., (Leguminosae) presentes en México. Bot. Sci. 58, 149–152.
2007. The tomato rhizosphere, an enviroment rich in nitrogen fixing Burkholderia Hartman, K., van der Heijden, M.G.A., Roussely-Provent, V., Walser, J.C., Schlaeppi, K.,
species with capabilities of interest for agriculture and bioremediation. Appl. 2017. Deciphering composition and function of the root microbiome of a legume
Environ. Microb. 73, 5308–5319. plant. Microbiome 5, 2. https://doi.org/10.1186/s40168-016-0220-z.
Cardoso, P., Alves, A., Silveira, P., Sá, C., Fidalgo, C., Freitas, R., Figueira, E., 2018. Ibáñez, F., Tonelli, M.L., Muñoz, V., Figueredo, M.S., Fabra, A., 2017. Bacterial en-
Bacteria from nodules of wild legume species: phylogenetic diversity, plant growth dophytes of plants: diversity, invasion mechanisms and effects on the host. In:
promotion abilities and osmotolerance. Sci. Total Environ. 645, 1094–1102. Maheswary, D.K. (Ed.), Endophytes: Biology and Biotechnology, Sustainable
Chávez-Ramírez, B., Kerber-Díaz, J.C., Acoltzi-Conde, M.C., Ibarra, J.A., Vásquez- Development and Biodiversity. Springer International Publishing, pp. 25–40.
Murrieta, M.S., Estrada-de los Santos, P., 2020. Inhibition of Rhizoctonia solani RhCh- Jain, D.K., Patriquin, D.G., 1985. Characterization of a substance produced by
14 and Pythium ultimum PyFr-14 by Paenibacillus polymyxa NMA1017 and Azospirillum which causes branching of wheat root hairs. Can. J. Microbiol. 31,
Burkholderia cenocepacia CACua-24: a proposal for biocontrol of phytopathogenic 206–210.
fungi. Microbiol. Res. 230, 126347. https://doi.org/10.1016/j.micres.2019.126347. Kenicer, G., 2005. In: In: Lewis, G., Schrire, B., MacKinder, B., Lock, M. (Eds.), Legumes of
Chun, J., Oren, A., Ventosa, A., Christensen, H., Ruiz Arahal, D., da Costa, M.S., Rooney, the World 62. Royal Botanic Gardens, Kew. Edinburgh J. Bot., pp. 195–196.
A.P., Yi, H., De Meyer, S., Trujillo, M.E., 2018. Proposed minimal standards for the Kumawat, K.C., Sharma, P., Sirari, A., Singh, I., Gill, B.S., Singh, U., Saharan, K., 2019.
use of genoma data for the taxonomy of prokaryotes. Int. J. Syst. Evol. Microbiol. 68, Synergism of Pseudomonas aeruginosa (LSE-2) nodule endophyte with Bradyrhizobium
461–466. https://doi.org/10.1099/ijsem.0.002516. sp. (LSBR-3) for improving plant growth, nutrient acquisition and soil health in
Clua, J., Roda, C., Zanetti, M., Blanco, F., 2018. Compatibility between legumes and soybean. World J. Microbiol. Biotechnol. 35, 47.
rhizobia for the establishment of a successful nitrogen-fixing symbiosis. Genes 9, 125. Lane, D.J., 1991. 16S/23S rRNA sequencing. In: Stackebrandt, E., Goodfellow, M. (Eds.),
https://doi.org/10.1128/AEM.71.11.7461-7471.2005. Nucleic Acid Techniques in Bacterial Systematics. John Wiley and Sons, New York,
Cordeiro, A.B., Ribeiro, R.A., Ferraz Helene, L.C., Hungría, M., 2017. Rhizobium esper- pp. 115–175.
anzae sp. nov., a N 2 -fixing root symbiont of Phaseolus vulgaris from Mexican soils. Lin, Q.H., Lv, Y.Y., Gao, Z.H., Qiu, L.H., 2019. Pararobbsia silviterrae gen. nov., sp. nov.,
Int. J. Syst. Evol. Microbiol. 67, 3937–3945. isolated from forest soil and reclassification of Burkholderia alpina as Pararobbsia al-
Da Silva, K., de Souza Cassetari, A., Lima, A.S., De Brandt, E., Pinnock, E., Vandamme, P., pina comb. nov. Int. J. Syst. Evol. Microbiol. https://doi.org/10.1099/ijsem.0.
de Souza Moreira, F.M., 2012. Diazotrophic Burkholderia species isolated from the 003932.
Amazon region exhibit phenotypical, functional and genetic diversity. Syst. Appl. López-López, A., Rogel, M.A., Ormeno-Orrillo, E., Martínez-Romero, J., Martínez-Romero,
Microbiol. 35, 253–262. E., 2010. Phaseolus vulgaris seed-borne endophytic community with novel bacterial
Da Silva Chaves, J., Baraúna, A.C., Mosqueira, C.A., Gianluppi, V., Zilli, J.É., Da Silva, K., species such as Rhizobium endophyticum sp. nov. Syst. Appl. Microbiol. 33, 322–327.
2016. Stylosanthes spp. from Amazon savanna harbour diverse and potentially ef- López-López, A., Rogel-Hernández, M.A., Barois, I., Ortiz-Ceballos, A.I., Ormeño-Orrillo,
fective rhizobia. Appl. Soil. Ecol. 108, 54–61. E., Martínez-Romero, E., 2012. Rhizobium grahamii sp. nov., from nodules of Dalea
Dall’Agnol, R.F., Plotegher, F., Souza, R.C., Mendes, I.C., dos Reis Junior, F.B., Béna, G., leporina, Leucaena leucocephala and Clitoria ternatea, and Rhizobium mesoamer-
Moulin, L., Hungria, M., 2016. Paraburkholderia nodosa is the main N2-fixing species icanum sp. nov. from nodules of Phaseolus vulgaris, sirtao, cowpea and Mimosa pu-
trapped by promiscuous common bean (Phaseolus vulgaris L.) in the Brazilian dica. Int. J. Syst. Evol. Microbiol. 62, 2264–2271.
‘Cerradão’. FEMS Microbiol. Ecol. 92, 1–13. Martínez, E., Pardo, M.A., Palacios, R., Miguel, A.C., 1985. Reiteration of nitrogen fixa-
Darriba, D., Taboada, G.L., Doallo, R., Posada, D., 2012. jModelTest 2: more models, new tion gene sequences and specificity of Rhizobium in nodulation and nitrogen fixation
heuristics and parallel computing. Nat. Methods 98, 772. in Phaseolus vulgaris. Microbiology 131, 1779–1786.
De Meyer, S.E., De Beuf, K., Vekeman, B., Willems, A., 2015. A large diversity of non- Martínez-Aguilar, L., Salazar-Salazar, C., Méndez, R.D., Caballero-Mellado, J., Hirsch,
rhizobial endophytes found in legume root nodules in Flanders (Belgium). Soil Biol. A.M., Vásquez-Murrieta, M.S., Estrada-de Los Santos, P., 2013. Burkholderia ca-
Biochem. 83, 1–11. balleronis sp. nov., a nitrogen fixing species isolated from tomato (Lycopersicon es-
De Meyer, S.E., Cnockaert, M., Moulin, L., Howieson, J.G., Vandamme, P., 2018. culentum) with the ability to effectively nodulate Phaseolus vulgaris. Antonie van
Symbiotic and non-symbiotic Paraburkholderia isolated from South African Lebeckia Leeuwenhoek 104, 1063–1071.
ambigua root nodules and the description of Paraburkholderia fynbosensis sp. nov. Int. Martínez-Hidalgo, P., Hirsch, A.M., 2017. The nodule microbiome: N2-fixing rhizobia do
J. Syst. Evol. Microbiol. 68, 2607–2614. not live alone. Phytobiomes 1, 70–82.
Estrada-de los Santos, P., Bustillos-Cristales, R., Caballero-Mellado, J., 2001. Burkholderia, Martínez-Romero, E., 2003. Diversity of Rhizboium-Phaseolus vulgaris symbiosis: over-
a genus rich in plant-associated nitrogen fixers with wide environmental and geo- view and perspectives. Plant Soil 252, 11–23.
graphic distribution. Appl. Environ. Microbiol. 67, 2790–2798. Martínez-Romero, E., Segovia, L., Mercante, F.M., Franco, A.A., Graham, P., Pardo, M.A.,
Estrada-de los Santos, P., Martínez-Aguilar, L., López-Lara, I.M., Caballero-Mellado, J., 1991. Rhizobium tropici sp. nov., a novel species nodulating Phaseolus vulgaris L. beans
2012. Cupriavidus alkaliphilus sp. nov., a new species associated with agricultural and Leucaenasp. Trees. Int. J. Syst. Bacteriol. 41, 417–426.
plants that grow in alkaline soils. Syst. Appl. Microbiol. 35, 310–314. Miranda-Sánchez, F., Rivera, J., Vinuesa, P., 2016. Diversity patterns of Rhizobiaceae
Estrada-de los Santos, P., Palmer, M., Chávez-Ramírez, B., Beukes, C., Steenkamp, E., communities inhabiting soils, root surfaces and nodules reveal a strong selection of
Briscoe, L., Khan, N., Maluk, M., Lafos, M., Humm, E., Arrabit, M., Cook, M., Gross, rhizobial partners by legumes. Environ. Microbiol. 18, 2375–2391.
E., Simon, M.F., Dos Reis Junior, F.B., Whitman, W.B., Shapiro, N., Poole, S.P., Muresu, R., Polone, E., Sulas, L., Baldan, B., Tondello, A., Delogu, G., Cappuccinelli, P.,
Hirsch, A.M., Venter, S.N., James, E.K., 2018. Whole genome analyses suggest that Alberghini, S., Benhizia, Y., Benhizia, H., Benguedouar, A., Mori, B., Calamassi, R.,
Burkholderia sensu lato contains two additional novel genera (Mycetohabitans gen. Dazzo, F.B., Squartini, A., 2008. Coexistence of predominantly nonculturable rhi-
nov., and Trinickia gen. nov.): implications for the evolution of diazotrophy and zobia with diverse, endophytic bacterial taxa within nodules of wild legumes. FEMS
nodulation in the Burkholderiaceae. Genes 9, 389. Microbiol. Ecol. 63, 383–400.
Estrada-de Los Santos, P., Vacaseydel-Aceves, N.B., Martínez-Aguilar, L., Cruz-Hernández, Muresu, R., Porceddu, A., Sulas, L., Squartini, A., 2019. Nodule-associated microbiome
M.A., Mendoza-Herrera, A., Caballero-Mellado, J., 2011. Cupriavidus and diversity in wild populations of Sulla coronaria reveals clues on the relative im-
Burkholderia species associated with agricultural plants that grow in alkaline soils. J. portance of culturable rhizobial symbionts and co-infecting endophytes. Microbiol.
Microbiol. 49, 867–876. Res. 221, 10–14.
Estrada-de Los Santos, P., Solano-Rodríguez, R., Matsumura-Paz, L.T., Vásquez-Murrieta, Nautiyal, C.S., 1999. An efficient microbiological growth medium for screening phos-
M.S., Martínez-Aguilar, L., 2014. Cupriavidus plantarum sp. nov., a plant-associated phate solubilizing microorganisms. FEMS Microbiol. Lett. 170, 265–270.
species. Arch. Microbiol. 196, 811–817. Ndlovu, J., Richardson, D.M., Wilson, J.R., Le Roux, J.J., 2013. Co‐invasion of South
Ferchichi, N., Toukabri, W., Boularess, M., Smaoui, A., Mhamdi, R., Trabelsi, D., 2019. African ecosystems by an Australian legume and its rhizobial symbionts. J. Biogeogr.
Isolation, identification and plant growth promotion ability of endophytic bacteria 40, 1240–1251.
associated with lupine root nodule grown in Tunisian soil. Arch. Microbiol. 201, Noori, F., Etesami, H., Zarini, H.N., Khoshkholgh-Sima, N.A., Salekdeh, G.H., Alishahi, F.,
1333–1349. 2018. Mining alfalfa (Medicago sativa L.) nodules for salinity tolerant non-rhizobial
Gardner, W.H., 1986. Water content. In: Klute, A. (Ed.), Methods of Soil Analysis, Part II. bacteria to improve growth of alfalfa under salinity stress. Ecotoxicol. Environ. Saf.
Physical and Mineralogical Methods, 2nd ed. Am. Soc. Agron., Madison, pp. 493–544 162, 129–138.
Agronomy Monograph. Pandya, M., Naresh Kumar, G., Rajkumar, S., 2013. Invasion of rhizobial infection thread

11
E.Y. Tapia-García, et al. Microbiological Research 239 (2020) 126522

by non-rhizobia for colonization of Vigna radiata root nodules. FEMS Microbiol. Lett. of Mimosa pudica with biotechnological potential. Microbiol. Res. 218, 76–86.
348, 58–65. Sarita, S., Sharma, P.K., Priefer, U.B., Prell, J., 2005. Direct amplification of rhizobial
Parker, M.A., 2002. Bradyrhizobia from wild Phaseolus, Desmodium, and Macroptilium nodC sequences from soil total DNA and comparison to nodC diversity of root nodule
species in Northern Mexico. Appl. Environ. Microbiol. 68, 2044–2048. isolates. FEMS Microbiol. Ecol. 54, 1–11.
Peix, A., Ramírez-Bahena, M.H., Velázquez, E., Bedmar, E.J., 2015. Bacterial associations Schwyn, B., Neilands, J.B., 1987. Universal chemical assay for the detection and de-
with legumes. Crit. Rev. Plant Sci. 34, 17–42. termination of siderophores. Anal. Biochem. 160, 47–56.
Pérez‐Fernández, M.A., Calvo‐Magro, E., Rodríguez‐Sánchez, J., Valentine, A., 2017. Segovia, L., Young, J.P., Martínez-Romero, E., 1993. Reclassification of american
Differential growth costs and nitrogen fixation in Cytisus multiflorus (L′ Hér.) Sweet Rhizobium leguminosarum biovar phaseoli type I as Rhizboium etli sp. nov. Int. J. Syst.
and Cytisus scoparius (L.) are mediated by sources of inorganic N. Plant Biol. 19, Bacteriol. 43, 374–377.
742–748. Selvakumar, G., Kundu, S., Anand, D.G., Yogesh, S.S., Hari, S.G., 2008. Isolation and
Rangel, W.D.M., de Oliveira Longatti, S.M., Ferreira, P.A.A., Bonaldi, D.S., Guimaraes, characterization of non-rhizobial plant growth promoting bacteria from nodules of
A.A., Thijs, S., Weyes, N., Vangrosveld, J., Moreira, F.M., 2017. Leguminosae native kudzu (Pueraria thunbergiana) and their effect on wheat seedling growth. Curr.
nodulating bacteria from a gold mine As-contaminated soil: multi-resistance to trace Microbiol. 56, 134–139.
elements, and possible role in plant growth and mineral nutrition. Int. J. Simon, M.F., Grether, R., de Queiroz, L.P., Särkinen, T.E., Dutra, V.F., Hughes, C.E., 2011.
Phytoremediation 19, 925–936. The evolutionary history of Mimosa (Leguminosae): toward a phylogeny of the sen-
Rincón-Rosales, R., Lloret, L., Ponce, E., Martínez-Romero, E., 2009. Rhizobia with dif- sitive plants. Am. J. Bot. 98, 1201–1221.
ferents symbiotic efficiencies nodulate Acaciella angustissima in Mexico, including Spaepen, S., Vanderleyden, J., Remans, R., 2007. Indole-3-acetic acid in microbial and
Sinorhizobium chiapanecum sp. nov. which has common symbiotic genes with microorganism-plant signaling. FEMS Microbiol. Rev. 31, 425–448.
Sinorhizobium mexicanum. FEMS Microbiol. Ecol. 67, 103–117. Stevens, G.G., Pérez-Fernádez, M.A., Morcillo, R.J., Kleinert, A., Hills, P.N., Brand, D.J.,
Rojas-Rojas, F.U., Tapia-García, E.Y., Maymon, M., Humm, E., Huntemann, M., Clum, A., Steenkamp, T., Valentine, A.J., 2019. Roots and nodules response differently to P
Pillay, M., Palaniappan, K., Varghese, N., Mikhailova, N., Stamatis, D., Reddy, T.B.K., starvation in the Mediterranean-type legume Virgilia divaricata. Front. Plant Sci.
Markowitz, V., Ivanova, N., Kyrpides, N., Woyke, T., Shapiro, N., Hirsch, A.M., 10, 73.
Estrada-de los Santos, P., 2017. Draft genome of Paraburkholderia caballeronis TNe- Thomas, G.W., 1996. Soil pH and soil acidity. In: Sparks, D.L. (Ed.), Methods of Soil
841T, a free-living, nitrogen-fixing, tomato plant-associated bacterium. Stand. Analysis. Chemical Methods. Soil Sci. Soc. of Am., Madison, pp. 475–490.
Genomic Sci. 12, 80. Torche, A., Benhizia, H., Rosselli, R., Romoli, O., Zanardo, M., Baldan, E., Alberghini, S.,
Rojas-Rojas, F.U., Salazar-Gómez, A., Vargas-Díaz, M.E., Vásquez-Murrieta, M.S., Hirsch, Tondello, A., Baldan, B., Benguedouar, A., Squartini, A., Benhizia, Y., 2016.
A.M., De Mot, R., Maarten, G.K., Ghequire, J., Ibarra, A., Estrada-de los Santos, P., Characterization of bacteria associated with nodules of two endemic legumes of
2018. Broad-spectrum antimicrobial activity by Burkholderia cenocepacia TAtl-371, a Algeria, Hedysarum naudinianum and H. perrauderianum. Ann. Microbiol. 64,
strain isolated from the tomato rhizosphere. Microbiology 164, 1072–1086. 1065–1071.
Román-Ponce, B., Zhang, Y.J., Vásquez-Murrieta, M.S., Sui, X.H., Chen, W.F., Padilla, Villa-Ruano, N., Zurita-Vásquez, G.G., Pacheco-Hernández, Y., Betancourt-Jiménez, M.G.,
J.C.A., Wang, E.T., 2016. Rhizobium acidisoli sp. nov., isolated from root nodules of Cruz-Durán, R., Duque-Bautista, H., 2013. Anti-Iipase and antioxidant properties of
Phaseolus vulgaris in acid soils. Int. J. Syst. Evol. Microbiol. 66, 398–406. 30 medicinal plants used in Oaxaca, México. Biol. Res. 46, 153–160.
Rosselló-Mora, R., Amann, R., 2015. Past and future species definitions for Bacteria and Wang, E.T., Rogel, M.A., García de los Santos, A., 1999. Rhizobium etli bv. mimosae, a
Archaea. Syst. Appl. Microbiol. 38, 209–216. https://doi.org/10.1016/j.syapm.2015. novel biovar isolated from Mimosa affinis. Int. J. Syst. Bacteriol. 49, 1479–1491.
02.001. Yan, J., Yan, H., Liu, X., Chen, W.F., Zhang, X.X., Verastegui-Valdes, M.M., Wang, E.T.,
Rzedowski, J., 1991. Diversidad y orígenes de la flora fanerogámica de México. Acta Bot. Han, X.Z., 2017. Rhizobium hidalgoense sp. nov., a nodule endophytic bacterium of
Mex. 14, 3–21. Phaseolus vulgaris in acid soil. Arch. Microbiol. 199, 97–104.
Sánchez-Cruz, R., Vázquez, I.T., Batista-García, R.A., Méndez-Santiago, E.W., del Rayo Yasmeen, S., Bano, A., 2014. Combined effect of phosphate-solubilizing microorganisms,
Sánchez-Carbente, M., Leija, A., Lira-Ruan, V., Hernández, G., Wong-Villarreal, A., Rhizobium and Enterobacter on root nodulation and physiology of soybean (Glycine
Folch-Mallol, J.L., 2019. Isolation and characterization of endophytes from nodules max L.). Commun. Soil Sci. Plant Anal. 45, 2373–2384.

12

You might also like