ArticlePDF Available

Abstract and Figures

Symptoms of leaf blight, stem canker, and pod rot were observed on T. cacao during a series of samplings conducted in several states of Malaysia from September 2018 to March 2019. The identity of the pathogen that was responsible for the diseases was determined using morphological characteristics, DNA sequences, and phylogenetic analyses of multiple genes, namely, internal transcribed spacer (ITS), elongation translation factor 1-alpha (tef1-α), β-tubulin (tub2), and RNA polymerase subunit II (rpb2). A total of 57 isolates recovered from diseased leaves of T. cacao (13 isolates), stems (20 isolates), and pods (24 isolates) showed morphological features that resembled Lasiodiplodia sp. The identity of the isolates was further determined up to the species level by comparing DNA sequences and phylogenetic analyses of multiple genes. The phylogenetic analysis of the combined dataset of ITS, tef1-α, tub2, and rpb2 elucidated that all of the isolates obtained were Lasiodiplodia theobromae as supported by 97% bootstrap value. The results of pathogenicity tests revealed L. theobromae as the causal pathogen of leaf blight, stem canker, and pod rot of T. cacao.
Pathogenicity of Lasiodiplodia theobromae on leaves, stems, and pods of Theobroma cacao. (A) Blighted leaf observed in the field, (B) Asymptomatic control inoculated leaf, (C) Irregular black lesions with yellow halo observed after 4 days of inoculation (D,E) The lesions enlarged after 6 and 9 days of inoculation, respectively, (F) Presence of conidiomata on the diseased area (red arrow), (G) Cankered stem observed in the field, (H) Asymptomatic control inoculated stem, (I–K) Black necrotic lesions observed on the inoculation sites after 7, 14, and 21 days of inoculation, respectively, (L) Black necrotic lesions extending upwards and downwards after 28 days of inoculation, (M) Black sunken lesion on the inoculation site, (N) Incision of the stem inoculated site showed reddish-brown to black necrotic lesion, (O) Formation of gummosis on the necrotic lesion, (P) Vertical section of control (left) and fungal inoculated stems (right) showed symptomless and dark brown to black necrotic lesion, respectively, (Q) Transverse section of control (below) and fungal inoculated stems (above) showed symptomless and necrotic lesion, respectively, (R) Rotted pod observed in the field showed external and internal rotting symptoms, (S) Asymptomatic control inoculated pod, (T) Brown to black lesions observed on the inoculation sites after 5 days of inoculation, (U) The lesions enlarged after 7 days of inoculation (V), The lesion rapidly expanded after 9 days of inoculation, (W) The inoculated pod completely covered by the fungal mycelia after 12 days of inoculation, (X) Presence of black conidiomata (red circle) on the fungal inoculated pod, (Y) Cross-section of fungal inoculated pod showed rotting of the internal tissue.
… 
This content is subject to copyright. Terms and conditions apply.
1
Vol.:(0123456789)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports
Lasiodiplodia theobromae
as a causal pathogen of leaf
blight, stem canker, and pod rot
of Theobroma cacao in Malaysia
Abd Rahim Huda‑Shakirah1, Nik Mohd Izham Mohamed Nor1, Latiah Zakaria1,
Yin‑Hui Leong2 & Masratul Hawa Mohd1*
Symptoms of leaf blight, stem canker, and pod rot were observed on T. cacao during a series
of samplings conducted in several states of Malaysia from September 2018 to March 2019. The
identity of the pathogen that was responsible for the diseases was determined using morphological
characteristics, DNA sequences, and phylogenetic analyses of multiple genes, namely, internal
transcribed spacer (ITS), elongation translation factor 1‑alpha (tef1-α), β‑tubulin (tub2), and RNA
polymerase subunit II (rpb2). A total of 57 isolates recovered from diseased leaves of T. cacao (13
isolates), stems (20 isolates), and pods (24 isolates) showed morphological features that resembled
Lasiodiplodia sp. The identity of the isolates was further determined up to the species level by
comparing DNA sequences and phylogenetic analyses of multiple genes. The phylogenetic analysis of
the combined dataset of ITS, tef1-α, tub2, and rpb2 elucidated that all of the isolates obtained were
Lasiodiplodia theobromae as supported by 97% bootstrap value. The results of pathogenicity tests
revealed L. theobromae as the causal pathogen of leaf blight, stem canker, and pod rot of T. cacao.
e cocoa tree (eobroma cacao) is an evergreen shrub that is recognized by several names, including kakaw,
pokok coklat, chocolate, cacao, koko, criollo, cacaoyer, and kakao1. Previously, T. cacao was classied under
Sterculiaceae family, before being reclassied as a member of Malvaceae. It is originated in the Neotropical
rainforest, particularly in the Amazon basin and on the Guyana plateau24. e word eobroma means “Food
of the Gods,” whereas cacao comes from the Mayans and Aztec languages, Kakaw and Cacahuatl, respectively5,6.
Furthermore, T. cacao is the recognized species among the 22 eobroma species that is commonly planted
beyond its natural range and have an economic value1,6. Besides T. cacoa, the other species of eobroma also
have economic value such as T. grandiorum in South America and T. bicolor in Mexico and Central America6.
Clone seedling is preferred for plantation over hybrid seedling in almost all cocoa-producing countries because
it will produce the same tree morphology, pod, and bean characteristics as the parent tree, where the clone tree
has greater pod bearing capacities, bigger and more uniform beans, richer butter content, withstand to pest and
pathogen attacks, and adaptable to a wide range of agro-climatic conditions1,7. e continued advancement of
Malaysias cocoa industry in the late 1970s and early 1980s resulted in the founding of the Malaysian Cocoa Board
(MCB) in 1989, which is overseen by the Ministry of Plantation Industries and Commodities. e Boards goal
was to grow Malaysia’s cocoa industry so that it could be incorporated in the global market, as well as to boost
the quality and performance of cocoa bean and downstream production8. Malaysia is now the leading country
in the cocoa grinding industry8.
In addition, cocoa and its products have various nutritional values owing to their rich amounts of alkaloids,
cardiac glycosides, catechin, enantiomer, epicatechin, avanol, methylxanthines, procyanidin B2, saponin, tan-
nins, and terpenoids9. Moreover, cocoa has several biological benets, including high antioxidant activity, blood
pressure reduction, anticancer activity, stress and depression reduction, reduced risk of heart attack and stroke,
cholesterol control, antiplatelet eect, and anti-inammatory activity1014.
eobroma cacao tree, similar to any other Malvaceae plants, has been shown to be fungus-prone. Among
the most important diseases aecting cacao in Malaysia are black pod rot, canker, and vascular streak dieback
(VSD), which aect the pod; trunk and stem; leaves and stems of the cacao tree, respectively1. Furthermore,
several previous studies on the diseases of T. cacao caused by fungal and fungal-like pathogens have been reported
OPEN
1School of Biological Sciences, Universiti Sains Malaysia, 11800 Minden, Penang, Malaysia. 2National Poison
Centre, Universiti Sains Malaysia, 11800 Minden, Penang, Malaysia. *email: masratulhawa@usm.my
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
Vol:.(1234567890)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
worldwide namely, Ceratobasidium theobromae15, Colletotrichum gloeosporioides6, Colletotrichum siamense16,17,
Colletotrichum theobromicola18, Colletotrichum tropicale17, Lasiodiplodia brasiliensis19, Lasiodiplodia pseudothe-
obromae17, Lasiodiplodia theobromae6,1925, Moniliophthora perniciosa26, Moniliophthora roreri27, Neofusicoccum
parvum28, Phytophthora palmivora6,25,29, and Phytophthora megakarya4,29.
In a series of samplings conducted from September 2018 to March 2019, the occurrences of leaf blight,
stem canker, and pod rots of T. cacao were observed in cocoa plantations in several states of Malaysia. From
observations during the sampling revealed the disease incidences of leaf blight, stem canker, and pod rots in
cocoa plantations were 15%, 20%, and 25%, respectively, which may reduce cocoa production. e diseased
samples were gathered and returned for further observation. erefore, the present study sought to nd the
causative agent of leaf blight, stem canker, and pod rot of T. cacao in Malaysia using morphological, molecular,
and pathogenicity analyses.
Results
Fungal isolation and morphological identication. In total, 57 fungal isolates were retrieved from dis-
eased leaves of T. cacao (13 isolates), stems (20 isolates), and pods (24 isolates). On PDA, the fungal isolates pro-
duced dense and fast-growing mycelia, white to pale greenish-gray colony and eventually becoming dark grayish
(Fig.1A). e pigmentation ranged from dark gray to black (Fig.1B). e conidiomata were solitary, globose
to subglobose, uniloculate, black, surrounded by dense grayish mycelia, and 3.32 ± 0.47 × 3.10mm ± 0.27mm
(mean ± standard deviation (SD)) (length (L) × width (W)) in size (Fig. 1C). e conidia were observed as
immature and mature conidia. Both immature and mature conidia were subovoid to ellipsoid-ovoid in shape,
with a broadly rounded apex and a tapering to the truncated base. e immature conidia were initially dou-
ble layered, hyaline, unicellular, and 25.0 ± 1.06 × 13.0 µm ± 0.48 µm (mean ± SD) (L × W) in size (Fig. 1D).
e mature conidia appeared light to dark brown color with typical striate formation, one-septate, and
25.7 ± 1.73 × 13.1µm ± 0.82µm (mean ± SD) (L × W) in size (Fig.1E). e conidiogenous cells were cylindrical,
hyaline, thin walled, holoblastic, and smooth. e structure of the paraphyses was aseptate and septate, with
rounded apex, hyaline, and cylindrical (Fig.1F). Based on the characterization of the morphological features
of the fungal isolates, it was tentatively identied as Lasiodiplodia sp., which is coherent with the morphology
described by Alves etal.30 and Phillips etal.31.
Figure1. Morphological characteristics of Lasiodiplodia sp. recovered from diseased leaves, stem, and pods
of eobroma cacao. (A) Upper view of the colony appearance, (B) Reverse view colony appearance, (C)
Conidiomata, (D) Immature conidia, (E) Mature conidia, (F) Conidiogenous cells and paraphyses. Scale bars:
(C) = 1mm; (DF) = 50µm.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
Vol.:(0123456789)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
Molecular identication and phylogenetic analysis. Molecular analysis of the sequences of ITS,
tef1-α, tub2, and rpb2 claried the species identication of all the 57 isolates of Lasiodiplodia sp. recovered from
T. cacao. BLAST searches in the GenBank database revealed that the isolates showed 98–100% sequence homol-
ogy to the KY473071 (ITS), JX464026 (tef1-α), EU673110 (tub2), and MT592333 (rpb2) of L. theobromae. A
multi-locus analysis was performed to explicate the phylogenetic positions of these L. theobromae isolates. To
construct the phylogenetic tree, the sequences of the isolates from the present study (57 isolates of L. theobro-
mae) were aligned with 38 reference isolates of Lasiodiplodia species and one outgroup taxon (Botryosphaeria
dothidea). Phylogenetic analysis revealed that the topologies of the ML trees generated from individual and con-
catenated genes (ITS, tef1-α, tub2, and rpb2) were similar (Figs.S1a–d and 2). e ML tree constructed from the
concatenated sequences conrmed that the phylogenetic positions of the 57 isolates from T. cacao were clustered
with the reference isolates of L. theobromae, supported by 97% bootstrap value (Fig.2). As a result, all the present
isolates were veried as L. theobromae by virtue of molecular identication and phylogenetic analysis.
Pathogenicity test. e pathogenicity analysis of 13, 20, and 24 fungal isolates on healthy leaves, stems,
and pods of T. cacao resulted in the production of typical symptoms of blight, canker, and rot, respectively as
observed in the elds (Fig.3A,G,R). ere were no visible symptoms produced on control points of leaves, stems,
and pods (Fig.3B,H,S).
Aer 4days of inoculation, the fungal inoculated leaves exhibited small irregular black lesions bounded
by yellow halos (Fig.3C). e lesions and yellowing areas enlarged gradually during the incubation period
(Fig.3D,E). Conidiomata formed on the inoculation site (Fig.3F). e lesion areas produced ranged from 3.0
to 4.6 cm2 (Table1). ere was no signicant dierence of lesion areas recorded among the tested isolates.
e fungal inoculated stems developed black necrotic lesions within the rst to the third week of inoculation
(Fig.3I–K). Aer 4weeks, the lesions extended longitudinally from the inoculation sites (Fig.3L). e incision
of the stem inoculated point displayed a reddish-brown to black necrotic lesion (Fig.3M,N). Formation of gum-
mosis on the necrotic lesion was also observed (Fig.3O). Vertical and transverse sections of control and fungal
inoculated stems showed symptomless and dark brown to black necrotic lesions, respectively (Fig.3P,Q). ere
were signicant dierences of lesion areas produced on the L. theobromae inoculated stems that ranged from
12 to 14 cm2 (Table1).
e fungal inoculated pods showed irregular brown to black lesions aer 5days of incubation (Fig.3T). As
the infection progressed, the lesions expanded and turned darker aer 7days of inoculation (Fig.3U). Aer
12days of inoculation, the lesions continued to expand, and the inoculated pods were completely colonized by
the fungal grayish mycelia (Fig.3V,W). Black conidiomata formed on the fungal inoculated pods (Fig.3X). A
cross-section of fungal inoculated pods showed rotting of the internal tissue (Fig.3Y). e lesion areas ranged
from 46.7 to 50.3 cm2 (Table1). e lesion areas recorded on the fungal inoculated pods were signicantly dif-
ferent compared to the control (Table1).
e repetition of the pathogenicity assessment yielded the same outcomes as the rst analysis. Kochs postu-
lates were achieved by reisolating the same fungal isolates from the symptomatic inoculated leaves, stems, and
pods of T. cacao and their identities were conrmed through morphological features.
Discussion
e present study identied L. theobromae isolates responsible to cause leaf blight, stem canker, and pod rot of T.
cacao in Malaysia based on the morphological features, sequence comparison, and phylogenetic analysis of four
genes (ITS, tef1-α, tub2, and rpb2). Fungi from genus Lasiodiplodia are cosmopolitan and belong to the Botry-
osphaeriaceae family, and most of the species can be primarily found in tropics and subtropics3133. e genus
consists of many phytopathogenic fungal species with widespread distribution33. Lasiodiplodia species responsible
to cause over 500 plant diseases, including fruit rot, root rot, collar rot, stem-end rot, dieback, canker, and leaf
necrosis32,3443. In Malaysia, Lasiodiplodia species have been attributed to various destructive diseases, such as
black rot of kenaf seeds44, leaf blight of Sansevieria trifasciata45, stem end-rot of Mangifera indica46, stem canker
on Jatropha curcas and Acacia mangium47,48, and fruit rot of mango and guava49,50. Apart from that, they can
act as secondary pathogens or endophytes, and they also can become pathogenic in response to a stressor34,36,40.
All the 57 fungal isolates recovered from diseased T. cacao in the present study was tentatively assigned as
Lasiodiplodia sp. based on their macroscopic and microscopic characteristics. According to Hyde etal.51, the
morphological approach has been widely used as the foundation for almost all studies of fungal taxonomy. Slip-
pers and Wingeld34 also stated that Botryosphaeriaceae members are easily recognized from most other fungi
through their colony appearance, aerial mycelium, and pigments, which can aid in the delimitation and rapid
identication. However, due to the signicant overlapping of key morphological characteristics among Lasiodip-
lodia species, clear-cut identication of the Lasiodiplodia isolates in the present study could not be achieved up
to the species level by using traditional morphological descriptions such as conidial shape30,40.
Attributable to unresolve identity of Lasiodiplodia isolates based on morphological characteristics that could
lead to uncertain and misleading results, phylogenetic analysis involving DNA sequences of multiple genes was
applied to delineate species boundaries. Consistent with previous studies that also highlighted the importance
of molecular work in dening Lasiodiplodia species34,39,40,52, the present study used several genes, namely, ITS,
tef1-α, tub2, and rpb2, to explicitly characterize Lasiodiplodia isolates. e ITS region has been proposed and
widely used in fungal taxonomic classication because of its straightforward amplication and it provides a
high probability of successful fungal recognition, with the barcoding dierence between inter- and intraspecic
variations53,54. Nonetheless, the ITS region lacks interspecies variety and may even be vague in the identication
of some fungi, thus the use of additional genes would provide better resolution in the phylogenetic analysis. Other
studies also showed that a single gene is incapable of determining species in the genus Lasiodiplodia, implying that
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
Vol:.(1234567890)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
Botryosphearia dothid ea CBS11547 6
Lasiodiplodia brasiliensis CBS115447
Lasiodiplodia brasiliensis CMM4015
98 Lasiodip lodia brasiliensis CSM11
NS2F
M3F
M4F
NS7F
NS8F
PP9F
PP11F
J13F
J15F
J16F
M19F
PE20F
PE22F
PP23F
K25F
K27F
S30F
PE31F
PE32F
S34F
S35F
97 PR36F
PR37F
PE39F
K41L
K42L
PR43L
PR44L
PE45L
PE46L
S47L
S48L
S49L
M50L
M51L
NS52L
NS53L
J54S
J55S
J56S
J57S
J58S
J59S
98 NS60S
NS61S
NS62S
M63S
M64S
S65S
S66S
PE67S
PE68S
PP69S
PP70S
PP71S
J72S
51 J73S
Lasiodiplodia theo bromae CBS164.69
Lasiodiplodia theo bromae CBS214.50
Lasiodiplodia theo bromae CMW13490
Lasiodiplodia theo bromae CMM4019
Lasiodiplodia theo bromae CSM57
Lasiodiplodia theo bromae M400
Lasiodiplodia brasiliensis CBS123095
100
99
Lasiodiplodia hormozga nensis CBS12470
9
100 Lasiodiplo dia hormozganensis CBS12470
8
100 Lasiod iplodia mahajangana CBS124925
Lasiodiplodia mahajang ana CBS124926
Lasiodiplodia viticola CBS1 28313
99 Lasiodiplodia viticola CBS128314
100 Lasiod iplodia iraniensis CBS214711
Lasiodiplo dia iraniensis
CB
S124710
91 Lasiodiplo dia iranensis CMW35881
99 Lasiodip lodia euphorbicola CMM3651
Lasiodiplodia euphorbicola CMW33268
Lasiodiplodia euphorbicola CMM3609
Lasiodiplodia citricola CBS1 24707
99 Lasiodiplo dia citricola CBS124706
100
Lasiodiplodia mediterran ea CBS137783
Lasiodiplodia mediterran ea CBS137784
100 Lasiodiplodia lignicola CBS134112
75 Lasiodiplodia lignicola MF
L
U
CC
110656
94 Lasiodiplodia pseudotheo bromae I46
Lasiodiplodia pseudo theobromae CBS116459
100
96 Lasiodiplodia pseudo theobromae
CB
S116460
Lasiodiplodia pseudo theobromae
CB
S130991
Lasiodiplodia crassisp ora CBS118741
Lasiodiplodia crassisp ora CBS125626
Lasiodiplodia crassisp ora CMW33262
98 Lasiodiplodia margar itacea CBS122519
Lasiodiplodia marg aritacea CBS138289
100 Lasiod iplodia margaritacea CBS138290
0.05
Figure2. e maximum likelihood (ML) tree was generated with 1000 bootstrap replications using the
Tamura-3-parameter model. e ML tree is inferred from concatenated sequence dataset of four genes (ITS,
tef1-α tub2, and rpb2). Bootstrap support values greater than 50% are pointed out at the nodes. Isolates in bold
represent isolates in the present study and Botryosphearia dothidea represents the outgroup. e bar indicates
the substitutions number per position.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
Vol.:(0123456789)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
Figure3. Pathogenicity of Lasiodiplodia theobromae on leaves, stems, and pods of eobroma cacao. (A) Blighted leaf observed in
the eld, (B) Asymptomatic control inoculated leaf, (C) Irregular black lesions with yellow halo observed aer 4days of inoculation
(D,E) e lesions enlarged aer 6 and 9days of inoculation, respectively, (F) Presence of conidiomata on the diseased area (red arrow),
(G) Cankered stem observed in the eld, (H) Asymptomatic control inoculated stem, (I–K) Black necrotic lesions observed on the
inoculation sites aer 7, 14, and 21days of inoculation, respectively, (L) Black necrotic lesions extending upwards and downwards aer
28days of inoculation, (M) Black sunken lesion on the inoculation site, (N) Incision of the stem inoculated site showed reddish-brown
to black necrotic lesion, (O) Formation of gummosis on the necrotic lesion, (P) Vertical section of control (le) and fungal inoculated
stems (right) showed symptomless and dark brown to black necrotic lesion, respectively, (Q) Transverse section of control (below)
and fungal inoculated stems (above) showed symptomless and necrotic lesion, respectively, (R) Rotted pod observed in the eld
showed external and internal rotting symptoms, (S) Asymptomatic control inoculated pod, (T) Brown to black lesions observed on the
inoculation sites aer 5days of inoculation, (U) e lesions enlarged aer 7days of inoculation (V), e lesion rapidly expanded aer
9days of inoculation, (W) e inoculated pod completely covered by the fungal mycelia aer 12days of inoculation, (X) Presence of
black conidiomata (red circle) on the fungal inoculated pod, (Y) Cross-section of fungal inoculated pod showed rotting of the internal
tissue.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
Vol:.(1234567890)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
Isolate code
aLesion area (cm2)
Leaf Stem Pod
K41L 3.3 ± 0.7b b– –
K42L 3.1 ± 0.1b– –
PR43L 3.3 ± 0.7b– –
PR44L 3.7 ± 1.0b– -
PE45L 3.3 ± 0.3b– –
PE46L 3.0 ± 0.3b– –
S47L 4.6 ± 1.2b– –
S48L 4.6 ± 1.2b– –
S49L 3.5 ± 1.0b– –
M50L 3.3 ± 0.4b– –
M51L 3.1 ± 0.3b– –
NS52L 4.0 ± 1.3b– –
NS53L 3.2 ± 0.6b– –
J54S 14 ± 0d
J55S 14 ± 0d
J56S 14 ± 0d
J57S 14 ± 0d
J58S 14 ± 0d
J59S 14 ± 0d
NS60S – 12.3 ± 0c
NS61S – 12.3 ± 0c
NS62S – 14 ± 0d
M63S – 13.1 ± 0cd
M64S – 13.1 ± 0cd
S65S 13.1 ± 0cd
S66S 13.1 ± 0cd
PE67S – 13.1 ± 0cd
PE68S – 13.1 ± 0cd
PP69S – 12.0 ± 0c
PP70S – 12.0 ± 0c
PP71S – 13.1 ± 0cd
J72S 13.1 ± 0cd
J73S 13.1 ± 0cd
NS2F – – 49.8 ± 5.3e
M3F – – 50.3 ± 3.5e
M4F – – 47.9 ± 4.0e
NS7F – – 49.2 ± 3.8e
NS8F – – 48.1 ± 4.4e
PP9F – – 48.0 ± 2.6e
PP11F – 47.1 ± 6.8e
J13F – – 49.8 ± 7.8e
J15F – – 47.8 ± 10.1e
J16F – – 49.9 ± 7.7e
M19F – – 46.7 ± 8.0e
PE20F – 48.3 ± 10.6e
PE22F – 46.7 ± 8.0e
PP23F – 49.4 ± 4.8e
K25F – – 48.3 ± 5.2e
K27F – – 47.8 ± 8.9e
S30F – – 47.2 ± 7.5e
PE31F – 50.8 ± 12.6e
PE32F – 49.2 ± 3.8e
S34F – – 46.8 ± 2.3e
S35F – – 46.2 ± 10.8e
PR36F – 49.2 ± 3.8e
Continued
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
Vol.:(0123456789)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
additional genes are required30,55. e tef1-α has become the marker of choice for fungal identication because
of its distinct polymorphisms among similar species and consists of non-orthologous copies of the gene that are
undetected in the genus56. e tub2 is another useful marker for delineating fungal species because it has fewer
obscure aligned regions and less homoplasy across genera57. e rpb2 gene which codes for the second-largest
protein subunit in fungi is a highly preserved single-copy gene54.
According to the results of phylogenetic analysis, it can be inferred that single gene analyses of ITS, tub2,
and rpb2 are unable to resolve the identity of Lasiodiplodia isolates in the present study (Fig.S1a,c,d). ose
phylogenetic trees displayed that L. theobromae was grouped with L. brasiliensis and L. hormozganensis. On the
contrary, phylogenetic analysis of tef1-α sequences was able to dierentiate isolates in the present study with
other species of Lasiodiplodia by clustering them with several reference sequences of L. theobromae from the
GenBank database with only 64% bootstrap value (Fig.S1b). Owing to the fact that single gene analysis could not
accurately identify the Lasiodiplodia isolates in the present study, the combination of ITS, tef1-α, tub2, and rpb2
sequences was used for better characterization. e phylogenetic inferences based on multiple gene sequences
revealed that the present isolates were grouped with L. theobromae with a higher bootstrap value (97%) (Fig.2).
e nding has been proven that phylogenetic analysis based on multigene provided robust resolution with
clear-cut fungal identity. is is in line with the ndings of Cruywagen etal.52.
Lasiodiplodia theobromae was conrmed to be the causal pathogen of leaf blight, stem canker, and pod rot of
T. cacao in Malaysia. In 1895, L. theobromae was rstly described and reported to cause minor charcoal rot on
cocoa in Ecuador31. Besides charcoal rot, L. theobromae was also reported to cause dieback on T. cacao since the
late 1980s20. In Malaysia, documentations of relationship between L. theobromae and T. cacao are still limited.
e present study represents the rst report of leaf blight, stem canker, and pod rot of T. cacao caused by L. theo-
bromae. Several studies have also found the incidence of L. theobromae causing foliar diseases in a wide range
of hosts, including Camellia sinensis42, Catasetum mbriatum58, Cocos nucifera59,60, Kadsura longipedunculata61,
and S. trifasciata45. Moreover, the present study also revealed the ability of L. theobromae isolates to cause stem
canker of T. cacao. Asman etal.24, previously reported L. theobromae as a causal agent of dieback and stem
canker of cocoa by demonstrating internal discoloration with visible brown streaks in the vascular cambium.
Furthermore, L. theobromae has been associated with cocoa dieback in Cameroon, India, and Venezuela1921. It
also responsible to cause dieback and stem canker on a number of plants, such as American ash (Fraxinus ameri-
cana)62, blueberry bushes (Vaccinium spp.)63, strawberry (Fragaria × ananassa)41, mango (M. indica)64, cashew
(Anacardium occidentale)65, sacha inchi (Plukenetia volubilis)66, Persian lime (Citrus latifolia)67, and grapevine
(Vitis vinifera)68. In addition to infecting the leaf and stem, cocoa pod was also found to be susceptible to L.
theobromae infection by showing rot symptoms. Several studies reported the occurrence of pod rot of T. cacao
caused by L. theobromae6,22,25. Other pathogens were also identied to cause the same disease on the cocoa pod,
namely C. gloeosporioides6, C. siamense17, C. tropicale17, L. pseudotheobromae17, N. parvum25, P. palmivora6,25,29,
and P. megakarya4,29. From the pathogenicity tests, isolates of L. theobromae required wound to initiate infection
and colonization on the host plant. Other studies have found that fungi from Botryosphaeriaceae can invade
plants via endophytic conquest, injuries, seed-to-seedling conquest, contaminated soil, and insect infestation34,36.
In conclusion, the current study emphasized the rst report of L. theobromae as a causal pathogen of leaf
blight, stem canker, and pod rot of T. cacao in Malaysia. e pathogen was identied using morphological fea-
tures supported by multigene DNA sequences and phylogenetic inference. e valid and precise identication
of phytopathogen is critical for quarantine purpose and disease management strategies.
Materials and methods
Collecting samples and isolating fungi. From September 2018 to March 2019, sampling was conducted
during rainy season in several states of Malaysia, including Johor, Kedah, Melaka, Negeri Sembilan, Perak, Perlis,
Pulau Pinang, and Selangor (Fig.4). e sampling sites and sampling activities were approved by the MCB com-
ply with relevant institutional, national, and international guidelines and legislation. During the sampling, 50
blighted leaves, cankered stems, and rotted pods of T. cacao from the Koko Mardi (KM) clone were collected. e
clone was used in the study because of its wide cultivation in Malaysia which showed susceptibility to a number
of fungal diseases. Symptomatic leaves showed blighted symptoms, including circular to irregular blackish lesion
surrounded by a yellow halo. e cankered stems were characterized as irregular blackish lesion, sometimes
accompanied by gummosis on the disease area, expanded longitudinally, and internally became reddish-brown.
e rotted pods were associated with dark brown to blackish lesions on the pods that eventually expanded and
rotted.
Table 1. Lesion area produced on the leaves, stems and pods of eobroma cacao inoculated with
Lasiodiplodia theobromae. a Means ± standard deviation followed by dierent letters are signicantly dierent
(p < 0.05) according to Tukey’s test. b Not applicable.
Isolate code
aLesion area (cm2)
Leaf Stem Pod
PR37F – 49.6 ± 12.8e
PE39F – 50.5 ± 7.1e
Control 0a0a0a
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
Vol:.(1234567890)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
e diseased and healthy margins of samples were cut into small pieces for fungal isolation. e small pieces
of samples were surface-sterilized in 70% ethanol (C2H5OH) and 1% sodium hypochlorite (NaOCl) separately
for 3min. e samples were then rinsed in sterile distilled water three times in succession for 1min each. e
sterilized sample was blotted dry on sterile lter paper, transferred onto potato dextrose agar (PDA), and incu-
bated at 25°C ± 2°C for 3–5days. Pure cultures of fungal isolates obtained from single spore isolation were used
for morphological and molecular assessments.
Morphological identication. In the present study, the fungal isolates obtained were provisionally exam-
ined based on morphological features, specically macroscopic and microscopic characteristics. Colony appear-
ance and pigmentation were observed at the macroscopic level. Under a dissecting microscope, the structure of
the conidiomata was observed and photographed (EZ4, Leica Microsystem, Germany). e microscopic features
such as conidia, conidiogenous cells, and paraphyses were observed using a light microscope (CX41, Olympus,
Japan) and a camera (KY-F55BE, JVC, Japan). e average size of 30 randomized conidia was measured and
recorded. Each fungal isolate was cultured onto carnation leaf agar (CLA) and incubated at 25°C ± 2°C for
7days to observe the structures of conidiomata, conidia, conidiogenous cells, and paraphyses.
Molecular identication and phylogenetic analysis. To corroborate the identity of the fungal iso-
lates of the present study, molecular identication and characterization was carried out. e fungal isolates
were cultured in potato dextrose broth (PDB) and subjected to incubation at 25°C ± 2°C for 5 to 7days. e
mycelia that grew on the surface of PDB were collected, placed on the sterile lter paper (Whatman No. 1),
and le to dry for 10min. Using a sterile mortar and pestle, the dried mycelia were ground to a ne powder in
liquid nitrogen. en, 0.05g of the ne powdered mycelia was placed in a 1.5ml microcentrifuge tube for DNA
extraction. e InnuPREP Plant DNA kit (Analytik Jena, Germany) was used to extract DNA by referring to the
manufacturer’s protocols. For amplication of internal transcribed spacer (ITS), elongation translation factor
1-alpha (tef1-α), β-tubulin (tub2), and RNA polymerase subunit II (rpb2), primer pairs of ITS1 (TCC GTA GGT
GAA CCT GCG G)/ITS4 (TCC TCC GCT TAT TGA TAT GC)69, EF1-688F (CGG TCA CTT GAT CTA CAA GTGC)/
EF1-1251R (CCT CGA ACT CAC CAG TAC CG)30, Bt2a (GGT AAC CAA ATC GGT GCT GCT TTC )/Bt2b (ACC
CTC AGT GTA GTG ACC CTT GGC )70, and rpb2-LasF (GGT AGC GAC GTC ACT CCT )/rpb2-LasR (GCG CAA
ATA CCC AGA ATC AT)52 were adopted, respectively. A reaction mixture of 50µl was prepared by adding 8µl of
green buer (Promega, USA), 8µl of MgCl2 (Promega, USA), 1µl of deoxynucleotide triphosphate polymerase
(dNTP) (Promega, USA), 8µl of each primer (Promega, USA), 0.3µl of Taq polymerase (Promega, USA), 1µl of
genomic DNA, and sterile distilled water to obtain a total volume of 50µl. e following conditions were used
in the polymerase chain reaction (PCR) with the MyCycler™ ermal Cycler (Bio-rad, Hercules, USA): Initial
denaturation at 95°C for 7min (ITS)/5min (tef1-α and tub2)/2min (rpb2), then 25 cycles (ITS)/30 cycles (tef1-α
and tub2)/35 cycles (rpb2) of denaturation at 94°C for 1min (ITS)/30s (tef1-α, tub2, and rpb2), annealing at
50°C for 1min (ITS)/55°C for 45s (tef1-α and tub2)/54°C for 30s (rpb2), extension at 72°C for 1min (ITS
and rpb2)/90s (tef1-α and tub2), and nal extension at 72°C for 10min (ITS, tef1-α, and tub2)/8min (rpb2).
Figure4. Sampling sites of diseased eobroma cacao in several states of Malaysia.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
9
Vol.:(0123456789)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
e PCR products were electrophoresed for 90min at 80V and 400mA in a 1.0% agarose gel (Promega, USA)
containing FloroSafe DNA stain (First Base) in a 1.0× Tris–borate EDTA buer. e Bio-Rad Molecular Imager®
Gel Doc™ XR System and Bio-Rad Quantity One® Soware were used to view and photograph the gel. e size of
the amplied PCR products was determined using a 100bp GeneRulers™ DNA ladder (ermo Scientic, USA).
e PCR products were sent to the First BASE Laboratories Sdn Bhd in Seri Kembangan, Malaysia, for DNA
purication and sequencing.
e sequences obtained were compared, and phylogenetic analysis was performed using the Molecular
Evolutionary Genetic Analysis (MEGA7) soware71. e nucleotide homogeneity of the resulting consensus
sequences was assessed by comparing with other sequence data in the GenBank database using Basic Local
Alignment Search Tools (BLAST) (https:// blas t. nc b i . nlm. nih. gov/ Blast. cgi). All sequences obtained were submit-
ted to the GenBank database. Table2 lists the sequences from the present study and the reference isolates used
for phylogenetic analysis. e phylogenetic classication of the isolates from the present study was performed
by analyzing the combination of multi-sequence alignments of fungal isolates and reference isolates in MEGA7
using the maximum likelihood (ML) method. e ML tree of combined genes was constructed using the Tamura
3-parameter model72 and 1000 bootstrap replicates73.
Pathogenicity tests. A total of 57 fungal isolates were assessed for pathogenicity on leaves (13 isolates),
stems (20 isolates), and pods (24 isolates) of T. cacao using KM clone. e 1-year-old healthy seedlings of T.
cacao grown using clay loam soil with a pH of 6.5–7 in polythene bags; and healthy mature pods (5months old
and 17cm in size) taken from 3-year-old trees were purchased from the MCB. e seedlings were placed in the
plant house of the School of Biological Sciences, Universiti Sains Malaysia (USM) at a temperature of 26°C to
32°C.
A fungal mycelial plug used as an inoculum was prepared from a 7-day-old PDA culture using a sterile cork
borer (5mm diameter). For control, the PDA plugs without fungal mycelia were prepared from the blank PDA
using the same methods. Pathogenicity tests for all fungal isolates were performed twice. e tests were carried
out on 84 healthy attached young leaves (84 seedlings), 126 stems (126 seedlings), and 150 detached pods of T.
cacao. e targeted plant parts were surface-sterilized with 70% ethanol prior to inoculation.
To inoculate 13 fungal isolates on leaves of T. cacao, a total of 84 healthy leaves (78 for the fungal treatment
and six for the control) from 84 seedlings of T. cacao were used for two pathogenicity tests. Each surface-sterilized
leaf was aseptically pricked at one point with a sterile toothpick represented a replicate. For each pathogenicity
test, three replicates were performed for each fungal isolate, using three dierent leaves from three dierent
seedlings. Controls were performed in the same ways but treated with the blank PDA plugs. A sterile scalpel
was used to inoculate control and mycelial plugs onto the control and treatment points, respectively. e plugs
were wrapped in sterile cotton wool and xed to the leaf with cellophane tape to avoid dryness. Each inoculated
leaf was covered in a sterile zip lock bag. e inoculated seedlings were kept in the plant house of the School of
Biological Sciences, USM for 9days at temperatures ranging from 26 to 32°C.
A total of 126 healthy stems of T. cacao (126 seedlings) were used to inoculate 20 fungal isolates for twice
pathogenicity tests. A small wound (0.5cm) was created on the sterilized surface of each stem by removing the
bark with a sterile scalpel. For each pathogenicity test, three wounded stems from three dierent seedlings were
used to inoculate each fungal isolate, representing triplicates. Control was treated similarly using the blank
PDA plugs. Using a sterile scalpel, the mycelial and control plugs were placed on the wounded points, with the
mycelium positioned towards the cambium. e moisture of the plugs was maintained by wrapping in sterilized
cotton and sealing with paralm. All the inoculated seedlings were incubated in the plant house of the School
of Biological Sciences, USM at temperatures ranging from 26 to 32°C.
Twice pathogenicity tests conducted on healthy detached cocoa pods involved 150 pods (144 for the fungal
treatment and six pods for the control). Control and fungal treatments were inoculated on dierent pods to avoid
symptoms overlapping if both were performed on the same pods. For each pathogenicity test, a wound point was
created on the three dierent pods for each fungal isolate by piercing the pod surface with a sterile cork borer.
en, 5mm mycelial plugs with the mycelium facing the surface of the pods were placed on the wounded points.
e three control pods were treated in the same way but using the blank PDA plugs. To retain moisture, all the
plugs were wrapped with sterilized cotton wool and the cotton was xed with cellophane tape. e inoculated
cocoa pods were incubated for 12days at 25°C ± 2°C in sterilized trays and covered with transparent plastic to
maintain humidity.
e area of the lesion developed on the inoculated leaves, stems, and pods of T. cacao was measured using
grid paper adopted by Parker etal.74 with slight modications. e area of diseased lesion was calculated by
multiplying the number of small squares covering the lesion with the value calculated for one small square. Dif-
ferences in lesion area were evaluated using the one-way method ANOVA and means were compared with the
Tukey’s test (p < 0.05) using the soware IBM SPSS Statistics version 26. To conrm Koch’s postulates, fungi from
symptomatic inoculated leaves, stems, and pods of T. cacao were reisolated and reidentied using morphological
characteristics.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
10
Vol:.(1234567890)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
Species Isolate Host Location
GenBank accession number
ReferencesITS tef1-α tub2 rpb2
Lasiodiplodia brasiliensis CBS123095 eobroma cacao Cameroon MT587423 MT592135 MT592615 MT592309 Zhang etal.75
L. brasiliensis CBS115447 Psychotria tutcheri Hong Kong MT587422 MT592134 MT592614 MT592308 Zhang etal. 75
L. brasiliensis CMM4015aMangifera indica Brazil JX464063 JX464049 MT592614 MT592308 Marques etal.76
L. brasiliensis CSM11 eobroma cacao Venezuela MF436018 MF436006 MF435998 MT592308 Mohali-Castillo and
Stewart19
Lasiodiplodia citricola CBS124707aCitrus sp. Iran GU945354 GU945340 KU887505 KU696351 Cruywagen etal.52;
Abdollahzadeh etal.55
L. citricola CBS124706 Citrus sp. Iran GU945353 GU945339 KU887504 KU696350 Cruywagen etal.52;
Abdollahzadeh etal.55
Lasiodiplodia crassispora CBS118741aSantalum album Australia DQ103550 DQ103557 KU887506 KU696353 Cruywagen etal.52
L. crassispora CBS125626 Vitis vinifera South Africa MT587424 DQ103557 MT592617 MT592312 Zhang etal.75
L. crassispora CMW33262 Adansonia sp. Unknown KU887068 DQ103557 KU887426 KU887364 Cruywagen etal.52
Lasiodiplodia euphor-
biicola CMM3609aJatropha curcas Brazil KF234543 KF226689 KF254926 KU887367 Machado etal.77
L. euphorbiicola CMM3651 Jatropha curcas Brazil KF234553 KF226711 KF254937 KU887367 Machado etal.77
L. euphorbiicola CMW33268 Adansonia sp. Unknown KU887131 KU887008 KU887430 KU887367 Cruywagen etal.52
Lasiodiplodia hormoz-
ganensis CBS124709aOlea sp. Iran GU945355 GU945343 KU887515 KU696361 Cruywagen etal.52;
Abdollahzadeh etal.55
L. hormozganensis CBS124708 Mangifera indica Iran GU945356 GU945344 KU887514 KU696360 Cruywagen etal.52;
Abdollahzadeh etal.55
Lasiodiplodia iraniensis CBS124710aSalvadora persica Iran GU945348 GU945336 KU887516 KU696363 Cruywagen etal.52;
Abdollahzadeh etal.55
L. iraniensis CBS124711 Juglans sp. Iran GU945347 GU945335 KU887517 KU696362 Cruywagen etal.52;
Abdollahzadeh etal.55
L. iraniensis CMW35881 Adansonia sp. Unknown KU887092 KU886970 KU887464 KU887388 Cruywagen etal.52
Lasiodiplodia lignicola CBS134112aDead wood ailand JX646797 KU887003 JX646845 KU696364 Cruywagen etal.52; Liu
etal.78
L. lignicola MFLUCC110656 Dead wood ailand JX646798 KU887003 JX646846 KU696364 Cruywagen etal.52; Liu
etal.78
Lasiodiplodia mahajan-
gana CBS124925aTerminalia catappa Madagascar FJ900595 FJ900641 KU887518 KU696365 Cruywagen etal.52;
Begoude etal.79
L. mahajangana CBS124926 Terminalia catappa Madagascar FJ900596 FJ900642 KU887519 KU696366 Cruywagen etal.52;
Begoude etal.79
Lasiodiplodia marga-
ritacea CBS122519aAdansonia gibbosa Australia EU144050 EU144065 KU887520 KU696367 Cruywagen etal.52
L. margaritacea CBS138289 Combretum elae-
agnoides Namibia KP872320 KP872349 KP872379 KP872429 Zhang etal.75
L. margaritacea CBS138290 Combretum collinum Zambia KP872321 KP872350 KP872380 KP872430 Zhang etal.75
Lasiodiplodia mediter-
ranea CBS137783aQuercus ilex Italy KJ638312 KJ638331 KU887521 KU696368 Cruywagen etal.52;
Linaldeddu etal.80
L. mediterranea CBS137784 Vitis vinifera Italy KJ638311 KJ638330 KU887522 KU696369 Cruywagen etal.52;
Linaldeddu etal.80
Lasiodiplodia pseudothe-
obromae CBS116459aGmelina arborea Costa Rica EF622077 EF622057 EU673111 KU696376 Alves etal.30; Phillips
etal.81
L. pseudotheobromae CBS116460 Acacia mangium Costa Rica MT587433 MT592145 KU198428 MT592322 Zhang etal.75
L. pseudotheobromae CBS130991 Mangifera indica Egypt MT587433 MT592145 MT592629 MT592325 Zhang etal.75
L. pseudotheobromae I46 eobroma cacao Puerto Rico MK693211 MK693707 MK693702 KU696376 Serrato-Diaz etal.17
Lasiodiplodia theo-
bromae CBS164.69aFruit on coral reef coast Indonesia: New Guinea AY640255 AY640258 EU673110 KU696383 Cruywagen etal.52
L. theobromae CBS214.50 Cajanus cajan India MT587440 MT592152 MT592637 MT592333 Zhang etal.75
L. theobromae CMW13490 Eucalyptus urophylla Venezuela: Acarigua KY473071 KY473019 KY472962 KY472888 Mehl etal.82
L. theobromae CMM4019 Mangifera indica Brazil JX464096 JX464026 EU673110 KU696383 Marques etal.76
L. theobromae CSM57 eobroma cacao Venezuela MF436029 MF436017 MF435999 KU696383 Mohali-Castillo and
Stewart19
L. theobromae M400 eobroma cacao USA: Puerto Rico MN446021 MN536705 MN536694 KU696383 Puig etal.25
L. theobromae NS2F eobroma cacao Malaysia: Negeri
Sembilan OL831055 OL863319 OL863262 OL863376 is study
L. theobromae M3F eobroma cacao Malaysia: Melaka OL831056 OL863320 OL863263 OL863377 is study
L. theobromae M4F eobroma cacao Malaysia: Melaka OL831057 OL863321 OL863264 OL863378 is study
L. theobromae NS7F eobroma cacao Malaysia: Negeri
Sembilan OL831058 OL863322 OL863265 OL863379 is study
L. theobromae NS8F eobroma cacao Malaysia: Negeri
Sembilan OL831059 OL863323 OL863266 OL863380 is study
Continued
Content courtesy of Springer Nature, terms of use apply. Rights reserved
11
Vol.:(0123456789)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
Species Isolate Host Location
GenBank accession number
ReferencesITS tef1-α tub2 rpb2
L. theobromae PP9F eobroma cacao Malaysia: Pulau Pinang OL831060 OL863324 OL863267 OL863381 is study
L. theobromae PP11F eobroma cacao Malaysia: Pulau Pinang OL831061 OL863325 OL863268 OL863382 is study
L. theobromae J13F eobroma cacao Malaysia: Johor OL831062 OL863326 OL863269 OL863383 is study
L. theobromae J15F eobroma cacao Malaysia: Johor OL831063 OL863327 OL863270 OL863384 is study
L. theobromae J16F eobroma cacao Malaysia: Johor OL831064 OL863328 OL863271 OL863385 is study
L. theobromae M19F eobroma cacao Malaysia: Melaka OL831065 OL863329 OL863272 OL863386 is study
L. theobromae PE20F eobroma cacao Malaysia: Perak OL831066 OL863330 OL863273 OL863387 is study
L. theobromae PE22F eobroma cacao Malaysia: Perak OL831067 OL863331 OL863274 OL863388 is study
L. theobromae PP23F eobroma cacao Malaysia: Pulau Pinang OL831068 OL863332 OL863275 OL863389 is study
L. theobromae K25F eobroma cacao Malaysia: Kedah OL831069 OL863333 OL863276 OL863390 is study
L. theobromae K27F eobroma cacao Malaysia: Kedah OL831070 OL863334 OL863277 OL863391 is study
L. theobromae S30F eobroma cacao Malaysia: Selangor OL831071 OL863335 OL863278 OL863392 is study
L. theobromae PE31F eobroma cacao Malaysia: Perak OL831072 OL863336 OL863279 OL863393 is study
L. theobromae PE32F eobroma cacao Malaysia: Perak OL831073 OL863337 OL863280 OL863394 is study
L. theobromae S34F eobroma cacao Malaysia: Selangor OL831074 OL863338 OL863281 OL863395 is study
L. theobromae S35F eobroma cacao Malaysia: Selangor OL831075 OL863339 OL863282 OL863396 is study
L. theobromae PR36F eobroma cacao Malaysia: Perlis OL831076 OL863340 OL863283 OL863397 is study
L. theobromae PR37F eobroma cacao Malaysia: Perlis OL831077 OL863341 OL863284 OL863398 is study
L. theobromae PE39F eobroma cacao Malaysia: Perak OL831078 OL863342 OL863285 OL863399 is study
L. theobromae K41L eobroma cacao Malaysia: Kedah OL831081 OL863343 OL863286 OL863400 is study
L. theobromae K42L eobroma cacao Malaysia: Kedah OL831082 OL863344 OL863287 OL863401 is study
L. theobromae PR43L eobroma cacao Malaysia: Perlis OL831083 OL863345 OL863288 OL863402 is study
L. theobromae PR44L eobroma cacao Malaysia: Perlis OL831084 OL863346 OL863289 OL863403 is study
L. theobromae PE45L eobroma cacao Malaysia: Perak OL831085 OL863347 OL863290 OL863404 is study
L. theobromae PE46L eobroma cacao Malaysia: Perak OL831086 OL863348 OL863291 OL863405 is study
L. theobromae S47L eobroma cacao Malaysia: Selangor OL831087 OL863349 OL863292 OL863406 is study
L. theobromae S48L eobroma cacao Malaysia: Selangor OL831088 OL863350 OL863293 OL863407 is study
L. theobromae S49L eobroma cacao Malaysia: Selangor OL831089 OL863351 OL863294 OL863408 is study
L. theobromae M50L eobroma cacao Malaysia: Melaka OL831090 OL863352 OL863295 OL863409 is study
L. theobromae M51L eobroma cacao Malaysia: Melaka OL831091 OL863353 OL863296 OL863410 is study
L. theobromae NS52L eobroma cacao Malaysia: Negeri
Sembilan OL831080 OL863354 OL863297 OL863411 is study
L. theobromae NS53L eobroma cacao Malaysia: Negeri
Sembilan OL831079 OL863355 OL863298 OL863412 is study
L. theobromae J54S eobroma cacao Malaysia: Johor OL831092 OL863356 OL863299 OL863413 is study
L. theobromae J55S eobroma cacao Malaysia: Johor OL831093 OL863357 OL863300 OL863414 is study
L. theobromae J56S eobroma cacao Malaysia: Johor OL831094 OL863358 OL863301 OL863415 is study
L. theobromae J57S eobroma cacao Malaysia: Johor OL831095 OL863359 OL863302 OL863416 is study
L. theobromae J58S eobroma cacao Malaysia: Johor OL831096 OL863360 OL863303 OL863417 is study
L. theobromae J59S eobroma cacao Malaysia: Johor OL831097 OL863361 OL863304 OL863418 is study
L. theobromae NS60S eobroma cacao Malaysia: Negeri
Sembilan OL831098 OL863362 OL863305 OL863419 is study
L. theobromae NS61S eobroma cacao Malaysia: Negeri
Sembilan OL831099 OL863363 OL863306 OL863420 is study
L. theobromae NS62S eobroma cacao Malaysia: Negeri
Sembilan OL831100 OL863364 OL863307 OL863421 is study
L. theobromae M63S eobroma cacao Malaysia: Melaka OL831101 OL863365 OL863308 OL863422 is study
L. theobromae M64S eobroma cacao Malaysia: Melaka OL831102 OL863366 OL863309 OL863423 is study
L. theobromae S65S eobroma cacao Malaysia: Selangor OL831103 OL863367 OL863310 OL863424 is study
L. theobromae S66S eobroma cacao Malaysia: Selangor OL831104 OL863368 OL863311 OL863425 is study
L. theobromae PE67S eobroma cacao Malaysia: Perak OL831105 OL863369 OL863312 OL863426 is study
L. theobromae PE68S eobroma cacao Malaysia: Perak OL831106 OL863370 OL863313 OL863427 is study
L. theobromae PP69S eobroma cacao Malaysia: Pulau Pinang OL831107 OL863371 OL863314 OL863428 is study
L. theobromae PP70S eobroma cacao Malaysia: Pulau Pinang OL831108 OL863372 OL863315 OL863429 is study
L. theobromae PP71S eobroma cacao Malaysia: Pulau Pinang OL831109 OL863373 OL863316 OL863430 is study
L. theobromae J72S eobroma cacao Malaysia: Johor OL831110 OL863374 OL863317 OL863431 is study
L. theobromae J73S eobroma cacao Malaysia: Johor OL831111 OL863375 OL863318 OL863432 is study
Continued
Content courtesy of Springer Nature, terms of use apply. Rights reserved
12
Vol:.(1234567890)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
Data availability
All sequence data are available in NCBI GenBank [https:// www. ncbi. nlm. nih. gov/ genba nk/] following the
accession numbers [OL831055–OL831111 (ITS); OL863319–OL863375 (tef1-α); OL863262–OL863318 (tub2);
OL863376–OL863432 (rpb2)] in the manuscript. All data analyzed during this study are included in this pub-
lished article and its supplementary information les.
Received: 18 February 2022; Accepted: 19 May 2022
References
1. Azhar, I. et al. Cocoa Planting Manual (Malaysian Cocoa Board, 2009).
2. Wood, G. A. R. & Lass, R. A. Cocoa (Tropical Agriculture) (Wiley Blackwell, 2001).
3. Lachenaud, P. H., Paulin, D., Ducamp, M. & evenin, J. M. Twenty years of agronomic evaluation of wild cocoa trees (eobroma
cacao L.) from French Guiana—review. Sci. Hortic. 113, 313–321 (2007).
4. Bailey, B. A. & Meinhardt, L. W. Cacao Diseases: A History of Old Enemies and New Encounters (Springer International Publishing,
2016).
5. Dillinger, T. L. et al. Food of the gods: Cure for humanity? A cultural history of the medicinal and ritual use of chocolate. J. Nutr.
130, 2057–2072 (2000).
6. Nair, K. P. e Agronomy and Economy of Important Tree Crops of the Developing World (Springer, 2021).
7. Beckett, S. T., Fowler, M. S. & Ziegler, G. R. Beckett’s Industrial Chocolate Manufacture and Use (Wiley, 2017).
8. Malaysian Cocoa Board (MCB). https:// www. koko. gov. my/ lkm/ loader. cfm? page=1 (2021).
9. Ishaq, S. & Jafri, L. Biomedical importance of cocoa (eobroma cacao): Signicance and potential for the maintenance of human
health. Matrix Sci. Pharm. 1, 1–5 (2017).
10. Steinberg, F. M., Bearden, M. M. & Keen, C. L. Cocoa and chocolate avonoids: Implications for cardiovascular health. J. Am. Diet
Assoc. 103, 215–223 (2003).
11. Dryden, G. W., Song, M. & McClain, C. Polyphenols and gastrointestinal diseases. Curr. Opin. Gastroenterol. 22, 165 (2006).
12. Selmi, C., Mao, T. K., Keen, C. L., Schmitz, H. H. & Gershwin, M. E. e anti-inammatory properties of cocoa avanols. J. Car-
diovasc. Pharmacol. 47, 163–171 (2006).
13. Taubert, D., Roesen, R. & Schömig, E. Eect of cocoa and tea intake on blood pressure: A meta-analysis. Arch. Intern. Med. 167,
626–634 (2007).
14. Latif, R. Health benets of cocoa. Curr. Opin. Clin. Nutr. Metab. Care 16, 669–674 (2013).
15. Guest, D. & Keane, P. Vascular-streak dieback: A new encounter disease of cacao in Papua New Guinea and Southeast Asia caused
by the obligate basidiomycete Oncobasidium theobromae. Phytopathology 97, 1654–1657 (2007).
16. James, R. S., Ray, J., Tan, Y. P. & Shivas, R. G. Colletotrichum siamense, C. theobromicola and C. queenslandicum from several plant
species and the identication of C. asianum in the northern territory, Australia. Australas. Plant Dis. Notes 9, 1–6 (2014).
17. Serrato-Diaz, L. M., Mariño, Y. A., Guadalupe, I., Bayman, P. & Goenaga, R. First report of Lasiodiplodia pseudotheobromae and
Colletotrichum siamense causing cacao pod rot, and rst report of C. tropicale causing cacao pod rot in Puerto Rico. Plant Dis. 104,
592 (2020).
18. Rojas, E. I. et al. Colletotrichum gloeosporioides sl associated with eobroma cacao and other plants in Panama: Multilocus phy-
logenies distinguish host-associated pathogens from asymptomatic endophytes. Mycologia 102, 1318–1338 (2010).
19. Mohali-Castillo, S. & Stewart, J.E. Microfungi associated with diseases on eobroma cacao L. in Merida State, Venezuela. In
Proceedings of the APS Annual Meeting (2017).
20. Mbenoun, M., Momo Zeutsa, E. H., Samuels, G., NsougaAmougou, F. & Nyasse, S. Dieback due to Lasiodiplodia theobromae, a
new constraint to cocoa production in Cameroon. Plant Pathol. 57, 381 (2008).
21. Kannan, C., Karthik, M. & Priya, K. Lasiodiplodia theobromae causes a damaging dieback of cocoa in India. Plant Pathol. 59, 410
(2010).
22. Twumasi, P., Ohene-Mensa, G. & Moses, E. e rot fungus Botryodiplodia theobromae strains cross infect cocoa, mango, banana
and yam with signicant tissue damage and economic losses. Afr. J. Agric. Res. 9, 613–619 (2014).
23. Alvindia, D. G. & Gallema, F. L. M. Lasiodiplodia theobromae causes vascular streak dieback (VSD)-like symptoms of cacao in
Davao Region, Philippines. Australas. Plant Dis. Notes 12, 1–4 (2017).
24. Asman, A. et al. La siodiplodia theobromae: an emerging threat to cocoa causes dieback and canker disease in Sulawesi. In Proceed-
ings of the Asia-Pacic Regional Cocoa IPM Symposium (2019).
25. Puig, A. S., Keith, L. M., Matsumoto, T. K., Gutierrez, O. A. & Marelli, J. P. Virulence tests of Neofusicoccum parvum, Lasiodiplodia
theobromae, and Phytophthora palmivora on eobroma cacao. Eur. J. Plant Pathol. 159, 851–862 (2021).
26. Meinhardt, L. W. et al. Moniliophthora perniciosa, the causal agent of witches’ broom disease of cacao: What’s new from this old
foe?Mol. Plant Pathol. 9, 577–588 (2008).
27. Phillips-Mora, W. & Wilkinson, M. J. Frosty pod of cacao: A disease with a limited geographic range but unlimited potential for
damage. Phytopathology 97, 1644–1647 (2007).
28. Puig, A. S., Marelli, J. P., Matsumoto, T. K., Keith, L. M. & Gutierrez, O. A. First report of Neofusicoccum parvum causing pod rot
on cacao in Hawaii. Plant Dis. 103, 1416 (2019).
29. Akro, A. Y., Amoako-Atta, I., Assuah, M. & Asare, E. K. Black pod disease on cacao (eobroma cacao L.) in Ghana: Spread of
Phytophthora megakarya and role of economic plants in the disease epidemiology. Crop Prot. 72, 66–75 (2015).
30. Alves, A., Crous, P. W., Correia, A. & Phillips, A. J. L. Morphological and molecular data reveal cryptic speciation in Lasiodiplodia
theobromae. Fungal Divers. 28, 1–13 (2008).
31. Phillips, A. J. L. et al. e Botryosphaeriaceae: Genera and species known from culture. Stud. Mycol. 76, 51–167 (2013).
Species Isolate Host Location
GenBank accession number
ReferencesITS tef1-α tub2 rpb2
Lasiodiplodia viticola CBS128313ahybrid grape Vignoles USA HQ288227 HQ288269 HQ288306 KU696385 Cruywagen etal.52
L. viticola CBS128314 Chardonel USA HQ288228 HQ288270 HQ288307 KU696386 Cruywagen etal.52
Botryosphearia dothidea CBS115476 Prunus sp. Switzerland KF766151 AY236898 MT592470 DQ677944 Slippers etal.83
Table 2. List of GenBank accession numbers of Lasiodiplodia species and the outgroup (Botryosphearia
dothidea) used in the phylogenetic analysis. a Ex-type isolates.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
13
Vol.:(0123456789)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
32. Punithalingam, E. Plant diseases attributed to Botryodiplodia theobromae. Bibl. Mycol. 71, 1–123 (1980).
33. Burgess, T. I., Sakalidis, M. L. & Hardy, G. E. S. Gene ow of the canker pathogen Botryosphaeria australis between Eucalyptus
globulus plantations and native eucalypt forests in Western Australia. Austral. Ecol. 31, 559–566 (2006).
34. Slippers, B. & Wingeld, M. J. Botryosphaeriaceae as endophytes and latent pathogens of woody plants: Diversity, ecology, and
impact. Fungal Biol. Rev. 21, 90–106 (2007).
35. Latha, P. et al. First report of Lasiodiplodia theobromae (Pat.) Gri. and Maubl causing root rot and collar rot of physic nut (Jatropha
curcas L.) in India. Australas. Plant Dis. Notes 4, 19–20 (2009).
36. Sakalidis, M. L., Ray, J. D., Lanoiselet, V., Hardy, G. E. S. & Burgess, T. I. Pathogenic Botryosphaeriaceae associated with Mangifera
indica in the Kimberley region of Western Australia. Eur. J. Plant Pathol. 130, 379–391 (2011).
37. Ismail, A. M. et al. Lasiodiplodia species associated with dieback disease of mango (Mangifera indica) in Egypt. Australas. Plant
Pathol. 41, 649–660 (2012).
38. Urbez-Torres, J. R. et al. Characterization of fungal pathogens associated with grapevine trunk diseases in Arkansas and Missouri.
Fungal Divers. 52, 169–189 (2012).
39. Netto, M. S. B. et al. Species of Lasiodiplodia associated with papaya stem-end rot in Brazil. Fungal Divers. 67, 127–141 (2014).
40. Slippers, B. et al. Confronting the constraints of morphological taxonomy in the Botryosphaeriales. Persoonia 33, 155–168 (2014).
41. Yildiz, A., Benlioglu, K. & Benlioglu, H. S. First report of strawberry dieback caused by Lasiodiplodia theobromae. Plant Dis. 98,
1579 (2014).
42. Li, H. L. et al. Lasiodiplodia theobromae and L. pseudotheobromae causing leaf necrosis on Camellia sinensis in Fujian Province,
China. Can. J. Plant Pathol. 41, 277–284 (2019).
43. Berraf-Tebbal, A. et al. Lasiodiplodia mitidjana sp. nov. and other Botryosphaeriaceae species causing branch canker and dieback
of Citrus sinensis in Algeria. PLoS ONE 15, 1–18. https:// doi. org/ 10. 1371/ journ al. pone. 02324 48 (2020).
44. Norhayati, M., Erneeza, M. H. & Kamaruzaman, S. Morphological, pathogenic and molecular characterization of Lasiodiplodia
theobromae: A causal pathogen of black rot disease on kenaf seeds in Malaysia. Int. J. Agric. Biol. 18, 80–85 (2016).
45. Kee, Y. J., Zakaria, L. & Mohd, M. H. Lasiodiplodia species associated with Sansevieria trifasciata leaf blight in Malaysia. J. Gen.
Plant Pathol. 85, 66–71 (2019).
46. Li, L., Mohd, M. H., Mohamed Nor, N. M. I., Subramaniam, S. & Latiah, Z. Identication of Botryosphaeriaceae associated with
stem-end rot of mango (Mangifera indica L.) in Malaysia. J. Appl. Microbiol. 130, 1273–1284 (2020).
47. Maid, M. et al. First report of stem canker disease on Acacia mangium induced by Lasiodiplodia theobromae and Lasiodiplodia
pseudotheobromae species in Sabah, Malaysia. Malays. Appl. Biol. 47, 147–151 (2018).
48. Sulaiman, R., anarajoo, S. S., Kadir, J. & Vadamalai, G. First report of Lasiodiplodia theobromae causing stem canker of Jatropha
curcas in Malaysia. Plant Dis. 96, 767 (2012).
49. Munirah, M. S., Azmi, A. R., Yong, S. Y. C. & Nur Ain Izzati, M. Z. Characterization of Lasiodiplodia theobromae and L. pseudothe-
obromae causing fruit rot on pre-harvest mango in Malaysia. Plant Pathol. Quar. 7, 202–213 (2017).
50. Z ee, K. Y., Asib, N. & Ismail, S. I. First report of Lasiodiplodia theobromae causing postharvest fruit rot on guava (Psidium guajava)
in Malaysia. Plant Dis. 105, 2716 (2021).
51. Hyde, K. D., Abd-Elsalam, K. & Cai, L. Morphology: Still essential in a molecular world. Mycotaxon 114, 439–451 (2010).
52. Cr uywagen, E. M., Slippers, B., Roux, J. & Wingeld, M. J. Phylogenetic species recognition and hybridisation in Lasiodiplodia: A
case study on species from baobabs. Fungal Biol. 121, 420–436 (2017).
53. Schoch, C. L. et al. Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for fungi. Proc.
Natl. Acad. Sci. USA. 109, 6241–6246 (2012).
54. Kashyap, P. L., Rai, P., Kumar, S., Chakdar, H. & Srivastava, A. K. DNA barcoding for diagnosis and monitoring of fungal plant
pathogens. In Molecular Markers in Mycology: Diagnostics and Marker Development (eds Singh, B. P. & Gupta, V. K.) 87–122
(Springer, 2017).
55. Abdollahzadeh, J., Javadi, A., Goltapeh, E. M., Zare, R. & Philip, A. J. L. Phylogeny and morphology of four new species of Lasi-
odiplodia from Iran. Persoonia 25, 1–10 (2010).
56. Geiser, D. M. et al. FUSARIUM-ID v. 1.0: A DNA sequence database for identifying Fusarium. Eur. J. Plant Pathol. 110, 473–479
(2004).
57. Udayanga, D. et al. Multi-locus phylogenetic evaluation of Diaporthe (Phomopsis). Fungal Divers. 56, 157–171 (2012).
58. Lopes, U. P., Zambolim, L. & Pereira, O. L. First report of Lasiodiplodia theobromae causing leaf blight on the orchid Catasetum
mbriatum in Brazil. Australas. Plant Dis. Notes 4, 64–65 (2009).
59. Ramjegathesh, R., Johnson, I., Hubballi, M. & Maheswarappa, H. P. Characterization of Lasiodiplodia theobromae causing leaf
blight disease of coconut. J. Plant Crops 47, 62–71 (2019).
60. Santos, P. H. et al. Is Lasiodiplodia theobromae the only species that causes leaf blight disease in Brazilian coconut palms?Trop.
Plant Pathol. 45, 434–442 (2020).
61. Fan, R . et al. First report of Lasiodiplodia theobromae causing leaf blight of Kadsura longipedunculata in China. Plant Dis. 104,
3063 (2020).
62. Chen, F., Zheng, X., Zhao, X. & Chen, F. First report of Lasiodiplodia theobromae causing stem canker of Fraxinus americana. Plant
Dis. 103, 3276 (2019).
63. Borrero, C., Pérez, S. & Avilés, M. First report of canker disease caused by Lasiodiplodia theobromae on blueberry bushes in Spain.
Plant Dis. 103, 2684 (2019).
64. Khanzada, M. A., Lodhi, A. M. & Shahzad, S. Mango dieback and gummosis in Sindh, Pakistan caused by Lasiodiplodia theobromae.
Plant Health Prog. 15, 13 (2004).
65. Cardoso, J. E., Vidal, J. C., dos Santos, A. A., Freir, F. C. O. & Viana, F. M. P. First report of black branch dieback of cashew caused
by Lasiodiplodia theobromae in Brazil. Plant Dis. 86, 558 (2002).
66. Wang, W. & Song, X. First report of Lasiodiplodia theobromae and L. pseudotheobromae causing canker disease of sacha inchi in
Hainan, China. Plant Dis. 105, 3757 (2021).
67. Bautista-Cruz, M. A. et al. Phylogeny, distribution, and pathogenicity of Lasiodiplodia species associated with cankers and dieback
symptoms of persian lime in Mexico. Plant Dis. 103, 1156–1165 (2019).
68. Úrbez-Torres, J. R. & Gubler, W. D. Pathogenicity of Botryosphaeriaceae species isolated from grapevine cankers in California.
Plant Dis. 93, 584–592 (2009).
69. White, T. J., Bruns, T., Lee, S. & Taylor, J. W. Amplication and direct sequencing of fungal ribosomal RNA genes for phylogenetics.
In PCR Protocols: A Guide to Methods and Applications (eds Innis, M. A. et al.) 315–322 (Academic Press, 1990).
70. Glass, N. L. & Donaldson, G. C. Development of primer sets designed for use with the PCR to amplify conserved genes from
lamentous ascomycetes. Appl. Environ. Microbiol. 61, 1323–1330 (1995).
71. Kumar, S., Stecher, G. & Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol.
Evol. 33, 1870–1874 (2016).
72. Tamura, K. Estimation of the number of nucleotide substitutions when there are strong transition-transversion and G + C-content
biases. Mol. Biol. Evol. 9, 678–687 (1992).
73. Felsenstein, J. Condence limits on phylogenies: An approach using the bootstrap. Evolution 39, 783–791 (1985).
74. Parker, S. R., Shaw, M. W. & Royle, D. J. e reliability of visual estimates of disease severity on cereal leaves. Plant Pathol. 44,
856–864 (1995).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
14
Vol:.(1234567890)
Scientic Reports | (2022) 12:8966 | https://doi.org/10.1038/s41598-022-13057-9
www.nature.com/scientificreports/
75. Zhang, W. et al. Evaluating species in Botryosphaeriales. Persoonia 46, 63–115 (2021).
76. Marques, M. W. et al. Species of Lasiodiplodia associated with mango in Brazil. Fungal Divers. 61, 181–193 (2013).
77. Machado, A. R., Pinho, D. B. & Pereira, O. L. Phylogeny, identication and pathogenicity of the Botryosphaeriaceae associated
with collar and root rot of the biofuel plant Jatropha curcas in Brazil, with a description of new species of Lasiodiplodia. Fungal
Divers. 67, 231–247 (2014).
78. Liu, J. K. et al. Towards a natural classication of Botryosphaeriales. Fungal Divers. 57, 149–210 (2012).
79. Begoude, B. D., Slippers, B., Wingeld, M. J. & Roux, J. Botryosphaeriaceae associated with Terminalia catappa in Cameroon,
South Africa and Madagascar. Mycol. Prog. 9, 101–123 (2010).
80. Linaldeddu, B. T. et al. Diversity of Botryosphaeriaceae species associated with grapevine and other woody hosts in Italy, Algeria
and Tunisia, with descriptions of Lasiodiplodia exigua and Lasiodiplodia mediterranea sp. nov. Fungal Divers. 71, 201–214 (2015).
81. Phillips, A. J. L. et al. Resolving the phylogenetic and taxonomic status of dark-spored teleomorph genera in the Botryosphaeriaceae.
Persoonia 21, 29–55 (2008).
82. Mehl, J., Wingeld, M. J., Roux, J. & Slippers, B. Invasive everywhere? Phylogeographic analysis of the globally distributed tree
pathogen Lasiodiplodia theobromae. Forests 8, 145 (2017).
83. Slippers, B. et al. Phylogenetic lineages in the Botryosphaeriales: A systematic and evolutionary framework. Stud. Mycol. 55, 35–52
(2006).
Acknowledgements
e authors thank the Malaysian Cocoa Board (MCB) for permission to collect samples and provide healthy
seedlings for pathogenicity tests. Special thanks to the MCB sta who have assisted in the eldwork activities.
Author contributions
A.R.H-S.: conceptualization, methodology, formal analysis, investigation, writing-original dra preparation.
N.M.I.M.N., L.Z., Y.-H.L.: writing-review and editing. M.H.M.: conceptualization, methodology, investigation,
writing-review and editing, supervision.
Funding
is research was funded by Fundamental Research Grant Scheme (FRGS/1/2019/WAB01/USM/02/1) from
Ministry of Higher Education, Malaysia.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary Information e online version contains supplementary material available at https:// doi. org/
10. 1038/ s41598- 022- 13057-9.
Correspondence and requests for materials should be addressed to M.H.M.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access is article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if changes were made. e images or other third party material in this
article are included in the articles Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this licence, visit http:// creat iveco mmons. org/ licen ses/ by/4. 0/.
© e Author(s) 2022
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Fungal diseases are the main limiting factor of the crop worldwide (Huda-Shakirah et al. 2022). Currently, the main fungal diseases reported on Peruvian cacao farms are frosty pod rot (caused by Moniliophthora roreri), witches' broom (M. ...
... PCR assays were performed to amplify the internal transcribed spacer region (ITS1, 5.8S and ITS2 rDNA regions; ITS), and the partial translation elongation factor 1-α (tef1), β tubulin (tub2) and RNA polymerase II subunit (rpb2) genes. We used the primer pairs ITS1F/ITS4, EF1-688F/EF1-1251R, Bt2a/Bt2b, and rpb2-LasF/rpb2-LasR, respectively, as in Huda-Shakirah et al. (2022). PCR products were sequenced at Macrogen (Seoul, South Korea). ...
... However, the fungus may show aseptate conidia when young, but these conidia may develop septa as they mature (Burgess et al. 2006;Phillips et al. 2013). In addition, the microscopic morphological characteristics of the conidia were consistent with those reported for this species in previous studies (Coutinho et al. 2017;Huda-Shakirah et al. 2022). The shape and color of mature conidia, as well as the presence of septa and longitudinal striae were important features for the identification of Lasiodiplodia spp. ...
Article
Lasiodiplodia spp. are known to cause canker, rot and dieback symptoms in several crops worldwide. In this study, two Lasiodiplodia species were identified as pathogens of native cacao accessions from the department of Amazonas, Peru, causing dieback and pod rot on young and old stems. We evaluated the macro and micro morphological characteristics, and conducted a molecular identification based on a phylogenetic analysis with a multilocus dataset with informative loci for the genus. Microscopic examination revealed the presence of immature conidia that were initially hyaline, subovoid, unicel-lular, and double-layered, which became reddish brown with a central septum and longitudinal grooves at maturity. In the phylogenetic analysis, we identified our isolates as L. theobromae and L. iraniensis with strong bootstrap support values. Koch's postulates were fulfilled after the re-isolation of the same species from diseased tissues of cacao fruits and stems after an artificial inoculation. Therefore, in this study, we report for the first time L. theobromae and L. iraniensis infecting native cacao plants in Amazonas, Peru.
... i v i cacao, L. theobromae là tác nhân gây b nh b c lá (leaf blight), th i thân (stem canker), th i qu (pod rot) và ch t khô (dieback) [34,35]. Ngoài ra, m t s báo cáo khác cho r ng L. theobromae còn là tác nhân gây b nh m lá (leaf spot), khô cành (branch dieback), th i thân (stem canker) cho cây g t ch [36], b nh ch t khô (dieback) cây nho [37], b nh b c lá (leaf blight) d a [38]. ...
... Nh ng nghiên c u trên cho th y L. theobromae là m t m m b nh th c v t ph bi n và thích nghi v i nhi u ký ch . i u ki n c n thi t L. theobromae gây b nh là thông qua v t th ng h trên ký ch [32,34,38], h n n a l p bi u bì c a OFI r t dày, c bao b c b i l p cutin ch ng thoát n c, vì v y t i nh ng v trí không gây v t th ng trên b thân x ng r ng, vi n m không xâm nh p gây b nh c. Do ó, vi c h n ch gây v t th ng nh h n ch các ng v t hút chích, v t th ng c h c do c t t a,…, trên ru ng OFI s giúp h n ch s lây lan b nh h i. ...
Article
Xương rồng Nopal (Opuntia ficus-indica) vừa là nguồn thực phẩm giàu dinh dưỡng chongười và gia súc, vừa là giải pháp lý tưởng cho những vùng đất hoang hóa, đất cát nhiễmmặn ven biển bỏ hoang không canh tác nông nghiệp được của Việt Nam, đặc biệt làNinh Thuận. Những năm gần đây, bệnh thối thân Xương rồng Nopal đã xuất hiện và nhanh chóng lây lan làm đổ gãy cây, ảnh hưởng đến đến năng suất và làm giảm hiệu quả canh tác. Dựa vào đặc điểm hình thái và trình tự DNA vùng ITS, vi nấm gây bệnh được xác định là Macrophomina sp. và Lasiodiplodia theobromae. Phân tích cây phát sinh loài dựa trên trình tự vùng ITS cho thấy 2 vi nấm này phân bố cùng nhóm với các phân lập khác của Macrophomina sp. và L. theobromae trên GenBank. Các thí nghiệmvề khả năng gây bệnh đã được thực hiện và đáp ứng định đề của Koch. Kết quả nghiên cứu đã chứng minh vai trò của cả 2 vi nấm trong việc gây bệnh thối thân trên Xương rồng Nopal, cung cấp thông tin quan trọng giúp nghiên cứu các biện pháp phòng trừ hợp lý, hiệu quả.® 2024 Journal of Science and Technology - NTTU
... La frecuencia de la aparición de cada hongo se muestra en la Figura 2 siendo los más recurrentes el Lasiodiplodia spp presente en casi el 80% de las muestras y el Aspergillus spp observado en más del 50 % de los frutos examinados. La presencia de estos hongos no parece tener un efecto patológico sobre los frutos, en especial la Lasiodiplodia spp, siendo un organismo saprófito, no suele ser sintomático en los frutos que infecta, sin embargo, se le considera un fitopatógeno latente que puede causar afectaciones cuando la homeostasis de la planta se ve alterada [23], [24]. ...
... Lasiodiplodia spp [23], [25] Mucor spp [26], [27] Aspergillus spp [28]- [30] Curvularia spp [31], [ ...
Article
Full-text available
Resumen El sulfato de cobre (II) pentahidratado (CuSO4·5H2O) es un compuesto común en la industria agrícola, usado para tratar la enfermedad del cacao denominada "moniliasis" ocasionada por la infección del fitopatógeno fúngico Moniliophthora roreri (M. roreri) en Theobroma cacao (T. cacao). El objetivo de este proyecto fue evaluar in vitro el efecto antifúngico del CuSO4·5H2O en M. roreri aislado de frutos de cacao infectados. Inicialmente, se aislaron los hongos endófitos de frutos de cacao sanos (T. cacao) en medios ADP y AV8, ambos identificados morfológicamente con azul de lactofenol. Posteriormente, M. roreri fue aislado de frutos de cacao infectados (T. cacao) e identificado morfológicamente con azul de lactofenol. El efecto fungicida del CuSO4·5H2O sobre M. roreri se evaluó por infusión en agares ADP y AV8 en concentraciones de 0.0625-1 mM y de 0.625-10 µM. De acuerdo con los resultados obtenidos, Lasiodiplodia spp fue el hongo endófito más frecuente en los frutos de cacao sanos analizados, mientras que la concentración mínima efectiva de CuSO4·5H2O sobre M. roreri fue 10 µM a 7 días de incubación. Palabras claves: sulfato de cobre, Moniliophthora roreri, Theobroma cacao. Abstract Copper (II) sulfate pentahydrate (CuSO4·5H2O) is a common compound in the agricultural industry, used to treat the cocoa disease called "moniliasis" caused by the infection of the fungal phytopathogen Moniliophthora roreri (M. roreri) on Theobroma cacao. The objective of this project was to evaluate in vitro the antifungal effect of CuSO4·5H2O on M. roreri isolated from infected cacao fruits. Initially, endophytic fungi were isolated from healthy cocoa fruits (T. cacao) on ADP and AV8 media, both morphologically identified with lactophenol blue. Subsequently, M. roreri was isolated from infected cocoa fruits (T. cacao) and morphologically identified with lactophenol blue. The fungicidal effect of CuSO4·5H2O on M. roreri was evaluated by infusion in ADP and AV8 agars at concentrations of 0.0625-1 mM and 0.625-10 µM. According to the results obtained, Lasiodiplodia spp was the most frequent endophytic fungus in the healthy cocoa fruits analyzed, while the minimum effective concentration of CuSO4·5H2O on M. roreri was 10 µM at 7 days of incubation.
... and N. clavispora, several studies have reported the occurrence of other species as plant pathogens associated with T. cacao. For example, Colletotrichum gloeosporioides, Colletotrichum tropicale, Colletotrichum siamense, D. pseudomangiferae, Lasiodiplodia pseudotheobromae, Lasiodiplodia theobromae, Neofusicoccum parvum, Nigrospora sphaerica, Phytophthora palmivora and Phytophthora megakarya(Akrofi et al., 2015;Asare et al., 2021;Huda-Shakirah et al., 2022;Huda-Shakirah & Mohd, 2023;Puig et al., 2021; ...
Article
Malaysia stands prominently among Asia’s key cocoa-producing countries. In the cocoa season of 2022/2023, Malaysia demonstrated its contribution to the industry, with an estimated production of around 364,000 tonnes of cacao bean grindings. Nonetheless, fungal diseases pose undeniable challenges to the cocoa sector. Extensive sampling conducted between September 2018 and March 2019 across multiple states in Malaysia revealed concerning symptoms of leaf blight and stem canker affecting Theobroma cacao plants. The aim of this study was to identify and characterize fungal species associated with leaf blight and stem canker of T. cacao in Malaysia through morphological, molecular and pathogenicity analyses. Morphological and molecular phylogenetic analyses using multiple DNA regions (rDNA internal transcribed spacer [ITS], TEF1 and TUB2) were performed and identified 40 fungal isolates found in this study as Diaporthe tulliensis (17 isolates), Fusarium solani (seven isolates), Fusarium proliferatum (six isolates) and Neopestalotiopsis clavispora (10 isolates). Pathogenicity tests with mycelial plugs and wound treatments showed that D. tulliensis and N. clavispora were responsible for causing leaf blight whereas D. tulliensis, F. solani and F. proliferatum caused stem canker of T. cacao. The present study provides insights into disease aetiology and symptomatology that may be useful in planning effective disease management for the host plant.
... It causes fungal gummosis of peaches-a disease that severely restricts the the growth and production of this fruit in orchards in southern China, the United States, and Japan [56]. It causes leaf blight, stem cancer, and fruit rot in Theobroma cacao in Malaysia [57]. In Brazil, L. theobromae is a serious threat to cashew cultivation areas, causing resinosis and black rot of the stem [58]. ...
Article
Full-text available
This study evaluates the activity of a recombinant chitinase from the leaf-cutting ant Atta sexdens (AsChtII-C4B1) against colloidal and solid α-and β-chitin substrates. 1 H NMR analyses of the reaction media showed the formation of N-acetylglucosamine (GlcNAc) as the hydrolysis product. Viscometry analyses revealed a reduction in the viscosity of chitin solutions, indicating that the enzyme decreases their molecular masses. Both solid state 13 C NMR and XRD analyses showed minor differences in chitin crystallinity pre-and post-reaction, indicative of partial hydrolysis under the studied conditions, resulting in the formation of GlcNAc and a reduction in molecular mass. However, the enzyme was unable to completely degrade the chitin samples, as they retained most of their solid-state structure. It was also observed that the enzyme acts progressively and with a greater activity on α-chitin than on β-chitin. AsChtII-C4B1 significantly changed the hyphae of the phytopathogenic fungus Lasiodiplodia theobromae, hindering its growth in both solid and liquid media and reducing its dry biomass by approximately 61%. The results demonstrate that AsChtII-C4B1 could be applied as an agent for the bioproduction of chitin derivatives and as a potential antifungal agent.
... Therefore, it is essential to identify Lasiodiplodia using a molecular approach. The combination of ITS, tub, rpb2 and tef1α genes has provided researchers with a powerful tool to identify Lasiodiplodia species (Hyde et al. 2014;Dou et al. 2017;Wang et al. 2021;Xiao et al. 2021;Huda-Shakirah et al. 2022). However, the use of a single gene did not resolve the identification of Lasiodiplodia at the species level (Hyde et al. 2014;de Silva et al. 2019;Wang et al. 2021). ...
Article
Full-text available
Botryosphaeriaceae (Botryosphaeriales) encompasses numerous endophytic, saprobic, and plant pathogenic fungal species associated with a wide range of hosts. They are commonly associated with woody plants and are recognized as opportunistic fungal pathogens. In this study, five endophytic Botryosphaeriaceae species were isolated from epiphytic plants of Ficus pumila and Lemmapphyllum microphyllum (Le.) and identified using both morphology and phylogeny. Botryosphaeria dothidea, B. ramosa, Lasiodiplodia citricola, L. mahajangana, and L. theobromae were identified and described. Notably, Lasiodiplodia citricola and L. mahajangana were isolated for the first time from F. pumila, while Botryosphaeria dothidea, B. ramosa, L. citricola, L. mahajangana, and L. theobromae were isolated for the first time from Le. microphyllum. We also synonymize Botryosphaeria tenuispora as B. dothidea and Lasiodiplodia chonburiensis as L. mahajangana. The results of this study significantly contribute to our understanding of the fungal diversity in F. pumila and Le. microphyllum. In addition, our findings expand the knowledge regarding the diversity and distribution of Botryosphaeriaceae species in South China.
Article
Full-text available
The tea plant, Camellia sinensis [L.] O. Kuntze, is a vital global agricultural commodity, yet faces challenges from fungal infections, which affects its production. To reduce the loss in the tea production, the fungal infections must be removed which is managed with fungicides, which are harmful to the environment. Leaf necrosis, which decreases tea quality and quantity, was investigated across Assam, revealing Lasiodiplodia theobromae as the causative agent. Pathogenicity tests, alongside morphological and molecular analyses, confirmed its role in leaf necrosis. Genome and gene analysis of L. theobromae showed multiple genes related to its pathogenicity. The study also assessed the impact of chemical pesticides on this pathogen. Additionally, the findings in this study highlight the significance of re-assessing management approaches in considering the fungal infection in tea.
Article
Pestalotiopsis leaf fall disease (PLFD) has been on the increase in recent years, resulting in diminished latex yield in many rubber producing countries. Thus far, some reports have pointed to the presence of fungal pathogen(s) other than Pestalotiopsis sp. in the diseased leaves. To decipher the diversity of fungal pathogens, a total of 110 early-stage and another 110 late-stage symptomatic leaves were sampled from 12 Hevea clones in seven different areas in Malaysia during wintering and wet seasons at different times of the year. A total of 487 fungal isolates were obtained; 241 were from the early-stage and 246 from the late-stage symptomatic leaves. These isolates were clustered into 11 morphotypes based on colony appearance and microscopic observation of the spore. The diversity analysis revealed significant differences in morphotype richness and evenness when comparing different rubber clones, seasonal effects, and locations by which the fungal isolates were sampled. Five isolates representative of the five predominant morphotypes associated with PLFD were analysed using the Internal Transcribed Spacer (ITS) markers and BLAST analysis. Their closest relative species were identified as Colletotrichum conoides, Neopestalotiopsis surinamensis, Lasiodiplodia theobromae, Phyllosticta fallopiae and Letendraea cordylinicola. This study enhanced our understanding of PLFD, particularly the diversity and distribution of fungal communities associated with the disease across different rubber clones, seasons, and locations in Malaysia. The identification of these fungal isolates associated with PLFD symptoms paves the way for further research into their pathogenicity and the development of targeted management strategies to mitigate the disease.
Article
Full-text available
During June 2018, several symptomatic fig (Ficus carica) cuttings, showing twig blight, subcortical discolouration and apical dieback were collected from a nursery in Catania province, Sicily (Italy). Isolations from diseased tissue consistently showed the presence of the same fungal colony. Morphology of the fungal isolates together with sequence data of the nuclear rDNA internal transcriber spacer (ITS) region, translation elongation factor 1-alpha (tef1) gene and partial beta-tubulin (tub2) gene of representatives isolates revealed the presence of the fungus Neofusicoccum parvum. Pathogenicity tests were conducted by inoculating fig cuttings with mycelial plugs. After 10 days, the inoculated plants developed cankers similar to those observed in the greenhouse and after 26 days all inoculated plants were dead. To the best of our knowledge, this is the first report worldwide of N. parvum causing disease on this host.
Article
Full-text available
Guava (Psidium guajava L.) is an economically important tropical fruit crop and is cultivated extensively in Malaysia. In September and October 2019, postharvest fruit rot symptoms were observed on 30% to 40% of guava fruit cv. Kampuchea in fruit markets of Puchong and Ipoh cities in the states of Selangor and Perak, Malaysia. Initial symptoms appeared as brown, irregular, water-soaked lesions on the upper portion of the fruit where it was attached to the peduncle. Subsequently, lesions then progressed to cover the whole fruit (Fig.1A). Lesions were covered with an abundance of black pycnidia and grayish mycelium. Ten symptomatic guava fruit were randomly collected from two local markets for our investigation. For fungal isolation, small fragments (5×5 mm) were excised from the lesion margin, surface sterilized with 0.5% NaOCl for 2 min, rinsed three times with sterile distilled water, placed on potato dextrose agar (PDA) and incubated at 25 °C with 12-h photoperiod for 2-3 days. Eight single-spore isolates with similar morphological characteristics were obtained and two representative isolates (P8 and S9) were characterized in depth. Colonies on PDA were initially composed of grayish-white aerial mycelium, but turned dark-gray after 7 days (Fig. 1B). Abundant black pycnidia were observed after incubation for 4 weeks. Immature conidia were hyaline, aseptate, ellipsoid, thick-walled, and mature conidia becoming dark brown and 1-septate with longitudinal striations, 25.0 − 27.0 ± 2.5 × 13.0 − 14.0 ± 1.0 μm (n = 30) (Fig.1C, D). On the basis of morphology, both representative isolates were identified as Lasiodiplodia theobromae (Pat.) Griffon & Maubl. (Alves et al. 2008). For molecular identification, genomic DNA of the two isolates was extracted using the DNeasy plant mini kit (Qiagen, USA). The internal transcribed spacer (ITS) region of rDNA and translation elongation factor 1-alpha (EF1-α) genes were amplified using ITS5/ITS4 and EF1-728F/EF1-986R primer set, respectively (White et al. 1990, Carbone and Kohn 1999). BLASTn analysis of the resulting ITS and EF1-α sequences indicated 100% identity to L. theobromae ex-type strain CBS 164.96 (GenBank accession nos: AY640255 and AY640258, respectively) (Phillips et al. 2013). The ITS (MW380428, MW380429) and EF1-α (MW387153, MW387154) sequences were deposited in GenBank. Phylogenetic analysis using the maximum likelihood based on the combined ITS-TEF sequences indicated that the isolates formed a strongly supported clade (100% bootstrap value) to the related L. theobromae (Kumar et al. 2016) (Fig.2). A pathogenicity test of two isolates was conducted on six healthy detached guava fruits per isolate. The fruit were surface sterilized using 70% ethanol and rinsed twice with sterile water prior inoculation. The fruit were wound-inoculated using a sterile needle according to the method of de Oliveira et al. (2014) and five-mm-diameter mycelial agar plugs from 7-days-old PDA culture of the isolates were placed onto the wounds. Six additional fruit were wound inoculated using sterile 5-mm-diameter PDA agar plugs to serve as controls. Inoculated fruit were placed in sterilized plastic container and incubated in a growth chamber at 25 ± 1 °C, 90% relative humidity with a photoperiod of 12-h. The experiment was conducted twice. Five days after inoculation, symptoms as described above developed on the inoculated sites and caused a fruit rot, while control treatment remained asymptomatic. L. theobromae was reisolated from all symptomatic tissues and confirmed by morphological characteristics and confirmed by PCR using ITS region. L. theobromae has recently been reported to cause fruit rot on rockmelon in Thailand (Suwannarach et al. 2020). To our knowledge, this is the first report of L. theobromae causing postharvest fruit rot on guava in Malaysia. The occurrence of this disease needs to be monitored as this disease can reduce the marketable yield of guava. Preventive strategies need to be developed in the field to reduce postharvest losses.
Article
Full-text available
Fig (Ficus carica L.) is among the earliest and widely cultivated fruit trees in the world due to its easy adaptation to diverse climates (Solomons et al. 2006). In July 2020, a rot disease was observed on multiple orchards located in Faisalabad- a region of Punjab Province. The symptoms appeared as light brown, circular to oval, and water-soaked lesions (4-8 mm in diameter). In more advanced stages of the disease, the lesions enlarged in size and leading to rot of the entire fruit. Disease incidence on fruit across the fields ranged from 23 to 29%. To isolate the causal agent, segments (5 mm2) were excised from 15 symptomatic fruit, surface disinfested with 70% ethanol for 1 min, washed in three changes of sterilized water, air dried, transferred aseptically to plates containing potato dextrose agar (PDA), and incubated at 25°C for 7 days with a 12-h photoperiod. Nine single spore isolates with similar morphology were isolated from the infected tissues. The cultured isolates consistently yielded dark brown to black colonies on PDA. Conidia were in chains (average conidial dimension 20 to 28 × 8 to 10 μm), olivaceous to dark brown, with a short conical beak with both transversal (two to five) and longitudinal (one to three) septa. Conidiophores were short, septate, hyaline to olivaceous brown, either branched or unbranched, 20 to 52 μm long, and 1 to 3 μm wide. These cultural and morphological characteristics were consistent with the descriptions of Alternaria alternata (Simmons 2007). The genomic DNA from three isolates was extracted using a PrepMan Ultra kit according to the manufacturer’s protocol and amplified using universal primers ITS1/4 (White et al. 1990) and the endopolygalacturonase gene using primers PG3/PG2b (Andrew et al. 2009), and sequenced. The amplified PCR products were deposited in GenBank (accession nos. MW261786, MW433689, MW439319 for ITS and MW249057, MW463344, MW463345 for PG3/PG2b). Blast searches against GenBank showed 99%-100% nucleotide identity with the reference sequences of various A. alternata isolates. The pathogenicity of the representative isolate (PDL 2021) was tested on Fig fruit cv. “Black Mission”. For that, 20 asymptomatic and mature fruit were surface-disinfected with 75% ethanol solution for 30 s. The fruit were inoculated by spraying a spore suspension (106 spores/ml) of A. alternata and stored at 25°C and 80% relative humidity. An equal number of fruit inoculated with sterile water were used as negative controls. Symptoms similar to those on the naturally infected fruits began after 4-5 days of inoculation. The negative controls remained healthy. Koch’s postulates were fulfilled by reisolating (100%) A. alternata from only the inoculated fruit. Previously, the pathogen has been reported to cause fruit rot of Lychee, Peach and Pomegranate in Pakistan (Alam et al 2017a; 2019b; 2019c). The pathogen has been reported to cause fig fruit rot in California (Michailides et al. 1994). Keeping in view the extent of disease on many fruits, further studies are needed on management options to combat the disease in Punjab Province of Pakistan.
Article
Full-text available
Neofusicoccum parvum is a recently reported pathogen affecting Theobroma cacao L., and has been isolated from symptomatic pods on Oahu and Hawaii Islands. Determining infection routes and virulence are essential for assessing the impact of N. parvum on cacao production and developing effective disease management strategies. Infection routes were determined by inoculating unwounded stems and pods with six isolates of N. parvum alongside Lasiodiplodia theobromae and Phytophthora palmivora . Fifty percent of unwounded stems inoculated with P. palmivora developed lesions, but only a single lesion developed following inoculation with N. parvum (isolate H44). L. theobromae and the remaining N. parvum isolates did not induce lesion development on unwounded stems. In contrast, all N. parvum and L. theobromae isolates produced lesions on 40–100% of unwounded pods of GNV 164 and GNV 360. Low incidences of infection were observed in unwounded pods of ICS 95 (0–66.7%), SHRS 21 (0–75%), and SHRS 33 (0–20%). On wounded pods, all pathogen species produced similar size lesions, ranging from 1.90 to 7.60 cm four days after inoculation. Results from this study show that all three species can produce high rates of pod infection on some clones in the absence of wounds, but stem infection is less likely. In addition, this is the first report of L. theobromae infecting cacao pods and P. palmivora infecting stems in the absence of wounds.
Article
Full-text available
The Botryosphaeriales (Dothideomycetes) includes numerous endophytic, saprobic, and plant pathogenic species associated with a wide range of symptoms, most commonly on woody plants. In a recent phylogenetic treatment of 499 isolates in the culture collection (CBS) of the Westerdijk Institute, we evaluated the families and genera accommodated in this order of important fungi. The present study presents multigene phylogenetic analyses for an additional 230 isolates, using ITS, tef1, tub2, LSU and rpb2 loci, in combination with morphological data. Based on these data, 58 species are reduced to synonymy, and eight novel species are described. They include Diplodia afrocarpi (Afrocarpus, South Africa), Dothiorella diospyricola (Diospyros, South Africa), Lasiodiplodia acaciae (Acacia, Indonesia), Neofusicoccum podocarpi (Podocarpus, South Africa), N. rapaneae (Rapanea, South Africa), Phaeobotryon ulmi (Ulmus, Germany), Saccharata grevilleae (Grevillea, Australia) and S. hakeiphila (Hakea, Australia). The results have clarified the identity of numerous isolates that lacked Latin binomials or had been deposited under incorrect names in the CBS collection in the past. They also provide a solid foundation for more in-depth future studies on taxa in the order. Sequences of the tef1, tub2 and rpb2 genes proved to be the most reliable markers. At the species level, results showed that the most informative genes were inconsistent, but that a combination of four candidate barcodes (ITS, tef1, tub2 and rpb2) provided reliable resolution. Furthermore, given the large number of additional isolates included in this study, and newly generated multigene DNA datasets, several species could also be reduced to synonymy. The study illustrates the value of reassessing the identity of older collections in culture collections utilising modern taxonomic frameworks and methods. Citation: Zhang W, Groenewald JZ, Lombard L, et al. 2021. Evaluating species in Botryosphaeriales. Persoonia 46: 63–115. https://doi.org/10.3767/persoonia.2021.46.03.
Article
Full-text available
Common fig (Ficus carica L.), an economically-important fruit species widely cultivated in the Mediterranean Basin, is gaining increasing interest in the tropical and subtropical regions. A vast genetic diversity of common fig is available; however, systematic studies of cultivar performance in a particular new environment are scarce. The objective of this study was to evaluate the performance of 17 common fig cultivars grown under the Malaysian tropical climate. Evaluation was based on 15 parameters relevant to fruit yield, fruit quality and pest susceptibility. The results showed variable performance among the cultivars tested; 10 cultivars performed well for orchard establishment. Cultivar ‘A134’ produced the highest fruit yield per plant (3410 g) compared to the average yield of the cultivars studied (2500 g). Cultivars ‘Fen Chan Huang’ and ‘Wuhan’ produced the highest number of fruits per plant, 119 and 82 fruits, respectively. Mealybug infestation and ostiole-end splitting are major limitations to fig cultivation in Malaysia. Cultivar ‘Longue d’Aout’ was the most susceptible to mealybug infestation; whereas cultivar ‘B110’ was the most resistant. Above 60% of the fruits of cvs. ‘Brunswick’ and ‘Qing Pi’ had split ostiole; however, this was less than 20% in cvs. ‘Longue d’Aout’ and ‘Alma’. This study facilitates successful farming of figs in Malaysia and provides insights for future development programs.
Article
Fig (Ficus carica) is an exotic deciduous plant that is grown worldwide including Malaysia. Fungal diseases pose a major threat to fig plants, affecting their fruit quality and production. This study was conducted to characterize the fungal isolates associated with leaf blight, stem rot, and fruit rot of F. carica in Malaysia through morphological analysis, DNA sequencing, multigene phylogenetic analysis, and pathogenicity tests. From September 2018 to March 2019, 30 blighted leaves and 30 rotted stems and fruits of F. carica were collected from several nurseries in Malaysia. Thirty fungal isolates which belonged to Lasiodiplodia theobromae (27 isolates) and Lasiodiplodia brasiliensis (3 isolates) were identified based on morphological characteristics, comparison of DNA sequences and phylogenetic analysis of the internal transcribed spacer (ITS), elongation translation factor 1‐alpha (tef1‐α), beta‐tubulin (tub2), and DNA‐directed RNA polymerase II subunit (rpb2). Among the 27 isolates of L. theobromae, 9 isolates were obtained from leaves, 8 isolates from stems, and 10 isolates from fruits, whereas the three isolates of L. brasiliensis were obtained from stems (2 isolates) and a leaf (1 isolate). The results of pathogenicity tests revealed that L. theobromae and L. brasiliensis isolates were responsible for leaf blight and stem rot of F. carica, whereas fruit rot was caused by L. theobromae isolates. The present study highlighted two different species, Lasiodiplodia theobromae and L. brasiliensis, as the causal agents of leaf blight and stem rot of F. carica. Additionally, L. theobromae caused fruit rot of F. carica in Malaysia.
Article
Sacha inchi (Plukenetia volubilis L.) belongs to the family Euphorbiaceae. It is a perennial wooden oilseed crop, and also exhibits a good source of polyunsaturated fatty acids, protein and other bioactive compounds, such as tocopherols, carotenes and phytosterols (Chirinos et al. 2013). During 2017-2018 survey, canker disease showing greyish-brown sunken lesions was observed on the branches of sacha inchi in Danzhou campus, Hainan University, China. The disease incidence is less than 5%. However, it can lead to leaf yellowing, wilt, and eventually the whole plant death. In Nov. 2017, twelve branches showing the typical canker symptoms were collected and covered with parafilm at both ends of all samples to prevent desiccation and placed in black plastic bags keeping at 4°C until isolations were made. Samples were rinsed with tap water and dried with paper towels. Fragments, 5mm in length and cut from the junction of diseased and healthy parts of branches, were surface-sterilized with 2% sodium hypochlorite solution for 2 min, rinsed with sterilized distilled water for 5 times, dried by sterilized filter paper, plated on PDA medium amended with 100 μg/mL streptomycin (PDA-str) and incubated in the dark for 4 days at 28°C. Pure cultures of fungal isolates were obtained by transferring mycelial fragments from colony margins onto fresh PDA plates and incubated as described before. The colonies of cultures were initially white, and eventually turned black after 4 days on PDA medium (Fig S1A). The morphology characterization of conidia produced by the isolates was initially hyaline and aseptate (Fig S1B), and a single median septum formed in the mature conidia (Fig S1B). The average size of 50 conidia was 16.39±1.46ⅹ 8.52±0.92μm for J6, and 15.64±1.73ⅹ 8.94±0.86μm for J3. Three genes were used for phylogenetic analysis (Alves et al. 2006). ITS regions and the partial of TUB (β-tubulin gene) were amplified using the primer pairs ITS1 and ITS4 (White et al. 1990), Bt2a and Bt2b (Glass and Donaldson 1995), respectively, and EF1-688F/EF1-1251R for J3 and EF1-728F/EF1-986R for J6 were used to amplify TEF (translation elongation factor 1-alpha) (Alves et al. 2008). The sequences of ITS, TUB and TEF from J3 and J6 were deposited in Gene-Bank (Table S1). The blast searches in Gene-Bank with ITS, TUB and TEF amplified from isolates J3, respectively, revealed 100, 99, and 100% identities with L. pseudotheobromae, and isolate J6 showed 100, 100 and 99% of identity with L. theobromae. The phylogenetic analysis of the combined ITS, TUB and TEF sequences of J3, J6 and 28 reference strains retrieved from Gene-Bank was performed using the program MEGA 6.0 evaluated by 1000 bootstrap replications, and the result was consistent with the conclusion above (Fig S2). With the phylogenic studies supported by morphological characters, J3 was identified as L. pseudotheobromae and J6 was L. theobromae. For the pathogenicity test, J3 and J6 were used to inoculate 4-week-old healthy sacha inchi potted seedlings. One wound about 5 mm in depth per seedling stem was made using a sterile blade. A 5-mm-diameter mycelium plug of each isolate taken from the edge of 4-day-old culture growing on PDA was placed to the freshly wound of each plant stem and the inoculated area was wrapped with Parafilm. Sterile PDA plugs were placed onto the wounds of control seedlings. Nine healthy seedlings were inoculated with each isolate or PDA plugs in a completely randomized design. After inoculation, plants were placed in a greenhouse at room temperature (26 to 30°C, 80% RH) and were irrigated when needed. The experiment was conducted twice. Five days later, black or dark-brown canker lesions formed on the stems of inoculated plants, and expended upward and downward from the inoculation points. Pycnidia produced on the necrotic regions and were used to to observe the morphology of conidia (Fig S3). The fungus L. pseudotheobromae or L. theobromae can be re-isolated from the inoculated plants, but not from the control ones. L. pseudotheobromae was recorded to be collected from dead leaves of P. volubilis in Yunnan Province, China, but did not prove this fungus to be pathogenic (Tennakoon et al. 2016). This is the first report that L. theobromae and L. pseudotheobromae causing stem canker in sacha inchi in Hainan, China. The results pave the way for the development of management strategies for canker disease in sacha inchi.
Article
Aims: To identify Botryosphaeriaceae fungal species that are associated with stem-end rot of mango, and to study their pathogenicity on mango fruit. Methods and results: Based on the sequences of ITS, TEF1-α, and β-tubulin, as well as on the phylogenetic analysis of combined sequences, four species of Lasiodiplodia (L. theobromae, L. pseudotheobromae, L. iranensis, L. mahajangana) and two species of Neofusicoccum (N. ribis, N. parvum) were identified. Pseudofusicoccum violaceum, Neoscytalidium dimidiatum and three species of Botryosphaeria (B. scharifii, B. dothidea, B. ramosa) were identified based on sequences of ITS and TEF1-α. Pathogenicity test of selected isolates were tested on Chok Anan, Waterlily, and Falan mango cultivars. Generally, all species were observed to be pathogenic on the three tested mango cultivars on wounded fruits, except for N. ribis and N. parvum, which were pathogenic on both wounded and unwounded fruits. However, N. ribis was only pathogenic on cultivar Falan, whereas B. ramosa were pathogenic on cultivars Waterlily and Falan. Conclusions: Eleven species of Botryosphaeriaceae were associated with mango stem-end rot in Malaysia. To the best of our knowledge, four species, namely L. mahajangana, B. ramosa, N. ribis, and P. violaceum are the first recorded Botryosphaeriaceae fungi associated with stem end rot of mango. Significance and impact of the study: The identification of Botryosphaeriaceae fungi is important to establish suitable control measures and quarantine requirements. Many species have a wide host range, which means that there is a possibility of cross infection from other infected plants.