Вы находитесь на странице: 1из 10

Plant Pathology (2003) 52, 476 485

Phylogeny of the frosty pod rot pathogen of cocoa


Blackwell Publishing Ltd.

H. C. Evansa*, K. A. Holmesa and A. P. Reidb


a
CABI Bioscience, Ascot Centre, Silwood Park, Ascot, Berkshire SL5 7TA; and bCABI Bioscience, Egham Centre, Bakeham Lane, Egham,
Surrey TW20 9TY, UK

Morphological, cytological and molecular evidence is presented which confirms that the frosty pod rot pathogen
of cocoa, formerly classified as the mitosporic fungus Moniliophthora roreri (Deuteromycota), belongs to the hymeno-
mycetous genus Crinipellis (Basidiomycota) and that two varieties should now be recognized: Crinipellis roreri var.
roreri and the new variety C. roreri var. gileri. The latter was collected on Theobroma gileri, an endemic tree of
submontane forests in north-west Ecuador, and can be distinguished from Ecuadorian and Peruvian isolates from cocoa
(T. cacao) on the basis of spore morphology, incompatibility and nucleotide sequence data. As with var. roreri, meiosis
is shown to occur within the dispersive and infective spore stage of var. gileri and these meiospores are interpreted
to represent a much modified probasidium. In addition, in a field inoculation experiment, an isolate from T. gileri
proved to be noninfective to cocoa pods when compared with positive control strains isolated from T. cacao in western
Ecuador and T. bicolor in eastern Ecuador. It is concluded that var. gileri is the vestigial progenitor of the frosty pod
rot pathogen of cocoa, with a host range and distribution restricted to T. gileri in the mesic forests of north-west South
America.

Keywords: characterization, Crinipellis, evolution, frosty pod rot, Moniliophthora, Theobroma

Leach et al., 2002), the disease has now been confirmed


Introduction in Nicaragua, Honduras and Guatemala (Evans, 2002),
Frosty pod rot of cocoa (Theobroma cacao) is a relatively and it must be only a matter of time before it reaches
minor disease in global terms, accounting for an estimated the more extensive and ancient cocoa plantations in
30 000 tonnes in loss of cocoa production, or less than Mexico. Wherever the pathogen has invaded, cocoa pro-
4% of total crop losses (Bowers et al., 2001). Thus, the duction has been severely affected, with, for example, an
disease has been of limited interest and, until recently, estimated 4050% fall in production in Peru (Evans et al.,
largely confined to its purported centre of origin in western 1998). The socioeconomic and ecological implications
Ecuador and /or Colombia (Holliday, 1971; Evans, could be profound. In Peru and Bolivia, cocoa constitutes
1981). It has been further speculated that the pathogen an important component in the campaign against illicit
evolved on the endemic forest host, Theobroma gileri drug crops. If cocoa farming were to become uneconomic
(Holliday, 1971; Evans, 2002). The disease, however, is due to disease pressure, then such commodity crops
currently in an invasive phase, having arrived in and would be abandoned in favour of the more lucrative coca
spread throughout Peru within the last decade; posing a (Erythroxylum coca). In both Central America and Brazil,
direct threat to the neighbouring cocoa-growing countries the loss of a sustainable, forest-shaded tree crop, such as
of Bolivia and Brazil (Evans et al., 1998; Evans, 2002). cocoa, and the predicted switch to annual, subsistence or
Similarly, following its arrival in Costa Rica in the late cash crops could spell ecological disaster as the forest
1970s (Enriquez & Suarez, 1978), and its subsequent and corridors for bird migration in Meso-America, and the
continuing dramatic impact on cocoa yields (Evans, 1986; forest refuges for endemic mammals in Brazil, especially
in Bahia state, disappear (Rice & Greenberg, 2000;
*To whom correspondence should be addressed. Saatchi et al., 2001).
E-mail: h.evans@cabi.org The main objective of the current study was to test the
theory that the pathogen which devastated Ecuadorian
Present address: Scottish Agricultural Science Agency,
cocoa plantations during the early part of the 20th
Diagnostics and Molecular Biology, 82 Craigs Road, East
century (Rorer, 1926) originated on T. gileri in the sub-
Craigs, Edinburgh, EH12 8NJ, UK
montane forests of north-west Ecuador (Evans, 1981;
Accepted 4 March 2003 Evans, 2002).

476 2003 BSPP


Frosty pod rot of cocoa 477

Table 1 Geographical and host origins of frosty pod rot isolates used in the study

Isolate code Locality Country Date Host Habitat

DIS 67 Nueva Primavera, Rio Napo, Ecuador March 1999 Theobroma bicolor Secondary forest
Napo Province
DIS 116 Guadual, Esmeraldas Ecuador September 1999 T. gileri Primary forest
Province
DIS 122 Anangu, Rio Napo, Ecuador September 1999 T. cacao Plantation
Napo Province
DIS 160 Juanjui, Rio Huallaga, Peru November 1999 T. cacao Plantation
San Martin Department
DIS 329a Guadual, Esmeraldas Ecuador November 2001 T. gileri Primary forest
Province
DIS 331 Guadual, Esmeraldas Ecuador November 2001 T. gileri Primary forest
Province

a
Isolated as an endophyte from an immature, symptomless pod; all other isolates from diseased pods with symptoms of frosty pod rot.

mutual repulsion (barrage reaction) occurred between the


Materials and methods various isolates. These were evaluated by taking 5-mm-
diameter plugs from 10-day-old colonies and transferring
Fungal isolates
them to MEA in 5-cm-diameter plates in paired or quad-
The origins of the isolates used in the present study are rupled combinations of the isolates from T. cacao and
listed in Table 1. Those from T. gileri, the suspected wild T. gileri listed in Table 1, followed by incubation at 25C
host, were collected from remnant, submontane forest in for 30 days in the dark.
north-west Ecuador, using published (Cuatrecasas, 1964)
as well as unpublished (herbarium) data to locate the
Field inoculation
holotype site. All isolates were made by plating pod tissues
directly on potato carrot agar (PCA) supplemented The experiment was carried out in April 2002 at the
with 10 mg mL1 penicillinstreptomycin solution (Sigma- Estacin Experimental Tropical (EET), Pichilingue, Los
Aldrich, Poole, UK) or malt extract agar (MEA) contain- Rios Province, Ecuador, in a 40- to 50-year-old, mixed
ing benomyl (5 mg mL1) and incubating at 25C in the hybrid cocoa plantation. Approximately 1-month-old
dark. pods (cherelles), 35 cm in length, were inoculated by
soaking a pad (2 cm in diameter) of absorbent cotton
wool in an aqueous spore suspension (1 108 spores
Spore morphology
mL1) of the selected isolate and wrapping this centrally
Mature spores were shaken from 10-day-old colonies around the pod. All inoculated pods, including controls
grown on MEA at 25C in the dark, and stained in treated with sterile distilled water only, were immediately
aceto-carmine. Measurements of the length and breadth enclosed in clear plastic bags (30 20 cm) and fastened
were made from 200 spores for each isolate and their to the peduncle with a plant tie, in order to maintain
mean size and range calculated. The length:breadth ratios humidity and prevent contamination during the 4- to 5-
were compared by analysis of variance (anova) after month growth and maturation period. Healthy mature
angular transformation using Genstat 5 (Payne et al., and infected pods were harvested and assessed macro-
1993). scopically for frosty pod rot symptoms. For those diseased
pods with no evident external sporulation, samples were
plated onto potato dextrose agar (PDA) to determine the
Spore cytology
identity of the causal agent.
A modified Giemsa method (Punithalingam, 1983) was
used to stain the nuclei of spores in various stages of devel-
PCR amplification of the ITS region
opment, including mature spores pregerminated on PCA
for 2448 h. In order to ensure that sufficient material Amplification of the ITS region of the rDNA operon was
adhered to the staining slides, spores were spread over the carried out by direct PCR from mycelia growing on agar
slides in a thin film of sterile distilled water and allowed plates. A small section of mycelium was scraped off each
to air dry slowly at 2025C for a minimum of 6 h. plate and placed into sterile 05 mL thin-walled PCR
tubes. An equal volume (20 L) of lysis buffer (LB;
10 mm Tris, pH 83, 50 mm KCl, 25 mm MgCl2, 045%
Compatibility reactions
NP40, 045% Tween 20, 001% gelatin, 60 g mL1
In order to detect genotype differences, compatibility or Proteinase K) was added and the tubes frozen at 80C
noncompatibility was assessed by observing if aversion or for 10 min. These were then incubated at 65C for 60 min

2003 BSPP Plant Pathology (2003) 52, 476 485


478 H. C. Evans et al.

followed by 10 min at 95C and 5 L added to a fresh Clones containing inserts were then grown overnight
thin-walled tube containing 45 L PCR reaction mix in 5 mL LB containing 50 g mL1 ampicillin at 37C
(10 mm Tris, pH 83, 50 mm KCl, 25 mm MgCl2, 2 mm in an orbital incubator at 200 rpm and the plasmids
dNTP (Gibco Life Sciences, Paisley, UK), 50 pmol of each purified using a Wizard Plus Minipreps DNA Purifica-
primer and two units of Taq polymerase (Sigma)). The tion System (Promega) according to the manufacturers
primers used were ITS 1F (White et al., 1990) and ITS 4 instructions. The purified plasmids were sequenced
(Gardes & Bruns, 1993) (Amersham Biosciences UK Ltd, using a SequiTherm Excel II DNA Sequencing Kit
Little Chalfont, UK). Cycling conditions were 95C for (Cambio Ltd, Cambridge, UK) with labelled M13 primers
5 min followed by 35 cycles of 95C for 1 min, 50C for obtained from MWG-Biotech (Ebersberg, Germany).
45 s and 72C for 2 min; a final extensions step of 72C Electrophoresis was carried out on a Li-Cor DNA Gene
for 10 min was included and the reactions held at 10C Reader 4200 (Li-Cor Biosciences UK Ltd, Cambridge,
until purification. Amplification was carried out in a UK).
Hybaid PCR Express Thermal Cycler (Hybaid Ltd,
Teddington, UK).
Data analysis
Forward and reverse sequences of clones from each strain
Cloning and sequencing of the PCR products
were assembled using Sequencer v4.0.5 (Gene Codes
The ITS PCR products were purified using a Wizard PCR Corporation, Ann Arbor, MI, USA). Alignment of the
Preps Purification System (Promega UK Ltd, Southampton, consensus sequences was carried out using the ClustalW
UK) and cloned using a pGEM-T Easy Vector System function in MacVector v70r2 (Accelrys Ltd, Cambridge,
II (Promega) according to the manufacturers instruc- UK) using the slow alignment speed. Pairwise matrices
tions. A number of colonies were picked from each plate were obtained by use of the Pustell DNA matrix function
and grown overnight in 1 mL LB containing 50 g mL1 within MacVector with a window size of 30 base pairs
ampicillin at 37C in an orbital incubator at 200 rpm. (bp) and a minimum percentage score of 60%. Phyloge-
The presence of the correct insert was checked by heating netic analysis was done by upgma analysis with 100 boot-
10 L from each culture at 95C in a thin-walled tube strap replicates. Details of the isolates used in the analysis,
for 10 min and adding 15 L PCR mix and amplifying with sequence data from GenBank, are presented in
as detailed earlier. Table 2.

Table 2 Details of the fungal isolates used in


Fungus Isolate code GenBank accession number
the ITS sequence phylogenetic analysis. All
Frosty pod rot* DIS 160 AY230254 obtained from GenBank, apart from those
Frosty pod rot* DIS 331 AY230255 marked with an asterisk (*)
Alternaria alternata AA1 AF314569
A. infectoriaa BMP 21-11-15 AF229458
A. sesamicola AZM AF314588
A. smyrnii EGS 37-093 AF229456
A. tenuissima ATCC 16423 AF229476
Cladosporium macrocarpum CBS 17562 AJ244229
C. sphaerospermum CBS 12247 AJ244228
Colletotrichum acutatum BBA 71427 AJ301987
C. gloeosporioides b BBA 65797 AJ301925
C. gloeosporioides BBA 71407 AJ301986
C. lindemuthianum BBA 65483 AJ301958
Crinipellis perniciosa AF335590
Gaeumannomyces sp. sp. 153 AJ010038
Magnaporthe poae C4 AF0774403
M. poae 1832 AJ010041
M. poae 2562 AJ010042
M. grisea 2690 U17328
M. grisea 2692 U17329
Phialophora graminicola P4 U17217
Pyricularia rabaulensis IMI 388508
Trichoderma reesii ATCC 56765 X93938
T. saturnisporum CBS 33592 X93973
T. virgatum ATCC 24961 Z48949
T. viride Tr 22 AJ230678

a
As teleomorph, Lewia infectoria.
b
As teleomorph, Glomerella cingulata.

2003 BSPP Plant Pathology (2003) 52, 476 485


Frosty pod rot of cocoa 479

Figure 1 Frosty pod rot pathogen on hosts and in culture: (a) on mature pod of Theobroma gileri, in understorey submontane forest, Ecuador (isolate
DIS 116), healthy green pod indicated by arrow in canopy background; (b) same locality, healthy pod of T. gileri from which DIS 329 was isolated as
an endophyte; (c) same locality, isolate DIS 331 on immature pod showing mature brown spore mass developing on white pseudostroma; (d) healthy
and diseased pods in cocoa plantation, Huallaga Valley, Peru, showing white pseudostroma covering pod on which immature (pink) spores are
developing; (e) same locality, external and internal pod symptoms, showing more advanced stage in sporogenesis with brown, flaking spore masses
beginning to obscure white pseudostroma, and compacted beans surrounded by white mycelium; (f ) dual plate culture (MEA) of two cocoa isolates,
DIS 122 ex Ecuador and DIS 160 ex Peru; (g) cocoa isolate (DIS 160) challenged with T. gileri isolate (DIS 331); (h) T. gileri isolate (DIS 116)
intermingling centrally from lateral inoculum plugs and inhibiting the growth of cocoa isolate (DIS 160) from proximal and distal plugs.

2003 BSPP Plant Pathology (2003) 52, 476 485


480 H. C. Evans et al.

Table 3 Microscopic analysis of frosty pod rot spores from Theobroma gileri and T. cacao

Isolate code Host Spore lengtha (m) Spore breadth (m) Overall mean (m)

DIS 116 T. gileri (75)95 160 (180) 55 90(100) 1113 701


DIS 329 T. gileri (75)110 160(220) 55 80(105) 1186 696
DIS 331 T. gileri (70)80 155(175) (50 )60 110 1071 844
DIS 122 T. cacao (55)65 105(120) 55 90(100) 812 720
DIS 160 T. cacao 55 105(125) 50 80(100) 822 735

a
Minimum of 200 spores.

Table 4 Comparison of spore sizes of frosty pod rot isolates from Table 5 Field evaluation of pathogenicity to cocoa (T. cacao) pods of
Theobroma gileri and T. cacao frosty pod rot isolates from Theobroma spp.

Mean spore Angular Isolate code Host Pods infectedb


Isolate length:breadth ratioa transformation
DIS 67 T. bicolor 9 (9)
T.gileri (DIS 329) 1730 7435ab DIS 116 T. gileri 0 (10)
T.gileri (DIS 116) 1651 7581a EET 1a T. cacao 16 (20)c
T.gileri (DIS 331) 1551 7288a Control 1d (10)
T.cacao (DIS 160) 1164 6063b
a
T.cacao (DIS122) 1143 6117b Source was an infected pod, collected on day of inoculation, within an
experimental plot.
a b
Means of 200 spores. Number in parentheses shows total pods inoculated.
b c
Values followed by the same letter are not significantly different Remainder infected by Crinipellis perniciosa.
d
according to the least significant difference test (P = 005). Natural infection from field inoculum.

bicolor, but the mistake was only detected when the seem-
Results ingly ambiguous results were being analysed. Thus, there
were insufficient data to assess pathogenicity critically,
Spore form and function
but provisional evidence indicated that T. gileri isolates
Macroscopic morphology of the frosty pod rot pathogen were nonpathogenic to cocoa, whilst cocoa isolates were
on T. cacao and T. gileri is illustrated in Fig. 1(a,ce). cross-infective within the genus Theobroma (Evans,
Microscopic analysis showed that spores of the isolates 1981).
from T. gileri differed significantly from the cocoa isolates
(Tables 3 and 4), a high proportion of the T. gileri spores
Phylogenetic analysis
being ellipsoidal to cylindrical and relatively thin-walled
(Fig. 2b), compared with the predominantly globose, The consensus sequence for the T. cacao strain (DIS 160)
thick-walled spores from T. cacao (Fig. 2a). was 795 bp in length and 796 bp for the T. gileri strain
Nuclear staining of the sporophores and the mature (DIS 331). The alignment of the two sequences (Fig. 3)
and germinating spores from T. gileri revealed evidence of showed that the two strains were 988% homologous
a meiotic cycle (Fig. 2ch), as recently reported for cocoa with only nine polymorphic sites; all of these sites were
isolates (Evans et al., 2002), and therefore that they func- consistent within the various clones sequenced. Most of
tion as meiospores rather than mitospores as previously these took the form of single nucleotide polymorphisms,
interpreted. although there was one region containing four polymor-
phisms in a 14-bp region. In the cocoa isolate, at bases
460 474, the sequence was 5-AAAAAAGCTTTTTC-3,
Isolate compatibility
but in T. gileri it was 5-AAAAAGCTTTTTTT-3. It is
Paired cocoa isolates produced a compatible reaction possible that these two regions form a hairpin structure
as did T. gileri isolates (Fig. 1f), whereas incompatibility, within the 58S rRNA gene. However, upon aligning
or a barrage reaction, was consistently demonstrated in both sequences with the ITS regions of numerous fungal
mixed combinations, with well-defined demarcation lines sequences from GenBank, this region was just outside
between isolates from the two different hosts (Fig. 1g h). the gene. The closest match was with Crinipellis perniciosa
(Stahel) Singer, the causal agent of witches broom disease
of cocoa, which displayed 89% homology.
Pathogenicity
Pairwise comparisons of the frosty pod isolates (Fig. 4a),
The results of the field inoculation experiment are shown showed a high degree of similarity, whilst those between
in Table 5. Due to confusion over isolate numbering, only the cocoa isolate (DIS 160) and Crinipellis perniciosa
one isolate from T. gileri was tested. The second isolate (Fig. 4b) revealed the polymorphisms, the majority of
turned out to be from another Theobroma host, T. which occurred in the ITS 2 region.

2003 BSPP Plant Pathology (2003) 52, 476 485


Frosty pod rot of cocoa 481

Figure 2 Spore morphology and sporogenesis of the frosty pod rot pathogen: (a) cocoa isolate (DIS 122) with predominantly globose, thick-walled
spores; (b) T. gileri isolate (DIS 116) with both globose and ellipsoidal, thinner-walled spores. (ch) Sporogenesis in T. gileri isolate DIS 116: (c) early
stage (prophase), showing erect sporophores surrounded by an amorphous layer (arrow) with newly formed diploid nuclei, the recent narrow hyphae
of the pseudostroma are binucleate; (d) swelling of sporophore with first meiotic division (anaphase); (e) binucleate, globose to ellipsoidal mature
spores with abnormally large cylindrical spore in foreground in late telophase stage of meiosis; (f) meiospore germinating after 48 h on PCA, showing
a four-celled metabasidium, with distal and apical cells reverting to a binucleate condition; (g, h) disrupted meiospore chains exhibiting various
stages of nuclear division; the majority of meiospores mature in the binucleate condition with the second meiotic divisions occurring during
germination. (All scale bars = 10 m.)

A phylogenetic analysis with 24 other fungal sequences


yielded a tree (Fig. 5) which clustered the frosty pod rot
Discussion
isolates with Crinipellis perniciosa, a relationship with a Until very recently, the phylogeny of the frosty pod rot
100% bootstrap support. pathogen of cocoa was based on a few facts, supported by

2003 BSPP Plant Pathology (2003) 52, 476 485


482 H. C. Evans et al.

Figure 3 Aligned nucleotide sequences of frosty pod


rot isolates DIS 160 (ex T. cacao, GenBank accession
no. AY230254) and DIS 331 (ex T. gileri, GenBank
accession no. AY230255).

several assumptions and some speculation. The facts unknown basidiomycete. Later, Evans (1981) speculated
were established by Evans et al. (1978), who reported that that the teleomorph could be a Crinipellis species since
the fungus had basidiomycetous origins rather than the the unusual pleomorphic, hemibiotrophic life-cycle is
ascomycetous (Leotiales) origin that the name Monilia mirrored by that of the other major South American
roreri Ciferri suggested. The evidence was obtained from cocoa pathogen, Crinipellis perniciosa (Evans, 1980): the
scanning and transmission electron microscopic studies former evolved on an endemic Theobroma host, or its
which revealed the unique sporogenesis, basipetal rather near relative Herrania, on the western or Pacific side of
than acropetal as in Monilia, and the presence of dolipore the Andean Cordillera (Rorer, 1926; Holliday, 1971;
septa in the mycelium. The new genus Moniliophthora was Evans, 1981), and the latter on T. cacao, in its native range
erected to accommodate the pathogen. It was assumed on the eastern or Amazonian side (Baker & Holliday,
that this represented the mitotic state (anamorph) of an 1957).

2003 BSPP Plant Pathology (2003) 52, 476 485


Frosty pod rot of cocoa 483

University of Wales Aberystwyth, Dyfed, UK, personal


communication, 2001), and also of mitochondrial DNA
(W. Phillips, University of Reading, UK, personal commu-
nication, 2001), has revealed that the fungus does indeed
lie very close to C. perniciosa in the respective phyloge-
netic trees. However, the assumption that the sporophores
and spores represent the anamorph has now been found
to be wanting. Using more traditional cytological methods,
it has been discovered that the purported conidia or mit-
ospores are, in fact, meiospores and that the frosty pod rot
pathogen of cocoa is a much modified agaric teleomorph.
The new combination Crinipellis roreri (Ciferri) H.C.
Evans has now been proposed, the genus Moniliophthora
being reduced to synonymy with Crinipellis (Evans et al.,
2002). The cytological results presented here (Fig. 2ch)
serve to confirm this behaviour, and also that a similar
meiotic cycle occurs during sporogenesis and germination
in the forest isolates from T. gileri.
The discovery of C. roreri on pods of T. gileri in its type
locality (Fig. 1a and c) gave initial support to the hypoth-
Figure 4 Pustell DNA matrices for pairwise comparisons of frosty pod
esis that the pathogen which ravaged cocoa plantations
rot isolates DIS 160 (ex T. cacao) and DIS 331 (ex T. gileri) (a) and frosty
in western Ecuador in the early 1900s originated on this
pod rot isolate DIS 160 and Crinipellis pernicosa (b).
forest host.
Subsequent morphological examination, however, casts
doubt on this and significant differences in spore form
The increasing invasiveness of the frosty pod rot were found between the isolates from T. gileri and T.
pathogen and its actual impact on and potential threat cacao (Tables 3 and 4, Fig. 1a,b). Those from the former
to cocoa cultivation in Latin America have offered the produce a much higher proportion (30 40%) of thin-
opportunity to test these assumptions and hypotheses walled, ellipsoidal to cylindrical spores compared with the
using modern molecular techniques. The work reported cocoa isolates (12%). Moreover, the molecular studies
here, as well as in recent studies on the ITS region of ribos- also showed that the Ecuadorian T. gileri strain differs
omal RNA (G.W. Griffith, Institute of Biological Sciences, in a number of base pairs from a selected T. cacao strain.

Figure 5 UPGMA tree derived from fungal ITS


sequence data. Bootstrap values displayed at
the nodes. *DIS 160, isolated from T. cacao;
**DIS 331, isolated from T. gileri.

2003 BSPP Plant Pathology (2003) 52, 476 485


484 H. C. Evans et al.

This morphological and molecular evidence, as well as endemic species. Thus, the Andean uplift which sepa-
that from ongoing work on the genetic diversity of cocoa rated the Choc and Amazonian regions in the mid-
isolates of frosty pod rot (W. Phillips, University of Read- Pleistocene (Gentry, 1982) resulted not only in the
ing, UK, personal communication, 2002), suggests that evolutionary isolation of the genus Theobroma, but also
the T. gileri strain in Ecuador has never moved from its of its mycobiota: the frosty pod rot pathogen evolved to
forest habitat. the west and witches broom in the much more extensive
The results of the field study lend support to this isola- forests to the east, possibly from a common endophytic
tion theory since there was no cross-infection. However, a Crinipellis ancestor. Indeed, isolate DIS 329 was isolated
statistical analysis was not undertaken because of back- as an endophyte from a healthy pod of T. gileri during the
ground contamination, mainly by C. perniciosa, and present study (Fig. 1b). Both pathogens appear to have
insufficient replicates in the T. gileri treatment. Initially, extended their host colonization abilities by invading mer-
the experimental design included two T. gileri isolates, but istematic tissues, causing auxin imbalances in the form of
due to confusion over code identification, an isolate from hypertrophy and hyperplasia. Although the frosty pod rot
T. bicolor was substituted erroneously. The latter (DIS 67) fungus can induce such symptoms in cocoa seedlings at
was obtained from a forest species endemic in eastern high spore loads (Evans, 1981), in the field situation it is
Ecuador, where the pathogen is thought to have arrived in strictly a pod pathogen, whereas C. perniciosa can invade
the late 1970s (Evans, 1986; Evans et al., 1998). A much all the meristems of cocoa leading to a multiplicity of
larger field experiment, involving 100 pods per treatment symptoms (Baker & Holliday, 1957).
and additional isolates from T. gileri, is being established The ice-age isolation of the Choc refuge (Gentry,
to confirm these results. 1982), resulting in the paucity of both Theobroma and
The reason why the cocoa strain or isolate can attack Herrania hosts compared with Amazonia (Schultes, 1958;
not only most of the cocoa clones and varieties against Cuatrecasas, 1964; Thorold, 1975), may have provided
which it has been screened, but also all of the species of the selection pressure whereby the agaric basidiomata
Theobroma and Herrania in germplasm collections where evolved to maximize the production of a multipurpose
the disease occurs (Evans, 1981), can now be explained by meiospore, which, unlike the basidiospore, has a survival
the presence of a sexual stage in the life-cycle. What was as well as a dispersal function, adapted for efficient long-
once thought to be a clonal pathogen, which cycled only distance movement. This would have been essential for a
through mitospores, has now been proven to be geneti- pathogen of a low-density, understorey forest tree such as
cally diverse within its native and exotic ranges, with the T. gileri, with a disparate range from northern Colombia
current T. gileri isolates forming a separate clade from all along the western slopes of the Andes to northern Ecua-
the cocoa isolates, including those from western Ecuador dor. Certainly in the latter country, T. gileri is confined to
(W. Phillips, University of Reading, UK, personal commu- submontane forests, between 500 and 700 m above sea
nication, 2002). Whether the sexual mechanism involves level (asl), and this species appears to be unknown either
outbreeding through heterothallism has yet to be estab- side of this narrow strip, as well as from the forests to the
lished; however, by virtue of the fact that many billions of south of Pichincha Province (unpublished observations).
meiospores are produced per pod (Fig. 1e) estimated at The thick white pseudostroma covering the diseased
44 106 cm2 (Evans, 1981) suggests that sufficient gene pod prior to sporogenesis, and which imparts the frosted
movement is possible through recombination and muta- appearance (Fig. 1b and c), may represent the vestigial pileus.
tion in order to respond to selection pressures. Indeed, this Additional circumstantial evidence that the fungus on
enormous inoculum potential demonstrates why frosty T. gileri is the progenitor of the cocoa pathogen, and that
pod rot has so rapidly displaced witches broom disease as this evolved in northern Colombia, comes from historical
the major constraint to cocoa production in Peru follow- accounts of cocoa cultivation in Latin America. Baker
ing its relatively recent arrival (Evans et al., 1998). et al. (1954) suggest that frosty pod rot may have been
The question remains: if not in Ecuador on T. gileri, recorded on cultivated cocoa in northern Colombia as
then where and on what host did the cocoa pathogen long ago as 1850, and that in this area cocoa may have
evolve? The genetic diversity data (W. Phillips, University been grown for many centuries or even millennia since
of Reading, UK, personal communication, 2002) show that wild cocoa was discovered in the dense forests bordering
by far the greatest variation occurs in northern Colombia the Rio Len. The origin of T. cacao remains controver-
and, significantly, it was from this region that Baker et al. sial, but past and recent evidence strongly indicates that
(1954) first confirmed the disease on T. gileri in the forests cocoa evolved in the Upper Amazon and later became
of north-west Antioquia Department. In an earlier report, naturalized in northern Colombia and Meso-America
this host had been confused with T. microcarpum Mart. (Cheesman, 1944; Baker et al., 1954; B.G.D. Bartley, Pao
(Cope, 1953) the Amazonian vicariant of T. gileri (Cuat- de Arcos, Portugal, personal communication, 2002). Thus,
recasas, 1964) since at that time the taxon T. gileri had the ancient cultivations in northern Antioquia would have
not yet been described (Cuatrecasas, 1953). This part of been planted near to or even amongst T. gileri populations
Antioquia, together with north-west Ecuador, is included and exposed to frosty pod inoculum, with inevitable
by Gentry (1982) in the Choc phytogeographic region, adaptation to the new host through sexual selection. The
now recognized by Myers et al. (2000) as a biodiversity added selection pressure of drier microclimates in the
hotspot, due to its exceptional concentrations of cocoa plantations, compared with the buffered humid

2003 BSPP Plant Pathology (2003) 52, 476 485


Frosty pod rot of cocoa 485

forests, may have contributed to the gradual loss of the of the genus Theobroma. Contributions of the United States
thinner-walled, ellipsoidal spores and the increased pro- National Herbarium 35, 379 614.
duction of thick-walled, globose spores which distinguish Enriquez GA, Suarez C, 1978. Monilia disease of cocoa in Costa
cocoa isolates from those of T. gileri. It is considered that Rica. Turrialba 28, 339 40.
these morphological characters, in addition to the molec- Evans HC, 1980. Pleomorphism in Crinipellis perniciosa, causal
ular and pathogenicity differences presented here, justify agent of witches broom disease of cocoa. Transactions of the
taxonomic recognition at the varietal level as follows: British Mycological Society 74, 515 32.
Evans HC, 1981. Pod Rot of Cacao Caused by Moniliophthora
Crinipellis roreri (Ciferri) H. C. Evans var. gileri H.C. (Monilia) roreri. Phytopathological Papers 24. Kew:
Evans & K.A. Holmes var. nov. (Figs 1a and c, and 2bh) Commonwealth Mycological Institute, 144.
Etym: derived from its host, Theobroma gileri. Evans HC, 1986. A reassessment of Moniliophthora (Monilia)
Meiosporae catenulata, ellipsoidea vel globose, 80 160 pod rot of cocoa. Cocoa Growers Bulletin 37, 34 43.
(220) 55110 m, plerumque tenuitunicata. Evans HC, 2002. Invasive neotropical pathogens of tree crops.
Holotypus: DIS 116, in Theobroma gileri (Sterculiaceae) In: Watling R, Frankland JC, Ainsworth AM, Isaac S,
in silvestri primiario, lectus in Guadual, Esmeraldas Robinson CH, eds. Tropical Mycology, Vol. 2. Wallingford,
Provincia, in Ecuador, 14 September 1999. UK: CABI Publishing, 83 112.
Material examined: holotype DIS 116 ex Theobroma gileri Evans HC, Holmes KA, Phillips W, Wilkinson MJ, 2002. Whats
(Sterculiaceae), in primary forest, Guadual, Esmeraldas in a name. Crinipellis, the final resting place for the frosty pod
pathogen of cocoa? Mycologist 16, 148 52.
Province, Ecuador, 550 m asl, 14 September 1999;
Evans HC, Krauss U, Rios RR, Zecevich TA, Arevalo-Gardini E,
paratypes DIS 329 and 331, same host and locality, 5
1998. Cocoa in Peru. Cocoa Growers Bulletin 51, 722.
November 2001.
Evans HC, Stalpers JA, Samson RA, Benny GL, 1978. On the
This variety is distinguished from C. roreri var. roreri taxonomy of Monilia roreri, an important pathogen of
by the significantly greater proportion of thin-walled, ellip- Theobroma cacao in South America. Canadian Journal
soidal spores, which raises the mean length from 8 m in of Botany 56, 2528 32.
var. roreri, range 5512 m, to 1112 m in var. gileri, Gardes M, Bruns TD, 1993. ITS primers with enhanced
range 8 22 m. In addition, var gileri is nonpathogenic specificity for basidiomycetes, application to the identification
to cocoa and differs in its genetic profile. of mycorrhizae and rusts. Molecular Ecology 2, 113 8.
Gentry AH, 1982. Phytogeographic patterns as evidence for a
Choc refuge. In: Prance GT, ed. Biological Diversification in
Acknowledgements the Tropics. New York, USA: Columbia University Press,
112 36.
This project was carried out within USDA-ARS Alterna-
Holliday P, 1971. Some tropical plant pathogenic fungi of
tive Crops Programme (project no. 0500-00008-001-21).
limited distribution. Review of Plant Pathology 50, 33748.
We would like to thank Dr Carmen Suarez and staff at Leach AW, Mumford JD, Krauss U, 2002. Modelling
the Estacin Experimental Tropical Pichilingue, Ecuador, Moniliophthora roreri in Costa Rica. Crop Protection 21,
for invaluable assistance with the field experiment. 31726.
Myers N, Mittermeier RA, Mittermeier CG, Fonseca GAB,
Kent J, 2000. Biodiversity hotspots for conservation
References
priorities. Nature, London 403, 853 8.
Baker RED, Cope FW, Holliday PC, Bartley BG, Taylor DJ, Payne RW, Lane PW, Digby PGN, Harding SA, Leech PK,
1954. The Anglo-Colombian Cacao Collecting Expedition. Morgan GW, Todd AD, Thompson R, Wilson GT,
Report on Cacao Research, 1953. St Augustine, Trinidad: Welham SJ, White RP (eds), 1993. Genstat 5 Release 3
Imperial College of Tropical Agriculture, 8 29. Reference Manual. Oxford, UK: Clarendon Press.
Baker RED, Holliday P, 1957. Witches Broom Disease of Punithalingam E, 1983. The nuclei of Macrophomina
Cacao ( Marasmius perniciosus Stahel). Phytopathological phaseolina (Tassi) Goid. Nova Hedwigia 38, 339 67.
Papers 2. Kew: Commonwealth Mycological Institute, 1 42. Rice RA, Greenberg R, 2000. Cacao cultivation and the
Bowers JH, Bailey BA, Hebbar PK, Sanogo S, Lumsden RD, conservation of biological diversity. Ambio 29, 16773.
2001. The impact of plant diseases on world chocolate Rorer JB, 1926. Ecuador cacao. Tropical Agriculture, Trinidad
production. Plant Health Progress On-line. (http:// 3, 46 7.
www.apsnet.cor/online/feature/cacao/top.html). Saatchi S, Agasti D, Algar K, Delabie J, Musinsky J, 2001.
Cheesman EE, 1944. Notes on the nomenclature, classification Examining fragmentation and loss of primary forest in the
and possible relationships of cacao populations. Tropical southern Bahian Atlantic forest of Brazil with radar imagery.
Agriculture, Trinidad 21, 144 59. Conservation Biology 15, 85775.
Cope FW, 1953. An interim report on the Anglo-Colombian Schultes RE, 1958. A synopsis of the genus Herrania. Journal
cocoa-collecting expedition. Report of the Cocoa Conference of the Arnold Arboretum 39, 216 78.
1953. London: Cocoa, Chocolate and Confectionery Thorold CA, 1975. Diseases of Cocoa. Oxford, UK: Clarendon
Alliance, 5762. Press.
Cuatrecasas J, 1953. Une nouvelle espce de Theobroma. Revue White TJ, Bruns TD, Lee S, Taylor J, 1990. Amplification
Internationale de Botanique Appliquee dAgriculture 33, and direct sequencing of fungal ribosomal DNA genes for
5625. phylogenetics. In: Innis MA, Sninsky DH, White TJ, eds.
Cuatrecasas J, 1964. Cacao and its allies: a taxonomic revision PCR Protocols. London: Academic Press, 315 22.

2003 BSPP Plant Pathology (2003) 52, 476 485

Вам также может понравиться