Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Medical and Veterinary Entomology (2021), doi: 10.1111/mve.

12531

High prevalence and low diversity of chigger infestation


in small mammals found in Bangkok Metropolitan parks
S. A. W U L A N D H A R I 1 , Y. P A L A D S I N G 2 , W. S A E S I M 3 ,
V. C H A R O E N N I T I W A T 2 , P. S O N T H A Y A N O N 4 , R. K U M L E R T 5 ,
S. M O R A N D 2,6 , S. S U M R U A Y P H O L 1 and K. C H A I S I R I 2
1
Department of Medical Entomology, Faculty of Tropical Medicine, Mahidol University, Bangkok, Thailand, 2 Department of
Helminthology, Faculty of Tropical Medicine, Mahidol University, Bangkok, Thailand, 3 Environment Department, Bangkok
Metropolitan Administration, Bangkok, Thailand, 4 Department of Molecular Tropical Medicine and Genetics, Faculty of Tropical
Medicine, Mahidol University, Bangkok, Thailand, 5 Division of Vector-Borne Diseases, Department of Disease Control, Ministry of
Public Health, Nonthaburi, Thailand and 6 CNRS-CIRAD, Faculty of Veterinary Technology, Kasetsart University, Bangkok, Thailand

Abstract. Chiggers are recognized as vectors of scrub typhus disease caused by


the bacteria, Orientia tsutsugamushi. The risk of disease exposure is mainly related
to chigger bites when humans or animals roam into vector-infested habitats. In big
cities, urban public parks could provide areas for the animal–human interface and
zoonotic pathogen transmission. The ecology and epidemiology of urban scrub typhus
are still poorly understood in Thailand. Small mammals were trapped and examined
for chigger infestation in urban public parks across metropolitan Bangkok, Thailand.
We found a high prevalence of infestation (76.8%) with surprisingly low diversity. Two
chigger species, Leptotrombidium deliense and Ascoschoengastia indica, were identified
using morphological characteristics and molecular confirmation. The generalized linear
model identified host intrinsic variables (i.e. body mass index) with host density,
habitat composition and open field as the extrinsic factors explaining the abundance
of chigger infestation. The bacteria O. tsutsugamushi was not detected in chiggers
(90 chigger-pooled samples) and animal host tissues (164 spleen samples). However,
the existence of chigger vectors calls for the Bangkok Metropolitan Administration
and public health authorities to develop a comprehensive scrub typhus monitoring and
prevention strategy in the parks and nearby communities.
Key words. Chiggers, Bangkok, Orientia, public park, small mammals, Thailand,
urban.

Introduction intracellular bacteria Orientia tsutsugamushi (Alphaproteobac-


teria: Rickettsiaceae), the causative agent of scrub typhus dis-
Chigger is a common term for the tiny larval stages of trom- ease in the Asia Pacific region (Kelly et al., 2009). In Thailand,
biculid mites (Acari: Trombiculidae), also known as harvest scrub typhus is highly prevalent and has been realized as one
mites, scrub itch mites or redbugs. Chiggers are parasitic only of the main causes of acute febrile illness (Wangrangsimakul
at the larval stage, feeding on a wide range of vertebrates et al., 2020). Sporadic human cases have also been reported
(and incidentally on humans). At other stages (nymph and recently in other regions, such as the Middle East (i.e. the United
adult), they prey on soft-bodied arthropods and soil nema- Arab Emirates) and South America (i.e. Chile), caused by two
todes (Elliott et al., 2019). The mites are the main vectors of other bacterial species: Orientia chuto and Candidatus Orientia

Correspondence: K. Chaisiri, Faculty of Tropical Medicine, Mahidol University, 420/6 Ratchavithi Rd., Ratchathewi, Bangkok, Thailand 10400.
E-mail: kittipong.cha@mahidol.ac.th

© 2021 The Royal Entomological Society 1


2 S. A. Wulandhari et al.

chiloensis, respectively (Izzard et al., 2010; Abarca et al., 2020). farming and outdoor activities, posing a high disease burden
These highlight the recent situation of global spread of scrub (Park et al., 2015; Wangrangsimakul et al., 2020). Human scrub
typhus outside the endemic region. In addition to the role of typhus incidence is not only reported from occupational-related
scrub typhus vector, chigger bites also result in intense irri- groups such as farmers in the countryside but also occurred
tation and dermatitis (trombiculiasis) in humans and animals in the others such as the elderly, children, pregnant women,
(Santibáñez et al., 2015; Elliott et al., 2019). tourists and even in the urban population (Kim et al., 2006;
Trombiculid mites are cosmopolitan, occurring across diverse Wei et al., 2014; Weitzel et al., 2018). In Bangkok, a rapid
landscapes and habitat types, e.g. forest, scrub vegetation, increase in population, capital investment and employment
grassland, plantation, agricultural land, swamp, seacoast and has expanded the city to surrounding areas. Regarding urban
urban settings (Dohany et al., 1977; Park et al., 2015; Peng expansion, recreational green spaces are essential for providing
et al., 2018; Chaisiri et al., 2019). There are over 3000 Trom- city dwellers a good quality of life. Big cities worldwide
biculidae species worldwide, with many described in the Ori- usually provide open space or public parks for recreation
ental, Australasian and Afrotropical regions compared to other and rest from the city’s hustle and busyness. Apart from
areas (Elliott et al., 2019). Chiggers in the genus Leptotrom- that, parks are also places for biodiversity hotspots (flora
bidium are the important vectors transmitting scrub typhus to and fauna). Some synanthropic small mammals (e.g. rats,
humans (Lv et al., 2018; Elliott et al., 2019). Leptotrombidium squirrels, shrews and tree shrews) are usually present in urban
akamushi, Leptotrombidium pallidum, Leptotrombidium scutel- parks, providing interaction between humans and wildlife.
lare and Leptotrombidium deliense are the major vectors in the Zoonotic pathogens, particularly vector-borne diseases, could be
Asia Pacific region (Stekolnikov, 2013; Santibáñez et al., 2015; exchanged between humans and animals. A serological survey
Elliott et al., 2019). The first three species tend to occupy the in Bangkok suburbs revealed high human exposure to scrub
temperate region of the endemic zone, e.g. Northern and East- typhus (21% prevalence) related to contact in orchards and farms
ern China, Japan and Korea (Zhang et al., 2013; Roh et al., 2014; (Strickman et al., 1994). Wei et al. (2014) suggested that people
Yoshimoto & Yoshimoto, 2019). The latter species, L. deliense is who had activities in the city park pose a higher risk than those
responsible for disease transmission in tropical areas represented who did not. Moreover, risk behaviours, e.g. sitting on the lawn,
by Southern China, South Asia and Southeast Asian countries, close contact with wildlife and sitting near rodent burrows,
including Thailand (Rodkvamtook et al., 2013; Chakraborty & significantly increased the risk for scrub typhus in a city park
Sarma, 2017; Lv et al., 2018). (Wei et al., 2014).
In Thailand, the geographic distribution of trombiculid mites Therefore, this study aims to understand the relationship
mainly follows the information and research on human scrub among the vector (chiggers), the maintenance host (small mam-
typhus’s local epidemiology (Kumlert et al., 2018). Lak- mals) and the causative bacterial agent of scrub typhus at public
shana (1973) reported that within 30 provinces of Thailand, parks in the Bangkok Metropolitan area. This study’s specific
there were two subfamilies of chigger mites, 27 genera and 121 objectives are to (1) to investigate the diversity, prevalence and
species. Subsequently, Chaisiri et al. (2016) revised the chig- abundance of chigger infestation in small mammals; (2) to inves-
ger checklist of Thailand, reporting 99 validated species across tigate the vectorial role of chiggers in carrying Orientia bacteria
the country’s six geographical regions and emphasized lack of and (3) to identify potential factors explaining chigger occur-
information for some particular regions, e.g. the central plain, rence in the public parks of Bangkok.
including Bangkok, the capital city.
Small mammals (e.g. rodents, shrews and tree shrews) and
ground-dwelling birds play important roles as the ‘maintenance Materials and methods
hosts’, by either contributing to the abundance of chiggers (pro-
vide feeding meals) or reservoirs of the Orientia bacteria (Elliott Study sites
et al., 2019). An individual host was estimated to carry numer-
ous chiggers, ranging from 1000 to 10 000 at one time and could A cross-sectional study was conducted in Bangkok, Thai-
support up to four generations of chiggers through its lifespan land, from September to the beginning of December 2018 (wet
(Audy, 1961). Chigger infestation status (e.g. prevalence, abun- season). Small mammals were collected using live-trapping
dance and species richness) on the maintenance hosts (partic- at the selected seven sites by focusing on the green space
ularly small mammals) is influenced by host intrinsic factors, or public parks in the Bangkok Metropolitan area (see Pal-
e.g. host taxonomy, body weight, maturity, behaviour, exploita- adsing et al., 2020). These included Suan Luang Rama IX
tion habit, home range and site of attachment on the host body (SL: lat 13.686583 and long 100.663942), Suan Lumpini (LP:
(Frances et al., 1999; Chaisiri et al., 2019; Matthee et al., 2020). lat 13.731026 and long 100.541407), Suan Serithai (ST: lat
Also, extrinsic factors, such as seasonal conditions, geographical 13.786245 and long 100.672598), Suan Taweevanarom (TW:
distribution, latitudinal gradient, landscape complexity, habitat lat 13.744377 and long 100.352274), Suan Thonburirom (TB:
type, ecotone and degree of human land use, also play signifi- lat 13.651955 and long 100.491626), Suan Vareepirom (VP: lat
cant roles in chigger infestation patterns (Frances et al., 1999; 13.850450 and long 100.774289) and Suan Wachirabenchatad
Kuo et al., 2015; Moniuszko & Makol, 2016; Peng et al., 2018; (WB: lat 13.812098 and long 100.554368), (see Fig. 1A). The
Chaisiri et al., 2019; Mathee et al., 2020). parks share some common characteristics, such as vegetation,
People engaged in agriculture and habiting forested areas water reservoirs and areas for human activities. However, each
are at high risk of infection with scrub typhus disease. In park’s habitat conditions differ depending on their internal archi-
endemic areas, the risk of infection is mainly associated with tecture and external characteristics around the parks; some are

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
Chiggers from Bangkok public parks 3

(A)

(B)

Fig. 1. Geographical distribution of the study sites for chigger infestation on small mammals in the Bangkok Metropolitan area. (A) The selected seven
public parks as the sampling sites across Bangkok. Pixel colours on the map represent associated land cover labels and RGB codes of the GlobCover2009
Land Cover Map at a resolution of 300 × 300 m (http://due.esrin.esa.int/page_globcover.php). (B) The abundance of chigger infestation at each study
site (size of the red circles indicates chigger abundance on individual hosts), and the pie charts showing the proportional occurrence of Leptotrombidium
deliense (in orange) and Ascoschoengastia indica (in blue) in the seven study sites. [Colour figure can be viewed at wileyonlinelibrary.com].

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
4 S. A. Wulandhari et al.

surrounded by dense human settlements and buildings, while whereas species identification followed the identification keys
others are in the middle of agricultural lands and fields. (Fernandes & Kulkarni, 2003; Stekolnikov, 2013). To estimate
The trapping line method was applied following Herbreteau chigger infestation status on small mammals, prevalence, mean
et al. (2011), i.e. 10 lines (each line comprising 10 traps) abundance and mean intensity were calculated using Quantita-
were randomly distributed throughout the area of a particular tive Parasitology 3.0 software (Rozsa et al., 2000).
park and retained for four nights within the same site. The The remaining chigger specimens from the same hosts, same
traps were checked every morning, and the captured animals species and study sites were pooled in 95% ethanol for further
were subsequently assigned the project ID number. Global molecular experiments. Chigger specimens were verified to the
Positioning System coordinates were immediately taken at each correct species using autofluorescence (ZEISS LSM 700 laser
positive trap. Trapping success and adjusted-trapping success scanning confocal microscope with fluorescein isothiocyanate
(Aplin et al., 2003) were calculated from each trapping site as a filter mode) and bright-field microscopy described by Kumlert
proxy for estimating small mammal density in the parks. et al. (2018) before pooling them in the same tube.

Animal trapping and handling Molecular procedures for chigger species confirmation
Animals were transported to the parasitological laboratory Genetic material was extracted from individual chigger speci-
at the Department of Helminthology, Faculty of Tropical
mens using Genomic DNA Mini Kit (Geneaid, Taiwan), follow-
Medicine, Mahidol University, on the same day of capture. They
ing the manufacturer’s protocol. The mitochondrial cytochrome
were euthanized and identified to species level based on either
oxidase subunit I (mtCOI) was selected as a genetic marker
the morphological characteristics following Aplin et al. (2003)
for the chigger taxonomic study. We used the universal primers
or the DNA barcoding approach (see Pagès et al., 2010; Pal-
derived from Folmer et al. (1994), which yield product size
adsing et al., 2020). Information on gender (male or female)
around 650–700 bp. Polymerase chain reaction (PCR) was
and maturity (adult or juvenile) of the small mammals was also
carried out in 50-μL reaction containing 5 μL of DNA tem-
recorded. We also calculated the animal body mass index (BMI)
plate, OnePCR master mix (GeneDireX, Taiwan) and 2 μL of
with the following formula: weight in gram/(length)2 in millime-
each primer (at 0.4 μM final concentration). PCR was run for
tres. Animals were humanely euthanized using 5% isoflurane
40 cycles: initial denaturation at 94 ∘ C for 1 min; 5 cycles of
inhalation. Biological specimens (e.g. blood, lung, liver, kid-
94 ∘ C for 1 min, 45 ∘ C for 90 s, 72 ∘ C for 90 s; 35 cycles of 94 ∘ C
ney, spleen, gastrointestinal tract, ears and faeces) were subse-
for 1 min, 50 ∘ C for 90 s, 72 ∘ C for 1 min and finishing with a
quently collected in ethanol, RNAlater and fresh frozen for fur-
final extension at 72 ∘ C for 5 min. PCR amplicons were visual-
ther parasite/pathogen investigations (other studies will be pub-
ized in 1% agarose gel electrophoresis containing SYBR Safe
lished elsewhere). The Faculty of Tropical Medicine—Animal
dye (Invitrogen, US) at 120 V for 40 min. PCR products were
Care and Use Committee, Mahidol University (FTM-ACUC)
sent for purification and Sanger sequencing at U2Bio (Thailand)
approved the protocol for animal trapping and handling under
Sequencing Service.
document number FTM-ACUC 016/2018E. Specifically, the
The DNA sequences were quality checked, edited and aligned
exemption protocol for the chigger study was also approved
using ClustalW multiple alignment, and a phylogenetic tree
under document number FTM-ACUC 015/2020E.
was created using MEGA-X software (Kumar et al., 2018).
The phylogenetic tree was constructed using the maximum
Investigation of chigger infestation on small mammals likelihood method based on the Hasegawa–Kishino–Yano
model (HKY) with Gamma distributed and invariant sites
After animal euthanization and blood collection, bodies were (G + I). The mtCOI sequences of trombiculid mites: L.
first checked by the naked eye for the presence of chiggers. deliense (GenBank accession no. HQ324977.1), Leptotrom-
Different body parts, e.g. head, pectoral region, axillary line, bidium imphalum (HQ324949.1), Leptotrombidium palpale
abdomen, legs and anogenital area, were carefully checked with (AB300499.1), L. scutellare (AB300498.1), Leptotrombidium
great attention inside the ears. Subsequently, the ears were fletcheri (AB300489.1), Leptotrombidium fuji (AB300496.1)
cut and examined for chigger infestation under a stereomicro- and Ascoschoengastia indica (KY930732.1) was compared with
scope. The chigger samples collected from the same hosts were chigger specimens from the Bangkok public parks. The DNA
counted to estimate the abundance of infestation and preserved sequence of house dust mite Dermatophagoides pteronyssinus
in 70–95% ethanol. To identify and estimate chigger species (KP406749.1) was an outgroup for phylogenetic reconstruction.
richness, around 10–20% of chiggers from each infested animal
were subsampled by differences in observed sizes and micro-
scopic appearance. Individual mites were put on glass slides and Molecular detection of O. tsutsugamushi in chiggers and small
mounted in Berlese’s fluid (clearing agent). The prepared spec- mammal hosts
imens were examined and measured using the PrimoVert light
microscope supplied with Axiocam 105 through ZEN Lite imag- Similar chigger species were selected for specimen pooling,
ing software (Carl ZEISS, Germany). based on different study site (public park) per pool. Ten pools of
Chigger mites were initially identified to subgenus level fol- chiggers (100 individuals from the same species per pool) and
lowing the pictorial key of Nadchatram & Dohany (1974), 80 pools (30 individuals from the same species within the same

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
Chiggers from Bangkok public parks 5

site per pool) were prepared. Increased numbers of chiggers per Multivariate analysis of independent variables explaining
pool were applied following previous Orientia detection studies, the abundance of chigger infestation
from one individual up to >100 chiggers per pool (Takhampunya
et al., 2014; Kuo et al., 2015; Elliott et al. 2019). Here, we Parasite and pathogen infections (i.e. diversity and abundance)
decided to prepare a majority of pool samples with 30 individual in wild rodents are commonly explained by intrinsic factors such
chiggers following Frances et al. (1999) and Shim et al. (2009). as host attributes (e.g. gender, maturity and body mass) or extrin-
In addition, we prepared pools of 100 chiggers to increase sic factors (e.g. habitat, land use type and other environmental
the amount of starting DNA template, which might improve factors). We hypothesized that such intrinsic and extrinsic fac-
the Orientia screening. DNA from each pooled sample was tors might explain chigger infestation status (in terms of chigger
extracted using the Genomic DNA Mini Kit (Geneaid, Taiwan), abundance) on small mammals from the Bangkok public parks.
following the manufacturer’s protocol. DNA was extracted from We investigated each public park’s land use characteristics
spleen tissues (n = 164) following the same protocol as chiggers using Global Land Cover Map (GlobCover2009 V2.3), a raster
for small mammals. in GEOTIFF format containing ID values for different land
Conventional nested PCR was performed to target the two cover categories at a resolution of 300 m × 300 m. Land use
gene markers: 56 and 47 kDa type-specific antigen (TSA) of Ori- boundaries of each study site were created, and proportion
entia bacteria in the chigger and small mammal samples. Ampli- of the land use types, i.e. (1) post-flooding/rain-fed land, (2)
fication of 56 kDa TSA gene (around 700 bp) was performed mosaic cropland/vegetation, (3) shrub/grassland, (4) close to a
with the following primers: CG56F (5′ -TTACAATGGATAA mixed forest and (5) artificial surface/urban within a certain
AACGCTTTGAA-3′ ) and CG56R (5′ -AGAAAAACCTAGA park were calculated using ‘PatchStat’ function in ‘SDMTools’
AGTTATAGCGTACA-3′ ) for the first round PCR; and package implemented in R freeware (R Core Team, 2019).
RTS8F (5′ -AGGATTAGAGTGTGGTCCTT-3′ ) and RTS9R Land use characteristics for each site were verified again
with Google Earth (Google Earth Pro v7.3.2) before using
(5′ -ACAGATGCACTATFAGGCAA-3′ ) for the second round
parameters in further analyses (see more details in Paladsing
PCR (Takhampunya et al., 2014). PCR composition and
et al., 2020). The total area (metre2 ) and boundary perimeter
thermocycler condition were run following the protocol of
(metre) of each park were also measured using Google Earth
Takhampunya et al. (2014). In brief, the first round PCR ampli-
Pro v7.3.2. All these variables were taken into account for
fication was carried out in 25 μL reaction containing 2.5 μL of
generating a composite habitat variable of each park through the
DNA template, OnePCR master mix (GeneDireX, Taiwan) and
principal components analysis (PCA) approach (component 1)
1 μL of each primer (at 0.4 μM final concentration). The PCR
using ‘princomp’ function in ‘stats’ package implemented in R
ran with an initial denaturation at 98 ∘ C for 2 min, followed by
freeware (R Core Team, 2019). The PCA was used to explore
40 cycles of 98 ∘ C for 10 s, 53 ∘ C for 20 s, 72 ∘ C for 1 min and
whether the habitat characteristics of each study site differ from
finishing with a final extension at 72 ∘ C for 5 min. The PCR
each other. We also adopted the score of principal component
product from the first round was diluted to one-tenth of the
1 as a composite habitat variable in subsequent multivariate
volume and used as a template (2.5 μL) in the second round analysis to understand the abundance of chigger infestation.
amplification. PCR ran following the same protocol described Also using Google Earth (Google Earth Pro v7.3.2), we man-
earlier, except that annealing temperature at 55 ∘ C for 15 s ually measured the distances (meter) from individual captured
was applied. PCR products were visualized in 1% agarose gel small mammals to the four habitat categories in/around the pub-
electrophoresis containing SYBR Safe dye (Invitrogen, US) at lic parks. These four habitat categories include (1) water body
120 V for 40 min. (e.g. surface water, water resources, ponds, river, lake and canal);
For 47 kDa TSA gene amplification, we performed nested (2) open field (e.g. grassland, fallow, shrub, cropland and parts of
PCR using the primers: Ot-145F (5′ -ACAGGCCAAGATATTG area that is not covered by shaded trees; (3) shaded tree (e.g. tree
GAAG-3′ ) and Ot-1780R (5′ -AATCGCCTTTAAACTAGATTT plantation and parts of the full-shaded area that contain peren-
ACTTATTA-3′ ) for the first-round amplification; and Ot-263F nial trees and (4) building (e.g. park toilet, office, greenhouse,
(5′ -GTGCTAAGAAARGATGATACTTC-3′ ) and Ot-1133R shelter, apartment, house, mall and other types of human build-
(5′ -ACATTTAACATACCACGACGAAT-3′ ) for the second- ings).
round amplification (Masakhwe et al., 2018). Both the first The generalized linear model (GLM) identified the potential
and second PCR ran in a 25-μL reaction containing 2 μL of effects of host attributes and environmental factors on the abun-
DNA template, OnePCR master mix (GeneDireX, Taiwan) and dance of chigger infestation. Variables included in the initial
0.7 μL of each primer (at 0.3 μM final concentration). Thermo- model were as follows: Chigger abundance = BMI + Sex +
cycler condition was performed similarly on both amplification Maturity + Host density (proxy) + Habitat composite + Dis-
round, starting with an initial denaturation at 95 ∘ C for 2 min, tance to human building + Distance to water body + Distance
followed by 40 cycles of 94 ∘ C for 30 s, 54 ∘ C for 30 s, 68 ∘ C to open field + Distance to shaded trees. Data distribution of
for 2 min and completed with a final extension at 72 ∘ C for the response variable (Chigger abundance) was tested using
7 min. The expected amplicon size was around 870 bp. All ‘descdist’ function in ‘fitdistrplus’ package implemented in R
the assays included negative control samples. The positive freeware (R Core Team, 2019). GLM with Akaike’s informa-
control samples, DNA of O. tsutsugamushi (Karp strain) tion criterion corrected for sample size (AICc) was performed
from Lao Oxford Mahosot Wellcome Trust Research Unit, with negative binomial distribution using ‘glm.nb’ and ‘glmulti’
Lao People’s Democratic Republic, were also applied in all functions in ‘gmulti’ and ‘lme4’ packages. We selected the
assays. best model for further discussion based on AICc, ΔAICc and

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
6 S. A. Wulandhari et al.

Table 1. Prevalence (%), mean abundance (CMA), mean intensity (CMI) and range of chigger infestation on small mammals (n = 164) from public
parks in the Bangkok Metropolitan area

No. of animal Total chiggers % Prevalence Range of


captured number (95%CI) CMA (95%CI) CMI (95%CI) infestation

Host species
Rattus rattus-complex 132 10 991 85.6 (78.4–90.7) 83.27 (64.4–109.7) 97.2 (76.5–125.4) 1–728
Rattus exulans 11 35 9.1 (0.47–40.4) 3.18 (0.00–9.55) 35 (0.0–0.0) 35
Rattus norvegicus 1 20 100 (5–100) 20 (0.0–0.0) 20 (0.0–0.0) 20
Tupaia belangeri 20 132 55 (32–75.5) 6.6 (2.9–14.15) 12 (5.4–22.8) 1–43
Site
WB 55 2455 87.3 (75.6–93.8) 44.6 (30.9–66.7) 51.1 (36.7–75.5) 1–334
SL 38 2789 92.1 (79.2–97.8) 73.4 (48.4–122.7) 79.7 (53.6–132.8) 1–497
LU 4 86 75 (24.8–98.7) 21.5(5.0–33.0) 28.7(20.0–36.3) 20–43
ST 17 213 29.4 (12.3–54-1) 12.5 (1.4–55.5) 42.6 (6.8–113.6) 6–182
VP 23 4522 91.3 (72.2–98.4) 196.6 (125.3–290.8) 215.4 (138.6–307.3) 6–728
TB 12 161 25 (7.2–54.2) 13.4 (0.4–49.7) 53.7 (5.0–140.0) 5–140
TW 15 952 73.3 (46.5–90.3) 63.5 (26.7–150.9) 86.5 (35.2–201.6) 3–406
Sex
Male 74 4959 83.8 (73.7–90.7) 67 (48.1–94.4) 79.9 (58.3–110.6) 1–508
Female 90 6219 71.1 (60.6–79.5) 69.1 (46.7–101) 91.7 (67.7–141.2) 1–728
Maturity
Adult 115 9787 81.7 (73.5–87.9) 85.1 (65.8–114.2) 104.1 (79.9–136.8) 1–728
Juvenile 49 1391 65.3 (51–77.8) 28.3 (15.8–57.2) 43.4 (25.5–84.2) 1–406

Akaike’s weight (W r ) scores (Burnham & Anderson, 2002). of 83.2 and 20, respectively, followed by the tree shrew, T. belan-
The variance inflation factor was computed using a ‘vif’ func- geri (CMA = 6.6) and lastly, the Polynesian rat, R. exulans
tion in the ‘car’ package implemented in R freeware (R Core (CMA = 3.18). Chigger abundance significantly differed among
Team, 2019) to identify the degree of multicollinearity among the public parks (Kruskal-Wallis test = 44.75, P < 0.001). Ani-
the explicative variables. Also, the quality of the selected model mals from VP had the highest mean abundance (CMA = 196.6),
was evaluated with goodness-of-fit and power analysis tests at least two times higher than the other parks, whereas ani-
using ‘chisq_gof’ and ‘modelPower’ functions in ‘sjstats’ and mals residing in TB and ST had a much lower abundance
‘lmSupport’ packages, respectively. (CMA = 13.4 and 12.5, respectively) (Fig. 1B). There was
no significant difference in chigger abundance between male
(CMA = 67) and female hosts (CMA = 69.1). However, adults
Results significantly had higher chigger abundance (CMA = 85.1) than
the young hosts (CMA =28.3) (Mann–Whitney U test = 1788,
Prevalence, abundance and species richness of chigger P < 0.001). Table 1 presents more details of chigger infestation
infestation on small mammals from Bangkok public parks status.
After 10–20% subsampling from each host (intentionally
We trapped 164 small mammals (Muridae: Rattus exulans, selected by differences in observed sizes and microscopic
Rattus norvegicus and Rattus rattus-complex and Scadentia: appearance), we prepared 1187 individual chiggers on per-
Tupaia belangeri) in seven public parks around Bangkok. A total manent slides for taxonomic identification, only two chigger
of 11 178 chiggers were collected from the ears of the animal species: L. deliense (n = 239; 20.1%) and A. indica (n = 948;
hosts. Among the mammals examined, 126 individuals were 79.9%) were identified (Fig. 2). Leptotrombidium deliense was
infested with chiggers, yielding a total prevalence of infestation present in four public parks: WB (n = 14), LR (n = 51), VP
at 76.8%. Host species prevalence of infestation varied from (n = 94) and TW (n = 80), whereas A. indica occurred in all
9.1% to 100%, noting the small number of R. norvegicus (n = 1) study sites. Leptotrombidium deliense was found on R. rat-
and R. exulans (n = 11). The highest prevalence of infestation for tus-complex and T. belangeri, while A. indica occurred widely
the study sites was found in small mammals from SL (92.11%), on the four host species.
followed by VP (91.3%), WB (87.3%), LP (75%), TW (73.3%) We conducted mtCOI gene amplification of selected 19 indi-
and ST (29.4%); whereas mammals from TB showed the lowest vidual chigger samples (9 L. deliense and 10 A. indica after
prevalence at 25%, no significant difference was observed on morphological identification) from various study sites. Phylo-
prevalence of infestation between male (83.8%) and female genetic reconstruction (see Fig. 3) illustrates the taxonomic
host (71.1%), (chi-square = 0.3274, P = 0.5672) and between classification and supports the two chigger species morpho-
adult (81.7%) and juvenile (65.3%), (chi-square = 0.5056, logical identification. Using nucleotide Basic Local Align-
P = 0.4771). ment Search Tool (BLASTn), L. deliense specimens from the
The murid rodents, R. rattus-complex and R. norvegicus had present study showed 98.60–98.90% sequence similarity to
the highest mean abundance of chiggers (CMA) on host species L. deliense: accession no. HQ324977.1, whereas A. indica

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
Chiggers from Bangkok public parks 7

(A) (B)

(C) (D)

(E) (F)

Fig. 2. Fluorescence and bright-field microscopy of chiggers from rodent hosts in Bangkok public parks, Leptotrombidium deliense (A) and
Ascoschoengastia indica (B). Bright-field microscopy showing scutum and whole body aspect of L. deliense (C and E) and A. indica (D and F).
[Colour figure can be viewed at wileyonlinelibrary.com].

showed 87.79–90.64% sequence similarity to A. indica: acces- (MW475716), LDP29 (MW475717), LDP30 (MW475718),
sion no. MG728110.1 and 86.74–88.65% to A. indica: accession LDP33 (MW475719) and LDP34 (MW475720).
no. KY930732.1. The nucleotide sequences of chiggers estab-
lished in this study were deposited in the nucleotide database
(GenBank) under accession numbers: ADP04 (MW478634), Non-detected O. tsutsugamushi in chiggers and small
ADP05 (MW478635), ADP06 (MW478636), ADP07 (MW478 mammals from Bangkok public parks
637), ADP08 (MW478638), ADP09 (MW478639), ADP10
(MW478640), ADP17 (MW478641), ADP18 (MW478642), Applying the O. tsutsugamushi positive control in all PCR
ADP19 (MW478643), LDP23 (MW475712), LDP24 (MW47 experiments (56 and 47 kDa TSA genes), we could not detect
5713), LDP26 (MW475714), LDP27 (MW475715), LDP28 any genetic material of the bacteria in both chiggers (90
© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
8 S. A. Wulandhari et al.

Fig. 3. Phylogenetic reconstruction based on mitochondrial COI sequences of individual chiggers on small mammals from public parks of Bangkok
compared with available DNA sequences in the NCBI database: KY930732.1 (Ascoschoengastia indica), MG728110.1 (A. indica), HQ324977.1 (Lep-
totrombidium deliense), AB300489.1 (Leptotrombidium fletcheri), AB300496.1 (Leptotrombidium fuji), HQ324949.1 (Leptotrombidium imphalum),
AB300499.1 (Leptotrombidium palpale), AB300498.1 (Leptotrombidium scutellare) and KP406749.1 (Dermatophagoides pteronyssinus) as the out-
group. Samples labelled with LDPxx and ADPxx are from the present study with abbreviated public park names. The percentage of trees in which
the associated taxa clustered together is shown next to the branches. Tree topology is drawn to scale, with branch lengths indicating the number of
substitutions per site.

chigger-pooled samples) and animal host tissues (164 spleen composite were the three variables that appeared in all of
samples). those top five models, suggesting that they contribute a strong
influential effect to explain chigger abundance. Using deviance
analysis (type II Wald Chi-square tests), we found significant
associations between chigger abundance and larger body mass
Associated variables explaining the abundance of chigger
of mammals and living in/closed to open field habitat (Fig. 5).
infestation on small mammals from Bangkok public parks

Regarding physical characteristics of the Bangkok public


parks, PCA revealed that the study sites appeared dissimilar Discussion
from each other in habitats and land use types inside and
around the park (Fig. 4). Accordingly, we derived the PCA The study’s findings showed that the prevalence (76.8%) of chig-
score (component 1) of each site as representative of habitat ger infestation on small mammal hosts was comparable to pre-
composite variable in GLM and some other extrinsic factors and vious reports in other areas of Thailand, e.g. 81.8% of small
host attributes to explain the abundance of chigger infestation mammals from a rural village of an ethnic minority in Chiang-
on small mammals. GLM with AICc selection computed 500 mai Province (Rodkvamtook et al., 2013), 88.9% of small mam-
candidate models; among these, the top 5 models are presented mals from military training sites in Chonburi Province (Rod-
in Table 2. The first model (Chigger abundance − BMI + Host kvamtook et al., 2018), and 23.8–95% of small mammals from
density (proxy) + Habitat composite + Distance to open field) the nation-wide surveys (Coleman et al., 2003; Takhampunya
was selected as the best-describing factors influencing chigger et al., 2018; Chaisiri et al., 2019). The prevalence was higher
abundance (see Table 3). BMI, Host density (proxy) and habitat than 33% reported in rats from Tak Province (Takhampunya

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
Chiggers from Bangkok public parks 9

habitats outside urban area (where biodiversity is limited in


urban), i.e. rodents living in forest, mountainous landscape, and
upland and lowland agricultural areas (Peng et al., 2018; Chaisiri
et al., 2019; Matthee et al., 2020). This trend perfectly sup-
ports our findings of only two chigger species: L. deliense and
A. indica recovered from small mammals in Bangkok’s urban
public parks.
The study revealed that host attributes and extrinsic environ-
mental factors influenced the abundance of chigger infestation in
mammalian hosts found in urban public parks. Adult hosts har-
boured a significantly higher number of infested chiggers than
younger ones; this is because adults live longer and roam fur-
ther, searching for food. With more activity and mobility, they
have a better chance for increased exposure to ectoparasites and
chigger foci in the environment than juvenile animals (Webber
et al., 2015; Matthee et al., 2020). Another consideration is that
animals with a larger body mass or BMI provide a larger sur-
face area for different ectoparasite species to feed on (Fernandes
et al., 2015).
Extrinsic environmental factors could also play an important
role in the abundance of chigger infestation on animal hosts.
Unlike endoparasites that generally live inside their host, and
Fig. 4. Principal component analysis (PCA) of habitat characteristics
whose abundance and species composition tend to vary due to
and their distribution among the selected seven public parks across
Bangkok, with the two first axes of the PCA explaining 61.17% of the host intrinsic factors (e.g. immunity and competition for nutri-
variance. [Colour figure can be viewed at wileyonlinelibrary.com]. ents from the host), ectoparasites live outside the fur or skin of
the host’s body. As such, internal host factors and off-host envi-
ronment conditions affect parasite abundance (Morand, 2015).
et al., 2018) and 44.1% in rats from orchards in Nonthaburi There was a significant effect of habitat characteristics in the
Province (Frances et al., 1999), suggesting that urban public public parks (i.e. composite habitat variable, including area and
parks provide a suitable habitat for chigger species to survive perimeter of each park, and the proportion of land use classi-
since they exploit the available hosts to maintain their population fication inside and around the parks) on chigger abundance in
in such a big city. Our findings are corroborated by a previ- the small mammal hosts. For example, we found the highest
ous study in South Africa, which showed a higher prevalence chigger abundance in mammal hosts from VP. The park pos-
of chigger infestation on rodents in urban settings than in agri- sesses the largest area and perimeter, located in Bangkok suburb,
cultural areas and natural habitats (Matthee et al., 2020). In con- and surrounded by open fields, e.g. cropland and paddy fields.
trast, greater chigger diversity was reported in biodiversity-rich In contrast, the parks with lower chigger abundance (e.g. ST,

Table 2. The top five generalized linear model (GLM, with negative binomial distribution) testing for intrinsic and extrinsic variables on chigger
abundance in small mammals from the public parks in Bangkok Metropolitan area. The best model (in bold) is selected for further discussion.

Model Dependent variable–independent variables K Log-like AICc ΔAICc Wr

1 Chigger abundance − BMI*** + Host density 4 −747.45 2009.47 0 0.0731


(proxy) + Habitat composite + Distance to open field***
2 Chigger abundance − BMI*** + Host density (proxy) + Habitat 3 −752.99 2009.57 0.1 0.0694
composite
3 Chigger abundance − BMI*** + Host density (proxy) + Habitat 4 −751.96 2010.17 0.7 0.0515
composite + Distance to human building
4 Chigger abundance − BMI** + Host density (proxy) + Habitat 5 −746.02 2010.22 0.75 0.0502
composite + Distance to human building + Distance to open
field***
5 Chigger abundance–Sex + BMI*** + Host density 5 −747.08 2011.53 2.06 0.0261
(proxy) + Habitat composite + Distance to open field***

Akaike’s Information Criterion corrected for sample size (AICc) was used for model selection after 500 computed models. The initial model for AICc
selection was Chigger abundance − BMI + Sex + Maturity + Host density (proxy) + Habitat composite + Distance to human building + Distance to
water body + Distance to open field + Distance to shaded trees. All the selected models were fitted with host taxonomy as a random effect. K, the
number of estimated variables; Log-like, maximized value of the logarithm of the likelihood function; ΔAICc, the difference between AICc value of
a given model and the model with minimum AICc and W r = Akaike weights. Analysis of deviance (type II test) significant level for each variable is
indicated as following: *P = 0.05–0.01; **P = 0.01–0.001 and ***P < 0.001.

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
10 S. A. Wulandhari et al.

Table 3. The final generalized linear model (GLM) (with negative et al. (2018) that abundance of the low-host specific ectopara-
binomial distribution) showing significant factors for chigger abundance sites (e.g. L. deliense and A. indica chiggers in our case) tends
on small mammals from the public parks across the Bangkok Metropoli- to decrease with an increasing urbanization gradient. Because
tan area (GLM with negative binomial distribution) chiggers are living most of their lifespan off the host in the
environment, their survival and fitness could largely depend
Variance
inflation on the environmental niche. They will not succeed in a higher
Parameters/effects Chi-square P-value factor level of disturbed habitat by urbanization. In addition to these
environmental factors, previous studies revealed that abundance
BMI*** 14.015 <0.001 1.024 and diversity of chiggers are largely influenced by climatic fac-
Host density (proxy) 0.405 0.524 1.011
tors, such as seasonality, temperature and humidity (reviewed
Habitat composite 1.412 0.234 1.066
Distance to open field*** 11.001 <0.001 1.037 by Elliott et al., 2019). Optimal temperature and humidity are
Observation 164 important for chigger activity and development, which also
Log-likelihood −747.45 determines chigger abundance. In tropical and temperate zones,
AICc 2009.47 high chigger diversity tends to be found in warmer or drier con-
Model power (power analysis) 0.876 ditions comparing to wet or cold season (Chaisiri et al., 2019;
The goodness of fit/coefficient of 0.999 Matthee et al., 2020).
determination (R2 )
Screening Orientia spp. bacteria, the causative agent of scrub
Selection of the model used Akaike’s Information Criterion corrected for typhus in vectors and maintenance hosts in the environment, is
sample size (AICc). Analysis of deviance table (type II Wald Chi-square an important strategy to understand disease transmission better
tests), significant level for each variable is indicated as following: and alert the potential risk of exposure (Kuo et al., 2015).
*P = 0.05–0.01; **P = 0.01–0.001 and ***P < 0.001. Although we intensively screened for Orientia spp. in chiggers
and small mammal hosts from Bangkok public parks using
TB and LP) are rather small in size (except the LP) and located standardized molecular methods, the bacteria were not detected
closer to Bangkok’s heart with dense buildings, human accom- from these samples. Chiggers and the mammalian hosts in these
modation and streets. urban parks may not carry Orientia spp., or the population
We also found that animals with high chigger abundance may harbour a very small amount of the pathogen. Previous
were significantly associated with open field habitat (usually publications documented that a higher risk of scrub typhus
peridomestic). This finding could be explained by Maaz exposure is associated with a landscape characterized by forest

Fig. 5. Effect plots of the four explicative variables after best model fitting with generalized linear model to estimate abundance of chigger infestation on
small mammals from public parks across Bangkok. Confidence intervals (95%) are shown in grey. Marks on the X-axis represent individual observations.
[Colour figure can be viewed at wileyonlinelibrary.com].

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
Chiggers from Bangkok public parks 11

and peridomestic habitats, e.g. fields, agricultural lands, fallows periods (or a kind of longitudinal study for a couple of years) to
in the vicinity of a forested area and high elevated land (Chaisiri determine the effect of seasonal variation on chigger abundance,
et al., 2017; Wangrangsimakul et al., 2020). Non-occurrence diversity and Orientia infection. Our results show no evidence of
of Orientia in the present study may be explained by Bangkok O. tsutsugamushi infection in chiggers and rodent tissues after
landscape characteristics, itself as the capital city is located in conventional nested PCR experiments, other molecular detec-
the large central plain (1–2 m elevation) of the country with tion methods may probably offer higher sensitivity, e.g. quan-
no elevated land or natural forest within the radius of 50 km. titative PCR is suggested for bacterial investigation in future
There is a clear difference in landscape prone scrub typhus studies (Jiang et al., 2004). Although the chigger specimens
compared to the two studies of urban scrub typhus incidences are very tiny and contain relatively low biomass (i.e. genetic
in Guangzhou, China (Wei et al., 2014) and Seoul, South Korea material), the potential issue of PCR inhibition in O. tsutsuga-
(Park et al., 2015). In Guangzhou city, there was a report of a mushi detection cannot be overlooked, particularly when work-
human scrub typhus outbreak in Haizhu District between 2006 ing with pooled samples. We also suggest that high-throughput
and 2012. The district is a short distance (<10 km) to the south sequencing technology, e.g. amplicon sequencing or metage-
of Baiyun Mountain (427 m in elevation). In Seoul, urban scrub nomics, could be alternatively applied to improve the bacterium
typhus cases were reported between 2010 and 2013. The capital Orientia’s detection and reveal some other potential pathogens
city of South Korea is surrounded by mountain peaks, e.g. the carried by chiggers. In addition, Orientia serological detection
Bukhan Mountain to the north and the Gwanak Mountain to of the maintenance hosts could also help to determine history
the south, within distances of around 8 km and 12 km from of exposure and may help investigating scrub typhus risk in the
the city centre, respectively. The majority of the human cases study areas.
were related to or near the Gwanak Mountain activities, e.g. Our research outcomes provide a basis for future studies in
trekking, gardening and harvesting agricultural products (Park this neglected field of vector-borne disease in an urban setting.
et al., 2015). With our results, the high prevalence of chiggers, as well as the
It can be hypothesized that lack of biodiversity-rich areas occurrence of important scrub typhus vector species in public
such as natural habitat or preserve forest (particularly on the parks of Bangkok, should provide a message to the Office of
mountainous landscape) in Bangkok may limit the abundance Public Parks, the Bangkok Metropolitan Administration and
and diversity of chiggers and probably affect Orientia bacte- other public health sectors to develop a plan to control vectors
ria’s ecological niche and transmission (particular horizontally). and prevent disease outbreaks.
This hypothesis has been postulated by Chaisiri et al. (2019),
with the evidence that (1) chigger diversity in murid rodents
from Thailand gradually decreased towards urbanization and Acknowledgements
(2) scrub typhus incidence was found to be positively corre-
lated with chigger species richness. This finding is in line with We would like to gratefully thank the Environment Department,
the low disease incidence in Bangkok; only 20 and 31 cases of Bangkok Metropolitan Administration (Director of the Environ-
scrub typhus were documented during 2018 and 2019, respec- ment Department, Mrs. Wullaya Wattanarat and the Director of
tively (Report 506: Bureau of Epidemiology, Ministry of Pub- Bangkok Public Park Office, Mrs. Arom Wongmaha) for grant-
lic Health, Thailand). However, we cannot conclude that pub- ing a permission to work in the area of Bangkok public parks. We
lic parks in Bangkok are safe from scrub typhus because the would like to deliver special appreciations also to the staffs from
important chigger vector species, L. deliense and A. indica have the seven public parks for their kind cooperation and facilitation
been recently found in the public parks. The former species, L. during the fieldworks conducted at their properties. Financial
deliense, is the most dominant scrub typhus vector in the tropi- support for the project was granted by the Office of the Higher
cal region of Asia and Pacific islands, widely spread in India, Education Commission and the Thailand Research Fund (Grant
South China, Taiwan, Thailand, Vietnam, Malaysia, Philip- no. MRG6180023). Funding for M.Sc. in Tropical Medicine of
pines, New Guinea and other countries (Stekolnikov, 2013; Lv Wulandhari from the Malaria Consortium’s Dr. Sylvia Meek
et al., 2018). The latter species is a generalist, broad host range Scholarship for Entomology. Also, we would like to acknowl-
widely found throughout Thailand and some other countries in edge the Faculty of Tropical Medicine - Animal Care and Use
the Asia Pacific region (Fernandes and Kulkarni, 2003; Chaisiri Committee, Mahidol University (FTM-ACUC) for an informa-
et al., 2019). Ascoschoengastia indica was one of over 45 chig- tive guideline and consultation on development of the animal
ger species harbouring O. tsutsugamushi, along with the follow- research protocols. We thank Warren, R. from ENAGO (https://
ing examples: Blankaartia acuscutellaris, Cheladonta ikaoen- university.enago.com/mahidol-university/) for English editing
sis, Eutrombicula wichmanni, Gahrliepia saduski, Helenicula throughout the manuscript (Reference No. MAHDLW-151).
miyagawai, L. akamushi, L. arenicola, L. chiangraiensis, L. All authors declare that we have no conflicts of interest
deliense, L. fletcheri, L. fuji, L. gaohuense, L. imphalum, L. regarding the content in this article.
intermedium, L. kitasatoi, L. orientale, L. pallidum, L. palpale,
L. pavlovskyi, L. scutellare, Neotrombicula japonica, Schoen-
gastiella ligula and Walchia pacifica (Frances et al., 1999; San- Author contributions
tibáñez et al., 2015; Elliott et al., 2019).
The present study was conducted under a cross-sectional Conceptualization: Chaisiri K.; Morand S.; Sumruayphol S.;
approach but did not cover the seasonal effect. Accordingly, fur- Sonthayanon P. Supervision: Sumruayphol S.; Chaisiri K.;
ther studies should be designed to cover both the wet and dry Sonthayanon P. Sample and data acquisition: Wulandhari S.A.;

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
12 S. A. Wulandhari et al.

Paladsing Y.; Saesim W.; Charoennitiwat V.; Kumlert R., Fernandes, S. & Kulkarni, S.M. (2003) Studies on the Trombiculid Mite
Chaisiri K. Data curation: Wulandhari S.A.; Charoennitiwat Fauna of India, p. 539. Zoological Survey of India, Kolkata.
V.; Chaisiri K. Formal analysis and interpretation: Wulandhari Fernandes, F.R., Cruz, L.D., Linhares, A.X. & Von Zuben, C.J. (2015)
S.A.; Sonthayanon P.; Morand S.; Sumruayphol S.; Chaisiri K. Effect of body size on the abundance of ectoparasitic mites on the wild
Manuscript writing—original draft: Wulandhari S.A.; Chaisiri rodent Oligoryzomys nigripes. Acta Parasitol, 60, 515–524.
K. Manuscript writing—review and edit: Wulandhari S.A.; Pal- Folmer, O., Black, M., Hoeh, W., Lutz, R. & Vrijenhoek, R. (1994) DNA
adsing Y.; Saesim W.; Charoennitiwat V.; Sonthayanon P.; Kum- primers for amplification of mitochondrial cytochrome c oxidase
lert R.; Morand S.; Sumruayphol S.; Chaisiri K. subunit I from diverse metazoan invertebrates. Molecular Marine
Biology and Biotechnology, 3, 294–299.
Frances, S.P., Watcharapichat, P., Phulsuksombati, D. & Tanskul, P.
(1999) Occurrence of Orientia tsutsugamushi in chiggers (Acari:
Data availability statement
Trombiculidae) and small animals in an orchard near Bangkok,
Thailand. Journal of Medical Entomology, 36, 449–453.
All DNA sequence data from the present study will be avail-
Herbreteau, V., Jittapalapong, S., Rerkamnuaychoke, W., Chaval, Y.,
able in the nucleotide database on NCBI website following an
Cosson, J.F. & Morand, S. (2011) Protocols for Field and Laboratory
embargo from the date of publication. Also, other data that sup-
Rodent Studies, p. 46. Kasetsart University Press, Bangkok.
port the findings of this study are available from the correspond-
Izzard, L., Fuller, A., Blacksell, S.D. et al. (2010) Isolation of a novel
ing author upon reasonable request.
Orientia species (O. chuto sp. nov.) from a patient infected in Dubai.
Journal of Clinical Microbiology, 48, 4404–4409.
References Jiang, J., Chan, T.-C., Temenak, J.J., Dasch, G.A., Ching, W.-M. &
Richards, A.L. (2004) Development of a quantitative real-time poly-
Abarca, K., Martínez-Valdebenito, C., Angulo, J. et al. (2020) Molecular merase chain reaction assay specific for Orientia tsutsugamushi. The
description of a novel Orientia species causing scrub typhus in Chile. American Journal of Tropical Medicine and Hygiene, 70, 351–356.
Emerging Infectious Diseases, 26, 2148–2156. Kelly, D.J., Fuerst, P.A., Ching, W.M. & Richards, A.L. (2009) Scrub
Aplin, K., Brown, P., Jacob, J., Krebs, C. & Singleton, G. (2003) typhus: the geographic distribution of phenotypic and genotypic
Field Methods for Rodent Studies in Asia and the Indo-Pacific, p. variants of Orientia tsutsugamushi. Clinical Infectious Diseases, 15,
220. ACIAR Monograph No. 100, Australian Centre for International 203–230.
Agricultural Research, Canberra. Kim, Y.S., Lee, H.J., Chang, M., Son, S.K., Rhee, Y.E. & Shim, S.K.
Audy, J.R. (1961) The ecology of scrub typhus. Studies in Disease (2006) Scrub typhus during pregnancy and its treatment: a case series
Ecology (ed. by J.M. May), pp. 389–432. Hafner Publishing Company and review of the literature. American Journal of Tropical Medicine
Inc., New York. and Hygiene, 75, 955–959.
Burnham, K.P. & Anderson, D.R. (2002) Model Selection and Multi-
Kumar, S., Stecher, G., Li, M., Knyaz, C. & Tamura, K. (2018)
model Inference: a Practical Information-Theoretic Approach, 2nd
MEGA X: molecular evolutionary genetics analysis across computing
edn, p. 488. Springer International Publishing, New York.
platforms. Molecular Biology and Evolution, 35, 1547–1549.
Chaisiri, K., Stekolnikov, A.A., Makepeace, B.L. & Morand, S. (2016)
Kumlert, R., Chaisiri, K., Anantatat, T. et al. (2018) Autofluorescence
A revised checklist of chigger mites (Acari: Trombiculidae) from
microscopy for paired-matched morphological and molecular identifi-
Thailand, with the description of three new species. Journal of
cation of individual chigger mites (Acari: Trombiculidae), the vectors
Medical Entomology, 53, 321–342.
of scrub typhus. PLoS One, 13, e0193163. https://doi.org/10.1371/
Chaisiri, K., Cosson, J.F. & Morand, S. (2017) Infection of rodents by
journal.pone.0193163.
Orientia tsutsugamushi, the agent of scrub typhus, in relation to land
Kuo, C.C., Lee, P.L., Chen, C.H. & Wang, H.C. (2015) Surveillance of
use in Thailand. Tropical Medicine and Infectious Diseases, 2, 53.
potential hosts and vectors of scrub typhus in Taiwan. Parasites &
Chaisiri, K., Gill, C.A., Stekolnikov, A.A. et al. (2019) Ecological
Vectors, 8, 611.
and microbiological diversity of chigger mites, including vectors
of scrub typhus on small mammals across stratified habitats in Lakshana, P. (1973) A checklist of the trombiculid mites of Thailand
Thailand. Animal Microbiome, 1, 18. https://doi.org/10.1186/s42523- (Prostigmata: Trombiculidae). United States Army Medical Compo-
019-0019-x. nent, South East Asia Treaty Organization, Bangkok 44p.
Chakraborty, S. & Sarma, N. (2017) Scrub typhus: an emerging threat. Lv, Y., Guo, X.G. & Jin, D.C. (2018) Research Progress on Leptotrom-
Indian Journal of Dermatology, 62, 478–485. bidium deliense. Korean Journal of Parasitology, 56, 313–324.
Coleman, R.E., Monkanna, T., Linthicum, K.J. et al. (2003) Occurrence Maaz, D., Krücken, J., Blümke, J. et al. (2018) Factors associated with
of Orientia tsutsugamushi in small mammals from Thailand. Ameri- diversity, quantity and zoonotic potential of ectoparasites on urban
can Journal of Tropical Medicine and Hygiene, 69, 519–524. mice and voles. PLoS One, 13, e0199385.
Core Team, R. (2019) R: a Language and Environment for Statistical Masakhwe, C., Linsuwanon, P., Kimita, G. et al. (2018) Identification
Computing. R Foundation for Statistical Computing, Vienna. and characterization of Orientia chuto in trombiculid chigger mites
Dohany, A.L., Phang, O.W. & Rapmund, G. (1977) Chigger (Acarina: collected from wild rodents in Kenya. Journal of Clinical Microbiol-
Trombiculidae) surveys of the west coast beaches of Sabah and ogy, 56, e01124–e01118.
Sarawak. The Southeast Asian Journal of Tropical Medicine and Matthee, S., Stekolnikov, A.A., Van der Mescht, L., Froeschke, G. &
Public Health, 8, 200–206. Morand, S. (2020) The diversity and distribution of chigger mites
Elliott, I., Pearson, I., Dahal, P., Thomas, N.V., Roberts, T. & Newton, associated with rodents in the south African savanna. Parasitology,
P.N. (2019) Scrub typhus ecology: a systematic review of Orientia 147, 1038–1047.
in vectors and hosts. Parasites & Vectors, 12, 513. https://doi.org/10 Moniuszko, H. & Makol, J. (2016) Host-parasite association in trombi-
.1186/s13071-019-3751-x. culid mites (Actinotrichida: Trombiculidae) of temperate zone—the

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531
Chiggers from Bangkok public parks 13

case of Hirsutiella zachvatkini (Schluger, 1948); are we dealing with Shim, S.K., Choi, E.N., Yu, K.O. et al. (2009) Characterisation of
prolonged contact with the host? Parasites & Vectors, 9, 61. Orientia tsutsugamushi genotypes from wild rodents and chigger
Morand, S. (2015) (macro-) evolutionary ecology of parasite diversity: mites in Korea. Clinical Microbiology and Infection, 15, 311–312.
from determinants of parasite species richness to host diversification. Stekolnikov, A.A. (2013) Leptotrombidium (Acari: Trombiculidae) of
International Journal for Parasitology: Parasites and Wildlife, 4, the world. Zootaxa, 3728, 1–173.
80–87. Strickman, D., Tanskul, P., Eamsila, C. & Kelly, D.J. (1994) Prevalence
Nadchatram, M. & Dohany, A.L. (1974) A Pictorial Key to the of antibodies to rickettsiae in the human population of suburban
Subfamilies, Genera and Subgenera of Southeast Asian Chiggers Bangkok. American Journal of Tropical Medicine and Hygiene, 51,
(Acari, Prostigmata, Trombiculidae), p. 67. Institute Penyelidikan 149–153.
Perubatan, Kuala Lumpur. Takhampunya, R., Tippayachai, B., Promsathaporn, S. et al. (2014)
Pagès, M., Chaval, Y., Herbreteau, V. et al. (2010) Revisiting the Characterization based on the 56-Kda type-specific antigen gene of
taxonomy of the Rattini tribe: a phylogeny-based delimitation of Orientia tsutsugamushi genotypes isolated from Leptotrombidium
species boundaries. BMC Evolutionary Biology, 10, 184. mites and the rodent host post-infection. American Journal of Tropical
Paladsing, Y., Boonsri, K., Saesim, W. et al. (2020) Helminth fauna Medicine and Hygiene, 90, 139–146.
of small mammals from public parks and urban areas in Bangkok Takhampunya, R., Korkusol, A., Promsathaporn, S. et al. (2018) Hetero-
metropolitan with emphasis on community ecology of infection in geneity of Orientia tsutsugamushi genotypes in field-collected trom-
synanthropic rodents. Parasitology Research, 119, 3675–3690. biculid mites from wild-caught small mammals in Thailand. PLoS
Park, S.W., Ha, N.Y., Ryu, B. et al. (2015) Urbanization of scrub Neglected Tropical Diseases, 12, e0006632.
typhus disease in South Korea. PLoS Neglected Tropical Diseases, Wangrangsimakul, T., Elliott, I., Nedsuwan, S. et al. (2020) The esti-
9, e0003814. https://doi.org/10.1371/journal.pntd.0003814. mated burden of scrub typhus in Thailand from national surveillance
Peng, P.Y., Guo, X.G., Jin, D.C. et al. (2018) Landscapes with differ- data (2003-2018). PLoS Neglected Tropical Diseases, 14, e0008233.
ent biodiversity influence distribution of small mammals and their https://doi.org/10.1371/journal.pntd.0008233.
ectoparasitic chigger mites: a comparative study from Southwest Webber, Q.M.R., McGuire, L.P., Smith, S.B. & Willis, C.K.R. (2015)
China. PLoS One, 13, e0189987. https://doi.org/10.1371/journal.pone Host behaviour, age and sex correlate with ectoparasite prevalence
.0189987. and intensity in a colonial mammal, the little brown bat. Behaviour,
Rodkvamtook, W., Gaywee, J., Kanjanavanit, S. et al. (2013) Scrub 152, 83–105.
typhus outbreak, northern Thailand, 2006–2007. Emerging Infectious Wei, Y., Luo, L., Jing, Q. et al. (2014) A city park as a potential
Diseases, 19, 774–777. epidemic site of scrub typhus: a case–control study of an outbreak
Rodkvamtook, W., Kuttasingkee, N., Linsuwanon, P. et al. (2018) in Guangzhou, China. Parasites & Vectors, 7, 513.
Scrub typhus outbreak in Chonburi Province, Central Thailand, 2013. Weitzel, T., Aylwin, M., Martínez-Valdebenito, C. et al. (2018) Imported
Emerging Infectious Diseases, 24, 361–365. scrub typhus: first case in South America and review of the literature.
Roh, J.Y., Song, B.G., Park, W.I. et al. (2014) Coincidence between Tropical diseases, Travel Medicine and Vaccines, 4, 10. https://doi
geographical distribution of Leptotrombidium scutellare and scrub .org/10.1186/s40794-018-0070-8.
typhus incidence in South Korea. PLoS One, 9, e113193. https://doi Yoshimoto, T. & Yoshimoto, T. (2019) Scrub typhus in Japan. American
.org/10.1371/journal.pone.0113193. Journal of Clinical Microbiology and Antimicrobials, 2, 1042.
Rozsa, L., Reiczigel, J. & Majoros, G. (2000) Quantifying parasites in Zhang, W.Y., Wang, L.Y., Ding, F. et al. (2013) Scrub typhus in mainland
samples of hosts. Journal of Parasitology, 86, 228–232. China, 2006-2012: the need for targeted public health interventions.
Santibáñez, P., Palomar, A.M., Portillo, A., Santibáñez, S. & Oteo, J.A. PLoS Neglected Tropical Diseases, 7, e2493. https://doi.org/10.1371/
(2015) The role of chiggers as human pathogens. In: Samie, A. (Eds), journal.pntd.0002493.
An Overview of Tropical Diseases. InTechOpen. 173–202. https://doi
.org/10.5772/61978. Accepted 10 May 2021

© 2021 The Royal Entomological Society, Medical and Veterinary Entomology, doi: 10.1111/mve.12531

You might also like