<<

MOLECULAR CHARACTERIZATION OF ADULT IN THE NORTHERN HOUSE , PIPIENS

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree Doctor of Philosophy in the Graduate

School of The Ohio State University

By

Rebecca M. Robich, M.S.

*****

The Ohio State University 2005

Dissertation Committee:

Professor David L. Denlinger, Advisor Approved by

Professor Donald H. Dean ______Professor Glen R. Needham Advisor Graduate Program in Entomology Professor Brian H. Smith

ABSTRACT

In the northern United States, Culex pipiens (L.), a major avian of several

-borne viruses, spends a good portion of the year in a state of developmental

arrest (diapause). Although the physiological and hormonal aspects of Cx. pipiens

diapause have been well-documented, there is little known on the molecular aspects of

this important stage. Using suppressive subtractive hybridization (SSH), 40 genes

differentially expressed in diapause were identified and their expression profiles were

probed by northern blot hybridization. These genes have been classified into 8 distinct

groupings: regulatory function, food utilization, stress response, metabolic function,

cytoskeletal, ribosomal, transposable elements, and genes with unknown functions.

Among 32 genes confirmed by northern blot hybridization, 6 are upregulated specifically in early diapause, 17 are upregulated in late diapause, and 2 are upregulated

throughout diapause. In addition, 2 genes are diapause downregulated and 5 remained

unchanged during diapause. Two regulatory genes upregulated in late diapause,

ribosomal protein (rp) S3A and rpS6, are of particular interest for their potential

involvement in developmental arrest. In other mosquito species, these genes are

upregulated prior to oogenesis, and their suppression leads to a disruption in ovarian

development. Since arrested ovarian development is a key characteristic of Cx. pipiens

diapause, the lack of expression in early diapause may be key to this developmental

ii arrest. Of equal interest is the upregulation of two stress-response genes in late diapause.

The molecular chaperone, heat shock protein 23 (hsp23) is slightly upregulated in late

diapause and may be involved in protecting females from environmental stresses. In

addition, a gene encoding an enzyme involved in oxidizing certain insecticides in

insecticide-resistant strains of Cx. quinquefasciatus, aldehyde oxidase, is strongly upregulated in Cx. pipiens late diapause.

SSH and northern blot analysis also demonstrated the strong upregulation of a

muscle-specific actin gene in Cx. pipiens early diapause. Although the upregulation of

this cytoskeletal gene in early diapause may seem counterintuitive, females actively

during diapause preparation in search of sugar meals and a protective site for

overwintering. In addition, two ribosomal genes (large ribosomal subunit L18 and large

ribosomal subunit) are upregulated in early diapause suggesting that their function is

restricted to diapause preparation. Curiously, two genes encoding transposable elements,

transposon T1-2 and Mimo-Cp2, are upregulated during late diapause in Cx. pipiens, but

the function of these genes during diapause is unknown. Nine genes were also isolated

by SSH with unknown identities, and all but one are upregulated in late diapause, as

identified by northern blot hybridization

In addition to the aforementioned genes, two genes encoding the blood digestive

enzymes, trypsin and chymotrypsin-like, are downregulated in early diapause, and

concurrently a gene associated with the accumulation of lipids, fatty acid synthase, is

highly upregulated. As females enter diapause, fatty-acid synthase is only sporadically

expressed, while the expression of trypsin and chymotrypsin-like remain undetectable

iii until late diapause, when females prepare to take a blood meal upon diapause break. This

is the first molecular evidence demonstrating that diapause in Cx. pipiens evokes a

molecular switch from blood-feeding in nondiapausing individuals to sugar feeding in

diapause-destined females.

The increase in sugar feeding in diapause-destined Cx. pipiens and subsequent

accumulation of lipid reserves likely requires a high amount of energy. This may be

reflected in the upregulation of genes encoding two respiratory enzymes, cytochrome c

oxidase subunit I (COI) and cytochrome c oxidase subunit III (COIII), during diapause

preparation. Although transcript levels decline in mid-diapause when females enter an

inactive state, levels again rise in late diapause, just prior to diapause break. There are no differences in mtDNA levels between nondiapausing and diapausing Cx. pipiens, thus suggesting that mitochondrial numbers are not reduced during diapause. Regulation of

COI and COIII is thus likely to be under transcriptional control.

In addition to genes obtained through SSH, heat shock protein 70 was isolated

from Cx. pipiens by 5- and 3- rapid amplification of cDNA ends. This gene is upregulated upon heat shock and recovery from cold shock in diapausing and nondiapausing females, but is not upregulated as a component of the diapause program.

However, diapausing females reared at 18°C survive cold exposure (-5°C) nearly twice as long as its nondiapausing counterparts reared at 18°C and 10 times as long as nondiapausing mosquitoes reared at 25°C. Diapausing females are also more desiccation resistant (1.6 to 2 fold) than nondiapausing females, regardless of rearing temperature.

iv These results may be useful when comparing the molecular aspects of diapause across different taxa and developmental stages. In addition, this work may be helpful in understanding the transseasonal maintenance of viruses that utilize overwintering .

v

For Michael

vi

ACKNOWLEDGMENTS

I am grateful to my advisor, David L. Denlinger, Ph.D., for his intellectual guidance and continued support for this project, and for his assistance in editing each chapter.

I would like to thank my committee members, Drs. Donald H. Dean, Glen R. Needham, and Brian H. Smith, for their ideas and comments in bettering this thesis.

I am also grateful for the assistance of Joseph P. Rinehart, Ph.D., and Linda Kitchen in helping to make this dissertation come together.

I would like to thank those involved in the surveillance of in overwintering mosquitoes, including Richard Gary and Robert A. Restifo of the Ohio Department of Health's Vector-Borne Disease Program, and Joseph Lynch of the Cuyahoga County Board of Health.

I appreciate the efforts of Woodbridge Foster, Ph.D., in helping to establish the Culex pipiens (Buckeye strain) colony, who collected the original larvae from his backyard in September 2000.

vii

VITA

September 20, 1974 .…...... Born – Painesville, Ohio

1992 ..…………………..… Bachelor of Science, Environmental Biology Ohio University

1997-1998 .………………. Graduate Teaching Associate Ohio University

1998 .……………………... Master of Science, Environmental Studies Ohio University

1999-2000 .………………. Pre-Doctoral Fellow Virology Division U.S. Army Medical Research Institute of Infectious Diseases

2000-2004 .………………. Graduate Teaching and Research Associate The Ohio State University

2001 .……………………... Summer Intern Ohio Department of Health

PUBLICATIONS

Moll, R.M., Romoser, W.S., Modrzakowski, M.C, Moncayo, A.C., and Lerdthusnee, K. 2001. Meconial peritrophic membranes and the fate of midgut bacteria during mosquito (Diptera: Culicidae) metamorphosis. Journal of Medical Entomology 38:29-32.

Romoser, W.S., Moll, R.M., Moncayo, A.C., and Lerdthusnee, K. 2000. The occurrence and fate of the meconium and meconial peritrophic membranes in pupal and adult mosquitoes (Diptera: Culicidae). Journal of Medical Entomology 37:893-6.

FIELDS OF STUDY

Major Field: Entomology viii

TABLE OF CONTENTS

Abstract ………………………………………………………………………………….. ii

Dedication ………………………………………………………………………………. vi

Acknowledgements …………………………………………………………………...... vii

Vita ……………………………………………………………………………………. viii

List of Tables …………………………………………………………………………… xi

List of Figures ………………………………………………………………………….. xii

Chapters:

1. Introduction ……………………………………………………………………… 1 References ……………………………………………………………… 12

2. Diapause-specific gene expression in adults of the northern house mosquito, Culex pipiens L., identified by suppressive subtractive hybridization. Abstract ………………………………………………………………… 16 Introduction …………………………………………………………….. 17 Materials and Methods …………………………………………………. 19 Results ………………………………………………………………….. 23 Discussion ……………………………………………………………… 27 References ……………………………………………………………… 38

3. Diapause in the mosquito Culex pipiens evokes a metabolic switch that shuts down genes encoding blood digestive enzymes and upregulates a gene associated with sugar utilization and lipid storage.

Abstract ………………………………………………………………… 49 Introduction …………………………………………………………….. 50 Materials and Methods …………………………………………….…… 52 Results ……………………………………………………………….…. 57 Discussion ……………………………………………………………… 63 References ……………………………………………………………… 69

ix

4. Downregulation of mitochondrial mRNA expression, but not mitochondrial number, during adult diapause in the northern house mosquito, Culex pipiens.

Abstract ………………………………………………………………… 82 Introduction ………………………………………………………….…. 83 Materials and Methods ……………………………………………….… 85 Results ………………………………………………………………….. 90 Discussion ……………………………………………………………… 94 References ……………………………………………………………… 99

5. Enhanced cold and desiccation tolerance in diapausing adults of Culex pipiens, and a role for hsp70 in response to cold shock but not as a component of the diapause program.

Abstract ……………………………………………………………..… 109 Introduction …………………………………………………………… 110 Materials and Methods ………………………………………………... 112 Results ………………………………………………………………… 120 Discussion …………………………………………………………….. 126 References ………………………………………………………….…. 132

Conclusions …………………………………………………………………………… 142

Appendix ……………………………………………………………………………… 149

Materials and Methods ……………………………………………...… 150 Results and Discussion ……………………………………………..… 155 References …………………………………………………………..… 159

Bibliography ………………………………………………………………………….. 149

x

LIST OF TABLES

Table Page

2.1 Diapause upregulated genes from Cx. pipiens, isolated by suppressive subtractive hybridization ………………………………………..… 44

2.2 Diapause downregulated genes from Cx. pipiens, isolated by suppressive subtractive hybridization ………………………………………….. 46

A.1 Ratio and percent of overwintering Cx. pipiens moving within Their sites from January to April, 2004 ………………………..… 170

xi

LIST OF FIGURES

Figure Page

2.1 Northern blot hybridization of Cx. pipiens diapause upregulated SSH clones … 47

2.2 Northern blot hybridization of Cx. pipiens diapause downregulated SSH clones …………………………………………………………….……….. 48

3.1 Nucleotide and deduced amino acid sequences of Cx. pipiens trypsin cDNA … 73

3.2 Nucleotide and deduced amino acid sequences of Cx. pipiens chymotrypsin-like serine protease ……………………………………………... 74

3.3 Nucleotide and deduced amino acid sequences of Cx. pipiens fatty acid synthase cDNA …………………………………………………….... 75

3.4 Multiple sequence alignment of the deduced Cx. pipiens trypsin with other trypsins retrieved from GenBank ..…………………………... 76

3.5 Multiple sequence alignment of the deduced Cx. pipiens chymotrypsin-like serine protease with other insect serine proteases retrieved from GenBank .….. 77

3.6 Multiple sequence alignment of the deduced Cx. pipiens fatty acid synthase with other insect fatty acid synthases retrieved from GenBank ...……………... 78

3.7 Northern blot hybridization of diapause-regulated genes involved in blood meal vs. sugar meal digestion in Cx. pipiens ……………………………. 79

3.8 Temporal pattern of expression of the genes encoding the digestive enzymes trypsin, chymotrypsin-like, and fatty acid synthase ……………………………. 80

3.9 Expression of trypsin, chymotrypsin-like, and fatty acid synthase throughout diapause and when diapause is broken …………………………….. 81

4.1 Nucleotide sequence and deduced amino acid sequence of Cx. pipiens cytochrome c oxidase subunit I cDNA …………………………... 102

xii 4.2 Nucleotide sequence and deduced amino acid sequence of Cx. pipiens cytochrome c oxidase subunit III cDNA ……………………….... 104

4.3 Multiple sequence alignment of the deduced Cx. pipiens cytochrome c oxidase subunit I (COI) with other insect COI sequences retrieved from GenBank ….. 105

4.4 Multiple sequence alignment of the deduced Cx. pipiens cytochrome c oxidase subunit III (COIII) with other insect COIII sequences retrieved from GenBank ……………………………………. 106

4.5 Temporal pattern of expression of the genes encoding the mitochondrial respiratory enzymes COI and COIII ………………………………………….. 107

4.6 Expression of COI and COIII throughout diapause and when diapause is broken ……………………………………………………………. 108

5.1 Percent survival (mean + SE) of Cx. pipiens females after exposure to different durations of low temperature (-5°C) ……………...……………… 136

5.2 Percent survival (mean + SE) of Cx. pipiens females at various relative humidities ……………………..…………...……………… 137

5.3 Complete nucleotide sequence and deduced amino acid sequences of Cx. pipiens hsp70 cDNA …………………………………………………... 138

5.4 Multiple sequence alignment of the deduced Cx. pipiens hsp70 with other hsp70 sequences retrieved from GenBank ………………………... 140

5.5 Northern blots showing expression of heat shock protein 70 in Cx. pipiens females at different times (hours) following a 4 h exposure to –5°C ………… 141

A.1 Diagrams and photos of three Ohio culverts …………………………………. 161

A.2 Average temperature recorded in three Ohio culverts from November 19, 2003 through April 30, 2004, at hourly intervals ……………... 164

A.3 Average relative humidity recorded in three Ohio culverts from November 19, 2003 through April 30, 2004, at hourly intervals ……………... 167

xiii

CHAPTER 1

INTRODUCTION

Many important mosquito vectors of transmissible disease that live in temperate

zones enter an overwintering (diapause). Though the period of dormancy

results in a significant decrease or absence of vector-borne diseases during this portion of

the year, the overwintering stage of the mosquito can harbor pathological agents during

this time and thus reinitiate disease transmission the following spring. This appears to be

a likely scenario that may have contributed to the reappearance of West Nile virus in the

New York City area following the initial outbreak in the summer of 1999 (Nasci et al.,

2001). In this case, diapausing adult females of Culex pipiens were found to harbor the

virus during the winter months, and such mosquitoes serve as one possible mechanism

for reintroducing the disease into populations the following season. More recently,

West Nile virus was noted in overwintering populations of Cx. pipiens collected during

February 2003 in spring houses in eastern Pennsylvania (Kristen Bardell, Pa. Dept. of

Environmental Protection, personal communication) and from populations collected in

2004 from culverts in Boston (Andrew Spielman, Harvard University, personal communication).

1 The diapausing stage thus emerges as an extremely important stage for

understanding certain disease transmission cycles. Not only is this stage critical for

understanding the seasonal occurrence of vector-borne diseases, but understanding the physiological and molecular basis for diapause could provide significant clues for understanding how the pathogen is regulated. The focus of this work is on the molecular mechanisms that serve to regulate the adult reproductive diapause in the northern house mosquito Cx. pipiens, the primary sylvatic vector of West Nile virus. Many aspects of diapause in this species have been well documented. There is a good database that describes the physiological features of this diapause, its environmental regulators, and the hormonal control mechanism, but what is currently lacking is an understanding of its molecular underpinning.

Diapause in Cx. pipiens

The northern house mosquito, Cx. pipiens L., is a member of the Cx. pipiens complex, which was once considered to be comprised of a number of different subspecies

(Mattingly et al., 1951). In the United States, the complex includes the southern house mosquito, Cx. quinquefasciatus Say, and two variants of the northern house mosquito: an autogenous form, Cx. pipiens var. molestus, which can produce an initial batch of eggs without a blood meal, and an anautogenous form, Cx. pipiens pipiens, which requires a blood meal for egg production. Both autogenous and anautogenous variants of Cx. pipiens exist north of 39o spanning across the temperate zones of , , and parts of (Barr, 1957), but it is only the anautogenous form (requires a blood

2 meal for egg production) that overwinters in an adult reproductive diapause (Spielman,

1964; 1971; Spielman and Wong, 1973a). Although males are occasionally found in hibernation sites, it is only the females that enter diapause and successfully overwinter.

Mating occurs prior to the entry into diapause, and diapausing females are rarely found without sperm (Spielman, 1964).

Environmental Regulation of Diapause in Cx. pipiens:

Like diapause in most temperate zone species, short daylength and low temperature are the primary environmental factors dictating the entry into diapause

(Eldridge, 1966; Sanburg and Larsen, 1973; Spielman and Wong, 1973b). Cx. pipiens reared under <12 h light per day and at temperatures below 20°C results in >80% diapausing females, while those reared at longer photoperiods and low temperature

(<20°C) leads to fewer than 45% of females in diapause (Spielman and Wong, 1973b.

Rearing females at 25°C or higher, however, produces all nondiapausing females, regardless of photoperiod (Eldridge, 1966; Spielman and Wong, 1973b; Sanburg and

Larsen, 1973). Thus, diapause in Cx. pipiens is determined by the combined effects of temperature and photoperiod.

The photosensitive stage used for programming diapause begins during late larval life and persists into the early days of adult life (Oda, 1968; Sandburg and Larsen, 1973;

Spielman and Wong, 1973b), but the period of maximum sensitivity occurs shortly after the larval-pupal molt (Spielman and Wong, 1973b). In the late summer and early fall, females that receive diapause-inducing conditions during the photosensitive stage

3 undergo a number of behavioral and physiological changes as adults; diapause-destined

females show a boost in sugar feeding, blood feeding ceases due to a lack of host-seeking

response, their fat body hypertrophies, and they seek well-protected sites for hibernation.

The Preparatory Phase of Cx. pipiens Diapause

In association with the entry into diapause, females increase their consumption of carbohydrates found in plant sources such as and rotting fruits, which leads to a hypertrophy of their fat body (Mitchell and Briegel, 1989a; Bowen, 1992). Diapause- destined females feed more readily on sugar and for a longer time than their nondiapausing counterparts during the first few weeks of adult life (Bowen, 1992), which leads to the accumulation of nearly twice as many lipid reserves (Mitchell and Briegel,

1989a). These lipid reserves are gradually depleted during the course of winter; laboratory and field studies indicate fat stores are depleted by 80% or more over a 5 month period (Onyeka and Boreham, 1987; Mitchell and Briegel, 1989a)

Another key characteristic of prehibernating females is an absence in host-seeking behavior, which results in a lack of blood feeding at this time (Mitchell, 1983; Mitchell and Briegel, 1989a; Bowen, 1992). Whether or not diapause-destined females will take a blood meal prior to entering hibernation has been a great debate in the literature. Females placed in close proximity to a host can be enticed to take a blood meal (Eldridge, 1966;

Eldridge and Bailey, 1979; Mitchell, 1983), but this bypasses the host-seeking step necessary under natural conditions. When laboratory-reared females are offered a blood meal in large containers where host-seeking is essential, only a small percentage will

4 actually take a blood meal (Mitchell, 1983). Bowen (1990) demonstrated that Cx. pipiens reared under diapause-inducing conditions lack high-sensitivity olfactory responsiveness to the host-attractant, lactic acid. Furthermore, force-fed diapausing Cx. pipiens eject most of the blood ingested within 24 h of taking a blood meal, and any remaining blood is incompletely digested and is not used to increase lipid reserves or to initiate vitellogenesis (Mitchell and Briegel, 1989b). In addition, winter-collected females show no signs of blood-feeding, and only an occasional blood-fed female can be found resting in sites after April, once diapause has been terminated (Onyeka and Boreham, 1987).

After the acquisition of sufficient lipid reserves for winter survival, Cx. pipiens seek well-protected sites for hibernation: caves, hollows, and various artificial shelters including culverts and unheated basements are commonly selected (Vinogradova, 2000).

Cx. pipiens do best in sites that remain above 0°C with a small fluctuation in temperature around 1-2°C per month, and have a high relative humidity over 90% (Minar and Ryba,

1971). In the Boston area, females first begin appearing in such sites in mid-August and reach peak density by early October (Spielman and Wong, 1973a). Similar patterns have been noted in England (Onyeka and Boreham, 1987). Females are active in their sites throughout the winter; considerable movement of diapausing Cx. pipiens was observed in each month tested from September to April in mosquitoes overwintering in southern

England (Onyeka and Boreham, 1987).

5 Arrest in Ovarian Development

Once in diapause, the female’s follicles are arrested at Christophers’ Stage I or

perhaps slightly earlier (Oda, 1968; Sanburg and Larsen, 1973; Spielman and Wong,

1973b). At this stage, the follicles contain distinct oocytes surrounded by a follicular

epithelium, but no lipid droplets are yet present in the ooplasm (Clements, 1992). The

most advanced follicles in diapausing females measure 40-50 µm in length, a length that

is no more than 1.5 x that of the germarium (Spielman and Wong, 1973b). In an experimental simulation of diapause, follicle size remained at Christophers’ Stage I for

the first 10 weeks and then gradually increased to Stage II by 22 weeks (Readio et al.,

1999).

Only parous females can be collected by host-attractant CO2 traps in the early fall,

while females collected from overwintering sites at this time are all nulliparous (Andrew

Spielman, personal communication). This suggests that females emerging after August that have follicles in arrested development do not seek a blood meal for further development at this time. Females collected in CO2 traps are older mosquitoes and will

likely produce diapause-destined progeny if a blood meal is available.

Hormonal Regulation of Diapause in Cx. pipiens

As is the case of most adult diapauses (Denlinger, 1985; 2002), the diapause of

Cx. pipiens appears to be initiated by a shut-down in the production of juvenile hormone

(JH) by the corpora allata (Spielman, 1974). During the first hours of adult life, there is

no production of juvenile hormone in either nondiapausing or diapausing females, but by

6 24 h nondiapausing females produce 10 times as much juvenile hormone as their

diapausing counterparts (Readio et al., 1999). Four weeks into diapause, juvenile

hormone titers increase slightly and plateau at this level until week 12; after this time

there is a gradual increase in JH production until levels comparable to 3-day old

nondiapausing females are reached by week 20 (Readio et al., 1999). This slow rise in

JH corresponds to the gradual increase in follicle size also seen in diapausing Cx. pipiens

throughout the winter period (Readio et al., 1999).

Moving diapausing females to nondiapause conditions (long daylength and high temperature) results in an increase in JH titers over a 5 day period (Readio et al., 1999)

In addition, allatectomized females of Cx. pipiens that were not programmed for diapause

enter a diapause-like state and cease host-seeking behavior (Meola and Petralia, 1980),

and diapausing females can be prompted to reinitiate ovarian development by topical

application of JH-III or by application of JH analogs (Spielman, 1974; Readio and Meola,

1985; Readio et al., 1988). All of this evidence points to this diapause being a classic

case of regulation by JH.

Termination of Diapause

Few Cx. pipiens can be found resting in overwintering sites after April, and

females have been observed to leave their sites as early as March (Minar and Ryba, 1971;

Onyeka and Boreham, 1987). The termination of diapause is dependent primarily on

temperature (Oktyabrskaya et al., 1965); females departed from a warm site in Moscow

nearly a month before females departed from a nearby, cooler site. In field-collected Cx.

7 pipiens from early winter, nearly 2 weeks of exposure to long-day length was required before diapause was broken, whereas a few days to less than a week of long-day conditions was required to terminate diapause in females collected from overwintering sites in March and April (Onyeka and Boreham, 1987). This suggests that females gradually come out of diapause through the course of winter and are prepared to resume nondiapausing activities once suitable environmental conditions are attained.

Transeasonal Maintenance of West Nile Virus

The failure to find evidence for blood feeding in overwintering Cx. pipiens suggests that a blood meal is an unlikely source of West Nile virus, and thus vertical transmission is the most likely scenario. Cx. pipiens is the primary sylvatic vector of

West Nile virus in the northern United States. For example, from 6,900 pools of mosquitoes tested for West Nile virus in the summer of 2002 in Ohio, 1,600 were positive by PCR, and 75% of these were identified as Cx. spp. ( pipiens or restuans)

(Richard Gary, Ohio Department of Health, personal communication). Laboratory experiments suggest that Cx. pipiens is a poor vector of West Nile virus since only 2% of orally infected mosquitoes obtained disseminated infections; however, 88% of those that became infected were able to subsequently transmit the virus to a naïve chicken (Turell et al., 2001).

In the laboratory, Cx. pipiens intrathoracically inoculated with West Nile virus showed low vertical transmission rates of <0.3%, which corresponds to a minimal filial infection rate ~ 1.8 (Dohm et al., 2002). Intrathoracic inoculation is not a natural means

8 of acquiring virus, and these numbers likely under-represent what happens in the field.

Although photoperiod has not been evaluated as a factor contributing to the ability of Cx.

pipiens to acquire and transmit West Nile virus, environmental temperature has been

tested and appears to play a role in the ability to recover live virus from orally-infected

females (Dohm and Turell, 2001). Live West Nile virus was recovered from 100% of

infected Cx. pipiens held at a high temperature (26°C), whereas no live virus was

recovered from mosquitoes held at 10°C (Dohm and Turell, 2001).

Molecular Regulation of Mosquito Diapause

The molecular characteristics of diapause in Cx. pipiens have not been explored.

Indeed, very little is known of these events in any mosquito vector, although such work has been initiated on the of Oclerotatus triseriatus, an important vector of La Crosse (Blitvich et al., 2002). Thus far, these workers have identified mRNA sequences present during embryonic diapause of Oc. triseriatus, using primers designed to amplify sequences that the La Crosse virus cap scavenges. Among cDNA fragments that have been identified are a mitochondrial cytochrome c oxidase subunit, 18S and 28S ribosomal RNAs, protein disulfide-isomerase, guanine nucleotide- binding protein, human N33 protein, and several novel transcripts. It is still too early to know if any of these genes are indeed involved in regulation of this diapause or in the regulation of viral transcription. Embryonic and reproductive diapauses are sufficiently

9 different in terms of hormonal regulation (Denlinger, 1985; 2002), thus it is not at all

certain whether the same or similar molecular mechanisms operate in diapauses of

different stages.

From our laboratory studies with the pupal diapause of flesh it is evident that

the expression of many genes is shut down during diapause, while a small cluster of

genes (approximately 4 %) are diapause upregulated (Denlinger, 2002). Several classes

of diapause upregulated genes have been noted. Some, such as those that encode heat shock protein (hsp) 23 and hsp70, are upregulated throughout diapause (Yocum et al.,

1998; Rinehart et al., 2000). Another group, represented by cystatin, is expressed only in early diapause and may be involved in initiating the arrest in development (Goto and

Denlinger, 2002). Yet another category, represented by ultraspiracle, is upregulated only in late diapause and is likely to be involved in the processes leading to diapause termination (Rinehart et al., 2001). Of potentially equal importance are key regulatory genes that are shut down during diapause. In flesh flies, one such gene encodes proliferating cell nuclear antigen, a key regulator of the cell cycle (Tammariello and

Denlinger, 1998). A shut down in expression of this gene is possibly responsible for the cell cycle arrest that characterizes pupal diapause in flesh flies. Experiments currently

underway with this system focus on a search for additional diapause-regulated genes and

the functions of those genes in relation to diapause.

Molecular studies on adult diapauses, such as that noted in Cx. pipiens, have thus far received little attention. The adult diapause of the Colorado potato beetle is arguably the best-understood adult diapause, but little work has focused on molecular aspects of

10 this diapause. The first paper of this nature (Yocum, 2001) documents the upregulation of hsp70 during adult diapause in the Colorado potato beetle, an intriguing result suggesting that there may indeed be some parallels between the molecular events regulating diapause in different species and stages.

Research Foci

This study investigates the molecular mechanisms regulating adult reproductive

diapause of Cx. pipiens, and also examines responses to environmental stress such as

temperature and relative humidity in the field and laboratory settings. Suppressive

subtractive hybridization was used to identify genes differentially expressed in early and

mid diapause, and expression of these genes was confirmed by northern blot

hybridization. From this, 9 categories of genes have been identified as either upregulated

or downregulated during diapause, and these include genes involved in regulatory

functions, genes associated with food utilization, stress-response genes, genes of

metabolic function, cytoskeletal genes, those of mitochondrial origin, genes encoding

ribosomal proteins, transposable elements, and genes with unknown function.

This research will provide a basis for understanding the molecular regulation of

diapause in this important species and may prove valuable for probing potential

commonality in the diapauses of different taxa and developmental stages. In addition,

these results may also be useful for understanding how the replication of West Nile virus

within Cx. pipiens can be seasonally shut down in the autumn and again reinitiated the

following spring.

11

REFERENCES

Barr, A.R. 1957. The distribution of Culex p. pipiens and Culex p. quinquefasciatus in North America. American Journal of Tropical Medicine and Hygiene 6:153-156.

Blitvich, B.J., Rayms-Keller, A., Blair, C.D., and Beaty, B.J. 2001. Identification and sequence determination of mRNAs detected in dormant (diapausing) Aedes triseriatus mosquito embryos. DNA Sequence 12:197-202.

Bowen, M.F. 1990. Post-diapause sensory responsiveness in Culex pipiens. Journal of Insect Physiology 36:923-929.

Bowen, M.F. 1992. Patterns of sugar feeding in diapausing and nondiapausing Culex pipiens (Dipetera: Culicidae) females. Journal of Medical Entomology 29:843-849.

Bowen, M.F., Davis, E.E., and Haggar, D.A. 1988. A behavioral and sensory analysis of host-seeking behavior in the diapausing mosquito Culex pipiens. Journal of Insect Physiology 34:805-813.

Clements, A.N. 1992. The biology of mosquitoes, volume I: development, nutrition, and . CABI Publishing, Cambridge, pp. 340-359.

Denlinger, D.L. 1985. Hormonal control of diapause. In: Kerkut, G.A., Gilbert, L.I. (Eds.), Comprehensive insect physiology, biochemistry and pharmacology, Vol. 8. Pergamon Press, Oxford, pp. 353-412.

Denlinger, D.L. 2002. Regulation of diapause. Annual Review of Entomology 47:93- 122.

Dohm, D.J., Sardelis, M.R., and Turell, M.J. 2002. Experimental transmission of West Nile virus by Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 39:640-644.

Dohm, D.J., and Turell, M.J. 2001. Effect of incubation at overwintering temperatures on the replication of West Nile virus in New York Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 38:462-464.

Eldridge, B.F. 1966. Environmental control of ovarian development in mosquitoes of the Culex pipiens complex. Science 151:826-828.

Eldridge, B.F., and Bailey, C.L. 1979. Experimental hibernation studies on Culex pipiens (Diptera: Culicidae): reactivation of ovarian development and blood-feeding in prehibernating females. Journal of Medical Entomology 15:462-467.

12

Goto, S.G., and Denlinger, D.L. 2002. Genes encoding two cystatins in the flesh fly Sarcophaga crassipalpis and their distinct expression patterns in relation to pupal diapause. Gene 292:121-127.

Mattingly, P.F., Rozeboom, L.E., Knight, K.E., Laven, H., Drummond, F.H., Christophers, S.R., and Shute, P.G. 1951. The Culex pipiens complex. Transactions of the Royal Entomological Society of London 7:331-343.

Meola, R.W., and Petralia, R.S. 1980. Juvenile hormone induction of biting behaviour in Culex mosquitoes. Science 209:1548-1550.

Minar, J. and J. Ryba. 1971. Experimental studies on overwintering conditions of mosquitoes. Folia Parasitologica (PRAHA) 18:255-259.

Mitchell, C.J. 1983. Differentiation of host-seeking behavior from blood-feeding behavior in overwintering Culex pipiens (Diptera: Culicidae) and observations on gonotrophic dissociation. Journal of Medical Entomology 20:157-163.

Mitchell, C.J., and Briegel, H. 1989a. Inability of diapausing Culex pipiens (Diptera: Culicidae) to use blood for producing lipid reserves for overwinter survival. Journal of Medical Entomology 26:318-326.

Mitchell, C.J., and Briegel, H. 1989b. Fate of the blood meal in force-fed, diapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 26:332-341.

Nasci, R.S., Savage, H.M., White, D.J., Miller, J.R., Cropp, B.C., Godsey, M.S., Kerst, A.J., Bennett, P., Gottfried, K., and Lanciotti, R.S. 2001. West Nile virus in overwintering Culex mosquitoes, New York City, 2000. Emerging Infectious Diseases 7:742-744.

Oda, T. 1968. Studies on the follicular development and overwintering of the house mosquito, Culex pipiens pallens in Nagasaki area. Tropical Medicine 10:195-216.

Oktyabrskaya, T.A., Astakhova, N.A., and Boiko, L.P. 1965. Materials on the species composition, biology and ecology of mosquitoes observed near Moscow. Meditsinskaia Parazitologiia i Parazitarnye Bolezni 34:510-513.

Onyeka, J.O.A., and Boreham, P.F.L. 1987. Population studies, physiological state and mortality factors of overwintering adult populations of females of Culex pipiens L. (Diptera: Culicidae). Bulletin of Entomological Research 77:99-111.

13 Readio, J., Chen, M., and Meola, R. 1999. Juvenile hormone biosynthesis in diapausing and nondiapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 36:355-360.

Rinehart, J.P., Cikra-Ireland, R.A., Flannagan, R.D., and Denlinger, D.L. 2001. Expression of ecdysone receptor is unaffected by pupal diapause in the flesh fly, Sarcophaga crassipalpis, while its dimerization partner, USP, is downregulated. Journal of Insect Physiology 47:915-921.

Rinehart, J.P., Yocum, G.D., and Denlinger, D.L. 2000. Developmental upregulation of inducible hsp70 transcripts, but not the cognate form, during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 30:515-521.

Sanburg, L.L., and Larsen, J.R. 1973. Effect of photoperiod and temperature on ovarian development in Culex pipiens pipiens. Journal of Insect Physiology 19:1173-1190.

Spielman, A. 1964. Studies on autogeny in Culex pipiens populations in nature. I. Reproductive isolation between autogenous and anautogenous populations. American Journal of Hygiene 80:175-183.

Spielman, A. 1971. Studies on autogeny in natural populations of Culex pipiens. II. Seasonal abundance of autogenous and anautogenous populations. Journal of Medical Entomology 8:555-561.

Spielman, A. 1974. Effect of synthetic juvenile hormone on ovarian diapause of Culex pipiens mosquitoes. Journal of Medical Entomology 11:223-225.

Spielman, A., and Wong, J. 1973a. Studies on autogeny in natural populations of Culex pipiens. III. Midsummer preparation for hibernation in anautogenous populations. Journal of Medical Entomology 10:319-324.

Spielman, A., and Wong, J. 1973b. Environmental control of ovarian diapause in Culex pipiens. Annals of the Entomological Society of America 66:905-907.

Tammariello, S.P., and Denlinger, D.L. 1998. G0/G1 cell cycle arrest in the brain of Sarcophaga crassipalpis during pupal diapause and the expression pattern of the cell cycle regulator, proliferating cell nuclear antigen. Insect Biochemistry and Molecular Biology 28:83-89.

Turell, M.J., O’Guinn, M.L., Dohm, D.J., and Jones, J.W. 2001. Vector competence in North American mosquitoes (Diptera: Culicidae) for West Nile virus. Journal of Medical Entomology 38:130-134.

14 Vinogradova, E.B. 2000. Culex pipiens pipiens mosquitoes: , distribution, ecology, physiology, genetics, applied importance and control. Pensoft Publishers, Sofia. pp. 46-115.

Yocum, G.D. 2001. Differential expression of two HSP70 transcripts in response to cold shock, thermoperiod, and adult diapause in the Colorado potato beetle. Journal of Insect Physiology 47:1139-1145.

Yocum, G.D., Joplin, K.H., and Denlinger, D.L. 1998. Upregulation of a 23 kDa small heat shock protein transcript during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 28:677-682.

15

CHAPTER 2

Diapause-specific gene expression in adults of the northern house mosquito, Culex pipiens L., identified by suppressive subtractive hybridization.

ABSTRACT

In the autumn, males of Cx. pipiens die and the inseminated females enter an overwintering reproductive diapause. In this study, we probe the molecular events underpinning diapause observed in these females. Using suppressive subtractive hybridization (SSH) we have identified 40 genes that are either upregulated or downregulated during this seasonal period of dormancy. These genes can be categorized into eight functional groups: genes with regulatory functions, metabolically-related genes, those involved in food utilization, stress response genes, cytoskeletal genes, ribosomal genes, transposable elements, and genes with unknown functions. Northern blot hybridizations have confirmed the expression of 32 of our SSH clones, including 6 genes that are upregulated specifically in early diapause, 17 that are upregulated in late diapause, and 2 upregulated throughout diapause. In addition, 2 genes are diapause down-regulated and 5 remain unchanged during diapause.

16 INTRODUCTION

One of the primary avian vectors of West Nile virus in the northern United States,

Culex pipiens (L.), enters an adult diapause in late summer and early fall in response to

short daylength and low temperature (Eldridge, 1966; Sanburg and Larsen, 1973;

Spielman and Wong, 1973). The mosquitoes first appear in overwintering sites such as

caves, culverts, and unheated basements (Vinogradova, 2000) as early as August

(Service, 1968; Spielman and Wong, 1973; Onyeka and Boreham, 1987) and remain there until spring when environmental conditions again become favorable for development. Only females enter diapause and most are inseminated prior to entering the hibernation site (Onyeka and Boreham, 1987).

In preparation for diapause, females increase their lipid reserves by feeding on sugar-rich sources such as nectar and rotting fruit (Mitchell and Briegel, 1989a; Bowen,

1992). Although females programmed for diapause can be enticed to take a blood meal by being placed in close proximity to a host (Mitchell, 1983; Mitchell and Briegel,

1989b), it appears this rarely, if ever, happens in the field. Failure of diapausing females to take a blood meal is presumably the reason that so few of the overwintering females harbor West Nile virus (Nasci et al., 2000).

Many aspects of diapause in Cx. pipiens have been well documented. There is a good database that describes the physiological features of this diapause, its environmental regulators, and the hormonal control mechanism. What is currently lacking is an understanding of its molecular underpinning. Very little is known of these events in any

mosquito vector, although such work has been initiated on the embryonic diapause of Oc.

17 triseriatus, an important vector of La Crosse encephalitis (Blitvich et al., 2002). In this

species, several cDNA fragments have been identified using primers designed to amplify

sequences that the La Crosse virus cap-scavenges. Among cDNA fragments that have

been identified are a mitochondrial cytochrome c oxidase subunit, 18S and 28S ribosomal

RNAs, protein disulfide-isomerase, and several novel transcripts, but it is not yet known if any of these genes are indeed involved in regulation of this diapause or in the regulation of viral transcription.

The molecular events involved in the pupal diapause of the flesh fly, are probably the best understood; while many genes are shut down during diapause, a small cluster of genes (approximately 4 %) are diapause upregulated (Denlinger, 2002). Several classes of diapause upregulated genes have been noted, including stress response genes,

developmental arrest genes, and genes involved in regulating specific physiological

activities that are unique to diapause. Although some genes are turned on at the onset of

diapause and remain upregulated until diapause has been broken, others are uniquely

expressed only in early or late diapause. For example, heat shock protein 70 is

upregulated throughout pupal diapause in the flesh fly, Sarcophaga crassipalpis

(Rinehart et al., 2000), while cystatin is upregulated only in early diapause (Goto and

Denlinger, 2002), and ultraspiricle is upregulated only in late diapause (Rinehart et al.,

2001). Other genes, such as the cell cycle regulator proliferating cell nuclear antigen,

are shut down during diapause (Tammariello and Denlinger, 1998).

In this study, suppressive subtractive hybridization (SSH) is used to identify genes

that are differentially expressed during the adult diapause of Cx. pipiens. Expression

18 patterns are confirmed by northern blot hybridization, and the regulated genes that have

been identified are discussed in the context of their possible functional contributions to

diapause.

MATERIALS AND METHODS

Insect Rearing

An anautogenous colony of Cx. pipiens L. was established in September, 2000,

from larvae collected in Columbus, Ohio (Buckeye strain). The colony was maintained

at 25°C, 75% r.h., with a 15hL:9hD daily light:dark cycle. Eggs and first instar larvae were kept under colony conditions until the second instar, and at that time larvae were moved to an environmental chamber at 18°C, 75% r.h., and 15L:9D (nondiapause, 18°C) or placed in an environmental room under diapause-inducing conditions of 18°C, 75% r.h., with a 9L:15D daily light:dark cycle (diapause, 18°C).

Larvae were reared in 5 x 18 x 28 cm plastic containers in de-chlorinated tap water, fed a diet of ground food (TetraMin), and maintained at a density of ~250 mosquitoes per pan. Adults were kept in 30.5 cm3 screened cages and provided constant

access to water and honey-soaked sponges. Honey sponges were removed from short-

day cages 10-13 days after adult eclosion to mimic the absence of sugar in the natural

environment during the overwintering period. None of the mosquitoes used in these

experiments were offered a blood meal. To confirm diapause status, primary follicle and

19 germarium lengths were measured, and the stage of ovarian development was determined

according to the methods described by Christophers (1911) and Spielman and Wong

(1973).

Suppressive Subtractive Hybridization

Total RNA was isolated from pools of 20 females by grinding with 4.5 mm

copper-coated spherical balls (“BB’s”) in 1 ml TRIzol® Reagent (Invitrogen). After homogenization, samples were spun at 12,404 g at 4°C for 10 min, and the supernatant was used for RNA extraction following standard protocol (Chomczynski and Sacchi,

1987). RNA pellets were stored in absolute ethanol at -70°C and dissolved in 30 µl ultraPURE™ water (GIBCO) for use in cDNA synthesis. Two rounds of suppressive subtractive hybridization (SSH) were performed using the Clontech PCR-Select™ cDNA

Subtraction Kit to select for genes upregulated in early and late diapause: the first round of SSH consisted of cDNA collected from females in early diapause (short daylength,

18°C, 7-10 days post adult eclosion; tester 1) and early nondiapause (long daylength,

18°C, 7-10 days after adult eclosion; driver 1); the second round of SSH was done using

cDNA from late-diapausing females (short daylength; 18°C, 56-59 days post-adult

eclosion) as tester 2 and early nondiapausing females (long daylength, 18°C, 7-10 days

after adult eclosion) as driver 2. Ovarian dissections of the late-diapausing females (56-

59 days after adult eclosion) indicated that the females were in a late stage of diapause

(Christophers, 1911), just prior to diapause termination.

20 During our initial round of SSH, mRNA was isolated with streptavidin-coupled

paramagnetic particles using the PolyATtract® mRNA Isolation System (Promega), and

this was directly followed by cDNA synthesis according to standard SSH protocol

(Clontech). This yielded a high percentage of clones with identity to fragments of the

16S large ribosomal subunit. To reduce the abundance of 16S during our second round of SSH, mRNA isolation and cDNA synthesis were performed using the BD SMART™

PCR cDNA Synthesis Kit (BD Biosciences) following standard protocol.

Forward and reverse subtracted libraries were cloned using the TOPO TA

Cloning™ Kit (Invitrogen). Transformed plasmids were inserted into competent

Escherichia coli cells and grown overnight on Luria-Bertani (LB) plates containing X-

Gal and ampicillin. For each library, over 100 white colonies were isolated and grown overnight in LB-ampicillin broth at 37°C. Colonies were then purified with QIAprep

Spin Miniprep (QIAGEN), run on a 1% agarose gel to determine concentration, and sequenced using the vector internal primer sites (T7 and M13R) at the Ohio State

University Plant-Microbe Genomics Facility on an Applied Biosystems 3730 DNA

Analyzer using BigDye® Terminator Cycle Sequencing chemistry (Applied Biosystems) following manufacturer’s protocol.

Northern Blot Analysis

Fifteen micrograms of denatured total RNA samples were separated by electrophoresis on a 1.4% agarose denaturing gel (0.41 M formaldehyde, 1X MOPS-

EDTA-sodium acetate). Visualization of ethidium bromide stained rRNA under UV light

21 exposure was used to confirm equal loading. Following the TURBOBLOTTER™ Rapid

Downward Transfer Systems protocol (Schleicher and Schuell), the RNA was transferred for 1.5 hours onto a 0.45 micron MagnaCharge nylon membrane (GE Osmonics) using downward capillary action in 3 M NaCl, 8 mM NaOH transfer buffer, followed by neutralization in a 1 M phosphate buffer solution and UV crosslinking. The membrane was then air-dried and either stored at -20°C or used immediately for hybridization.

Digoxigenin (DIG)-labeled cDNA probes were developed from genes of interest in our forward and reverse subtracted SSH libraries. PCR was performed on each clone using the SSH nested primers (Clontech PCR-Select™ cDNA Subtraction Kit) according to the following parameters: 94°C for 3 min and 35 cycles of 94°C for 30 sec, 60°C for

30 sec, and 72°C for 2 min, followed by a 7 min extension at 72°C and a 4°C hold. The

PCR products were run on a 1% TAE agarose gel and the band of interest was excised from any remaining vector, extracted with Ultrafree®-DA (Millipore), and re-amplified by PCR. To confirm clone identity, PCR products were sequenced using the forward and reverse nested primers (Clontech) by the methods described above. The cDNAs were individually labeled in an overnight DIG reaction using 100ng of template DNA and the

Dig High Prime DNA Labeling and Detection Starter Kit II (Roche Applied Sciences).

Probes were stored at -20°C.

Hybridization was carried out overnight followed by stringency washes and immunological detection using the Dig High Prime DNA Labeling and Detection Starter

Kit II (Roche Applied Sciences) according to manufacturer’s protocol. Blots were then exposed to chemiluminescence film (Kodak Biomax). Each northern was replicated three

22 or more times. To confirm equal transfer of RNA, each membrane was stripped with 0.2

M NaOH/0.1% SDS and re-probed using DIG-labeled 28S cDNA, according to

manufacturer’s instructions.

Bioinfomatics Analyses

Sequences were edited and assembled using dnaLIMS (dnaTools) and the Baylor

College of Medicine Search Launcher: Sequence Utilities

(http://dot.imgen.bcm.tmc.edu/seq-util/seq-util.html). Similar sequences and percent identities were determined by BLASTn and BLASTx searchs in GenBank

(http://www.ncbi.nlm.nih.gov/). Low percent identities (<40%) and short matches (<30 bp) were considered not significantly similar and are thus listed as “genes with unknown function”. Only BLASTn results were listed if this search produced a high percent identity, otherwise BLASTx results were used. Both BLASTn and BLASTx identifications were listed in Tables 1 and 2 if this provided additional useful information.

Nucleotide sequences were deposited in GenBank and assigned accession numbers listed in Tables 1 and 2.

RESULTS

Early and Late Diapause Subtraction

Our first round of SSH (early diapause – early nondiapause), yielded only 5 unique diapause-upregulated clones from our initial screening of 48 clones, while the rest were identified as fragments of 16S large ribosomal subunit (with high identity to Aedes

23 aegypti large ribosomal subunit RNA gene, AY431935). The 16S sequences were all of similar length, corresponding to two major bands on the final SSH PCR gel. Therefore, before continuing with sequencing of additional clones, we constructed new libraries from selected regions of this gel, excluding the putative 16S large ribosomal subunit fragment bands. Sequencing of 95 randomly chosen clones from each forward and reverse subtracted library was much more successful, resulting in 18 unique clones that were further analyzed by northern blot hybridization.

Our second round of SSH utilized cDNA from Cx. pipiens in late-diapause (56-

59 days post adult eclosion) as the tester and cDNA from nondiapausing females (7-10 days after adult eclosion) as the driver. When constructing the late-diapause subtracted libraries, a slightly different method was employed. Rather than rely on mRNA purification to eliminate ribosomal RNA, the SMART cDNA synthesis kit (BD

Biosciences) was used to create full-length enriched cDNA pools, which were then used in the subsequent subtraction procedures. While 10 of these clones had sequences with high similarity to the 16S large ribosomal subunit gene, the remaining clones appeared to be of mRNA origin. Out of 95 clones sequenced (80 from the forward-subtracted library and 15 from the reverse-subtracted library), 22 were unique clones, and all but one were detectable by northern blot hybridization.

Confirmation by Northern Blot Analysis

Northern blot hybridizations were used to confirm the putative upregulation or downregulation of 40 cDNAs obtained in our early- and late-diapause subtracted libraries

24 (Figures 2.1 and 2.2). In most cases, northern blot hybridization confirmed the upregulation or downregulation of the cDNAs that were isolated, however, in a few cases northern blots showed a different pattern of expression than that obtained by SSH. All clones produced bands of expected size, as determined by information retrieved from

GenBank. Our early-diapause clones ranged in size from 177 to 408 bp while our late- diapause clones ranged from 133 to 1,450 bp. The highest matching sequences retrieved from GenBank, including percent identities, organism, and accession numbers, are listed in Tables 2.1 and 2.2. Several clones isolated by SSH were undetectable by northern blots, suggesting that a more sensitive technique such as real time PCR will be needed to confirm the diapause expression pattern suggested by SSH.

Diapause Upregulated Genes

Our SSH library results for early diapause yielded 18 unique clones, 11 of which were verifiable by northern blot hybridization (Figure 2.1). The following genes appear to be upregulated only in early diapause: 2 genes with regulatory functions (CpiED-A38, cytosolic small ribosomal subunit S3A; CpiED-L01, ribosomal protein S6), one gene

involved in food utilization (CpiED-A47, fatty acid synthase), one with a metabolic

function (CpiED-A07, L-malate dehydrogenase), a cytoskeletal gene (CpiED-A09,

muscle-specific actin 2), and one ribosomal gene (CpiED-A32, large ribosomal subunit

L18). Three genes encoding ribosomal protein S24 (CpiMD-C04), cytochrome oxidase subunit I (CpiMD-H06), and large ribosomal subunit (CpiMD-A01) were obtained from

our late-diapause library, but northern blots indicated that they are only upregulated in

25 early diapause. In addition, 7 genes are putatively upregulated according to SSH, but

these genes were expressed at levels undetectable by northern blot hybridization. These

include transcription elongation factor polypeptide B (CpiED-B03), thyroid hormone

receptor-associated protein TRAP170 (CpiED-D03), Methoprene-tolerant protein

(CpiED-M01), selenoprotein (CpiED-A24), disease resistance Cf-2 protein (CpiED-

C09), small ribosomal subunit 27A (CpiED-E08), and a gene with unknown function

(CpiED-A01).

We have also obtained 22 clones from our late-diapause upregulated library

(Figure 2.1), all of which were verifiable by northern blot hybridization, except for one clone, calreticulin (CpiMD-G10). Although some clones were expressed in both early and late diapause, others were upregulated only in late diapause. Genes upregulated only in late-diapause include two stress response genes (CpiMD-A06, aldehyde oxidase and

CpiMD-A11, heat shock protein 23), a gene with metabolic function (CpiMD-E09, methylmalonate-semialdehyde dehyrogenase), a ribosomal gene from the endosymbiont

Wolbachia pipientis (CpiMD-C02, 23S ribosomal RNA gene), two transposable elements

(CpiMD-D02, trasposon T1-2; CpiMD-D11, transposable element Mimo-Cp2), and 8 genes with unknown functions (CpiMD-B02, -B11, -D06, -F01, -F03, -F07, -and –H02).

In addition, 2 genes were obtained from our late-diapause library but are upregulated in both early and late-diapause. These genes encode cytochrome oxidase subunit 3

(CpiMD-H04) and a gene with unknown function (CpiMD-M43).

26 Diapause Downregulated Genes and Genes Unchanged in Diapause

Only two genes of interest were obtained from our reverse subtracted library as

being downregulated in early diapause (Figure 2.2): CpiED-A15 encoding trypsin and

CpiED-A34 encoding serine protease. These results were confirmed by northern blot

hybridization; both yielded a strong signal in nondiapausing females, no signal in early

diapause, and only a weak signal in late diapause.

In addition, several obtained from our forward and reverse subtracted libraries

showed no change in expression levels in all three stages tested (Figure 2.2). These include genes encoding poly A binding protein (CpiED-A18), ubiquitin extension protein

(CpiED-A29), cecropin A (CpiMD-A10), isoform C beta tubulin 56D (CpiED-B06), and

28S large subunit ribosomal RNA gene (CpiMD-H09). The consistency of 28S

expression prompted us to use this gene as a control for the northern blot hybridizations.

DISCUSSION

The results presented here provide some first clues about the molecular events that characterize the adult diapause of Cx. pipiens. Using SSH, we have identified 40 genes differentially expressed in diapause, and confirmed the following expression patterns by northern blot hybridization: 6 genes are upregulated specifically in early diapause, 17 genes are upregulated in late diapause, and 2 genes are upregulated throughout diapause. In addition, we have identified 2 genes that are diapause down- regulated and 5 that remained unchanged during diapause. We have categorized these genes into 8 distinct groupings: regulatory function, food utilization, stress response,

27 metabolic function, cytoskeletal, ribosomal, transposable elements, and genes with unknown functions. Northern blot hybridization has confirmed the expression of 32 of the 40 genes obtained by SSH, while the others appear to be expressed at levels undetectable by this method.

Regulatory Genes

The diapause of Cx. pipiens is characterized by a state of inactivity and a dramatically slow rate of ovarian maturation (Readio et al., 1999). Genes regulating these and other molecular events may prove useful in understanding how Cx. pipiens can survive in a prolonged inactive state. Certain ribosomal proteins have functions in regulating cell growth and death in addition to their roles in translation (Naora and Naora,

1999). The appearance of three such ribosomal proteins in our SSH libraries suggests a possible contribution of these proteins in regulating the adult diapause of Cx. pipiens.

Ribosomal protein (rp) S3A, rpS6, and rpS24 are all components of the 40S ribosomal subunit mRNA binding domain and are involved in the initiation of translation

(Takahashi et al., 2002). Two of these ribosomal proteins are expressed at low levels in early diapause and all three become highly expressed in late diapause, shortly before diapause is terminated.

The highly conserved gene encoding rpS3A is found in high concentration in the ovaries of Anopheles gambiae (Zurita et al., 1997) and in the follicular epithelial cells of

Drosophila melanogaster (Reynaud et al., 1997). Suppression of rpS3A in D. melanogaster leads to a disruption of the follicular epithelium and an inhibition of ovarian development (Reynaud et al., 1997). Since ovarian development requires high

28 protein synthesis, suppression of this gene likely leads to a disruption in this process.

Since arrested ovarian development is a prominent characteristic of Cx. pipiens diapause

(Spielman and Wong, 1973), the lack of expression of rpS3A in early diapause may be key to this developmental arrest.

Of equal interest is the undetectable level of expression of ribosomal protein S6

(rpS6) in early diapause and its subsequent upregulation in late diapause. This is another

gene that is upregulated prior to oogenesis in Ae. aegypti; rpS6 mRNA accumulates 24-

48 h after adult eclosion, remains stable until the blood meal prompts protein synthesis in

the fat body (Niu and Fallon, 2000). Of particular interest to the diapause of Cx. pipiens

is the fact that suppression of rpS6 activity has been implicated in other models of

developmental arrest. In the encysted embryos of the brine shrimp Artemia franciscana,

S6 kinase, which is required for S6 phospohrylation, shows a rapid accumulation of

mRNA 4 h after embryos are placed in hatching conditions, and the active enzyme is

detectable within 15 minutes of diapause break (Santiago and Sturgill, 2001). While

embryonic diapause differs greatly from an adult diapause, the two models suggest that

downregulation of rpS6 may be a key element in arrested development, and its

upregulation may be essential for diapause termination.

The upregulation of our clone with high similarity to the Methoprene tolerant

gene (Met) from is of particular interest because it is possibly a juvenile

hormone receptor (Wilson, 2003) and may function as a JH-dependent transcription

factor (Miura et al., 2005). The diapause of Cx. pipiens is known to be regulated by an

absence of juvenile hormone (Readio et al., 1999), thus the upregulation of Met during

29 diapause is puzzling because one might have anticipated that, if anything, Met would be downregulated at this time. The link between Met expression and the juvenile hormone mediation of diapause is unclear at this point, but it raises intriguing scenarios that may need to be considered for future work on the hormonal control of diapause in this species.

In addition, several other genes obtained by SSH may also have regulatory functions. A gene encoding calreticulin, a multi-functional Ca2+ binding protein, was obtained from our late-diapause library. Calreticulin acts as a molecular chaperone in

Ca2+-dependent pathways through calcium binding (Michalak et al., 2002), functions in the immune system (Henson et al., 2001; Johnson et al., 2001; Gao et al., 2002), and can regulate gene expression (Michalak et al., 1999). Two other putative diapause- upregulated genes encoding thyroid hormone receptor-associated protein (TRAP 170) and predicted transcription elongation factor polypeptide B are likely involved in signal transduction (Malik and Roeder, 2000) and mRNA processing (Shilatifard, 2004), respectively.

Food Utilization

A key characteristic of diapausing Cx. pipiens is that they lack the host-seeking response and will not take a blood meal under natural conditions (Mitchell, 1983; Bowen et al., 1988). Although diapausing females can be enticed to take a blood meal if the host-seeking step is bypassed (Mitchell, 1983), most of the blood ingested is ejected within 24 h and blood that remains in the gut is used neither for sequestration of lipid reserves nor for vitellogenesis (Mitchell and Briegel, 1989b). Instead, females are

30 programmed to sequester lipid reserves (Bowen, 1992; Mitchell and Briegel, 1989a) and

do so by feeding on plant sources rich in carbohydrates such as nectar and rotting fruits.

Our results show the downregulation of two genes encoding the blood digestive enzymes

trypsin and serine protease in early diapause, while the gene encoding the enzyme

involved in the conversion of sugars to lipid stores, fatty acid synthase, is highly

upregulated at this time. The evidence presented here supports the contention that even if

a blood meal is taken by diapausing Cx. pipiens, they lack the molecular machinery

required for blood digestion and are instead programmed to sequester lipid reserves. In late diapause, the accumulation of mRNAs encoding trypsin and serine protease may indicate that females are preparing to terminate diapause by regaining competence to digest a blood meal.

Stress Response

Overwintering insects confront harsh environmental conditions including low temperature, varying relative humidity, and invasion by pathogenic organisms. In several insect species, heat shock proteins are highly upregulated upon entry into diapause

(Denlinger et al., 2001). These proteins act as molecular chaperones by preventing abnormal protein folding during environmental stresses such as extreme heat, cold, or desiccation and have also been implicated in playing a role in cell cycle arrest (Feder et al., 1992). In the pupal diapause of the flesh fly Sarcophaga crassipalpis, hsp23 and

hsp70 are developmentally upregulated upon the entry into diapause and remain

expressed until diapause has been broken (Rinehart et al., 2000; Yocum et al., 1998),

31 while hsp90 is downregulated at this time but remains responsive to environmental stress

(Rinehart and Denlinger, 2000). In Cx. pipiens, hsp23 is upregulated in late diapause, but the upregulation is slight by comparison with the strong upregulation of hsp23 noted in S. crassipalpis (Yocum et al., 1998). A slight elevation of hsp70 was also noted in the adult diapause of the Colorado potato beetle (Yocum, 2001), suggesting that upregulation of hsps is not a major component of the diapause syndrome in adults. In adults of

Drosophila triauraria, hsps do not appear to be at all upregulated during diapause (Goto and Kimura, 2004)

A second stress response gene identified in late diapause is aldehyde oxidase, which encodes a multifunctional molybdo-flavoenzyme with broad substrate specificity involved in the oxidation of aromatic N-heterocycles and aldehydes (Garattini et al.,

2003). Several functions have been proposed for this enzyme including its involvement in catalyzing metabolic pathways, vitamin degradation, and detoxification of environmental pollutants (Gerattini et al., 2003). In addition, aldehyde oxidase may play an important role in insecticide resistance in the common house mosquito, Cx. quinquefasciatus (Coleman et al., 2002); our clone from Cx. pipiens shares 93% identity with aldehyde oxidase from Cx. quinquefasciatus. In certain insecticide-resistant strains of Cx. quinquefasciatus, the aldehyde oxidase gene is amplified in conjunction with two resistance-associated esterases, and the enzyme shows high substrate specificity for insecticide oxidation (Hemingway et al., 2000). Thus, aldehyde oxidase may also function in diapausing Cx. pipiens to oxidize environmental pollutants found in overwintering sites.

32 Two additional stress-response genes, selenoprotein and a disease resistance Cf-2 protein, are putatively upregulated in Cx. pipiens diapause and may confer protection against environmental stress. In D. melanogaster, selenoproteins function as antioxidants and can decrease lipid peroxidation (Morozova et al., 2003), functions that may be especially important in long-lived individuals that are in diapause. In Arabidopsis thaliana, the Cf-2-dependent disease resistance protein is involved in a fungal defense pathway (Kruger, et al., 2002). Field studies with diapausing Cx. pipiens indicate that two fungi, Cephalosporium sp. and Entomophthera sp., cause considerable mortality in overwintering populations from England (Service, 1968), thus suggesting the importance of this gene in overwintering mosquito populations.

Although cecropin A and ubiquitin extension protein were obtained from our late- diapause upregulated library, northern blot hybridizations indicate that their expression levels remain unchanged in diapause. The immune peptide, cecropin A, was detectable at very low levels in all stages tested. Further studies are needed to demonstrate if it is upregulated in response to a bacterial infection in overwintering females. The low level of expression is consistent with that observed in other species: Bartholomay et al. (2003) demonstrated that cecropin A transcripts are not detectable in naïve mosquitoes, but are rapidly transcribed after bacterial inoculation. Likewise, in spite of its isolation from our diapause library, ubiquitin extension protein, a gene associated with protein degradation and stress responses (Esser et al., 2004), was expressed equally in nondiapausing and diapausing mosquitoes, as noted with northern blots.

33 Metabolic Genes

Four genes with metabolic functions are upregulated during diapause: L-malate

dehydrogenase, methylmalonate-semialdehyde dehydrogenase, cytochrome oxidase (CO)

subunit III and COI. Two of these, COI and COIII, are of mitochondrial origin and serve

an essential role in aerobic oxidation. Although metabolic rates in insects are typically

suppressed during diapause, the metabolic suppression in adult diapauses is not as

extensive as in other stages such as the egg or pupa (Danks, 1987). The upregulation of

the two mitochondrial genes, COI and COIII, in early Cx. pipiens diapause may, at first,

seem counterintuitive, but Cx. pipiens adults are quite active prior to hibernation. They

actively seek sugar meals (Bowen, 1992), and they also must fly to their hibernation site.

Diapause preparation thus requires considerable energy. Similar results have been noted

in the early phase of larval diapause in the Japanese beetle, Popillia japonica, where

cytochrome oxidase activity actually increases during early diapause (Ludwig, 1953). In

Cx. pipiens, COI expression is depressed in late diapause, while COIII transcripts remain high.

The concurrent upregulation of L-malate dehydrogenase (MDH) and

methylmalonate semialdehyde dehydrogenase may also be involved in specific metabolic

events associated with diapause. MDH has been implicated in increased cold tolerance;

certain forms of this enzyme function more efficiently at low temperatures (Kim et al.,

1999). MDH upregulation in conjunction with increased cold tolerance has been

observed in organisms as diverse as the channel catfish Ictalurus punctatus (Seddon and

34 Prosser, 1997) and the potato Solanum sogarandinum (Rorat et al., 1997). It is not clear

what unique function methylmalonate semialdehyde dehydrogenase, an enzyme involved

in amino acid , may play during diapause.

Cytoskeletal Genes

Our experiments indicate that the expression of cytoskeletal genes is affected by

diapause: an actin is upregulated in early diapause and returns to low levels by late

diapause, while a beta tubulin is unchanged during diapause. The upregulation of an

actin in early diapause is in contrast to reports from other species, but this may reflect the

increased flight activity of females preparing for hibernation. A brain-specific actin is

downregulated during the pharate larval diapause of the gypsy moth Lymantria dispar

(Lee et al., 1998). Data from plant models indicate that actin downregulation may

contribute to the increased cold tolerance associated with dormancy. In wheat (Triticum aestivum), actin depolymerizing factor (ADF) accumulation is a major component of cold

acclimation (Ouellet et al., 2001). Upon activation, this protein sequesters actin and

induces actin depolymerization (Ouellet et al., 2001), and removal of actin from the

cytoskeleton increases membrane fluidity and thus increases resistance to cold. In Cx.

pipiens, however, not only do females actively fly during diapause preparation, but they

continue to move around within their during the winter months (Minar and

Ryba, 1971; Buffington, 1972).

35 Ribosomal Genes

In addition to the three ribosomal genes thought to serve regulatory functions (see

“Regulatory Genes”), three other ribosomal genes are upregulated in early diapause:

large ribosomal subunit L18, small ribosomal subunit 27A, and large ribosomal subunit

(16S). The fact that two of these ribosomal genes are downregulated in late diapause

(L18 was undetectable by northern blots) suggests that their function is restricted to the events of early diapause. By contrast, the 23S ribosomal RNA gene was recovered in late

diapause from the obligate intracellular bacteria of Cx. pipiens, . The strong upregulation of this gene in late diapause indicates that Wolbachia is active in late- diapausing Cx. pipiens. The differential expression of this Wolbachia gene in association with the diapause of its host suggests that the diapause status of Cx. pipiens may affect development of this bacterial parasite, as demonstrated in Wolbachia-infected eggs during the diapause of another mosquito, Ae. albopictus (Ruang-areerate et al., 2004).

Transposable Elements

Curiously, two genes encoding fractions of transposable elements, transposon

T1-2 and Mimo-Cp2, are upregulated during late diapause in Cx. pipiens. Although the function of transposable elements in diapause is unclear, diapause regulation of transposons has also been noted in two other species: two genes encoding retroviral envelope proteins are expressed during the embryonic diapause of Bombyx mori

(Yamashita et al., 2001), and a gene encoding a retrotransposon is highly expressed in the

36 early pupal diapause of S. crassipalpis (Denlinger, 2002). That transposable elements

would be diapause upregulated in all three of these species suggests an intriguing, but

still unknown, role for transposable elements in the regulation of diapause.

Genes with Unknown Function

Nine genes with unknown functions are upregulated in late diapause, as confirmed by northern blots, and one of these (CpiMD-M43) is also expressed in early diapause. This gene is of particular interest since its high level of expression in early

diapause suggests it may play a role in initiating the diapause program.

In summary, this study represents the first large-scale investigation of the

molecular aspects of diapause in any mosquito species. By suppressive subtractive

hybridization, we have demonstrated the differential regulation of genes specifically involved in early and late diapause and have categorized these genes into several distinct functional groups. Future work will certainly reveal additional genes and possibly additional gene categories that are involved in the diapause of Cx. pipiens. We anticipate that these results will be useful in probing potential molecular commonality in the diapauses of different taxa and developmental stages. We also anticipate that this type of work will prove helpful in understanding the transseasonal maintenance of West Nile virus in diapausing Cx. pipiens and may contribute to an understanding of the dynamic relationships between other pathogens and their vectors during the overwintering season.

37 REFERENCES

Bartholomay, L.C., Farid, H.A., Ramzy, R.M., and Christensen, B.M. 2003. Culex pipiens pipiens: characterization of immune peptides and the influence of immune activation on development of Wuchereria bancrofti. Molecular and Biochemical Parasitology 130:43-50.

Blitvich, B.J., Rayms-Keller, A., Blair, C.D., and Beaty, B.J. 2001. Identification and sequence determination of mRNAs detected in dormant (diapausing) Aedes triseriatus mosquito embryos. DNA Sequence 12:197-202.

Bowen, M.F. 1992. Patterns of sugar feeding in diapausing and nondiapausing Culex pipiens (Dipetera: Culicidae) females. Journal of Medical Entomology 29:843-849.

Bowen, M.F., Davis, E.E., and Haggar, D.A. 1988. A behavioral and sensory analysis of host-seeking behavior in the diapausing mosquito Culex pipiens. Journal of Insect Physiology 34:805-813.

Buffington, J.D. 1972. Hibernaculum choice in Culex pipiens. Journal of Medical Entomology 9:128-132.

Chomczynski, P., and Sacchi, N. 1987. Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction. Analytical Biochemistry 162:156-159.

Christophers, S.R., 1911. The development of the egg follicle in Anopheles. Paludism 2:73-88.

Coleman, M., Vontas, J.G., and Hemingway, J. Molecular characterization of the amplified aldehyde oxidase from insecticide resistant . European Journal of Biochemistry 269:768-779.

Danks, H.V. 1987. Insect dormancy: an ecological perspective. Biological Survey of Canada, Ottawa, pp. 19-45.

Denlinger, D.L. 2002. Regulation of diapause. Annual Review of Entomology 47:93- 122.

Denlinger, D.L., Rinehart, J.P., and Yocum, G.D. 2001. Stress proteins: a role in insect diapause? In: Denlinger, D.L., Giebultowicz, J., Sauders, D.S. (Eds.), Insect timing: circadian rhythmicity to seasonality, Elsevier Press, Oxford. pp. 155-171.

38 Eldridge, B.F. 1966. Environmental control of ovarian development in mosquitoes of the Culex pipiens complex. Science 151:826-828.

Eldridge, B.F. 1968. The effect of temperature and photoperiod on blood-feeding and ovarian development in mosquitoes of the Culex pipiens complex. American Journal of Tropical Medicine and Hygiene 17:133-140.

Eldridge, B.F., and Bailey, C.L. 1979. Experimental hibernation studies on Culex pipiens (Diptera: Culicidae): reactivation of ovarian development and blood-feeding in prehibernating females. Journal of Medical Entomology 15:462-467.

Esser, C., Alberti, S., Hohfeld, J. 2004. Cooperation of molecular chaperones with the ubiquitin/proteasome system. Biochimica et Biophysica Acta-Molecular Cell Research 1695:171-188.

Feder, J.H., Rossi, J.M., Solomon, J., Solomon, N., and Lindquist, S. 1992. The consequences of expressing hsp70 in Drosophila cells at normal temperatures. Genes and Development 6:1402-1413.

Gao, B., Adhikari, R., Howarth, M., Nakamura, K., Gold, M.C., Hill, A.B., Knee, R., Michalak, M., and Elliott, T. 2002. Assembly and antigen-presenting function of the MHC class I molecules in cells lacking the ER chaperon calreticulin. Immunity 16:99- 109.

Garattini, E., Mendel, R., Romao, M.J., Wright, R., and Terao, M. 2003. Mammalian molybdo-flavoenzymes, an expanding family1 of proteins: structure, genetics, regulation, function and pathophysiology. Biochemical Journal 372:15-32.

Goto, S.G., and Denlinger, D.L. 2002. Genes encoding two cystatins in the flesh fly Sarcophaga crassipalpis and their distinct expression patterns in relation to pupal diapause. Gene 292:121-127.

Goto, S.G., and Kimura, M.T. 2004. Heat-shock-responsive genes are not involved in the adult diapause of Drosophila triauraria. Gene 326:117-122.

Hemingway, J., Coleman, M., Paton, M., McCarroll, L., Vaughan, A., and DeSilva, D. 2000. Aldehyde oxidase is coamplified with the World’s most common Culex mosquito insecticide resistance-associated esterases. Insect Molecular Biology 9:93-99.

Henson, P.M., Bratton, D.L., and Fadok, V.A. 2001. Apoptotic cell removal. Current Biology 11:R795-R805.

39 Johnson, S., Michalak, M., Opas, M., and Eggleton, P. 2001. The ins and outs of calreticulin: from the ER lumen to the extracellular space. Trends in Cell Biology 11:122-129.

Kim, S.-Y., Hwang, K.Y., Kim, S.-H., Sung, H.-C., Han, Y.S., and Cho, Y. Structural basis for cold adaptation. Journal of Biological Chemistry 274:11761-11767.

Kruger, J., Thomas, C.M., Golstein, C., Dixon, M.S., Smoker, M., Tang, S., Mulder, L., and Jones, J.D.G. 2002. A tomato cysteine protease required for Cf-2-dependent disease resistance and suppression of autonecrosis. Science. 296:744-747.

Lee, K-Y., Hiremath, S., and Denlinger, D.L. 1998. Expression of actin in the central nervous system is switched off during diapause in the gypsy moth, Lymantria dispar. Journal of Insect Physiology 44:221-226.

Ludwig, D. 1953. Cytochrome oxidase activity during diapause and metamorphosis of the Japanese beetle (Popillia Japonica Newman). The Journal of General Physiology 36:751-757.

Malik, S., and Roeder, R.G. Transcriptional regulation through mediator-like coactivators in yeast and metazoan cells. TIBS 25:277-283.

Michalak, M., Corbett, E.F., Mesaeli, N., Nakamura, K., and Opas, M. 1999. Calreticulin: one protein, one gene, many functions. Biochemistry Journal 344:281-292.

Michalak, M., Robert Parker, J.M., and Opas, M. 2002. Ca2+ signaling and calcium binding chaperones of the endoplasmic reticulum. Cell Calcium 32:269-278.

Minar, J. and J. Ryba. 1971. Experimental studies on overwintering conditions of mosquitoes. Folia Parasitologica (PRAHA) 18:255-259.

Mitchell, C.J. 1983. Differentiation of host-seeking behavior from blood-feeding behavior in overwintering Culex pipiens (Diptera: Culicidae) and observations on gonotrophic dissociation. Journal of Medical Entomology 20:157-163.

Mitchell, C.J., and Briegel, H. 1989a. Inability of diapausing Culex pipiens (Diptera:Culicidae) to use blood for producing lipid reserves for overwinter survival. Journal of Medical Entomology 26:318-326.

Mitchell, C.J., and Briegel, H. 1989b. Fate of the blood meal in force-fed, diapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 26:332-341.

40 Miura, K., Oda, M., Makita, S., and Chinzei, Y. 2005. Characterization of the Drosophila methoprene-tolerant gene product: juvenile hormone binding and ligand- dependent gene regulation. FEBS Journal 272:1169-1178.

Morozova, N., Forry, E.P., Shahid, E., Zavacki, A.M., Harney, J.W., Kraytsberg, Y., and Berry, M.J. 2003. Antioxidant function of a novel selenoprotein in Drosophila melanogaster. Genes to Cells 8:963-971.

Naora, H., and Naora, H. 1999. Involvement of ribosomal proteins in regulating cell growth and apoptosis: translational modulation or recruitment for extraribosomal activity? Immunology and Cell Biology 77:197-205.

Nasci, R.S., Savage, H.M., White, D.J., Miller, J.R., Cropp, B.C., Godsey, M.S., Kerst, A.J., Bennett, P., Gottfried, K. and Lanciotti, R.S. 2001. West Nile virus in overwintering Culex mosquitoes, New York City, 2000. Emerging Infectious Diseases 7:742-744.

Niu, L.L., and Fallon, A.M. 2000. Differential regulation of ribosomal protein gene expression in Aedes aegypti mosquitoes before and after the blood meal. Insect Molecular Biology 9:613-623.

Onyeka, J.O.A., and Boreham, P.F.L. 1987. Population studies, physiological state and mortality factors of overwintering adult populations of females of Culex pipiens L. (Diptera: Culicidae). Bulletin of Entomological Research 77:99-111.

Ouellet, F., Carpentier, E., Cope, M.J.T.V., Monroy, A.F., and Sarhan, F. 2001. Regulation of a wheat actin-depolymerizing factor during cold acclimation. Plant Physiology 125:360-368.

Readio, J., Chen, M., and Meola, R. 1999. Juvenile hormone biosynthesis in diapausing and nondiapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 36:355-360.

Reynaud, E., Bolshakov, V.N., Barajas, V., Kafatos, F.C., and Zurita, M. 1997. Antisense suppression of the putative ribosomal protein S3A gene disrupts ovarian development in Drosophila melanogaster. Molecular and General Genetics 256:462- 467.

Rinehart, J.P., and Denlinger, D.L. 2000. Heat-shock protein 90 is down-regulated during pupal diapause in the flesh fly, Sarcophaga crassipalpis, but remains responsive to thermal stress. Insect Molecular Biology 9:641-645.

41 Rinehart, J.P., Yocum, G.D., and Denlinger, D.L. 2000. Developmental upregulation of inducible hsp70 transcripts, but not the cognate form, during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 30:515-521.

Rinehart, J.P., Cikra-Ireland, R.A., Flannagan, R.D., and Denlinger, D.L. 2001. Expression of ecdysone receptor is unaffected by pupal diapause in the flesh fly, Sarcophaga crassipalpis, while its dimerization partner, USP, is downregulated. Journal of Insect Physiology 47:915-921.

Rorat, T., Irzykowski, W., and Grygorowicz, W.J. 1997. Identification and expression of novel cold induced genes in potato (Solanum sogarandium). Plant Science 124:69-78.

Ruang-areerate, T., Kittayapong, P., McGraw, E.A., Baimai, V., and O’Neill, S.L. 2004. Wolbachia replication and host cell division in . Current Microbiology 49:1-12.

Sanburg, L.L., and Larsen, J.R. 1973. Effect of photoperiod and temperature on ovarian development in Culex pipiens pipiens. Journal of Insect Physiology 19:1173-1190.

Santiago, J., and Sturgill, T.W. 2001. Identification of the S6 kinase activity stimulated in quiescent brine shrimp embryos upon entry to preemergence development as p70 ribosomal protein S6 kinase: isolation of Artemia franciscana p70S6K cDNA. Biochemistry and Cell Biology 79:141-152.

Seddon, W.L., and Prosser, C.L. 1997. Seasonal variations in the temperature acclimation response of the channel catfish, Ictalurus punctatus. Physiological Zoology 70:33-44.

Service, M.W. 1968. Observations on the ecology of some British mosquitoes. Bulletin of Entomological Research 59:161-194.

Shilatifard, A. 2004. Transcriptional elongation control of RNA polymerase III: a new frontier. Biochimica et Biophysica Acta 1677:79-86.

Spielman, A., and Wong, J. 1973. Environmental control of ovarian diapause in Culex pipiens. Annals of the Entomological Society of America 66:905-907.

Takahashi, Y., Mitsuma, T., Hirayama, S., and Odani, S. 2002. Identification of the ribosomal proteins present in the vicinity of globin mRNA in the 40S initiation complex. Journal of Biochemistry 132:705-711.

42 Tammariello, S.P., and Denlinger, D.L. 1998. G0/G1 cell cycle arrest in the brain of Sarcophaga crassipalpis during pupal diapause and the expression pattern of the cell cycle regulator, proliferating cell nuclear antigen. Insect Biochemistry and Molecular Biology 28:83-89.

Vinogradova, E.B. 2000. Culex pipiens pipiens mosquitoes: taxonomy, distribution, ecology, physiology, genetics, applied importance and control. Pensoft Publishers, Sofia. pp. 46-115.

Wilson, T.G. 2003. Methoprene-tolerant, a bHLH-PAS gene essential for insect endocrinology. In: Crews, S.T. (Ed.), PAS proteins: regulators and sensors of development and physiology. Kluwer Academic, Boston, pp. 109-132.

Yamashita, O., Shiomi, K., Ishida, Y., Katagiri, N., and Niimi, T. 2001. Insights for future studies on embryonic diapause promoted by molecular analyses of diapause hormone and its action in Bombyx mori. In: Denlinger, D.L., Giebultowicz, J., and Saunders, D.S. (Ed.), Insect timing: circadian rhythmicity to seasonality. Elsevier Science B.V., New York, pp. 145-153.

Yocum, G.D. 2001. Differential expression of two HSP70 transcripts in response to cold shock, thermoperiod, and adult diapause in the Colorado potato beetle. Journal of Insect Physiology 47:1139-1145.

Yocum, G.D., Joplin, K.H., and Denlinger, D.L. 1998. Upregulation of a 23 kDa small heat shock protein transcript during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 28:677-682.

Zurita, M., Reynaud, E., and Kafatos, F.C. 1997. Cloning and characterization of cDNAs preferentially expressed in the ovary of the mosquito Anopheles gambiae. Insect Molecular Biology 6:55-62.

43

X X X X X X X X X X X XX X X X northern northern

gov/BLAST/). In the blast , AF202953 , AAG02462 , T13418 , AE017256

, AAC63387 r r r , AY441061 , AAM14384.1 , U23710 , AF457547 , AF417833 , AF417833 ractive hybridization. Percent ractive hybridization. Percent

, AY432783 , XP_143238.2 , AY431319 , AY431545 , AY431935 , AY431732 , AY431617 , AAX55681 , AY064700 , AY289764 , AY433139 , AF466603 , AF425847 (18°C, short daylength; 7-10 days post LASTn) and/or translation (BLASTx) LASTn) and/or translation (BLASTx)

diapause, ED and LD, respectively. edes aegypti us musculus us Armigeres subalbatus Drosophila melanogaste Drosophila Anopheles gambiae Arabidopsis thaliana Arabidopsis Culex pipiens quinquefasciatus Culex pipiens Aedes aegypti Aedes Aedes aegypti Aedes Aedes aegypti Aedes melanogaste Drosophila M Wolbachia pipientis Wolbachia Aedes aegypti Aedes aegypti Aedes Culex tarsalis Uranotaenia lowii Uranotaenia Sarcophaga crassipalpis Sarcophaga Aedes aegypti Aedes Aedes aegypti Aedes Aedes aegypti Aedes Aedes aegypti Aedes melanogaste Drosophila Aedes aegypti Aedes

identity, organism, accession # accession identity, organism, ED LD 85%, 59%, 87%, 42%, 93%, 91%, 86%, 100%, A 98%, 97%, 97%, 97%, 87%, 92%, 87%, 49%, 40%, 77%, 98%, 89%, 92%, 85%, 94%,

Table 1.1 (continued). (continued). 1.1 Table

GenBank (http://www.ncbi.nlm.nih. through a nucleotide (B , isolated by suppressive subt

Cx. pipiens n blot is indicated by an “X” in early

ylength; 56-59 days post adult eclosion)

sion #’s were retrieved from 23S ribosomal RNA generibosomal 23S n

l tress response tress ood utilization CpiED-A47 230 synthase fatty-acid n CpiED-D03 614TRAP170 protein receptor-associated hormone thyroid x CpiED-A24 306 selenoprotein n CpiED-C09 306 Cf-2 protein resistance disease putative x CpiMD-A06 1,450 oxidase aldehyde n regulatory CpiED-A38 469 S3A subunit ribosomal small cytosolic n Diapause Upregulated Genes clone size (bp) putative identityf s blast % CpiED-A32 177 L18 subunit ribosomal large cytosolic n CpiED-E08CpiMD-A01CpiMD-C02 105 231 324 27A subunit ribosomal small cytosolic RNA subunit ribosomal large endosymbiont Wolbachia x n n CpiED-L01 296 S6 protein ribosomal CpiMD-H04 805 3 subunit oxidase cytochrome CpiMD-H06ribosoma 912 I subunit oxidase cytochrome n CpiMD-H09 291 RNA gene ribosomal subunit large 28S n CpiMD-A11 483 protein shock 23kDa heat x CpiED-B03 356 B polypeptide factor elongation transcription predicted x CpiED-M01 70 protein methoprene-tolerant putative x CpiMD-C04 295 S24 protein ribosomal n cytoskeletal CpiED-A09 408 2 actin muscle-specific n metabolic function metabolic CpiED-A07CpiMD-E09 199 1,055 dehydrogenase L-malate dehydrogenase putative methylmalonate-semialdehyde x n CpiMD-G10 519 calreticulin putative

adult eclosion) and late (18°C, short da identities, organisms, and acces column, “n” and “x” indicate putative identities obtained query, respectively. Confirmation by norther Table 2.1. Diapause upregulated genes from 44

X X X X X X X X X X X

XX northern

, B34751 , EAA13087 y y y y y y y y y , AF217612

Anopheles gambiae Anopheles Culex pipiens gambiae Anopheles 44%, 87%, 65%, n,xn,x significant no similarit n,x significant no similarit n,x significant no similarit n,x significant no similarit significant no similarit n,xn,x significant no similarit n,x significant no similarit n,x significant no similarit significant no similarit

y y y y y y y y y

44

Diapause Upregulated Genes Continued Genes Upregulated Diapause clone elements transposable CpiMD-D02 (bp) size identity putative 897 T1-2 transposon blast % # accession identity,organism, x ED LD CpiMD-D11 function unknown with genes CpiED-A01 889 Mimo-Cp2 element transposable 859 similarit significant no n CpiMD-B02CpiMD-B11CpiMD-D06CpiMD-F01CpiMD-F03 312CpiMD-F07 841 1,047CpiMD-H02 similarit significant no CpiMD-H03 445 similarit significant no similarit CpiMD-A43 significant no 311 250 similarit significant no 312PEST str. gambiae Anopheles 133 similarit significant no 561 similarit significant no similarit significant no similarit significant no x Table 2.1 (continued).

45

Diapause Downregulated and Genes Unchanged in Diapause northern clone size (bp) putative identity blast % identity, organism, accession # ND ED LD

Diapause Downregulated Genes food utilization CpiED-A15 294 serine protease x 40%, Anopheles gambiae , AAA73920 X CpiED-A34 264 trypsin n 98%, Culex pipiens pallens , AY034060 X

Genes Unchanged in Diapause regulatory CpiED-A18 292 putative poly A binding n 91%, Aedes aegypti , AY431644 XXX stress response CpiED-A29 295 ubiquitin extension protein n 94%, Drosophila melanogaster , X59943 XXX CpiMD-A10 333 cecropin A n 79%, Culex pipiens pipiens , AY189808 XXX cytoskeletal CpiED-B06 241 isoform C (betaTub56D) n 89%, Drosophila melanogaster , NM_166357 XXX

Table 2.2. Diapause downregulated genes from Cx. pipiens, isolated by suppressive subtractive hybridization, and genes unchanged in diapause. Percent identities, organisms, and accession #’s were retrieved from GeneBank (http://www.ncbi.nlm.nih.gov/BLAST/). In the blast column, “n” and “x” indicate putative identities obtained through a nucleotide (BLASTn) and/or translation (BLASTx) query, respectively. Confirmation by northern blot is indicated by an “X” in nondiapause (18°C, long daylength; 7-10 days post adult eclosion) and early diapause (18°C, short daylength; 7-10 days post adult eclosion), and late diapause (18°C, short daylength; 56- 59 days post adult eclosion).

46

short Mimo-Cp2 28S CpiMD-D11 on). Each on). Each

CpiMD-M43

LD ED ND CpiMD-D02 transposon T1-2 LD ED ND LD ED ND G. Transposable Elements G. Transposable actin CpiMD-H03 ND ED LD ED ND CpiED-A09 ND ED LD ED ND D. Cytoskeletal

CpiMD-C02 23S (Wolbachia) ND ED LD ED ND

are listed above each nd putative identities CpiMD-H02 ND ED LD ED ND hsp23 CpiMD-A11 after adult eclosi days 56-59 t daylength; ND ED LD ED ND CpiMD-A01 ND ED LD ED ND large ribosomal subunit ribosomal large CpiMD-F07

LD ED ND

CpiMD-A06 ng was confirmed by northern blot hybridization with a hybridization northern blot ng was confirmed by ND ED LD ED ND aldehyde oxidase C. StressC. Response CpiED-A32 ys after adult eclosion), ED = females in early diapause (18°C,

ND ED LD ED ND ribosomal subunit L18 subunit ribosomal

CpiMD-F03 LD ED ND

upregulated SSH clones. Clone ID a

CpiED-A47 ND ED LD ED ND fatty acid synthase acid fatty

B. Food Utilization Food B. CpiMD-F01 CpiMD-H06 ND ED LD ED ND LD ED ND cytochrome c oxidase I Cx. pipiens LD = females in late diapause (18°C, shor

CpiMD-C04 CpiMD-H04 CpiMD-D06 ND ED LD ED ND LD ED ND ribosomal protein S24 protein ribosomal cytochrome oxidase c 3

CpiED-L01 CpiMD-B11 CpiMD-E09 ND ED LD ED ND LD ED ND ribosomal protein S6 protein ribosomal meth. dehydrogenase meth. g of total RNA pooled from 20 females. Equal loadi g of total

µ

CpiED-A38 CpiED-A07 CpiMD-B02

ND ED LD ED ND LD ED ND LD ED ND LD ED ND LD ED ND malate dehydrogenase ribosomal subunit S3A subunit ribosomal A. Regulatory FunctionE. Metabolic Function Unknown with Genes H. F. Ribosomal Diapause Upregulated Genes Upregulated Diapause 28S 28S 28S of hybridization Northern blot Figure 2.1. cDNA probe. lane contains 15 blot. ND = nondiapausing females (18°C, long daylength; 7-10 da 7-10 daylength; females (18°C, long blot. ND = nondiapausing days after adult eclosion), 7-10 daylength;

47

Diapause Downregulated Genes

A. Food Utilization CpiED-A15 CpiED-A34 serine protease trypsin

28S

ND ED LD ND ED LD

Genes Unchanged in Diapause

B. Regulatory C. Stress Response D. Cytoskeletal CpiED-A18 CpiED-A29 CpiMD-A10 CpiED-B06 poly A binding ubiquitin cecropin A beta tubulin

28S

ND ED LD ND ED LD ND ED LD ND ED LD

Figure 2.2. Northern blot hybridization of Cx. pipiens diapause downregulated genes and genes unchanged in diapause. Clone ID and putative identities are listed above each blot. ND = nondiapausing females (18°C, long daylength; 7-10 days after adult eclosion), ED = females in early diapause (18°C, short daylength; 7-10 days after adult eclosion), LD = females in late diapause (18°C, short daylength; 56-59 days after adult eclosion). Each lane contains 15 µg of total RNA pooled from 20 females. A 28S cDNA probe was used to confirm equal loading.

48

Diapause Downregulated and Genes Unchanged in Diapause northern clone size (bp) putative identity blast % identity, organism, accession # ND ED LD

Diapause Downregulated Genes food utilization CpiED-A15 294 serine protease x 40%, Anopheles gambiae , AAA73920 X CpiED-A34 264 trypsin n 98%, Culex pipiens pallens , AY034060 X

Genes Unchanged in Diapause regulatory CpiED-A18 292 putative poly A binding n 91%, Aedes aegypti , AY431644 XXX

stress response CpiED-A29 295 ubiquitin extension protein n 94%, Drosophila melanogaster , X59943 XXX CpiMD-A10 333 cecropin A n 79%, Culex pipiens pipiens , AY189808 XXX

cytoskeletal CpiED-B06 241 isoform C (betaTub56D) n 89%, Drosophila melanogaster , NM_166357 XXX

Table 2. Diapause downregulated genes from Cx. pipiens, isolated by suppressive subtractive hybridization, and genes unchanged in diapause. Percent identities, organisms, and accession #’s were retrieved from GeneBank (http://www.ncbi.nlm.nih.gov/BLAST/). In the blast column, “n” and “x” indicate putative identities obtained through a nucleotide (BLASTn) and/or translation (BLASTx) query, respectively. Confirmation by northern blot is indicated by an “X” in nondiapause (18°C, long daylength; 7-10 days post adult eclosion) and early diapause (18°C, short daylength; 7-10 days post adult eclosion), and late diapause (18°C, short daylength; 56-59 days post adult eclosion).

49

CHAPTER 3

Diapause in the mosquito Culex pipiens evokes a metabolic switch that shuts down genes

encoding blood digestive enzymes and upregulates a gene associated with sugar

utilization and lipid storage.

ABSTRACT

A key characteristic of diapause in Culex pipiens is the switch from blood-feeding in nondiapausing individuals to sugar feeding in diapause-destined females. We present evidence demonstrating that genes encoding enzymes needed to digest a blood meal

(trypsin and a chymotrypsin-like protease) are downregulated in prehibernating females, and concurrently a gene associated with the accumulation of lipid reserves (fatty acid synthase) is highly upregulated. As females then enter the hibernation state (diapause) fatty acid synthase is only sporadically expressed, and expression of trypsin and chymotrypsin-like remain undetectable. Late in diapause (2-3 months at 18°C) the genes

encoding the digestive enzymes begin to be expressed as the female prepares to take a

blood meal upon termination of diapause. The results thus underscore a molecular switch

that either directs the female towards blood feeding (nondiapause) or sugar feeding and

lipid sequestration (diapause).

50 INTRODUCTION

The northern house mosquito, Culex pipiens (L.), plays a prominent role in maintaining the natural reservoirs of several arthropod-borne viruses including St. Louis

and West Nile viruses in North America. The ecology and physiology of Cx. pipiens

have therefore been extensively studied, with particular interest being paid to the

overwintering stage. Winter is typically characterized by a significant decrease or

absence of vector-borne diseases, and several studies have shown that the overwintering

stage of the mosquito can harbor pathological agents during this time and thus reinitiate

disease transmission the following spring (Bellamy et al., 1958; Watts et al., 1974;

Reisen et al., 2002). This appears to be a likely scenario explaining the reappearance of

West Nile virus in the New York City area following the initial outbreak in the summer

of 1999 (Nasci et al., 2001). In this case, diapausing adult females of Cx. pipiens were

found to harbor the virus during the winter months, and such mosquitoes may have

reintroduced the virus into bird populations the following season.

Cx. pipiens is a temperate zone species that enters an overwintering dormancy

(diapause) in response to short daylength and low temperatures received in the fourth

larval instar and early pupal stage (Eldridge, 1966; Sandburg and Larsen 1973; Spielman

and Wong 1973). Diapausing individuals first appear in overwintering sites such as

caves, culverts and unheated basements (Vinogradova, 2000) as early as August

(Spielman and Wong, 1973; Onyeka and Boreham, 1987). Only adult females enter

diapause; males die after fertilizing the females in the autumn (Service, 1968; Onyeka

and Boreham, 1987). One of the main features of diapause is that the primary ovarian

51 follicles remain in a state of arrested development (Eldridge, 1966; Sanburg and Larsen,

1973; Spielman and Wong 1973). In addition, a number of behavioral changes occur,

including a lack of host-seeking behavior and the concurrent increase in feeding on

carbohydrate-rich nectar, rotting fruits, and other plant products, which leads to a hypertrophy of the fat body prior to the onset of diapause (Mitchell, 1983; Mitchell and

Briegel, 1989a; Bowen, 1992).

Females programmed for diapause can be enticed to take a blood meal under laboratory conditions if placed in close proximity to a host (Eldridge and Bailey, 1979;

Bailey et al., 1982; Mitchell and Briegel, 1989b), but most of the blood ingested by diapausing females is ejected. Blood that remains in the midgut is not used to increase lipid reserves, and only a few females have been observed to use this blood to initiate vitellogenesis (Mitchell and Briegel, 1989b). In diapause, females of Cx. pipiens lack high-sensitivity olfactory responsiveness to the host-attractant, lactic acid (Bowen et al.,

1988; Bowen, 1990), suggesting that the lack of blood feeding is associated with a shut- down of the host-seeking response. Thus, the arrest in ovarian development observed in diapausing Cx. pipiens is normally accompanied by a halt in blood feeding until diapause has been broken (Swellengrebel, 1929; Washino, 1977). Like most adult diapauses

(Denlinger, 1985), the diapause of Cx. pipiens appears to be the consequence of a shut- down in juvenile hormone synthesis by the corpora allata (Spielman, 1974; Readio et al.,

1999).

The hypertrophy of the fat body and elevation of lipid reserves that are associated with diapause are linked to a boost in sugar feeding that accompanies the entry into

52 diapause. The accumulation of lipid reserves occurs after adult eclosion; within a week,

females programmed for diapause by short daylength accumulate twice as much lipid as

their nondiapause counterparts, and these reserves are largely depleted during the course

of the winter (Mitchell and Briegel 1989a).

Although the physiological and ecological aspects of blood and sugar feeding in diapausing Cx. pipiens have been well described, the molecular events that contribute to

this metabolic decision have not been explored. In this study, we isolate and sequence three clones encoding digestive enzymes by suppressive subtractive hybridization: fatty

acid synthase, trypsin, and chymotrypsin-like serine protease, and then use these clones

to probe the metabolic pathways associated with the decision by the mosquito to enter

and terminate diapause. Since diapause in Cx. pipiens is programmed by both short

daylength and low temperature, we also distinguish between temperature and

photoperiodic effects. We conclude that the short-day programming of diapause results

in the downregulation of the genes encoding the blood digestive enzymes and the

upregulation of a gene associated with lipid sequestration.

MATERIALS AND METHODS

Insect Rearing

An anautogenous colony of Cx. pipiens L. was established in September, 2000,

from larvae collected in Columbus, Ohio (Buckeye strain). The colony was maintained

at 25°C, 75% r.h., with a 15hL:9hD daily light:dark cycle. Eggs and 1st instar larvae

were kept under colony conditions until the 2nd instar, and at that time larvae were either

53 kept in the colony rearing room (nondiapause, 25°C), moved to an environmental

chamber at 18°C, 75% r.h., and 15L:9D (nondiapause, 18°C), or placed in an

environmental room under diapause-inducing conditions of 18°C, 75% r.h., with a

9L:15D daily light:dark cycle (diapause, 18°C).

Larvae were reared in 18 x 28 x 5 cm plastic containers in de-chlorinated tap

water, fed a diet of ground fish food (TetraMin), and maintained at a density of ~250

mosquitoes per pan. Adults were kept in 30.5 x 30.5 x 30.5 cm screened cages and

provided constant access to water and honey-soaked sponges. Honey sponges were

removed from short-day cages 10-13 days after adult eclosion to mimic the absence of

sugar in the natural environment during the overwintering period. None of the

mosquitoes used in these experiments were offered a blood meal. To confirm diapause status, primary follicle and germarium lengths were measured and the stage of ovarian development was determined according to the methods described by Christophers (1911) and Spielman and Wong (1973).

Suppressive Subtractive Hybridization

Total RNA was isolated from pools of 20 females by grinding with 4.5 mm copper-coated spherical balls (“BB’s”) in 1 ml TRIzol® Reagent (Invitrogen). After homogenization, samples were spun at 12,404 g at 4°C for 10 min and the supernatant was used for RNA extraction following standard protocol (Chomczynski and Sacchi,

1987). RNA pellets were stored in absolute ethanol at -70°C and dissolved in 30 µl ultraPURE™ water (GIBCO) for use in cDNA synthesis (BD SMART™ PCR cDNA

54 Synthesis Kit, BD Biosciences) following standard protocol. Suppressive subtractive

hybridization (SSH) was performed using the Clontech PCR-Select™ cDNA Subtraction

Kit: the forward subtracted library was constructed using females in early diapause (7-10

days post adult eclosion, short daylength, 18°C) and the reverse subtracted library was

constructed using nondiapausing females (7-10 days post adult eclosion, long daylength,

18°C). Forward and reverse libraries were cloned using the TOPO TA Cloning™ Kit

(Invitrogen). Transformed plasmids were inserted into competent Escherichia coli cells

and grown overnight on Luria-Bertani (LB) plates containing X-Gal and ampicillin. For

each library, over 100 white colonies were isolated and grown overnight in LB-ampicillin

broth at 37°C. Colonies were then purified with QIAprep Spin Miniprep (QIAGEN) and

sequenced using the vector internal primer sites (T7 and M13R) at the Ohio State

University Plant-Microbe Genomics Facility on an Applied Biosystems 3730 DNA

Analyzer using BigDye® Terminator Cycle Sequencing chemistry (Applied Biosystems)

following manufacturer’s protocol.

Northern Blot Analysis

RNA was extracted from adults and pupae following the methods described

above. Pupae were sexed as females based on their large size and prolonged

development. Fifteen micrograms of denatured total RNA samples were separated by

electrophoresis on a 1.4% agarose denaturing gel (0.41 M formaldehyde,

1X MOPS-EDTA-sodium acetate). Visualization of ethidium bromide stained rRNA

under UV light exposure was used to confirm equal loading. Following the

55 TURBOBLOTTER™ Rapid Downward Transfer Systems protocol (Schleicher and

Schuell), the RNA was transferred for 1.5 hours onto a 0.45 micron MagnaCharge nylon

membrane (GE Osmonics) using downward capillary action in 3 M NaCl, 8 mM NaOH

transfer buffer, followed by neutralization in a 1 M phosphate buffer solution and UV crosslinking. The membrane was then air-dried and either stored at -20°C or used immediately for hybridization.

Digoxigenin (DIG)-labeled cDNA probes were developed from the three metabolically related genes generated in our forward and reverse subtracted SSH libraries. PCR was performed on each clone using the SSH nested primers (Clontech

PCR-Select™ cDNA Subtraction Kit) according to the following parameters: 94°C for 3 min and 35 cycles of 94°C for 30 sec, 60°C for 30 sec, and 72°C for 2 min, followed by a

7 min extension at 72°C and a 4°C hold. The PCR products were run on a 1% TAE agarose gel and the band of interest was excised from any remaining vector, extracted with Ultrafree®-DA (Millipore), and re-amplified by PCR. Fatty acid synthase, trypsin, and cymotrypsin-like serine protease cDNAs were individually labeled in an overnight

DIG reaction using 100ng of template DNA and the Dig High Prime DNA Labeling and

Detection Starter Kit II (Roche Applied Sciences). Probes were stored at -20°C.

Hybridization was carried out overnight followed by stringency washes and immunological detection using the Dig High Prime DNA Labeling and Detection Starter

Kit II (Roche Applied Sciences) according to manufacturer’s protocol. Blots were then exposed to chemiluminescence film (Kodak Biomax). Each northern was replicated three or more times. To confirm equal transfer of RNA, each membrane was stripped with 0.2

56 M NaOH/0.1% SDS and re-probed using DIG-labeled 28S cDNA, according to

manufacturer’s instructions.

3’ and 5’ RACE

The SMART™ RACE cDNA Amplification Kit (Clontech) was used to perform

both 5’ and 3’-rapid amplification of cDNA ends (RACE). For 3’ RACE, first strand

cDNA was synthesized from 5 µg total RNA using the manufacturer’s provided adaptor primer. Target cDNA was then amplified using the Universal Amplification Primer and a forward, gene-specific primer based on the sequence of the original clones: fatty acid

synthase (5’-AAT TAC GCC AAA CTG C AA GG-3’), trypsin (5’-CAA CTT CCT CTC

GTC CGG TA-3’), and chymotrypsin-like (5’GAT GAT CTG CCC AAG GAC TC-3’).

PCR consisted of a “hot start” at 94°C for 3 min and 35 cycles of 94°C for 30s, 60°C for

30s, and 72°C for 2 min, followed by an additional 7 min at 72°C.

The 5’ RACE was carried out using two gene specific reverse primers according

to manufacturer’s protocol. Template cDNA was synthesized with trypsin (5’-ACG TTG

GAG TAA ATT C-3’) and chymotrypsin-like (5’-TAT CAC AAC TCT TAA TTT C-3’)

reverse primers. After purification and TdT tailing of the cDNA, PCR was used to

amplify each gene using a second set of nested, reverse gene-specific primers and the

provided Abridged Anchor Primer. The PCR gene-specific primers were as follows:

trypsin (5’-GGT GAC CTC GCG ATC ATA GT-3’) and chymotrypsin-like (5’-TCC

CAA CCT TTT CGT TGA AG-3’). Two rounds of PCR were done according to the

parameters described above. The 3’- and 5’-RACE products were cloned and double-

57 sequenced as described above. Although several sets of gene-specific primers were used

in attempts to amplify the 5’ end of fatty acid synthase, full-length sequence was not

obtained.

Bioinfomatics Analyses

The 5’ and 3’ RACE products were edited and assembled using dnaLIMS

(dnaTools) and BioEdit Sequence Alignment Editor (Isis Pharmaceuticals). Similar sequences were identified by performing a BLASTn and BLASTx search in GenBank

(http://www.ncbi.nlm.nih.gov/). The deduced amino acid sequences were assembled, analyzed, and aligned with similar sequences using Blastp (NCBI), the Baylor College of

Medicine Search Launcher: Sequence Utilities (http://dot.imgen.bcm.tmc.edu/seq- util/seq-util.html), and BoxShade Server 3.21

(http://www.ch.embnet.org/software/BOX_form.html). Percent identities were obtained by blasting two sequences with the Blastp server using BLOSUM62 matrix in NCBI’s web server. The full-length nucleotide sequences for trypsin and chymotrypsin-like serine protease and the 3’ end of fatty acid synthase were deposited in GenBank and assigned accession numbers AY958426, AY958427, and AY958428, respectively.

RESULTS

Clone Identification

Three metabolically-related genes were identified by suppressive subtractive hybridization at the onset of diapause (7-10 days post adult eclosion): two blood digestive

58 enzymes in the class of serine proteases, trypsin and chymotrypsin-like serine protease,

were among the diapause-regulated genes that were putatively downregulated (obtained

from our reverse subtracted library), and fatty acid synthase was putatively diapause upregulated (obtained from our forward subtracted library). The cDNA matching trypsin is a 267 bp fragment matching with 96% identity the complete coding sequence of Cx. pipiens pallens trypsin mRNA. This portion of the clone corresponds to an 88 residue

deduced amino acid sequence that matches Cx. pipiens pallens amino acids 1 through 76.

The second clone, encoding a chymotrypsin-like gene, was also identified in the reverse

subtracted (downregulated) library. The chymotrypsin-like clone is 294 bp long and

encodes a deduced amino acid sequence of 97 residues, with 40% identity and 51%

positive matches to a serine protease from Anopheles gambiae. The forward subtracted

library (upregulated) yielded a 230 bp cDNA fragment with 85% identity to an Armigeres

subalbatus fatty acid synthase mRNA expressed sequence tag (EST). The deduced

amino acid sequence is 76 residues long and matches amino acids 2408 through 2468 in

fatty acid synthase from An. gambiae. The original clones of trypsin, chymotrypsin-like,

and fatty acid synthase were used to generate DIG-labeled probes for northern blot hybridization, producing bands of 0.9, 0.9, and ~8.0 Kb, respectively.

To obtain full-length cDNAs of trypsin, chymotrypsin-like, and fatty acid

synthase, the original SSH partial cDNA sequences were used to design gene-specific

primers for 5’ and 3’ RACE amplification. The resulting trypsin PCR products for 5’ and

3’ RACE were 397 and 501 bp, respectively. Both sequences overlapped the initial SSH

clone and yielded a total product size of 898 bp with a 783 bp open reading frame, as

59 shown in Figure 3.1. Our clone has a 46 and 27 base pair 5’ and 3’ untranslated region, respectively, with a putative polyadenylation signal (AATAAA) occurring at position

839. The deduced protein has an open reading frame starting at nucleotide 91 and is 261 residues long. Figure 3.2 shows the full-length sequence of the chymotrypsin-like cDNA, along with the predicted 240 residue amino acid sequence. The 5’ RACE yielded a 609 bp product, and 3’ RACE produced a segment 435 bp long. Together they form the complete chymotrypsin-like cDNA, which is 881 bp with an open reading frame of 720 bp. The 5’ untranslated region is 47 base pairs long, and the 3’ untranslated region is 94 bp with a polyadenylation site occurring at position 811.

The 3’ end of fatty acid synthase was obtained by RACE and resulted in a 954 bp clone that encodes 48 amino acids (Figure 3.3). Our clone has a large 3’ untranslated region 810 bp long. The putative polyadenylation signal was identified at position 906.

A Blastx search revealed that this segment includes a portion of the thioesterase domain, as determined in other known protein sequences (Holzer et al., 1989). Attempts at 5’

RACE were unsuccessful, thus the full-length sequence of fatty acid synthase was not obtained.

Comparison and Analysis of the Deduced Protein Sequences

The Cx. pipiens trypsin and chymotrypsin-like open reading frames include predicted mature active peptides in the family of serine proteases, identified by the histidine, aspartic acid, and serine residues that form the characteristic catalytic triad

60 (Walsh and Wilcox, 1970). The three cysteine bridges that form the disulphide bonds

essential for holding the polypeptide chains together are also conserved (Figures 3.4 and

3.5).

A multiple sequence alignment of the deduced trypsin amino acid sequence with

the sequences of other insect trypsins is shown in Figure 3.4. The Cx. pipiens trypsin

ORF shares 92% identity with trypsin from a close relative, Cx. pipiens pallens. When

compared with the well-described trypsins from Ae. aegypti, our sequence aligns closest with early trypsin (51%), compared to 31% identity with late trypsin. In addition, our trypsin aligns with 53% and 51% identities to an An. gambiae trypsin and Drosophila

melanogaster trypsin-like protease, respectively. The deduced amino acid trypsin sequence has a predicted activation peptide 16 amino acids long, assuming a signal peptide cleavage site after Gly-18 and a putative activation peptide cleavage site after

Lys-34. The putative signal peptide cleavage site was determined using the PSORT

program through the www at http://psort.nibb.ac.jp/form2.html (McGeoch, 1985; Von

Heijne, 1986), and the activation peptide cleavage site was determined from the highly

conserved IVGG sequence, common at the start of most active serine protease enzymes

(Yan et al., 1997; Muller et al., 1995). Our putative active chymotrypsin-like peptide,

however, begins with the tetrapeptide IFGG. The Asp residue characteristic of trypsin-

like serine proteases was also identified in Figure 3.4, along with the three residues that

make up the zymogen triad (Ser/His/Asp) that are involved in stabilization of the

proenzyme (Madison et al., 1993).

61 Our full-length clone encoding the second downregulated digestive enzyme is most similar to serine proteases with chymotrypsin activity, as indicated by a Blastp search. The serine protease ORF aligns most closely with an undescribed protein from the An. gambiae genome project with which it shares a 45% identity (Figure 3.5). A multiple sequence alignment highlights the identities with other insect serine proteases as follows: An. gambiae serine protease, 45%; An. darlingi chymotrypsin 1, 36%; and An. darlingi chymotrypsin 2, 37%. Using the PSORT program, we predicted the cleavage site of the signal peptide to be between Ala-20 and Arg-21. This would leave a dipeptide activation segment prior to the tryptic cleavage at Arg-22. In addition, the conserved Gly residue characteristic of chymotrypsin-like serine proteases (Kraut, 1977; Warshel et al.,

1989) is noted in Figure 3.5.

Our Cx. pipiens fatty acid synthase ORF is multi-aligned at the 3’ end with other fatty acid synthase sequences retrieved from GenBank (Figure 3.6). The Cx. pipiens fatty acid synthase deduced amino acid sequence is 80% identical to fatty acid synthase from

Ar. subalbatus and 65% identical to an undescribed expressed sequence tag from An. gambiae. In addition, our alignment shows 47% identity with the well-described fatty acid synthase from the chicken Gallus gallus and 37% identity with the chimpanzee Pan troglodytes. Our 3’ end also overlaps a portion of the thioesterase domain as described in

G. gallus (Holzer et al., 1989).

62 Confirmation of Diapause Up- and Downregulated Genes

Northern blot hybridizations confirmed the SSH results showing the downregulation of mRNA encoding trypsin and chymotrypsin-like and the upregulation of fatty acid synthase in early diapause (Figure 3.7). Our SSH comparison utilized mosquitoes reared under diapause-inducing (short daylength) and nondiapause–inducing

(long daylength) conditions at 18°C. The only environmental variable was daylength, and thus we can conclude that the distinctions we observed were in direct response to photoperiod. To further evaluate the role of temperature, we also used northern blots to compare nondiapausing mosquitoes reared at both 18°C and 25°C. The same results were observed at the two temperatures (Figure 3.7), thus the rearing temperature of the mosquitoes does not appear to be a primary environmental factor regulating the expression of fatty acid synthase, trypsin, and chymotrypsin-like serine protease.

Expression Patterns at the Onset of Diapause

The two blood digestive enzymes, trypsin and chymotrypsin-like, have identical patterns of mRNA expression in the early days following adult eclosion (Figure 3.8). In nondiapausing females, the transcripts were first detected by northern blot hybridization 2 days after adult eclosion, and a strong signal persisted from day 3 through to our final observation on day 7. This pattern is consistent with the onset of host-seeking behavior in our laboratory colony: females were fully ready to take a blood meal 2-3 days after adult eclosion. By contrast, neither trypsin nor chymotrypsin-like gene expression was detectable in the diapausing individuals at this early stage.

63 Fatty acid synthase, the gene encoding a key enzyme in the conversion of sugars

to fat, was expressed at a low level in nondiapausing females, and then only on days 2

and 3 after eclosion. By contrast, this gene was highly expressed in diapausing adults

beginning on day 3, and expression persisted throughout the remainder of the 7 day

observation period. In all six of the independent diapause replicates, the signal was

highest on day 4 and was reduced on day 5 and 6. This decrease in detectable message

varied in intensity, but it was consistently observed in all replicates on either day 5 or 6.

Thus, in diapausing females the onset of expression of fatty acid synthase occurred one

day later, but expression was higher and persisted longer than in nondiapausing females.

Expression Patterns throughout Diapause and at Diapause Termination

Expression of mRNAs encoding these three enzymes was also monitored

throughout diapause, beginning one week (7-10 days) post adult eclosion and then at 30-

day intervals thereafter for up to four months. Though trypsin was not expressed early in

diapause (Figure 3.8), a signal was evident by day 90 and persisted through day 120

(Figure 3.9). A similar pattern of expression was observed for chymotrypsin-like, but in

this case a weak signal was first noted on day 60. When diapause was broken at 2

months by transferring the females to long daylength and high temperature, both genes were highly expressed within one week (Figure 3.9).

Our northern blots showed that the mRNA encoding fatty acid synthase was highly expressed in diapausing females during the first week following adult eclosion

(Figure 3.8), but the expression was sporadic thereafter (Figure 3.9): the mRNA was

64 consistently undetectable on day 30, was strongly present on days 60 and 90, but was gone again on day 120. Expression was consistently high when diapause was broken.

DISCUSSION

Suppressive subtractive hybridization yielded three clones of potential interest for probing feeding responses in nondiapausing and diapausing individuals of Cx. pipiens.

Genes encoding two blood digestive enzymes, trypsin and chymotrypsin-like serine protease, are downregulated in early diapause, and fatty acid synthase, an enzyme involved in lipid sequestration, is highly upregulated at this time. We have confirmed these results by northern blot hybridization and have also demonstrated that the regulation of these genes is under photoperiodic control (short daylength) and not temperature. This is the first molecular evidence demonstrating that diapause-destined females are programmed to express a gene associated with the accumulation of lipid reserves and that these females have shut down the expression of genes associated with digestion of a blood meal.

We obtained the 3’ end of fatty acid synthase, which contains a deduced amino acid sequence that overlaps a portion of the thioesterase domain of fatty acid synthase in

G. gallus. As is typical of this gene, our clone also contains a long (807bp), 3’- untranslated region with a nucleotide sequence of low homology to the other known insect fatty acid synthase sequences. The translated region, however, is conserved

(>65%) among the three mosquito species examined. We have also identified full-length cDNA clones that encode trypsin and a chymotrypsin-like protein, two proteolytic blood

65 digestive enzymes in the class of serine proteases. Both deduced amino acid sequences contain the characteristic catalytic triad and the six cysteine residues typical of serine proteases (Wilcox, 1970). Our clones differ, however, in the residues involved in substrate specificity: the predicted active trypsin enzyme contains a negatively charged carboxylate (Asp 210) located at the bottom of the substrate binding pocket which is typical of trypsin-like serine proteases, while our second clone contains a hydrophobic substrate binding pocket (Gly-189) characteristic of chymotrypsin-like serine proteases.

Our trypsin clone also contains the three residues that make up the zymogen triad

(Ser/His/Asp) that is involved in stabilizing the inactive serine protease proenzyme

(Madison et al., 1993). However, our chymotrypsin-like serine protease contains only one of the three zymogen triad residues in the conserved location, a feature similar to a chymotrypsin-like protease from the human vector An. gambiae (Han et al.,

1997). Our first serine protease clone thus appears to be a trypsin-like serine protease, and our second clone is most similar to a chymotrypsin-like serine protease.

The expression of these three genes investigated by northern blot hybridization revealed distinct patterns of expression at the onset of diapause, during four months in diapause, as well as at diapause termination. During the first seven days of adult life in diapause-destined females, fatty acid synthase is more highly expressed than in nondiapausing individuals and the expression persists for a longer period. These results are consistent with the pattern of sugar feeding in Cx. pipiens (Bowen, 1992): diapause- destined females fed on sugar more readily and for a longer period of time than their nondiapausing counterparts during the first 15 days of adult life. Although fatty acid

66 synthase is undetectable one month into diapause, it is sporadically expressed thereafter until diapause has been broken. In contrast, the genes encoding the blood digestive enzymes trypsin and chymotrypsin-like are completely “shut down” at the onset of diapause and remain downregulated until mid to late diapause, when females are preparing for diapause break.

In nondiapausing females, the upregulation of trypsin and chymotrypsin-like 2-3 days after adult eclosion corresponds with the expression patterns of chymotrypsin (Jiang, et al., 1997) and early trypsin (Kalhok et al., 1993; Noriega et al., 1996) in nondiapausing individuals of Ae. aegypti. Since our mosquitoes were not blood-fed, we would expect the trypsin we observed to be most similar to early trypsin in Ae. aegypti because early trypsin mRNA is abundant prior to blood feeding (Kalhok et al., 1993). The translation of early trypsin upon blood feeding is essential in activating the transcription of late trypsin, the major midgut endoprotease (Noriega et al., 1996). Indeed, our trypsin aligns most closely with early trypsin in Ae. aegypti (51%) , compared to 31% for late trypsin

(31%). In addition, our chymotrypsin-like clone is present prior to blood-feeding and is likely to be involved in blood digestion. Jian et al. (1997) characterized a female-specific chymotrypsin from Ae. aegypti and demonstrated the accumulation of mRNA 24 h after adult eclosion with translation being induced following blood feeding. Unlike early trypsin, chymotrypsin remained highly active during blood meal digestion in Ae. aegypti.

Our results suggest that it is unlikely that diapause-destined Cx. pipiens females take a blood meal prior to entering hibernation, since the molecular machinery necessary to process the blood meal is not functional in these individuals. The absence of trypsin

67 and chymotrypsin-like mRNA in diapause-destined Cx. pipiens is likely a result of a lack

in juvenile hormone (JH) at this time, since the transcriptional regulation of early trypsin

is under the control of JH in Ae. aegypti (Noriega et al., 1997). In Ae. aegypti, abdominal

ligations 1 h post adult eclosion led to a complete inhibition of early trypsin transcription

(Noriega et al., 1997). Instead of blood feeding, prehibernating females feed on nectar

and other plant products early in diapause and have the ability to convert the

carbohydrates into extra lipid reserves, a feature that is essential for survival throughout

the long winter. After 7 days of feeding on sugar, diapausing females accumulate nearly

twice as many lipid reserves as nondiapausing females reared at the same temperature

(Mitchell and Briegel, 1989a).

Once females enter diapause, fatty acid synthase continues to be expressed in mid to late diapause, thus females are likely to be capable of processing a sugar meal throughout diapause. This raises the possibility that Cx. pipiens may search for and

utilize a readily available sugar meal during the winter. This is plausible; several reports

have noted that diapausing females are active even in mid-winter and often leave their

hibernaculum during this time (Berg and Lang, 1948; Service, 1968; Buffington, 1972;

Onyeka and Boreham, 1987). We have also observed mid-winter movement in our field

sites and have observed that our laboratory-reared females will readily take a sugar meal

when honey-soaked sponges are placed into their cages 2-3 months after the onset

diapause (Robich et al., unpublished observations). An occasional sugar meal in mid-

winter would enable females to replenish lipid reserves and may enhance survival.

68 As the end of diapause approaches, the accumulation of trypsin and chymotrypsin- like mRNA indicates that females are preparing for blood feeding and subsequent egg production. This pattern follows the gradual increase of JH observed in diapausing females; by the end of winter JH titers reach levels equivalent to those observed in nondiapausing females (Readio et al., 1999). It is evident from the patterns of gene expression that regulation of fatty acid synthase, trypsin, and chymotrypsin-like is a part of the diapause program, and not due to other factors such as temperature, host availability, or feeding activities. It has been suggested that an occasional warm spell

(Indian Summer) in the fall may lead to Cx. pipiens taking a blood meal (Eldridge and

Bailey, 1979). But, previous behavioral data (Mitchell, 1983; Bowen et al., 1988;

Bowen, 1992), together with our current molecular evidence, suggest that diapause- destined females not only lack the host-seeking response but they are unable to process a blood meal. Instead, such females exposed to short daylength are programmed to feed on sugar and garner lipid reserves. This metabolic switch is thus a component of the diapause program.

69

REFERENCES

Bailey, C.L., Faran, M.E., Gargan, T.P., and Hayes, D.E. 1982. Winter survival of blood-fed and nonblood-fed Culex pipiens L. American Journal of Tropical Medicine and Hygiene 31:1054-1061.

Bellamy, R.E., Reeves, W.C., and Scrivani, R.P. 1958. Relationships of mosquito vectors to winter survival of encephalitis viruses. II. Under experimental conditions. American Journal of Hygiene 67:90-100.

Berg, M., and Lang, S. 1948. Observations of hibernating mosquitoes in Massachusetts. Mosquito News 8:70-71.

Bowen, M.F. 1990. Post-diapause sensory responsiveness in Culex pipiens. Journal of Insect Physiology 36:923-929.

Bowen, M.F. 1992. Patterns of sugar feeding in diapausing and nondiapausing Culex pipiens (Dipetera: Culicidae) females. Journal of Medical Entomology 29:843-849.

Bowen, M.F., Davis, E.E., and Haggar, D.A. 1988. A behavioral and sensory analysis of host-seeking behavior in the diapausing mosquito Culex pipiens. Journal of Insect Physiology 34:805-813.

Buffington, J.D. 1972. Hibernaculum choice in Culex pipiens. Journal of Medical Entomology 9:128-132.

Chomczynski, P., and Sacchi, N. 1987. Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction. Analytical Biochemistry 162:156-159.

Christophers, S.R., 1911. The development of the egg follicle in Anopheles. Paludism 2:73-88.

Denlinger, D.L. 1985. Hormonal control of diapause. In: Kerkut, G.A., Gilbert, L.I. (Eds.), Comprehensive insect physiology, biochemistry and pharmacology, Vol. 8. Pergamon Press, Oxford, pp. 353-412.

Denlinger, D.L. 2002. Regulation of diapause. Annual Review of Entomology 47:93- 122.

Eldridge, B.F. 1966. Environmental control of ovarian development in mosquitoes of the Culex pipiens complex. Science 151:826-828.

70

Eldridge, B.F., and Bailey, C.L. 1979. Experimental hibernation studies on Culex pipiens (Diptera: Culicidae): reactivation of ovarian development and blood-feeding in prehibernating females. Journal of Medical Entomology 15:462-467.

Han, Y.S., Salazar, C.E., Reese-Stardy, S.R., Cornel, A., Gorman, M.J., Collins, F.H., and Paskewitz, S.M. 1997. Cloning and characterization of a serine protease from the human malaria vector, Anopheles gambiae. Insect Molecular Biology 6:385-395.

Holzer, K.P., Liu, W., and Hammes, G.G. 1989. Molecular cloning and sequencing of chicken liver fatty acid synthase cDNA. Proceedings of the National Academy of Sciences of the United States of America 86:4387-4391.

Jiang, Q., Hall, M., Noriega, F.G., and Wells, M. cDNA cloning and pattern of expression of an adult, female-specific chymotrypsin from Aedes aegypti midgut. Insect Biochemistry and Molecular Biology 27:283-289.

Kalhok, S.E., Tabak, L.M., Prosser, D.E., Brook, W., Downe, A.E., and White, B.N. 1993. Isolation, sequencing and characterization of two cDNA clones coding for trypsin- like enzymes from the midgut of Aedes aegypti. Insect Molecular Biology 2:71-79.

Kraut, J. 1977. Serine proteases: structure and mechanism of catalysis. Annual Review of Biochemistry 46:331-358.

Reisen, W.K., Kramer, L.D., Chiles, R.E., Wolfe, T.M., and Green, E.N. 2002. Simulated overwintering of encephalitis viruses in dipausing female Culex tarsalis (Diptera: Culicidae). Journal of Medical Entomology 39:226-233.

Spielman, A., and Wong, J. 1973b. Environmental control of ovarian diapause in Culex pipiens. Annals of the Entomological Society of America 66:905-907.

Madison, E.L., Kobe, A., Gething, M.J., Sambrook, J.F., Goldsmith, E.J. 1993. Converting tissue plasminogen activator to a zymogen: a regulatory triad of Asp-His-Ser. Science 262:419-421.

Madison, E.L., Kobe, A., Gething, M-J, Sambrook, J.F., and Goldsmith, E.J. 1993. Converting tissue plasminogen activator to a zymogen: A regulatory triad of Asp-His- Ser. Science 262:419-421.

McGeoch, D.J. 1985. On the predictive recognition of signal peptide sequences. Virus Research 3:271-286.

71 Mitchell, C.J. 1983. Differentiation of host-seeking behavior from blood-feeding behavior in overwintering Culex pipiens (Diptera: Culicidae) and observations on gonotrophic dissociation. Journal of Medical Entomology 20:157-163.

Mitchell, C.J., and Briegel, H. 1989a. Inability of diapausing Culex pipiens (Diptera:Culicidae) to use blood for producing lipid reserves for overwinter survival. Journal of Medical Entomology 26:318-326.

Mitchell, C.J., and Briegel, H. 1989b. Fate of the blood meal in force-fed, diapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 26:332-341.

Muller, H-M, Catteruccia, F., Vizioli, J., Della Torre, A., and Crisanti, A. 1995. Constitutive and blood meal-induced trypsin genes in Anopheles gambiae. Experimental Parasitology 81:371-385.

Nakai, K., and Kanehisa, M. 1992. A knowledge base for predicting protein localization in eukaryotic cells. Genomics 14:897-911.

Nasci, R.S., Savage, H.M., White, D.J., Miller, J.R., Cropp, B.C., Godsey, M.S., Kerst, A.J., Bennett, P., Gottfried, K., and Lanciotti, R.S. 2001. West Nile virus in overwintering Culex mosquitoes, New York City, 2000. Emerging Infectious Diseases 7:742-744.

Noriega, F.G., Wang, X., Pennington, J.E., Barillas-Mury, C.V., and Wells, M.A. 1996. Early trypsin, a female-specific midgut protease in Aedes aegypti: Isolation, amino- terminal sequence determination, and cloning and sequencing of the gene. Insect Molecular Biology 26:119-126.

Onyeka, J.O.A., and Boreham, P.F.L. 1987. Population studies, physiological state and mortality factors of overwintering adult populations of females of Culex pipiens L. (Diptera: Culicidae). Bulletin of Entomological Research 77:99-111.

Readio, J., Chen, M., and Meola, R. 1999. Juvenile hormone biosynthesis in diapausing and nondiapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 36:355-360.

Sanburg, L.L., and Larsen, J.R. 1973. Effect of photoperiod and temperature on ovarian development in Culex pipiens pipiens. Journal of Insect Physiology 19:1173-1190.

Service, M.W. 1968. Observations on the ecology of some British mosquitoes. Bulletin of Emtomological Research 59:161-194.

Spielman, A. 1974. Effect of synthetic juvenile hormone on ovarian diapause of Culex pipiens mosquitoes. Journal of Medical Entomology 11:223-225.

72

Spielman, A., and Wong, J. 1973. Environmental control of ovarian diapause in Culex pipiens. Annals of the Entomological Society of America 66:905-907.

Swellengrebel, N.H. 1929. La dissociation des fonctions sexuelles et nutritives (dissociation gono-trophique) d’ Anopheles maculipennis comme cause du paludisme dans les Pays-Bas et ses rapports avec “l’infection domiciliaire”. Annales de l'Institut Pasteur 43:370-1389.

Vinogradova, E.B. 2000. Culex pipiens pipiens mosquitoes: taxonomy, distribution, ecology, physiology, genetics, applied importance and control. Pensoft Publishers, Sofia. pp. 46-115.

Von Heijne, G. 1986. A new method for predicting signal sequence cleavage sites. Nucleic Acid Research 14:4683-4690.

Walsh, K.A., and Wilcox, P.E. 1970. Serine proteases. Methods in Enzymology 19:64- 108.

Warshel, A., Naray-Szabo, G., Sussman, F., and Hwang, J.K. 1989. How do serine proteases really work? Biochemistry 28:3629-3637.

Washino, R.K. 1977. The physiological ecology of gonotrophic dissociation and related phenomena in mosquitoes. Journal of Medical Entomology 13:381-388.

Watts, D.M., Thompson, W.H., Yuill, T.M., DeFoliart, G.R., and Hanson, R.P. 1974. Overwintering of La Crosse virus in Aedes triseriatus. American Journal of Tropical Medicine and Hygiene 23:694-700.

Wilcox, P.E. 1970. The serine proteases. Methods in Enymology 19:64-108.

Yan, Y.S., Salazar, C.E., Reese-Stardy, S.R., Cornel, A., Gorman, M.J., Collins, F.H., and Paskewitz, S.M. 1997. Cloning and characterization of a serine protease from the human malaria vector, Anopheles gambiae. Insect Molecular Biology 6:385-395.

73

ATCATTCGATCAACTTCCTCTCGTCCGGTAAACCGAATCCAACGCA 46

ATG GCC AAG TTA TTA GTG TTG ACC ACC TGT GCC CTC CTG GGC TTA 91 M A K L L V L T T C A L L G L 15

ACA TCT GGT GCC TCC CTC AAG TCC ACC TTG ATG CCG AGC TTT TCT 136 T S G A S L K S T L M P S F S 30

CGC GCA GGC AAA ATC GTC GGA GGG TTC CAG ATC GAC GTC GTC GAC 181 R A G K I V G G F Q I D V V D 45

GTC CCG TAC CAG GTG TCG CTG CAG CGC AAT AAC CGT CAC CAC TGC 226 V P Y Q V S L Q R N N R H H C 60

GGC GGA TCG ATT ATC GAC GAG AGA TGG GTG CTG ACG GCG GCC CAC 271 G G S I I D E R W V L T A A H 75

TGT ACG GAG AAT ACC GAC GCC GGT ATC TAC AGT GTG CGC GTC GGT 316 C T E N T D A G I Y S V R V G 90

TCG TCG GAA CAC GCC ACC GGA GGG CAG CTG GTC CCG GTG AAG ACC 361 S S E H A T G G Q L V P V K T 105

GTT CAC AAC CAT CCG GAC TAT GAT CGC GAG GTC ACC GAG TTT GAC 406 V H N H P D Y D R E V T E F D 120

TTT TGC TTG CTG GAG TTG GGC GAG CGT TTG GAG TTT GGC CAC GCC 451 F C L L E L G E R L E F G H A 135

GTT CAA CCG GTT GAC CTG GTT CGG GAC GAA CCG GCT GAC GAG AGT 496 V Q P V D L V R D E P A D E S 150

CAG TCG CTG GTT TCC GGC TGG GGA GAC ACG AGA TCG CTG GAG GAA 541 Q S L V S G W G D T R S L E E 165

TCC ACC GAT GTC CTG AGG GGT GTT TTA GTG CCG TTG GTG AAC CGC 586 S T D V L R G V L V P L V N R 180

GAG GAA TGT GCC GAA GCT TAC CAG AAG CTT GGT ATG CCG GTT ACG 631 E E C A E A Y Q K L G M P V T 195

GAG AGC ATG ATC TGC GCT GGA TTC GCC AAG GAA GGA GGC AAG GAC 676 E S M I C A G F A K E G G K D 210

GCC TGC CAA GGA GAC AGC GGT GGT CCC CTG GTC GTG GAC GGT CAA 721 A C Q G D S G G P L V V D G Q 225

CTG GCT GGA GTA GTT TCT TGG GGA AAG GGT TGC GCT GAA CCT GGA 766 L A G V V S W G K G C A E P G 240

TTT CCG GGA ATT TAC TCC AAC GTA GCG TAC GTC CGC GAT TGG ATC 811 F P G I Y S N V A Y V R D W I 255

AAA AAG GTG GCC AAG GTT TAAAAACCCAATAAAGTCTTAAAATTTAAAAAAAA 864 K K V A K V * 261

AAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAA 898

Figure 3.1. Nucleotide and deduced amino acid sequences of Cx. pipiens trypsin cDNA (GenBank accession no. AY958426). The predicted active enzyme is shaded. Primers used in RACE are shown with arrows. The polyadenylation signal is boxed, and an asterisk (*) represents the stop codon.

74

CACTATCGTTCCCCATCGAAGTCGGTTCAGTACCTGCCCGGATCATC 47

ATG AAT AAG GTC GCG ATC GTT TGC CTC CTT CTA GCG ACC CTT GTT 92 M N K V A I V C L L L A T L V 15

GGT CCT AGT TTG GCC CGT CGG ATC TTT GGA GGG CAG TTC GCA GAG 137 G P S L A R R I F G G Q F A E 30

GAG CGC CAG TTT CCC TAC CAG GTG GCG CTC TTC CAC AAC GGA CAC 182 E R Q F P Y Q V A L F H N G H 45

TTT GAC TGC GGG GGA TCG ATC ATC GAC AAC CGG TGG ATT TTC ACG 227 F D C G G S I I D N R W I F T 60

GCT GCG CAT TGC GTT CTG GAA CTG AAT GGA TCG GTT GCC ACG AAC 272 A A H C V L E L N G S V A T N 75

CTC TCC GTC TTA GTC GGT TCC CAG CAC CTG GTT GAG GGT GGC CGT 317 L S V L V G S Q H L V E G G R 90

CGC TTC GAG CCG GAA GCA ATC TTC GCG CAC GAA TCG TAC GGA AAC 362 R F E P E A I F A H E S Y G N 105

TTC CAG AAC GAT ATC GCG CTG ATC AAG CTG GGT GAG TCG ATC GAG 407 F Q N D I A L I K L G E S I E 120

TAC GAC GAG CAG AGC CAA CCG ATC GCG CTG TAC GAG GGC GAT GAT 452 Y D E Q S Q P I A L Y E G D D 135

CTG CCC AAG GAC TCC GTG GTG GTC ATT TCC GGA CAC GGT CGA ACC 497 L P K D S V V V I S G H G R T 150

GAG GAT CAT GAC TTC TCC GAG CTG CTC AAG TTC AAC CGG ATG TTG 542 E D H D F S E L L K F N R M L 165

GTG GAC ACG CAG GAG TCT TGC GGG AAG GAC CGG GAG GGG CTG ATT 587 V D T Q E S C G K D R E G L I 180

TGC TTC AAC GAA AAG GTT GGG AAC GGC GCT TGC CAC GGG GAT TCC 632 C F N E K V G N G A C H G D S 195

GGT GGT CCG GCG GTG TTC GAG GGA CGC CAG GTT GGG GTG GCC AAT 677 G G P A V F E G R Q V G V A N 210

TTT GTG CAG GGA TCT TGT GGG TCG AAG TTT GCG GAT GGT TAT GCC 722 F V Q G S C G S K F A D G Y A 225

AAG GTG ACG CAC TAC CGG GAG TGG ATC GAT AGG ACG AAG AGA AAT 767 K V T H Y R E W I D R T K R N 240

TAAGAGTTGTGATAGCAAAATGAATGTCTGGGAATGATCGTGAAATAAATGTCTGCTCA 826 * ACGAAACACGCCATGGCGGAAGGCAATTATTATCGAAAAAAAAAAAAAAAAAAAA 881

Figure 3.2. Nucleotide and deduced amino acid sequences of Cx. pipiens chymotrypsin- like serine protease cDNA (GenBank accession no. AY958427). The predicted active enzyme is shaded. Primers used in RACE are shown with arrows. The polyadenylation signal is boxed, and an asterisk (*) represents the stop codon.

75

AAT TAC GCC AAA CTG CAA GGC GAC TAC GGT CTT TCA GAG CTC TGC 45 N Y A K L Q G D Y G L S E L C 15

ACC AAG GAG GTT AAG GTG ACC ACC GTC AAG GGT GAC CAT CGG TCA 90 T K E V K V T T V K G D H R S 30

ATC CTG ACG GGT GAA TCG ATG CAG AAC ATT TCC AAA CTG CTG CTG 135 I L T G E S M Q N I S K L L L 45

GAA ACG AAT TAAGGGACCATGCAGTCAGTCCGGTGCATCCTGTACAAAGACACACA 191 E T N * 48

AAAACCACAAACAGAAAATCAAACAATGAACGTGAAAAGTCTACATCGGCCGCATTTAC 250

AGGCAGGCGAGAAAGAAGCATCGTAATTTATTCATTTTTCTCCGCGATTTTTTTTTTGT 309

TTATAGTAGTTTAAATTTGCTTCCTCTTTTTTTTATACATTTTTTTGTATAAGAATCTT 368

AAAACCTACTGTTCGAGAAGCCAACACATAATTGGTCATACTAATATGTCAGCACGCGC 427

GCGTGATGGTAGCAAGATCAGATCTTACTATGAATCGTCAGCACTCAGCGAAGGAACCT 486

AGCCGCAAAGAGAACTGCCACTAATATCTGGGAGCCCAAGTCCAAATATCCCAGGGATA 545

GTGCCACTTTTCTTTGCTTTGCAGCGTTCTTCAAAAACACTGAAAACATGTGATGTTTG 604

TAACGTTTATAACCGTTTCTCTTGCGTAGCAATTAGTTGCCTTATTTCTGTGCGAACAA 663

AAAAAACGGCTTCGAAAAACAAAAACAAAAAAAAACATACGTTCCAACTTTAGAGTAAC 722

GCTGTGTTATCTCTCTTGCTAATTGATATGATTAGGTAGCTAGTGATAGATAAGTTTTA 781

GTCAGCAACAACAAACGGGGGATTGGGATCCTTTCAAACACTGAACTGTGTGTACTGTG 840

TAGTGTGGTTAGCACGCTTTAGAGCAGAAGGAAGAAAACAAATATTTTTACATCACAAA 899

ACACAAAATAAATAAATAAATAAAAATAATAAAACAAGCAAAAAAAAAAAAAAAA 954

Figure 3.3. Nucleotide and deduced amino acid sequences of Cx. pipiens fatty acid synthase cDNA (GenBank accession no. AY958428). The primer used in 3’ RACE is shown with an arrow. The polyadenylation signal is boxed, and an asterisk (*) represents the stop codon.

76 ▼ ▼ CpiTry 1 ------MAKLLVLTTCALLGLTSGASLKS------TLMPSFSRAGKIVGGFQIDVV CpaTry 1 ------MAKLLVLTTCALLGLTSGASLKS------TLMPSFSRAGKIVGGFQIDVV AaeETry 1 ------MNQFLFVSFCALLGLS------QVSAATLSSGRIVGGFQIDIA AgaTry 1 MSNKIAILLLAVLVAVVACAQAQPSGRHHLVHPLLPRFLPRLHRDSNGHRVVGGFQIDVS DmeTryL 1 ------MRSSIGLTGMAKTILHLFIGGIPPGKSELRSHCKAPTLDGRIVGGQVANIK AaeLTryP 1 ------MFTSTVVFASLMALAS------AFPSLDNGRVVNGQTATLG

○ ○ C1 *C1 CpiTry 45 DVPYQVSLQRN---NRHHCGGSIIDERWVLTAAHCTENTDAGIYSVRVGSSEHAT--GGQ CpaTry 45 DVPYQVSLQRN---NRHHCGGSIIDERWVLTAAHCTENTDAGIYSVRVGSSVHAT--GGQ AaeETry 38 EVPHQVSLQRS---GRHFCGGSIISPRWVLTAAHCTTNTDPAAYTIRAGSTDRTN--GGI AgaTry 61 DAPYQVSLQYF---NSHRCGGSVLDNKWVLTAAHCTQGLDPSSLAVRLGSSEHAT--GGT DmeTryL 52 DIPYQVSLQR----TYHFCGGSLIAQGWVLTAAHCTEGSAILLSKVRIGSSRTSV--GGQ AaeLTryP 36 QFPFQVLLKVELSQGRALCGGSLLSDQWVLTAGHCTDGAKSFEVTLGAVDFEDTTNDGRV

* CpiTry 100 LVPVKTVHNHPDYDREVTEFDFCLLELGERLEFGHAVQPVDLVRDEP--ADESQSLVSGW CpaTry 100 LVPVKAVHNHPDYDREVTEFDFCLLELGERLEFGHAVQPVDLVRDEP--ADESQSLVSGW AaeETry 93 IVKVKSVIPHPQYNGDTYNYDFSLLELDESIGFSRSIEAIALPEASETVADGAMCTVSGW AgaTry 116 LVGVLRTVEHPQYDGNTIDFDFSLMELETELTFSDLVQPVELPEHEEPVEPGTMATVSGW DmeTryL 106 LVGIKRVHRHPKFDAYTIDFDFSLLELEEYSAKNVTQAFVGLPEQDADISDGTPVLVSGW AaeLTryP 96 VLTATEYHRHEKYNPLFATNDVAVVKLPTPVAFNDRVQPVKLPTGSDT-FTDREVVVSGW

C2 C2 ↓ C3 ○* CpiTry 158 GDTRSLEESTDVLRGVLVPLVNREECAEAYQKLGMPVTESMICAGFAKEGGKDACQGDSG CpaTry 158 GDTRSLEESTDILRGVLVPLVNREECAEAYQKLGMPVTESMICAGFAKEGVKDACQGDSG AaeETry 153 GDTKNVFEMNTLLRAVNVPSYNQAECAAALVNV-VPVTEQMICAGYAAGG-KDSCQGDSG AgaTry 176 GNTQSAVESSDFLRAANVPTVSHEDCSDAYMWF-GEITDRMLCAGYQQGG-KDACQGDSG DmeTryL 166 GNTQSAQETSAVLRSVTVPKVSQTQCTEAYGNF-GSITDRMLCVITEGGK--DACQGDSG AaeLTryP 155 GLQKNGGNVADKLQYAPLTVISNNECSKAYSPL--VIKKTTLCAKGENKE--SPCQGDSG

C3 CpiTry 218 GPLVVDG--QLAGVVSWGKGCAEPG-FPGIYSNVAYVRDWIKKVAKV CpaTry 218 GPLVVDG--QLAGVVSWGKGCAEPG-YPGIYSNVVYVRGWIKKVAKV AaeETry 211 GPLVFGD--ELVGVVSWGKGCALPN-LPGVYARVSTVRQWIREVSEV AgaTry 234 GPLVADG--KLVGVVSWGYGCAQPG-YPGVYGRVASVRDWVRENSGV DmeTryL 223 GPLAADG--VLWGVVSWGYGCARPN-YPGVYSRVSAVRDWISSVSGI AaeLTryP 211 GPLVLEGENVQVGVVSFGHAVGCEQGYPGAFARLTSFVDWIKQKTGL

Figure 3.4. Multiple sequence alignment of the deduced Cx. pipiens trypsin with other insect trypsins retrieved from GenBank. Amino acids identical to Cx. pipiens are shaded. The predicted cleavage sites of the signal peptide and the putative activation peptide are denoted with a triangle (▼). The conserved three pairs of cysteines are in bold and labeled C1-C3. The residues of the catalytic triad (His/Asp/Ser) are denoted by an asterisk (*), and the Asp residue characteristic of trypsin-like serine proteases is marked with an arrow (↓). Open circles (○) denote the three residues that make up the zymogen triad (Ser/His/Asp). CpiTry: Cx. pipiens trypsin, AY958426, CpaTry: Cx. pipiens pallens trypsin, AAK67462; AaeETry: Ae. aegypti early trypsin, AAM34268, AgaTry: An. gambiae trypsin, CAA80518; DmeTryL: D. melanogaster trypsin-like protease, AAC47304; AaeLTryP: Ae. aegypti late trypsin precursor, AF266757. 77 .▼ ▼ CpiChyL 1 ------MNKVAIVCLLLATLVGPSLAR------RIFGGQFAEER AgaUnk 1 ------RRSSVVVGLVLVAFLGTVLSVP--IW------NRIVGGQLAEDT AgaSer 1 -MTLADRVPLALAALAYLALVSGVRFHLSEQNDVLPGGSQARRPFFQGARIVGGSVASEG AdaChy1 1 -----MVRGITVLAAVCLMVGANNIPKLVPDDHYV------NRVVGGQEAEEG AdaChy2 1 ------MKAIITVLAVISAIVDAQCKVPSRQH------NRVVGGQDAEES

C1 *C1 CpiChyL 33 QFPYQVALFHNG-HFDCGGSIIDNRWIFTAAHCVLELNGSV-ATNLSVLVGSQHLVEGGR AgaUnk 46 QMPYQIALFYQG-SFRCGGSIIGDRHVLTAAHCVMDDDVLLPAFKFGVHAGSAHLNAGGK AgaSer 60 QFPHQVALLRGN-ALTCGGSLIESRWVLTAAHCVYNGALVVPASSIVVVAGSVSLSNG-V AdaChy1 43 SAPYQVSLQVALWGHNCGGSILSERWVLTAAHCLVGTD----AEELEVLVGTNSLKEGGQ AdaChy2 39 SAPYQISLQLADRGHFCGGSILNERWILTAAHCIKEID----AADLEVLAGTNNLQEGGQ

* CpiChyL 91 RFEPEAIFAHESYGN--FQNDIALIKLGESIEYDEQSQPIALYEGDDLPKDSVVVISGHG AgaUnk 105 LFKVRAVYPHEGYGN--FQHDIAVMEMKEPFAFDKYIQPIELMDE-EVPLGGEVVISGYG AgaSer 118 RRAVARVITHERYGN--FKNDVALLQLQLSLPSSAYIRPIALRTS-SVPAGSEVVISGWG AdaChy1 99 RYKADKLLYHSRYNSPQFHNDIGLVRLATPIKFSSTVKSIEYSEN-VVPVNATVRLTGWG AdaChy2 95 RYRVDRLFSHSRYNRPQFHNDIALVHLAAPIRFSSKIKSIEYSEQ-ALPANVTVRLTGWG

C2 C2 ↓ C3 * CpiChyL 149 RTEDHDFS-ELLKFNRMLVDTQESCGKDREG------LICFNEKVGNGACHGDSGGPA AgaUnk 162 RVGSNGPVSPALLYTSMFVVEDENCNSISEG------LMCIDKEGSYGACNGDSGGPA AgaSer 175 VCTKVAPYQTCFDTTVLPVVADQQCRDPTGISTG-----LICFTSPVNNGACNGDSGGPA AdaChy1 158 RTSAGGSVPTKLQTIDLRTLSNEDCKKKSGN-PGNVDIGHVCTLTRTGEGACNGDSGGPL AdaChy2 154 MLDVWGPSPTQLQTIDLRTLTNKDCKAKLMLNPHNVDIGHVCTLTKKGEGACNGDSGGPL

C3 CpiChyL 200 VFEGRQVGVANFVQGSCGSKFADGYAKVTHYREWIDRTKRN-- AgaUnk 214 VYDGKLAGVANFIIDQCGGNFADGYAKVSFYLDWIRQFLE--- AgaSer 230 ILNNQLVGRPNFIINYCGSASPDGYAKVSDFVTWIQTTMRRY- AdaChy1 217 VYEDKVIGVVNFGVP-CALGYPDGFARVSYYHDWIRTTIRNN- AdaChy2 214 VFGNKLVGVVNFGMP-CATGMPDMFARVSYYHDWIRTTIANNS

Figure 3.5. Multiple sequence alignment of the deduced Cx. pipiens chymotrypsin-like serine protease with other insect trypsins retrieved from GenBank. Amino acids identical to Cx. pipiens are shaded. The predicted cleavage sites of the signal peptide and the putative activation peptide are denoted with a triangle (▼). The conserved three pairs of cysteines are in bold and labeled C1-C3. The residues of the catalytic triad (His/Asp/Ser) are denoted by an asterisk (*), and the Gly residue characteristic of chmotrypsin-like serine proteases is marked with an arrow (↓). CpiChyL: Cx. pipiens chymotrypsin-like serine protease, AY958427, AgaUnk: An. gambiae unknown protein, EAA09456; AgaSer: An. gambiae serine protease, AAA73920; AdaChy1: An. darlingi chymotrypsin 1, AAD17493; Adachy2: An. darlingi chymotrypsin 2, AAD17494.

78 CpiFas 1 NYAKLQGDYGLSELCTKEVKVTTVKGDHRSILTGESMQNISKLLLETN------AsuFas 1 NYAKLQGDYGLSELCTKDVKVTTVKGDHRSIXVGDSMLQISSILHELL------AgaUnk 1 NYAKLQGDYGLSDLCQQKVELFTVEGDHRSMLLGDSMKKISDVLQK------GgaFas 1 YEEGLGGDYRLSEVCDGKVSVHIIEGDHRTLLEGDGVESIIGIIHGSLAEPRVSVREG PtrPFas 1 YGEDLGADYNLSQVCDGKVSVHVIEGDHRTLLEGSGLESIVSIIHSSLAEPRVSVREG

Figure 3.6. Multiple sequence alignment of the deduced Cx. pipiens fatty acid synthase with other insect fatty acid synthases retrieved from GenBank. Amino acids identical to Cx. pipiens are shaded. CpiFas: Cx. pipiens fatty acid synthase, AY958428; AsuFas: Ar. subalbatus fatty acid synthase, AY441061; Agaunk: An. gambiae str. PEST, EAA15087; GgaFas: Gallus gallus fatty acid synthase, AAA48767; PtrPFas: P. troglodytes predicted fatty acid synthase, XP_511758.

79 A. trypsin C. fatty acid synthase 25°C 18°C 18°C 25°C 18°C 18°C nondiapause nondiapause diapause nondiapause nondiapause diapause

28S 28S

B. serine protease 25°C 18°C 18°C nondiapause nondiapause diapause

28S

Figure 3.7. Northern blot hybridization of diapause-regulated genes involved in blood meal vs. sugar meal digestion in Cx. pipiens. These northern hybridizations confirm the SSH results showing early diapause downregulation of trypsin and chymotrypsin-like serine protease and upregulation of fatty acid synthase. The results further indicate that it is the diapause-inducing photoregime (short daylength) rather than temperature that elicits the distinction. Each lane contains 15 µg of total RNA pooled from 20 females. Equal loading was confirmed by Northern blot hybridization with a 28S cDNA probe.

80 A. trypsin C. fatty acid synthase nondiapause nondiapause

28S diapause diapause

28S P 1 2 3 4 5 6 7 P 1 2 3 4 5 6 7 days after adult eclosion days after adult eclosion

B. chymotrypsin-like nondiapause

28S diapause

28S P 1 2 3 4 5 6 7 days after adult eclosion

Figure 3.8. Temporal pattern of expression of the genes encoding the digestive enzymes trypsin, chymotrypsin-like, and fatty acid synthase in late pupae (P) and during the first 7 days post adult eclosion in nondiapausing and diapause-destined females reared at 18°C. Each lane contains 15 µg of RNA isolated from pools of 20 females. Each membrane was stripped and re-probed with dig-labeled 28S cDNA to confirm equal loading.

81

A. trypsin C. fatty acid synthase

28S ND 10 30 60 90 120 break ND 10 30 60 90 120 break days in diapause days in diapause

B. chymotrypsin-like

28S ND 10 30 60 90 120 break days in diapause

Figure 3.9. Expression of trypsin, chymotrypsin-like, and fatty acid synthase throughout diapause (short daylength; 18°C), and when diapause is broken at two months. ND represents 10 day-old females reared under nondiapausing (short daylength; 18°C) conditions. Each lane contains 15 µg of RNA isolated from pools of 20 females. A 28S cDNA probe was used to confirm equal loading.

82

CHAPTER 4

Downregulation of mitochondrial mRNA expression, but not mitochondrial number,

during adult diapause in the northern house mosquito, Culex pipiens.

ABSTRACT

Two genes encoding the mitochondrial respiratory enzymes, cytochrome c

oxidase subunit I (COI) and cytochrome c oxidase subunit III (COIII) were identified as

being diapause regulated in Culex pipiens, using suppressive subtractive hybridization

(SSH). Two SSH clones were used to obtain a large portion of the coding region of COI

comprising 1,541 bp of the cDNA sequence that encodes a 504 residue deduced amino

acid sequence. A third SSH clone yielded the full-length sequence of COIII, which is

812 bp with a 262 deduced amino acid sequence. Northern blot analysis shows the

upregulation of COI and COIII in Cx. pipiens preparing for diapause, and transcript levels decline in mid-diapause and again return to high levels in late diapause, just prior

to diapause break. High levels of mRNA expression in early and late diapause likely

reflect the high energy requirements for diapause preparation and preparation for

resumption of reproduction. There are no differences in mtDNA levels between

nondiapausing and diapausing females, suggesting that mitochondrial numbers are not

reduced during diapause, and regulation of CO transcripts under transcriptional control.

83 INTRODUCTION

Most insects enter an overwintering dormancy (diapause) that is characterized by

a depressed metabolic rate. This “shut-down” in development allows the organism to

withstand long periods of adverse environmental conditions by entering a state that requires significantly less energy for maintenance. The extent of metabolic depression in diapause varies greatly depending on the developmental stage (Danks, 1987). Although metabolic suppression is especially dramatic for pupal diapause, in which case the metabolic rates may be 10-20% of the rates in comparable nondiapausing stages, the level of suppression is usually not quite so extreme in adult diapause where rates may be as high as 65% of the nondiapause rates (Danks, 1987).

Since mitochondria serve an essential role in aerobic oxidation, several investigators have investigated mitochondria during diapause by measuring enzyme activity, quantifying mitochondrial DNA (mtDNA), and examining mRNA expression.

Reductions in mitochondrial enzyme activity (Joanisse and Storey, 1994) and mRNA expression (Uno et al., 2004) have been noted in dormant insects. For example, in the larval diapause of the goldenrod gall fly Eurosta solidaginis and the gall moth Epiblema scudderiana the activities of three mitochondrial enzymes (citrate synthase, glutamate dehydrogenase, and NAD-isocitrate dehydrogenase) decreased by 50% during the winter

(Joanisse and Storey, 1994). Joanisse and Storey (1994) suggested that the reduction in enzyme activity may be the result of an overall decrease in mitochondrial numbers as an adaptation to the reduced energy demands during diapause. A later study, however, demonstrated an increase in the mitochondrial enzyme involved in ,

84 β-hydroxybutyrate dehydrogenase, in diapausing E. solidaginis larvae, suggesting that the decline previously noted in the other three enzymes may have been due to enzyme degradation (Joanisse and Storey, 1996).

In some insects, mitochondrial DNA decreases during cold adaptation, a feature often associated with diapause (Denlinger, 1991). In the freeze-tolerant high Arctic wooly caterpillar, Gynaephora groenlandica, the number of mitochondria in the

brain cells and fat body decrease in parallel with low respiration and cold adaptation

(Kukal et al., 1989), and Levin et al. (2003) showed similar results in cold-adapted E.

solidaginis larvae, despite the observed increase in β-hydroxybutyrate dehydrogenase

(Joanisse and Storey, 1996). Returning E. solidaginis larvae to high temperatures elicits

a rapid increase in mtDNA (<5 hrs) and respiration, indicating the capacity of this insect

to quickly change mitochondrial numbers (Levin et al., 2003).

In this study, we examine females of the northern house mosquito, Culex pipiens

(L.), a species that enters a reproductive diapause in late summer and early fall in

response to low temperature and short daylength (Eldridge, 1966; Sandburg and Larsen,

1973; Spielman and Wong, 1973). The diapause is characterized by an arrest in ovarian

development, a shut-down in blood feeding, a boost in sugar feeding, and a search for a

protective site for overwintering (Eldridge, 1966; Spielman and Wong, 1973; Sanburg

and Larsen, 1973; Mitchell, 1983; Bowen, 1992; Vinogradova, 2000). Although the

metabolic rate has not been reported previously for Cx. pipiens that are in diapause,

suppression of metabolic rate is well documented for other diapausing species (Danks,

1987).

85 We isolated and sequenced two genes encoding the mitochondrial respiratory

enzymes cytochrome c oxidase subunit I (COI) and cytochrome c oxidase subunit III

(COIII). To investigate the molecular events underlying the suppressed metabolic rates observed during diapause in Cx. pipiens, we used these two clones to probe mRNA levels at the onset of diapause and for 3 months into diapause, and we contrast these levels with those observed in nondiapausing adult females. We have also examined mitochondrial numbers throughout diapause by measuring mtDNA by dot blot hybridization. We conclude that mtDNA levels are unchanged during the course of diapause, while the mRNAs encoding the two respiratory enzymes, COI and COIII, are downregulated once diapause has been initiated.

MATERIALS AND METHODS

Insect Rearing

Mosquitoes were obtained from an anautogenous colony of Cx. pipiens L. that was established in September, 2000, from larvae collected in Columbus, Ohio (Buckeye strain). Colony mosquitoes were maintained at 25°C, 75% r.h., with a 15L:9D daily light:dark cycle. Larvae were reared in 18 x 28 x 5 cm plastic containers in de- chlorinated tap water, fed a diet of ground fish food (TetraMin), and maintained at a density of ~250 mosquitoes per pan. Adults were provided with water and honey-soaked sponges and kept in 30.5 x 30.5 x 30.5 cm screened cages.

Three experimental rearing groups were created by moving second instar larvae to one of three environmental regimes. Larvae were either kept in the colony rearing room

86 (nondiapause, 25°C), moved to an environmental chamber at 18°C, 75% r.h. and 15L:9D

(nondiapause, 18°C), or placed in an environmental room under diapause-inducing

conditions of 18°C, 75% r.h., with a 9L:15D daily light:dark cycle (diapause, 18°C). To

mimic the absence of sugar during the winter, honey sponges were removed from short- day cages 10-13 days after adult eclosion. Diapause status was confirmed by measuring

primary follicle and germarium lengths, and the stage of ovarian development was

determined according to the methods described by Christophers (1911) and Spielman and

Wong (1973). None of the mosquitoes used in these experiments were offered a blood

meal.

Suppressive Subtractive Hybridization

Total RNA was isolated from pools of 20 females by grinding with 4.5 mm

copper-coated spherical balls (“BB’s”) in 1 ml TRIzol® Reagent (Invitrogen). After homogenization, samples were spun at 12,404 g at 4°C for 10 min, and the supernatant was used for RNA extraction following standard protocol (Chomczynski and Sacchi,

1987). RNA pellets were stored in absolute ethanol at -70°C and dissolved in 30 µl ultraPURE™ water (GIBCO) for use in cDNA synthesis (BD SMART™ PCR cDNA

Synthesis Kit, BD Biosciences) following standard protocol. Suppressive subtractive hybridization (SSH) was performed using the Clontech PCR-Select™ cDNA Subtraction

Kit: the forward subtracted library was constructed using females in late diapause (56-59 days post adult eclosion; short daylength, 18°C) and the reverse subtracted library was constructed using nondiapausing females (7-10 days post adult eclosion; long daylength,

87 18°C). Forward and reverse libraries were cloned using the TOPO TA Cloning™ Kit

(Invitrogen). Transformed plasmids were inserted into competent Escherichia coli cells

and grown overnight on Luria-Bertani (LB) plates containing X-Gal and ampicillin. For

each library, over 100 white colonies were isolated and grown overnight in LB-ampicillin

broth at 37°C. Colonies were then purified with QIAprep Spin Miniprep (QIAGEN) and

sequenced using the vector internal primer sites (T7 and M13R) at the Ohio State

University Plant-Microbe Genomics Facility on an Applied Biosystems 3730 DNA

Analyzer using BigDye® Terminator Cycle Sequencing chemistry (Applied Biosystems)

following manufacturer’s protocol.

Northern Blot Analysis

RNA was extracted from adults and pupae according to the methods described

above. Pupae were sexed as females based on their large size and delayed time of

development. Fifteen micrograms of denatured total RNA samples were separated by electrophoresis on a 1.4% agarose denaturing gel (0.41 M formaldehyde, 1X MOPS-

EDTA-sodium acetate). Visualization of ethidium bromide stained rRNA under UV light

exposure was used to confirm equal loading. Following the TURBOBLOTTER™ Rapid

Downward Transfer Systems protocol (Schleicher and Schuell), the RNA was transferred for 1.5 hours onto a 0.45 micron MagnaCharge nylon membrane (GE Osmonics) using downward capillary action in 3 M NaCl, 8 mM NaOH transfer buffer, followed by neutralization in a 1 M phosphate buffer solution and UV crosslinking. The membrane was then air-dried and either stored at -20°C or used immediately for hybridization.

88 Digoxigenin (DIG)-labeled cDNA probes were developed from two

mitochondrial genes generated in our forward and reverse subtracted SSH libraries. PCR

was performed on each clone using the SSH nested primers (Clontech PCR-Select™

cDNA Subtraction Kit) according to the following parameters: 94°C for 3 min and 35

cycles of 94°C for 30 sec, 60°C for 30 sec, and 72°C for 2 min, followed by a 7 min extension at 72°C and a 4°C hold. The PCR products were run on a 1% TAE agarose gel, and the band of interest was excised from any remaining vector, extracted with

Ultrafree®-DA (Millipore), and re-amplified by PCR. Cytochrome c oxidase subunit I

(COI) and cytochrome c oxidase subunit III (COIII) cDNAs were individually labeled in an overnight DIG reaction using 100ng of template DNA and the Dig High Prime DNA

Labeling and Detection Starter Kit II (Roche Applied Sciences). Probes were stored at

-20°C.

Hybridization was carried out overnight followed by stringency washes and immunological detection using the Dig High Prime DNA Labeling and Detection Starter

Kit II (Roche Applied Sciences) according to manufacturer’s protocol. Blots were then exposed to chemiluminescence film (Kodak Biomax). Each northern blot was replicated three or more times. To confirm equal transfer of RNA, each membrane was stripped with 0.2 M NaOH/0.1% SDS and re-probed using DIG-labeled 28S cDNA, according to manufacturer’s instructions.

89 Quantification of mitochondrial DNA (mtDNA)

Total DNA was extracted from pools of 15 mosquitoes by homogenization with a

plastic pestle in DNAzol (Invitrogen) following standard protocol. Samples were purified by phenol-chloroform extraction using the methods described by Sambrook et al.

(1989). DNA pellets were then resuspended in 100 µl 10mM Tris-HCL/1mM EDTA/pH

8.0 (TE) buffer containing 4 µg of DNase-free RNase A (Invitrogen) and incubated for two hours at 37oC. To purify the samples, an additional phenol-chloroform extraction

step was performed, and pellets were resuspended in 10 µl TE buffer for use in dot blot hybridization. The concentration of DNA in each sample was determined by measuring the absorbance at 260 nm using a BioSpec-mini spectrophotometer (Shimadzu). DNA samples were diluted with 0.6N NaOH to a final concentration of 125 ng in a total volume of 5 µl. This was then spotted onto a nylon membrane in sets of three.

Membranes were neutralized in 0.5M Tris buffer (pH 7.5) for 5 min., rinsed briefly in dH2O, and crosslinked by UV exposure. Hybridization, stringency washes, and

immunological detection were carried out according to the methods described above

using our DIG-labeled COI probe. Equal loading of DNA was confirmed using the 28S

cDNA probe, as described above.

Chemiluminescence-exposed films were digitized and the densitometry of each

spot was determined using Kodak 1D image analysis software (Kodak). To eliminate any

variation in DNA loading, dot blot values were calculated as the ratio of COI to 28S

90 densitometry. The resulting data was then analyzed by one-way analysis of variance

(ANOVA) and Tukey’s HSD multiple comparison using Statistica data analysis program

(Statsoft). P values < 0.01 were considered statistically significant.

Bioinfomatics Analyses

The SSH cDNAs were edited and assembled using dnaLIMS (dnaTools) and

BioEdit Sequence Alignment Editor (Isis Pharmaceuticals). Similar sequences were

identified by performing BLASTn and BLASTx searches in GenBank

(http://www.ncbi.nlm.nih.gov/). The deduced amino acid sequences were assembled,

analyzed, and aligned with similar sequences using Blastp (NCBI), the Baylor College of

Medicine Search Launcher: Sequence Utilities (http://dot.imgen.bcm.tmc.edu/seq-

util/seq-util.html), and BoxShade Server 3.21

(http://www.ch.embnet.org/software/BOX_form.html). Percent identities were obtained

by blasting two sequences with the Blastp server using BLOSUM62 matrix in NCBI’s

web server.

RESULTS

Clones

Two mitochondrial genes encoding the respiratory enzymes cytochrome c oxidase

subunit I (COI) and cytochrome c oxidase subunit III (COIII) were identified by

suppressive subtractive hybridization in our forward subtracted library as being putatively

upregulated during late diapause. Out of 96 SSH clones sequenced, we obtained 4 clones

91 encoding COI and 2 clones matching with high similarity to COIII. Two of the COI clones were used to obtain a large portion of the coding region: a 912 bp segment at the

5’ end and a 629 bp segment at the 3’ end, including the poly-A tail. Together, they comprise 1, 541 bp of the COI cDNA sequence (Figure 4.1) that matches with 92% identity the partial sequence from Cx. tarsalis cytochrome oxidase subunit I gene

(accession #: AF425847). This portion of the clone corresponds to a 504 residue deduced amino acid sequence that matches amino acids 9 through 512 in COI from Aedes aegypti

(accession #: AAK56378).

The full-length sequence of cytochrome c oxidase subunit III was obtained from one of our 4 COIII cones in our late-diapause (forward) library. Figure 4.2 shows the complete Cx. pipiens COIII cDNA which is 812 bp with 87% identity to a portion of the complete mitochondrial genome from Anopheles quadrimaculatus (accession #:

LO4272). The deduced amino acid sequence has an open reading frame starting at nucleotide 8 and is 262 residues long, matching amino acids 1 through 262 in cytochrome oxidase subunit IIII from An. quadrimaculatus. To generate dig-labeled probes for northern blot hybridization, we used our 912 bp COI SSH clone and our full-length COIII cDNA which produced bands of 1.5 and 0.8 Kb, respectively.

Comparison and Analysis of the Deduced Protein Sequences

A multiple sequence alignment of the deduced COI amino acid sequence with the sequences of other insect COIs is shown in Figure 4.3. The Cx. pipiens COI ORF shares

95% identity with COI from the mosquito Ae. aegypti. In addition, our COI aligns with

92 94% and 88% identities to an An. gambiae COI and Drosophila melanogaster COI,

respectively. A Blastp search revealed that this segment includes the entire putative

functional COI domain, as determined in other known protein sequences. Our full-length

clone encoding the second mid-diapause upregulated respiratory enzyme has an open

reading frame that aligns most closely with COIII from An. quadrimaculatus, with which

it shares a 90% identity (Figure 4.4). A multiple sequence alignment highlights the

identities with other insect species (88% identity with Ae. albopictus COIII and 85%

identity with D. melanogaster COIII).

Expression Patterns at the Onset of Diapause

The two respiratory enzymes, COI and COIII, had similar patterns of mRNA expression in the early days following adult eclosion in diapausing and nondiapausing females (Figure 4.5). The transcripts were most highly expressed in short-day pupae, and the signal remained high from day 1 through day 7 after adult eclosion. This differed from the pattern observed in mosquitoes reared under long daylength; although a strong signal was seen in late pupae, it was weaker than that observed in pupae programmed for diapause. In addition, 7 days after adult eclosion in nondiapausing females the expression levels of both COI and COIII transcripts were reduced, whereas the signal remained high in diapausing individuals. For COIII, the strongest signal was seen 2 days after adult eclosion in nondiapausing females.

93 Expression Patterns throughout Diapause and at Diapause Termination

Expression of mRNAs encoding the mitochondrial respiratory enzymes COI and

COIII was also monitored throughout diapause, beginning one week (7-10 days) post

adult eclosion and then at 30-day intervals thereafter for up to three months. The mRNA

encoding COI was only weakly expressed in nondiapausing females during the first week

following adult eclosion (Figure 4.6), but the expression was highly upregulated in their

diapausing counterparts. The signal decreased slightly on days 30 and 60, and was up

again by 120 days into diapause. A similar pattern of expression was observed for COIII

(Figure 4.6) where the highest signal was seen 30 and 120 days into diapause, but in this

case the signal dropped to low (day 30) or undetectable (day 60) levels in mid-diapause.

The dip in mitochondrial gene expression in mid diapause was consistently seen in all

replicates with both COI and COIII, but the intensity of the bands 30 and 60 days into diapause varied from low or undetectable levels to a moderately strong signal. The blots shown represent these two extremes. When diapause was broken 60 days into diapause by transferring females to long daylength and high temperature, both genes were downregulated within one week (Figure 4.6).

Quantification of Whole-Body mtDNA

Pupae reared under either short or long daylength had significantly more mitochondrial DNA than any of the adults observed. One-way ANOVA revealed a statistically significant difference in mitochondrial DNA ratios (COI/28S) between at

least two of the 8 rearing stages (df = (7, 16); F = 22.7; p < 0.01), and analysis by

94 Tukey’s HSD post-hoc test demonstrated that the significant difference is between ND18 pupae and all other adult stages, as well as between D18 pupae and all other adult stages

(dfwithin = 16; treatment = 8; all p < 0.01). No differences were detected between diapausing and nondiapausing adults.

DISCUSSION

Genes encoding the two mitochondrial respiratory enzymes, cytochrome c oxidase I and cytochrome c oxidase III, varied in mRNA intensity during the course of diapause, thus demonstrating the complex nature of this important dormant stage. Our results differ from the unchanged mitochondrial mRNA levels seen in the larval diapause

of the goldenrod gall fly E. solidaginis (Levin et al., 2003) and the absence of cytochrome

c oxidase I transcript in the pupal diapause of the sweet potato hornworm, Agrius

convolvuli (Uno et al., 2004). These differences clearly indicate that not all diapausing

insects “shut down” their metabolism during diapause in the same way.

COI and COIII clones were isolated from our forward subtracted (late-diapause)

SSH library, and their expression was confirmed by northern blot hybridization. Out of

96 clones tested, 6 were genes encoding subunits of cytochrome c oxidase. Our

partial-length COI nucleotide sequence showed high similarity to COIs from other

mosquito species. Unlike COI from Ae. aegypti, where a single “T” represents the stop

codon (Morlais and Severson, 2002), our COI clone ends with the conserved asparagine

(AAC) followed by the poly-A tail. In contrast, the full-length sequence obtained for

COIII includes the typical start (ATG) and stop (TAA) codons and matches with high

95 identity to the COIII gene from An. quadrimaculatus (accession #: NP_008691). Both

Cx. pipiens CO sequences are A-T rich, a feature observed in most insect cytochrome c oxidase sequences (Garcia-Machado et al., 1999).

The adult diapause of Cx. pipiens is programmed in the late fourth instar and early pupal stages by daylength and low temperature (Eldridge, 1966; Spielman and Wong,

1973; Sanburg and Larsen, 1973). This is followed by behavioral and physiological changes evident in young adult females preparing to enter hibernation. At this time, females increase their lipid reserves for winter survival, and they do so by feeding more readily and for a longer period of time on plant sources rich in carbohydrates (Mitchell and Briegel, 1989; Bowen et al., 1992). In addition, they actively seek a well-protected site for hibernation (Vinogradova, 2000). Thus, the upregulation of COI and COIII transcripts for the first 7 days after adult eclosion may indicate an increased demand in energy that is required during preparation for hibernation.

In contrast, nondiapausing females exhibit high levels of CO mRNA only during the first 5 days after adult eclosion, with the highest peaks seen on days 2-4 for COI and on day 2 for COIII. During the first few days of adult life, females undergo physiological changes to prepare for blood feeding, including an elevation in juvenile hormone (Readio et al., 1999) and the subsequent increase in mRNA encoding enzymes required for blood digestion (Noriega et al., 1997). JH titers peak 3 days after adult eclosion (Readio et al.,

1999), corresponding to high host-seeking activity 3-10 days after adult eclosion (Bowen et al., 1988). JH levels drop sharply by day 4 (Readio et al., 1999), are further suppressed after a blood meal has been taken, and rise again within 48-72 hours after blood digestion

96 (Li et al., 2003). Thus, the increase in CO mRNA early on in adult life in nondiapausing females may reflect physiological changes occurring in preparation of blood feeding.

Since our mosquitoes were not blood fed, 5-7 day old nondiapausing females likely entered a state of inactivity that correlates with the decrease in CO expression at this time.

Once females entered diapause, CO transcripts were less strongly expressed.

Field observations, however, suggest that Cx. pipiens remains somewhat active during diapause; they have even been observed flying freely in their overwintering sites in mid- winter (Service, 1968; Buffington, 1972; Onyeka and Boreham, 1987). Such activity would require a higher metabolic rate than observed in diapauses that occur in other developmental stages (e.g. egg and pupae) and might be the reason for the reduction, rather than absence, of CO mRNA during the overwintering period. The presence of COI and COIII transcripts during diapause is in contrast with cytochrome oxidase enzyme activity observed in some other species. In the pupal diapause of the giant silkmoth

Hyalophora cecropia, cytochrome c oxidase activity is greatly reduced, while cytochromes b5 and a+a3 become the most active enzymes (Pappenheimer and Williams,

1953). In contrast, the larval diapause of the Japanese beetle Popillia japonica has increased levels of cytochrome c oxidase activity (Ludwig, 1953), a result that is similar to the results obtained in this study. Thus, cytochrome c oxidase activity and mRNA expression patterns differ markedly in different insects, possibly as a result of the different developmental stages used for overwintering.

97 By late diapause (3 months at 18°C, short daylength), Cx. pipiens begins to come

out of diapause, as indicted by the elevated levels of CO transcript expression. Other

studies also show Cx. pipiens reared in short daylength and low temperature have

increasing follicle size after 3-4 months, corresponding to a rise in the JH titer (Readio et

al., 1999).

The lack of a decrease in mtDNA during diapause in Cx. pipiens suggests that the

differential expression of mRNA is controlled at the level of transcription. This is consistent with observations in the brine shrimp, Artemia franciscana, where mitochondria quantities are equivalent in diapausing and postdiapausing embryos

(Reynolds and Hand, 2004), but our observation with Cx. pipiens differ from the reduction in mitochondria observed in cold-adapted larvae of E. solidaginis (Levin et al.,

2003) and diapausing G. groenlandica (Kukal et al., 1989).

In conclusion, the genes from Cx. pipiens that encode cytochrome c oxidase subunit I and cytochrome c oxidase subunit III show high identity with sequences reported from other closely related species of mosquitoes. Both of these genes are downregulated during mid diapause in Cx. pipiens, but expression is high in early and

late diapause. High levels of mRNA expression at these times most likely reflect the high

energy demands needed during diapause preparation and preparation for the resumption

of reproduction. The fact that no differences were observed in mtDNA between

nondiapausing and diapausing females suggests that the number of mitochondria is not

reduced during diapause, but instead the low mRNA levels observed are likely

determined by a low rate of transcription during diapause.

98 REFERENCES

Bowen, M.F. 1992. Patterns of sugar feeding in diapausing and nondiapausing Culex pipiens (Dipetera: Culicidae) females. Journal of Medical Entomology 29:843-849.

Bowen, M.R., Davis, E.E., and Haggart, D.A. 1988. A behavioural and sensory analysis of host-seeking behaviour in the diapausing mosquito Culex pipiens. Journal of Insect Physiology 34:805-813.

Buffington, J.D. 1972. Hibernaculum choice in Culex pipiens. Journal of Medical Entomology 9:128-132.

Chomczynski, P. and Sacchi, N. 1987. Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction. Analytical Biochemistry 162:156-159.

Christophers, S.R., 1911. The development of the egg follicle in Anopheles. Paludism 2:73-88.

Danks, H.V. 1987. Insect dormancy: an ecological perspective. Biological Survey of Canada, Ottawa, pp. 19-45.

Denlinger, D.L. 1991. Relationship between cold hardiness and diapause. In: Lee, R.E., Denlinger, D.L. (Eds.), Insects at low temperatures, Chapman and Hall, New York. pp. 174-198.

Eldridge, B.F. 1966. Environmental control of ovarian development in mosquitoes of the Culex pipiens complex. Science 151:826-828.

Garcia-Machado, E., Pempera, M., Dennebouy, N, Oliva-Suarez, M., Mounolou, J.C., and Monnerot, M. 1999. Mitochondrial genes collectively suggest the paraphyly of Crustacea with respect to Insecta. Journal of Molecular Evolution 49:142-149.

Joanisse, D.R. and Storey, K.B. 1994. Mitochondrial enzymes during overwintering in two species of cold-hardy gall insects. Insect Biochemistry and Molecular Biology 24:145-150.

Joanisse, D.R., and Storey, K.B. 1996. Fatty acid content and enzymes of fatty acid metabolism in overwintering cold-hardy gall insects. Physiological Zoology 69:1079- 1095.

Kukal, O., Duman, J.G., and Serianni, A.S. 1989. Cold-induced mitochondrial degradation and cryoprotectant synthesis in the freeze-tolerant arctic caterpillars. Journal of Comparative Physiology B. 158:661-671. 99

Levin, D.B., Danks, H.V., and Barber, S.A. 2003. Variations in mitochondrial DNA and gene transcription in freezing-tolerant larvae of Eurosta solidaginis (Diptera:Tephritidae) and Gynaephora groenlandica (Lepidoptera: Lymantriidae). Insect Molecular Biology 12:281-289.

Li, Y., Hernandez-Martinez, S., Unnithan, G.C., Feyereisen, R., and Noriega, F.G. 2003. Activity of the corpora allata of adult female Aedes aegypti: effects of mating and feeding. Insect Biochemistry and Molecular Biology 33:1307-1315.

Ludwig, D. 1953. Cytochrome oxidase activity during diapause and metamorphosis of the Japanese beetle (Popillia Japonica Newman). The Journal of General Physiology 36:751-757.

Mitchell, C.J. 1983. Differentiation of host-seeking behavior from blood-feeding behavior in overwintering Culex pipiens (Diptera: Culicidae) and observations on gonotrophic dissociation. Journal of Medical Entomology 20:157-163.

Mitchell, C.J., and Briegel, H. 1989. Inability of diapausing Culex pipiens (Diptera:Culicidae) to use blood for producing lipid reserves for overwinter survival. Journal of Medical Entomology 26:318-326.

Morlais, I., and Severson, D.W. 2002. Complete mitochondrial DNA sequence and amino acid analysis of the cytochrome c oxidase subunit I (COI) from Aedes aegypti. DNA Sequence 13:123-127.

Noriega, F.G., Shah, D.K, and Wells, M.A. 1997. Juvenile hormone controls early trypsin gene transcription in the midgut of Aedes aegypti. Insect Molecular Biology 6:63-66.

Onyeka, J.O.A., and Boreham, P.F.L. 1987. Population studies, physiological state and mortality factors of overwintering adult populations of females of Culex pipiens L. (Diptera: Culicidae). Bulletin of Entomological Research 77:99-111.

Pappenheimer, Jr., A.M., and Williams, C.M. 1954. Cytochrome b5 and the dihydro- coenzyme I-oxidase system in the Cecropia silkworm. Journal of Biological Chemistry 209:915-929.

Readio, J., Chen, M., and Meola, R. 1999. Juvenile hormone biosynthesis in diapausing and nondiapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 36:355-360.

100 Reynolds, J.A., and Hand, S.C. 2004. Differences in isolated mitochondria are insufficient to account for respiratory depression during diapause in Artemia franciscana embryos. Physiological and Biochemical Zoology 77:366-377.

Sambrook, J., Fritsch, E.F., and Maniatis, T. 1989. Molecular cloning: a laboratory manual. Cold Spring Harbor Laboratory Press, New York.

Sanburg, L.L., and Larsen, J.R. 1973. Effect of photoperiod and temperature on ovarian development in Culex pipiens pipiens. Journal of Insect Physiology 19:1173-1190.

Service, M.W. 1968. Observations on the ecology of some British mosquitoes. Bulletin of Emtomological Research 59:161-194.

Spielman, A., and Wong, J. 1973. Environmental control of ovarian diapause in Culex pipiens. Annals of the Entomological Society of America 66:905-907.

Uno, T., Nakasuji, A., Shimoda, M., and Aizono, Y. 2004. Expression of cytochrome c oxidase subunit I gene in the brain at an early stage in the termination of pupal diapause in the sweet potato hornworm, Agrius convolvuli. Journal of Insect Physiology 50:35-42.

Vinogradova, E.B. 2000. Culex pipiens pipiens mosquitoes: taxonomy, distribution, ecology, physiology, genetics, applied importance and control. Pensoft Publishers, Sofia, pp. 46-115.

101 A AAT CAT AAA GAT ATT GGA ACA TTA TAT TTT ATT TTT GGG GCT 43 N H K D I G T L Y F I F G A 15

TGA GCT GGA ATA GTT GGA ACT TCT TTA AGT TTA CTA ATT CGA GCA 88 W A G M V G T S L S L L I R A 29

GAA TTA AGT CAA CCA GGT GTA TTT ATT GGA AAT GAT CAA ATT TAT 133 E L S Q P G V F I G N D Q I Y 44

AAT GTT ATT GTA ACT GCT CAT GCT TTT ATT ATA ATT TTT TTT ATA 178 N V I V T A H A F I M I F F M 59

GTA ATA CCA ATC ATA ATT GGA GGA TTT GGA AAT TGA TTA GTT CCT 223 V M P I M I G G F G N W L V P 74

TTA ATG TTA GGA GCT CCA GAT ATG GCC TTT CCT CGA ATA AAT AAT 268 L M L G A P D M A F P R M N N 89

ATA AGT TTT TGA ATA CTA CCT CCT TCA TTG ACA CTA CTA CTT TCA 313 M S F W M L P P S L T L L L S 104

AGT AGT TTA GTA GAA AAT GGA GCT GGG ACT GGA TGA ACA GTG TAT 358 S S L V E N G A G T G W T V Y 119

CCC CCT CTT TCA TCT GGA ACA GCT CAT GCT GGA GCT TCA GTA GAC 403 P P L S S G T A H A G A S V D 134

TTA GCT ATT TTT TCT TTA CAT TTA GCA GGA ATT TCA TCA ATT TTA 448 L A I F S L H L A G I S S I L 149

GGT GCA GTA AAT TTT ATT ACA ACA GTA ATT AAT ATA CGA TCT TCA 493 G A V N F I T T V I N M R S S 164

GGA ATT ACT CTT GAT CGA ATA CCT TTA TTT GTT TGA TCA GTA GTA 538 G I T L D R M P L F V W S V V 179

ATT ACT GCA GTT TTA TTA CTT CTT TCT TTA CCT GTT TTA GCT GGT 583 I T A V L L L L S L P V L A G 194

GCT ATT ACT ATG CTA TTA ACA GAT CGA AAT TTA AAT ACT TCA TTC 628 A I T M L L T D R N L N T S F 209

TTT GAT CCA ATT GGA GGA GGA GAT CCA ATT TTA TAT CAA CAT TTA 673 F D P I G G G D P I L Y Q H L 224

TTT TGA TTC TTT GGA CAT CCA GAA GTT TAT ATT TTA ATT CTT CCA 718 F W F F G H P E V Y I L I L P 239

GGG TTT GGA ATA ATT TCT CAT ATT ATT ACT CAA GAA AGA GGA AAA 763 G F G M I S H I I T Q E S G K 254

Figure 4.1 (continued on next page).

Figure 4.1. Nucleotide sequence and deduced amino acid sequence of Cx. pipiens cytochrome c oxidase subunit I (COI) cDNA. Shading indicates the putative conserved COI domain. The invertebrate mitochondrial genetic code was used for translation.

102 Figure 4.1 (continued).

AAG GAA ACA TTT GGA ACT TTA GGA ATA ATT TAT GCT ATA TTA GCT 808 K E T F G T L G M I Y A M L A 269

ATT GGT TTA TTA GGG TTT ATT GTT TGA GCT CAT CAT ATA TTT ACG 853 I G L L G F I V W A H H M F T 284

GTT GGA ATA GAT GTT GAT ACA CGA GCT TAT TTT ACA TCT GCT ACA 898 V G M D V D T R A Y F T S A T 299

ATA ATT ATT GCT GTA CCA ACA GGA ATT AAA ATT TTC AGT TGA TTA 943 M I I A V P T G I K I F S W L 314

GCT ACT CTT CAT GGA ACA CAA TTA AAC TAT ACA CCT GCT TTA TTA 988 A T L H G T Q L N Y T P A L L 329

TGA TCA TTA GGA TTT GTA TTT TTA TTT ACT GTT GGA GGA TTA ACA 1033 W S L G F V F L F T V G G L T 344

GGA GTT GTA TTA GCT AAT TCT TCT ATT GAT ATT GTT CTT CAT GAT 1078 G V V L A N S S I D I V L H D 359

ACA TAT TAT GTT GTT GCT CAT TTC CAT TAT GTA TTA TCT ATA GGA 1123 T Y Y V V A H F H Y V L S M G 374

GCT GTA TTT GCT ATT ATA GCA GGA TTT ATT CAC TGA TAT CCT TTA 1168 A V F A I M A G F I H W Y P L 389

TTA ACA GGA TTA GTA ATA AAC CCT ACA TGA TTA AAG ATT CAA TTT 1213 L T G L V M N P T W L K I Q F 404

ACT ATT ATA TTT ATT GGA GTA AAT TTA ACA TTT TTC CCA CAA CAT 1258 T I M F I G V N L T F F P Q H 419

TTC TTA GGA TTA GCA GGA ATA CCA CGA CGA TAT TCT GAT TTT CCA 1303 F L G L A G M P R R Y S D F P 434

GAT AGT TAC TTA GCA TGA AAT ATT GTT TCA TCA TTA GGT AGA ACA 1348 D S Y L A W N I V S S L G S T 449

ATT TCA TTA TTT GGA ATT GTA TTC TTT TTA TTT ATT ATT TGA GAA 1393 I S L F G I V F F L F I I W E 464

AGT ATA ATT TCT CAA CGA ACA CCT TCA TTC CCT ATA CAA TTA TCA 1438 S M I S Q R T P S F P M Q L S 479

TCA TCA ATT GAA TGA TAT CAT ACT CTT CCA CCT GCA GAA CAT ACA 1483 S S I E W Y H T L P P A E H T 494

TAT GCA GAA CTT CCA TTA TTA TCA TCT AAC AAAAAAAAAAAAAAAAAAA 1532 Y A E L P L L S S N 509

AAAAAAAAA 1541

103

ACGCGGG ATG CCA ACA CAC GCA AAT CAC CCC TTT CAT TTA GTC GAT 46 M P T H A N H P F H L V D 13

TAT AGC CCT TGA CCT TTA ACA GGA GCT ATT GGA GCT ATA ACA ACT 91 Y S P W P L T G A I G A M T T 28

GTT ACT GGT CTT GTT CAA TGA TTT CAT CAA TAT GAT TTA ACA CTA 136 V T G L V Q W F H Q Y D L T L 43

TTT ACT TTA GGA AAT ATT ATT ACT TTA ATA ACT ATA TAT CAA TGA 181 F T L G N I I T L M T M Y Q W 58

TGA CGA GAT ATC TCT CGA GAA GGA ACT TTT CAA GGA TTA CAT ACA 226 W R D I S R E G T F Q G L H T 73

TTA CCA GTT ACT TTA GGT TTA CGA TGA GGA ATA ATT TTA TTT ATT 271 L P V T L G L R W G M I L F I 88

GTT TCT GAA ATT TTC TTC TTT ATT TCT TTT TTT TGA GCT TTT TTT 316 V S E I F F F I S F F W A F F 103

CAT AGT AGT CTT TCA CCA ACA ATT GAA TTA GGA ATA ACT TGA CCC 361 H S S L S P T I E L G M T W P 118

CCA GTT GGA ATT ATT GCT TTT AAC CCC TTT CAA ATT CCT CTT TTA 406 P V G I I A F N P F Q I P L L 133

AAT ACT GCT ATT TTA TTA GCA TCA GGA GTT ACT GTA ACA TGG GCT 451 N T A I L L A S G V T V T W A 148

CAT CAT AGT TTA ATA GAA AAT AAT CAC ACT CAA GCA ACA CAA AGT 496 H H S L M E N N H T Q A T Q S 163

TTA TTT TTT ACT GTA TTA TTA GGA ATT TAT TTT TCG ATT CTT CAA 541 L F F T V L L G I Y F S I L Q 178

GGT TAT GAA TAT ATT GAA GCT TCA TTT ACA ATT GCA GAT AGT GTT 586 G Y E Y I E A S F T I A D S V 193

TAT GGT TCA ACA TTT TTT ATA GCA ACA GGA TTC CAT GGG CTT CAT 631 Y G S T F F M A T G F H G L H 208

GTA TTA ATT GGA ACA TCT TTT TTA TTA GTA TGT TTA CTA CGA CAT 676 V L I G T S F L L V C L L R H 223

ATT AAT TAT CAT TTT TCA AAA AGT CAT CAT TTT GGA TTT GAA GCT 721 I N Y H F S K S H H F G F E A 238

GCA GCA TGA TAT TGA CAT TTT GTT GAT GTA GTT TGA TTA TTT TTA 766 A A W Y W H F V D V V W L F L 253

TAT ATT TCA ATT TAC TGA TGA GGT AGA TAA AAAAAAAAAAAAAAAA 812 Y I S I Y W W G S * 262

Figure 4.2. Nucleotide sequence and deduced amino acid sequence of Cx. pipiens cytochrome c oxidase subunit III (COIII) cDNA. Shading indicates the putative conserved COIII domain. The invertebrate mitochondrial genetic code was used for translation. 104 CpiCOI 1 ------NHKDIGTLYFIFGAWAGMVGTSLSLLIRAELSQPGVFIGNDQIYNVIVTAHA AaeCOI 1 SRQWLFSTNHKDIGTLYFIFGVWSGMVGTSLSILIRAELSHPGMFIGNDQIYNVIVTAHA AgaCOI 1 SRQWLFSTNHKDIGTLYFIFGAWAGMVGTSLSILIRAELGHPGAFIGDDQIYNVIVTAHA DmeCOI 1 SRQWLFSTNHKDIGTLYFIFGAWAGMVGTSLSILIRAELGHPGALIGDDQIYNVIVTAHA

CpiCOI 53 FIMIFFMVMPIMIGGFGNWLVPLMLGAPDMAFPRMNNMSFWMLPPSLTLLLSSSLVENGA AaeCOI 61 FIMIFFMVMPIMIGGFGNWLVPLMLGAPDMAFPRMNNMSFWMLPPSLTLLLSSSMVENGA AgaCOI 61 FIMIFFMVMPIMIGGFGNWLVPLMLGAPDMAFPRMNNMSFWMLPPSLTLLISSSMVENGA DmeCOI 61 FIMIFFMVMPIMIGGFGNWLVPLMLGAPDMAFPRMNNMSFWLLPPALSLLLVSSMVENGA

CpiCOI 113 GTGWTVYPPLSSGTAHAGASVDLAIFSLHLAGISSILGAVNFITTVINMRSSGITLDRMP AaeCOI 121 GTGWTVYPPLSSGTAHAGASVDLAIFSLHLAGISSILGAVNFITTVINMRSSGITLDRLP AgaCOI 121 GTGWTVYPPLSSGIAHAGASVDLAIFSLHLAGISSILGAVNFITTVINMRSPGITLDRMP DmeCOI 121 GTGWTVYPPLSAGIAHGGASVDLAIFSLHLAGISSILGAVNFITTVINMRSTGISLDRMP

CpiCOI 173 LFVWSVVITAVLLLLSLPVLAGAITMLLTDRNLNTSFFDPIGGGDPILYQHLFWFFGHPE AaeCOI 181 LFVWSVVITAILLLLSLPVLAGAITMLLTDRNLNTSFFDPIGGGDPILYQHLFWFFGHPE AgaCOI 181 LFVWSVVITAVLLLLSLPVLAGAITMLLTDRNLNTSFFDPAGGGDPILYQHLFWFFGHPE DmeCOI 181 LFVWSVVITALLLLLSLPVLAGAITMLLTDRNLNTSFFDPAGGGDPILYQHLFWFFGHPE

CpiCOI 233 VYILILPGFGMISHIITQESGKKETFGTLGMIYAMLAIGLLGFIVWAHHMFTVGMDVDTR AaeCOI 241 VYILILPGFGMISHIITQESGKKETFGTLGMIYAMLTIGLLGFIVWAHHMFTVGMDVDTR AgaCOI 241 VYILILPGFGMISHIITQESGKKETFGNLGMIYAMLAIGLLGFIVWAHHMFTVGMDVDTR DmeCOI 241 VYILILPGFGMISHIISQESGKKETFGSLGMIYAMLAIGLLGFIVWAHHMFTVGMDVDTR

CpiCOI 293 AYFTSATMIIAVPTGIKIFSWLATLHGTQLNYTPALLWSLGFVFLFTVGGLTGVVLANSS AaeCOI 301 AYFTSATMIIAVPTGIKIFSWLATLHGTQLTYSPALLWSLGFVFLFTVGGLTGVVLANSS AgaCOI 301 AYFTSATMIIAVPTGIKIFSWLATLHGTQLTYSPAMLWAFGFVFLFTVGGLTGVVLANSS DmeCOI 301 AYFTSATMIIAVPTGIKIFSWLATLHGTQLSYSPAILWALGFVFLFTVGGLTGVVLANSS

CpiCOI 353 IDIVLHDTYYVVAHFHYVLSMGAVFAIMAGFIHWYPLLTGLVMNPTWLKIQFTIMFIGVN AaeCOI 361 IDIVLHDTYYVVAHFHYVLSMGAVFAIMAGFIHWYPLLTGMVMNPSWLKAQFSMMFIGVN AgaCOI 361 IDIVLHDTYYVVAHFHYVLSMGAVFAIMAGFVHWYPLLTGLTMNPTWLKIQFSIMFVGVN DmeCOI 361 VDIILHDTYYVVAHFHYVLSMGAVFAIMAGFIHWYPLFTGLTLNNKWLKSHFIIMFIGVN

CpiCOI 413 LTFFPQHFLGLAGMPRRYSDFPDSYLAWNIVSSLGSTISLFGIVFFLFIIWESMISQRTP AaeCOI 421 LTFFPQHFLGLAGMPRRYSDFPDSYLTWNIISSLGSTISLFAVIFFLFIIWESMITQRTP AgaCOI 421 LTFFPQHFLGLAGMPRRYSDFPDSYLTWNVVSSLGSTISLFAILYFLFIIWESMITQRTP DmeCOI 421 LTFFPQHFLGLAGMPRRYSDYPDAYTTWNIVSTIGSTISLLGILFFFFIIWESLVSQRQV

CpiCOI 473 SFPMQLSSSIEWYHTLPPAEHTYAELPLLSSN AaeCOI 481 SFPMQLSSSIEWYHTLPPAEHTYSELPLLSSN AgaCOI 481 AFPMQLSSSIEWYHTLPPAEHTYAELPLLTNN DmeCOI 481 IYPIQLNSSIEWYQNTPPAEHSYSELPLLTN-

Figure 4.3. Multiple sequence alignment of the deduced Cx. pipiens cytochrome c oxidase subunit I (COI) with other insect COI sequences retrieved from GenBank. Amino acids identical to Cx. pipiens are shaded. CpiCOI: Cx. pipiens COI, XXXXXX; AaeCOI: Ae. aegypti COI, AAK56378; AgaCOI: An. gambiae COI, AAD12191, DmeCOI: D. melanogaster COI, NP_008278.

105 CpiCOIII 1 MPTHANHPFHLVDYSPWPLTGAIGAMTTVTGLVQWFHQYDLTLFTLGNIITLMTMYQWWR AquCOIII 1 MSAHANHPFHLVDYSPWPLTGAIGAMTTVSGLVQWFHQYTMTLFILGNIITILTMYQWWR AalCOIII 1 MSTHANHPFHLVDYSPWPLTGAIGAMTTVTGLVQWFHQYNNSLFLLGNIITMLTMYQWWR DmeCOIII 1 MSTHSNHPFHLVDYSPWPLTGAIGAMTTVSGMVKWFHQYDISLFVLGNIITILTVYQWWR

CpiCOIII 61 DISREGTFQGLHTLPVTLGLRWGMILFIVSEIFFFISFFWAFFHSSLSPTIELGMTWPPV AquCOIII 61 DISREGTFQGLHTFPVTIGLRWGMILFIVSEIFFFISFFWAFFHSSLSPTIELGMTWPPV AalCOIII 61 DISREGTFQGLHTIPVTLGLRWGMILFIISEVFFFISFFWAFFHSSLSPTIELGMIWPPI DmeCOIII 61 DVSREGTYQGLHTYAVTIGLRWGMILFILSEVLFFVSFFWAFFHSSLSPAIELGASWPPM

CpiCOIII 121 GIIAFNPFQIPLLNTAILLASGVTVTWAHHSLMENNHTQATQSLFFTVLLGIYFSILQGY AquCOIII 121 GIIAFNPFQIPLLNTAILLASGVTVTWAHHALMESNHSQATQGLFFTIVLGIYFSILQAY AalCOIII 121 GIIPFNPFQIPLLNTAILLASGVTVTWAHHTLMESNHSQTTQGLFFTIMLGIYFSILQAY DmeCOIII 121 GIISFNPFQIPLLNTAILLASGVTVTWAHHSLMENNHSQTTQGLFFTVLLGIYFTILQAY

CpiCOIII 181 EYIEASFTIADSVYGSTFFMATGFHGLHVLIGTSFLLVCLLRHINYHFSKSHHFGFEAAA AquCOIII 181 EYIEAPFTIADAVYGSTFYMATGFHGLHVLIGTTFLLICFLRHINFHFSKNHHFGFEAAA AalCOIII 181 EYIEAPFTIADSVYGSTFYIATGFHGLHVLIGTTFLLICLLRHLNYHFSKNHHFGFEAAA DmeCOIII 181 EYIEAPFTIADSIYGSTFFMATGFHGIHVLIGTTFLLVCLLRHLNNHFSKNHHFGFEAAA

CpiCOIII 241 WYWHFVDVVWLFLYISIYWWGS AquCOIII 241 WYWHFVDVVWLFLYISIYWWGS AalCOIII 241 WYWHFVDEVWLFLYIPIYWWGN DmeCOIII 241 WYWHFVDVVWLFLYITIYWWGG

Figure 4.4. Multiple sequence alignment of the deduced Cx. pipiens cytochrome c oxidase subunit III (COIII) with other insect COIII sequences retrieved from GenBank. Amino acids identical to Cx. pipiens are shaded. CpiCOIII: Cx. pipiens COIII, XXXXXX; AquCOIII: An. quadrimaculatus COIII, NP_008691; AalCOIII: Ae. albopictus COIII, AAL61977, DmeCOIII: D. melanogaster COIII, NP_008282.

106 A. cytochrome c oxidase I B. cytochrome c oxidase III nondiapause nondiapause

28S diapause diapause

28S P 1 2 3 4 5 6 7 P 1 2 3 4 5 6 7 days after adult eclosion days after adult eclosion

Figure 4.5. Temporal pattern of expression of the genes encoding the mitochondrial respiratory enzymes cytochrome c oxidase I and cytochrome c oxidase III in late pupae (P) and during the first 7 days post adult eclosion in nondiapausing and diapausing females reared at 18°C. Each lane contains 15 µg of RNA isolated from pools of 20 females. Each membrane was stripped and re-probed with dig-labeled 28S cDNA to confirm equal loading.

107

A. cytochrome c oxidase I B. cytochrome c oxidase III

28S

ND 10 30 60 90 break ND 10 30 60 90 break days in diapause days in diapause

Figure 4.6. Expression of cytochrome c oxidase subunit I and cytochrome c oxidase subunit III throughout diapause (short daylength; 18°C), and when diapause is broken at two months. ND represents 10 day old females reared under nondiapausing (short daylength; 18°C) conditions. Each lane contains 15 µg of RNA isolated from pools of 20 females. A 28S cDNA probe was used to confirm equal loading.

108

CHAPTER 5

Enhanced cold and desiccation tolerance in diapausing adults of Culex pipiens,

and a role for hsp70 in response to cold shock but not as a component of the diapause

program.

ABSTRACT

Cx. pipiens reared under diapause-inducing conditions (short daylength; 18°C) were more cold-tolerant and desiccation resistant than their nondiapausing counterparts

(long daylength; 18°C). Upon cold exposure (-5°C), diapausing mosquitoes reared at

18°C survived nearly twice as long as nondiapausing mosquitoes reared at 18°C and 10 times as long as nondiapausing mosquitoes reared at 25°C. Thus, rearing temperature provided partial protection against low temperature injury in nondiapausing mosquitoes, but maximum resistance to cold was attained by the diapause state. In this species, the supercooling point is not a good indicator of cold tolerance. Both diapausing and nondiapausing females had supercooling points of approximately -16°C, but diapausing as well as nondiapausing females died at temperatures well above the supercooling point suggesting that mortality from low temperature was due to indirect chilling injury.

Diapause conferred greater resistance to desiccation (1.6 to 2 fold increase in survival) in

109 comparison to the nondiapause state. The gene encoding a 70kDa heat shock protein,

hsp70, was not upregulated as a part of the diapause program, nor was it upregulated by

desiccation stress, but it was upregulated during recovery from cold shock. Cx. pipiens

thus differs from a number of other diapausing insect species that are known to

developmentally upregulate hsp70 during diapause.

INTRODUCTION

The northern house Mosquito, Culex pipiens (L.), is a major sylvatic vector of

West Nile virus in North America and has a wide distribution spanning much of the

temperate zone of North America, Europe, and parts of Asia (Mattingly et al., 1951).

Short daylength and low temperature received by 4th instar larvae and early pupae

program newly-eclosed adult females to enter a reproductive diapause, characterized by

an arrest in development of the primary ovarian follicles (Eldridge, 1966; Sanburg and

Larsen, 1973; Spielman and Wong, 1973). Prior to diapause, females feed only on carbohydrate sources, which they use to generate the lipid reserves needed for overwintering (Mitchell and Briegel, 1989; Bowen, 1992). They then seek well-protected sites such as caves, culverts, or unheated basements that typically contain standing or running water (Vinogradova, 2000), and survival is highest in sites that remain above 0°C with a stable humidity above 90% (Minar and Ryba, 1971).

The ability of an insect to survive low temperatures through metabolic and

physical change is referred to as cold hardiness (for reviews see Zachariassen, 1985;

Bale, 2002; Sinclair et al., 2003). Two main strategies of cold hardiness have been well-

110 described in the literature: freeze tolerance in which case the insect survives the internal

formation of ice, and freeze avoidance which characteristically involves extensive

supercooling that prevents ice formation. In this study we provide evidence that Cx. pipiens is a freeze-avoiding species that relies on supercooling for low temperature survival.

We also examine the relationship between diapause and cold hardiness in Cx. pipiens. For cold hardiness to be considered a part of the diapause program, insects reared under diapause conditions should consistently be more cold tolerant than their nondiapausing counterparts reared at the same temperature (Denlinger, 1991). In Aedes albopictus, a mosquito that overwinters as a pharate 1st instar , both diapause and

cold acclimation increase cold hardiness in the field and laboratory (Hanson and Craig,

1994). This linkage has been well-documented in a number of other insects including, for example, diapausing pupae of the flesh flies Sarcophaga crassipalpis and S. bullata

(Adedokun and Denlinger, 1984) and the adult diapause of the Colorado potato beetle

Leptinotarsa decemlineata (Lefevere et al., 1989).

Desiccation is another significant environmental stress that confronts overwintering insects. Dormant insects can resist desiccation by limiting water loss, by tolerating a low content of body water, or by mechanisms of water uptake; habitat choice is also an important feature for minimizing this form of stress (Danks, 2000). One possible option for Cx. pipiens is to avoid desiccation by moving to a more humid location or by drinking water. Alternatively, they may be able to tolerate low relative humidities through some physiological adjustments.

111 One mechanism that could possibly play a role in both desiccation resistance and

cold hardiness is the upregulation of the molecular chaperone heat shock protein 70

(hsp70). Hsp70 is upregulated during diapause in a number of insect species, and it is

well known to be upregulated by temperature extremes and desiccation in nondiapausing

individuals (Denlinger et al., 2001). For example, in the pupal diapause of the flesh fly S.

crassipalpis, hsp70 is expressed immediately upon entry into diapause and remains

upregulated until diapause is terminated (Rinehart et al., 2000), and in nondiapausing

flesh flies hsp70 is upregulated in response to desiccation (Tammariello et al., 1999;

Hayward et al., 2004). We thus test the possibility that hsp70 is expressed during

diapause and/or in response to temperature or desiccation stress.

In this study we show that (1) diapausing adults of Cx. pipiens are consistently more cold-tolerant and desiccation resistant than their nondiapausing counterparts, (2) low temperature also enhances cold tolerance to some extent in nondiapausing individuals, and (3) hsp70 is not developmentally upregulated during diapause nor responsive to desiccation stress, but is upregulated during recovery from cold shock.

MATERIALS AND METHODS

Insect Rearing

Autogenous Cx. pipiens L. (Buckeye strain) were derived from mosquitoes collected in Columbus, Ohio in September, 2000. The colony was maintained at 25°C,

112 75% r.h., with a 15hL:9hD daily light:dark cycle. Larvae were reared in 18 x 28 x 5 cm

plastic containers in de-chlorinated tap water, fed a daily diet of ground fish food

(TetraMin), and held at a density of ~250 mosquitoes per pan.

When larvae reached the 2nd instar, the rearing containers were placed under one

of three environmental conditions. Two nondiapause groups were created: the first group remained in the colony rearing room (nondiapause, 25°C), and a second nondiapause group was reared at 18°C, 75% r.h., and 15L:9D daily light:dark cycle (nondiapause,

18°C). By rearing nondiapausing mosquitoes at two temperatures, we were able to

distinguish between the effects of photoperiod and rearing temperature. To induce

diapause, a third group of mosquitoes was placed at 18°C, 75% r.h., with a 9L:15D daily

light:dark cycle (diapause, 18°C). The three rearing groups will be referred to as ND25,

ND18, and D18, respectively, throughout the remainder of the text.

Adults were provided with water and honey-soaked sponges and kept in 30.5 x

30.5 x 30.5 cm screened cages. To mimic the absence of sugar in the natural

environment during the overwintering period, honey sponges were removed from

diapause cages 10-13 days after adult eclosion. None of the mosquitoes used in these

experiments were offered a blood meal. Diapause status was confirmed by measuring the

primary follicle and germarium lengths, and the stage of ovarian development was

determined according to the methods described by Christophers (1911) and Spielman and

Wong (1973). All experiments were conducted on females, 7-14 days post adult

eclosion.

113 Monitoring Environmental Conditions at Field Sites

We recorded temperature and relative humidity from November 1, 2003 to March

27, 2004 in three culverts located in Columbus, Ohio. Culverts were chosen based on the

large number (>500) of diapausing Cx. pipiens females found in each site. Each culvert

was constructed from cement with one main opening facing east. Culverts were 1.5

meters in diameter, extended > 65 meters in length with a constant flow of running water.

Two HOBO® H8 Family Data Loggers (Onset Computer) were placed in each culvert, one at 27-39 meters and another at 54-115 meters from each entrance, and data was recorded at hourly intervals.

Low Temperature Exposure

To evaluate low temperature survival, groups of 15 laboratory-reared females were aspirated into a modified 50 ml polypropylene tube (bottom of tube was cut off, nylon screen was inserted and glued 2 cm from bottom, mosquitoes were added, tube was

capped and inverted, and the space above the screen was plugged with cotton). The

inverted tube was then submerged to the level of the screen in a Lauda model RM20

circulating glycerol-water bath that maintained a temperature of -5°C. ND18 and D18

mosquitoes were removed from the water bath at 24 hour intervals until 100% mortality

was attained. The ND25 group did not survive even 24 hours at -5°C, thus they were

removed from the bath at two hour intervals. Cold-treated mosquitoes were transferred to

750 ml plastic holding cages and returned to their original environmental chambers.

Cages were supplied with a water source, and mosquito survival was assessed 24 hours

114 after exposure to -5°C. Survival was defined as the ability of the mosquito to right itself.

Experiments for each of the three rearing groups were replicated six times.

Desiccation

For each rearing group six replicates of fifteen mosquitoes each were transferred by aspiration to 750 ml plastic holding cages and placed into 250 mm Nalgene plastic desiccators at 18°C under their respective light regimes. Six different relative humidities

(r.h.) were attained by filling the bottom of the desiccators with one of the following solutions (Winston and Bates, 1960): Drierite (0% r.h.), saturated MgCl2 (33% r.h.),

saturated Ca(NO3)2 (50% r.h.), saturated NaCl (75% r.h.), saturated KCl (85% r.h.), and

water saturated towels (100% r.h.). Survivorship, defined as the ability of a mosquito to

right itself, was assessed at daily intervals. All chambers were monitored until 100%

mortality was attained.

Supercooling points

Supercooling points were determined for 20 females from each rearing group. A

type-T 30 gauge copper-constantan thermocouple, coated with a thin film of petroleum

jelly, was placed in contact with the female’s abdomen. The mosquito was then lowered

into a thin-walled 13 x 100 mm glass test tube that was plugged with cotton, and the tube

was placed in a beaker containing 800 ml of isopropanol. The beaker was then placed in

a -70°C freezer, thereby achieving a constant cooling rate of -1°C/min. The

thermocouples were attached to an Omega HH506R datalogger (Omega Engineering),

115 which recorded the temperature of the probe at 1 second intervals. The supercooling point was defined as the last temperature recorded before observing the temperature spike generated by the latent heat of crystallization.

Statistical Analysis of Survivorship Data

LT50 values were calculated for each replicate of low temperature and desiccation treatment groups from the resulting equation of the arcsin square root transformed survival curve. These values were then averaged for each treatment group and either directly used in statistical analysis or further transformed. The temperature data required square root transformation to accommodate the wide variance according to Sokal and

Rohlf (1995) and was analyzed by one-way analysis of variance (ANOVA) and Tukey’s

HSD multiple comparison using Statistica data analysis program (Statsoft). Data from the desiccation experiments did not require further transformation and was therefore directly analyzed by two-way ANOVA with replication. If the two-way ANOVA indicated a significant difference, further analysis was conducted by identifying non- overlapping standard error bars. For both datasets, a p<0.001 was considered statistically significant.

Prior to statistical analysis, the supercooling point data was square root transformed to accommodate the variance. Analysis was performed by one-way

ANOVA. The lack of statistical differences precluded the use of post-hoc testing.

116 Cloning and Sequencing of Hsp70

Hsp 70 was initially sequenced from Cx. pipiens by PCR using primers designed

to amplify a portion of the conserved 5’ region. Full length sequence was obtained using

the SMART™ RACE cDNA Amplification Kit (Clontech) by 5’ and 3’-rapid

amplification of cDNA ends (RACE). For 3’ RACE, first strand cDNA was synthesized

from 5 µg total RNA using the manufacturer’s provided adaptor primer. Target cDNA

was then amplified using the Universal Amplification Primer and a forward gene-specific

primer (5’ -AAG GAA ACT GCT GAG GCG TA- 3’). PCR consisted of a “hot start” at

94°C for 3 min and 35 cycles of 94°C for 30s, 58°C for 30s, and 72°C for 2 min, followed by an additional 7 min at 72°C.

The 5’ RACE was carried out using two gene specific reverse primers following

manufacturer’s protocol. Template cDNA was synthesized in reverse transcription with

the primer 5’ -GTA GAC GTC TCA GAG C- 3’. After purification and TdT tailing of

the cDNA, PCR was performed using a second nested, reverse gene-specific primer (5’ –

TTC GAC GAG ACG TCC TTC TT- 3’) and the provided Abridged Anchor Primer.

The 3’- and 5’-RACE products were cloned using the TOPO TA Cloning™ Kit

(Invitrogen). Transformed plasmids were inserted into competent Escherichia coli cells

and grown overnight at 37°C on ampicillin/X-Gal treated Luria-Bertani (LB) Agar plates.

Individual white colonies were then isolated and grown overnight in LB-ampicillin broth

at 37°C and purified with QIAprep Spin Miniprep (QIAGEN). Clones were sequenced

using the vector internal primer sites (T7 and M13R) at the Ohio State University

117 Plant-Microbe Genomics Facility on an Applied Biosystems 3730 DNA Analyzer using

BigDye® Terminator Cycle Sequencing chemistry according to manufacturer’s protocol.

The 5’ and 3’ RACE products were edited and assembled using dnaLIMS

(dnaTools) and BioEdit Sequence Alignment Editor (Isis Pharmaceuticals). Similar sequences were identified by performing a BLASTn and BLASTx search in GenBank

(http://www.ncbi.nlm.nih.gov/). The deduced amino acid sequences were assembled, analyzed, and aligned with similar sequences using BLASTp (NCBI), the Baylor College of Medicine Search Launcher: Sequence Utilities (http://dot.imgen.bcm.tmc.edu/seq- util/seq-util.html), and Boxshade 3.21

(http://www.ch.embnet.org/software/BOX_form.html). Percent identities were obtained by blasting two sequences using the BLASTp server with BLOSUM62 matrix in NCBI’s web server. The full-length nucleotide sequence for Cx. pipiens heat shock protein 70

was deposited in GenBank and assigned accession number AY974355.

Northern blot analysis

Total RNA was isolated from pools of 20 females, by grinding with 4.5 mm

copper-coated spherical balls (“BB’s”) in 1 ml TRIzol® Reagent (Invitrogen). Insoluble material was removed by spinning at 12,404 g at 4°C for 10 min, and the supernatant was used in RNA extraction following standard protocol (Chomczynski and

Sacchi, 1987). RNA pellets were stored in absolute ethanol at -70°C and then dissolved

in 30 µl DEPC-treated water for use in northern blot analysis.

118 Twenty micrograms of denatured total RNA samples were separated by

electrophoresis on a 1.4% agarose denaturing gel (0.41 M formaldehyde, 1X MOPS-

EDTA-sodium acetate). The RNA was transferred onto a 0.45 micron MagnaCharge

nylon membrane (GE Osmonics) for 1.5 hours using downward capillary action in 3 M

NaCl, 8 mM NaOH transfer buffer (Schleicher and Schuell), neutralized in 1 M

phosphate buffer solution, and crosslinked with UV irradiation. The crosslinked

membrane was air-dried and either stored at -20°C or used immediately for hybridization.

Digoxigenin (DIG)-labeled hsp70 cDNA probe was prepared from a PCR product generated using the original hsp70 primers in RT-PCR. RNA expressing hsp70 was obtained by placing 15 mosquitoes in a 38°C water bath for 30 min, and the RNA was extracted with TRIzol® Reagent as described above. To obtain cDNA, random hexamers were used in reverse transcription followed by PCR consisting of a “hot start” at 94°C for

2 min and 35 cycles of 94°C for 30s, 45°C for 30s, and 72°C for 2 min, followed by an

additional 7 min at 72°C. Hsp70 cDNA was then labeled in an overnight DIG reaction

using 100ng of template DNA and the Dig High Prime DNA Labeling and Detection

Starter Kit II (Roche Applied Sciences). Probes were stored at -20°C.

Overnight hybridization was carried out at 37°C using the Dig High Prime DNA

Labeling and Detection Starter Kit II (Roche Applied Sciences). Stringency washes and

immunological detection were done according to manufacturer’s protocol, and the blots

were subsequently exposed to chemiluminescence film (Kodak Biomax). Northern

119 blotting was preformed in triplicate. To confirm equal loading of RNA, the membrane

was stripped with 0.2 M NaOH/0.1% SDS and re-probed using DIG-labeled 28S cDNA,

according to manufacturer’s instructions.

RESULTS

Environmental Conditions Recorded at Field Sites

During the interval of December 2003 to April 2004, temperatures recorded in

three Ohio culverts inhabited by diapausing females of Cx. pipiens ranged from

-8.9°C to 16.4°C, with the lowest and highest temperatures recorded on January 31, 2004

and April 25, 2004, respectively. During the winter season, two of the three culverts had

4-6 days with temperatures below -5°C. The longest continuous stretches of days with

temperatures dipping below -5°C were two 3-day periods from January 23 to January 25,

2004 and January 30 to February 01, 2004. Temperatures in the third culvert remained

above -5°C for the entire winter season, with -2.9°C as the lowest temperature recorded.

By late February, well before the termination of diapause, the warmest culvert had an

abundance of mosquitoes remaining, few remained in the coldest culvert, and an

intermediate number remained in the culvert with an intermediate mean temperature, thus

suggesting that females in the coldest sites either died or moved to another location.

Relative humidity varied greatly in each site, fluctuating 40% or more during any

given month. Relative humidity reached as high as 90% and dropped below 50% for

each month recorded, with the lowest relative humidity (36%) recorded in January. This

winter season included 23 days with a relative humidity lower than 50%. Mean relative

120 humidity was 69% in December, 64% in January, 74% in February, 80% in March, and

86% in April. By mid-April, few mosquitoes remained in any of the overwintering sites,

suggesting that they had departed by this time.

Cold Tolerance

Both diapause and low rearing temperature increased cold tolerance

(D18>ND18>ND25) (Figure 5.1). One-way ANOVA showed a statistically significant difference in cold tolerance between at least two mean LT50 values (df = (15, 2); F =

231.7; p < 0.001), and analysis by the Tukey’s HSD post-hoc test revealed that the mean

LT50 values different significantly between all three rearing groups (dfwithin = 15;

treatment = 3; all p < 0.001). Mosquitoes reared under long daylength at 25°C (ND25)

died quickly when exposed to -5°C. Initial mortality was realized with as little as two

hours of exposure, with no individuals surviving more than 12 hrs of cold treatment. The

calculated LT50 value for the ND25 mosquitoes was 4.9 ± 0.5 hrs (mean ± SE; n=6

groups of 15 mosquitoes each). Lowering the rearing temperature to 18°C for

nondiapausing mosquitoes significantly increased cold tolerance: nearly 60% of females

survived a 24 hour exposure to -5°C, and 100% mortality was not reached until 72 hours

of cold treatment. The calculated LT50 for the ND18 group was 28.9 ± 0.8 hrs, a five- fold increase in cold survival when compared to the ND25 mosquitoes. Even greater

cold tolerance was observed in mosquitoes reared under diapause-inducing conditions

(D18): 86% of the females survived a 24 hour exposure to -5°C, and 100% mortality was

121 not attained until 120 hours of cold treatment. D18 females exhibited an over ten-fold increase in the LT50 value (50.3 ± 3.5 hrs) when compared to ND25 mosquitoes and a

two fold increase when compared to ND18 mosquitoes.

Desiccation Tolerance

Within all experimental groups (ND25, ND18, D18), a two-way ANOVA indicated a statistically significant difference (df = (75, 4); F = 180.4; p < 0.001) in LT50

values as a function of relative humidity, as shown in Figure 5.2. Increasing relative

humidity resulted in an increase in survival of female mosquitoes for each rearing group tested. Nondiapausing mosquitoes reared at 25°C died quickly when placed in 0% r.h., but survival increased progressively at higher humidities. More than 90% of the females placed in 100% r.h. survived the 24 day duration of our experiment, indicating that

mortality in the other groups was due to desiccation stress, rather than factors such as lack of food or water. Since a >90% survival rate was observed for each rearing group

placed at 100% r.h., the 100% r.h. exposure was considered as our control and was not

included in the statistical analysis.

Contrary to what was observed for cold stress, nondiapausing Cx. pipiens reared

at 18°C did not have an increased ability to withstand desiccative stress when compared

to nondiapausing mosquitoes reared at 25°C (Figure 5.2). For 4 of the 5 relative

humidities tested, there were no significant differences in survival to desiccation between

the two nondiapausing groups reared at different temperatures (ND25 and ND18), as

indicated by overlapping error bars. At 0% relative humidity, however, the error bars do

122 not overlap, suggesting that ND18 females are less resistant to 0% r.h. than their ND25

counterparts. The biological significance of this difference is doubtful, and clearly a

distinction between ND18 and ND25 can not be made at the other relative humidities.

In contrast, females reared in diapause-inducing conditions had a significant

increase in the ability to survive at all 5 relative humidities tested when compared to

nondiapausing mosquitoes reared at the same temperature, as indicated by the non-

overlapping error bars. This is supported by a two-way ANOVA, which showed a significant difference between the three rearing groups in survival at each relative humidity (df = (75, 2); F = 117.5; p < 0.001). D18 females showed LT50 survival with

values 1.6 to 2.0 fold higher than the survival of the ND18 group at all five

relative humidities. The two-way ANOVA also indicates a significant (df = (8, 75);

F = 10.7; p < 0.001) interaction between the diapause program and relative humidity

exposure: as relative humidity increases, there is a widening gap in survivorship between

nondiapausing (ND25 and ND18) and diapausing (D18) female mosquitoes.

Supercooling Point

Neither low rearing temperature nor diapause status significantly affected the

supercooling point (SCP) of adult Cx. pipiens. Females reared at ND25 had a mean SCP

of -16.1 ± 0.5 °C (mean ± SE; each n=20), while those reared at ND18 and D18 had

SCPs of -15.4 ± 0.2 °C and -16.0 ± 0.3 °C, respectively. The mean supercooling points

did not differ among the three rearing groups (p = 0.22; analysis of variance F-test; F =

1.55). The SCPs for all three groups were substantially lower than the experimental

123 temperature (-5°C) that caused mortality, thus indicating that this mortality was due to indirect chilling injury (Lee, 1991), rather than the direct freezing of tissues.

Cloning and Analysis of the Deduced Amino Acid Protein Sequence for Hsp70

Our initial hsp70 cDNA clone, attained using the universal primers in PCR, resulted in a 232 bp clone that aligned with 100% identity to the Anopheles gambiae hsp70 (AAM94344) amino acid sequence. To obtain full-length sequence, we used our hsp70 cDNA fragment to design gene-specific primers for 5’ and 3’ RACE. This

resulted in 735 (5’ end) and 1,725 (3’ end) bp-long fragments, both of which overlapped

the initial hsp70 clone and yielded a total product size of 2,274 bp (Figure 5.3).

Conceptual translation showed that the Cx. pipiens hsp70 open reading frame is 638

residues long, starting at nucleotide 178. The deduced amino acid sequence is highly

conserved when compared with those from An. albimanus, Drosophila melanogaster,

Ceratitis capitata, and Manduca sexta, sharing 89, 82, 80, and 76 percent identity,

respectively (Figure 5.4).

Expression of Hsp70

Northern blot analysis revealed that diapausing Cx. pipiens does not express

detectable levels of hsp70 transcript in early diapause (Figure 5.5), nor at any other time

during diapause (data not shown). In addition, the transcript was not upregulated in

response to prolonged desiccation (1 day at 0% r.h. and 4 days at 75% r.h.) in either diapausing or nondiapausing females (data not shown). Hsp70, however, was

124 upregulated following a 4 h cold shock at -5°C (Figure 5.5). The transcript was

detectable in all the experimental groups (ND25, ND18, and D18) 1 and 2h after the

mosquitoes were removed from the -5°C bath. Twenty-four hours later, the signal was

once again undetectable. No apparent differences could be seen in the intensity of the

signal in the three groups, but in all cases hsp70 was responsive to low temperature.

DISCUSSION

Cold Hardiness

Rearing Cx. pipiens at low temperature enhanced cold hardiness in nondiapausing

individuals, but maximum was attained if the mosquitoes reared at low

temperature were also programmed for diapause by being reared under short daylength.

Lowering the rearing temperature in nondiapausing mosquitoes from 25°C to 18°C resulted in a greater than 5 fold increase in the adult female’s resistance to cold.

Diapause status further increased cold hardiness: such females were able to withstand exposure to -5°C almost twice as long as their nondiapausing counterparts reared at 18°C and 10 times as long as those reared at 25°C. Cx. pipiens thus appears to be a species in which cold hardiness is a component of the diapause program, as noted for a number of other species (Denlinger, 1991; Kostal et al., 2004), including embryos of Ae. albopictus

(Hanson and Craig, 1994; Hawley et al., 1989). In both of these mosquito species, the

programming of diapause elicits cold hardening, but the cold hardening can be further

enhanced by low temperature exposure.

125 Adults of Cx. pipiens died at sub-zero temperatures that were well above their supercooling point, indicating that lethality at -5°C was due to chilling injury rather than freezing. Nondiapausing mosquitoes reared at 25°C died within hours of exposure to low temperature, and thus would be considered chill-susceptible by Bale (1996): this category includes insects that die following a short exposure to temperatures well above the supercooling point. Cold acclimation and diapause, however, shift Cx. pipiens into the

“chill tolerant” category since these individuals have a much higher level of cold tolerance (i.e. can withstand days at -5°C).

The supercooling point of Cx. pipiens is not a good indicator of low temperature tolerance. Nondiapausing adults reared at 18°C or 25°C and diapausing adults reared at

18°C all had similar SCPs around -16°C, yet even diapausing adults died following prolonged exposure to -5°C, a temperature well above the recorded supercooling point.

Although some species can tolerate temperatures close to the supercooling point, many insects are like Cx. pipiens and are unable to do so (Lee, 1991; Bale, 1996). The supercooling points we noted for Cx. pipiens are similar to those recorded for diapausing adults of An. quadramaculatus (-17.2°C) and An. punctapennis (-20.1°C) (Wallace and

Grimstad, 2002), both of which are frequently found in the same overwintering sites as

Cx. pipiens in the American Midwest. Danks (1978) suggests that the SCP often approaches the climatic minima in the species habitat, but this does not appear to be true for Cx. pipiens in central Ohio. The minimum temperature we observed in culverts used for overwintering in Ohio during the winter of 2003-04 was -8.9°C. That particular winter had lower than average temperatures in January, thus it is unlikely that

126 temperatures in these protected sites ever drop as low as -16°C. Possibly Cx. pipiens that

are in diapause can survive very brief exposure to temperatures just above the SCP, but

this has not been tested.

Desiccation Resistance

Diapausing Cx. pipiens were more resistant to desiccation than nondiapausing

females reared at either high or low temperatures. Unlike some other insect species

(Zachariassen, 1991; Ring and Danks, 1994; Block, 1996), cold acclimation of

nondiapausing females did not provide protection against exposure to dry conditions.

Although Cx. pipiens overwinters in sites that are well-protected and often contain

standing water, the relative humidity in such sites can fluctuate greatly over the course of

a winter. Three Ohio culverts monitored during the winter of 2003-04 showed large

variation in relative humidity, ranging from a low of 36% to a high of 100%. Since Cx.

pipiens diapauses in the adult stage, they have the ability to move within the

hibernaculum, and our own unpublished observations confirm that the adults do indeed

move around within the culverts during the winter. This may allow the females to

continually select optimum conditions by choosing a resting place with a high relative humidity, or they may even drink water during this time. Our unpublished laboratory observations suggest that it is critical for diapausing Cx. pipiens to have access to water:

D18 survivorship was approximately 3 weeks at 85% r.h. with no access to water or food,

while D18 females held at 75% r.h. with access to water but not to food survived more

than 4 months. Our results demonstrate that resistance to water loss is a component of

127 the diapause syndrome, but free access to water is essential for long-term survival. The

factors contributing to desiccation resistance in diapausing adults of Cx. pipiens remain

unknown.

A Role for Hsp70?

Heat shock proteins are best known for their role in the protection of normal

cellular function during exposure or recovery from environmental stress, but curiously they are also upregulated in a number of insect species as a component of the diapause program (Denlinger et al., 2001). To test the possibility of diapause upregulation of heat shock proteins in Cx. pipiens, we cloned hsp70 and determined whether it was upregulated during diapause or by environmental stresses (high or low temperature, desiccation) during diapause.

The Cx. pipiens hsp70 deduced amino acid sequence is highly conserved when compared with those from other insects. It is most similar (89% amino acid identity) to hsp70 from the mosquito An. albimanus (Benedict, et al., 1993). In Cx. pipiens, hsp70 is not expressed as a component of diapause. Thus, Cx. pipiens is unlike a number of other species that upregulate hsps upon entry into diapause, even in the absence of thermal stress. Diapause upregulation of hsps has been well-documented in the pupal diapause of

the flesh fly S. crassipalpis (Yocum et al., 1998; Rinehart et al., 2000), as well as a

number of other including diapausing embryos of the brine shrimp Artemia

franciscana (McRae, 2003), and diapausing adults of the Colorado potato beetle L.

decemlineata (Yocum, 2001). But, upregulation of hsps does not occur in all species.

128 For example, none of the hsps appear to be upregulated during the adult dipauase of

Drosophila triauraria (Goto et al., 1998; Goto and Kimura, 2004). Although one of the hsp70 transcripts in diapausing adults of L. decemlineata is upregulated (Yocum, 2001), the upregulation is rather modest in comparison to that observed in flesh fly pupae

(Rinehart et al., 2000). Thus, the hsp70 expression pattern we observed in Cx. pipiens in association with diapause is quite similar to what has been observed previously in other adult diapauses. Although relatively few species have been examined, it would appear that diapause upregulation of hsp70 is less common in adult diapause than it is in diapauses occurring in pre-adult stages of development.

Although hsp70 is not expressed as a component of the diapause program in Cx. pipiens, hsp70 does remain responsive to cold shock (4 h at -5°C) during diapause.

Expression was not observed during the cold shock, but only after a 1-2 hour recovery period from cold shock, and the signal again returned to undetectable levels after 24 hours. There were no major differences in the intensity of response by all three groups tested (ND25, ND18, and D18), indicating that hsp70 expression elicited by cold shock was not affected by the diapause state or cold acclimation. The signal detected by northern blotting following cold shock was also weak when compared to the heat shock response. Together, the weak intensity of the response, coupled with the brevity of expression, suggests that hsp70 is most likely not the major factor contributing to the increased cold resistance seen in diapausing females. In diapausing embryos of the gypsy moth, Lymantria dispar, hsp70 is expressed strongly in response to low temperature and

129 expression then persists for the duration of diapause (Yocum et al., 1991; Denlinger et al.,

1992), but such is clearly not the pattern observed in Cx. pipiens. In Cx. pipiens, the gene

is turned on briefly in response to a cold shock but expression does not persist.

Conclusions

In this study we demonstrated that diapausing Cx. pipiens are more cold-tolerant

and desiccation resistant than their nondiapausing counterparts. Nondiapausing

mosquitoes reared at a low temperature are partially protected against cold stress, but not

to the extent of those in diapause; rearing temperature does not confer resistance to

desiccation in nondiapausing females. The supercooling point is not a good indicator of cold tolerance in Cx. pipiens, since diapausing and nondiapuasing females died when exposed to -5°C, a temperature well above their supercooling point (approximately -

16°C). This suggests that low temperature mortality was due to indirect chilling injury.

In addition, hsp70 is not upregulated as a part of the diapause program nor in response to desiccation stress, but it is upregulated during recovery from cold shock in all three rearing groups (ND25, ND18, D18), suggesting that hsp70 plays at least a transient role in protecting the mosquito from low temperature injury.

130 REFERENCES

Adedokun, T.A.., and Denlinger, D.L. 1984. Cold-hardiness: a component of the diapause syndrome in pupae of the flesh flies, Sarcophaga crassipalpis and Sarcophaga bullata. Physiological Entomology 9:361-364.

Bale, J.S. 1996. Insect cold hardiness: a matter of life and death. European Journal of Entomology 93:369-382.

Bale, J.S. 2002. Insects and low temperatures: from molecular biology to distribution and abundance. Philosophical Transactions of the Royal Society of London Series B 357:849-862.

Benedict, M.Q., Cockburn, A.F., and Seawright, J.A. 1993. The Hsp70 heat-shock gene family of the mosquito Anopheles albimanus. Insect Molecular Biology 2:93-102.

Block, W. 1996. Cold or drought – the lesser of two evils for terrestrial arthropods? European Journal of Entomology 93:325-339.

Bowen, M.F. 1992. Patterns of sugar feeding in diapausing and nondiapausing Culex pipiens (Dipetera: Culicidae) females. Journal of Medical Entomology 29:843-849.

Christophers, S.R., 1911. The development of the egg follicle in Anopheles. Paludism 2:73-88.

Chomczynski, P., and Sacchi, N. 1987. Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction. Analytical Biochemistry 162:156-159.

Danks, H.V. 1978. Insect dormancy: an ecological perspective. Biological Survey of Canada, Ottawa, Ontario.

Danks, H.V. 2000. Dehydration in dormant insects. Journal of Insect Physiology 46:837-852.

Denlinger, D.L., Rinehart, J.P., and Yocum, G.D. 2001. Stress proteins: a role in insect diapause? In: Denlinger, D.L., Giebultowicz, J., Sauders, D.S. (Eds.), Insect timing: circadian rhythmicity to seasonality, Elsevier Press, Oxford. pp. 155-171.

Denlinger, D.L. 1991. Relationship between cold hardiness and diapause. In: Lee, R.E., Denlinger, D.L. (Eds.), Insects at low temperatures, Chapman and Hall, New York. pp. 174-198.

131 Denlinger, D.L., Lee, R.E., Yocum G.D., and Kukal, O. 1992. Role of chilling in the acquisition of cold tolerance and the capacitation to express stress proteins in diapausing pharate larvae of the gypsy moth, Lymantria dispar. Archives of Insect Biochemistry and Physiology 21:271-280.

Eldridge, B.F. 1966. Environmental control of ovarian development in mosquitoes of the Culex pipiens complex. Science 151:826-828.

Goto, S.G., Yoshida, K.M., and Kimura, M.T. 1998. Accumulation of Hsp70 mRNA under environmental stresses in diapausing and nondiapausing adults of Drosophila triauraria. Journal of Insect Physiology 44:1009-1015.

Goto, S.G., and Kimura, M.T. 2004. Heat-shock-responsive genes are not involved in the adult diapause of Drosophila triauraria. Gene 326:117-122.

Hanson, S.M., and Craig, G.B. Jr.. 1994. Cold acclimation, diapause, and geographic origin affect cold hardiness in eggs of Aedes albopictus (Diptera: Culicidae). Journal of Medical Entomology 31:192-201.

Hawley, W.A., Pumpuni, C.B., Brady, R.H., and Craig, Jr., G.B. 1989. Overwintering survival of Aedes albopictus (Diptera: Culicidae) eggs in Indiana. Journal of Medical Entomology 26:122-129.

Hayward, S.A.L., Rinehart, J.P., and Denlinger, D.L. 2004. Desiccation and rehydration elicit distinct heat shock protein transcript responses in flesh fly pupae. Journal of Experimental Biology 207:963-971.

Kostal, V., Vambera, J., and Bastl, J. 2004. On the nature of pre-freeze mortality in insects: water balance, ion homeostasis and energy charge in the adults of Pyrrhocoris apterus. Journal of Experimental Biology 207:1509-1521.

Lee, R.L. 1991. Principles of insect low temperature tolerance. In: Lee, R.E., Denlinger, D.L. (Eds.), Insects at low temperatures, Chapman and Hall, New York. pp. 17-46.

Lefevere, K.S., Koopmanschap, A.B., and Dekort, C.A.D. 1989. Changes in the concentrations of metabolites in hemolymph during and after diapause in female Colorado potato beetle, Leptinotarsa decemlineata. Journal of Insect Physiology 35:121- 128.

Mattingly, P.F., Rozeboom, L.E., Knight, K.E., Laven, H., Drummond, F.H., Christophers, S.R., and Shute, P.G. 1951. The Culex pipiens complex. Transactions of the Royal Entomological Society of London 7:331-343.

132 McRae, T.H. 2003. Molecular chaperones, stress resistance and development in Artemia franciscana. Seminars in Cell and 14:251-258.

Minar, J. and J. Ryba. 1971. Experimental studies on overwintering conditions of mosquitoes. Folia Parasitologica (PRAHA) 18:255-259.

Mitchell, C.J., and Briegel, H. 1989. Fate of the blood meal in force-fed, diapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 26:332-341.

Rinehart, J.P., Yocum, G.D., and Denlinger, D.L. 2000. Developmental upregulation of inducible hsp70 transcripts, but not the cognate form, during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 30:515-521.

Ring, R.A., and Danks, H.V. 1994. Desiccation and cryoprotection: overlapping adaptations. Cryo-Letters 15:181-190.

Sanburg, L.L., and Larsen, J.R. 1973. Effect of photoperiod and temperature on ovarian development in Culex pipiens pipiens. Journal of Insect Physiology 19:1173-1190.

Sokal, R.R., and Rohlf, F.J. 1995. Biometry: The Principles and Practice of Statistics in Biological Research. 3rd ed. New York, W.H. Freeman.

Sinclair, B.J., Addo-Bediako, A., and Chown, S.L. 2003. Climatic variability and the evolution of insect freeze tolerance. Biological Reviews of the Cambridge Philosophical Society 78:181-195.

Spielman, A., and Wong, J. 1973. Environmental control of ovarian diapause in Culex pipiens. Annals of the Entomological Society of America 66:905-907.

Tammariello, S.P., Rinehart, J.P., and Denlinger, D.L. 1999. Desiccation elicits heat shock protein transcription in the flesh fly, Sarcophaga crassipalpis, but does not enhance tolerance to high or low temperatures. Journal of Insect Physiology 45:933-938.

Vinogradova, E.B. 2000. Culex pipiens pipiens mosquitoes: taxonomy, distribution, ecology, physiology, genetics, applied importance and control. Pensoft Publishers, Sofia. pp. 46-115.

Wallace, J.R., and Grimstad, P.R. 2002. A preliminary characterization of the physiological ecology of overwintering Anopheles mosquitoes in the Midwestern USA. Journal of the American Mosquito Control Association 18:126-127.

Winston, P.W., and Bates, D.H. 1960. Saturated solutions for the control of humidity in biological research. Ecology 41:232-237.

133 Yocum, G.D. 2001. Differential expression of two HSP70 transcripts in response to cold shock, thermoperiod, and adult diapause in the Colorado potato beetle. Journal of Insect Physiology 47:1139-1145.

Yocum, G.D., Joplin, K.H., and Denlinger, D.L. 1991. Expression of heat-shock proteins in response to high and low-temperature extremes in diapausing pharate larvae of the gypsy moth, Lymantria dispar. Archives of Insect Biochemistry and Physiology 18:239-249.

Yocum, G.D., Joplin, K.H., and Denlinger, D.L. 1998. Upregulation of a 23 kDa small heat shock protein transcript during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 28:677-682.

Zachariassen, K.E. 1985. Physiology of cold tolerance in insects. Physiological Reviews 65:799-832.

Zachariassen, K.E. 1991. The water relations of overwintering insects. In: Lee, R.E. Jr., Denlinger, D.L. (Eds.), Insects at low temperature, Chapman and Hall, New York. pp. 47-63.

134 100

)

E LT50 (hours) S 80 ● 4.9 ± 0.5 h

±

n ▲ 28.9 ± 0.8 h

a

e ■ 50.3 ± 3.5 h

m 60

(

l

a

v

i

v

r 40 ND25 ND18 D18

u

s

t

n

e

c

r 20

e

p

0 0 24487296120 hours at -5°C

Figure 5.1. Percent survival (mean + SE) of Cx. pipiens females after exposure to different durations of low temperature (-5°C). LT50 indicates the time (hours) when 50% of females were unable to right themselves. Nondiapausing females reared at 25°C (●), nondiapausing females reared at 18°C (▲), and diapausing females reared at 18°C (■). N=6 groups of 15 individuals for each data point.

135 100

0% R.H. 75% R.H. 80 LT50 (days) LT50 (days) ND25 ● 1.7 ± 0.1 ND25 ● 5.2 ± 0.4 ND18 ▲ 1.3 ± 0.1 ND18 ▲ 5.6 ± 0.4 60 D18 ■ 2.9 ± 0.2 D18 ■ 10.1 ± 0.5

40

20

0

100

33% R.H. 85% R.H. 80 LT 50 LT50 ND25 ● 2.6 ± 0.4 ND25 ● 7.6 ± 0.6 ND18 ▲ 2.8 ± 0.2 ND18 ▲ 7.8 ± 0.3 60 D18 ■ 4.4 ± 0.3 D18 ■ 14.4 ± 0.9

40

20 percent survival (mean ± SE) ± (mean survival percent 0

100

50% R.H. 100% R.H. 80 LT50 ND25 ● 3.9 ± 0.4 ND18 ▲ 3.9 ± 0.2 60 D18 ■ 6.4 ± 0.3

40

20

0 0 2 4 6 8 1012141618202224 0 2 4 6 8 10 12 14 16 18 20 22 24 days of exposure

Figure 5.2. Percent survival (mean + SE) of Cx. pipiens females at various relative humidities. LT50 indicates the time (hours) when 50% of females were unable to right themselves. Nondiapausing females reared at 25°C (●), nondiapausing females reared at 18°C (▲), and diapausing females reared at 18°C (■). N=6 groups of 15 individuals for each data point. 136 ATCAGTTCAAATCCAACAAGCGAGCAAAGCACTAGCAGAAGAGAGAAAATAACGTGAGC 59

AAGTCCTACACCAGAGGCAAACCAAAAAAGTTATCCCAAGTGATTCGAAAGTAAAGTGA 118

AAAAGATCACATAAAAGTTTTACATCAAAGATCAAGTGAAAATCAACAGTAGAGCAAAA 177

ATG TCT GCG ATT GGA ATC GAT TTG GGC ACG ACG TAT TCG TGC GTT 222 M S A I G I D L G T T Y S C V 15

GGA GTG TTC CAG CAT GGC AAG GTT GAA ATC ATC CCG AAT GAT CAG 267 G V F Q H G K V E I I P N D Q 30

GGC AAC CGG ACG ACT CCC AGC TAT GTG GCC TTT TCG GAC ACG GAA 312 G N R T T P S Y V A F S D T E 45

CGG TTG ATT GGC GAC GCG GCA AAG AAC CAG GTT GCC ATG AAC CCA 357 R L I G D A A K N Q V A M N P 60

CGC AAC ACG GTC TTC GAT GCC AAG CGA CTG ATT GGG CGC CGA TTC 402 R N T V F D A K R L I G R R F 75

GAC GAT CCG AAG ATC CAG GCC GAC TTG AAG CAC TGG CCA TTC CAG 447 D D P K I Q A D L K H W P F Q 90

GTG ATC AGT GAC GGT GGC AAG CCA AAG ATC GAG ATC GAG TTC AAA 492 V I S D G G K P K I E I E F K 105

GGC GAA CGT AAG CGG TTT GCA CCG GAA GAG ATC AGT TCC ATG GTG 537 G E R K R F A P E E I S S M V 120

TTG ACC AAG ATG AAG GAA ACT GCT GAG GCG TAT CTG GGC AAG TCG 582 L T K M K E T A E A Y L G K S 135

GTT AAG AAC GCG GTC ATC ACC GTA CCG GCG TAC TTC AAC GAT TCT 627 V K N A V I T V P A Y F N D S 150

CAA CGC CAG GCC ACT AAG GAT GCT GGA GCC ATC GCC GGG CTG AAC 672 Q R Q A T K D A G A I A G L N 165

GTT ATG AGA ATT ATC AAC GAA CCG ACG GCT GCG GCG TTG GCC TAC 717 V M R I I N E P T A A A L A Y 180

GGA CTG GAC AAG AAC CTA AAG GGT GAA CGA AAT GTG CTG ATC TTC 762 G L D K N L K G E R N V L I F 195

GAT TTG GGT GGT GGC ACC TTC GAC GTA TCC ATC TTG ACC ATT GAC 807 D L G G G T F D V S I L T I D 210

GAG GGC TCA CTG TTT GAG GTG CGT TCC ACG GCC GGT GAT ACT CAC 852 E G S L F E V R S T A G D T H 225

CTG GGA GGA GAA GAC TTT GAT AAC CGA ATG GTG TCC CAC TTT GTG 897 L G G E D F D N R M V S H F V 240

GAC GAG TTC AAG CGC AAA TAC AAG AAG GAC GTC TCG TCA AAT CCA 942 D E F K R K Y K K D V S S N P 255

CGT GCT CTG AGA CGT CTA CGA ACG GCC TGC GAG CGA GCC AAG CGT 987 R A L R R L R T A C E R A K R 270

ACA CTG TCC TCG AGC ACT GAA GCG ACT GTC GAG ATC GAT GCC CTG 1032 T L S S S T E A T V E I D A L 285

Figure 5.3 (continued on next page).

Figure 5.3. Complete nucleotide sequence and deduced amino acid sequence of Cx. pipiens hsp70 cDNA (GenBank accession no. AY974355). Shading indicates the conserved hsp70 domain. Arrows denote the primers used in 3’ and 5’ RACE. The polyadenylation site (AATAAA) is outlined by a box, and an asterisk represents the stop codon. 137 Figure 5.3 (continued).

CTG GAC GGA ATC GAC TAT TAC ACC AAG ATT TCA AGG GCT CGA TTT 1077 L D G I D Y Y T K I S R A R F 300

GAG GAA CTG TGC TCG GAC TTG TTC CGT AAC ACG CTG CAA CCG GTG 1122 E E L C S D L F R N T L Q P V 315

GAG CGA GCC CTC TCG GAT GCC AAA ATG GAC AAG AGC GCC ATC CAT 1167 E R A L S D A K M D K S A I H 330

GAC ATT GTC CTA GTT GGC GGA TCG ACC CGA ATC CCG AAA GTG CAG 1212 D I V L V G G S T R I P K V Q 345

TCA CTG CTG CAA AAC TTC TTC TGT GGA AAA GCT CTG AAC CTT TCG 1257 S L L Q N F F C G K A L N L S 360

ATT AAT CCG GAC GAG GCC GTA GCT TAC GGT GCA GCA GTT CAG GCA 1302 I N P D E A V A Y G A A V Q A 375

GCT ATT CTG AAC GGA GAC AAG GAT GAA AAG ATT CAG GAT GTT TTG 1347 A I L N G D K D E K I Q D V L 390

CTG GTG GAT GTG GCT CCT CTG TCG CTT GGA ATT GAA ACT GCC GGA 1392 L V D V A P L S L G I E T A G 405

GGA GTG ATG ACC AAG CTG ATC GAA CGC AAC AGT CGC ATC CCA TGC 1437 G V M T K L I E R N S R I P C 420

AAG CAA ACT CAA ACC TTC TCA ACG TAC GCG GAC AAC CAA CCA GGA 1482 K Q T Q T F S T Y A D N Q P G 435

GTC TCG ATT CAG GTG TTT GAG GGA GAA CGA GCC ATG ACC AAG GAC 1527 V S I Q V F E G E R A M T K D 450

AAC AAC CGG CTG GGT CAG TTT GAT TTG TCC GGA ATT CCT CCG GCA 1572 N N R L G Q F D L S G I P P A 465

CCA CGT GGC GTT CCA CAG ATT GAG GTC ACT TTC GAC TTG GAT GCC 1617 P R G V P Q I E V T F D L D A 480

AAC GGA ATC TTG AAC GTG TCG GCC AAG GAA ATG AGC TCT GGC AAG 1662 N G I L N V S A K E M S S G K 495

GAG AAG AAC ATC ACC ATC AAG AAC GAC AAG GGA CGA CTC AGC CAG 1707 E K N I T I K N D K G R L S Q 510

GCC GAC ATT GAC CGG ATG GTT TCG GAA GCG GAC CGA TTC CGC GAG 1752 A D I D R M V S E A D R F R E 525

GAA GAC GAA AAA CAG AGA GAA CGC ATT GCG GCC AGA AAC CAA CTG 1797 E D E K Q R E R I A A R N Q L 540

GAG GGC TAT TGC TTC CAG CTG AAA CAG ACG CTG GAC ACG GCC GGG 1842 E G Y C F Q L K Q T L D T A G 555

GAC AAA CTG AGC GAT TCG GAT CGG AAC ACG GTC AAG GAC AAA TGC 1887 D K L S D S D R N T V K D K C 570

GAC GAA ACG CTT CGA TGG CTG GAC GGA AAC ACG ATG GCC GAG AAG 1932 D E T L R W L D G N T M A E K 585

GAC GAG TTT GAG CAC AAG ATG AAG GAG CTG AAC CAG GTG TGC AGT 1977 D E F E H K M K E L N Q V C S 600

CCA ATC ATG ACG CGA TTG CAC CAG GGA TCG ATG CCT GGA GCT GAG 2022 P I M T R L H Q G S M P G A E 615

GCC ACC AGC TGT GGA CAG CAG GCG GGA GGA TTT GGT GGA CGA GGT 2067 A T S C G Q Q A G G F G G R G 630

GGT CCC ACC GTT GAG GAG GTT GAC TAAACGATGTCTTGATTTTATAAATTG 2118 G P T V E E V D * 638

ATTATGCTATTAGGATTTATTTAAGATTCGTTGACTGATTAATTTGTAAAGATTAATTT 2177

TAAGTTCAGATTTTGAGTTGTAACTAATTATTTTAATTTAAAAAATAATAAAGTAATGG 2236

ATAGTATCAGCTCTAAAAAAAAAAAAAAAAAAAAAAAA 2274

138 CpiHsp70 1 -MSAIGIDLGTTYSCVGVFQHGKVEIIPNDQGNRTTPSYVAFSDTERLIGDAAKNQVAMN AalHsp70A2/B2 1 MPSAIGIDLGTTYSCVGVFQHGKVEIIANDQGNRTTPSYVAFSDTERLIGDAAKNQVAMN DmeHsp70Ba/b 1 -MPAIGIDLGTTYSCVGVYQHGKVEIIANDQGNRTTPSYVAFTDSERLIGDPAKNQVAMN CcaHsp70 1 -MVAIGIDLGTTYSCFGVFQHGKVEIIANDQGNRTTPSYVAFTDSERLIGDAAKNQVAMN MseHsp70 1 -MPAIGIDLGTTYSCVGVWQHSNVEIIANDQGNRTTPSYVAFTDTERLIGDAAKNQVALN

CpiHsp70 60 PRNTVFDAKRLIGRRFDDPKIQADLKHWPFQVISDGGKPKIEIEFKGERKRFAPEEISSM AalHsp70A2/B2 61 PTNTVFDAKRLIGRKFDDPKIQADMKHWPFTVVNDGGKPKIRVEFKGERKTFAPEEISSM DmeHsp70Ba/b 60 PRNTVFDAKRLIGRKYDDPKIAEDMKHWPFKVVSDGGKPKIGVEYKGESKRFAPEEISSM CcaHsp70 60 PKNTVFDAKRLIGRKYDDPKIMEDVKHWPFKVVSDGGKPKISVEYKEENKQFAPEEISSM MseHsp70 60 PNNTVFDAKRPIGRKFDDPKIQQDMKHWPFKVVSDGGKPKIQVEFKGEMKRFAPEEISSM

CpiHsp70 120 VLTKMKETAEAYLGKSVKNAVITVPAYFNDSQRQATKDAGAIAGLNVMRIINEPTAAALA AalHsp70A2/B2 121 VLTKMKETAEAYLGQSVKNAVITVPAYFNDSQRQATKDAGAIAGLNVMRIINEPTAAALA DmeHsp70Ba/b 120 VLTKMKETAEAYLGESITDAVITVPAYFNDSQRQATKDAGHIAGLNVLRIINEPTAAALA CcaHsp70 120 VLTKMKETAEVILGTTVTDAVITVPAYFNDSQRQATKDAGRIAGLNVLRIINEPTAAALA MseHsp70 120 VLTKMKETAEAYLGSAVRDAVITVPAYFNDSQRQATKDAGAIAGLNVLRIINEPTAAALA

CpiHsp70 180 YGLDKNLKGERNVLIFDLGGGTFDVSILTIDEGSLFEVRSTAGDTHLGGEDFDNRMVSHF AalHsp70A2/B2 181 YGLDKNLKGERNVLIFDLGGGTFDVSILTIDEGSLFEVRSTAGDTHLGGEDFDNRMVGHF DmeHsp70Ba/b 180 YGLDKNLKGERNVLIFDLGGGTFDVSILTIDEGSLFEVRSTAGDTHLGGEDFDNRLVTHL CcaHsp70 180 YGLDKNLKGERNVLIFDLGGGTFDVSILTIDEGSLFEVRATAGDTHLG-EDFDNRLVSHL MseHsp70 180 YGLDKNLKGERNVLIFDLGGGTFDVSILSIDEGSLFEVKATAGDTHLGGEDFDNRLVNHL

CpiHsp70 240 VDEFKRKYKKDVSSNPRALRRLRTACERAKRTLSSSTEATVEIDALLDGIDYYTKISRAR AalHsp70A2/B2 241 VEEFKRKHKKDLSKNARALRRLRTACERAKRTLSSSTEATIEIDALMDGIDYYTKISRAR DmeHsp70Ba/b 240 ADEFKRKYKKDLRSNPRALRRLRTAAERAKRTLSSSTEATIEIDALFEGQDFYTKVSRAR CcaHsp70 239 AEEFKRKYKKDLRSNPRALRRLRTAAERAKRTLSSSTEATIEIDALFEGIDLYTKVSRAR MseHsp70 240 AEEFQRKFKKDLRSSPRALRRLRTAAERAKRTLSSSTEATIEIDALYEGIDFYTRVSRAR

CpiHsp70 300 FEELCSDLFRNTLQPVERALSDAKMDKSAIHDIVLVGGSTRIPKVQSLLQNFFCGKALNL AalHsp70A2/B2 301 FEELCSDLFRSTLQPVEKALSDAKMDKSSIHDIVLVGGSTRIPKVQSLLQNFFAGKSLNL DmeHsp70Ba/b 300 FEELCADLFRNTLQPVEKALNDAKMDKGQIHDIVLVGGSTRIPKVQSLLQDFFHGKNLNL CcaHsp70 299 FEELCADLFRQTLEPVEKALNDAKMDKNQIHVYVLVGGSTRIPKVQRLLQSFFCGKSLNL MseHsp70 300 FEELNADLFRGTLDPVEKALKDAKMDKSQIHDVVLVGGSTRIPKVQSLLQNFFCGKKLNL

CpiHsp70 360 SINPDEAVAYGAAVQAAILNGDKDEKIQDVLLVDVAPLSLGIETAGGVMTKLIERNSRIP AalHsp70A2/B2 361 SINPDEAVAYGAAVQAAILSGDKDDKIQDVLLVDVAPLSLGIETAGGVMTKLIERNSRIP DmeHsp70Ba/b 360 SINPDEAVAYGAAVQAAILSGDQSGKIQDVLLVDVAPLSLGIETAGGVMTKLIERNCRIP CcaHsp70 359 SINPDEAVAYGAAVQAAILSGDKSTEIQDVLLVDVAPLSLGIETAGGVMAKIIERNCRIP MseHsp70 360 SINPGPRRSRTAPPCSRLLRGATDSKIQDVLLVDVAPLSLGIETAGGVMPKIVERNSKIP

CpiHsp70 420 CKQTQTFSTYADNQPGVSIQVFEGERAMTKDNNRLGQFDLSGIPPAPRGVPQIEVTFDLD AalHsp70A2/B2 421 CKQTQIFSTYADNQPGVSIQVFEGERAMTKDNNLLGQFDLSGIPPAPRGVPQIEVTFDLD DmeHsp70Ba/b 420 CKQTKTFSTYADNQPGVSIQVYEGERAMTKDNNALGTFDLSGIPPAPRGVPQIEVTFDLD CcaHsp70 419 CKQTQTFSTYSDNQPGVNIQVYEGERVMTKDNNRLGTFDLSGIPPAPRGVPQIEVTFDVD MseHsp70 420 CNSRKRSLPYSDNQPAVTIQVYEGERAMTKDNNLLGTFDLTGIPPAPRGVPKIDVTFDMD

CpiHsp70 480 ANG-ILNVSAKEMSSGKEKNITIKNDKGRLSQADIDRMVSEADRFREEDEKQRERIAARN AalHsp70A2/B2 481 ANG-ILNVAAKEKSTGKEKNITIKNDKGRLSQADIDRMVSEAEKFREEDEKQRERISARN DmeHsp70Ba/b 480 ANG-ILNVSAKEMSTGKAKNITIKNDKGRLSQAEIDRMVNEAEKYADEDEKHRQRITSRN CcaHsp70 479 ANGNNLNVSAKEMSSGNAKNITIKNDKGRLSQAEIDRMVNEAGRYAEEDERQRNKIAARN MseHsp70 480 ANG-ILNVSAKENSTGRSKNIVIKNDRGRLSQAEIERMLAEAERYKEEDEKQRQRVAARN

CpiHsp70 539 QLEGYCFQLKQTLDT-AGDKLSDSDRNTVKDKCDETLRWLDGNTMAEKDEFEHKMKELNQ AalHsp70A2/B2 540 QLEAYCFNLKQSLDGEGASKLSDADRKTVQDRCEETLRWIDGNTMADKEEFEHKMQELTK DmeHsp70Ba/b 539 ALESYVFNVKQSVEQAPAGKLDEADKNSVLDKCNETIRWLDSNTTAEKEEFDHKMEELTR CcaHsp70 539 NLESYVLAVKQAWTT-LVDKLSEREKSEVTKACDDTIKWLDATRLADKEEYEDKMNTLTK MseHsp70 539 QLEAYVFSVQQALDD-AGDKLSESDKSTARSACAAALRWLDNNTLAEQEEYEHKLKDLQR

CpiHsp70 598 VCSPIMTRLHQGSMPGA--EATSCGQQAGGFGG-RGGPTVEEVD AalHsp70A2/B2 600 ACSPIMTKLHQQAAGGP--SPSSCAQQAGGFGG-RTGPTVEEVD DmeHsp70Ba/b 599 HCSPIMTKMHQQGAGAAGGPGANCGQQAGGFGG-YSGPTVEEVD CcaHsp70 598 LCTPIMTKLHSGGGAGQG---ASCGQQAGGFNGGHTGPTVEEVD MseHsp70 598 VCSPVMPKMHGGAGAGAAP-----GGQQ--H-GRGAGPTVEEVD

Figure 5.4. Multiple sequence alignment of the deduced Cx. pipiens hsp70 with other hsp70 sequences retrieved from GenBank. Amino acids identical to Cx. pipiens are shaded. CpiHsp70: Cx. pipiens hsp70, AY974355; AalHsp70A2/B2: An. albimanus hsp70 A2, AAC41543 and B2, P41827; DmeHsp70Ba/b: D. melanogaster hsp70 Ba, AAG26905 and Bb, AAG26901; Ccahsp70: C. capitata Hsp70: CAA70153; MseHsp70: M. sexta Hsp70, AAO65964. 139 heat shock protein 70 nondiapause 25°C nondiapause 18°C diapause 18°C

28S

C 0 1 2 24 C 0 1 2 24 C 0 1 2 24 HS hours after cold exposure (-5°C)

Figure 5.5. Northern blots showing expression of heat shock protein 70 in Cx. pipiens females at different times (hours) following a 4 h exposure to -5°C. A 28S cDNA probe was used to confirm equal loading. C = untreated females. HS = females heat shocked at 38°C for 30 min.

140

CONCLUSIONS

This dissertation investigates the molecular events underlying the diapause of the northern house mosquito, Culex pipiens L. Using suppressive subtractive hybridization,

we have identified genes upregulated, downregulated, and unchanged during diapause

and confirmed these results by northern blot hybridization. We have also obtained full-

length sequencing for a subset of these genes and probed their expression levels at

different time intervals throughout diapause. In addition, diapausing females were tested

for their resistance to environmental stresses, including low temperature and relative

humidity.

I. Diapause-specific gene expression during adult diapause in the northern house

mosquito, Culex pipiens L., identified by suppressive subtractive hybridization

(SSH).

1. Forty genes were isolated by suppressive subtractive hybridization as

differentially expressed in Cx. pipiens diapause and 32 of these were verifiable by

northern blot hybridization. Among these genes are 6 upregulated specifically in

141 early diapause, 17 upregulated in late diapause, and 2 upregulated throughout

diapause. In addition, we have identified 2 genes that are diapause down-

regulated and 4 that remain unchanged in diapause.

2. Among genes identified as having regulatory functions, three encode

ribosomal proteins [ribosomal protein (rp) S3A, rpS6, and rpS24] and are upregulated in late diapause. The low expression of these genes in early diapause suggests a possible role in suppression of ovarian development, since these genes have been implicated as being involved in oogenesis in other mosquito species.

3. Three food utilization genes were isolated by SSH; two genes encoding the blood digestive enzymes trypsin and serine protease, are downregulated in early

diapause, while the gene encoding an enzyme involved in the conversion of sugars to lipid reserves, fatty acid synthase, is highly upregulated at this time.

Previous studies indicate females preparing for diapause lack the host-seeking response and are instead are programmed to accumulate lipid reserves for winter survival.

4. Several stress response genes were isolated from our late-diapause

(upregulated) SSH library including heat shock protein 23, aldehyde oxidase,

selenoprotein, and disease resistance protein Cf-2. The upregulation of these

142 genes in late diapause may contribute to the female’s defense against extreme

cold, invasion by pathogenic organisms, or other environmental adversities.

5. Cx. pipiens preparing to enter hibernation are active compared to other insect

species that diapause in a different developmental stage, such as an egg or pupa.

Cx. pipiens must actively seek sugar meals and subsequently find a place to

overwinter prior to hibernation. The upregulation of genes involved in

metabolism in diapause-destined females (malate dehydrogenase,

methylmalonate-semialdehyde dehydrogenase, cytochrome oxidase subunit I and

cytochrome oxidase subunit III) may reflect the higher energy needs at this time.

6. One cytoskeletal gene, a muscle-specific actin, is also upregulated in early

diapause and may be involved in the high requirement for flight in diapause-

destined Cx. pipiens, since these individuals must fly to find sugar meals and an

adequate site for hibernation.

7. Several ribosomal genes were isolated from our late-diapause library, and

northern blot hybridizations confirm their upregulation at this time. These include

large ribosomal subunit L18, small ribosomal subunit 27A, and a large ribosomal

subunit. In addition, the 23S ribosomal RNA gene was recovered in late diapause

from the obligate intracellular bacteria of Cx. pipiens, Wolbachia.

143 8. Curiously, two genes encoding transposable elements (transposon T1-2 and

Mimo Cp2) are upregulated during late diapause in Cx. pipiens, although their

function remains unknown.

9. In addition to genes obtained by SSH with known identities, 9 genes were

isolated with unknown functions.

II. Diapause in the mosquito Culex pipiens evokes a metabolic switch that shuts

down genes encoding blood digestive enzymes and upregulates a gene associated

with sugar utilization and lipid storage.

1. Full-length sequences of genes encoding the two blood-digestive enzymes

trypsin and chymotrypsin-like were obtained by 5’- and 3’-RACE. Our Cx.

pipiens trypsin clone is 898 bp with a 783 bp open reading frame and has a 46 and

27 bp 5’ and 3’ untranslated regions, respectively. The complete chymotrypsin-

like cDNA is 881 bp with an open reading frame of 720 bp. The 5’ untranslated

region is 47 bp long, and the 3’ untranslated region is 94 bp long.

2. The 3’ end of fatty acid synthase was obtained by RACE and resulted in a 954

bp clone that encodes 48 amino acids. Our clone has a large 3’ untranslated

region, 810 bp long.

144 3. Genes encoding the blood digestive enzymes, trypsin and a chymotrypsin-like

protease, are downregulated in prehibernating females, and concurrently a gene

associated with the accumulation of lipid reserves (fatty acid synthase) is highly

upregulated.

4. As females enter diapause, fatty acid synthase is only sporadically expressed,

and the expression of trypsin and chymotrypsin-like remain undetectable.

5. Late in diapause (2-3 months at 18° C) trypsin and chymotrypsin-like begin to

be expressed as the female prepares for blood feeding upon diapause break.

III. Downregulation of mitochondrial mRNA expression, but not mitochondrial

number, during adult diapause in the northern house mosquito, Culex pipiens.

1. Using clones derived from suppressive subtractive hybridization, a large

portion of the coding region of cytochrome c oxidase subunit I (COI) was

obtained comprising 1,541 bp of the cDNA sequence that encodes a 504 residue

deduced amino acid sequence. The full-length sequence of cytochrome c oxidase

subunit III (COIII) was also obtained, which is 812 bp with a 262 deduced amino

acid sequence.

145 2. Northern blot hybridization showed the upregulation of COI and COIII in

females preparing for diapause (7-10 days post adult eclosion), as well as just

prior to the termination of diapause (90 days). Cx. pipiens are highly active prior

to entering hibernation and just before diapause break, thus the upregulation of

CO transcripts may reflect an increased energy requirement at these times.

3. In mid-diapause (30-60 days), CO transcripts are weakened, corresponding to

the most inactive state of diapausing Cx. pipiens.

IV. Enhanced cold and desiccation tolerance in diapausing adults of Culex pipiens,

and a role for hsp70 in response to cold shock but not as a component of the

diapause program.

1. Diapausing mosquitoes survived cold-exposure (-5°C) nearly twice as long as

their nondiapausing counterparts (18°C) and 10 times as long as nondiapausing

females reared at 25°C. Thus, rearing temperature provided partial protection

against chill injury, but maximum cold hardiness was attained if the mosquitoes

were programmed for diapause.

146 2. The supercooling point of Cx. pipiens was not a good indicator of low

temperature tolerance. Regardless of rearing conditions, all females had similar

supercooling points around -16°C.

3. Diapausing Cx. pipiens were more resistant to desiccation than nondiapausing

females reared at either high or low temperature, thus cold acclimation did not

confer partial protection to this type of stress.

4. The full-length sequence for heat shock protein 70 (hsp70) was acquired

through 5’ and 3’ RACE resulting in a 2, 274 bp product with an open reading

frame 638 residues long. The deduced amino acid sequence is highly conserved

when compared to other insect species.

5. Hsp70 is not upregulated as a part of the diapause program nor in responsive to

desiccation stress, but hsp70 does remain responsive to cold shock (4 h at -5°C)

during diapause.

147

APPENDIX

Field observations on diapausing Cx. pipiens: overwintering sites, environmental conditions, movement, and the absence of West Nile virus.

In late summer and early autumn, diapause-destined Culex pipiens seek well- protected sites for hibernation including natural habitats such as caves and hollows, and artificial shelters including mines, culverts, sewers, and unheated basements

(Vinogradova, 2000). Overwintering females do best in sites with a small fluctuation in temperature (< 2°C per month) and a high relative humidity over 90% (Minar and Ryba,

1971). Cx. pipiens first appear in such sites as early as August (Service, 1969; Spielman and Wong, 1973), reaching peak densities in late October (Service, 1969). Considerable movement within and between overwintering sites has been noted in diapausing females; all studies thus far have demonstrated movement in any given month tested, even in mid- winter (Berg and Lang, 1948; Service, 1969; Minar and Ryba, 1971; Buffington, 1972).

In addition, Cx. pipiens have been observed exiting their hibernaculum during the winter months (Buffington, 1972).

Mortality in overwintering Cx. pipiens often reaches as high as 75% or more, and has been documented as being due to predation by spiders and fungal disease

(Service, 1969). In addition, chill injury, desiccation, and depletion of lipid reserves may

148 also contribute to mortality in diapausing females. Service (1969) noted peak mortality

from November to December; while mortality was not quite as severe in the following

months. The termination of diapause is dependent primarily on temperature

(Oktyabrskaya et al., 1965): females departed from a warm site in Moscow nearly a

month before females departed from a nearby, cooler site.

In this study, we monitor environmental conditions in three overwintering sites of

Cx. pipiens in Ohio. Movement was also recorded for diapausing females along with the

survival of caged females placed in field sites. In addition, overwintering females were

tested for the presence of West Nile (WN) virus by RT- and TaqMan-PCRs, since

previous studies demonstrated Cx. pipiens can harbor WN virus through the winter

(Nasci et al., 2001).

MATERIALS AND METHODS

Field Sites

This study was carried out in two suburbs of Columbus, Ohio from November 19,

2003 to April 30, 2004. Three culverts were chosen as field-study sites based on their

large number (>500) of diapausing Cx. pipiens. Each culvert was constructed from

cement and had a constant flow of running water. Culverts were 1.5 meters in diameter

and extended > 65 meters in length. HOBO® H8 Family Data Loggers (Onset Computer)

were programmed to record temperature and relative humidity at hourly intervals.

HOBO®s were placed inside 14 x 14 x 5 cm plastic containers (Ziploc® Brand) that had lids modified with 10 x 10 cm nylon screening, and the caged HOBO®s were fastened to

149 the walls of the culverts with caulking. The design of the plastic container was such that

it allowed free exchange of air while protecting the HOBO® from direct contact with

water. The total length of each culvert was not measured, but each culvert extended well

beyond our study site. Since temperature and relative humidity readings from the two

HOBO®s in each culvert were quite similar, this data was reported as an average. Figure

A.1 depicts a drawing of each site, along with a photograph.

Culvert A (Mill Run), located in Upper Arlington, OH, had one main entrance

opening west and ran underneath several paved parking lots of shopping and business

complexes. The culvert opened to a small stream that contained large boulders on each

side with small trees and light brush growing amongst it. Within 5 meters, the stream

disappeared underneath a major highway. Inside the culvert, our study section ran 105 m

from the main entrance and contained a sewer grate at 52 m, a manhole at 105 m, and 2

sets of small pipes (<0.5 m diameter) entering from various locations. Two HOBO®s

were placed in the culvert at 37 and 71 meters from the main entrance. In addition, two

sugar traps were hung from the ceiling of the culvert in mid-March: one just near the

main entrance and another by the manhole.

Culvert B (Slyh Run) opened east to a small park that was heavily forested with a

stream running through it and was located in Upper Arlington, OH. This culvert

extended away from the park and ran underneath a well-established residential area. Our study section in Slyh run ran 66 m from the entrance of the culvert and contained several

150 small pipes (<0.5m diameter) and a manhole located at the end of our study site.

HOBO®s were place at 28 m and 55 m from the main entrance of the culvert. Sugar traps were not place in this site since few mosquitoes remained by March.

Culvert C (Walcutt Road) was located in Hilliard, OH in a newly developed residential area. The main entrance of the culvert opened south to a stream that led underneath a main road to an undeveloped, heavily forested area. This culvert had several openings from the main entrance including two sewer grates (22 and 61 m), two manholes (55 and 122 m), and several small pipes (<0.5m). Our study section ran 122 m from the main entrance. HOBO®s were fastened to the walls of the culvert at 39 and 115

m. Three sugar traps were placed in this site in mid-March: two were hung from the

ceiling of the culvert by each sewer grate and the third was placed near the last manhole,

at the end of our study site.

Movement of Diapausing Cx. pipiens within Hibernaculum

To determine whether diapausing Cx. pipiens move within their hibernation

sites in mid-winter, resting females were circled with colored chalk at monthly intervals beginning in December, and the number of females remaining inside each circle was

counted each subsequent month. Two groups of females were circled in each culvert:

one towards the entrance of the culvert surrounding the first HOBO® and another towards

the back of our study site on both sides of the second HOBO®. Resting females were

151 circled from December 2003 to March 2004, and a different color of chalk was used for

each month tested. If females were disturbed during this process, the circles were not

counted in our movement data.

Survivorship of Caged Females within the Culverts

To determine whether movement within the culverts was important in survival of

diapausing Cx. pipiens, 12 groups of 20 laboratory reared females (short daylength,

18°C) were placed in plastic/screened containers, constructed from the materials

described above. Four containers were fastened to the walls of each culvert with

caulking (two by each HOBO) on November 18 and 20, 2003. One month later

(December 18, 2003), the number of alive females was counted in each plastic container.

Since no laboratory-reared females survived this interval, the remaining experiments were done using wild-caught Cx. pipiens. These females were collected from sections of the culvert beyond our study site, so that our study population would not be depleted.

Ten to twenty females were collected by battery-powered aspiration and blown lightly into the screened cages from December 2003 to April 2004. Survivorship was checked at monthly intervals. If all females were dead at this point, these were removed and replaced with newly caught females. Slyh Run was not included in the study, since the mosquito population in this site was rapidly declining by mid-winter.

152 Sugar Traps

In mid-March, sugar traps were placed in each culvert to determine when the

overwintering females would begin to seek a sugar meal at the end of hibernation. Sugar

traps were constructed from 2 liter pop containers: the top of the pop containers were cut

off and a plastic funnel was inserted into this end in such a way that females could easily

enter, but not exit, the traps. Honey-soaked sponges and an apple slice was placed in

each sugar trap and hung from the ceiling of the culverts at the locations described above.

Testing for West Nile Virus in Overwintering Females

We collected overwintering Cx. pipiens from culverts located in and around

Columbus Ohio, from January to March, 2003. Diapausing mosquitoes were also

obtained from Cleveland, Ohio, and were collected by individuals from the Cuyahoga

County Board of Health. Mosquitoes were immediately frozen after collection and sent to the Ohio Department of Health in Columbus, Ohio for species identification and PCR analysis. Females were sorted by species and pooled into groups of 1-10 according to collection date. Total RNA was isolated from each pool of females by grinding with 4.5 mm copper-coated spherical balls (“BB’s”) in 1 ml TRIzol® Reagent (Invitrogen) and

RNA was extracted following standard protocol (Chomczynski and Sacchi, 1987). RNA

pellets were stored in absolute ethanol at -70°C and dissolved in 30 µl ultraPURE™

water (GIBCO) for use in reverse transcription - polymerase chain reaction (PCR) using

153 the primers and methods described by Lanciotti et al. (2000). Samples were also tested

by the Ohio Department of Health using the more sensitive TaqMan PCR, according to

the methods described by Lanciotti et al. (2000).

RESULTS AND DISCUSSION

Environmental Conditions Recorded at Field Sites

During the interval of November 19, 2003 to April 30, 2004, average

temperatures recorded in three Ohio culverts inhabited by diapausing females of Cx.

pipiens ranged from -8.4°C to 15.1°C, with the lowest and highest temperatures recorded

on January 31, 2004 and April 25, 2004, respectively (Figure A.2). Contrary to Minar and Ryba’s (1971) data, our temperatures fluctuated greater than 5°C during any given

month. Culvert C (Walcutt Road) remained above -3°C during the entire interval tested,

and this site had the largest number of mosquitoes remaining by spring. Culvert A (Mill

Run) had 2 days with temperatures below -5°C on January 1 and January 2, 2004, while

Culvert B (Slyh Run) had the longest continuous stretches of days with temperatures dipping below -5°C from January 30 to February 01, 2004. These two culverts had fewer mosquitoes remaining by spring: less than 10 mosquitoes were found resting in the coldest culvert (Slyh Run) by this time, while an intermediate number remained in the culvert with an intermediate mean temperature (Mill Run), thus suggesting that females in the coldest sites either died or moved to another location.

Relative humidity varied greatly in each site, fluctuating 25% or more during any given month (Figure A.3). Average relative humidity reached as high as 85% and

154 dropped below 60% for each month recorded, with the lowest relative humidity (37%)

recorded in January. Mill Run had one day with relative humidity dropping below 40%

on January 17, 2003, and Slyh Run had two days below 40% r.h. on January 16 and 23,

2004. The third culvert, Walcutt Road, remained above 40% r.h. during the entire

interval recorded. The winter season included 23 days with average relative humidity

lower than 50% in all three culverts. Mean relative humidity for all three culverts was

72% in November, 69% in December, 64% in January, 74% in February, 80% in March, and 86% in April. By mid-April, few mosquitoes remained in any of the overwintering sites, suggesting that they had departed by this time.

Three HOBO®s were lost during our overwintering field study: Walcutt Road

HOBO®s were lost after December 18, 2003 and April 6, 2004, and a Slyh Run HOBO

was lost after April 19, 2004. Some environmental data is thus missing in Figures A.2

and A.3.

Movement of Diapausing Cx. pipiens within the Hibernaculum

Since there was no apparent difference in movement between mosquitoes circled

near the entrance of the culvert to those deep into the study site, these two number sets

were combined for each culvert and are presented in Table A.1. In all three culverts,

greater than 97% of females moved out of the chalked circle during each month tested.

This is in agreement with other movement data reported for overwintering Cx. pipiens

(Berg and Lang, 1948; Minar and Ryba, 1971; Buffington, 1972). In England, more than

50% of marked mosquitoes moved within their hibernation site within one week of being

155 marked, and some mosquitoes were even observed to leave their sites even in mid-winter

(Service, 1968). Buffington (1972) noted females were more restless in November and

December than in the following months of winter. In our study, we noted that females

were more easily disturbed on warm days (>5°C) than on cold days (< 0°C). The number of mosquitoes found in the overwintering sites gradually declined from January to April, and in one site (Slyh Run) few to no mosquitoes remained in March, thus no movement data was obtained for this culvert for the month of April.

Survivorship of Caged Females within the Culverts

All colony-reared, caged Cx. pipiens were dead after one month in the culverts during the interval of November to December, 2003 (n = 80 per culvert), data not shown.

Wild-caught mosquitoes had slightly higher survival rates. Out of 35 wild-caught

females caged on January 29, 2004 in the Mill Run culvert, 10 (29%) were alive after 20

days, while none survived past 50 days. Thirty-two wild-caught, mosquitoes from

Walcutt Road culvert were caged on February 11, 2004. While 1 (3%) survived 30 days

under these conditions, none were alive by day 60. Although a more complete study is required to determine if Cx. pipiens require movement for overwintering survival, this preliminary data suggests movement is necessary for the females to make it through the winter. Overwintering females may require movement to find more optimal environmental conditions, or, they may need to drink water at this time. To our knowledge, no other such studies have been reported.

156 Sugar Traps

Two females were collected from our sugar traps in mid-April, one from Mill Run

and another from Walcutt Road. We anticipate that putting the traps out earlier (before

March) may result in a higher number of collected mosquitoes, since many females

appeared to have already departed the sites by the time the sugar traps were set.

Testing for West Nile virus in Overwintering Females

From January to March, 2003, 2,923 mosquitoes were collected from culverts,

sewers, and unheated basements in Cuyahoga and Deleware Counties, Ohio. The Ohio

Department of Health provided the following species identifications: 13 were Anopheles

quadrimaculatus, 75 were Anopheles punctipennis, and the rest (2,835) were identified as

Cx. pipiens. None of the mosquito pools tested were positive for West Nile virus by rt-

PCR or TaqMan PCR. West Nile virus has been found in low numbers in other

populations of overwintering Cx. pipiens. Following the initial outbreak of WN virus in

the New York City area in 1999, WN RNA was detected by TaqMan RT-PCR assay in

3% of the mosquito pools tested (3/91), and 2 of these pools were identified by a species-

diagnostic PCR to contain only Cx. pipiens mosquitoes (Nasci et al., 2001). It was not

possible to determine the identity of the third pool of mosquitoes (Nasci et al., 2001).

West Nile virus was noted in overwintering populations of Cx. pipiens collected during

February 2003 in spring houses in eastern Pennsylvania (Kristen Bardell, Pa. Dept. of

157 Environmental Protection, personal communication) and from populations collected in

2004 from culverts in Boston (Andrew Spielman, Harvard University, personal communication).

158 REFERENCES

Beaty, B.J., and Bishop, D.H.L. 1988. Bunyavirus-vector interactions. Virus Research 10:289-302.

Berg, M., and Lang, S. 1948. Observations of hibernating mosquitoes in Massachusetts. Mosquito News 8:70-71.

Buffington, J.D. 1972. Hibernaculum choice in Culex pipiens. Journal of Medical Entomology 9:128-132.

Chomczynski, P., and Sacchi, N. 1987. Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction. Analytical Biochemistry 162:156-159.

Lanciotti, R.S., Kerst, A.J., Nasci, R.S., Godsey, M.S., Mitchell, C.J., Savage, H.M., Komar, N., Panella, N.A., Allen, B.C., Volpe, K.E., Davis, B.S., and Roehrig, J.T. 2000. Rapid detection of West Nile virus from human clinical specimens, field-collected mosquitoes, and avian samples by a TaqMan reverse transcriptase-PCR assay. Journal of Clinical Microbiology 38:4066-4071.

Minar, J. and J. Ryba. 1971. Experimental studies on overwintering conditions of mosquitoes. Folia Parasitologica (PRAHA) 18:255-259.

Nasci, R.S., Savage, H.M., White, D.J., Miller, J.R., Cropp, B.C., Godsey, M.S., Kerst, A.J., Bennett, P., Gottfried, K., and Lanciotti, R.S. 2001. West Nile virus in overwintering Culex mosquitoes, New York City, 2000. Emerging Infectious Diseases 7:742-744.

Oktyabrskaya, T.A., Astakhova, N.A., and Boiko, L.P. 1965. Materials on the species composition, biology and ecology of mosquitoes observed near Moscow. Meditsinskaia Parazitologiia i Parazitarnye Bolezni 34:510-513.

Service, M.W. 1968. Observations on the ecology of some British mosquitoes. Bulletin of Entomological Research 59:161-194.

Spielman, A., and Wong, J. 1973. Environmental control of ovarian diapause in Culex pipiens. Annals of the Entomological Society of America 66:905-907.

Vinogradova, E.B. 2000. Culex pipiens pipiens mosquitoes: taxonomy, distribution, ecology, physiology, genetics, applied importance and control. Pensoft Publishers, Sofia. pp. 46-115.

159

Wesenberg-Lund, C. 1921. Contributions to the biology of the Danish Culicidae. K. danske Vidensk. Selsk. Skr.

160 A. Mill Run

Legend HOBO® data logger entrance sugar trap sewer grate

manhole cover

10 meters

N

W E S

sugar trap

Figure A.1 (continued on next page).

Figure A.1. Diagrams and photos of three Ohio culverts included in our field study on overwintering populations of Cx. pipiens. Figure A.1 (continued). 161

B. Slyh Run

sugar trap

Legend

N HOBO® data logger

W E sewer grate manhole cover S

10 meters

entrance

Figure A.1 (continued on next page).

162 Figure A.1 (continued).

C. Walcutt Road

sugar sugar trap trap

sugar trap

Legend N

HOBO® data logger W E

sewer grate S

manhole cover

10 meters entrance

163

A. Mill Run 25

20

15

10

5 C) ° 0

-5

-10

Temperature ( -15 culvert -20 air -25 11/19/03 11/26/03 12/03/03 12/10/03 12/17/03 12/24/03 12/31/03 01/07/04 01/14/04 01/21/04 01/28/04 02/04/04 02/11/04 02/18/04 02/25/04 03/03/04 03/10/04 03/17/04 03/24/04 03/31/04 04/07/04 04/14/04 04/21/04 04/28/04

date (mm/dd/yr) Figure A.2 (continued on next page).

Figure A.2. Average temperature recorded in three Ohio culverts from November 19, 2003 through April 30, 2003 at hourly intervals. Air readings were obtained from the Ohio Agricultural Research and Development Center (http://www.oardc.ohio- state.edu/centernet/stations/dehome.html).

164 Figure A.2 (continued).

B. Slyh Run 25

20

15

10

5 C) ° 0

-5

-10

Temperature ( -15 culvert -20 air -25 11/19/03 11/26/03 12/03/03 12/10/03 12/17/03 12/24/03 12/31/03 01/07/04 01/14/04 01/21/04 01/28/04 02/04/04 02/11/04 02/18/04 02/25/04 03/03/04 03/10/04 03/17/04 03/24/04 03/31/04 04/07/04 04/14/04 04/21/04 04/28/04 date (mm/dd/yr)

Figure A.2 (continued on next page).

165 Figure A.2 (continued).

C. Walcutt Road 25

20

15

10

5 C) ° 0

-5

-10

Temperature ( -15 culvert -20 air -25 11/19/03 11/26/03 12/03/03 12/10/03 12/17/03 12/24/03 12/31/03 01/07/04 01/14/04 01/21/04 01/28/04 02/04/04 02/11/04 02/18/04 02/25/04 03/03/04 03/10/04 03/17/04 03/24/04 03/31/04 04/07/04 04/14/04 04/21/04 04/28/04 date (mm/dd/yr)

166 A. Mill Run 100

80

60

40 % relative% humidity culvert air 20 11/19/03 11/26/03 12/03/03 12/10/03 12/17/03 12/24/03 12/31/03 01/07/04 01/14/04 01/21/04 01/28/04 02/04/04 02/11/04 02/18/04 02/25/04 03/03/04 03/10/04 03/17/04 03/24/04 03/31/04 04/07/04 04/14/04 04/21/04 04/28/04

date (mm/dd/yr) FigureA.3 (continued on next page).

Figure A.3. Average relative humidity recorded in three Ohio culverts from November 19, 2003 through April 28, 2004 at hourly intervals. Relative humidity recorded from the air was obtained from the Ohio Agricultural Research and Development Center (http://www.oardc.ohio-state.edu/centernet/stations/dehome.html).

167

Figure A.3 (continued).

B. Slyh Run 100

80

60

40 % relative humidity relative % culvert air 20 11/19/03 11/26/03 12/03/03 12/10/03 12/17/03 12/24/03 12/31/03 01/07/04 01/14/04 01/21/04 01/28/04 02/04/04 02/11/04 02/18/04 02/25/04 03/03/04 03/10/04 03/17/04 03/24/04 03/31/04 04/07/04 04/14/04 04/21/04 04/28/04

date (mm/dd/yr) Figure A.3 (continued on next page).

168

Figure A.3 (continued).

C. Walcutt Road 100

80

60

40 % relative humidity relative % culvert air 20 11/19/03 11/26/03 12/03/03 12/10/03 12/17/03 12/24/03 12/31/03 01/07/04 01/14/04 01/21/04 01/28/04 02/04/04 02/11/04 02/18/04 02/25/04 03/03/04 03/10/04 03/17/04 03/24/04 03/31/04 04/07/04 04/14/04 04/21/04 04/28/04

date (mm/dd/yr)

169

A. Mill Run B. Slyh Run C. Walcutt Road

January 141/143 (99%) 151/151 (100%) 197/202 (98%) February 121/122 (99%) 39/40 (98%) 173/176 (98%) March 93/94 (99%) 11/11 (100%) 118/120 (98%) April 59/59 (100%) n/a 52/52 (100%)

Table A.1. Ratio (#moved/total circled) and percent of overwintering Cx. pipiens moving within their hibernation sites from January to April, 2004.

170

BIBLIOGRAPHY

Adedokun, T.A., and Denlinger, D. L. 1985. Metabolic reserves associated with pupal diapause in the flesh fly, Sarcophaga crassipalpis. Journal of Insect Physiology 31:229- 233.

Ashok, M., Turner, C., and Wilson, T.G. 1998. Insect juvenile hormone resistance gene homology with the bHLH-PAS family of transcriptional regulators. Proceedings of the National Academy of Sciences USA 95:2761-2766.

Bailey, C.L., Faran, M.E., Gargan, T.P., and Hayes, D.E. 1982. Winter survival of blood-fed and nonblood-fed Culex pipiens L. American Journal of Tropical Medicine and Hygiene 31:1054-1061.

Barr, A.R., 1957. The distribution of Culex p. pipiens and Culex p. quinquefasciatus in North America. American Journal of Tropical Medicine and Hygiene 6:153-156.

Beaty, B.J., and Bishop, D.H.L. 1988. Bunyavirus-vector interactions. Virus Research 10:289-302.

Bellamy, R.E., Reeves, W.C., and Scrivani, R.P. 1958. Relationships of mosquito vectors to winter survival of encephalitis viruses. II. Under experimental conditions. American Journal of Hygiene 67:90-100.

Berg, M., and Lang, S. 1948. Observations of hibernating mosquitoes in Massachusetts. Mosquito News 8:70-71.

Blandin, S., Moita, L.F., Kocher, T., Wilm, M., Kafatos, F.C., and Levashina, E.A. 2002. Reverse genetics in the mosquito Anopheles gambiae: targeted disruption of the defensin gene. EMBO Reports 3:852-856.

Blitvich, B.J., Rayms-Keller, A., Blair, C.D., and Beaty, B.J. 2001. Identification and sequence determination of mRNAs detected in dormant (diapausing) Aedes triseriatus mosquito embryos. DNA Sequence 12:197-202.

Bowen, M.F. 1992. Patterns of sugar feeding in diapausing and nondiapausing Culex pipiens (Dipetera: Culicidae) females. Journal of Medical Entomology 29:843-849. 171

Bowen, M.F., Davis, E.E., and Haggar, D.A. 1988. A behavioral and sensory analysis of host-seeking behavior in the diapausing mosquito Culex pipiens. Journal of Insect Physiology 34:805-813.

Buffington, J.D. 1972. Hibernaculum choice in Culex pipiens. Journal of Medical Entomology 9(2):128-132.

Caplen, N.J., Fleenor, J., Fire, A., and Morgan, R.A. 2000. dsRNA-mediated gene silencing in cultured Drosophila cells: a tissue culture model for the analysis of RNA interference. Gene 252:95-105.

Chandler, L.J., Wasieloski, L.P., Blair, C.D., and Beaty, B.J. 1996. Analysis of La Crosse virus S-segment RNA and its positive- transcripts in persistently infected mosquito tissues. Journal of Virology 70:8972-8976.

Chirala, S.S., Chang, H., Matzuk, M., Abu-Elheiga, L., Mao, J.Q., Mahon, K., Finegold, M., and Wakil, S.J. 2003. Fatty acid synthesis is essential in embryonic development: fatty acid synthase null mutants and most of the heterozygotes die in utero. Proceedings of the National Academy of Sciences of the United States of America 100:6358-6363.

Christophers, S.R. 1911. The development of the egg follicle in Anopheles. Paludism 1:73-88.

Clements, A.N. 1992. The biology of mosquitoes, volume I: development, nutrition, and reproduction. CABI Publishing, Cambridge, pp. 340-359.

Chomczynski, P., and Sacchi, N. 1987. Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction. Analytical Biochemistry 162:156-159.

Dahl, C. 1988. Taxonomic studies on Culex pipiens and C. torrentium. Biosystematics of haematophagous insects. Oxford systematics Assoc., Special Vol. 37, pp. 149-175.

Daibo,S., Kimura, M.T., and Goto, S.G. 2001. Upregulation of genes belonging to the drosomycin family in diapausing adults of Drosophila triauraria. Gene 278:177-184.

Denlinger, D.L. 1985. Hormonal control of diapause. In: Kerkut, G.A., Gilbert, L.I. (Eds.), Comprehensive insect physiology, biochemistry and pharmacology, Vol. 8. Pergamon Press, Oxford, pp. 353-412.

Denlinger, D.L. 2002. Regulation of diapause. Annual Review of Entomology 47:93- 122.

172 Denlinger, D.L., Giebultowicz, J., and Adedokun, T. 1988. Insect diapause: dynamics of hormone sensitivity and vulnerability to environmental stress. In: Sehnal, F., Zabza, A.,

Denlinger, D.L. (Eds.), Endocrinological Frontiers in Physiological Insect Ecology, Wroclaw Technical University Press, Wroclaw, Poland, pp. 309-324.

Denlinger, D.L., Lee, R.E., Yocum, G.D., and Kukal, O. 1992. Role of chilling in the acquisition of cold tolerance and the capacitation to express stress proteins in diapausing pharate larvae of the gypsy moth, Lymantria-dispar. Archives of Biochemistry and Physiology 21:271-280.

Denlinger, D.L., Rinehart, J.P., and Yocum, G.D. 2001. Stress proteins: a role in insect diapause? In: Denlinger, D.L., Giebultowicz, J., Sauders, D.S. (Eds.), Insect timing: circadian rhythmicity to seasonality, Elsevier Press, Oxford. pp. 155-171.

Edney, E.B. 1977. Water balance in land arthropods. New York, Springer-Verlag.

Elbashir, S.M., Harborth, J., Lendeckel, W., Yalcin, A., Weber, K., and Tuschl, T. 2001. Duplexes of 210 nucleotide RNAs mediate RNA interference in cultured mammalian cells. Nature 411:494-498.

Eldridge, B.F. 1966. Environmental control of ovarian development in mosquitoes of the Culex pipiens complex. Science 151:826-828.

Eldridge, B.F., and Bailey, C.L. 1979. Experimental hibernation studies on Culex pipiens (Diptera: Culicidae): reactivation of ovarian development and blood-feeding in prehibernating females. Journal of Medical Entomology 15:462-467.

Flannagan, R.D., Tammariello, S.P., Joplin, K.H, Cikra-Ireland, R.A., Yocum, G.D., and Denlinger, D.L., 1998. Diapause-specific gene expression in pupae of the flesh fly, Sarcophaga crassipalpis. Proceedings of the National Academy of Sciences USA 95:5616-5620.

Franks, R., and R.H.M. Hatley. 1985. Low temperature unfolding of chymotripsinogen. Cryo Letters 6:171-180.

Goto, S.G., and Denlinger, D.L. 2002. Genes encoding two cystatins in the flesh fly Sarcophaga crassipalpis and their distinct expression patterns in relation to pupal diapause. Gene 292:121-127.

Goto, S.G., and Denlinger, D.L. 2002. Short-day and long-day expression patterns of genes involved in the flesh fly clock mechanism: period, timeless, cycle and cryptochrome. Journal of Insect Physiology 48:803-816. 173

Goto, S.G., and Kimura, M.T. 1998. Heat- and cold-shock responses and temperature adaptations in subtropical and temperate species of Drosophila. Journal of Insect Physiology 44:1233-1239.

Han, Y.S., Salazar, C.E., Reese-Stardy, S.R., Cornel, A., Gorman, M.J., Collins, F.H., and Paskewitz, S.M. 1997. Cloning and characterization of a serine protease from the human malaria vector, Anopheles gambiae. Insect Molecular Biology 6(4):385-395.

Holzer, K.P., Liu, W., and Hammes, G.G. 1989. Molecular cloning and sequencing of chicken liver fatty acid synthase cDNA. Proceedings of the National Academy of Sciences of the United States of America 86(12):4387-4391.

Jiang, Q., Hall, M., Noriega, F.G., and Wells, M. cDNA cloning and pattern of expression of an adult, female-specific chymotrypsin from Aedes aegypti midgut. Insect Biochemistry and Molecular Biology 27(4):283-289.

Johnston, S.L., and R.E. Lee. 1990. Regulation of supercooling and nucleation in a freeze intolerant beetle (Tenebrio mollitor). 267:562-568.

Jung, M.K., May, G.S., and Oakley, B.R. 1998. Mitosis in wild-type and beta-tubulin mutant strains of Aspergillus nidulans. Fungal Genetics and Biology 24:146-160.

Kalhok, S.E., Tabak, L.M., Prosser, D.E., Brook, W., Downe, A.E., and White, B.N. 1993. Isolation, sequencing and characterization of two cDNA clones coding for trypsin- like enzymes from the midgut of Aedes aegypti. Insect Molecular Biology 2(2):71-79.

Kawasaki, H., Sugaya, K., Quan, G.X., Nohata, J., and Mita, K. 2003. Analysis of alpha- and beta-tubulin genes of Bombyx mori using an EST database. Insect Biochemistry and Molecular Biology 33:131-137.

Kennerdell, J.R., and Carthew, R.W. 1998. Use of dsRNA-mediated genetic interference to demonstrate that frizzled and frizzled 2 act in the wingless pathway. Cell 95:1017- 1026.

Kimura, K.D., Tissenbaum, H.A., Liu, Y., and Ruvkun, G. 1997. Daf-2, an insulin receptor-like gene that regulates longevity and diapause in Caenorhabditis elegans. Science 277:842-946.

Kraut, J. 1977. Serine proteases: structure and mechanism of catalysis. Annual review of biochemistry 46:331-358.

174 Lee, K-Y., Hiremath, S., and Denlinger, D.L. 1998. Expression of actin in the central nervous system is switched off during diapause in the gypsy moth, Lymantria dispar. Journal of Insect Physiology 44:221-226.

Lee, K.Y., Horodyski, F.M., Valaitis, A.P., and Denlinger, D.L. 2002. Molecular characterization of the insect immune protein hemolin and its high induction during embryonic diapause in the gypsy moth, Lymantria dispar. Insect Biochemistry and Molecular Biology 32:1457-1467.

Lee, S.S., Kennedy, S., Tolonen, A.C., and Ruvkun, G. 2003. DAF-16 target genes that control C. elegans life-span and metabolism. Science 300:644-647.

Liang P., and MacRae, T.H. 1997. The synthesis of a small heat shock/alpha-crystallin protein in Artemia and its relationship to stress tolerance during development. Developmental Biology 207:445-456.

Madison, E.L., Kobe, A., Gething, M.J., Sambrook, J.F., Goldsmith, E.J. 1993. Converting tissue plasminogen activator to a zymogen: a regulatory triad of Asp-His-Ser. Science 262:419-421.

McGeoch, D.J. 1985. On the predictive recognition of signal peptide sequences. Virus Research 3:271-286.

Meola, R.W., and Petralia, R.S. 1980. Juvenile hormone induction of biting behaviour in Culex mosquitoes. Science 209:1548-1550.

Mitchell, C.J., 1983. Differentiation of host-seeking behavior from blood-feeding behavior in overwintering Culex pipiens (Diptera: Culicidae) and observations on gonotrophic dissociation. Journal of Medical Entomology 20:157-163.

Mitchell, C.J., and Briegel, H. 1989a. Fate of the blood meal in force-fed, diapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 4:332-341.

Mitchell, C.J., and Briegel, H. 1989b. Inability of diapausing Culex pipiens (Diptera:Culicidae) to use blood for producing lipid reserves for overwinter survival. Journal of Medical Entomology 26:318-326.

Modig, C., Olsson, P.E., Barasoain, I., de Ines, C., Andreu, J.M., Roach, M.C., Luduena, R.F., and Wallin, M. 1999. Identification of beta (III)- and beta (IV)-tubulin isotypes in cold-adapted microtubules from Atlantic cod (Gadus morhua): Antibody mapping and cDNA sequencing. Cell Motility and the Cytoskeleton 42:315-330.

175 Modig, C., Wallin, M., and Olsson, P.E. 2000. Expression of cold-adapted beta-tubulins confer cold-tolerance to human cellular microtubules. Biochemical and Biophysical Research Communications 269:787-791.

Moll, R.M., Romoser, W.S., Modrzakowski, M.C., Moncayo, A.C., and Lerdthusnee, K. 2001. Meconial peritrophic membranes and the fate of midgut bacteria during mosquito (Diptera : Culicidae) metamorphosis. Journal of Medical Entomology 38:29-32.

Muller, H-M, Catteruccia, F., Vizioli, J., Della Torre, A., and Crisanti, A. 1995. Constitutive and blood meal-induced trypsin genes in Anopheles gambiae. Experimental Parasitology 81:371-385.

Nakai, K., and Kanehisa, M. 1992. A knowledge base for predicting protein localization in eukaryotic cells. Genomics 14:897-911.

Naora, H., 1999. Involvement of ribosomal proteins in regulating cell growth and apoptosis: translational modulation or recruitment for extraribosomal activity? Immunology and Cell Biology 77:197-205.

Nasci, R.S., Savage, H.M., White, D.J., Miller, J.R., Cropp, B.C., Godsey, M.S., Kerst, A.J., Bennett, P., Gottfried, K., and Lanciotti, R.S. 2001. West Nile virus in overwintering Culex mosquitoes, New York City, 2000. Emerging Infectious Diseases 7:742-744.

Niu, L.L., and Fallon, A.M. 2000. Differential regulation of ribosomal protein gene expression in Aedes aegypti mosquitoes before and after the blood meal. Insect Molecular Biology 9:613-623.

Noriega, F.G., Wang, X., Pennington, J.E., Barillas-Mury, C.V., and Wells, M.A. 1996. Early trypsin, a female-specific midgut protease in Aedes aegypti: Isolation, amino- terminal sequence determination, and cloning and sequencing of the gene. Insect Molecular Biology 26:119-126.

Oda, T. 1968. Studies on the follicular development and overwintering of the house mosquito, Culex pipiens pallens in Nagasaki area. Tropical Medicine 10:195-216.

Oktyabrskaya, T.A., Astakhova, N.A., and Boiko, L.P. 1965. Materials on the species composition, biology and ecology of mosquitoes observed near Moscow. Meditsinskaia Parazitologiia i Parazitarnye Bolezni 34:510-513.

176 Onyeka, J.O.A., Boreham, P.F.L., 1987. Population studies, physiological state and mortality factors of overwintering adult populations of females of Culex pipiens L. (Diptera, Culicidae). Bulletin of Entomological Research 77, 99-112.

Ouellet, F., Carpentier, E., Cope, M.J.T.V., Monroy, A.F., and Sarhan, F. 2001. Regulation of a wheat actin-depolymerizing factor during cold acclimation. Plant Physiology 125:360-368.

Paluh, J.L., Nogales, E., Oakley, B.R., McDonald, K., Pidoux, A.L., and Cande, W.Z. 2000. A mutation in gamma-tubulin alters microtubule dynamics and organization and is synthetically lethal with the kinesin-like protein Pkl1p. Molecular Biology of the Cell 11:1125-1239.

Pucciarelli, S., Ballarini, P., and Miceli, C. 1997. Cold-adapted microtubules: Characterization of tubulin posttranslational modifications in the Antarctic ciliate Euplotes focardii. Cell Motility and the Cytoskeleton 38:329-340.

Pucciarelli, S., and Miceli, C. 2002. Characterization of the cold-adapted alpha-tubulin from the psychrophilic ciliate Euplotes focardii. Extremophiles 6:385-389.

Quan, G.X., Kanda, T., and Tamura, T. 2002. Induction of the white egg 3 mutant phenotype by injection of the double-stranded RNA of the silkworm white gene. Insect Molecular Biology 11:217-222.

Raikhel, A.S., and Lea, A.O. 1983. Previtellogenic development and vitellogenin synthesis in the fat-body of a mosquito- An ultrastructural and immunocytochemical study. Tissue and Cell 15:281-300.

Readio, J., Chen, M., and Meola, R. 1999. Juvenile hormone biosynthesis in diapausing and nondiapausing Culex pipiens (Diptera: Culicidae). Journal of Medical Entomology 36:355-360.

Readio, J., and Meola, R. 1985. Two stages of juvenile hormone-mediated growth of secondary follicles in Culex pipiens. Journal of Insect Physiology 31:559-562.

Readio, J., Peck, K., Meola, R., and Dahm, K.H. 1988. Corpus allatum activity (in vitro) in female Culex pipiens during adult life cycle. Journal of Insect Physiology 34:131-135.

Reisen, W.K., Kramer, L.D., Chiles, R.E., Wolfe, T.M., and Green, E.N. 2002. Simulated overwintering of encephalitis viruses in dipausing female Culex tarsalis (Diptera: Culicidae). Journal of Medical Entomology 39:226-233.

177 Reiter, P., Spielman, A, Pollack, R., Shah, S., Reddy, M., Lepore, T.J., and Gubler, D. 2002. Evidence for vertical transmission of West Nile virus to overwintering Culex pipiens mosquitoes. Program and abstracts of the 51st annual meeting of the American Society of Tropical Medicine and Hygiene 67:247.

Reynaud, E., Bolshakov, V.N., Barajas, V., Kafatos, F.C., and Zurita, M. 1997. Antisense suppression of the putative ribosomal protein S3A gene disrupts ovarian development in Drosophila melanogaster. Molecular and General Genetics 256:462- 467.

Rinehart, J.P., Cikra-Ireland, R.A., Flannagan, R.D., and Denlinger, D.L. 2001. Expression of ecdysone receptor is unaffected by pupal diapause in the flesh fly, Sarcophaga crassipalpis, while its dimerization partner, USP, is downregulated. Journal of Insect Physiology 47:915-921.

Rinehart, J.P., Yocum, G.D., and Denlinger, D.L. 2000. Developmental upregulation of inducible hsp70 transcripts, but not the cognate form, during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 30:515-521.

Rivers D.B., and Denlinger D.L. 1995. Venom-induced alterations in fly lipid metabolism and its impact on larval development on the ectoparasitoid Nasonia vitripennis (Walker) (Hymenoptera, Pteromalidae). Journal of Invertebrate Pathology 66:104-110.

Rivers D.B., and Denlinger D.L. 1994. Redirection of metabolism in the flesh fly, Sarcophaga bullata, following envenomation by the ectoparasitoid, Nasonia vitripennis. Journal of Insect Physiology 40:121-127

Romoser, W.S., Moll, R.M., Moncayo, A.C., and Lerdthusnee, K. 2000. The occurrence and fate of the meconium and meconial peritrophic membranes in pupal and adult mosquitoes (Diptera : Culicidae). Journal of Medical Entomology 37:893-896.

Rorat, T., Irzykowski, W., and Grygorowicz, W.J. 1997. Identification and expression of novel cold induced genes in potato (Solanum sogarandium). Plant Science 124: 69-78.

Salt, R.W. 1961. Principles of insect cold-hardiness. Annual Review of Entomology 6:55-74.

Sanburg, L.L., and Larsen, J.R. 1973. Effect of photoperiod and temperature on ovarian development in Culex pipiens pipiens. Journal of Insect Physiology 19:1173-1190.

Santiago, J., and Sturgill, T.W. 2001. Identification of the S6 kinase activity stimulated in quiescent brine shrimp embryos upon entry to preemergence development as p70 ribosomal protein S6 kinase: Isolation of Artemia franciscana p70 (S6K) cDNA.

178 Biochemistry and Cell Biology-Biochemie et Biologie Cellulaire. 79:141-152.

Saunders, D. S. 2002. Insect Clocks, 3rd Edition. Elsevier, New York.

Seddon, W.L., and Prosser, C.L. 1997. Seasonal variations in the temperature acclimation response of the channel catfish, Ictalurus punctatus. Physiological Zoology 70:33-44.

Service, M.W. 1968. Observations on the ecology of some British mosquitoes. Bulletin of Entomological Research 59:161-194.

Schatten H., Chakrabarti A., and Hedrick J. 1999. Centrosome and microtubule instability in aging Drosophila cells. Journal of Cellular Biochemistry 74:229-241.

Spielman, A. 1964. Studies on autogeny in Culex pipiens populations in nature. I. Reproductive isolation between autogenous and anautogenous populations. American Journal of Hygiene 80:175-183.

Spielman, A. 1971. Studies on autogeny in natural populations of Culex pipiens. II. Seasonal abundance of autogenous and anautogenous populations. Journal of Medical Entomology 8:555-561.

Spielman, A. 1974. Effect of synthetic juvenile hormone on ovarian diapause of Culex pipiens mosquitoes. Journal of Medical Entomology 11:223-225.

Spielman, A, and Wong, J. 1973a. Studies on autogeny in natural populations of Culex pipiens. III. Midsummer preparation for hibernation in anautogenous populations. Journal of Medical Entomology 10:319-324.

Spielman, A., and Wong, J., 1973b. Environmental control of ovarian diapause in Culex pipiens. Annals of the Entomological Society of America 66:905-907.

Swellengrebel, N.H. 1929. La dissociation des fonctions sexuelles et nutritives (dissociation gono-trophique) d’ Anopheles maculipennis comme cause du paludisme dans les Pays-Bas et ses rapports avec “l’infection domiciliaire”. Annales de l'Institut Pasteur 43:370-1389.

Tammariello, S.P., and Denlinger, D.L. 1998a. Cloning and sequencing of proliferating cell nuclear antigen (PCNA) from the flesh fly, Sarcophaga crassipalpis, and its expression in response to cold shock and heat shock. Gene 215:425-429.

179 Tammariello, S.P., and Denlinger, D.L. 1998b. G0/G1 cell cycle arrest in the brain of Sarcophaga crassipalpis during pupal diapause and the expression pattern of the cell cycle regulator, proliferating cell nuclear antigen. Insect Biochemistry and Molecular Biology 28:83-89.

Tekle, A. 1960. The physiology of hibernation and its role in the geographical distribution of populations of the Culex pipiens complex. American Journal of Tropical Medicine and Hygiene 9:321-330.

Vinogradova, E.B. 2000. Culex pipiens pipiens mosquitoes: taxonomy, distribution, ecology, physiology, genetics, applied importance and control. Pensoft Publishers, Sofia. pp. 46-115.

Von Heijne, G. 1986. A new method for predicting signal sequence cleavage sites. Nucleic Acid Research 14:4683-4690.

Walsh, K.A., and Wilcox, P.E. 1970. Serine proteases. Methods in Enzymology 19:64- 108.

Warshel, A., Naray-Szabo, G., Sussman, F., and Hwang, J.K. 1989. How do serine proteases really work? Biochemistry 28:3629-3637.

Washino, R.K. 1977. The physiological ecology of gonotrophic dissociation and related phenomena in mosquitoes. Journal of Medical Entomology 13:381-388.

Watts, D.M., Thompson, W.H., Yuill, T.M., DeFoliart, G.R., and Hanson, R.P. 1974. Overwintering of La Crosse virus in Aedes triseriatus. American Journal of Tropical Medicine and Hygiene 23:694-700.

Wilson, T.G. 2003. Methoprene-tolerant, a bHLH-PAS gene essential for insect endocrinology. In: Crews, S.T. (Ed.), PAS proteins: regulators and sensors of development and physiology. Kluwer Academic, Boston, pp. 109-132.

Wilson, T.G., and Ashok, M. 1998. Insecticide resistance resulting from an absence of target-site gene product. Proceedings of National Academy of Sciences USA 95:14040- 14044.

Yan, Y.S., Salazar, C.E., Reese-Stardy, S.R., Cornel, A., Gorman, M.J., Collins, F.H., and Paskewitz, S.M. 1997. Cloning and characterization of a serine protease from the human malaria vector, Anopheles gambiae. Insect Molecular Biology 6:385-395.

180 Yocum, G.D. 2001. Differential expression of two HSP70 transcripts in response to cold shock, thermoperiod, and adult diapause in the Colorado potato beetle. Journal of Insect Physiology 47:1139-1145.

Yocum, G.D. Joplin, K.H., Denlinger, D.L., 1998. Upregulation of a 23 kDa small heat shock protein transcript during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 28:677-682.

Zamore, P.D., Tuschl, T., Sharp, P.A., and Bartel, D.P. 2000. RNAi: double-stranded RNA directs the ATP-dependent cleavage of mRNA at 21-23 nucleotide intervals. Cell 101:25-33.

Zurita, M., Reynaud, E., and Kafatos, F.C. 1997. Cloning and characterization of cDNAs preferentially expressed in the ovary of the mosquito Anopheles gambiae. Insect Molecular Biology 6:55-62.

181