Louisiana State University LSU Digital Commons

LSU Master's Theses Graduate School

2004 Identification of bacterial pathogens causing panicle blight of rice in Louisiana Xianglong Yuan Louisiana State University and Agricultural and Mechanical College, [email protected]

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_theses Part of the Plant Sciences Commons

Recommended Citation Yuan, Xianglong, "Identification of bacterial pathogens causing panicle blight of rice in Louisiana" (2004). LSU Master's Theses. 2883. https://digitalcommons.lsu.edu/gradschool_theses/2883

This Thesis is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Master's Theses by an authorized graduate school editor of LSU Digital Commons. For more information, please contact [email protected]. IDENTIFICATION OF BACTERIAL PATHOGENS CAUSING PANICLE BLIGHT OF RICE IN LOUISIANA

A Thesis Submitted to the Graduate Faculty of the Louisiana State University and Agricultural and Mechanical College In partial fulfillment of the Requirements for the degree of Master of Science In The Department of Plant Pathology and Crop Physiology

By

Xianglong Yuan B.S., University of Liaoning, 1985 M.S., Shenyang Agriculture University, 1990 May 2004

ACKNOWLEDGMENTS

I wish to express my sincerest gratitude to Dr. M. C. Rush, my advisor, for giving me this opportunity to pursue my graduate studies in the Department of Plant Pathology and Crop Physiology at Louisiana State University. It was Dr. Rush that led me to enter the world of plant pathology, widening my field of vision in academia. I sincerely appreciate his guidance, advice, understanding, and patience throughout my time of study here, particularly in the research work and writing. His working attitude will necessarily be a beacon lighting my future voyage.

I greatly appreciate Dr. A. K. M. Shahjahan’s help throughout my research efforts. His generous, patient guidance in my laboratory research, and his many excellent suggestions made during the past 2.5 years have helped my research efforts greatly.

Special thank go to my committee members, Dr. D. E. Groth, Dr. S. D. Linscombe, and Dr. J. P. Jones for their suggestions on conducting my research and writing my thesis. It was their influence that greatly inspired me in developing my thesis.

I acknowledge all of Chinese students that helped me at this university. Their great and generous support helped me to overcome many difficulties in my life and during my studies in Baton Rouge, LA.

The most sincere appreciation belongs to my mom and dad. Their encouragement, love and care were the main source of power that prompted me to accomplish this thesis. I also express thankfulness to my other family members for their generous support.

ii TABLE OF CONTENTS

ACKNOWLEDGEMENTS ...... ii ABSTRACT...... v CHAPTER 1. INTRODUCTION AND LITERATURE REVIEW ...... 1 1.1 Disease Distribution and Economic Importance...... 1 1.2 Symptoms and Epidemiology...... 3 1.2.1 Symptoms ...... 3 1.2.2 Epidemiology...... 5 1.3 Detection and Identification Methods for Pathogenic on Rice ...... 7 1.4 Classification, Nomenclature, and Pathological Characteristics of and Related Genera ...... 8 1.4.1 Classification and Nomenclature ...... 8 1.4.2 Characteristics of spp...... 9 1.4.2.1 B. glumae ...... 9 1.4.2.2 B. Plantarii...... 10 1.4.2.3 The B. gladioli and B. cepacia Complex ...... 10 1.4.2.4 P. fuscovaginae, P. syringae, Acidovorax avenae ( P. avenae) ...... 11 1.5 Virulence...... 12 1.6 Disease Control...... 13 1.6.1 Chemical Control...... 13 1.6.2 Biological Control...... 13 1.6.3 Developing Resistant Cultivars and Lines Using Genetics and Biotechnology Techniques...... 13

CHAPTER 2. CLASSIFICATION, IDENTIFICATION, AND PATHOGENICITY OF BACTERIA ASSOCIATED WITH PANICLE BLIGHT IN RICE ...... 15 2.1 Materials and Methods...... 15 2.1.1 Purification of Isolates ...... 15 2.1.2 Identification of Isolates Using the BiologTM GN2 Microplate System ...... 16 2.1.3 Plant Materials ...... 17 2.1.4 Pathogenicity Tests……………………………...... 17 2.1.5 Functional Classification of the Identified Isolates ...... 21 2.2 Results and Discussion ...... 22 2.2.1 Purification and Identification of Isolates with the BiologTM GN2 System ...... 22 2.2.2 Pathogenicity Tests and Symptoms Produced ...... 38 2.2.3 Colony and Culture Characteristics of the Bacterial Pathogens ...... 47 2.2.4 Clustering of Bacterial Strains Isolated from Rice Showing Panicle Blight into Functional Groups ...... 50

iii 2.3 Summary and Conclusions ...... 56

LITERATURE CITED ...... 61 APPENDIX: BIBLIOGRAPHY OF RELATED LITERATURE...... 80 VITA...... 97

iv ABSTRACT

Four hundred and two bacterial isolates were isolated on the semi-specific S-PG medium from diseased rice tissues showing symptoms of panicle blighting. These isolates were purified using serial dilution in sterile water and replating on S-PG medium. A total of 420 single isolates were obtained. These isolates were subjected to pathogenicity tests on rice (Oryza sativa L. cv. Cypress). Based on these tests, 339 isolates were used in BiologTM tests and identified to the species level. Bacterial strains from 39 species in 16 genera were identified, including 52 isolates representing 15 Pseudomonas species and 261 isolates representing six Burkholderia species. The remaining 26 isolates included 14 other genera. Of 261 Burkholderia strains, 103 isolates were B. gladioli, 68 isolates were B. glumae, 60 isolates were B. multivorans, 25 isolates were B. plantarii, three isolates were B.cocovenenans and three isolates were B. vietnamiensis. The pathogenicity tests revealed that 69% or 234/339 isolates caused seedling infection, sheath rot, and/or panicle blighting. Most of the pathogenic strains were in the genera Burkholderia and Pseudomonas. The four most common species, B. glumae, B. gladioli, B. multivorans, and B. plantarii, comprised 90% of all of the pathogenic bacteria, suggesting that a complex of Burkholderia species were causing the panicle blight/sheath rot syndrome recently found in Louisiana. Five Pseudomonas species, with total of 16 isolates, were found to be associate with this disease, including two isolates of P. syringae pv zizanize, three isolates of P. fluorescens, three isolates of P. pyrrocinia, one isolate of P. spinosa and seven isolates of P. tolaasii. The symptoms of the disease on rice were only produced by the indicated bacterial strains. Symptoms included brown margined flag leaf sheath lesions, grain rot, sterile florets, grain discoloration, and leaf and sheath rot on inoculated plants. It was impossible to distinguish among Burkholderia species based on symptoms and colony morphology. Pathogenic Burkholderia strains produced a yellow pigment in King’s B medium. Avirulent strains did not produce this pigment. The isolates from rice were grouped by species and pathogenicity. It appeared that B. glumae and B. gladioli were the most common pathogenic species.

v CHAPTER 1 INTRODUCTION AND LITERATURE REVIEW

1.1 Disease Distribution and Economic Importance Panicle blighting has been an important sporadic problem in the southern rice production area of the United States for many years. A somewhat similar disease called “ear blight”, pecky rice, or grain discoloration has been attributed to fungal causal agents (Lee, 1992a; Lee, 1992b). This problem is characterized by discoloration of the grain and panicle branches, usually with distinct lesions, and many fungi have been described as causing this disease (Atkins, 1974; Ou, 1985; Lee, 1992b). A symptom commonly called “panicle blight” has been observed for many years in the southern rice producing areas of the United States. It was considered to be a disorder of undetermined cause as no pathogen was isolated from diseased tissues (LSU Agricultural Center, 1987; LSU Agricultural Center, 1999; Groth, et al., 1991). The disease/disorder was characterized as having panicles with brown or straw-colored discoloration of florets (Groth et al., 1991). Panicle branches remained green, without lesions or discoloration, and affected florets stopped developing or aborted. Affected panicles had few to all of the florets diseased. When the disease was severe the panicle remained upright as the grain did not fill. In 1996 a bacterial pathogen was recognized as the cause of the rice panicle blight syndrome (Rush, 1998; Shahjahan, 1998, 2000a, 2000b, and 2000c). Rush and Shahjahan revealed that (formerly Pseudomonas glumae) was the agent causing the bacterial panicle blight (BPB) disease in Louisiana and adjacent states. Surveying showed that the BPB occurred in 60 percent of Louisiana fields, suggesting that the bacterial pathogen was prevalent throughout the rice production areas of the state (Shahjahan, 2000b). The world wide distribution of this disease has been become clear as recent survey reports became available. Studies of the epidemiology, etiology and control of this disease have attracted the attention of rice plant pathologists from several countries. According to S.H. Ou (Ou, 1985) bacterial sheath rot and grain blighting on rice was first reported in Hungary (Klement, 1955). The disease was described as caused by Pseudomonas oryzicola (later shown by Visnyovszky et al. (1971) to be a synonym of

1 Pseudomonas syringae pv. syringae (Ou, 1985). According to Ou (1985), research on bacterial grain rot was initiated by Goto in Japan (Goto and Ohata, 1956; Goto et al.,1987). Grain rot in rice was reported to be caused by several bacteria, including Pseudomonas glumae (Burkholderia glumae), but only P. glumae caused seedling blight on inoculated plants (Goto et al., 1987). According to Hashioka (1969); Kurita and Tabei, 1967 named the pathogen of grain rot Pseudomonas glumae Kurita et Tabei. The temperature optimum for growth of P. glumae was reported as 30-35C with a range of 11-40C and a thermal death point at 70C (Kurita et al., 1964). Tanii, et al. (1974) isolated several bacteria from rice grains showing grain rot. They identified three Pseudomonas species, but not P. glumae. P. glumae was isolated from plants in nursery flats with seedlings showing bacterial seedling rot symptoms (Uematsu, et al., 1976). Since the 1980’s, the disease has been reported from several different countries in the world, including Korea (Jeong et al., 2003), Taiwan (Chien and Chang, 1987), Latin America (Zeigler et al., 1987; Zeigler and Alvarez, 1989a; 1989b; 1989c; 1990 and Zeigler, 1990), Vietnam (Trung et al., 1993), Philippines (Cottyn et al., 1996a; 1996b) and the Gulf of Mexico rice production area in the U.S. (Rush, et al.,1998; Shahjahan et al., 1998; 2000b; 2000c). It is reasonable to believe that this disease frequently occurs in rice around the world, particularly in tropical countries. It is not surprising as the pathogen is readily seedborne. Rice is considered to be one of the most important food crops in the world (Lu and Chang, 1980). It is clear that wide-spread development of a major rice disease would be disadvantageous to the world’s economy. The direct agronomic consequences of such an event would be a severe decrease in crop yield and stability of production. BPB has the potential to reduce the yield of rice as much as 75% in severely affected regions due to reduction in grain weight, floret sterility, inhibition of seed germination, reduction of stands, as well as the year-to-year transmission because of the seedborne nature of the pathogen (Trung, et al., 1993).

2 1.2 Symptoms and Epidemiology 1.2.1 Symptoms The reported symptoms for this disease syndrome were observed in Louisiana on the leaf sheath of seedlings, flag leaf sheath, and panicle florets. The disease was described as sheath rotting (brown, necrotic lesion with a distinct margin); leaf stripe; flag sheath browning (Figure 1); grain rot (Figure 2); spikelet discoloration (grayish or straw-colored, with bottom half of the floret darkening); sterile florets and/or seedling blighting ( Shahjahan, 2000b). Several bacterial species have been proposed as causing similar symptoms. The suspected pathogens are mainly distributed in two taxonomically and physiologically close genera, Pseudomonas and Burkholderia. B. glumae causes rice grain rot and seedling rot (Uematsu et al., 1976), sheath rot complex, and grain/sheath rot (Cottyn et al., 1996a, 1996b, and 2001). The bacterial stripe disease, with grain discoloration, was reported as caused by Pseudomonas avenae (Kadota and Ohuchi, 1983). Sheath rot and grain discoloration in Latin America was believed to be caused by Pseudomonas fuscovaginae and Pseudomonas glumae (Zeigler and Alvarez, 1987; Zeigler and Alvarez, 1989b). However, the different pathogens share similar symptoms among them. For instance, P. syringae, P. fluorescens and B. glumae caused similar early symptoms (Goto et al., 1987). Zeigler reported the symptoms produced by Pseudomonas fuscovaginae as extensive water-soaking and necrosis with poorly defined margins and glume discoloration before panicles emerge from boots. Infected grain varied from completely discolored and sterile to nearly symptomless with only small brown spots (Zeigler and Alvarez, 1987; Zeigler et al., 1987; Zeigler and Alvarez, 1989a; Zeigler and Alvarez, 1989c). Symptoms typically caused by B. glumae in Louisiana were panicle blighting with floret discoloration (with a gray-brown color), usually on the lower half of the developing grain, with a clear deep brown border followed by sterility or partial filling of the florets causing the panicles to stand erect (Shahjahan et al., 1998; Shahjahan et al., 2000b). It is clear that accurate identification of the pathogens and epidemiological surveys are essential before proposing practical control schemes.

3

Figure 1.1. Flag leaf sheath rot, caused by Burkholderia glumae on inoculated Cypress rice.

4

Figure 1.2. Panicle blighting and grain rot, caused by Burkholderia glumae, on inoculated Cypress rice.

1.2.2 Epidemiology Bacterial pathogens of plants exist ubiquitously in the plant’s ecosystems and are usually found in the air, soil, and water. Soil distributions of these pathogens are affected by soil type, pH value, plants cultivated, and weather conditions. It has been established that pathogenic bacteria exist on the phylloplanes of rice plants during the growing season (Matsuda and Sato, 1987; Hikichi, 1993a; Hikichi, 1993b; Hikichi et al., 1993; Otofuji et al., 1988; Tsushima et al., 1996), on or in rice seeds stored at room temperature during the winter (Tsushima et al.1987; Tsushimi et al., 1989), on weeds in the field and

5 in rice tissues from the previous crop buried in soil (Sogou and Tsuzaki, 1983). The disease can develop from inoculum in infected seeds from the previous year, in soil, and on weeds in the field. In the process of infection, host susceptibility, inoculum density, and climatic factors play the key roles (Tsushima et al., 1985; Tsushima et al., 1987; Tsushima and Naito, 1991; Tsushima et al., 1995a; Tsushima et al., 1996; Tsushima, 1996). This disease tends to break out under conditions of unusually high temperatures, especially at night, and frequent rains (Tushima et al., 1985; Mew, 1992; Zeigler and Alvarez, 1990). In 1995 and 1998 there were severe incidences of rice BPB in Louisiana as well as the other Southern rice production areas. Yield losses as high as 40% were observed in some fields (Shahjahan, 2000b). Record high temperatures were recorded during these seasons, with high temperatures extending into the night. High humidity levels at the flowering stages were reported to be particularly conductive to spikelet infections (Tsushima et al., 1995b). The most susceptible period for floret infections appears to be during panicle emergence and flowering. Inoculation during flowering gives the highest rates of floret infection and production of diseased spikelets (Shahjahan et al., 1998). In these tests using B. glumae, a high level of floret infection continued for up to 3 days after flowering in inoculation experiments. The infection rate of the pathogen also depends on inoculum density. Pot experiments conducted by spraying emerging panicles revealed that infection increased proportionally as the lognormal volumes of inoculum density (Tsushima et al., 1985). Based on their experiments, 102 to 104 cfu was proposed as the minimum inoculum density when spraying in the field and not less than1 cfu/ml was an appropriate dose for the injection method (Hikichi, et al. 1994). Pathogen cells present on leaf sheaths play an essential role in primary infection. Infection on the leaf sheath provides the primary source of inoculum to the emerging panicle (Tsushima et al., 1991 and Tsushima et al., 1996). We have also observed that that following injection of the pathogen into boots the first visible diseased tissue was always on the flag leaf sheath followed by panicle infection. When rice plants are developing from apparently healthy seeds, the uninfected flag leaf sheath forecasts that the panicle should be healthy. Therefore, the frequency and severity of infection of the

6 flag leaf sheath is an important indicator of the eventual infection on the panicle because the sheath is spatially close to panicles and the primary time of infection was at heading time (Tsushima, 1996; Tsushima et al., 1996). The primary infection site by B. glumae is apparently through the plumules (Hikichi, 1993a; Hikichi, 1995b). The bacterium enters the lemma and paleae through stomata, multiplies in intercellular spaces of the parenchyma (Tabei et al, 1989), and moves towards the surrounding healthy cells and tissues. The bacteria were detected in the epidermis, parenchyma, and sclerenchyma of glumes of naturally infected rice seeds using antiserum (Hikichi, 1993b). The long-distance movement of bacteria was accomplished via vascular systems. Bacteria were observed, by the author (X. Yuan), flooding out of vascular systems of severely infected rice by tissue sectioning techniques.

1.3 Detection and Identification Methods for Pathogenic Bacteria on Rice After evaluating the efficacy of the various methods, it appeared important to select an appropriate combination of methods to recognize the pathogens. A good identification scheme depends not only on developing a satisfactory resolution level of methods, but also on the group of bacteria studied (Welch, 1991). Apparently, it is difficult to differentiate pathogens that are closely related physiologically and taxonomically by the symptoms they produce and by their growth on selective media. A semi-selective medium for B. glumae was developed by Tsuschima et al., 1986; however, other Burkholderia species will also grow on the SP-G medium. Traditional biochemical and serological methods still have powerful advantages. Cottyn et al., 1996a and 1996b detected 204 pathogens strains distributed among seven species by using a comprehensive scheme including antiserum, API 20NE, BiologTM, and cellular fatty acid methyl ester-fingerprinting. The best differentiation was given by the BiologTM system (Cottyn et al., 1996b). Although immunological methods possess quiet high resolution, specific antibodies to some antigens are not readily obtained and this may limit the wide application of these methods. Recently, PCR-based DNA-typing methods based on both conservation and variance of rDNA genes among species and pathovars, and other molecular based techniques, have become popular due to the rapidity, accuracy, and relative simplicity of operation. These methods will undoubtedly become powerful tools

7 in the field of classification and identification of microorganisms (Furuya et al., 2002; Karpati and Jonasson, 1996; Mizuno et al., 1995; O’Callaghan et al., 1994; Pelt et al., 1999; Reiter et al., 2000; Tyler et al., 1995; Ura et al., 1998; Whitby et al., 2000a; Whitby et al., 2000b.

1.4 Classification, Nomenclature and Pathological Characteristics of Pseudomonas and Related Genera

1.4.1 Classification and Nomenclature It has long been known that some species in the genus Pseudomonas were plant pathogens. According to Burkholder (1949), as early as 1921 Pseudomonas gladioli was identified as the causal agent of flower rot of . Burkholder described a new bacterial pathogen causing sour skin and rots and proposed the bacterium as Pseudomonas cepacia n. sp. (Burkholder, 1949)., However it was not recognized that some Pseudomonas spp. could serve as severe human pathogens until the 1960’s. Recently, it has been confirmed that B. cepacia and B. gladioli can produce human cystic fibrosis disease (CF) and pulmonary disease (Baxter et al., 1997; Parke and Gurian- Sherman, 1998; Vandamme, 1996, Vandamme et al., 1997, and Vandamme et al., 2000). The genus Burkholderia was not established until 1992 when Yabuuchi proposed to transfer seven species previously in the Pseudomonas homology group II to the new genus Burkholderia (Yabuuchi et al., 1992). P. plantarii and P. glumae were transferred to this genus in 1994 (Urakami et al., 1994). Yabuuchi et al. reclassified two Burkholderia species, B. solanacearum and B. pickettii, and Alcaligenes eutrophus, into the novel genus Ralstonia (Yabuuchi et al., 1995). Until 2002, there were 28 species in the genus Burkholderia (Table 1.1). Many of the species were believed to be plant pathogens causing rice, tobacco, maize and other crop diseases. The established rice pathogens included P. avenae (Goto and Ohata, 1956), P. fuscovaginae (Zeigler, 1987, 1989, 1990; Rott, 1989 and 1991; Cottyn, 1996 and 2001), P. syringae, B. glumae (Goto and Ohata, 1956; Goto,1965 and Goto and Ohata,1987; Hikichi et al., 1993d, Hikichi et al.,1994, and HIkichi et al.,1995a; Tsushima et al., 1985, Tsushima et al.,1986, and Tsushima, 1991; Cottyn et al., 1996a, 1996b and 2001), B. plantarii (Azegami et al.

8 1985, Azegami et al.,1987 and Azegami et al.1988), B. gladioli (Cheng, et al., 2001), and B. vietnamiensis (Trung et al., 1993).

Table 1.1. Bacterial species described in the genus Burkholderia.*

B. andropogonis B. cepacia genomovar I B. anthina B. multivorans (formerly genomovar II) B. ambifaria B. cepacia genomovar III B. caledonica B stabilis (formerly genomovar IV) B. caribensis B. vietnamiensis (formerly genomovar V) B. caryophylli B. cepacia genomovar VI B. cocovenenans B. plantarii B. fungorum B. pseudomallei B. gladioli B. phenazinium B. glathei B. pyrrocinia B. graminis B. sacchari B. glumae B. thailandensis B. kururiensis B. ubonensis B. mallei B. vandii * Coenye et al., 2001; Estrada-de los Santos et al., 2001; Gillis et al., 1995; Vandamme, 1996; Vandamme et al.,1997; Vandamme et al., 2000; Yabuuchi et al., 1992.

1.4.2 Characteristics of Burkholderia spp. 1.4.2.1 B. glumae B. glumae was first recorded as a rice pathogen in the genus Pseudomonas (Goto and Ohata, 1956), and is considered the major bacterial agent causing seedling and grain rot in Japan. The typical symptoms were described as sheath brown rot (Cottyn et al., 1996a and Cottyn et al., 1996b), grain rot and/or discoloration (Cottyn et al., 1996; Goto and Ohata, 1956 and Goto et al., 1987; Tsushima et al., 1986; Mew et al., 1990; and Zeigler and Alvarez,1989a.), and seedling rot (Uematsu et al., 1976a and Uematsu et al 1976a; Tsushima et al., 1986). B. glumae is a nonfluorescent bacterium producing a yellow-green, water-soluble pigment on various media. The organism is rod-shaped, with 1-3 polar flagella. The colony is grayish white or yellow due to the pigment. Arginine dehydrolase, oxidase reaction and nitrate reduction express a negative reaction. Lecithinase production and the utilization of L-arginine and inositol are positive.

9 1.4.2.2 B. plantarii The existence and pathogenicity of B. plantarii was first recognized in northern rice production areas in Japan (Azegami et al., 1987). The pathogen was found to be distributed on the weeds in fields and in seed stored at room temperature. B. plantarii is often isolated when isolating B. glumae, suggesting that they may share a similar transmission path and life cycle. They also show similar symptoms on rice (Azegami et al., 1987 and Azegami et al., 1988; Tanaka et al., 1994). However, so far this organism has only been reported from two countries. Azegami et al., (1987) reported that B. plantarii produced a reddish-brown pigment on media. 1.4.2.3 The B. gladioli and B. cepacia complex

B. gladioli and B. cepacia are organisms with wide distribution, and have been isolated from soil, water, plant roots, plant rhizospheres, and the lungs of CF patients (Holmes et al., 1998; Segonds et al., 1999; Vandamme et al., 1997). For instance, B. multivorans (B. cepacia genomovar III) was found to be a common plant-associated bacterium (Balandreau et al., 2001), and can colonize rice, corn, maize, pea, sunflower, and radish. In natural niches, B. gladioli and B. cepacia regularly coexist with each other, with B. glumae, and/or occur in the form of hybrids of B. gladioli and B. cepacia. These hybrids may have a significant pathogenic importance (Baxter et al., 1997). Cheng et al. (2001) revealed that B. gladioli was one of the causal agents of rice BPB in Louisiana. In Japan, the symptoms on rice infected with B. gladioli were reported as necrosis or chlorotic spots on the leaf (Furuya et al., 1997). In addition, necrosis with water-soaked lesions on tobacco leaves was recorded (Furuya et al., 1997), revealing that this bacterium has a wide host range. In addition to being a human pathogen, B. cepacia has been reported to promote maize growth (Bevivino et al., 1998), to enhance crop yields (Chiarini et al., 1998; Tabacchioni et al., 1993), to suppress many soilborne plant pathogens (Bevivino et al., 1998; Hebbar et al., 1998; McLoughlin et al., 1992), and to degrade diverse pesticides (Daubaras et al., 1996; McLouglin et al., 1992). B. cepacia has demonstrated an effective role in biological control of soilborne, foliar, and post-harvest plant diseases. However, it has recently been recognized that this application may pose risks to human health and

10 create environmental problems (Holmes et al., 1998). Aspects of symptomology and epidemiology on rice concerned with this organism still remained little reported. 1.4.2.4 P. fuscovaginae, P. syringae and Acidovorax avenae (P. avenae) Another widely observed pathogenic bacterium on rice is the nonfluorescent bacterium P. fuscovaginae, which produces symptoms similar to those produced by the previously mentioned bacterial pathogens on rice, and has a worldwide distribution. Ziegler and Alvarez, (1987a) reported that the bacteria P. fuscovaginae and P. syringae pv. syringae caused sheath brown rot and grain and sheath discoloration on rice in Latin America. P. fuscovaginae, P. syringae pv. syringae, A. avenae, and B. glumae were all identified from rice in Colombia and Central America (Zeigler and Alvarez, 1987b; Ziegler and Alvarez, 1989a; Ziegler and Alvarez, 1989b; Ziegler, 1990).. Of them, P. fuscovaginae and A. avenae were the most common pathogens (Zeigler, 1990). Rott et al., (1989) showed that there were intra-species differences among the native strains of P. fuscovaginae in Madagascar and reference strains by serological techniques. This suggested that epidemiology studies and surveys could be complicated by strain differences. P. syringae was first detected by Kuwata (1985) on rice in Japan where it caused the halo blight disease on rice. This new rice bacterial disease had symptoms of small brown foci surrounded by large, yellow halos on leaf blades (Kuwata, 1985). The disease on rice caused by A. avenae (formerly P. avenae) shows brown stripes on leaf blades and sheaths extending over the whole leaf blade in the severest infections (Ou, 1972; Kodota and Ohuchi, 1983; Shakya and Chung, 1983; Shakya et al., 1985; Tominaga et al., 1983). The pathogen has a wide host range, causing diseases on barley, maize, millet, oats, and rice (Kadota, 1996). The research of Cottyn et al. (1996a and 1996b) demonstrated the existence and pathogenicity of Philippines strains of P. fuscovaginae, P. syringae pv. syringae, A. avenae, and B. glumae. However, his reports simultaneously pointed out that none of these bacteria could be associated with distinct symptoms and B. glumae could not be differentiated from P. fuscovaginae, P. syringae, and A. avenae based on symptoms, suggesting that the limitations of this method indicated that a comprehensive detection scheme and more advanced techniques appeared to be necessary.

11 1.5 Virulence Sato et al. (1989) isolated two bioactive substances from a culture filtrate of B. glumae, identified as toxoflavin and fervenulin, and indicated that these phytotoxins could induce chlorotic spots on detached rice leaves. Toxoflavin, a yellow crystalline solid, first isolated in 1933 from B. cocovenenans (Buckle et al., 1990), reduces the elongation of sprouts and roots of rice seedlings (Suzuki et al., 1998; Yoneyama et al. 1998). When rice panicles were treated with the substance a brown band on the palea and lemma, a characteristic symptom of bacterial grain rot, was observed (Liyama et al., 1995). This phytotoxin was also detected in B. glumae, several strains of B. gladioli (Suzuki et al, 1998), and B. cocovenenans, not only from culture filtrates, but also from rice seedlings infected with the bacteria (Liyama et al. 1994; Liyama et al., 1995; Liyama et al., 1998). Wang compared the virulent effects of pigment producing strains and non- producing strains of B. glumae on rice seedlings and potato tuber slices. The results indicated that all of the 28 pigment producing strains used were virulent to rice seedlings, whereas all of non-producing strains were avirulent with two exceptions (Wang et al., 1991). Liyama et al.(1994 and 1995) also demonstrated that virulence to rice was closely correlated with the toxin after making comparisons between the phytotoxin producing strains and non-producing strains. By using a transposon mutagenesis technique that destroys the genes encoding toxin-biosynthesis or virulence-related pathways, transposon mutants of B. glumae were obtained. The results showed that the non-toxin producing mutants were unable to induce rice bacterial rot disease (Yoneyama et al., 1998). In addition, there is a report indicating that the virulence of B. glumae declined rapidly with continued transfers on PSA medium, suggesting that the loss of virulence occurs from subculture to a subculture (Tsushima et al., 1991). The virulence of B. plantarii was associated with production of the phytotoxin tropolone, a water-soluble substance, which is produced by some Pseudomonas and Burkholderia spp. including B. glumae. This compound is bioactive to plants, fungi, and bacteria (Lindberg et al., 1980; Lindberg,1981). This toxin is a non-benzenoid aromatic compound with a seven-member ring and gives red crystals of ferric tropolone when

12 chelated with iron. It is not clear whether the reddish-brown pigment on B. plantarii growth media (Azegami et al., 1985 is associated with tropolone production.

1.6 Disease Control 1.6.1 Chemical Control Many bactericides are able to effectively control or suppress the occurrence of seedling rot and panicle rot caused by plant pathogenic Burkholderia spp. including antibiotics, copper, and copper-containing compounds (Katsubi and Takeda, 1998; Shahjahan et al., 2000b). Oxolinic acid, a synthetic bactericide developed from quinoline derivatives, inhibits disease development by Gram negative bacteria. This compound was highly efficacious for the control of this rice disease, either as seed treatments or foliar sprays. It exhibited both preventive and curative effects (Hikichi, 1993). A study of the efficacy of oxolinic acid, with FITC-conjugated antibody and fluorescence microscopy, demonstrated that only 3% of seedlings expressed measurable symptoms after treatment, compared with 92% of seedlings from untreated seeds (Hikichi, 1995). 1.6.2 Biological Control The ability of some avirulent strains of Burkholderia to restrict the development of rice grain rot has been confirmed. Spraying rice panicles with a mixed suspension of a virulent strain of B. glumae and an avirulent strain of B. gladioli almost completely controlled the occurrence of the disease in pot and field experiments (Miyagawa and Takaya, 2000). Similar efficacy was expressed when the seedlings were sprayed with the avirulent B. gladioli strain prior to inoculation with the B. glumae strain; whereas no suppression was observed when the order of inoculation was reversed (Miyagawa, 2000a). Similar results were also observed using avirulent strains of B. glumae (Furuya, et al., 1991) and of B. plantarii (Ohno, et al., 1992), indicating that pre-treatment of seeds with avirulent strains is an effective biological control method.

1.6.3 Developing Resistant Cultivars and Lines using Genetics and Biotechnology Techniques Developing resistant cultivars and lines using genetic and biotechnological means is one of the major trends in plant disease control. Transforming rice with the oat thionin gene gave a measurable resistance against grain rot caused by B. plantarii (Iwai, et al.,

13 2002). Due to a high level of accumulation of thionin in cell walls, this transgenic rice was protected against the attack of the bacterium. The stained and marked bacteria distributed only on the surface of stomata, whereas, all of the inoculated wild-type rice Seedlings, developed from seeds germinated after treatment with bacterial suspension, were wilted and showed characteristic blight symptoms (Iwai, et al., 2002). Rush revealed that B. cepacia and B. gladioli were common inhabitants of the phylloplane of rice in Louisiana (Rush et al., 1998). Previous studies showed that B. glumae (Shahjahan, 2000a and 2000b) and B. gladioli (Cheng, et al, 2001) were the causal agents of BPB on rice in Louisiana. However, it is still unclear whether other bacterial pathogens exist that can cause this disease. In this study more than 400 bacterial isolates obtained from infected rice panicles collected from rice fields in Louisiana, were used to conduct pathogenicity tests and the cultures were grouped into functional, biochemically characteristic strains for proper identification with the BiologTM system.

14

CHAPTER 2

CLASSIFICATION, IDENTIFICATION, AND PATHOGENICITY OF BACTERIA ASSOCIATED WITH BACTERIAL PANICLE BLIGHT IN RICE

Several hundred strains of bacteria isolated from diseased tissues of rice collected in Louisiana and showing BPB symptoms were provided by M.C. Rush and A.K.M. Shahjahan for further purification, for identification to species, to conduct pathogenicity tests to determine which strains were pathogens, and to determine the virulence of pathogenic isolates. The objectives of this study were to: 1. Replate by serial dilution and streaking all of the primary cultures provided by Dr. A.K.M. Shahjahan. Obtain single bacterium cultures from the primary isolates, and preserve these cultures for further study.

2. Identify all single bacterium cultures to species using the Biolog TM GN MicroPlate System (Jones, 1993).

3. Conduct pathogenicity tests on rice seedlings and panicles in the boot by injection of bacterial suspensions of all of the identified bacterial cultures..

4. Identify bacterial isolates pathogenic on rice and quantify virulence of the pathogenic isolates.

2.1 Materials and Methods

2.1.1 Purification of Isolates Bacterial isolates isolated previously on the selective medium S-PG (Tsushima et al., 1986), from diseased grains or rice sheaths (Oryza sativa L.) in samples collected from Louisiana rice production areas, were used in this study. All of the bacterial isolates were subcultured on King’s Medium B with agar (KBA) (20g protease peptone #3, 1.5g . KH2PO4, 1.5g MgSO4 7 H2O, 15 g Bacto-agar, and 15 ml glycerol). All isolates were purified by serial dilution with sterile distilled water and plating on KBA medium. Plates

15 were incubated at 30C for 48 hr. The bacterial colonies on these plates were examined for their morphological characteristics. Single colonies, with different cultural characteristics, from each culture plate were touched with a flamed bacteriological loop and streaked on KBA medium, incubated at 30C for 48 hr, and then stored in microcentrifuge tubes at -70oC in 30% glycerol. Each isolate was given a culture number.

Figure 2.1.1. Typical colonies of Burkholderia sp. on S-PG medium.

Bacterial strains were removed from the freezer as needed for identification or pathogenicity tests, plated on KBA medium, and incubated at 30C for 48 hr. The development or failure of development of yellow pigment in the medium was noted for selected B. glumae and B. gladioli isolates for comparison of pigment production with pathogenicity of these cultures (Figures 2.25, 2.26, and 2.27).

2.1.2 Identification of Isolates Using the BiologTM GN2 Microplate System Test isolates were removed from the freezer, plated on KBA medium, and incubated at 30C for 48 hr. Bacteria were subjected to Gram staining tests (Brock et al, 1994). Gram-negative cultures were transferred onto BUG Agar (BUG Agar with 5% sheep blood), and incubated at 30C for 24hr. Bacterial cells were removed from the plates with sterile cotton-tipped swabs, suspended in a GN/GP-IF containing 18 to 20 ml of sterile “gelling” inoculating fluid (0.40% NaCl, 0.03% Pluronic F-689, 0.02% Gellan Gum). Cell densities were adjusted with a turbidimeter to 52+/-4% transmittance. The BiologTM GN2 MicroPlate test panels (Biolog, Hayward, CA) were inoculated and incubated at 30C for 24-48 hr. Results were analyzed with BiologTM GN2 database

16 version 4.20 to determine the identity of each isolate. Data on each isolate was printed out and kept in ring binders.

2.1.3 Plant Materials Ten to 15 rice seeds of the variety Cypress were planted in each 15-cm plastic pot with soil consisting of a mixture of 2:1:1 soil, sand, and peat moss, respectively. The plants were grown under natural light with a day temperature of 33C – 41C and 75% - 95% relative humidity in a greenhouse on the Louisiana State University campus in Baton Rouge, LA. Fertilizer was applied in the form of 1 tablespoon of slow release OsmocoteTM with NPK of 19-5-8 for general plant maintenance. After seedlings were established they were thinned to a population of 5-8 seedlings per plot (Figure 2.1.2).

2.1.4 Pathogenicity Tests Bacterial inoculum grown on KBA medium incubated at 30C for 48 hr were harvested with a sterile cotton swab and suspended in a vial containing 0.9ml of sterile distilled water (ca. 107-108 cfu/ml). All seedling inoculations were by injection with sterile B-D TM 1ml syringes using B-DTM 23G1 needles. Inoculation preparation and injection of bacteria into plants was conducted using sterile technique. The person inoculating plants wore rubber gloves and rubbed his hands in the gloves with 95% ethanol. Three to five Cypress rice seedlings in the same pot were inoculated on stems before the plants began tillering. Inoculum consisted of 0.5ml of each bacterial isolate per stem. The inoculated seedlings were maintained in the greenhouse. After 2 weeks the disease reaction was determined on each inoculated plant. Inoculated seedlings were rated with a four category scale. In this scale 0 = no symptoms produced after inoculation (Figure 2.1.3), + = slight browning around the injection site, ++ = a brown lesion 1-2 cm in diameter or spreading up and down the stem from the injection site with a distinct darker brown lesion, +++ = brown lesion spreading up the stem from the injection site and with the leaf blade on the inner leaf yellowing or blighted and turning necrotic (Figure 2.1.4). Plants with a “0” rating were listed as NP or non-pathogenic in the data tables.

17

Figure 2.1.2. Four-leaf-stage Cypress rice seedlings ready for inoculation in pathogenicity tests.

Figure 2.1.3. Inoculated seedling showing “0”or NP reaction.

18

Figure 2.1.4. Inoculated seedlings showing symptoms typical of a +++ rating (picture courtesy of A.K.M. Shahjahan).

Cypress plants produced as above were fertilized again at the maximum tillering stage with 1 tablespoon of OsmacoatTM, and inoculated just before panicle emergence (Figure 2.1.5) with 1.0ml of bacterial suspension injected into the boot using B-DTM 3ml syringes with B-DTM 23G1 needles. Inoculated panicles were rated 3-4 weeks after inoculation and emergence with a similar rating where 0 = no signs of disease on the panicle or florets, + = some browning of the florets with most grain filling normally; ++ = browning on most florets and some grain not filling, lesions formed on sheaths at inoculation points, +++ = panicles poorly emerging, or fully emerging and florets discolored and brown with many florets sterile and panicles tending to remain upright due to lack of grain filling ( Figure 2.1.6). An elongated lesion usually developed on the flag-leaf sheath at the inoculation point with virulent strains (Figure 2.1.7).

19

Figure 2.1.5. Plants at the boot stage ready for inoculation.

Figure 2.1.6. Typical +++ reaction for panicles inoculated with a virulent Burkholderia strain.

20

Figure 2.1.7. Flag leaf sheath rot caused by Burkholderia glumae on inoculated Cypress rice.

All final pathogenicity ratings and evaluations were made during July to October in the greenhouse with day-time temperatures ranging from 37-43C and relative humidity ranging from 75-95%. The high temperatures and humidity favored disease development. Please note that seedling reactions made in the summer of 2002 were not rated with the four reaction scale, but were assigned a not pathogenic (NP) or pathogenic (P) rating.

2.1.5 Functional Classification of the Identified Isolates The isolates of the four most common bacteria, B. glumae, B. gladioli, B. multivorans, and B. plantarii, were clustered into functional groups based on two clustering systems used independently. One was based on the four unit scale of severity and the other was a clustering system based upon the similarity index of an isolate to a typical strain in the BiologTM database. This was done to evaluate the correlation between the strains and identification accuracy. An isolate was separated into a group of high-similarity index or a group of low-similarity indexes based on the BiologTM identification index.. The threshold value of high or low similarity for B. plantarii was assigned to be 0.6. Considering larger number of isolates in B. glumae, B. multivorans

21 and B. gladioli, a weight was added to the threshold values of each of these species. The threshold value of high-similarity or low-similarity for B. multivorans and B. glumae was shifted up to 0.65 and the index was shifted to 0.7 for B. gladioli. By combining both clustering systems, 8 groups of bacteria in B. glumae, B. gladioli, B. multivorans, and B. plantarii were produced; that is, high virulence + high similarity index, high virulence + low similarity index, low virulence + high similarity index, low virulence + low similarity index, not virulent + high similarity index, not virulent + low similarity index, moderate virulence + high similarity index, and moderate virulence + low similarity index. The results of classifications were represented in Tables 2.2.7, 2.2.8, 2.2.9, and 2.2.10. The distribution characteristics of isolates of B. gladioli and B. glumae after such classifications were illustrated in Figures 2.2.24, 2.2.25, and 2.2.26.

2. 2 Results and Discussion 2.2.1 Purification and Identification of Isolates with the BiologTM GN2 System Based upon the previous isolation of bacteria using the S-PG semi-selective medium, a total of 402 bacterial isolates were isolated and further purified using serial dilution in sterile water. This procedure gave 420 bacterial isolates. All of these isolates were subjected to pathogenicity tests based on inoculating Cypress seedlings at the four- leaf stage of growth and panicles in the boot just before emergence. Of these isolates, 339 were identified to the species or pathovar level using the BiologTM GN2 system (Figures 2.2.1, 2.2.2, 2.2.3, 2.2.4, and 2.2.5). The bacteria belonged to 39 species distributed among 16 genera (Table 2.2.1). Burkholderia was the most common genus, giving 262 isolates. One hundred and three isolates were B. gladioli, 68 isolates were B. glumae, 25 isolates were B. plantarii, 60 isolates were B. multivorans, 3 isolates were B. cocovenenans, and 3 isolates were B. vietnamiensis. Fifty two isolates were placed in the genus Pseudomonas; and another 26 isolates were distributed among 14 other genera (Table 2.2.1) Pathogenicity tests on the variety Cypress confirmed that the bacteria that were pathogenic consisted of 234 of 339 strains tested or 65% of the isolates (Table 2.2.2). Most of the pathogenic strains were in the two genera Burkholderia and Pseudomonas. The four most commonly isolated species, B. glumae, B. gladioli, B. multivorans, and B.

22 plantarii, comprised 90% of the pathogenic bacterial strains, indicating that a complex of Burkholderia species may be the causal agents of the BPB and sheath rot disease recently identified in Louisiana (Table 2.2.2). This is apparently the first report of B. multivorans and P. tolaasii ( Table 2.2.3) as bacterial pathogens of rice. Also, this is the first time that B. gladioli and B. plantarii were clearly determined to be pathogens of rice in Louisiana. Among the B. glumae isolates tested, 91% were pathogenic on Cypress, suggesting that there is widespread pathogenicity among natural strains of this species (Table 2.2.2). Three isolates were tentatively identified as B. cocovenenans and three as B. vietnamiensis. If these are valid identifications, this suggests that these species only rarely infect rice in Louisiana. No isolates of P. fuscovaginae, which was previously reported as a causal agent of grain discoloration and grain rot in Latin America (Zeigler and Alvarez, 1987a and Ziegler and Alvarez, 1990) and in Africa (Rott et al., 1989 and Rott et al., 1991), was detected among bacteria isolated from diseased rice in Louisiana. It was evident that most of the Pseudomonas species, although associated with Burkholderia species in epiphytic communities on diseased rice tissues, were not the causal agents. Five of 15 Pseudomonas species showed pathogenicity on Cypress rice (Table 2.2.3). Two strains of P. syringae pv zizanize, isolated from diseased grains, showed pathogenic characteristics to both rice seedlings and panicles similar to that reported in the literature. P. syringae pv tagetis isolated from a diseased rice sheath was not pathogenic in inoculation tests. Three strains of P. pyrrocinia were rice pathogens consistent with previous reports in the literature (Table 2.2.3) (Rott et al., 1989). Seven strains of Acidovorax avenae, including two isolates of A. avenae subsp. avenae and five isolates of A. avenae subsp. cattleyae, were not pathogenic on rice seedlings or panicles although Zeigler (1990) reported A. avenae as pathogenic in Latin America. These two subspecies were non-pathogens. In addition, an isolate identified as Vibrio tubiashii by the BiologTM System, caused intermediate pathogenicity on inoculated rice seedlings in this study. Whether it is a new plant pathogenic species needs further research. For most of the bacterial strains tested, isolates were pathogenic to both seedlings and panicles (Table 2.2.2). However, there were 26 instances among Burkholderia strains where the strain was pathogenic on panicles, but not on seedlings. Only one strain was

23 pathogenic on seedlings but not on panicles. This suggests that panicles in the boot or emerging are more susceptible than 4-leaf-stage seedlings.

TM Figure 2.2.1. Biolog GN2 Profile of B. gladioli isolate 382gr-1.

Figure 2.2.2. BiologTM GN2 profile of B. multivorans isolate 99sh-2.

24

Figure 2.2.3. Biolog GN2 Profile of P. syringae pv. zizaniae isolate 319gr-4.

Figure 2.2.4. BiologTM GN2 Profile of B. plantarii isolate 260gr-3.

Figure 2.2.5. BiologTM GN2 Profile of B. glumae isolate156sh-1-b.

25 Table 2.2.1. Genera of bacteria isolated from panicle blighted rice plants collected in Louisiana. Genera isolated Number of species in each genus Burkholderia 6 Pseudomonas 15 Acidovorax 2 Agrobacterium 1 Aquaspirillum 1 Aeromonas 1 Commonas 1 Chryseobacterium 1 Flavimonas 1 Flavobacterium 1 Herbaspirillum 1 Pasteurella 1 Rhizobium 1 Sphingomonas 2 Stennotrophomonas 1 Vibrio 3

Table 2.2.2. The frequency of isolation of Burkholderia species from rice plants showing symptoms of bacterial panicle blight.* Frequency of distribution Number of isolates Species Identified of the species among of each genus Burkholderia isolates

B. glumae 68 26% B. gladioli 103 39.3% B. multivorans 60 22.9% B. plantarii 25 9.5% B. vietnamiensis 3 1% B.cocovenenans 3 1% * Isolations on SP-G semi-selective medium (Tsushima et al., 1986).

26 Table 2.2.3. Results of BiologTM identification tests and seedling and panicle inoculation tests to determine pathogenicity of bacterial isolates from bacterial panicle blight infected rice collected in Louisiana. Bacterial species Pathogenicity Tests* Isolate number according to BiologTM seedling panicle 99gr-8-a B. glumae ++ +++ 99gr-2-a B. glumae +++ +++ 98gr-5 B. multivorans +++ +++ 99gr-3-b B. glumae ++ +++ 99gr-4-a B. glumae ++ +++ 6-1 P. resinovorans NP NP 98gr-1 B. gladioli NP NP 98gr-8-a P. tolaasii NP NP 99sh-2 B. multivorans + + 99gr-11 B. glumae +++ +++ 99gr-8-b No ID NP NP 98gr-12-a B. glumae ++ ++ 99gr-6-a B. glumae +++ ++ 99sh-6 B. glumae +++ +++ 80-1 B. glumae +++ +++ 98gr-13-b B. gladioli +++ + 99sh-9 B. glumae + ++ 32-5 P. putida NP +/NP** 99sh-18 B. glumae ++ +++ 101gr-2 B. glumae + ++ 99sh-3 B. glumae +++ +++ 99gr-2-b B. glumae +++ +++ 99gr-4-b B. glumae +++ +++ 99gr-7-a B. plantarii +++ +++ 99sh-15-a B. multivorans +++ +++ 99sh-12 B. multivorans ++ ++ 101gr-1 B. glumae +++ ++ 98gr-6-a B. glumae ++ ++ P. putida biotype 15R-1 NP A NP 99sh-16 B. plantarii +++ +++ 99gr-5-b B. multivorans NP ++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

27 Table 2.2.3. (continued) Bacterial species according Pathogenicity Tests* Isolate number TM to Biolog seedlings panicles 98gr-6-b B. glumae +++ ++ 99sh-10 B. gladioli +++ +++ 99sh-5 B. glumae ++ +++ 34-1 P. putid biotype A NP NP 98gr-2 P. putida NP NP 98gr-3 B. multivorans NP ++ 99sh-11 B. glumae +++ +++ 99sh-17 B. multivorans ++ +++ 98gr-13-a B. gladioli NP + 99gr-1-b B. multivorans ++ ++ 99gr-6-b B. glumae ++ +++ 50-9 P. aurantiaca NP NP 99gr-1-a B. glumae +++ +++ 99sh-7 B. gladioli NP NP 99sh-15-b B. multivorans +++ +++ 99gr-9 B. gladioli ++ +++ 99gr-4-b B. gladioli +++ +++ 99gr-3-a A. dispar NP NP 99gr-7-b B. glumae ++ +++ 11sh-5 B. multivorans NP NP 35-2 P. putida NP NP 25-2 P. aeruginosa NP NP 99gr-5-a B. glumae ++ + 10-4 S. maltophilia NP NP 99sh-14 B. glumae +++ +++ 98gr-12-b B. glumae ++ ++ 99sh-4-a B. vietnamiensis ++ +++ 99gr-4-b P. tolaasii +++ +++ 170gr-4 A.Tumefaciens/Radiobacter NP NP 191gr-6 B. plantarii ++ ++ 146sh-3 P. boreopolis NP NP *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

28 Table 2.2.3. (continued) Bacterial species Pathogenicity Tests* according to Isolate number BiologTM seedlings panicles 167gr-1 P. spinosa + ++ 189sh-7 B. multivorans ++ ++ 192gr-2 B. gladioli NP NP 191gr-1 V. fluvialis NP NP 189sh-1 A. a. ss avenae NP NP 190gr-4 P. aeruginosa NP NP 133sh-3 P. resinovorans NP NP 191sh-10 B. glumae NP +++ 127sh-7 P. spinosa NP NP 189gr-8 B. glumae NP NP 171gr-2 B. glumae + ++ 190gr-1 B. glumae + +++ 191sh-1 B. multivorans NP NP 156sh-1 P. resinovorans NP NP 157sh-1 B. multivorans + + 158sh-2 P. resinovorans NP NP 154sh-1-a B. plantarii ++ +++ 127sh-6 P. spinosa NP NP 170sh-1 B. gladioli + +/NP* 190gr-2 B. gladioli ++ ++ 156sh-1-a P. resinovorans NP NP 171gr-1-a B. gladioli ++ ++ 191gr-4 B. gladioli + + 170gr-3-a B. gladioli ++ ++ 190gr-3 B. gladioli ++ ++ 166gr-6 P. aurantiaca NP NP 171gr-3 B. gladioli ++ +++ 189gr-9 B. gladioli ++ ++ 191sh-2 P. corrugata NP NP 191sh-12 B. glumae ++ +++ 166gr-1 B. gladioli NP +++ 189gr-4 B. glumae +++ +++ 189gr-2 B. gladioli ++ ++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

29 Table 2.2.3. (continued) Bacterial species Isolate number according to Pathogenicity Tests* BiologTM seedlings panicles 189sh-6 A. a. ss avenae NP NP 147sh-2 A. a. ss cattleyae NP NP/+** 190gr-1 B. glumae + +++ 167sh-3 H. seropedicae NP NP 188gr-1 B. plantarii NP + 188gr-2 B. glumae ++ +++ 192gr-1 B. gladioli + +++ 166gr-2 B. gladioli ++ ++ 184sh+gr-2 S. maltophilia NP NP 191gr-2 B. plantarii ++ +++ 180gr-2 F. oryzihabitans -*** NP 154sh-1-b S. macrogoltabidus NP NP 171gr-1-b B. gladioli ++ ++ 156sh-1-b B. glumae + + 170gr-3-b B. gladioli +++ ++ #5: 193gr-1 B. multivorans + +++ 213sh-2 P. putida biotype B NP NP 198gr-2 P. tolaasii + +++ 201gr-1 B. multivorans NP +++ 192gr-7 B. plantarii ++ ++ 206gr-3 B. plantarii + ++ 207gr-2 B. multivorans + ++ 217gr-1 B. multivorans ++ +++ 210gr-1 P. agarici NP NP 202gr-2 B. plantarii + ++ 203gr-2 B, multivorans NP +++ 209gr-2 B. gladioli + +++ 201sh-1 B. multivorans NP NP 195gr-7 B. plantarii + +++ 196gr-4 B. multivorans + ++ 216gr-3-a B. multivorans + +++ 197gr-1-a B. multivorans ++ ++ 216gr-8 B. multivorans ++ +++ 200gr-4 P. tolaasii +++ +++ P.fluorescens 200gr-5 biotype G NP +++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants. *** - = no data available

30 Table 2.2.3. (continued) Bacterial species Pathogenicity Tests* Isolate number according to BiologTM seedlings panicles 208gr-3 B. multivorans ++ +++ 196gr-5 B. multivorans ++ +++ 210gr-3 B. gladioli +++ +++ 198gr-3 B. multivorans + +++ 216gr-3-b B. multivorans +++ ++ 209gr-1 B. multivorans +++ +++ 201gr-2 B. gladioli + +++ 216sh-1 B. gladioli + +++ 217gr-2 B. multivorans ++ +++ 207gr-4 B. gladioli ++ NP 206sh-1 P. fluorescens +++ ++ 192gr-5 B. plantarii + +++ 197gr-2 B. multivorans ++ +++ 195gr-11 B. glumae ++ ++ 197gr-1-b B. multivorans + +++ 213gr-1 P. fluorescens + +++ 195gr-5 B. multivorans ++ +++ 216gr-1 B. glumae ++ ++ 199gr-2 B. multivorans NP ++ 193gr-5 B. multivorans ++ +++ 196gr-1 B. multivorans NP +++ 200gr-3 B. multivorans NP +++ 216gr-6 B. multivorans + +++ 217gr-3 B. plantarii ++ ++ 203gr-1 B. multivorans NP ++ 195gr-6 B. plantarii NP +++ 201gr-2 B. multivorans + +++ P. fluorescens 210gr-4 + biotype G +++ 200sh-1 B. multivorans ++ +++ #6: 221gr-3 +++ +++ 237gr-3 B. gladioli +++ +++ 217sh-1 B. gladioli + ++ 252gr-5-b B. gladioli +++ +++ 250sh-1 B. gladioli +++ +++ 221sh-1 B. multivorans NP ++ 219sh-4 B. gladioli +++ +++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

31 Table 2.2.3. (continued) Bacterial species according Pathogenicity Tests* Isolate number TM to Biolog on seedlings on panicles 241gr-1 B. gladioli +++ +++ 220gr-1-b B. glumae ++ +++ 250gr-1 B. gladioli NP ++ 249sh-1 B. gladioli + + 218sh-2 B. glumae ++ ++ 222gr-1-a B. gladioli ++ +++ 227gr-2-a B. gladioli +++ +++ 220gr-2-b B. gladioli ++ ++ 250gr-2 B. plantarii ++ ++ 237gr-1 B. gladioli +++ +++ 227gr-1 B. gladioli +++ +++ 228gr-1 B. gladioli +++ +++ 221gr-1 B. multivorans +++ +++ 252gr-5-a B. multivorans +++ +++ 222gr-2 P. syringae pv zizanize ++ +++ 230sh-1 P. pyrrocinia ++ ++ 252gr-2 B. gladioli + ++ 237gr-5 B. gadioli NP NP 235gr-2 B. gladioli ++ ++ 218gr-1 B. glumae +++ ++ 252gr-1 B. gladioli +++ +++ 227gr-2-b B. glumae +++ +++ 252gr-7 B. multivorans ++ + 222gr-1-b B. glumae + + 249gr-2 B. gladioli +++ ++ 230gr-1 B. gladioli +++ ++ 241gr-3 B. gladioli +++ +++ 218sh-1 B. gladioli ++ +++ 236gr-1 B. gladioli ++ ++ 243gr-1 B. gladioli ++ ++ 249gr-3 P. tolaasii ++ +++ 223gr-1 B. gladioli ++ ++ 219sh-2 P. syringae pv tagetis NP ++ 238gr-3 B. gladioli +++ ++ 236gr-3 B. gladioli +++ ++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

32 Table 2.2.3. (continued) Bacterial species Pathogenicity Tests* Isolate number TM according to Biolog seedlings panicles 235gr-1 B. gladioli ++ + +++ 252gr-4 B. plantarii ++ +++ 252gr-2 B. gladioli +++ +++ 219sh-1 B. gladioli ++ +++ 230sh-2 B. gladioli +++ +++ 235sh-1 B. multivorans +++ +++ 220gr-1-a P. pyrrocinia +++ +++ 223gr-2 B. gladioli ++ +++ 220gr-2-a B. gladioli +++ +++ 218gr-3 B. plantarii + ++ 243gr-3 B. gladioli +++ +++ 226gr-1 B. glumae +++ +++ 237gr-5 B. gladioli NP NP #7: 258gr-5 B. multivorans +++ ++ 261gr-9 B. multivorans NP +/NP 253gr-2 A. facilis NP NP 261gr-1 B. multivorans +++ +++ 258gr-8 B. glumae +++ ++ 257gr-2 B. gladioli ++ +++ 257sh-1 B. gladioli NP NP 270sh-4 B. plantarii +++ +++ 273sh-1 B. multivorans +++ +++ 259gr-1 B. gladioli ++ ++ 253sh-1 B. multivorans +++ +++ 266gr-1 S. macrogoltabidus NP ++

256sh-2 R. radiobacter NP NP 256gr-2 B. multivorans NP + 261gr-2 B. multivorans ++ +++ 266gr-5 B. multivorans +++ +++ 259gr-4 B. gladioli NP ++ 261gr-5 V. tubiashii ++ ++ 273sh-2 B. multivorans + +++ 258gr-4 B. gladioli + ++ 254sh-4 A. avenae. ss cattleyae NP NP 266sh-2 P. tolaasii +++ +++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

33 Table 2.2.3. (continued) Bacterial species Pathogenicity Tests* Isolate number TM according to Biolog seedlings panicles 259gr-5 B. gladioli ++ ++ 261gr-4 B. multivorans +++ +++ 272gr-4 B. multivorans NP + 254sh-1 P. tolaasii +++ +++ 260gr-1 B. glumae + +++ 258gr-1 B. multivorans NP ++ 258gr-5 B. multivorans +++ ++ 256sh-3 P. tolaasii NP +++ 256gr-1 B. phenazinium + ++ 253sh-2 A. a ss cattleyae NP NP 255sh-1 P. tolaasii + ++ 269gr-1-a B. gladioli +++ ++ 257sh-2 B. glumae ++ +++ 257gr-1 C. meningosepticum NP ++ 254gr-1 B. gladioli + +++ 267gr-1-b S. sanguinis NP NP/+ 273gr-1 B. glumae + +++ 260gr-3 B. plantarii NP + 257gr-2 B. gladioli NP +++ 254gr-5 B. plantarii ++ +++ 254gr-3 F. ferrugineum NP NP/+ 259gr-3 B. plantarii + +++ 265gr-1 B. glumae + ++ 270sh-2 B. plantarii NP +++ 255gr-2 B. multivorans ++ +++ 273gr-4 B. glumae ++ + 267gr-2 B. plantarii ++ + 270gr-1 P. tolaasii ++ +++ 254gr-2 B. gladioli +++ +++ 269gr-1-b B. glumae ++ +++ 272gr-2 B. multivorans +++ +++ 106sh-11 B. glumae +++ +++ 396gr-2 B. gladioli NP NP 363gr-1 B. gladioli NP + 407gr-6 B. gladioli +++ +++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

34 Table 2.2.3. (continued) Bacterial species Pathogenicity Tests* according to Bacterial isolates BiologTM seedlings panicles 387gr-2-b B. gladioli +++ +++ 406gr-3 B. gladioli +++ +++ 408gr-4 B. cocovenenans +++ +++ 357-8 B. gladioli ++ ++ 366gr-1 B. multivorans +++ +++ 379gr-1-a-(1) B. glumae +++ +++ 106sh-7 B. vietnamiensis + ++ 378g-3 B. gladioli ++ + 106sh-9 B. plantarii NP NP 357gr-1 B. gladioli NP NP 106sh-5 B. glumae NP NP 384gr-1 P. viridilivida NP NP 395-2 B. gladioli NP NP 408gr-7 B. gladioli + ++ 403gr-2 B. glumae ++ +++ 980007-1(5) B. glumae +++ +++ 405gr-4 B. cocovenenans +++ +++ 961149-4(4) B. gladioli NP NP 407gr-3 B. plantarii +++ +++ 357-3 B. gladioli ++ + 106gr-3 B. glumae ++ + 106sh-4-b B. glumae +++ +++ 387gr-2-a B. glumae ++ +++ 367gr-1 B. glumae ++ +++ 395gr-2 B. gladioli +++ +++ 367gr-3 B. glumae ++ +++ 254gr-2 B. multivorans +++ +++ 255gr-2 B. glumae ++ ++ 373gr-1 B. gladioli +++ +++ 318gr-4 B. glumae ++ +++ 366gr-2 B. gladioli NP NP 318gr-1 B. multivorans +++ +++ 379gr-1-b B. gladioli NP NP 980007-1(6) B. gladioli +++ +++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

35 Table 2.2.3. (continued) Bacterial species Pathogenicity Tests* Isolate number TM according to Biolog seedlings panicles 382gr-1 B. gladioli +++ ++ 385gr-1 B. gladioli ++ + 106sh-4-a B. glumae NP NP 366gr-6 B. vietnamiensis ++ ++ 398gr-3 B. glumae +++ +++ 373gr-2 P. viridilivida NP NP 398gr-1 B. glumae +++ +++ 407gr-8 B. gladioli + ++ 407gr-1 B. gladioli +++ +++ 379gr--1-a-(2) B. gladioli ++ ++ 951886-4(3) B. glumae ++ +++ 372gr-1 B. gladioli ++ +++ 379gr-1-a-(1) B. multivorans ++ ++ 393gr-7 P. tolaasii - NP/+** 321gr-9 B. gladioli + + 321gr-2-a B. gladioli +++ + 321gr-3 B. gladioli ++ ++ 951886-4(3) B. gladioli +++ ++ 321gr-6 B. gladioli ++ + 325gr-1 B. gladioli + + 321gr-8-a B. gladioli ++ + 324gr-2 B. gladioli ++ +++ 324gr-3 B. gladioli ++ +++ 319gr-4 P. syringae pv zizanize ++ + 321gr-2-b B. gladioli ++ ++ 321gr-4 B. plantarii ++ ++ 325gr-2 B. gladioli ++ ++ 321gr-8-b B. gladioli ++ ++ 325gr-4 B. gladioli ++ +++ 321gr-7 B. gladioli ++ ++ 319gr-1 B. gladioli +++ +++ 319gr-6 - *** + + 154sh-1-a B. gladioli + ++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants. *** - = no data available

36 Table 2.2.3. (continued) Bacterial species Pathogenicity Test* Isolate number TM according to Biolog seedling panicle 113sh2-7 B. glumae P +++ 113sh2-4 B. plantarii P ++ 113sh2-5 B. glumae P + 113sh2-2-a B. glumae P + 113sh2-8 B. glumae P +++ 117sh1-3 B. glumae P +++ 117sh1-8 B. glumae P +++ 117sh2-3 B. glumae P +++ 117g1-7-a B. glumae NP NP 116g1-1 B. glumae P +++ 117g1-8 B. glumae P +++ 117g1-9 B. glumae P +++ 117sh1-1 S. maltophilia NP NP 120g1-1 B. glumae P ++ 119g1-5 B. glumae P +++ 123sh-3 - - - 118sh-6 A. avenae ss cattleyae NP NP 119g1-4 B. glumae NP ++ 119g1-3 P. pyrrocinia P +++ 120g1-2 B. glumae P +++ 122sh-1 - NP NP 117g1-11 B. glumae P +++ 118sh-5 P. group2 (B.-like) NP NP 114sh1-4 B. cepacia P NP 122sh-5 A. jandaei DNA group 9 NP NP 113sh2-9 B. glumae P +++ 113sh2-3 B. plantarii P NP 113sh1-10 C. acidovorans NP NP 124sh-1 - - - 124sh-4 - - - 113sh2-2-b - P + 117g1-7-b - P ++ *NP = not pathogenic, ratings represent degree of virulence when strain was pathogenic with + = weakly virulent, ++ = moderately virulent, and +++ = highly virulent. ** NP/+ = NP or + on different inoculated plants.

37 Table 2.2.4. The frequency of pathogenic bacterial strains of each species isolated from rice based on inoculation of Cypress rice plants. Number of Pathogenic Total number of Relative pathogenic bacterial isolates of each frequency isolates of each species species (%) identified species Burkholderia spp. B. glumae 62 68 91 B. gladioli 85 103 83 B. multivorans 44 60 73 B. plantarii 20 25 80 B. cocovenenans 3 3 100 B. vietnamiensis 3 3 100 Pseudomonas spp. P. syringae pv zizanize 2 2 100 * P. fluorescens (biotype G) 3 4 75 P. tolaasii 7 12 58 P. pyrrocinia 3 3 100 P. spinosa 1 3 33 Vibrio spp. V. tubiashii 1 1 100 * P. fluorescens or P. fluorescens biotype G

2.2.2 Pathogenicity Tests and Symptoms Produced Pathogenicity tests were applied using the method of injection of bacterial suspensions into the stems of 4-leaf-stage rice seedlings 1 – 2 days before tillering or into boots just before panicles emerged using the variety Cypress. Continuous observations and records for the progressive symptoms were made with a digital camera and notes were taken for 20 – 25 days after inoculation. A range of leaf and leaf sheath symptoms was observed on seedlings (Figures 2.2.6, 2.2.7, 2.2.8, and 2.2.9) and the boot leaf sheath ( Figure 2.2.11, 2.2.18, and 2.2.20), emerging panicles including poor panicle emergence ( Figure 2.2.10) grain discoloration, sterile spikelets, and complete blacking of hulls (Figure 2.2.19). Panicle damage and grain discoloration existed in various degrees of severity. In the severest case, a poor or total lack of panicle emergence occurred. The lesions on sheaths developed longitudinally into with brown/black necrosis on sheaths, with grayish-white centers and dark brown, distinct margins. Symptoms produced by identified B. glumae cultures were similar among all pathogenic Burkholderia species. In infections with intermediate virulence, a restricted foci around the injection point was

38 formed on the sheath (Figure 2.2.8). For inoculated rice seedlings, a widespread symptom was yellow or light-brown rot of the leaf sheath. The development of the disease depended strongly on weather conditions. Cooler temperatures of 23oC – 25oC significantly reduced lesion development on affected leaf sheaths. No infection was expressed as a yellow or small light brown area around the injection site (Figure 2.1.3). An experiment where seedlings and emerging panicles were simultaneously inoculated with the same bacterial strains under, under cool temperature conditions, confirmed that unlike rice seedlings, emerged panicles were still severely damaged. In a few cases, B. plantarii and B. cocovenenans expressed distinct symptoms. A brown leaf stripe along mid-veins was produced by B. plantarii (Figure 2.2.7) and with B. cocovenenans leaf discoloration spread along the leaf blade edge to the tip of the leaf (Figure 2.2.9). However, whether symptoms can be used to distinguish pathogenic strains needs further study. Typical infection by virulent or moderately virulent strains of several species of Burkholderia are shown in Figures 2.2.12 to 2.2.17.

Figure 2.2.6. Leaf sheath rot on Cypress seedlings caused by B. glumae isolate106sh-4-b on the left and B. gladioli isolate 980007-1(6) on the right. Please note that the whole inoculated leaf sheath rots when the bacterial isolate is highly virulent.

39

Figure 2.2.7. Brown and tan leaf stripes or lesions on seedling leaf blades infected by B. glumae isolate 106sh-11on the left and B. plantarii isolate 407gr-3 on the right.

Figure 2.2.8. Low virulence reaction produced by B. gladioli (isolate 357gr-1 on the left) and B. gladioli (isolate 406gr-3 on the right). Symptoms were expressed as a brown lesion of less than 2cm length around the inoculation site with a distinct darker brown margin.

40

Figure 2.2.9. Leaf discoloration extending along the edge of the leaf blade down to the leaf tip caused by B. cocovenenans (isolate 405gr-4 on the left) and the uninoculated Cypress control on the right.

Figure 2.2.10. Poor emergence of panicle with spilelet sterility by caused by B. multivorans isolate 366gr-1on the left and a dark brown sheath and grain Rot caused by B. glumae isolate 367gr-3 on the right.

41

Figure 2.2.11. Extended brown necrotic lesion caused by a highly virulent strain of B. glumae (367gr-3 on the left) and a dark brown lesion surrounding the inoculation site on the sheath caused by the moderately virulent B. glumae strain (398gr-3 on the right).

Figure 2.2.12. Strain 98gr-5 of Burkholderia multivorans showing a virulent (+++) reaction on Cypress rice after seedling inoculation.

42

Figure 2.2.13. Strain 98gr-5 of Burkholderia tolaasii, on Cypress rice showing a highly virulent (+++) reaction after seedling inoculation.

Figure 2.2.14. Seedling disease reaction (+++) caused by bacterial strain 99gr-2-b of Burkholderia glumae on Cypress rice.

43

Figure 2.2.15. Highly virulent seedling disease reaction (+++) of bacterial strain 99sh-10, on Cypress rice.

Figure 2.2.16. Highly virulent seedling disease reaction (+++) of bacterial Strain 99sh-16, Burkholderia plantarii on Cypress rice.

44

Figure 2.2.17. Seedling disease ++ reaction of bacterial strain 99sh-4-a, Burkholderia vietnamiensis on Cypress rice.

Figure 2.2.18. Highly virulent panicle and flag leaf sheath reaction (+++) on isolate 210gr-3 of Burkholderia gladioli on Cypress rice after boot inoculation (overview).

45

Figure 2.2.19. Close-up view of severe panicle blighting symptoms on Cypress rice resulting from inoculation with isolate 210gr-3 of Burkholderia gladioli.

46

Figure 2.2.20. Close-up view of flag leaf sheath lesions produced when Cypress rice boots were inoculated with isolate 210gr-3 of Burkholderia gladioli.

2.2.3 Colony and Culture Characteristics of the Bacterial Pathogens Examinations of colony morphology showed that the Burkholderia species could not be distinguished simply by colony or culture characters. All of the species formed convex colonies with smooth margins or flat colonies with pleated margins on KBA medium, releasing yellow pigment or no pigment into the medium. Most cultures had gray-white or cream-white colonies. In a very few cultures there was a reddish tint to the colonies. Yellow or gray-white colonies were the most common. B. gladioli colonies were yellow, gray-white, or yellow-green in color, but all of these colony-types were pathogenic to rice plants, suggesting that there may be toxins other than the phytotoxin associated with the yellow pigment. All avirulent colonies of B. glumae, B. gladioli, B. multivorans and B. plantarii were gray-white or cream white colonies that did not release

47 pigments into the medium (Figures 2.2.21, 2.2.22, and 2.2.23). When selected cultures of B. glumae and B. gladioli were grown on King’s B medium and observed for production of pigment (toxin) in the medium, it was clear that cultures that produced the yellow or green pigments were pathogenic and virulent (Table 2.2.5). Cultures that did not produce pigment were avirulent. Several non-pathogenic and non-pigment producing cultures of both B. glumae and B. gladioli were identified in this experiment. These cultures will be useful for future experiments to determine what controls toxin production and if this character can be transferred from virulent cultures to avirulent cultures. It’s possible that the gene(s) for toxin production are carried by a plasmid that can be transferred among Burkholderia spp.

Figure 2.2.21. Cultures of B. plantarii, virulent strain (left) and avirulent strain (right). Note the yellow toxin in the virulent culture and the lack of yellow toxin in the medium of the avirulent strain.

48 Table 2.2.5. Production of yellow or green pigment (toxin) in King’s B medium by pathogenic and nonpathogenic Burkholderia glumae and Burkholderia gladioli isolates. Seedling Panicle Pigment Isolate Number Bacterial species pathogenicity pathogenicity (toxin)* rating rating 357gr-1 B. gladioli NP NP N 106sh-5 B. glumae NP NP N 395-2 B. glumae NP NP N 366gr-2 B. gladioli NP NP N 237gr-5 B. gladioli NP NP N 99sh-7 B. gladioli NP NP N 257sh-1 B. gladioli NP NP N 321gr-9 B. gladioli + + W 379gr-1-b B. gladioli NP NP N 396gr-2 B. gladioli NP NP N 98gr-13-a B. gladioli NP + N 259gr-4 B. gladioli NP ++ W 363gr-1 B. gladioli NP + N 189gr-8 B. glumae NP NP W 191sh-10 B. glumae NP +++ W 250gr-1 B. gladioli NP ++ W 321gr-9 B. gladioli + + W 192gr-2 B. gladioli NP NP W 166gr-1 B. gladioli NP +++ IM 98gr-1 B. gladioli NP NP N 106sh-4-a B. glumae NP NP N 319gr-1 B. gladioli +++ +++ G 398gr-1 B. glumae +++ +++ I 106sh-11 B. glumae +++ +++ I 980007-1(5) B. glumae +++ +++ I 321gr-2-a B. gladioli +++ + IM 406gr-3 B. gladioli +++ +++ I 382gr-1 B. gladioli +++ ++ IM 407gr-1 B. gladioli +++ +++ I 99gr-6-a B. glumae +++ ++ I 99gr-2-a B. glumae +++ +++ I * Production of pigment (toxin) is indicated as “I” = intense, “IM” = intermediate, “W” = weak (very faint color produced), and “N” = none. “G” = no yellow pigment produced, but green pigment was produced.

49

Figure 2.2.22. Cultures of Burkholderia glumae with a virulent strain on the left and an avirulent strain on the right. Note the lack of toxin in the medium on the right.

Figure 2.2.23. Cultures of Burkholderia gladioli, virulent strain (left) and avirulent strain (right). Note the lack of yellow pigment (toxin) in the medium on the right and the presence of yellow pigment on the left.

2.2.4. Clustering of Bacterial Strains Isolated from Rice Showing Panicle Blight into Functional Groups In order to compare the similarity of strains of the different species to representative strains stored in the BiologTM GN database, bacterial isolates were clustered into functional groups based upon data from BiologTM identification tests and pathogenicity tests. Each isolate was therefore assigned two characteristic parameters: a virulence parameter and a typicality or similarity parameter. Evaluation using the

50 virulence system produced four groups: non-virulent strains, low virulence strains, middle-virulent strains, and high virulence strains. Evaluation using the similarity system formed two groups: low similarity strains and high similarity strains (Table 2.2.6). The threshold value of high / low similarity of B. plantarii was assigned at 0.6. Considering the larger number of strains in B. glumae, B. multivorans and B. gladioli, a weight was added to the threshold values; shifting the values of B. multivorans (Table 2.2.7) and B. glumae to 0.65 and B. gladioli strains were shifted to 0.7. The results indicated that B. glumae (Table 2.2.9) and B. gladioli (Table 2.2.8) possessed more virulent strains and higher similarity indexes as strain populations than that of B. plantarii, suggesting these two species may be the main pathogens in this study. A higher similarity index implied higher detection accuracy and high similarity to reference strains. However, low similarity might not mean low detection accuracy, but imply that they were different types, considering that bacterial metabolic types are variable (Jones, 1993). B. multivorans expressed an intermediate pattern of population distribution between B. plantarii, B. gladioli, and B. glumae according to the above evaluation system, suggesting that they took second place in the importance of pathogens (Table 2.2.7).

Table 2.2.6. Clustering of Burkholderia plantarii into functional and typical strains. High Virulence* Low Virulence Non-Virulent Mid Virulence High Low High Low High Low High Low similarity** similarity similarity similarity similarity similarity similarity similarity

407gr-3 195gr-7 206gr-3 260gr-3 195gr-6 217gr-3 192gr-7 270sh-4 154sh-1- 192gr-5 202gr-2 188gr-1 270sh-2 252gr-4 a 99sh-16 218gr-3 321gr-4 217gr-3 99gr-7-a 259gr-3 191gr-2 254gr-5 267gr-2 191gr-6 250gr-2 * The three-scale system of disease severity was determined by pathogenicity tests. ** A threshold value of similarity to typical strains was assigned with a similarity index of 0.6 in this table. The similarity was considered as high similarity if the index was more than 0.6; otherwise as low similarity.

51

Table 2.2.7. Clustering of Burkholderia multivorans into functional and typical strains.* High Virulence Low Virulence Non-Virulent Mid Virulence High Low High Low High Low High Low similarity similarity similarity similarity similarity similarity similarity similarity 98gr-5 366gr-1 198gr-3 99sh-2 258gr-1 11sh-5 252gr-7 99gr-1-b 261gr-4 318gr-1 201gr-2 273sh-2 261g-9 99gr-5-b 216gr-8 99sh-17 99sh-15- 197gr-1- 197gr-1- 261gr-1 157sh-1 221sh-1 272gr-4 99sh-12 a b a 216gr-3- 273sh-1 266gr-5 196gr-4 201sh-1 256gr-2 196gr-5 261gr-2 a 252gr-5- 253sh-1 193gr-1 203gr-1 191sh-1 208gr-3 255gr-2 a 216gr-3- 272gr-2 207gr-2 196gr-1 201gr-1 379gr-1 200sh-1 b 209gr-1 235sh-1 216gr-6 200gr-3 203gr-2 a(1) 217gr-1 254gr-2 199gr-2 217gr-2 258gr-5 197gr-2 221gr-1 195gr-5 193gr-5 * A threshold value of similarity to typical strains was assigned with a similarity index of 0.65 in this table

52 Table 2.2.8. Clustering of Burkholderia gladioli into functional and typical strains.*

High Virulence Low Virulence Non-Virulence Mid Virulence

High Low High Low High Low High Low similarity similarity similarity similarity similarity similarity similarity similarity

321gr-2- 379gr-1- 325gr-1 408gr-7 321gr-9 321gr-6 99gr-9 a b 99sh-10 220gr-2- 319gr-1 191gr-4 192gr-1 166gr-1 99sh-7 325gr-4 b 170gr-3- 98gr-13- 170sh-1 209gr-2 237gr-5 98gr-1 321gr-3 235gr-1 b b 254gr-2 98gr-13- 243gr-3 217sh-1 216sh-1 357gr-1 325gr-2 243gr-1 a 241gr-3 241gr-1 249sh-1 363gr-1 257sh-1 324gr-2 219sh-1 249gr-2 219sh-4 258gr-4 396gr-2 257gr-2 324gr-3 166gr-2 252gr-2 321gr-8- 237gr-3 254gr-1 366gr-2 192gr-2 207gr-4 a 252gr-5- 321gr-8- 170gr-3- 237gr-1 407gr-8 250gr-1 b b a 227gr-2- 236gr-3 154sh-1- 321gr-2- 259gr-4 235gr-2 a a b 252gr-1 228gr-1 321gr-7 99gr-4-a 210gr-3 171gr-1- 230sh-2 b 250sh-1 227gr-1 171gr-3 220gr-2- 230gr-1 189gr-2 a

269gr-1- 190gr-2 a 238gr-3 190gr-3 382gr-1 189gr-9 980007- 171gr-1-

1(6) a 407gr-6 223gr-1 406gr-3 223gr-2 222gr-1- 407gr-1 a 395gr-2 387gr-2- 236gr-1

b 218sh-1 259gr-5 259gr-1 372gr-1 378gr-3 385gr-1 379gr-1-

a(2) 357-8 357-3 *A threshold value of similarity to typical strains was assigned with a similarity index of 0.70 in this table.

53 Table 2.2.9. Clustering of Burkholderia glumae into functional and typical strains.* High Virulence Low Virulence Non-Virulent Mid Virulence

High Low High Low High Low High Low similarity similarity similarity similarity similarity similarity similarity similarity 252gr-5- 156sh-1- 189gr-4 171gr-2 189gr-8 106sh-5 191gr-2 a b 222gr-1- 216gr-1 218gr-1 258gr-8 265gr-1 191sh-10 188gr-2 b 195gr-11 106sh-4- 189sh-7 226gr-1 99sh-6 260gr-1 99sh-9 218sh-2 a 273gr-4 227gr-2- 220gr-1- 99gr-11 273gr-1 101gr-2 395-2 b b 98gr-12- 80-1 98gr-6-b 190gr-1 257sh-2 b 106sh-4- 269gr-1- 101gr-1 99gr-7-b b b 99gr-6-a 398gr-3 99gr-8-a 99sh-18 99gr-2-a 99gr-3-b 99gr-5-a 99gr-4-b 99sh-14 99sh-5 367gr-1 951886 99gr-2-b 98gr-6-a -4(3) 99sh-3 99gr-6-b 98gr-12- 99sh-11 a 99gr-1-a 318gr-4 99sh-15- 367gr-3 b 398gr-1 403gr-2 980007- 255gr-2 1(5) 387gr-2- 106sh-11 a 387gr-2- 106gr-3 b * A threshold value of similarity to typical strains was assigned with a similarity index of 0.65 in this table.

Based on the percentage of isolates of each Burkholderia species among the bacterial isolates from rice showing BPB symptoms, B. gladioli, B. glumae, and B. multivorans appear to be the most important species causing BPB in Louisiana rice (Table 2.2.24).

54 120 The Number of Isolates 100 80 Relative Frequency (%) 60 40 20 0

e i ii a iol s r s s d ta si m a ran n n lu gl la e nan . ivo mi e . g B lt . p a n B B n ve t o mu e c B. vi o . .c B B

Figure 2.2.24. The Frequency of distribution of Burkholderia species isolated from rice plants (Oryza sativa L. cv. Cypress) showing BPB symptoms.

18 16 ates

l 14 o

s 12 I f

o 10

er 8 b

m 6 u 4 e N

h 2 T 0 i i i i i im im im im im imi -s -s s -s -s s h h- g w- g g ow o Low i L Hi L Hi H + + + u+ u ru ir ru+ ir i iru+ iru i v -v v -v -v v - - h h w on ig w- on g Lo N N Hi H Lo

Figure 2.2.25. Functional classification of bacterial isolates from rice plants showing BPB symptoms and identified as B. glumae.

Based on the similarity/pathogenicity groupings, it looks like B. glumae (Figure 2.2.25) and B. gladioli (Table 2.2.26) may have races based on the distribution of pathogenicity and closeness of relatedness.

55 25

ates 20 l o s I

f 15 o er

b 10 m u

N 5 e h T 0 i i i i i i im im im m im im s -s s -si -s -s h- ig w igh- ig ow H o H Low H L + L + u + u u+ ru+ ir ru ir ir iru+ i v i -v v -v - -v -v h - n n h g w w o g Lo N No Hi Hi Lo

Figure 2.2.26. Functional classification of the bacterial isolates from rice plants showing BPB symptoms and identified as B. gladioli.

2.3 Summary and Conclusions Several genera and species of bacteria were isolated from Louisiana-grown rice showing symptoms of panicle blighting. Thirty nine bacterial species in 10 genera could grow on the semi-specific medium, SP-G, suggesting that this medium was not unique for selecting B. glumae and thus the species growing on the medium partly represented the general distribution of bacteria in diseased rice tissues. Among the Burkholderia isolates B. gladioli and B. glumae had the largest numbers of isolates highly virulent to Cypress rice. This suggested that they were the species in the Burkholderia complex most important in causing BPB when compared to B. multivorans, B. plantarii, and other isolated Burkholderia and pathogenic Pseudomonas species. The components of populations and geographic distributions of bacterial pathogens are determined by many factors, including weather conditions, cultivar and lines, root location, and soil types. The cultivar Cypress was previously reported to be susceptible to B. glumae (Shahjahan et al., 2000b). Whether the cultivar was equally susceptible to pathogenic Pseudomonas spp. needs further study. Soil type is an important factor affecting not only the components of populations of plant-associated bacteria, but also the variability of genetics in bacteria. It has been shown that the B. cepacia complex displays a wide variability in genetics. Dalmastri et al. (1999)studied the associations between the genetic diversity of maize root-associated B. cepacia

56 populations with soil type and the cultivars grown. They confirmed that soil type played a more important role in determining the genetic diversity of B. cepacia populations than did the cultivars grown. B. multivorans, B. gladioli, and B. glumae were found to be the most common isolates from panicle blighted rice plants in Louisiana in this study. The effect of soil types, cultivars, weather, and other elements affecting population dynamics of the complex needs further study. The fact that many non-pathogenic Psudomonas species were isolated from diseased tissue needs further exploration, including an evaluation of the roles of Pseudomonas species in the process of development of the syndrome and ecological relationships with the pathogenic bacteria. B. glumae, B. gladioli, B. multivorans, and B. plantarii strains, isolated from diseased rice in the southern United States, caused flag leaf sheath lesions, spikelet sterility, and poor emergence of panicles after inoculation. In most cases, they induced discoloration and rot of leaf sheaths on rice seedlings, consistent with the literature. Similar symptoms were reported for P. fuscovaginae causing brown-black necrosis on stems and poor emergence of panicles (Rott et al., 1989), suggesting that some functional characteristics, including pathogenicity, are similar to each other even though they belong to different genera and tend to have different geographic distributions. P. syringae pv. aptata was confirmed to attack developing florets of rice plants, producing similar early symptoms to B. glumae and P. fuscovaginae (Goto et al., 1987). Two pathovars of P. syringae were found in this investigation. P. syringae pv. zizanize was pathogenic and produced symptoms similar to those produced by Burkholderia spp. on both seedlings and panicles of rice. Variability of bacteria for pathogenicity is a universal phenomenon, with several possible causes including heterogeneity within strains, biochemical variability, variance of pathogens, and loss of disease agents during inoculation. This study showed that continued subculturing of isolates could cause loss of virulence of B. glumae. In this survey, there were 26 Burkholderia isolates expressing non-pathogenicity on seedlings, but pathogenicity on panicles inoculated at the late boot stage. The reverse only occurred once. It may be that weakly virulent strains can infect the more susceptible immature panicle tissues, but not the rapidly developing seedling tissues. As the BPB disease is

57 seedborne, avoiding contaminated seeds was essential for inoculation studies. No infection of non-inoculated rice at the seedling or panicle stages was observed in this study. Therefore, in this study the results of pathogenicity tests were mainly based on the experiments on seedlings in order to ensure the reliability of results. The BPB syndrome is poorly defined. The development of symptoms and severity of disease not only depends on virulence of the strain, but also on environmental factors, particularly weather conditions. Klement (1955) pointed out that P. fuscovaginae and P. syringae grow well in relative cool, wet weather. The frequency in Japan of the occurrence of the seedling rot disease caused by B. glumae, where the weather is cooler and drier than Louisiana, was due to the introduction of new planting methods where seedlings were grown in seedling boxes in warehouses maintained at high temperatures and humidity (Goto et al., 1987). Our research showed that a temperature range of 37oC - 41oC was a critical environmental factor affecting the occurrence and development of the disease in the greenhouse, not only for B. glumae, but also for B. gladioli, B. multivorans, and B. plantarii. Therefore, in inoculation tests the descriptions of typical symptoms and the determinations of disease severity only had significance when these environmental conditions were met. In this study, all of the identified isolates obtained an acceptable similarity index, over 0.5 as assigned by the BiologTM System, and the identities could therefore be reported according to “GN2 MicroPlateTM Instructions for Use” (Biolog Co., Hayward, CA ). The highest indexes were received by some B. gladioli strains, which reached up to 1.0, indicating a full match with a reference strain in the database library. The 4.20 version of BiologTM database used in this study has been greatly improved as more strains were added to the library, offering a powerful technique to support identification of bacteria. Some Pseudomonas species were identified to the pathovar level. However, the identification of bacteria is a long process and cross-referencing among two or more detection methods is desirable to ensure that an isolate can be identified accurately. Fatty acid analysis for the isolated reported on in this research is being made separately by our laboratory.

58 Furthermore, less than 0.5 similarity indexes may not necessarily mean low detection accuracy. It should be considered that some atypical strains have not been collected into the database. It should also be considered that natural variability among strains of the same bacterial species may give differences in metabolic types. There also may be loses of some biochemical capabilities during continuing culturing and storage of strains. An effective control measure for plant diseases is to eradicate sources of contamination with the pathogen. The BPB pathogens are seedborne in rice seeds harvested the previous year (Shahjahan et al., 1998b). B. glumae could not be detected from seeds kept outdoors for 5 months ((Tsusima et al., 1989), suggesting that low temperature treatment may kill the bacterium. Dry heat treatment using 65oC for 6 days or salt water selection could not ensure pathogen-free seeds in their studies. Monitoring the contamination of commercial seed-lots seems to be more practical than chemical control or using cold or heat treatments at this time based on current seed production, storage, and planting practices in Louisiana. A seed treatment pesticide to control BPB is not presently labeled for use in the United States. Two hundred and sixty two bacterial isolates from diseased rice tissues showing BPB symptoms were identified as Burkholderia species using the BiologTM GN2 test, comprising 75% of the identified bacteria. Six Burkholderia species were identified including B. glumae, B. gladioli, B. multivorans (B. cepacia genomovar II), B. plantarii, B. vietnamiensis (B. cepacia genomovar V) and B. cocovenenans. Of them, B. gladioli, B. multivorans, B. glumae, and B. plantarii comprised about 90 % of the Burkholderia isolates. Pathogenicity tests on seedlings and emerging panicles showed that six species were rice pathogens causing sheath rot or/and panicle blight. Eighty three percent of the B. gladioli, 91% of B. glumae, 73% of B. multivorans, and 80% of B. plantarii were pathogenic strains. It is therefore suggested that a complex of Burkholderia species caused the BPB syndrome recently found in Louisiana. These species caused seeding blight, seedling leaf sheath rot, and flag leaf sheath brown-black rot, sterile florets, grain discoloration, and/or poor emergence of panicles, similar to the symptoms reported in the literature. Symptom expression alone was not sufficient to distinguish Burkholderia pathogens to species level.

59 This appears to be the first report of B. multivorans as a pathogen on rice. Pathogenicity tests and data analysis showed that B. gladioli and B. glumae had more highly virulent strains than B. multivorans and B. plantarii, suggesting that they played the critical role in pathogenicity on rice plants. Morphology studies showed that convex colonies with smooth margins and yellow or gray-white in color on KB medium were common morphological characteristics among these species. Universally, avirulent strains were associated with gray-white colonies, supporting reports in the literature. Pseudomonas species were shown to be one of the causes of the syndrome. They included two P. syringae pv. zizanize, three P. fluorescens, three P. pyrrocinia, one P. spinosa and seven P. tolaasii strains, suggesting that Pseudomonas species could also be involved in the complex of bacterial species causing BPB on rice in Louisiana. The factors determining whether a given bacterium is the main casual agent may include weather conditions or cultivar and line susceptibility.

60 LITERATURE CITED Atkins, J.G. 1974. Rice Diseases of the Americas: A Review of the Literature. pp. 55-58. Agr. Handbook No. 448, ARS, USDA, Washington, D.C. 106p.

Ashizawa, T., S. Tsushima, H. Nemoto, N. Hayashi and H. Naito. 1997. Effect of cleaning brown rice infected by Pseudomonas glumae. Proceedings of the Kanto-Tosan Plant Protection Society. 44: 23-24.

Ashura, L. K., R. B. Mabagala, C. N. Mortensen. 1999. Isolation and characterization of seed-borne pathogenic bacteria from rice (Oryza sativa L.) in Tanzania. Tanzanian Journal of Agricultural Sciences. 2 (1): 71-80

Azegami, K., K. Nishiyama, Y. Watanabe, T. Suzuki, M. Yoshida, K. Nose and S. Toda. 1985. Tropolone as a root growth-inhibitor produced by a plant pathogenic Pseudomonas sp. Causing seedling blight of rice. Ann. Phyto. Soc. Japan. 51: 315-317.

Azegami, K., K. Nishiyama, Y. Watanabe, I. Kadota, A. Ohuchi, and C. Fukazame. 1987. Pseudomonas plantarii sp. Now., the causal agent of rice seedling blight. Int. J. Syst. Bacteriol. 17: 144-152.

Azegami, K., K. Nishiyama, and H. Tabei. 1988a. Infection courts of rice seedlings with Pseudomonas plantarii and Pseudomonas glumae. Annals of the Phytopathological Society of Japan. 54 (3): 337-341

Azegami, K., H. Tabei, and T. Fukuda. 1988b. Entrance into rice grains of Pseudomonas plantarii, the causal agent of seedling blight of rice. Annals of the Phytopathological Society of Japan. 54 (5): 633-636

Azegami, K. 1996. Media and temperatures favorable for the light production by Burkholderia glumae and B. plantarii transformed with the bioluminescence genes of Vibrio fischeri. Annals of the Phytopathologicl Society of Japan, vol. 62 (2): 161-166

Balandreau, J., Viallard. V, B. Cournoyer, T. Coenye, S. Laevens, and P. Vandamme. 2001. Burkholderia cepacia genomovar III is a common plant-associated bacterium. Appl. Environ. Microbiol. 67: 982-985

Baldani, V.L. D., J. I. Baldni, and J. Dobereiner. 2000. Inoculation of rice plants with the endophytic diazotrophs Herbaspirillum seropedicae and Burkholderia spp. Biology and Fertility of Soils 30 (5/6): 485-491.

61 Ball, S. F. L., and J. C. Reeves. 1991. The application of new techniques in the rapid testing for seed-borne pathogens. Plant Varieties and Seeds 4: 169-176.

Batoko, H. 1994. Inhibition of rice (Oryza-satival) seedling elongation by a Pseudomonas-fuscovaginae toxin. Euphytica 76 (1-2): 139-143.

Batoko, H. J. Bouharmont, and H. Maraite. 1994. Inhibition of rice (Oryza-sativa L.) seedling elongation by a Pseudomonas fuscovaginae toxin. Euphytica 76(1-2): 139-143.

Batoko, H. 1997. Involvement of toxins produced by Pseudomonas fuscovaginae in aetiology of rice bacterial sheath brown rot. Journal of Phytopatology- Phytopathologische Zeitschrift. 145(11-12): 525-531.

Bauernfeind, A., I. Schneider, R. Jungwirth, and C. Roller. 1998. Discrimination of Burkholderia gladioli from other Burkholderia species detectable in cystic fibrosis patients by PCR. Journ. of Clinical Microbiology. 36 (9): 2748-2751

Baxter, I. A., P. I. Lamber and I. N. Simpson. 1997. Isolation from clinical sources of Burkholderia cepacia possessing characteristics of Burkholderia gladioli. Journ. of Antimicrobial Chemotherapy. 39: 169-175

Bevivino, A., S. Sarrocco, C. Dalmastri, S. Tabacchioni, C. Cantale, and L. Chiarini. 1998. Characterization of a free-living maize-rhizosphere population of Burkholderia cepacia: effect of seed treatment on disease suppression and growth promotion of maize. FEMS Microbiol. Ecol. 27:225-237

Brock, T. D., Madigan, M. T., Martinko, J. M., and J. Parker. 1994. Biology of Microbiology. 7th edition. Prentice-Hall, Inc. A Paramount Communications Company Englewood Cliffs, New Jersey 07632. pp. 45-46.

Buckle, K. A. and E. Kartadarma. 1990. Inhibition of bongkrek acid and toxoflavin production in tempe bongkrek containing Pseudomonas cocovenenans. Journal of Applied Bacteriology 68: 571-576.

Burkholder, W. H. 1949. Sour skin, a bacterial rot of onion bulbs. Phytopathology 32: 141-149.

Campbell, P. W., J. A. Phillips, G. J. Heidecker, M. R. S. Krishnamani, R. Zahorchak, and T. L. Stull. 1995. Detection of Pseudomonas (Burkholderia) cepacia using PCR. Pediatr. Pulmonol. 20: 44-49

62 Candlish, A. A. G., J. D. Taylor, and J. Cameron. 1988. Immunological methods as applied to bacterial pea blight. Brighton Crop Protection Conference- Pests and Diseases (1988). 2: 787-794.

Chen, Z. and T. W. Mew. 1998. Relationship between the colonization, concentration and spray timing of antagonistic bacteria and sheath blight of rice. Jiangsu Journ. of Agr. Sci. 14 (1): 31-35.

Cheng, G., A. K. M. Shahjahan, X. L. Yuan, and M. C. Rush. 2001. Distinguishing among isolates of Burkholderia glumae, B. gladioli, B. plantarii, and B. cepacia from Rice. Proc. 29th RTWG. 29: 94.

Chiarini, L., A. Bevivino, S. Tabacchioni, and C. Dalmastri. 1998. Inoculation of Burkholderia cepacia, Pseudomonas fluorescens and Enterobacter sp. on Sorghum bicolor: root colonization and plant growth promotion of dual strain inocula. Soil Biol. Biochem. 30: 81-87

Chien, C. C., and Y. C. Chang. 1987. The susceptibility of Rice Plants at different growth stages and of 21 commercial rice varieties to Pseudomonas glumae. Journ. of Agric. Res. of China 36 (3): 302-310

Coenye, T., P. Vandamme, J. R. W. Govan, and J. J. LiPuma. 2001. and identification of the Burkholderia cepacia complex. Journ. of Clinical Micro. 39: 3427- 3436.

Cottyn, B., M. T. Cerez, M. F. Van Outryve, J. Barroga, J. Swings, and T. W. Mew. 1996a. Bacterial diseases of rice. I. Pathogenic bacteria associated with sheath rot complex and grain discoloration of rice in the Philippines. Plant Dis. 80 (4): 429-437

Cottyn, B., M. F. van. Outryve, M. T. Cerez, M. de. Cleene, J. Swings, and T. W. Mew. 1996b. Bacterial disease of rice. II. Characterization of pathogenic bacteria associated with sheath rot complex and grain discoloration of rice in the Philippines. Plant Dis. 80 (4): 438-445.

Cottyn, B., E. Regalado, B. Lanoot, M. de. Cleene, T. W. Mew, and J. Swings. 2001. Bacterial populations associated with rice seed in the tropical environment. Phytopathology 91 (3): 282-292.

Dalmastri, C., L. Chiarini, C. Cantale, A. Bevivino, and S. Tabacchioni. 1999. Soil type and maize cultivar affect the genetic diversity of maize root-associated Burkholderia cepacia populations. Microbial Ecology 38: 273-284.

63 Daubaras, D. L., C. E. Danganan, A. Hübner, R. W. Ye, W. Hendrickson, and A. M. Chakrabarty. 1996. Biodegradation of 2, 4, 5-trichlorophenoxyacetic acid by Burkholderia cepacia strain AC1100: evolutionary insight. Gene 179:1-8

Duveiller, E., K. Miyajima, F. Snacken, A. Autrique, and H. Maraite. 1988. Characterization of Pseudomonas fuscovaginae and differentiation from other fluorescent pseudomonas occurring on rice in Burundi. J. Phytopathology 122: 97-107.

Estrada-de los santos, P., R. Bustillos-cristales, and J. Caballero-mellado. 2001. Burkholderia, a genus rich in plant-associated nitrogen fixers with wide environmental and geographic distribution. Applied and Environ. Micro. 67: 2790-2798.

Flamand, M. C. 1996. Production of syringotoxin and other bioactive peptides by Pseudomonas fuscovaginae. Physiol. and Mole. Plant Path. 48 (4): 217-231.

Furuya, N., T. Okamoto, Y. Kori, N. Matsuyama, and S. Wakimoto. 1991. Control of bacterial seedling rot of rice by avirulent strains of Pseudomonas glumae. Ann. of the Phytopath. Soc. of Jap. 57 (3): 371-376

Furuya, N., H. Matsuo, K. Koga, N. Matsuyama. 1992. Control of bacterial seedling rot of rice by pretreatment with avirulent strains of Pseudomonas glumae and its mechanisms. Proc. of the Kanto-Tosan Plant Prot. Soc. 38: 4-6.

Furuya, N. 1992. Production of antibacterial substances by Pseudomonas-glumae. Journ. of the Fac. of Agric. Kyushu Univ. 37 (2): 149-158

Furuya, N., K. Liyama, Y. Ueda, and N. Matsuyama. 1997. Reaction of tobacco and rice leaf tissue infiltrated with Burkholderia glumae or B. gladioli. Journ. of the Fac. of Agric. Kyushu Univ. 42 (1-2): 43-51.

Furuya, N., H. Ura, K. Ilyama, M. Matsumoto, M. Takeshita and Y. Takanami. 2002. Specific oligonucleotide primers based on sequences of the 16S-23S rDNA spacer region for the detection of Burkholderia gladioli by PCR. J. Gen. Plant Pathol. 68: 220-224.

Gillis, M., T. V. Van, R. Bardin, M. Goor, P. Hebra, A. Willems, P. Segers, K. Kersters, T. Heulin, and M. P. Fernandez. 1995. Polyphasic taxonomy in the genus Burkholderia leading to an emended description of the genus and proposition of Burkholderia vietnamiensis sp. nov. for N2-fixing isolates from rice in Vietnam. Int. Journ. of Systematic Bact. 45 (2): 274-289.

64 Goto, K., and K. Ohata. 1956. New bacterial disease of rice (brown stripe and grain rot). Ann. Phyto. Soc. Jap. 21: 46-47.

Goto, T., K. Nishiyama and K-I. Ohata. 1987. Bacteria causing grain rot of rice. Ann. Phytopath. Soc. Jap. 53: 141-149.

Groth, D. E. and D. M. Brandon. 1984. Effect of N. fertilizer on disease development in rice. In. Ann. Prog. Rep., Rice Exp. Stn., Crowley, LA 76: 233-243.

Groth, D. E. and C. A. Hollier. 1984. Rice disease survey. In. Ann. Prog. Reg., Rice Exp. Stn., Crowley, LA 77: 250-257.

Groth, D.E., M.C. Rush, and C.A. Hollier. 1991. Rice Diseases and Disorders in Louisiana. Bulletin No. 828, Louisiana State University Agricultural Center. Baton Rouge, Louisiana. 37 p.

Hashioka,Y. 1969. Rice diseases in the world – III Bacterial diseases. Il Riso18: 189-204.

Hanzawa, S. and K. Kaji. 1991. Secondary infection of rice seedling rot in a nursery box, caused by grain rot bacterium, Pseudomonas glumae Kurita and Tabei 1967 and rot of diseased seedlings in paddy field. Ann. Rept. Plant Prot. North Jap. 42: 27-30.

Hebbar, K. P., M. H. Martel, and T. Heulin. 1998. Suppression of pre- and postemergence damping-off in corn by Burkholderia cepacia. Eur. J. Plant Pathol. 104: 29-36

Hikichi, Y., C. Noda, and K. Shimizu. 1989. Oxolinic acid. Jap. Pest. Info. 55: 21-23.

Hikichi, Y. 1993a. Mode of action of oxolinic acid against bacterial seedling rot of rice caused by Pseudomonas glumae. I. Relationship between population dynamics of P. glumae on seedling of rice and disease severity of bacterial seedling rots of rice. Ann. Phyto. Soc. Jap. 59: 441-446.

Hikichi, Y. 1993b. Relationship between population dynamics of Pseudomonas glumae on rice plants and disease severity of bacterial grain rot of rice. J. Pesticide Sci.18: 319- 324.

Hikichi, Y. 1993c. Antibacterial activity of oxolinic acid on Pseudomonas glumae. Annals of the Phytopath. Soc. of Jap. 59 (4): 369-374.

65 Hikichi, Y., T. Okuno, and I. Furusawa. 1993. Immunofluorescent antibody technique for detecting Pseudomonas glumae on rice plants. Annals of the Phytopath. Soc. of Jap. 59 (4): 477-480

Hikichi, Y., T. Okuno, and I. Furusawa. 1994. Susceptibility of rice spikelets to infection with Pseudomonas glumae and its population-dynamics. Journ. of Pest. Sci. 19 (1): 11- 17.

Hikichi, Y. 1995a. Mode-of-action of oxolinic acid against bacterial seedling rot of rice caused by Pseudomonas glumae. II. Efficacy of oxolinic acid against secondary infection. Annals of the Phytopath. Soc. of Jap. 59 (4): 447-451

Hikichi, Y. 1995b. The spread of bacterial grainrot of rice and its control in the paddy fields. J. Pest. Sci. 20: 329-331.

Hikichi, Y. and H. Egami. 1995. Control systems for bacterial grain rot of rice with oxolinic acid and seed selection with salt solution. Annals of the Phytopath. Soc. of Jap. 61 (4): 405-409

Hikichi, Y., T. Okuno, and I. Furusawa. 1995a. Mode of action of oxolinic acid against bacterial seedling rot of rice caused by Pseudomonas glumae. I: Infection with P. glumae into plumules. Annals of the Phytopath. Soc. of Jap. 61 (2): 134-136

Hikichi, Y., T. Okuno and I. Furusawa. 1995b. Mode of action of oxolinic acid against bacterial seedling rot of rice caused by Pseudomonas glumae. III. Infection with P. glumae into plumules. Ann. Phyto. Soc. Japan. 61: 134-136.

Hikichi, Y., H. Egami, Y. Oguri, and T. Okuno. 1998. Fitness for survival of Burkholderia glumae resistant to oxolinic acid in rice plants. Ann. Phyto. Soc. Jap. 64 (3): 147-152.

Holmes, A., J. Govan, and R. Goldstein. 1998. Agricultural use of Burkholderia (Pseudomonas) cepacia: a threat to human health? Emerg. Infect. Dis. 4: 221-227

Hu, F. P., J. M. Young, and C. M. Triggs. 1991. Numerical analysis and determinative tests for nonfluorescent plant-pathogenic Pseudomonas spp. and genomic analsysis and reclassification of species related to Pseudomonas avenae manns 1909. Int. Journ. of Systematic Bact. 41: 516-525.

66 Iguchi, K. 1991. Growth development of rice plant suffering from bacterial seedling blight after transplanting and the population changes of the causal agent, Pseudomonas plantarii in rice plant. Proceed. of the Kanto-Tosan Plant Protect. Soc. 38: 23-24.

Iguchi, K. 1992. SEM observation of the infection site on rice seedling infected with Pseudomonas plantarii. Proceed. of the Kanto-Tosan Plant Protec. Soc. 39: 17-19.

Iiyama, K., N. furuya, Y. Takanami and N. Matsuyama. 1995. A role of phytotoxin in virulence of Pseudomonas glumae Kurita et Tabei. Ann. Phyto. Soc. Japan. 61: 470-476.

Iiyama, K., N. Furuya, H. Ura, and N. Matsuyama. 1998. Role of phytotoxin in the pathogenesis of Burkholderia species. Journ. of the Fac. Agric. Kyushu Univ. 42 (3/4): 289-293.

Isogawa, Y., S. Kato, and S-H. Li. 1989. Studies on environmental factors and control methods of bacterial grain rot of rice caused by Pseudomonas glumae Kurita te Tabei II. Res. Bull. Aichi. Agric. Res. Ctr. 21: 85-90.

Iwai, T., H. Kaku, R. Honkura, S. Nakamura, H. Ochiai, T. Sasaki, and Y. Ohashi. 2002. Enhanced resistance to seed-transmitted bacterial diseases in transgenic rice plants overproducing an oat cell-wall-bound thionin. Mole. Plant Microbe Interaction 15: 515- 521

Jaunet, T. 1996. Pathogenicity of process of Pseudomonas fuscovaginae, the causal agent of sheath brown rot of rice. Journ. of Phytopathology-Phytopathologische Zeitschrift 144 (9-10): 425-430

Jeong, Y., J. Kim, S. Kim, Y. Kang, T. Nagamatsu, and I. Hwang. 2003. Toxoflavin produced by Burkholderia glumae causing rice grain rot is responsible for inducing bacterial wilt in many field crops. Plant Dis. 87:890-895.

Jones, J. B. 1993. Evaluation of the Biolog GN MicroPlate system for identification of some plant-pathogenic bacteria. Plant Dis. 77(6): 553-558

Jurita, T. and H. Tabei. 1967. On the pathogenic bacterium of bacterial grain rot of rice. Ann. Phyto. Soc. Jap. 33: 111.

Kadota, I. And A. Ohuchi. 1983. Symptoms of bacterial brown stripe of rice seedlings in nursery boxes. Ann. Phyto. Soc. Jap. 49: 561-564.

67 Kadota, I. 1996. Studies on the pathogen of bacterial brown stripe of rice and its ecology. Bull. of the Hokuriku Nat. Agr. Expt. Sta. 38: 113-171

Kadota, I., A. Mizuno and K. Nishiyama. 1996. Detection of a protein specific to the strain Pseudomonas avenae Manns 1909 pathogenic to rice. Ann. Phyto. Soc. Jap. 62: 425-428.

Karpati, F., and J. Jonasson. 1996. Polymerase chain reaction for the detection of Pseudomonas aeruginosa, Stenotrophomonas maltophilia and Burkholderia cepacia in sputum of patients with cystic fibrosis. Mol. Cell. Probes 10:397-403

Katoh, T., Y. Fujita and H. Sakuma. 1990. Tropolone method to find rice seed infected with bacterial seedling blight, Pseudomonas plantarii. Ann. Rept. Plant Prot. North Jap. 41: 47-49.

Katsube, K. and S. Takeda. 1997. Suppression of bacterial seedling rot and bacterial seedling blight of rice on the pool nursery. Ann. Rept. of the Soc. of Plant Protect. of North Jap. 48: 43-46

Katsube, K., and S. I. Takeda. 1998. Effect of copper content in seed disinfectants on control efficacy against bacterial seedling rot [Burkholderia glumae] and bacterial seedling blight [B. plantarii] of rice. Ann. Repo. Soc. Plant Prot. North Jap. 49: 33-36.

Katsube, K. and S. Takeda. 1998. Increase in severity of bacterial seedling blight of rice with soil drenching of fungicides. Ann. Rept. Plant Prot. North Jap. 49: 37-39.

Kawaradani, M., K. Okada and S. Kusakari. 2000. New selective medium for isolation of Burkholderia glumae from rice seeds. J. Gen. Plant Pathol. 66: 234-237.

Klement, A. 1955. A new bacterial disease of rice caused by Pseudomonas oryzicola n. sp. Acta Microbiol. Acad. Sci. Hung. 2: 265-274.

Klement, Z. 1983. Detection of seed-borne bacteria by hypersensitive reaction. Seed Sci. and Tech. 11: 589-593

Kodama, K., N. Kimura, and K. Komagata. 1985. Two new species of Pseudomonas: P. oryzihabitans isolated from rice paddy and clinical specimens and P. luteola isolated from clinical specimens. Int. Journ. of Systematic Bact. 35 (4): 467-474.

Kurita, T., H. Tabei, and T. Sato. 1964. A few studies on factors associated with infection of bacterial grain rot of rice. Ann. Phyto. Soc. Jap. 29: 60

68 Kurita, T. and H.Tabei. 1967. On the pathogenic bacterium of bacterial grain rot of rice. Ann. Phyto. Soc. Jap. 33: 111.

Kuwata, H. 1985. Pseudomonas syringae pv. oryzae pv. Nov., causal agent of bacterial halo blight of rice. Ann. Phytopath. Soc. Jap. 51: 212-218.

Lee, F.N. 1992a. Ear blight. p.32. In: Compendium of Rice Diseases. (Ed.) Webster, R.K. and P.S. Gunnell. APS Press. St. Paul, MN. 62p.

Lee, F.N. 1992b. Grain discoloration. pp.31-32. In: Compendium of Rice Diseases. (Ed.) Webster, R.K. and P.S. Gunnell. APS Press. St. Paul, MN. 62p.

Lindberg, G. D., J. M. Larkin and H. A. Whaley. 1980. Production of tropolone by a Pseudomonas. J. Nat. Prod. 43: 592-594.

Lindberg, G. D. 1981. An antibiotic lethal to fungi. Plant Dis. 65: 680-683.

LiPuma, J. 1998. International Burkholderia cepacia Working group. Available: online- http://allserv.rug.ac.be/~tcoenye/cepacia/miami.pdf

Liyama, K., N. Furuya, K. Hara, N. Nakashima, Y. Takanami, and N. Matsuyama. 1994. Phytotoxin produced by Pseudomonas glumae Jurita et Tabei, a causal bacterium of the grain and seedling rot of rice. Journ. of the Fac. of Agr., Kyushu Univ. 38 (3-4): 175-181.

Liyama, K., N. Furuya, Y. Takanami, and N. Matsuyama. A role of phytotoxin in virulence of Pseudomonas glumae Jurita et Tabei. 1995. Annals of the Phytol. Soc. of Jap. 61 (5): 470-476.

Liyama, K., N. Furuya, H. Ura, and N. Matsuyama. 1998. Role of phytotoxins in the pathogensis of Burkholderia species. Lab. of Plant Path., Fac. of Agr., Kyushu Univ. 42 (3/4): 289-293.

LSU Agricultural Center. 1987. Pub. 2321, Rice Production Handbook. p.39. Louisiana Agric. Expt. Sta. Baton Rouge, LA. 63p.

LSU Agricultural Center. 1999. Louisiana Rice Production Handbook. p.81. La. State Univ. Agr. Center. Baton Rouge, LA. 116p.

69 Lu, J.J. and T.T. Chang. 1980. Rice in its temporal and spatial perspectives. In: Rice: Production and Utilization. (Ed.) B.S. Luh. AVI Publishing Co., Inc. Westport, CN. 925p.

Matsuda, I. And Z. Sato. 1987. Ecology of P. glumae, cause bacterial grain rot of rice, from planting to the mature stage. Ann. Phyto. Soc. Jap. 53:122.

McLoughlin, T. J., J. P. Quinn, A. Bettermann, and R. Booklands. 1992. Pseudomonas cepacia suppression of sunflower wilt fungus and role of antifungal compounds in controlling the disease. Appl. Environ. Microbiol. 58: 1760-1763.

Mew, T. W. 1992. Grain Rot. p.9. In: Compendium of Rice Diseases. (Ed.) Webster, R.K. and P.S. Gunnell. APS Press, St. Paul, MN. 62p.

Miyagawa, H. and S. Takaya. 2000. Biological control of bacterial grain rot of rice by avirulent strain of Burkholderia gladioli. Bull. of the Chugoku Nat. Agr. Expt. Sta. 21: 1- 21.

Miyagawa, H. 2000. Biocontrol of bacterial seedling blight of rice caused by Burkholderia gladioli using with its avirulent isolate. Ann. Phyto. Soc. Jap. 66: 232-238.

Miyagawa. H. 2000. Specific detection of pigment-producing strains of Burkholderia glumae by ammonia treatment. Ann. Phyto. Soc. Jap. 66: 246-251.

Miyajima, K., A. Tanii, and T. Akita. 1983. Pseduomonas fuscovaginae sp. Nov., nom. Rev. Int. J. Syst. Bacteriol. 33: 656-657.

Mizuno, A., S. Tsushima, I. Kadota and K. Nishiyama. 1995. A cloned DNA probe for detection of Pseudomonas gladioli. Ann. Phyto. Soc. Jap. 61: 22-26.

O'Callaghan, E. M., M. S. Tanner, and G. J. Boulnois. 1994. Development of a PCR probe test for identifying Pseudomonas aeruginosa and Pseudomonas (Burkholderia) cepacia. J. Clin. Pathol. 47: 222-226.

Ohno, Y., S. Okuda, T. Natsuaki, and M. Teranaka. 1992. Control of bacterial seedling blight of rice by fluorescent Pseudomonas spp. Proceedings of the Kanto-Tosan Plant Protect. Soc. 39: 9-11.

Ohta, K. and R. Tei. 1994. Chemical control of bacterial seedling blight and bacterial grain rot of rice in nursery stage. Proceed. of the Kanto-Tosan Plant Protect. Soc. 41: 9- 11.

70 Ohta, K. 1995. A simple detection method for seed infected with bacterial seedling blight pathogen. Proceed. of the Kanto-Tosan Plant Protect. Soc. 42: 15-17.

Ohta, k. 1995. Control of bacterial seedling blight of rice by vacuum soaking method. Proceed. of the Kanto-Tosan Plant Protect. Soc. 42: 19-21.

Otofuji, M., K. Kadoshige and K. Yoshida. 1988. Persistent part of Pseudomonas glumae Jurita et Tabei in rice plant. Proc. Assoc. Plant Prot. Kyushu. 34: 1-4.

Ou, S. M. 1985. Rice Disease. 2nd. Ed. Commonwealth Mycological Institute, Kew, Surrey, England. 380p.

Palleroni, N. J. and B. Holmes. 1981. Pseudomonas cepacia sp. nov., nom. rev. Int. Journ. of Systematic Bact.. 31 (4): 479-481.

Parke, J. L. and D. Gurian-Sherman. 2001. Diversity of the Burkholderia cepacia complex and implications for risk assessment of biological control strains. Annu. Rev. Phytopathol. 39: 225-258.

Parke, J. L. 1998. Burkholderia cepacia: Friend or Foe? Available: Online: http://www.apsnet.org/online/feature/.

Pelt, C. V., C. M. Verduin, V. H. F. Goessens, M. C. Vos, B. Tummler, C. Segonds, F. Reubsaet, H. Verbrugh, and A. V. Belkum. 1999. Identification of Burkholderia spp. in the clinical microbiology laboratory. Comparison of Conventional and Molecular Methods. 37: 2158-2164.

Reiter, B., A. Glieder, D. Talker, and H. Schwab. 2000. Cloning and characterization of EstC from Burkholderia gladioli, a novel-type esterase related to plant enzymes. Appl. Microbiol. Biotechnol. 54: 778-785.

Rott, P., J. L. Notteghem, and P. Frossard. 1989. Identification and characterization of Pseudomonas fuscovaginae, the causal agent of bacterial sheath brown rot of rice, from Madagascar and other countries. Plant Dis. 73: 133-137.

Rott, P., J. Honegger, J.-L. Notteghem, and S. Ranomenjanahary. 1991. Identification of pseudomonas-fuscovaginae with biochemical, serological, and pathogenicity tests. Plant Dis. 75 (8): 843-846.

Rush, M. C., C. Clark, and D. E. Groth. 1998. Rice bacterial leaf blight found in Louisiana rice. Proc. 22nd RTWG. 22: 65.

71 SAS. 1982. SAS Users Guide: Statistics. Statistical Analysis System, Car, North Carolina. 584 pp.

Sato, Z., Y. Koiso, S. Iwasaki, I. Matsuda, and A. Shirata. 1989. Toxins produced by Pseudomonas glumae. Ann. Phyto. Soc. Jap. 55: 353-356.

Segonds, C., T. Heulin, N. Marty, and G. Chabanon. 1999. Differentiation of Burkholderia species by PCR-restriction fragment length polymorphism analysis of the 16S rRNA gene and application to cystic fibrosis isolates. J. Clin. Microbiol. 37: 2201- 2208

Seki, M. 1959. Study on the bacterial grain rot of rice. Bull. Saga Pref. Agric. Exp. Stn. 2: 131-147.

Shahjahan, A. K. M., M. C. Rush, C. E. Clark, and D. E. Groth. 1998. Bacterial sheath rot and panicle blight of rice in Louisiana. Proc. 27th RTWG. 27: 31-32.

Shahjahan, A. K. M., D. E. Groth, C. Clark, S. D. Linscombe, and M. C. Rush. 2000a. Critical stage of infection and the effect of infected seeds on disease development and yield of rice. Proc. 28th RTWG. 28: 77.

Shahjahan, A. K., M. C. Rush, D. Groth, and C. Clark. 2000b. Panicle Blight. Rice Journ. 15: 26-29.

Shahjahan, A. K. M., D. E. Groth, C. A. Clark, S. D. Linscombe, and M. C. Rush. 2000c. Epidemiological studies on panicle blight of rice: critical stage of infection and the effect of infected seeds on disease development and yield of rice. Proc. 28th RTWG. 28: 77.

Shakya, D.D. and H. S. Chung. 1983. Detection of Pseudomonas avenae in rice seed. Seed Sci. & Tech. 11: 583-587.

Shakya, D. D., F. Vinther and S. B. Mathur. 1985. World wide distribution of a bacterial stripe pathogen of rice identified as Pseudomonas avenae. Phytopath. Z. 114: 256-259.

Shubba, K., S. A. Shetty, H. S. Shetty and N. G. K. Karanth. 1992. Interactions between seed-borne bacteria and fungi of paddy. Int. J. Trop. Plant Dis. 10: 125-130.

Simpson, I. N., J. Finlay, D. J. Winstanley, N. Dewhurst, J. W. Nelson, S. L. Butler and J. R. W. Govan. 1994. Multi-resistance isolates possessing characteristics of both Burkholderia (Pseudomonas) cepacia and Burkholderia gladioli from patients with cystic fibrosis. Journ. of Antimicrobial Chemotherapy 34: 353-361.

72 Sogou, K. and Y. Tsuzaki. 1983. Overwintering of grain rot bacterium Pseudomonas glumae and its infection to rice plant in the paddy field. Proc. Assoc. Plant Prot. Shikotu. 18: 15-20.

Song, W. Y. and K. HyungMoo. 1999. Current status of bacterial grain rot of rice in Korea. Res. in Plant Dis. 5 (1): 1-7.

Stead, D. E. 1992. Grouping of plant-pathogenic and some other Pseudomonas spp. by using cellular fatty acid profiles. Int. Journ. of Systematic Bact.. 42 (2): 281-295.

Sthapit, B. 1995. Inheritance and selection of field-resistance to sheath brown-rot disease in rice. Plant Dis. 79 (11): 1140-1144.

Suslow, T. V., M. N. Schroth, and M. Isaka. 1982. Application of a rapid method for Gram differentiation of plant pathogenic and saprophytic bacteria without staining. Phytopathology 72: 917-918.

Suzuki, F., Y. F. Zhu, H. Sawada, and I. Matsuda. 1998. Identification of proteins involved in toxin production by Pseudomonas glumae. Annals of the Phytopath. Soc. of Jap. 64 (2): 75-79

Tabacchioni, S., A. Bevivino, L. Chiarini, P. Visca, and M. Del Gallo. 1993. Characteristics of two rhizosphere isolates of Pseudomonas cepacia and their potential plant-growth-promoting activity. Microb. Rel. 2:161-168.

Tabei, H., K. Azegami, T. Fukuda, and T. Goto. 1989. Stomatal infection of rice grain with Pseudomonas glumae, the causal agent of the bacterial grain rot of rice. Ann. Phyto. Soc. Jap. 55 (2): 224-228.

Takao, G. 1987. Bacteria causing grain rot of rice. Annals of the Phytopath. Soc. of Jap. 53: 141-149.

Takeuchi, T. 1994. A method for testing the effect of seed disinfection for bacterial seedling blight of rice. Ann. Rept. Plant Prot. North Jap. 45: 21-26.

Takeuchi, T. Influence of cultural conditions in nursery bed on occurrence of bacterial seedling blight of rice. Ann. Rept. Plant Prot. North Jap. 45: 17-20.

Takeuchi, T., H. Sawada, F. Suzuki, and I. Matsuda. 1997. Specific detection of Burkholderia plantarii and B. glumae by PCR using primers selected from the 16S-23S rDNA spacer regions. Annals of the Phytopath. Soc. of Jap. 63 (6): 455-462.

73 Tanaka, T., T. Katoh, and Y. Fujita. 1994. Isolation of Pseudomonas plantarii, pathogen of bacterial seedling blight of rice, from the weeds around the paddy field. Annals of the Phytopath. Soc. of Jap. 60 (5): 640-643.

Tanii, A., T. Baba and T. Haruki. 1974. Bacteria isolated from black rot of rice grains. Ann. Phytopath. Soc. Jap. 40: 309-318.

Tanii, A., K. Miyajima, and T. Akita. 1976. The sheath brown rot disease of rice plant and its causal bacterium, Pseudomonas fuscovaginae, A. Tanii, K. Miyajima, et T. Akita sp. Nov. Ann. Phyto. Soc. Jap. 42: 540-548.

Taxahash, C. 1993. Effect of autoclaving of nursery bed soil on occurrence of bacterial seedling blight Pseudomonas plantarii Vzegara et al. Ann. Rept. Plant Prot. North Jap. 44: 14-15.

Tominaga, T., K. Kimura, and N. Goh. 1983. Bacterial brown stripe of rice in nursery box, caused by Pseudomonas avenae. Ann. Phytopath. Soc. Jap. 49: 463-466.

Tran Van, V. 1994. Growth promotion of rice after inoculation with the nitrogen-fixing bacterium Burkholderia vietnamiensis isolated from an acid-sulfate soil in Vietnam. Agronomie 14 (10): 697-707

Tran Van, V., O. berge, S. Ngo Ke, J. Balandreau and T. Heulin. 2000. Repeated beneficial effects of rice inoculation with a strain of Burkholderia vietnamiensis on early and late yield components in low fertility sulphate acid soils of Vietnam. Plant and Soil 218: 273-284.

Trung, H. M., N. V. Van, N. V. Vien, D. T. Lam, and M. Lien. 1993. Occurrence of rice grain rot disease in Vietnam. Int. Rice Res. Notes 18 (3): 30

Tsushima, S., S. Mogi and H. Saito. 1985. Effects of inoculum density, incubation temperature and incubation period on the development of rice bacterial grain rot. Proc. Assoc. Plant Prot. Kyushu. 31: 11-12.

Tsushima, S., S. Wakimoto and S. Mogi. 1986. Selective medium for detecting Pseudomonas glumae Jurita et Tabei, the causal bacterium of grain rot of rice. Ann. Phytopath. Soc. Japn. 52: 253-259.

Tsushima, S., K. Tsuno, S. Mogi, S. Wakimoto and H. Saito. 1987. The multiplication of Pseudomonas glumae on rice grains. Ann. Phyto. Soc. Japan. 53: 663-667.

74 Tsushima, S., S. Mogi, and H. Naito, and H. Saito. 1989. Existence of Pseudomonas glumae on the rice seeds and development of the simple method for detecting P. glumae from the rice seeds. Bull. of the Kyushu Nat. Agric. Expt. Sta. 25 (3): 261-270.

Tsushima, S. 1991. Reduction of virulence and colonial variation of Pseudomonas glumae cultured on PSA medium. Bull. of the Jyushu Nat. Agric. Expt. Sta. 26 (4): 361- 379.

Tsushima, S. and H. Naito. 1991. Spatial distribution and dissemination of bacterial grain rot of rice caused by Pseudomonas glumae. Ann. Phyto. Soc. Jap. 57 (2): 180-187.

Tsushima, S., S. Mogi, H. Naito, and H. Saito. 1991. Populations of Pseudomonas glumae on rice plants. Ann. Phyto. Soc. Jap. 57: 145-152.

Tsushima, S., C. Narimatsu, A. Mizuno, and R. Kimura. 1994. Cloned DNA probes for detection of Pseudomonas glumae causing bacterial grain rot of rice. Ann. Phyto. Soc. Jap. 60: 576-584.

Tsushima, S., H, Naito, and M. Koitabashi. 1995a. Change in panicle susceptibility associated with flowering rate of spikelets in bacterial grain rot of rice caused by Pseudomonas glumae. Annals of the Phytopath. Soc. of Jap. 61 (2): 109-113.

Tsushima, S., H. Naito, and M. Koitabashi. 1995b. Forecast of yield loss suffered from bacterial grain rot of rice in paddy fields by severely diseased panicles. Annals of the Phytopath. Soc. of Jap. 61 (5): 419-424.

Tsushima, S. 1996. Epidemiology of bacterial grain rots of rice caused by Pseudomonas glumae. JARQ 30 (2): 85-89.

Tsushima, S., H. Naito, and M. Koitabashi. 1996. Population dynamics of Pseudomonas glumae, the causal agent of bacterial grain rot of rice, on leaf sheaths of rice plants in relation to disease development in the field. Ann. Phyto. Soc. Japan. 62: 108-113.

Tyler, S. D., C. A. Strathdee, K. R. Rozee, and W. M. Johnson. 1995. Oligonucleotide primers designed to differentiate pathogenic Pseudomonads on the basis of the sequence of genes coding for 16S-23S rRNA internal transcribed spacers. Clin. Diagn. Lab. Immunol. 2: 448-453.

Uematsu, T., D. Yoshimura, K. Nishiyama, T. Ibaraki, and H. Fuji. 1976. Pathogenic bacterium causing seedling rot of rice. Ann. Phyto. Soc. Jap. 42: 461-471.

75 Uematsu, T. D. Yoshimura, K. Nishiyama, T. Ibaraki, and H. Fujii. 1976. Occurrence of bacterial seedling rot in nursery flat, caused by grain rot bacterium Pseudomonas glumae. Ann. Phyto. Soc. Jap. 42: 310-312.

Ura, H., M. Matsumoto, K. Liyama, N. Furuya, and N. Matsuyama. 1998. PCR-RFLP analysis for distinction of Burkholderia species, causal agents of various rice diseases. Proceed. of the Assoc. for Plant Protect. of Kyushu Univ. 44 (11): 33-66.

Urakami, T., C. Ito-Yoshida, H. Araki, T. Kijima, K-I. Suzuki, and K. Komagata. 1994. Transfer of Pseudomonas plantarii and Pseudomonas glumae to Burkholderia as Burkholderia spp. and description of Burkholderia vandii sp. nov. Int. Journ. of Systematic Bact. 44 (2): 235-245.

Valter, H. B. 1949. Sour skin, a bacterial rot of onion bulbs. Phytopathology 40: 115- 117.

Vandamme, P. A. R. 1996. Polyphasic taxonomy in practice: the Burkholderia cepacia challenge. Available: online: http://www.wfcc.info/NEWSLETTER/.

Vandamme, P., B. Holmes, M. Vancanneyt, T. Coenye, B. Hoste, R. Coopman, H. Revets, S. Lauwers, M. Gillis, K. Kersters, and J. R. W. Govan. 1997. Occurrence of multiple genomovars of Burkholderia cepacia in cystic fibrosis patients and proposal of Burkholderia multivorans sp. nov. Int. J. Syst. Bacteriol. 47: 1188-1200.

Vandamme, P., E. Mahenthiralingam, B. Holmes, T. Coenye, B. Hoste, P. D. Vos, D. Henry, and D. P. Speert. 2000. Identification and population structure of Burkholderia stabilis sp. nov. (formerly Burkholderia cepacia genomovar IV). Journ. of Clinical Microbiol. 38 (3): 1042-1047.

Wakimoto, S., A. Makoto, and K. Tscchiya. 1987. Serological specificity of Pseudomonas glumae, the pathogenic bacteria of grain rot disease of rice. Annals of the Phytopath. Soc. of Jap. 53: 150-158.

Wang, Z., R. Yanagita, K. Tsuchiya, N. Matsuyama and S. Wakimoto. 1991. Relationship between productivity and some other bacteriological properties in the mutant strains of Pseudomonas glumae induced by nitrosoguanidine-treatment. Ann. Phyto. Soc. Japan. 57: 219-224.

Wasano, K. and S-I. Okuda. 1994. Evaluation of resistance of rice cultivars to bacterial grain rot by the syringe inoculation method. Breeding Sci. 44: 1-6.

76 Watanabe, Y., S. Okuda, T. Natsuaki and M. Teranaka. 1991. Chemical control of bacterial seedling blight of rice and the influence of cultural conditions on its occurrence. Proceed. of the Kanto-Tosan Plant Protect. Soc. 38: 19-21.

Welch, D. F. 1991. Applications of cellular fatty acid analysis. Clinical Microb. Rev. 4: 422-438.

Weller, D. M. 2002. Microbial populations responsible for specific soil suppressivness to plant pathogens. Annu. Rev. Phytopathol. 40: 309-348.

Whitby, P. W., K. B. Carter, K. L. Hatter, J. J. Lipuma, and T. L. Stull. 2000. Identification of members of the Burkholderia cepacia complex by species-species PCR. Journ. of Clinical Microbiol. 38 (8): 2962-2965.

Whitby, P. W., L. C. Pope, K. B. Carter, J. J. Lipuma, and T. L. Stull. 2000. Species- specific PCR as a tool for the identification of Burkholderia gladioli. Journ. of Clinical Microbiol. 38 (1): 282-285.

Willems, A., M. Gorr, S. Thielemans, M. Gillis, K. Kersters, and J. De Ley. 1992. Transfer of several phytopathogenic Pseudomonas species to Acidovorax as Acidovorax avenae subsp. avenae subsp. nov., comb. nov., Acidovorax avenae subsp. citrulli, Acidovorax avenae subsp. cattleya, and Acidovorax konjaci. Int. J. Syst. Bacteriol. 42: 107-119.

Xie, G. L., J. C. Zheng, and T. W. Mew. 1999. Pathogenic bacteria associated with rice seed in the subtropic and the tropic areas. Acta Agriculturae Zhjiangensis. 11 (3): 127- 132.

Yabuuchi, E., Y. Kosako, H. Oyaizu, I. Yano, H. Hotta, Y. Hashimoto, T. Ezaki, and M. Arakawa. 1992. Proposal of Burkholderia gen. Nov. and transfer of seven species of the genus Pseudomonas homology group II to the new genus, with the type species Burkholderia cepacia (Palleroni and Holmes 1981) comb. Nov. Microbial. Immun. 36: 1251-1275.

Yabuuchi, E., Y. Kosako, I. Yano, H. Hotta, and Y. Nishiuchi. 1995. Transfer of two Burkholderia and an Alcaligenes species to Ralstonia gen. Nov.: proposal of Ralstonia pickettii (Ralston, Palleroni and Doudoroff 1973) comb. Nov., Ralstonia solanacearum (Smith 1896) comb. Nov. and Ralstonia eutropha (Davis 1969) comb. Nov. Microbiol. Immunol. 39: 897-904.

77 Yamashita, T., N. Kobayashi, N. Eguchi and Y. Saitoh. 1998. Occurrence of bacterial disease (Burkholderia glumae, Burkholderia plantarii) in nursery boxes of rice plants in Nagano prefecture. Proceed. of the Kanto-Tosan Plant Protect. Soc. 45: 11-13.

Yamasaki, S., et al. 1998. Effect of specific ions in agar on antibiotic production by Burkholderia glumae. Journ. of the Faculty of Agri. Kyushu Univ. 43 (3-4): 309-314.

Yasufumi, H. 1995. Mode of action of Oxolinic acid against bacterial seedling rot of rice caused by Pseudomonas glumae. III. Infection with P. glumae into plumules. Annals of the Phytopath. Soc.of Jap. 61: 134-136.

Yasufumi, H. and H. Egami. 1995. Control system for bacterial grain rot of rice with oxolinic acid and seed selection with salt solution. Ann. Phyto. Soc. Jap. 61: 405-409.

Yasunaga, T., H. Matsumoto, and Y. Shigematsu. 1985. Infection and control of bacterial grain rot of rice. Bull. Ehime Pref. Agric. Exp. Stn. 24: 21-28.

Yoneyama, K., Y. Kono, I. Yamaguchi, M. Horikoshi, and T. Hirooka. 1998. Toxoflavin is an essential factor for virulence of Burkholderia glumae causing rice seedling rot disease. Annals of the Phytopath. Soc. of Jap. 64 (2): 91-96.

Yu, X-F., G. Xie, J. Swings, T. W. Mew. 2002. Relationships of different cultural patterns of a predominant japonica rice variety and population variation of antagonistic bacteria against rice sheath blight and the bacterial species. Chinese Journ. of Rice Sci. 16 (4): 351-355.

Zeigler, R. S., and E. Alvarez. 1987. Bacterial sheath brown rot of rice caused by Pseudomonas fuscovaginae in Latin America. Plant Dis. 71: 592-597.

Zeigler, R. S., G. Aricapa, and E. Hoyos. 1987. Distribution of fluorescent Pseudomonas spp. causing grain and sheath discoloration of rice in Latin American. Plant Dis. 71: 896- 900.

Zeigler, R. S. and E. Alvarez. 1989a. Pseudomonas species causing rice sheath rot (ShR) and grain discoloration (GID). Int. Rice Res. Newsletter 14 (1): 26.

Zeigler, R. S. and E. Alvarez. 1989b. Grain discoloration of rice caused by Pseudomonas glumae in LatinAmerica. Plant Dis. 73 (4): 368.

78 Zeigler, R. S. and E. Alvarez. 1989c. Differential culture medium for Pseudomonas species causing sheath rot (ShR) and grain discoloration (GID) of rice. Int. Rice Res. Inst. Newsl. 14: 27-28.

Ziegler, R. S. 1990. Nonfluorescent Pseudomonas strains causing rice sterility and grain discoloration in Colombia. Int. Rice Res. Newsletter 15 (3): 28-29.

Zeigler, G. S. and E. Alvarez. 1990. Characteristics of Pseudomonas spp. causing grain discoloration and sheath rot of rice, and associated Pseudomonas epiphytes. Plant Dis. 74 (11): 917-922.

Many papers were reviewed for this research that covered research that was indirectly related to the research area. As these papers were not cited, but contained information of use in future research with pathogenic bacteria on rice, they were listed in an Appendix as “BIBLIOGRAPHY OF RELATED LITERATURE”.

79 APPENDIX BIBLIOGRAPHY OF RELATED LITERATURE

Abed, Y., A. Davin-Regli, C. Bollet, and P. De Micco. 1995. Efficient discrimination of mycobacterium tuberculosis strains by 16S-23S spacer region-based random amplified polymorphic DNA analysis. J. Clin. Microbiol. 33: 1418-1420.

Aguilar, C., I. Bertani, and V. Venturi. 2003. Quorum-sensing system and stationar-phase sigma factor (rpoS) of the onion pathogen Burkholderia cepacia genomovar I type strain, ATCC 25416. Applied and Environmental Microbiology. 69 (3): 1739-1747.

Aunpad, R., S. P. Muench, P. J. Baker, S. Sedelnikova, W. Panbangred, N. Doukyu, R. Aono, D. W. Rice. 2002. Crystallization and preliminary X-ray crystallographic studies on the class II cholesterol oxidase from Burkholderia cepacia containing bound flavin. ACTA Crystallographica Section D-Biological Crystallography. 58: 2182-2183.

Balandreau, J. A moratorium about Burkholderia use in the environment? Available: Online: http://www.ag.uidaho.edu/bacteriology/Archives/.

Balashova, N. V., I. A. Kosheleva, N. P. Golovchenko and A. M. Boronin. 1999. Phenanthrene metabolism by POseudomonas and Burkholderia strains. Process Biochemistry. 35 (3-4): 291-296.

Baldani, V. L. D., J. I. Baldani, and J. Dobereiner. 2000. Inoculation of rice plants with the endophytic diazotrophs Herbaspirillum seropedicae and Burkholderia spp. Biology and Fertility of Soils. 30 (5-6): 485-491.

Ballio, A., F. Bossa, L. Camoni, D. Di Giorgio, M. C. Flamand, H. Maraite, G. Nitti, P. Pucci, and A. Scaloni. 1996. FEBS Lett 381: 213-216.

Bare, S., V. M. Coiro, A. Scaloni, A. Di Nola, M. Paci, A. L. Segre, and A. Ballio. 1999. Conformations in solution of the fuscopeptins – Phytotoxic metabolites of Pseudomonas fuscovaginae. European Journal of Biochemistry. 266 (2): 484-492.

Batoko, H., A. D. d’Exaerde, J. M. Kinet, J. Bouharmont, R. A. Gage, H. Maraite, and M. Boutry. 1998. Modulation of plant pasma membrane H+ATPase by phytotoxic lipodepsipeptides produced by the plant pathogen Pseudomonas fuscovaginae. Biochimica et Biophysica ACTA-Biomembranes. 1327 (2): 216-226.

Batoko, H., J. Bouharmont, and H. Maraite. 1994. Inhibition of rice (Oryza-sativa L) seedling elongation by a Pseudomonas-Fuscovaginae toxin. Euphytica. 76 (1-2): 139- 143.

80 Batoko, H., J. Bouharmont, J. M. Kinet, and H. Maraite. 1997. Involvement of toxins produced by Pseudomonas fuscovaginae in aetiology of rice bacterial sheath brown rot. Journal of Phytopathology-Phytopathologische Zeitschrift. 145 (11-12): 525-531.

Bauernfeind, A., I. Schneider, R. Jungwirth, and C. Roller. 1998. Discrimination of burkholderia gladioli from other Burkholderia species detectable in cystic fibrosis patients by PCR. J. Cli. Microbiol. 36: 2748-2751.

Bertilsson, S., C. M. Cavandaugh, M. F. Polz. 2002. Sequencing-independent method to generate oligonucleotide probes targeting a variable region in bacterial 16S rRNA by PCR with detachable primers. Appl. Environ. MIcrobiol. 68: 6077-6086.

Bevivino, A., S. Sarrocco, C. Dalmastri, S. Tabacchioni, C. Cantale, and L. Chiarini. 1998. Characterization of a free-living maize-rhizosphere population of Burkholderia cepacia effect of seed treatment on disease suppression and growth promotion of maize. FEMS Microbiology Ecology. 27 (3): 225-237.

Brett, P. J., D. Deshazer, and D. E. Woods. 1997. Characterization of Burkholderia pseudomallei and Burkholderia pseudomallei-like strains. Epidemiol. Infect. 118: 137- 148.

Brett, P. J., M. N. Burtnick, and D. E. Woods. 2003. The wbiA locus is required for the 2-O-acetylation of lipopolysaccharides expressed by Burkholderia pseudomallei and Burkholderia thailandensis. FEMS Microbiology Letters. 218 (2): 323-328.

Brisse, S., C. M. Verduin, D. Milatovic, A. Fluit, J. Verhoef, S. Laevens, P. Vandamme, B. Tummler, H. A. Verbrugh, and A. Belkum. 2000. Distinguishing species of the burkholderia cepacia complex and burkholderia gladioli by automated ribotyping. Journal of Clin. MIcrobiol. 38: 1876-1884.

Bultreys, A., I. Gheysen, B. Wathelet, H. Maraite, and H. Hoffmann. 2003. High- performance liquid chromatography analyses of pyoverdin siderophores differentiate among phytopathogenic fluorescent Pseudomonas species. Appl. Environ. Microbiol. 69: 1143-1153.

Campbell, P. W., J. A. Phillips, G. J. Heidecker, M. R. Krishnamani, R. Zahorchak, T. L. Stull. 1995. Detection of pseudomonas (Burkholderia) cepacia using PCR. Pediatr Pulmonol. 20: 44-49.

Carpena, X., J. Switala, S. Loprasert, S. Mongkolsuk, I. Fita, and P. C. Loewen. 2002. Crystallization and preliminary X-ray analysis of the catalase-peroxidase KatG from Burkholderia pseudomallei. ACTA Crystallographica Section D-Biological Crystallography. 58: 2184-2186.

Cartwright, R. and F. Lee. 2001. Bacterial panicle blight. Available: online: http://www.uaex.edu/.

81

Cerantola, S., J-D. Bounery, C. Segonds, N. Marty and H. Montrozier. 2000. EXopolysaccharide production by mucoid and non-mucoid strains of Burkholderia cepacia. FEMS Microbiology Letters. 185 (2): 243-246.

Cerantola, S. and H. Montrozier. 2001. Production in vitro, on different solid culture media, of two distinct exopolysaccharid by a mucoid clinical strain of Burkholderia capacia. FEMS Microbiology Letters. 202 (1): 129-133.

Clode, P. E., M. E. Kaufmann, H. Malnick, and T. L. Pitt. 1999. Evaluation of three oligonucleotide primer sets in PCR for the identification of Burkholderia cepacia and their differentiation from Burkholderia gladioli. J. of Clin. Pathol. 52 (3): 173-176.

Chen, Z. and T. W. Mew. 1998. Relationship between the colonization, concentration and spray timing of antagonistic bacteria and sheath blight of rice. Jiangsu J. of Agr. Sci. 14 (1): 31-35.

Cheneby, D., L. Philipport, A. Hartmann, C. Henault and J-C, Germon. 2000. 16S rDNA analysis for characterization of denitrifying bacteria isolated from three agricultural soils. FEMS Microbiology Ecology. 34 (2): 121-128.

Cho, Y-S., S-H. Park, C-K. Kim, and K-H. Oh. 2000 Induction of stress shock proteins DnaK and GroEL by phenoxyherbicide 2, 4-D in Burkholderia sp. YK-2 isolated from rice field. Curr. Microbiol. 41: 33-38.

Chung, J. W., E. Altman, T. J. Beveridge, and D. P. Speert. 2003. Colonial morphology of Burkholderia cepacia complex genomovar III: Implications in exopolysaccharide production, pilus expression, and persistence in the mouse. Infection and immunity. 71 (2): 904-909.

Coenye, T., E. Mahenthiralingam, D. Hernry, J. J. LiPuma, S. Laevens, M. Gillis, D. P. Speert, and P. Vandamme. 2001. Burkholderia ambifaria sp. nov., a novel member of the Burkholderia cepacia complex including biocontrol and cystic fibrosis-related isolates. J. Syst. Evol. Microbiol. 51: 1481-1490.

Coenye, T., B. Holmes, K. Kersters, J. R. Govan, and P. Vandamme. 1999. Burkholderia cocovenenans (van Damme et al. 1960) Gillis et al. 1995 and Burkholderia vandii Urakami et al. 1994 are junior synonyms of Burkholderia gladioli (Severini 1913) Yabuuchi et al. 1993 and Burkholderia plantarii (Azegami et al. 1987) Urakami et al 1994, respectively. Int. J. Syst. Bacteriol. 49: 37-42.

Coenye, T., M. Gillis and P. Vandamme. 2000. Pseudomonas antimicrobica attafuah and Bradbury 1990 is a junior synonym of Burkholderia gladioli (Severini 1913) Yabuuchi et al. 1993. Int. J. of Syst. And Evol. Microbiol. 50: 2135-2139.

82

Coenye, T., T. Spilker, A. Martin, and J. J. LiPuma. 2002. Comparative assessment of genotyping methods for epidemiologic study of Burkholderia cepacia genomovar III. J. of Clin. Microbiol. 40 (9): 3300-3307.

Coenye, T. and J. J. LiPuma 2003. Molecular epidemiology of Burkholderia species. Frontiers in Bioscience. 8: 55-67.

Coenye, T. and J. J. LiPuma. 2003. Population structure analysis of burkholderia cepacia genomovar III: varying degrees of genetic recombination characterize major clonal complexes. Microbiology. 149: 77-88.

Cottyn, B. E. Regalado, B. Lanoot, M. De Cleene, T. W. Mew, and J. Swings. 2001. Bacterial populations associated with rice seed in the tropical environment. Phytopathology. 91 (3): 282-292.

Derakshani, M., T. Lukow, and W. Liesack. 2001. Novel bacterial lineages at the (sub) division level as detected by signature nucleotide-targeted recovery of 16S rRNA sgenes from bulk soil and rice roots of flooded rice microcosms. Appl. Environ. Microbiol. 67: 623-631.

Duveiller, F. Snacken, H. Maraite, and A. Autrique. 1989. 1st Detection of Pseudomona- fuscovaginae on maize and sorghum in Burundi. Plant Dis. 73 (6): 514-517.

Duveiller, E., J. L. Notteghem, P. Rott, F. Snacken, and H. Maraite. 1990. Bacterial sheath brown rot of rice caused by pseudomonas-fuscovaginae in Malagasy. Trop. Pest Mgnt. 36 (2): 151-153.

El Khattabi, M., C. Ockhuijsen, W. Bitter, K.-E. Jageger, and J. Tommassen. 1999. Specificity of the lipase-specific foldases of gram-negative bacteria and the role of the membrane anchor. Molecular and General Genetics. 261 (4-5): 770-776.

El Khattabi, M., P. Van Gelder, W. Bitter, and J. Tommassen. 2000. Role of the lipase- specific foldase of Burkholderia glumae as a steric chaperone. J. of Biological Chemistry. 275 (35): 26885-26891.

Engelhard, M., T. Hurek, and B. Reinhold-Hurek. 2000. Preferential occurrence of diazotrophic endophytes, Azoarcus spp., in wild rice species and land races of Oryza sativa in comparison with modern races. Environmental Microbiology. 2 (2): 131-141.

Estrada, P., P. Mavingui, B. Cournoyer, F. Fontaine, J. Balandreau, and J. Caballero- Mellado. 2002. A N2-fixing endophytic Burkholderia sp. associated with maize plants cultivated in Mexico. Can. J. MIcrobiol. 48: 285-294.

83 Estrada-De Los Santos, P., R. Bustillos-Cristales, and J. Caballero-Mellado. 2001. Burkholderia, a genus rich in plant-associated nitrogen fixers with wide environmental and geographic distribution. Appl. Environ. Microbiol. 67: 2790-2798.

Fehlner-Gardiner, C. C., T. M.-H. Hopkins and A. M. Valvano. 2002. Identification of a general secretory pathway in a human isolate of Burkholderia vietnamiensis (formerly B. cepacia complex genomovar V) that is required for the secretion of hemolysin and phospholipase C activities. Microbial. Pathogenesis. 32 (5): 249-254.

Flamand, M.C., S. Pelsser, E. Ewbank, and H. Maraite. 1996. Production of syringotoxin and other bioactive peptides by Pseudomonas fuscovaginae. Physiological and Molecular Plant Pathology. 48 (4): 217-231.

Flore, A., S. Laevens, A. bevivinao, C. Dalmastri, S. Tabacchioni, P. Vandamme, and L. Chiarini. 2001. Burkholderia cepacia complex: distribution of genomovars among isolates from the maize rhizosphere in Italy. Environ. Microbiol. 3: 137-143.

Furuya, N., K. Iiyama, Y. Ueda, and N. Matsuyama. 1997. Reaction of tobacco and rice leaf tissue infiltrated with Burkholderia glumae or B-gladioli. J. of the Faculty of Agr. Kyushu University. 42 (1-2): 43-51.

Furuya, N., K. Iiyama, N. Shiozaki, and N. Matsuyama. 1997. Phytotoxin produced by Burkholderia gladioli. J. of the Faculty of Agr. Kyushu University. 42 (1-2): 33-37.

Furuya, N., Y. Kushima, and N. Matsuyama. 1992. Production of antibacterial substances by Pseudomonas-glumae. J. of the Faculty of Agr. Kyushu University. 37 (2): 149-158.

Furuya, N., U. Hiroyuki, K. Iiyama, M. Matsumoto, M. Takeshita, and Y. Takanami. 2002. Specific oligonucleotide primers baqsed on sequences of the 16S-23S rDNA spacer region for the detection of Burkholderia gladioli by PCR. J. Gen. Plant Pathol. 68: 220- 224.

Galbrath, L., H. Martina, L. Jonsson, C. Rudhe and S. G. Wilkinson. 1999. Lipids and fatty acids of Burkholderia and Ralstonia species. FEMS Microbiol. Letters. 173 (2): 359-364.

Gangadhara, K. P. and A. A. Kunhi. 2000. Protection of tomato seed germination from the inhibitory effect of 2, 4, 5-trichlorophenoxyacetic acid by inoculation of soil with Burkholderia cepacia AC1100. J. Agric. Food Chem. 48: 4314-4319.

Gardan, L., P. Bella, J. M. Meyer, R. Christen, P. Rott, W. Achouak, R. Samson. Int. Syst. Evol. Microbiol. 52: 2065-2074.

Garrec, G. M-L., i. Artaud and C. Capeillere-blandin. 2001. Purification and catalytic properties of the chlorophenol 4-monooxygenase from Burkholderia cepacia strain

84 AC1100. Biochimica et Biophysica Acta (BBA)/Protein Structure and Molecular Enzymology. 1547 (2): 288-301.

Gillis, M., T. Van Van, R. Bardin, M. Goor, P. Hebbar, A. Willems, P. Segers, K. Kersters, T. Heulin, and M. P. Fernandez. 1995. Polyphasic taxonomy in the genus Burkholderia leading to an emended description of the genus and proposition of Burkholderia vietnamiensis sp. nov. for N inferior 2-fixing isolates from rice in Vietnam. Int. J. of Syst. Bacteriol. 45(2): 273-289.

Graves, M., T. Robin, A. M. Chipman, J. Wong, S. Khashe, and J. M. Janda. 1997. Four additional cases of Burkholderia gladioli infection with microbiological correlates and review. Clin. Infect. Dis. 25:838-842.

Gronow, S., C. Noah, A. Blumenthal, B. Lindner, and H. Brade. 2003. Construction of a deep-rough matant of Burkyholderia cepacia ATCC 25416 and characterization of its chemical and biological properties. J. of Biological Chemistry. 278 (3): 1647-1655.

Hanzawa, S. and K. Kaji. 1991. Secondary infection of rice seeding rot in a nursery box, caused by grain rot bacterium, Psedomonas glumae Kurita and Tabei 1967 and rot of diseased seedling in paddy field. Ann. Rept. Plant Prot. North Japan. 42: 27-30.

Hasebe, A., S. Tsushima, and S. Iida. 1998. Isolation and characterization of IS1416 from Pseudomonas glumae, a new member of the IS3 family. Plasmid. 39: 196-204.

Hasebe, A. and S. Iida. 2000. The novel insertion sequence IS 1417, IS1418, and IS1419 from Burkholderia glumae and their strain distribution. Plasmid. 44: 44-53.

Hayashi, K., J. Oyama, A. Kikuia, and K. Fuijj. 1997. The pool-nursery, as a counter- measure of rice bacterial seedling blight. Ann. Rept. Plant Prot. North Japan. 48: 47-49.

Henry, D. A., M. E. Campbell, J. J. LiPuma, and D. P. Speert. 1997. Identificationof Burkholderia cepacia isolates from patients with cystic fibrosis and use of a simple new selective medium. J. Clin. MIcrobiol. 35: 614-619.

Henry, D. A., E. Mahenthiralingam, P. Vandamme, T. Coenye, and D. P. Speert. 2001. Phenotypic methods for determining genomovar status of the Burkholderia cepacia complex. J. Clin. Microbiol. 49: 1073-1078.

Heuer, H. and K. Smalla. 1997. Evaluation of community-level catabolic profiling using Biolog GN microplates to microbial community changes in potato phyllosphere. J. of Microbiological Methods. 30 (1): 49-61.

Hikichi, Y., T. Okuno, and I. Furusawa. 1994. Susceptibility of rice spikelets to infection with pseudomonas-glumae and its population-dynamics. J. of Pest. Sci. 19 (1): 11-17.

85 Hopfl, P., W. Ludwig, K. H. Schleifer, and N. Larsen. 1989. The 23S ribosomal RNA higher-order structure of Pseudomonas cepacia and other prokaryotes. European J. Biochem. 185: 355-364.

Hu, F. P. and J. M. Young. 1998. Biocidal activity in plant pathogenic acidovorax, Burkholderia, Herbaspirillum, Ralstonia and Xanthomonas spp. J. Appl. Microbiol. 84: 263-271.

Hu, F. P., J. M. Young, and M. J. Fletcher. 1998. Preliminary description of biocidal (syringomycin) activity in fluorescent plant pathogenic Pseudomonas species. J. Appl. Microbiol. 85: 365-371.

Iguchi, K. 1991. Growth development of rice plant suffering from bacterial seedling blight after transplanting and the population changes of the causal agent, Pseudomonas plantarii in rice plant. Proceedings of the Kanto-Tosan Plant Protection Society. 38: 23- 24.

Iiyama, K., N. Furuya, K. Hara, N. Nakashima, Y. Takanami, and N. Matsuyama. 1994. Phytotoxin Produced by Pseudomonas-glumae Kurita-et-Tabei, a causal bacterium of the grain and seedling rot of rice. J. of the Faculty of Agr. Kyushu University. 38 (3-4): 175- 181.

Iiyama, K., N. Furuya, H. Ura, and N. Matsuyama. 1998. Role of phytotoxins in the pathogenesis of Burkholderia species. J. of the Faculty of Agr. Kyushu University. 42 (3- 4): 289-293.

Inose, K., M. Fujikawa, T. Yaniazaki, K. Kojima, and K. Sode. 2003. Cloning and expression of the gene encoding catalytic subunit of thermostable glucose dehydrogenase from Burkholderia cepacia in Escherichia coli. Biochimica et Biophysica Acta-Proteins and Proteomics. 1645 (2): 133-138.

Isogawa, Y., S. Kato, and S-H. Li. 1989. Studies on environmental factors and control methods of bacterial grain rot of rice caused by Pseudomonas glumae Kurita et Tabei II. REs. Bull. Aichi. Agric. REs. Ctr. 21: 85-90.

Iwai, T., H. Kaku, R. Honkura, S. Nakamura, H. Ochiai, T. Sasaki, and Y. Ohashi. 2002. Enhanced resistance to seed-transmitted bacterial diseases in transgenic rice plants overproducing an oat cell-wall-bound thion. Mol. Plant Microbe Interact. 15: 515-512.

Jaeger, K-E. and M. T. Reetz. 1998. Microbial lipases form versatile tools for biotechnology. A TRENDS Guide to Proteomics II. 16 (9): 396-403.

Jaunet, T., G. Laguerre, P. Lemanceau, R. Frutos, and J. L. Notteghem. 1995. Diversity of pseudomonas fuscovaginae and other fluorescent pseudomonads isolated from diseased rice. Phytopathology 85 (12): 1534-1541.

86 Jaunet, T., J. L. Notteghem, and F. Rapilly. 1996. Pathogenicity process of Pseudomonas fuscovaginae, the causal agent of sheath brown rot of rice. J. of Phytopathology- Phytopathologische Zeitschrift. 144 (9-10): 425-430.

Kadota, I. 1993. Studies on the infection cycle and control of bacterial brown stripe of rice caused by Pseudomonas avenae. Ann. Phytopath. Soc. Japan 59: 242.

Katoh, T., Y. Fujita and H. Sakuma. 1990. Tropolone method to find rice seed infected with bacterial seedling blight. Ann. Rept. Plant Prot North Japan. 41: 47-49.

Katsube, K. and S. Takeda. 1998. Increase in severity of bacterial seedling blight of rice with soil drenching of fungicides. Ann. Rept. Plant Prot. North Japan. 49: 37-49.

Katsuyoshi, Y., K. Yoshiki, Y. Isamu, H. Mamoru, and H. Takashi. 1998. Toxoflavin is an essential factor for virulence of Burkholderia glumae causing rice seedling rot disease. Ann. of the Phyto. Soc. of Japan. 64 (2): 91-96.

Kenna, D. T., V. A. Barcus, R. J. Langley, P. Vandamme, and J. R. W. Govan. 2003. Lack of correlation between O-serotype, bacteriophage susceptibility and genomovar status in the Burkholderia cepacia complex. FEMS Immunology and Med. Microbiology. 35 (2): 87-92.

Knee, E. M., F. C. Gong, M. Gao, M. Teplitski, A. R. Jones, A. Foxworthy, A. J. Mort, and W. D. Bauer. 2001. Root mucilage from pea and its utilization by rhizosphere bacteria as a sole carbon sourc. Molecular Plant-Microbe Interactions. 14 (6): 775-784.

Koga, K., N. Furuya, Z. W. Wang, S. Wakimoto, and N. Matsuyama. 1993. Pathogenicity of the Tn5 inserted mutant strains of Pseudomonas-glumae. J. of the Faculty of Agr. Kyushu Univ. 37: 359-369.

Kong, H., M. Shimosaka, Y. Ando, K. Nishiyama, T. Fujii, and K. Miyashita. 2001. Species-specific distribution of a modular family 19 chitinase gene in Burkholderia gladioli. FEMS Microbiology Ecology. 37 (2): 135-141.

Koshino, H., Y. Kono, K. Yoneyama, and J. Uzawa. 2000. N-15 NMR study of toxoflavin and fervenulin. Heterocycles. 52 (2): 811-817.

Kotilainen, P., J. Jalava, O. Meurman, O. P. Lehtonen, E. Rintala, O. P. Seppala, E. Eerola, and S. Nikkari. Diagnosis of meningococcal meningitis by broad-range bacterial PCR with cerebrospinal fluid. J. Clin. Microbiol. 36: 2205-2209.

Kreader, C. A. 1995. Design and evaluation of bacteroides DNA probes for the specific detection of human fecal pollution. App.Environ. Microbiol. 61: 1171-1179.

Lagatolla, C., S. Skerlavaj, L. Dolzani, E. A. Tonin, C. M. Bragadin, M. Bosco, R. Rizzo, L. Giglio and P. Cescutti. 2002. Microbiological characterization of Burkholderia

87 cepacia isolates from cystic fibrosis patients: investigation of the exopolysaccharides produced. FEMS Microbiology Letters. 209 (1): 95-102.

Lefebre, M. D. and M. A. Valvano. 2002. Construction and evaluation of plasmid vectors optimized for constitutive and regulated gene expression in Burkholderia cepacia complex isolates. Applied and Environmental Microbiology. 68 (12): 5956-5964.

Leelayuwat, C., A. Romphruk, A. Lulitanond, S. Trajulsomboon and V. Thamlikitkul. 2000. Genotype analysis of Burkholderia pseudomallei using randomly amplified polymorphism DNA (RAPD): indicative of genetic differences amongst environmental and clinical isolates. Acta Tropica. 77 (2): 229-237.

Leu. L. S. and H. C. Yang. 1985. Distribution and survival of sclerotia of rice sheath blight fungus, Thanate phorus cucumeris, in Taiwan. Ann. Phytopath. Soc. Japan 51: 1-7.

Li, H. Q., I. Matsuda, Y. ujise, and A. Ichiyama. 1999. Short-chain acyl-CoA-dependent production of oxalate from oxaloacetate by burkholderia glumae, a plant pathogen which causes grain rot and seedling rot of rice via the oxalate production. J. Biochem. (Tokyo) 126: 243-253.

LiPuma, J. J., B. J. Dulaney, J. D. McMenamin, P. W. Whitby, T. L. Stull, T. Coenye, and P. Vandamme. 1999. Development or rRNA-based pcr assays for identification of Burkholderia cepacia complex isolates recovered from cystic fibrosis patients. J. Clin. Microbiol. 37: 3167-3170.

Liu, P. Y., Z. Y. Shi, Y. J. Lau, B. S. Hu, J. M. Shyr, W. S. Tsai, Y. H. Lin, and C. Y. Tseng. 1995. Comparson of different PCR approaches for characterization of Burkholderia (Pseudomonas) cepacia isolates. J. Clin. MIcrobiol. 33: 3304-3307.

Mattos, K. A., C. Jones, N. Heise, J. O. Previato, and L. Mendonca-Previato. Structure of an acidic exopolysaccharide produced by the diazotrophic endophytic bacterium Burkholderia brasiliensis. European J. of Biochemistry. 268 (11): 3174-3179.

Meyer, J. M., V. T. Van, A. Stintzi, O. Berge, and G. Winkelmann. 1995. Ornibactin production and transport-properties in strains of Burkholderia-vietnamiensis and Burkholderia-cepacia (formerly Pseudomonas-cepacia). Biometals. 8 (4): 309-317.

Meyer, J. M., V. A. Geoffroy, N. Baida, L. Gardan, D. Izard, P. Lemanceau, W. Achouak, N. J. Palleroni. 2002. Siderophore typing, a powerful tool for the identification oflfuorescent and nonfluorescent pseudomonads. Appl. Environ. Microbiol. 68: 2745- 2753.

Michel, V. V., G. L. Hartman, and D. J. Midmore. 1996. Effect of previous crop on soil populations of Burkholderia solanacearum, bacterial wilt, and yield of tomatoes in Taiwan. Plant disease. 80 (12): 1367-1372.

88 Miche, L. and J. Balandreau. 2001. Effects of rice seed surfaces sterilization with hypochlorite on inoculated Burkholderia vietnamiensis. Applied and Environmenal Microbiology. 67 (7): 3046-3052.

Mizuno, A., S. Tsushima, I. Kadota, and K. Nishiyama. 1995. A cloned DNA probe for detection of Pseudomonas gladioli. Ann. Phytopathol. Soc. Japan. 61: 22-26.

Nacamulli, C., A. Bevivino, C. Dalmastri, S. Tabacchioni, and L. Chiarini. 1997. Perturbation of maize rhizosphere microflora following seed bacterization with Burkholderia cepacia MCI 7. FEMS Microbiology Ecology. 23 (3): 183-193.

Nagamatsu, T., H. Yamasaki, T. Hirota, Y. Kido, M. Shibata, and F. Yoneda. 1992. Synthesis of 6-phenyl analogs of toxoflavin and their 4-oxides and antimicrobial activities of toxoflavin and its analogs. J. of Pharmacobio-Dynamics. 15 (7): 93.

Nakata, P. A. 2002. The generation of a transposon-mutagenized Burkholderia glumae library to isolate novel mutants. Plant Science. 162 (2): 267-271.

Nzula, S, P. Vandamme, and J. R. W. Govan. 2002. Influence of taxonomic status on the in vitro antimicrobial susceptibility of the Burkholderia cepacia complex. J. of Antimicrobial Chemotherapy. 50: 265-269.

Oh, C. J., H. B. Kim, and C. S. An. 2003. Molecular cloning and complementation analysis of nifV gene from Frankia EuIK1 strain. Molecules and cells. 15 (1): 27-33.

Ohta, K. and R. Tei. 1994. Chemical control of bacterial seedling blight and bacterial grain rot of rice in nursery stage. Proceedings of the Kanto-Tosan Plant Protection Society. 41: 9-11.

Ohta, K. 1995. A simple detection method for seed infected with bacterial seedling blight pathogen. Proceedings of the Kanto-Tosan Plant Protection Society. 42: 15-17.

Ohta, K. 1995. Control of bacterial seedling blight of rice by vacuum soaking method. Proceedings of the Kanto-Tosan Plant Protection Society. 42: 19-21.

Pallud, C., V. Viallard, J. Balandreau, P. Normand, and G. Grundmmann. 2001. Combined use of aspecific probe and PCAT medium to study Burkholderia in soil. J. Microbiological Methods. 47 (1): 25-34.

Parke, J. L. 1998. Burkholderia cepacia: Friend or Foe? Available: online: http://www.apsnet.org/online/feature/BurkholderiaCepacia/.

Parke, J. L. and D. Gurian-Sherman. 2001. Diversity of the Burkholderia cepacia complex and implications for risk assessment of biological control strains. Annu. Rev. Phytopathol. 39: 225-258.

89 Parsons, Y. N., K. J. Glendinning, V. Thornton, B. A. Hales, C. A. Hart, and C. Winstanley. 2001. A putative type III secretion gene cluster is widely distributed in the Burkholdeia cepacia complex but absent from genomovar I. FEMS Microbiol. Letters. 203 (1): 103-108.

Perez-Prat, E. and M. M. van L. Campagne. 2002. Hybrid seed production and the challenge of propagating male-sterile plants. Trends in Plant Science. 7 (5): 199-203.

Pitt, T. L., S. Trakulsomboon and D. A. B. Dance. 2000. Molecular phylogeny of Burkholderia pseudomallei. Acta Tropica. 74 (2-3): 181-185.

Porteous L. A., F. Widmer, and R. J. Seidler. 2002. Multiple enzyme restriction fragment length polymorphism analysis for high resolution distinction distinction of Pseudomonas (sensu stricto) 16S rRNA genes. J. Microbiol Methods. 51: 337-348.

Putra, S. R., A. Disch, J-M. Bravo, and M. Rohmer. 1998. Distribution of mevalonate and glyceraldehydes 3-phosphate/pyruvate routes for isoprenoid biosynthesis in some Gram- negative bacteria and mycobacteria. FEMS Microbiology Letters. 164 (1): 169-175.

Rathi, P., V. K. Goswami, V. Sahai, and R. Gupta. 2002. Statistical medium optimization and production of a hyperthermostable lipase from Burkholderia cepacia in a bioreactor. J. of Applied Microbiology. 93 (6): 930-936.

Reetz, M. T. and K-E. Jaeger. 1998. Overexpression, immobilization and biotechnological application of Pseudomonas lipases. Chemistry and Physics of Lipids. 93 (1-2): 3-14.

Reiter, B., A. Glieder, D. Talker, and H. Schwab. 2000. Cloning and characterization of EstC from Burkholderia gladioli, a novel-type esterase related to plant enzyme. Appl. Microbiol. Biotechnol. 54: 778-785.

Rehm, B. H. A. 2002. Molecular characterization of the poly (3-hydroxybutyrate) (PHB) SYNTHASE FROM Ralstonia eutropha: in vitro evoluation, site-specific mutagenesis and development of a PHB synthase protein model. Proteins and Proteomics. 1594 (1): 178-190.

Richardson, J., D. E. Stead, J. G. Elphinstone, and R. H. Coutts. 2002. Diversity of Burkholderia isolates from voodland rhizosphere environments. J. Appl. Microbiol. 93: 616-630.

Riedel, K., C. Arevalo-Ferro, G. Reil, A. Gorg, F. Lottspeich, and L. Eberl. 2003. Analysis of the quorum-sensing regulon of the opportunistic pathogen Burkolderia capacia H111 by proteomics. Electrophoresis. 24 (4): 740-750.

Rosenau, F. and K-E. Jaeger. 2000. Bacterial lipases from Pseudomonas: Regulation of gene expression and mechanisms secretion. Biochimie. 82 (11): 1023-1032.

90

Rosenstein, R. and F. Gotz. 2000. Staphylococcal lipases: Biochemical and molecular characterization. Biochimie. 82 (11): 1005-1014.

Rott, P., J. L. Notteghem, and P. Frossard. 1989. Identification and characterization of Pseudomonas-fuscovaginae, the causal agent of bacterial sheath brown rot of rice, from Madagascar and other countries. Plant Disease. 73 (2): 133-137.

Rott, P., J. Honegger, J. L. Notteghem, and S. Ranomenjanahary. 1991. Identification of Pseudomonas-fuscovaginae with biochemical, serological, and pathogenicity tests. Plant Disease. 75 (8): 843-846.

Rumbaugh, K. P., J. A. Griswold, and A. N. Hamood. 2000. The role of quorum sensing in the in vitro virulence of Pseudomonas aeruginose. Microbes and Infection. 2 (14): 1721-1731.

Ryley, H. C., L. Millar-Jones, A. Paull, and J. Weeks. 1995. Characterization of Burkholderia cepacia from cystic fibrosis patients living in Wales by PCR ribotyping. J. Med. Microbiol. 43: 436-441.

Salles, J. F., F. A. De Souza, and J. D. van Elsas. 2002. Molecular method to assess the diversity of Burkholderia species in environmental samples. Appl. Environ. Microbiol. 68: 1595-1603.

Schlacher, A., T. Stanzer, I. Osprian, M. Mischitz, E. Klingsbichel, K. Faber, and H. Schwab. 1998. Detection of a new enzyme for stereoselective hydrolysis of linalyl acetate using sim. Plate assays for the characterization of cloned esterase from Burkholderia gladioli. J. of Biotechnology. 62 (1): 47-54.

Schmitz, F. J., C. R. Mackenzie, B. Hofmann, J. Verhoef, M. Finken-Eigen, H. P. Heinz, and K. Kohrer. 1997. Specific information concerning taxonomy, pathogenicity and methicillin resistance of staphylococci obtained by a multiplex PCR. J. Med. Microbiol. 46: 773-778.

Schloter, M. M. Lebuhn, T. Heulin, and A. Hartmann. 2000. Ecology and evolution of bacterial microdiversity. FEMS. Microbiology Review. 24 (5): 647-660.

Segonds, C., T. Heulin, N. Marty, and G. Chabanon. 1999. Differentiationof Burkholderia species by PCR-restriction fragment length polymorphism analysis of the 16S rRNA gene and application to cystic fibrosis isolates. J. Clin. Microbiol. 37: 2201- 2208.

Segonds, C. and G. Chabanon. 2001. Burkholderia cepacia: dangers of a phytopathogen organism for patients with cystic fibrosis. Ann. Biol. Clin. 59: 259-269.

91 Sermswan, R. W., S. Wongratanacheewin, S. Trakulsomboon, and V. Thamlikitful. 2001. Ribotyping of Burkholderia psedomallei fromclinical and soil isolates in Thailand. Acta Tropica. 80 (3): 237-244.

Silva, L. F., J. G. C. Gomez, M.S. Oliveira, and B. B. Torres. 2000. Propionic acid metabolism and poly-3-hydroxybutyrate-co-3-hydroxyvalerate (P3HB3HV) production by Burkholderia sp. J. of Biotechnology. 76 (2-3): 165-174.

Simpson, I. N., J. Finlay, D. J. Winstanley, N. Dewhurst, J. W. Nelson, S. L. Butler, and J. R. W. Govan. 1994. Multi-resistance isolate possessing characteristics of both Burkholderia (Pseudomonas) cepacia and Burkholderia gladioli from patients with cystic fibrosis. J. Antimicrobial Chemotherapy. 34: 353-361.

Smith-Vaughan, H. C., D. Gal, P. M. Lawrite, C.Winstanley, K. S. Sriprakash, and B. J. Currie. 2003. Ubiquity of putative type III secretion genes among clinical and environmental Burkholderia pseudomallei isolates in Northern Australia. J. of Clin. Microbiol. 41 (2): 883-885.

Smith, M. D., V. Wuthiekanun, A. L. Walsh, and N. J. White. 1995. Quantitative recovery of Burkholderia pseudomallei from soil in Thailand. Transactions of the Royal Society of Tropical Medicine and Hygiene. 89 (5): 488-490. de Souza, J. T. and J. M. Raaijmakers. 2003. Polymorphisms within the prnD and pltC genes from pyrrolnitrin and pyoluteorin-producing Pseudomonas and Burkholderia spp. FEMS Microbiol. Ecology. 43 (1): 21-34.

Sprague, L. D., G. Zysk, R. M. Hagen, H. Meyer, J. Ellis, N. Anuntagool, Y. Gauthier, and H. Neubauer. 2002. A possible pitfall in the identification of Burkholderia mallei using molecular identification systems based on the sequence of the flagellin flic gene. FEMS Immunology and Medical Microbiology. 34 (3): 231-236.

Stead, D. E. 1992. Grouping of plant-pathogenic and some other Pseudomonas spp. by using cellular fatty acid profiles. International Journal of Systematic Bacteriology. 42 (2): 281-295.

Sthapit, B. R., P. M. Pradhanang, and J. R. Witcombe. 1995. Inheritance and selection of field-resistance to sheath brown-rot disease in rice. Plant Disease. 79 (11): 1140-1144.

Stull, T. L., J. J. LiPuma and T. D. Edlind. 1988. A broad-spectrum probe for molecular epidemiology of bacteria: ribosomal RNA. The J. of Infectious Diseases. 157 (2): 280- 285.

Takahash, C. 1993. Effect of Autoclave of nursery bed soil on occurrence of bacterial seedling blight, Pseudomonas plantarii, vergara et al. Ann. Rept. Plant Prot. North Japan. 44: 14-15.

92 Takeuchi, T. 1994. A method for testing the effect of seed disinfection for bacterial seedling blight of rice. Ann. Rept. Plant Prot. North Japan. 45: 21-26.

Takeuchi, T. 1994. Influence of cultural conditions in nursery bed on occurrence of bacterial seedling blight of rice. Ann. Rept. Plant Prot. North Japan. 45: 17-20.

Takikawa, Y., M. Jojima, S. Adachi, M. Kobayashi, T. Jumakura and Y. Inoue. 2000. Cloning and characterization of hrp genes from Burkholderia glumae, Acidovorax avenae, Plantoea agglomerans PV. millettiae and Pseudomonas fluorescens. Available: online: http://www.bspp.org.uk/icpp98/.

Thao, M. L., P. J. Gullan, and P. Baumann. 2002. Secondary () endosymbionts infect the primary (proteobacteria) endosymbionts of mealybugs multiple times and coevolve with their hosts. Appl. And Enviro. Microbiol. 68 (7): 3190-3197.

Tan, M-W. 2002. Cross-species infections and their analysis. Annu. Rev. Microbiol. 56: 539-565.

Tomich, M., A. Griffith, C. A. Herfst, J. L. Burns and C. D. Mohr. 2003. Attenuated virulence of a Burkholderia cepacia type III secretion mutant in a murine model of infection. Infection and Immunity. 73 (3): 1405-1415.

Tsushima, S., C. Narimatsu, A. Mizuno, and R. Kimura. 1994. Cloned DNA probes for detection of Pseudomonas glumae causing bacterial grain rot of rice. Ann. Phytopath. Soc. Japan. 60: 576-584.

Tsushima, S., H. Naito, and M. Koitabashi. 1995. Forecast of yield loss suffered from bacterial grain rot of rice in paddy fields by severely diseased panicles. Ann. Phyto. Soc. Japan. 61: 419-424.

Tungpradabkul, S., S. Wajanarogana, S. Tunpiboonsak, and S. Panyim. 1999. PCR-RFLP analysis of the flagellin sequences for identification of Burkholderia pseudomallei and Burkholderia cepacia from clinical isolates. Mol. Cell Probes. 13: 99-105.

Ulett, G. C., B. J. Currie, T. W. Clair, M. Mayo, N. Ketheesan, J. Labroc, D. Gal, R. Norton, C. A. Smith, J. Barnes, J. Warner, and R. G. Hart. 2001. Burkholderia pseudomallei virulence: definition, stability and association with clonalies. Microbes and Infection. 3 (8): 621-631.

Van Ovevelen, S., R. De Wachter, P. Vandamme, E. Robbrecht, and E. Prinsen. 2002. Identification of the bacterial endosymbionts in leaf galls of Psychotria (Rubiaceae, angiosperms) and proposed of ‘Candidatus Burkholderia Kirkii’ sp. nov. International J. of Syst. And Evolutionary Microbiology. 52 (6): 2023-2027.

Van Pelt, C., C. M. Verduin, W. H. F. Goessens, M. C. Vos, B. Tummler, C. Segonds, F. Reubsaet, H. Verbrugh, and A. Van Belkum. 1999. Identification of Burkholderia spp. in

93 the clinical microbiology laboratory: comparison of conventional and molecular methods. J. of Clin. Microbio. 37: 2158-2164.

Van, V. T., O. Berge, S. N. Ke, J. Balandreau, and T. Heulin. 2000. Repeated beneficial effects ofrice inoculationwith a strain of Burkholderia vietnamiensis on early and late yield components in low fertility sulphate acid soils of Vietnam. Plant and Soil. 218 (1- 2): 273-284.

Van, V. T., O. Berge, J. Balandreau, S. N. Ke, and T. Heulin. 1996. Isolation and nitrogenase activity of Burkholderia vietnamiensis, a nitrogen-fixing bacterium associated with rice (Oryza sativa L) on a sulphate acid soil of Vietnam. Agronomie. 16 (8): 479-491.

Vandamme, P., D. Henry, T. Coenye, S. Nzula, M. Vancanneyt, J. J. LiPuma, D. P. Speert, J. R. W. Govan, and E. Mahenthiralingam. 2002. Burkholderia antina sp. nov. and Burkholderia pyrrocinia, two additional Burkholderia cepacia complex bacteria, may confound results of new molecular diagnostic tools. FEMS Immunology and Medical Microbiology. 33 (2): 143-149.

Vandamme, P., E. Mahenthiralingam, B. Holmes, T. Coenye, B. Hoste, P. De Vos, D. Henry, and D. P. Speert. 2000. Identification and population structure of Burkholderia stabilis sp. nov. (formerly Burkholderia cepacia genomovar IV). J. of Clin. Microbiol. 38 (3): 1042-1047.

Vandamme, P., D. Henry, T. Coenye, S. Nzula, M. Vancanneyt, J. J. LiPuma, D. P. Speert, J. R. Govan, and E. Mahenthiralingam. 2002. Burkholderia anthina sp. nov. and Burkholderia pyrrocinia, two additional Burkholderia cepacia complex bacteria, may confound results of new molecular diagnostic tools. FEMS Immunol. Med. Microbiol. 33: 143-149.

Vandamme, P., J. Goris, W. M. Chen, P. de Vos, and A. Willems. 2002. Burkholderia tuberum sp. nov. and Burkholderia phymatum sp. nov., nodulate the roots of tropical legumes. Systematic and Applied Microbiology. 25 (4): 507-512.

Vandamme, P., B. Holmes, T. Coenye, J. Goris, E. Mahenthiralingam, J. J. LiPuma, and J. R. W. Govan. 2003. Burkholderia cenocepacia sp. nov.- a new twist to an old story. Research in Microbiology. 154 (2): 91-96.

Vandamme, P., B. Pot, M. Gillis, P. de Vos, K. Kersters, and J. Swings. 1996. Polyphasic taxonomy, a consensus approach to bacterial systematics. Microbiol. Rev., 60 (2): 407- 438.

Vermis, K., C. Vandekerckhove, H. J. Nelis, and P. A. R. Vandamme. 2002. Evaluation of restriction fragment length polymorphism analysis of 16S rDNA as a tool for genomovar characterization within the Burkholderia cepacia complex. FEMS Microbiol. Letters. 214 (1): 1-5.

94

Viallard, V., I. Poirier, B. Cournoyer, J. Haurat, S. Wiebkin, K. Ophel-Keller, and J. Balandreau. Burkholderia graminis sp. nov., a rhizospheric Burkholderia species and reassessment of [Pseudomonas] phenazinium, [Pseudomona] pyrrocinia and [Pseudomonas] glathei as Burkholdeia. Int. J. Syst. Bacteriol. 48 (2): 549-563.

Wang, Z. Z., M. S. Ma, and R. Wang. 1996. Enhanced vasocontraction of rat tail arteries by toxoflavin. British J. of Pharmacology. 117 (2): 293-298.

Weller, D. M. 2002. Microbial populations responsible for specific soil suppressiveness to plant pathogens. Annu. Rev. Phytopathol. 40: 309-348.

Whitby, P. W., H. L. Dick, P. W. Campbell, D. E. Tullis, A. Matlow, and T. L. Stull. 1998. Comparison of culture and PCR for detection of Burkolderia cepacia in sputum samples of patients with cystic fibrosis. J. Clin. Microbiol. 36: 1642-1645.

Whitby, P. W., L. C. Pope, K. B. Carter, J. J. LiPuma, and T. L. Stull. 2000. Species- specific PCR as a tool for the identification of Burkholderia gladioli. J. Clin. Microbiol. 38: 282-285.

Whitby, P. W., K. B. Carter, K. L. Hatter, J. J. LiPuma, and T. L. Stull. 2000. Identification of members of the Burkholderia cepacia complex by species-specific PCR. J. Clin. Microbiol. 38: 2962-2965.

Winstanley, C., B. A. Hales, J. A. Morgan, M. J. Gallagher, S. D. Puthucheary, M. F. Cisse, and C. A. Hart. 1999. Analysis of flic variation among clinical isolates of Burkholderia cepacia. J. Med. Microbiol. 48: 657-662.

Woo, P. C., G. K. Woo, S. K. Lau, S. S. Wong, and K. Yuen. 2002. Single gene target bacterial identification. groEL gene sequencing for discriminating clinical isolates of Burkholderia pseudomallei and Burkholderia thailandensis. Diagn. Microbiol. Infect. Dis. 44: 143-149.

Yamasaki, S., N. Furuya, and N. Matsuyama. 1998. Effect of specific ions in agar on antibiotic production by Burkholderia glumae. J. of the Faculty of Agr. Kyushu University. 42 (3-4): 309-314.

Yamashita, T., N. Kobayashi, N. Eguchi, and Y. Saitoh. 1998. Occurrence of Bacterial disease in nursery boxes of rice plants in Nagano prefecture. Proceedings of the Kanto- Tosan Plant Protection Society. 45: 11-13.

Yao, F., H. Zhou, and T. G. Lessie. 2002. Characterization of N-acyl homoserine lactone overproducing mutants of Burkholderia multivorans ATCC 17616. FEMS Microbiol. Letters. 206 (2): 201-207.

95 Yuanbo, D., C. J. Zheng, C. S. Rong, S. X. Xea, and F. C. Tai. 1983. A technique for the detection of Xanthomonas campestris pv. oryzae using bacteriophage. Seed Sci. & Technol. 11: 579-582.

96 VITA

Xianglong Yuan was born on Juan 12, 1962, in Shenyang, Liaoning, P. R., China, where he accomplished his elementary and high school education. In 1985, he enrolled in the Department of Biology at the University of Liaoning. He received a Bachelors of Science honors degree in biology in 1985. Subsequently, he was employed by the Institute of Environment Protection at Shenyang. In 1987, he returned to school and enrolled as a M.S graduate student in the Department of Plant Physiology and Biochemistry in the Division of Basic Science (currently called the College of Biotechnology) at Shenyang Agriculture University. He received a M.S degree in July, 1990. Upon graduation, he was employed by the Division of Basic Science in the College of Traditional Chinese Medicine at Liaoning as a biochemistry lecturer until 1996. Subsequently, he enrolled again as a graduate student in the plant physiology Ph.D. program of the Biology Department at South China Normal University until 1998. In the fall of 2001, he transferred to the Department of Plant Pathology and Crop Physiology at Louisiana State University in Baton Rouge, Louisiana as a graduate student to pursue a Masters degree under the guidance of Dr. M. C. Rush.

97