Frontiers in Science

p-ISSN: 2166-6083    e-ISSN: 2166-6113

2019;  9(2): 33-40

doi:10.5923/j.fs.20190902.01

 

Impacts and Management Strategies of Common Potato (Solanum tuberosum L.) Pests and Diseases in East Africa

Hillary M. O. Otieno1, 2

1Department of Plant Science and Crop Protection, University of Nairobi (UoN), Nairobi, Kenya

2Department of Agricultural Research, One Acre Fund (OAF), Nairobi Kenya

Correspondence to: Hillary M. O. Otieno, Department of Plant Science and Crop Protection, University of Nairobi (UoN), Nairobi, Kenya.

Email:

Copyright © 2019 The Author(s). Published by Scientific & Academic Publishing.

This work is licensed under the Creative Commons Attribution International License (CC BY).
http://creativecommons.org/licenses/by/4.0/

Abstract

Potatoes are attacked by numerous pests and diseases in the farmer fields with high potential to cause low to high yield losses. The direct and indirect nature of impact and the ability to attack at the fields and stores make potato tuber moth the most important pest in the production of potatoes. Moreover, the capacity of pathogens to have a wide range of host plants and ability to live in the soil make diseases such as Bacterial wilt, Late blight, and Verticillium wilt difficult to offer effective and lasting management solutions. For farmers to sustainably manage these constraints, the control must begin right from the initial stage of selection of fields and planting materials- all these must be free of pests and diseases. During production, farmers should manage these pests and diseases through improved soil and nutrient management and chemical application. Use of chemicals should be done with much care to avoid pollution and effect on beneficial organisms that would otherwise cause ecological imbalance.

Keywords: Disease management, Late blight, Meloidogyne spp, Potato tuber moth, Potato virus, Pest management, Solanum tuberosum, Verticillium Wilt

Cite this paper: Hillary M. O. Otieno, Impacts and Management Strategies of Common Potato (Solanum tuberosum L.) Pests and Diseases in East Africa, Frontiers in Science, Vol. 9 No. 2, 2019, pp. 33-40. doi: 10.5923/j.fs.20190902.01.

1. Introduction

Potato (Solanum tuberosum L.) is one of the most important crops in East Africa with a potential to abate hunger due to higher yield per unit area of land compared to maize (UARC, 1990). According to the International Potato Center (CIP) (2008), one hectare of Potato can yield 2-4 times the food value of grain crops and produces more food per unit of water than any other major crop and being up to seven times more efficient in using water than cereals. However, the yields have remained low at 5-9 t/ha compared to the achievable farmer yield of about 20 t/ha in the region (Gildemacher et al., 2009; Al-Dalain, 2009; Namwata et al., 2010). Such a high yield gap is could be linked to pest and disease attacks, among other constraints. Potato yield losses due to pest and disease attacks could be as high as 100% depending on crop tolerance level, climatic conditions, soil, type of pest and disease (Olanya et al., 2002). The prevalence is high due to the warm and wet climatic conditions experienced in the region. A survey carried out on potatoes sold at rural markets in Kenya showed that 99.6% of all sampled tubers had viral diseases (Gildemacher et al. 2007). Such high levels of incidences are due to the constant use of farmer-saved seeds which have caused build-up and transfer of pathogens from generation to the other. Gildemacher et al. (2009) reported that only 41%, 26% and 44% of farmers in Kenya, Uganda, and Ethiopia renew their seeds, a practice which is done after 6, 7 and 3 seasons respectively. Early blight, Common scab, Verticillium wilt, Late blight, Bacterial wilt diseases have been confirmed to affect potato farmers in East Africa (Low, 1997; Tumwine et al., 2002; Nyankanga et al. 2004; Wakahiu et al. 2007; Gildemacher et al., 2009; Muthomi et al., 2012). On the other hand, pest attacks have been confirmed and reported across all potato producing zones. Most important pests reported in the region include Thrips, potato tuber moth (Phthorimaea operculella), cutworm and nematodes (Das & Raman, 1994; Kfir, 2003; Capinera, 2008; Nyasani et al., 2012; Onkedi et al., 2014; Ghebreslassie, 2017). Potato tuber moth is one of the most devastating pests causing up to 100% yield loss due to its capacity to attack at the field and continue affecting tubers at the store (Ojero & Mueke, 1985; Okonya & Kroschel, 2016). The review, therefore, aimed to present economic impacts and management strategies of these common potato pests and diseases for adoption by farmers in East Africa Region.

2. Management of Common Potato Pests and Diseases

Potatoes are affected by several insect pests and diseases causing both direct and indirect yield losses. The general pest and disease management strategy involves proper implementation of integrated pest management (IPM) - a broad-based approach that integrates practices for economic control of pests in the most promising and environmentally safe option. The use of synthetic pesticides under IPM should be done under great care and only in situations where other strategies have proved ineffective or under emergency (Otieno, 2019). Some of the most effective pest and disease management strategies wide wider application include;
Proper monitoring and scouting: The frequency of scouting depends on the pest, crop, weather, and stage of growth. Better scouting process ensures early detection and execution of control strategy. During field scouting, farmers should move in a traverse using the “W” pattern while identifying and recording any symptom and sign of crop attack.
Adoption of better cultural practices: Practices such as proper field selection to avoid areas with already existing potato pests and diseases is a very crucial initial step towards constraint management. This means that proper keeping of farm history records (on diseases and pests) is required for better field selection. The field selection practice is, however, facing threat from the ever-shrinking farm sizes due to subdivision leading to limited space for selection. Proper fields preparation early in the dry season helps kill pathogens before planting crops.
Use of certified and disease and pest free seeds: During production, farmers need to select tubers from healthy mothers, in the case of farmer-saved seeds, or purchase seeds from certified dealers to avoid the spread of pests and diseases. Again, farmers need to reduce the level of seed tuber recycling, a practice that has always been linked with increased pest and disease build-up. Farmers should ensure that only varieties tolerant to these pests and disease and have better yield levels are selected for production.
Practice legume-brassica-potato-cereal rotation system: Crop rotation has been widely accepted as a means of keeping potato pathogens below economic thresholds. The common mechanisms of control are through breaking the lifecycle and repelling of pests. In addition to pathogen control, the program also improves soil structure and fertility through biological nitrogen fixation and soil organic matter build-up when legumes are incorporated.
Use of well-decomposed manure: Use of partially decomposed manure/compost derived from plant materials especially crops residues previously infected by diseases could provide a route for reinfection as pathogens are potentially transferred back into the field. Therefore, proper composting of organic materials including potato residues to kill pathogens and pests before applying back into the farm as manure should be ensured.

2.1. Impacts and Management of Important Potato Pests

Potatoes are affected by a number of insect pests causing both direct and indirect yield losses. Weekly scouting for caterpillars during the first 30 days after planting; throughout the vegetative phase scouting for whiteflies, aphids, and thrips; and during tuber bulking scouting for millipedes and tuber moths is important for prompt initiation of a better control strategy.
2.1.1. Nematode (Meloidogyne spp)
Several nematode species affect potato yields (Medina et al., 2017). The pest damage roots through feeding thereby reducing water, nutrient uptake and accumulation. The attacked tubers have lesions, rot and shrivel leading to far and wide economic losses (Hay et al., 2016; Coyne et al., 2018).
Several control strategies have been recommended: Tuber seed treatment with hot water before planting has been found effective (Youssef, 2013). Use of antagonistic plant species such as Crotalaria spectabilis, C. juncea, Tagetes patula, T. minuta, T. erecta, and Estizolobium spp could be adapted as intercrop or edge plants in the farm (Embrapa, 2015). Rotating potato with barley is effective in controlling Globodera rostochiensis and has reported up to 87% reduction in attack (Senasica, 2013). Retention of crop residues under no-till has also been found to reduce the effect of nematode (Otieno, 2017). Application of biopesticides, use of products developed from microorganisms such as, Pochonia chlamydosporia, Bacillus firmus, Paecillomyces lilacinus, and Trichoderma spp., with the capacity to attach and parasitize on the nematode or eggs leading to their death have been recommended (Atkins et al., 2003; Yang et al., 2012; Castillo et al., 2013; KARI, 2014). Under heavy build-up, application of nematicide chemicals may be necessary- chemicals such as methyl bromide, Aldicarb, fenamiphos, oxamyl, 1, 3 dichloropropene (1, 3‐D), dazomet and metam‐sodium have been found to offer some level of control (Onkendi et al., 2014). These nematicides could be applied about 2 weeks to planting or at planting in the planting holes (KARI, 2014).
2.1.2. Cutworm (Agrotis spp)
The impact of cutworms is high during droughts and in new and succulent sprouts. Due to their sporadic nature, the impact is not easily quantified. However, under heavy infestation, yield loss of about 20-37% has been reported by Atwal, (1976).
To manage the pest, farmers may use family labour to physically collect and kill cutworms from the fields. This should be done in the morning or evening when the temperature is low and pest is actively feeding. This practice could be aided using rice brans placed strategically in the field to attract the worms which are then collected and killed. The crops could also be sprayed continuously with neem leaves or neem seed extract for eco-friendly control (Campos et al., 2016). Use of synthetic chemical compounds such as Chlorpyriphos 20 EC at 2.0 kg/ha, quinalphos 25 EC at 2.0 kg/ha has been found effective towards cutworm control (Tripathi et al., 2003). Poison Bait (Dipterex +Sugar + Rice husk) has also proved effective- these baits are placed strategically on permanent potato plots where they attract and kill the worms that feed on them (Shakur et al., 2007).
2.1.3. Potato Aphids
The most important Aphids species include Myzus persicae, Macrosiphum euphorbiae, and Aulacorthum solani (Harrington et al., 1986). Aphids directly cause yield losses through leaf curl caused by feeding and black sooty mould that grows on the sugary excrement thereby lowering photosynthetic capacity of plants. Indirectly, the pest transmits viral diseases which cause serious losses in potato fields (Robert et al., 2000; Radcliffe & Ragsdale, 2002).
The build-up of this pest could be controlled naturally through common enemies such as ladybird beetles, both adult and grub and syrphid larvae in potato fields (Tschumi et al., 2016). Farmers may encourage survival and multiplication of these natural enemies through reduction in the use of broad-spectrum pesticides and increased cultivation of nectar-producing edge plants to attract these enemies. Under heavy infestation, pesticides could be sprayed to reduce the population. Application of bio-pesticide products (e.g. Lecanicillium; Lecanicillium longisporum (Vertalec®), and Lecanicillium attenuatum (CS625) (Kim et al., 2007); plant extract (e.g. Azatrol (1.2% azadirachtin), Triple Action Neem Oil (70% neem oil) and Pure Neem Oil (100% neem oil) (Shannag et al., 2014); and synthetic chemical formulations (e.g. phorate 10G, carbofuran 3G, acephate 75SP, lindane 6.5 WP) (Konar & Paul, 2005) have been reported to be effective.
2.1.4. Thrips
The Frankliniella schultzei is the commonly found potato thrip species in the region (Nyasani et al., 2012). Thrips suck potato plant sap from soft tissues causing leaves to curl and dry under heavy attack. This pest has also been associated with the transmission of viral diseases in potato fields (Marchoux et al., 1991; Jones, 2005).
The effect of thrips is highly influenced by crop health. Therefore, keeping crops healthy is of great importance towards reducing the level of damage- adequate water and nutrient supply to crops are highly recommended in maintaining crop healthy. Use of reflective mulches (such as silver polyethylene and pieces of cardboard with aluminum foil) to illuminate underneath of leaves has been found to help in repelling these insects (Riley & Pappu, 2000; Reitz & Funderburk, 2012; Lal Bhardwaj, 2013). The use of chemical method is somehow difficult and less effective since thrips hide in leaf sheaths and underneath and flower bracts which are hard to reach with chemical sprays. However, if done well, application of Azadirachta indica formulation (azadirachtin at 0.03% EC), and imidacloprid at 0.02% could be effective (Anuj, 2009).
2.1.5. Potato Tuber Moth (Phthorimaea operculella)
Potato tuber moth is the most important pest affecting production of potatoes in warm and dry conditions experienced in East Africa. The attack begins from the field and proceeds into the store with the caterpillars creating tunnels leading to 100% tuber yield loss if no intervention is made (Alvarez et al. 2005; Ahmed et al., 2013). The excrement in the tunnels also attracts fungal and bacterial growth leading to further infections and damages. The holes created provide secondary infection- entry points for pathogens.
The pest could be controlled naturally through conservation of natural enemies (e.g. wasps and Larvae of lacewings, big-eyed bugs, ground beetles, earwigs, and rove beetles) which prey on all stages of tuber moth (Symington, 2003, Alvarez et al. 2005). Cultural practices such as deep planting (10-15 cm) coupled with 2-3 hilling during growing periods and proper irrigation to avoid cracks during dry periods have been found to reduce pest attack (Hanafi, 1999; Clough et al., 2010). Neem plant has repellent property hence when leaves are used together with mulch could help keep the tuber moths ways. Farmers could also practice improved cropping systems- intercropping potatoes with pepper, onion or peas and rotation with non-solanaceous crops have been found to help in breaking the life cycle of potato tuber moth resulting to low population and build-up on the farm (Lal, 1991). Use of plant-extracts like oil from neem and sunflower seeds could reduce storage losses to 25% (Salem, 1991). However, care must be taken to avoid over-application of these oils that may reduce tuber respiration resulting in low quality as they turn black and become flaccid. Control of potato tuber moth using synthetic chemical products is ineffective once the pest is inside the tubers. However, farmers could spray the plants with neem leaf extracts, Bacillus thuringiensis, fenvalerate, methamidophos, acetamiprid, imidacloprid, and methomyl products which have been found to be effective (Kroschel & Koch, 1996; Rondon, 2010; Vaneva-Gancheva & Dimitrov, 2013).

2.2. Impact and Management of Important Potato Diseases

Diseases are the most important component of biological constraint to potato production as they affect both above and below ground parts. Once in the tubers, the disease continues to develop and multiply even when the products are in stores- meaning 100% yield loss is possible depending on the period and conditions of storage, type of disease and multiplication rate. Therefore, scouting of fungal, viral and bacterial diseases should begin immediately after sprout emergence and continue until harvesting.
2.2.1. Early Blight (Alternaria solani)
The pathogen thrives in warm and wet conditions and attacks both above and below ground potato parts. It lowers the quality of the products through the formation of brown to black, dry corky rot in the tubers. Yield losses of up to 80% have been reported (Stevenson et al. 2001; Olanya et al., 2009; Horsfield et al., 2010; Tsedaley, 2014).
Control of Early blight is difficult due to its capacity to produce huge amounts of secondary inoculum and survival in the soil (Campo et al., 2007; Pasche et al., 2004). Being a soilborne disease, the pathogen could easily be dispersed by irrigation water. Therefore, farmers should regulate water and avoid splashing soils onto plant leaves during irrigation (Olanya et al., 2009). At harvesting, farmers should avoid causing injuries to the tubers that could lead to further contamination and spread in the store. The storage structures should provide cool and aerated conditions that promote rapid suberization of bruises and cut edges to keep away the pathogens as they are unable to infect through intact periderm (Tsedaley, 2014). Chemical application is also a feasible strategy- use of Ridomil Gold, Antracol, and difenoconazole from 30 days after planting has been found effective (Horsfield et al., 2010).
2.2.2. Late Blight (Phytophthora infestans)
The disease attacks crops, mostly during cool, cloudy and wet conditions and may results in 100% yield loss under heavy attack on susceptible varieties. Under good management but without fungicide sprays, yield losses could be 40-50% for the moderately resistant varieties and 50-70% for the more susceptible varieties (Ojiambo et al., 2001; Rahman et al., 2008). Attacks on the fields reduce the photosynthetic capacity of the plants. Due to both field and postharvest losses and control requirement, Late blight is classified as the most expensive disease of potato crop (Cooke et al., 2011).
Proper irrigation management through minimized wetting on potato leaves and ensuring air circulation to dry leaves is important in disease management. Chemical control should be used cautiously and only under heavy attack- Products such as Ridomil Gold, Dithane M-45, Filthane M-45, Secure, Melody Duo have been found effective and recommended for use (Nyankanga et al., 2004; Rasheed & Khan, 2008; Rahman et al., 2008).
2.2.3. Common Scab (Streptomyces scabies)
Common scab incidence has been shown to be related to soil pH, though the extent and optimum level is a point for further debate. Most researchers have leaned towards alkaline conditions (pH 7-8) to reduce disease incidences and yield losses but this needs to be done cautiously without affecting the next crop in the subsequent season (Waterer, 2002). Use of manure has also been floated as a strategy for the management of scab disease (Conn & Lazarovits, 1999). This could be, in parts, due to the ability of manure to raise soil pH and attract natural enemies (Otieno et al., 2018). But the optimum rate of manure application is not clear and is likely to be affected by the quality and availability. Use of Pseudomonas species and vermicompost are some of the technologies which have shown good efficacy and being considered for the management of the disease (Singhai et al., 2011).
2.2.4. Verticillium Wilt (Verticillium albo-atrum/ Verticillium dahlia)
Verticillium wilt disease is caused by a soil-borne pathogen with broad host range – annual to perennial plants- and could cause up to 50% yield loss in the affected fields (Johnson & Dung, 2010). The attacked potato plants wilt and dry up with time. Tubers have discoloration, hence low market value, though they are edible.
Controlling Verticillium wilt disease is very challenging due to the wide range of host plants that exist at any given time in the farmer fields. However, rotating potatoes with cereals and non-host plants have shown a significant reduction of the disease depending on the period of rotation. According to Larkin et al. (2011), significant reduction requires 4-6 years of rotation and the use of green manure. Under irrigation system, farmers are required to maintain soil moisture at 70-75% during the vegetative stage and 80% during tuberization to reduce the amount of infection (Powelson & Rowe, 2008; Johnson & Dung, 2010). Soil fumigation with fumigants such as Metam sodium has been recommended, though the efficacy is highly influenced by soil type, temperature, physical properties, pH, and water holding capacity (Woodward et al., 2011). These conditions are difficult to achieve and regulate under smallholder rainfed systems. Again, soil fumigation is an expensive practice and exposes farmers to highly toxic compounds. For the safety of the applicator, full personal protective equipment should be put on as most of these fumigants are very toxic (Otieno, 2019). Biological control using bacterial antagonists (Bacillus pumilus (M1), Pseudomonas fluorescens Biotype F (DF37), and plant extracts (Astragalus canadensis L) have also been found effective in controlling Verticillium wilt (Uppal et al., 2007). Some of these products are out in the market though likely to be expensive for most of the resource-constrained farmers.
2.2.5. Bacterial Wilt (Ralstonia solanacearum)
Bacterial wilt is a soil-borne disease which can spread by water, farm tools, infected seeds, previously infested crops residue, and volunteer potato crops. When attacked, the whole plant wilts and dries up, tubers get brown-black stain and rot, (hence reduced market quantity and quality). Yield losses of between 50-100% have been estimated and reported in the region (Muthoni et al., 2012; Mwankemwa, 2015).
The pathogen has common host plants such as chili, tomato, tobacco, and eggplant, as well as several species of weeds that are commonly found in farmer fields. Management of this disease using available agrochemicals is difficult. However, farmers should be encouraged to use disease-free planting material of tolerant varieties and select fields that are free from bacterial wilt for crop production (Muthoni et al., 2012). During rotation, cabbage/cauliflower should come immediately before planting potatoes due to their capacity to significantly reduce the pathogen (Larkin et al., 2011).
2.2.6. Viral Disease (Leaf Roll Virus, Virus Y, and Potato Mosaic Virus)
These potato virus diseases have been reported across potato producing zones in the region (Okeyo et al., 2019). The Leaf roll, Virus Y and Mosaic virus have been reported to cause up to 90, 80 and 40%, respectively, reduction on yield from cultivated plants (cipotato.org; Kaniewski et al., 1990). Yield losses of up to 100% have also been quoted in the region (Chiunga, 2013).
There are no clear chemicals that directly control viral diseases in potatoes; however, management could be achieved through the selection of disease-free fields and planting of disease-free and certified seeds from tolerant varieties (Okeyo et al., 2019). Management of vectors, such as whiteflies, aphids, and thrips through judicious use of selective pesticides could help (refer to the section under potato pests for better control of these vectors).

3. Conclusions

Potatoes are attacked by numerous pests and diseases in the farmer fields with high potential to cause low to high yield losses. The direct and indirect nature of impacts and the ability to attack at the fields and stores make potato tuber moth the most important pest in the region. Also, the capacity of pathogens to have a wide range of host plants and live in the soil make it difficult and unstainable to control diseases such as Late blight, Bacterial wilt, and Verticillium wilt. For farmers to sustainably manage these constraints, control must begin right from the selection of appropriate field and planting materials- all the must be free from pest and disease. However, land subdivision due to the high population growth rate is threatening the capacity of the farmers to select fields and rotate potato crops as farms become limited. Proper soil and nutrient management is important as the crops’ tolerance levels are increased with better nutrition. Chemical application should be done with much care to avoid pollution and effect on beneficial insects that would otherwise cause ecological imbalance. Again, most of the farm chemicals are toxic to farmers and should be applied with maximum protection.

References

[1]  Ahmed, A. A. I., Hashem, M. Y., Mohamed, S. M., & Khalil, S. S. (2013). Protection of potato crop against Phthorimaea operculella (Zeller) infestation using frass extract of two noctuid insect pests under laboratory and storage simulation conditions. Archives of Phytopathology and Plant Protection, 46(20), 2409-2419.
[2]  Al-Dalain SA (2009). Effect of intercropping of Zea mays with potato Solanum tuberosum, L. on potato growth and on the productivity and land equivalent ratio of potato and Zea mays. Agric. J. 4(3): 164-170.
[3]  Alvarez JM, Dotseth E, Nolte P. (2005). Potato tuberworm a threat for Idaho potatoes. University of Idaho Extension, Idaho Agricultural Experiment Station, Moscow, ID. (31 Jan 2014).
[4]  Anuj, B. (2009). Efficacy and economics of insecticides and bio-pesticides against thrips on potato. Annals of Plant Protection Sciences, 17(2), 501-503.
[5]  Atkins, S. D., Hidalgo‐Diaz, L., Kalisz, H., Mauchline, T. H., Hirsch, P. R., & Kerry, B. R. (2003). Development of a new management strategy for the control of root‐knot nematodes (Meloidogyne spp) in organic vegetable production. Pest Management Science: formerly Pesticide Science, 59(2), 183-189.
[6]  Atwal, A.S. 1976. Agricultural Pests of India and South East Asia. Kiliani Publishers Delhi. 502p.
[7]  Campo, A.R.O., Zambolim L, and Costa, L.C. (2007). Potato early blight epidemics and comparison of methods to determine its initial symptoms in a potato field. Rev. Fac. Nal. Agr. Medellin. 60-2:3877-3890.
[8]  Campos, E. V., de Oliveira, J. L., Pascoli, M., de Lima, R., & Fraceto, L. F. (2016). Neem oil and crop protection: from now to the future. Frontiers in plant science, 7, 1494.
[9]  Capinera, J. L. (2008). Potato tuberworm, Phthorimaea operculella (Zeller)(Lepidoptera: Gelechiidae). Encyclopedia of Entomology, 3027-3030.
[10]  Castillo, J. D., Lawrence, K. S., & Kloepper, J. W. (2013). Biocontrol of the reniform nematode by Bacillus firmus GB-126 and Paecilomyces lilacinus 251 on cotton. Plant Disease, 97(7), 967-976.
[11]  Chiunga, E. (2013). Viruses occurring in potatoes (Solanum tuberosum) in Mbeya region, Tanzania.
[12]  Clough, G. H., Rondon, S. I., DeBano, S. J., David, N., & Hamm, P. B. (2010). Reducing tuber damage by potato tuberworm (Lepidoptera: Gelechiidae) with cultural practices and insecticides. Journal of economic entomology, 103(4), 1306-1311.
[13]  Conn, K. L., & Lazarovits, G. (1999). Impact of animal manures on verticillium wilt, potato scab, and soil microbial populations. Canadian Journal of Plant Pathology, 21(1), 81-92.
[14]  Cooke, L. R., Schepers, H. T. A. M., Hermansen, A., Bain, R. A., Bradshaw, N. J., Ritchie, F., ... & Andersson, B. (2011). Epidemiology and integrated control of potato late blight in Europe. Potato Research, 54(2), 183-222.
[15]  Coyne, D. L., Cortada, L., Dalzell, J. J., Claudius-Cole, A. O., Haukeland, S., Luambano, N., & Talwana, H. (2018). Plant-parasitic nematodes and food security in sub-Saharan Africa. Annual Review of Phytopathology, 56, 381-403.
[16]  Das, G. P., & Raman, K. V. (1994). Alternate hosts of the potato tuber moth, Phthorimaea operculella (Zeller). Crop Protection, 13(2), 83-86.
[17]  Embrapa (2015) Nematoides na cultura da batata. Circular Técnica;143:1-12.
[18]  Ghebreslassie, B. M. (2017). Characterization of Potato (Solanum tuberosum L.) Cultivars Grown in Eritrea Using Morphological, Molecular and Nutritional Descriptors (Doctoral dissertation, Faculty of Agriculture, JKUAT).
[19]  Gildemacher, P. R., Kaguongo, W., Ortiz, O., Tesfaye, A., Woldegiorgis, G., Wagoire, W. W., ... & Leeuwis, C. (2009). Improving potato production in Kenya, Uganda, and Ethiopia: a system diagnosis. Potato research, 52(2), 173-205.
[20]  Gildemacher, P.R., J. Mwangi, P. Demo, and I. Barker. (2007). Prevalence of potato viruses in Kenya and consequences for seed potato system research and development. In7th triennial African PotatoAssociation conference, ed. A. Khalf-Allah, 84–92. Alexandria: African Potato Association.
[21]  Hanafi, A. (1999). Integrated pest management of potato tuber moth in field and storage. Potato Research, 42(2), 373-380.
[22]  Harrington, R., Katis, N., & Gibson, R. W. (1986). Field assessment of the relative importance of different aphid species in the transmission of potato virus Y. Potato Research, 29(1), 67-76.
[23]  Hay, F. S., Herdina, Ophel-Keller, K., Hartley, D. M., & Pethybridge, S. J. (2016). Prediction of potato tuber damage by root-knot nematodes using Quantitative DNA Assay of Soil. Plant disease, 100(3), 592-600.
[24]  Horsfield, A., Wicks, T., Davies, K., Wilson, D., & Paton, S. (2010). Effect of fungicide use strategies on the control of early blight (Alternaria solani) and potato yield. Australasian Plant Pathology, 39(4), 368-375.
[25]  International Potato Center (CIP), (2008). 2008 – The International Year of Potato. Available from http://www.cipotato.org/pressroom/facts_figures/2008_international_year_of_the_potato.asp] site visited on 1/5/2009.
[26]  International Potato Center: https://cipotato.org/crops/potato/
[27]  Johnson, D. A., & Dung, J. K. (2010). Verticillium wilt of potato–the pathogen, disease and management. Canadian Journal of Plant Pathology, 32(1), 58-67.
[28]  Jones, D. R. (2005). Plant viruses transmitted by thrips. European journal of plant pathology, 113(2), 119-157.
[29]  Kaniewski, W., Lawson, C., Sammons, B., Haley, L., Hart, J., Delannay, X., & Tumer, N. E. (1990). Field resistance of transgenic Russet burbank potato to effects of infection by potato virus X and potato virus Y. Bio/technology, 8(8), 750.
[30]  Kenya Agricultural & Livestock Research Organization (2014) Root knot nematode. KARI e‐Mimi Fact Sheet No.10/2014. Available at https://0x9.me/1PNiZ.
[31]  Kfir, R. (2003). Biological control of the potato tuber moth Phthorimaea operculella in Africa. Biological control in IPM systems in Africa, 77-85.
[32]  Kim, J. J., Goettel, M. S., & Gillespie, D. R. (2007). Potential of Lecanicillium species for dual microbial control of aphids and the cucumber powdery mildew fungus, Sphaerotheca fuliginea. Biological Control, 40(3), 327-332.
[33]  Konar, A., & Paul, S. (2005). Comparative field efficacy of synthetic insecticides and bio-pesticides against aphids on potato. Annals of Plant Protection Sciences, 13(1), 34-36.
[34]  Kroschel, J., & Koch, W. (1996). Studies on the use of chemicals, botanicals and Bacillus thuringiensis in the management of the potato tuber moth in potato stores. Crop Protection, 15(2), 197-203.
[35]  Lal Bhardwaj, R. (2013). Effect of mulching on crop production under rainfed condition-a review. Agricultural Reviews, 34(3).
[36]  Lal, L. (1991). Effect of intercropping on the incidence of potato tuber moth, Phthorimaea operculella (Zeller). Agriculture, ecosystems & environment, 36(3-4), 185-190.
[37]  Larkin, R. P., Honeycutt, C. W., & Olanya, O. M. (2011). Management of Verticillium wilt of potato with disease-suppressive green manures and as affected by previous cropping history. Plant Disease, 95(5), 568-576.
[38]  Low JW (1997) Potato in southwest Uganda: threats to sustainable production. Afr Crop Sci J 5:395–412.
[39]  Marchoux, G., Gebre‐Selassie, K., & Villevieille, M. (1991). Detection of tomato spotted wilt virus and transmission by Frankliniella occidentalis in France. Plant Pathology, 40(3), 347-351.
[40]  Medina, I. L., Gomes, C. B., Correa, V. R., Mattos, V. S., Castagnone-Sereno, P., & Carneiro, R. M. (2017). Genetic diversity of Meloidogyne spp. parasitising potato in Brazil and aggressiveness of M. javanica populations on susceptible cultivars. Nematology, 19(1), 69-80.
[41]  Muthoni, J., Shimelis, H., & Melis, R. (2012). Management of Bacterial Wilt [Rhalstonia solanacearum Yabuuchi et al., 1995] of Potatoes: Opportunity for Host Resistance in Kenya. Journal of Agricultural Science, 4(9), 64.
[42]  Mwankemwa, Z. (2015). Occurrence and distribution of potato bacterial wilt disease and variability of its causal agent in southern highlands of Tanzania (Doctoral dissertation, Sokoine University of Agriculture).
[43]  Namwata BML, Lwelamira J, Mzirai OB (2010). Adoption of improved agricultural technologies for Irish potatoes (Solanum tuberosum) among farmers in Mbeya Rural district, Tanzania: A case of Ilungu ward. J. Animal and Plant Sci. 8:1:927- 935.
[44]  Nyankanga RO, Wien HC, Olanya OM, Ojiambo PS (2004) Farmers' cultural practices and management of potato late blight in Kenya highlands: implications for development of integrated disease management. Int J Pest Manage 50:135–144.
[45]  Nyasani, J. O., Meyhöfer, R., Subramanian, S., & Poehling, H. M. (2012). Effect of intercrops on thrips species composition and population abundance on French beans in Kenya. Entomologia Experimentalis et Applicata, 142(3), 236-246.
[46]  Ojero, M. F. O., & Mueke, J. M. (1985). Resistance of four potato varieties to the potato tuber moth, Phthorimaea operculella (Zell.) in storage. International Journal of Tropical Insect Science, 6(2), 205-207.
[47]  Ojiambo, P. S., Namanda, S., Olanya, O. M., El-Bedewy, R., Hakiza, J. J., Adipala, E., & Forbes, G. (2001). Impact of fungicide application and late blight development on potato growth parameters and yield in the tropical highlands of Kenya and Uganda. African Crop Science Journal, 9(1), 225-233.
[48]  Okeyo, G. O., Narla, R. D., Otieno, H. M. O., & Schulte-Geldermann, E. (2019). Response of Selected Potato Genotypes to Natural Virus Infection in the Field. Asian Journal of Research in Crop Science, 1-13.
[49]  Okonya, J. S., & Kroschel, J. (2016). Farmers’ knowledge and perceptions of potato pests and their management in Uganda. Journal of Agriculture and Rural Development in the Tropics and Subtropics (JARTS), 117(1), 87-97.
[50]  Olanya, O. M., El-Bedewy, R., Adipala, E., Hakiza, J. J., Namanda, S., Kakuhenzire, R., Wagoire, W. W., Angiyah, T., Karinga, J., Ewell, P. & Lungaho, C. (2002). Estimation of yield loss caused by late blight and the effects of environmental factors on late blight severity in Kenya and Uganda. African Crop Science Proceedings, 5, 455—460.
[51]  Olanya, O. M., Honeycutt, C. W., Larkin, R. P., Griffin, T. S., He, Z., & Halloran, J. M. (2009). The effect of cropping systems and irrigation management on development of potato early blight. Journal of general plant pathology, 75(4), 267-275.
[52]  Onkendi, E. M., Kariuki, G. M., Marais, M., & Moleleki, L. N. (2014). The threat of root‐knot nematodes (Meloidogyne spp.) in Africa: a review. Plant Pathology, 63(4), 727-737.
[53]  OTIENO, H. M. O. (2017). NUTRIENT MANAGEMENT OPTIONS FOR ENHANCING PRODUCTIVITY OF MAIZE AND BEANS UNDER CONSERVATION AND CONVENTIONAL TILLAGE SYSTEMS (Doctoral dissertation, UNIVERSITY OF NAIROBI).
[54]  Otieno, H. M. O. (2019). Pesticide Training Tool: A Simplified Guide for Agricultural Extension Officers and Farmers. Asian Journal of Research in Crop Science, 3(4), 1-5. https://doi.org/10.9734/ajrcs/2019/v3i430056.
[55]  Otieno, H. M., Chemining’Wa, G. N., & Zingore, S. (2018). Effect of farmyard manure, lime and inorganic fertilizer applications on soil pH, nutrients uptake, growth and nodulation of soybean in acid soils of western Kenya. J Agr Sci, 10, 199-208.
[56]  Pasche J.S., Wharam C.M., and Gudmestad, N.C. (2004). Shift in sensitivity of Alternaria solani in response to Q (0) I fungicides. Plant Dis. 88: 181–187.
[57]  Powelson, M., and R.C. Rowe. (2008). Managing diseases caused by seedborne and soilborne fungi and fungus-like pathogens. In Potato Health Management pp. 183–195, 2nd ed. D.A. Johnson, ed. The American Phytopathological Society. St. Paul, MN.
[58]  Radcliffe, E. B., & Ragsdale, D. W. (2002). Aphid-transmitted potato viruses: the importance of understanding vector biology. American journal of potato research, 79(5), 353-386.
[59]  Rahman, M. M., Dey, T. K., Ali, M. A., Khalequzzaman, K. M., & Hussain, M. A. (2008). Control of late blight disease of potato by using new fungicides. Int. J. Sustain. Crop Prod, 3(2), 10-15.
[60]  Rasheed, A., & Khan, S. A. (2008). Relative efficacy of various fungicides, chemicals and biochemical against late blight of potato. Pakistan Journal of Phytopathology (Pakistan).
[61]  Reitz, S. R., & Funderburk, J. (2012). Management strategies for western flower thrips and the role of insecticides. In Insecticides-Pest Engineering. IntechOpen.
[62]  Riley, D. G., & Pappu, H. R. (2000). Evaluation of tactics for management of thrips-vectored tomato spotted wilt virus in tomato. Plant disease, 84(8), 847-852.
[63]  Robert, Y., Woodford, J. T., & Ducray-Bourdin, D. G. (2000). Some epidemiological approaches to the control of aphid-borne virus diseases in seed potato crops in northern Europe. Virus Research, 71(1-2), 33-47.
[64]  Rondon, S. I. (2010). The potato tuberworm: a literature review of its biology, ecology, and control. American Journal of Potato Research, 87(2), 149-166.
[65]  Salem, S. A. (1991). Evaluation of neem seed oil as tuber protectant against Phthorimaea operculella Zell. (Lepidoptera, Gelechiidae). Annals of Agricultural Science, Moshtohor, 29(1), 589-595.
[66]  Senasica (2013) Nematodo dorado de la papa- Globodera rostochiensis – Mexico – Segarpa. Ficha Técnica;19:1-24.
[67]  Shakur, M., Ullah, F., Naem, M., Amin, M., Saljoqi, A. U. R., & Zamin, M. (2007). Effect of various insecticides for the control of potato cutworm (Agrotis ipsilon huf, Noctuidae: Lepidoptera) at Kalam Swat. SARHAD JOURNAL OF AGRICULTURE, 23(2), 423.
[68]  Shannag, H. S., Capinera, J. L., & Freihat, N. M. (2014). Efficacy of different neem-based biopesticides against green peach aphid, Myzus persicae (Hemiptera: Aphididae). International Journal of Agricultural Policy and Research, 2(2), 61-68.
[69]  Singhai, P. K., Sarma, B. K., & Srivastava, J. S. (2011). Biological management of common scab of potato through Pseudomonas species and vermicompost. Biological control, 57(2), 150-157.
[70]  Stevenson WR, Loria R, Franc GD, Weingartner DP (Eds) (2001) ‘Compendium of potato diseases.’ 2nd edn. (American Phytopathological Society: St. Paul, MN, US).
[71]  Symington, C. A. (2003). Lethal and sublethal effects of pesticides on the potato tuber moth, Phthorimaea operculella (Zeller) (Lepidoptera: Gelechiidae) and its parasitoid Orgilus lepidus Muesebeck (Hymenoptera: Braconidae). Crop Protection, 22(3), 513-519.
[72]  Tripathi, D. M., Bisht, R. S., & Mishra, P. N. (2003). Bio-efficacy of some synthetic insecticides and bio-pesticides against black cutworm, Agrotis ipsilon infesting potato (Solanum tuberosum) in Garhwal Himalaya. Indian Journal of Entomology, 65(4), 468-473.
[73]  Tschumi, M., Albrecht, M., Collatz, J., Dubsky, V., Entling, M. H., Najar‐Rodriguez, A. J., & Jacot, K. (2016). Tailored flower strips promote natural enemy biodiversity and pest control in potato crops. Journal of applied ecology, 53(4), 1169-1176.
[74]  Tsedaley, B. (2014). Review on early blight (Alternaria spp.) of potato disease and its management options. J Biol Agric Healthc, 4(27), 191-199.
[75]  Tumwine, J., Frinking, H. D., & Jeger, M. J. (2002). Tomato late blight (Phytophthora infestans) in Uganda. International journal of pest management, 48(1), 59-64.
[76]  Uppal, A. K., El Hadrami, A., Adam, L. R., Tenuta, M., & Daayf, F. (2008). Biological control of potato Verticillium wilt under controlled and field conditions using selected bacterial antagonists and plant extracts. Biological Control, 44(1), 90-100.
[77]  Uyole Agricultural Research Centre (UARC) (1990). Kilimo Bora cha Viazi Mviringo. Extension Leaflet No. 49. Shirika la Kilimo Uyole, Mbeya, Tanzania. 20pp.
[78]  Vaneva-Gancheva, T., & Dimitrov, Y. (2013). Chemical control of the potato tuber moth Phthorimaea operculella (Zeller) on tobacco. Bulg J Agric Sci, 19(5), 1003-1008.
[79]  Wakahiu MW, Gildemacher PR, Kinyua ZM, Kabira JN, Kimenju AW, Mutitu EW (2007) Occurrence of potato bacterial wilt caused by Ralstonia solanacearum in Kenya and opportunities for intervention. In: 7th triennial African Potato Association conference. African Potato Association, Alexandria, pp 267–271.
[80]  Waterer, D. (2002). Impact of high soil pH on potato yields and grade losses to common scab. Canadian Journal of plant science, 82(3), 583-586.