Review Article

Potato Production under Drought Conditions: Identification of Adaptive Traits  

Jane Muthoni , J.N. Kabira
Kenya Agricultural and Livestock Research Organisation (KALRO),Tigoni, Kenya
Author    Correspondence author
International Journal of Horticulture, 2016, Vol. 6, No. 12   doi: 10.5376/ijh.2016.06.0012
Received: 22 Feb., 2016    Accepted: 01 Apr., 2016    Published: 21 May, 2016
© 2016 BioPublisher Publishing Platform
This is an open access article published under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Preferred citation for this article:

 Muthoni J., and Kabira J.N., 2016, Potato production under drought conditions: identification of adaptive traits, International Journal of Horticulture, 6(12): 1-9 (doi: 10.5376/ijh.2016.06.0012)

Abstract

Globally, potato is the third most important food crop after rice and wheat in terms of consumption. It is fast maturing, versatile in use and grows in a wide range of environments from sea level up to about 4700 metres above sea level and from Southern Chile to Greenland. Currently, potato production is rapidly expanding beyond the traditional high potential areas. This is mostly due to expanding world population accompanied by increasing food demand. However, potato is relatively susceptible to yield loss due to drought. Unfortunately, the extent, frequency and severity of drought occurrences have been increasing globally due to climate change. In order to feed the ever-increasing global population, increasing drought tolerance or resistance in potato through breeding is essential. For this to be successful, plant traits associated with drought tolerance or resistance need to be accurately identified and selected. Previously, water use efficiency (WUE) was identified as the most promising trait for selecting drought tolerant potato clones. This review also looks at the effect of moisture stress on potato growth, development and production; plant emergence and tuberization were identified as two critical periods when water stress mostly affect final tuber yield. Possibilities of conventional breeding as well as marker assisted breeding to increase drought resistance or tolerance were explored.

Keywords
Breeding; Drought tolerance; Potatoes; Water use efficiency

Introduction

Moisture stress is one of the major abiotic factors that affect potato production worldwide (Yuan et al., 2003). Vulnerability of potato to drought has been attributed mainly to the crop’s shallow root system and low capacity of recuperation after a period of water stress (Iwama and Yamaguchi, 2006). Potatoes have sparse and shallow root system (Kashyap and Panda, 2003; Onder et al., 2005) with a depth ranging from 0.5 to 1.0 m (Vos and Groenwold, 1989). About 85 % of the total root length is concentrated in the upper 0.3 m of soil. Due to this, potato extracts less of the available water from the soil compared to other crops (Weisz et al., 1994). Even short periods of water shortage can reduce tuber production and quality (Miller and Martin, 1987). The relative inability of potato to withstand drought limits its productive range to areas with adequate rainfall or reliable irrigation. Consequently, the ideal conditions for potato production include high and nearly constant soil matric potential, high soil oxygen diffusion rate, adequate incoming radiation and optimal soil nutrients (Yuan et al., 2003). It is estimated that the average global potato yield could be increased by at least 50 % if water supply to the crop could be optimised.

 

The negative impacts of moisture stress on potato production are likely to increase over the next decades due to climate change and the expansion of potato production in drought prone areas. Due to drought, it is estimated that potential potato yield will decrease by 18 to 32 % between 2040 and 2069 (Hijmans, 2003). Recently, drought affected Russia and led to losses of around 30% on industrial potato farms in the Central and the Volga Valley in 2010 (GAIN, 2010). Furthermore, crop models predict reduction in potato yields by about 30 % in Poland as a result of water deficit (http://www.climateadaptation.eu/poland/agriculture-and-horticulture/). Insufficient water supply may occur almost anywhere in the world; in arid tropical and sub-tropical zones where crop production is only possible with irrigation, short periods of drought often arise because of inappropriate irrigation techniques or shortage of rainfall. In the temperate zones, both short and long periods of drought may occur due to irregular rainfall particularly on soils with low water holding capacity. In warm tropical areas, the negative effect of water stress is exacerbated by high temperature. Even with good irrigation practices, water stress may occur because of high transpiration rates especially during mid-day when root system cannot completely meet the water requirements of the plant leading to increased water potential and consequent reduction in the rate of photosynthesis (Minhas and Sukumaran, 1988).

 
In the face of climate change, there is need to develop potato varieties that are well adapted to drought conditions. To achieve this, there is need to accurately identify plant traits that confer tolerance or resistance to water deficits and to understand the interaction between these traits and tuber yields. This review discusses the potato plant traits that have already been identified and which can be employed in an effective breeding programme.
 
Effects of moisture stress on plant growth

In most plants, moderate drought leads to a reduction in stem height, number of green leaves and leaf length (Deblonde and Ledent, 2001). Drought limits crop productivity by affecting photosynthetic processes at the canopy, leaf or chloroplast level, either directly, or by feedback inhibition if transport of photosynthates to sink organs is limited (Jones and Corlett, 1992). Plants grown under drought conditions tend to have lower stomatal conductance, thus helping to conserve water and maintain an adequate leaf water status; however, this reduces leaf internal carbon dioxide (CO2) concentration and photosynthesis (Chaves et al., 2002). The effects of drought on plants depend on intensity, duration and rate of progression of imposed drought (Pinheiro and Chaves, 2011).  Moisture stress first causes stomatal closure thus reducing CO2 uptake for photosynthesis; this leads to reduced plant growth and yield (Serraj et al., 2004; Mafakheri et al., 2010; ScienceDaily, 2008). Sugar concentration within the leaf tissue increases to increase the osmotic potential of the plant; this leads to feedback inhibition of photosynthesis (Basu et al., 1999). Drought also leads to increased accumulation of reactive oxygen species (ROS) in plants such as superoxide radical (O2-) and hydrogen peroxide (H2O2). Overproduction of ROS can disrupt normal plant metabolism through impaired enzyme activity due to oxidative damage, protein degradation, DNA and RNA damage and membrane lipid peroxidation, which can ultimately culminate in cell death (Finkel and Holbrook, 2000).

 

Effects of drought on potato crop

Drought may affect potato crop in a number of ways: 1) by reducing the amount of productive foliage thereby decreasing plant growth (Deblonde and Ledent, 2001), 2) by decreasing the rate of photosynthesis per unit of leaf area and 3) by shortening the growth cycle/vegetative period (Kumar et al., 2007). This result in reduced number (Eiasu et al., 2007) and size (Schafleitner et al., 2007) of tubers produced Potato yields and quality are influenced by the timing, duration and intensity of rainfall or irrigation (Ekanayakeand Midmore, 1989; Jeffery, 1995); it is possible to increase yields through well-scheduled irrigation programmes throughout the growing season (King and Stark, 1997). Sensitivity of potato to water stress varies with the developmental stage of the crop; plant emergence and tuberization are two critical periods when water stress most affects final tuber yield (Martínez and Moreno, 1992). Drought after planting may delay or even inhibit plant emergence while insufficient water supply between plant emergence and beginning of tuber bulking may lead to slow growth rate of the foliage, small leaves and small plants. Minhas and Bansal (1991) showed that tuber initiation is the most sensitive stage to water stress; drought during this period can reduce the number of tubers produced per plant (King and Stark, 1997).Moisture stress at stolon and tuber initiation not only restrains foliage and plant development but also limits the number of stolons formed leading to reduced tuber numbers and therefore a reduction in final yields (Deblonde and Ledent, 2001). Tuber size and quality are closely related to moisture supply during tuber bulking period and total yield of potatoes is most sensitive to water stress during this period. Water stress during tuber bulking stage leads to a reduction in the leaf expansion rate, inhibits the development of new leaves and encourages plant senescence resulting in decreased leaf area index (LAI) (Kumar and Minhas, 1999; Susnoschi and Shimsi, 1985).  Gandergander and Tanner (1976) showed that mild water stress of –3 to –5 bars greatly reduce leaf expansion in potatoes. For best tuber yields, a 120-150 day potato crop requires 20-27.5 inches (508-698.5 mm) of water. Moisture stress can reduce yields, produce misshapen tubers, negatively affect processing quality and increase common scab incidence (Mane et al., 2008). Tuber characteristics such as shape, dry matter and reducing sugars contents can be influenced by water stress during the vegetative period. Shape defects such as dumb-bell shaped, knobby or pointed end tubers can be caused by short periods of moisture stress during the tuber bulking stage (MacKerron and Jefferies, 1988). Misshapen tubers can also occur due to secondary growth which mainly occurs in dry soils when temperatures rise (Lugt, 1964). These hot and dry conditions may also result in poor cooking quality (glassiness) of the tubers, jelly end or translucent tuber ends. They also result in high content of reducing sugars in tubers which causes difficulties during processing. Tubers from water stressed plants often have higher contents of total sugars and dry matter than well-watered plants (Levy, 1983). Steckel and Gray (1979) found that the dry matter content of potato tubers grown under low soil moisture was much higher than in tubers from well-watered plants. However, Levy (1983) found that tubers from stressed plants of cultivar up-to-date and cultivar Troubadour had a lower dry matter content while cultivar Alpha had higher dry matter content than well-watered controls.

 

Plant response to water stress

Reduction in plant size and leaf area, early plant senescence, and prolonged stomata closure are plant responses to drought (Vos and Groenwold, 1989). Drought resistance in plants can be classified according to the mechanisms exhibited; these include drought escape, avoidance, tolerance and recovery. According to Levitt (1980), drought avoidance includes closure of stomata or possession of a large root system while drought tolerance includes capacity for osmotic adjustment or rapid resumption of photosynthetic activity. Drought tolerance is primarily attributed to maintenance cell turgor and it includes osmotic adjustment and cellular or tissue elasticity (Obidiegwu et al., 2015). Responses to drought stress can also be partitioned into (i) avoidance of tissue water deficits/dehydration, (ii) tolerance of tissue water deficits, and (iii) efficiency mechanisms (Turner, 1986; Jones, 2014). Avoidance of tissue water deficits can be achieved by means of “drought escape,” where plants grow only during periods of ample moisture and often involve rapid phenological development. The drought escape process is significant in arid regions where adapted annuals might combine short life spans with high rates of growth and gas exchange while utilizing the maximum moisture content in soil (Maroco et al., 2000). Avoiding tissue dehydration can also be achieved by enhanced water uptake as a result of increased root depth or altered rooting patterns (Jackson et al., 2000) or by reduced water loss due to stomatal closure or adjustments of the leaf energy balance through reduction in light absorption or modifications to heat and mass transfer in the leaf boundary layer (Larcher, 2000; Mitra, 2001; Jones, 2014). Drought tolerance has been associated with control of plant growth and carbon transfer under water stressed conditions (Tourneux et al., 2003), enhanced water use efficiency (Alva et al., 2012), and osmotic adjustment (Heuer and Nadler, 1998). Tolerance of tissue water deficits most commonly involves maintenance of turgor either through osmotic adjustment (OA) (Morgan, 1984) or as a result of rigid cell walls or decreased cell size (Wilson et al., 1980) even when the tissue water potential declines. In an agricultural context, farmers and breeders tend to define drought tolerant cultivars as those that maintain yield under drought conditions. Potential efficiency mechanisms for improvement of crop drought tolerance include improvements in the water use efficiency (WUE) and improvements in the efficiency with which assimilate is converted to harvestable yield (HI).

Drought tolerant cultivars employ multiple strategies to survive under water-limited conditions and to produce higher yields than sensitive cultivars. One such strategy is to obtain as much water as possible from the soil by forming a well-developed root system (Lahlou and Ledent, 2005). The second strategy is to retain more water in the plant. Relative water content (RWC) is one of the most reliable indicators for defining water retention in plants (Rampino et al., 2006; Sanchez-Rodriguez et al., 2010). Studies have shown that RWC decreases in response to drought stress (Bürling et al., 2013; Shaw et al., 2002). The third strategy of drought tolerant cultivars is to increase the capacity to defend against oxidative damage caused by drought stress. Oxidative damage is characterized by overproduction of ROS such as O2- and H2O2 resulting in lipid peroxidation and even cell death (Imlay, 2003; Ashraf, 2009; Ashraf, 2010). To counteract ROS, plants produce various types of antioxidants; activation of antioxidants is associated with the degree of drought tolerance of the plant species (Sunkar et al., 2006). Partial root zone drying has been shown to enhance antioxidant activity in potato tubers (Jovanovic et al., 2010).

Aspects of drought tolerance that are important and should be considered in potato breeding programme are: 1) the effect of short periods of moisture stress on productivity and tuber quality, 2) survival and recovery of the plants after water stress and 3) water use efficiency. Regarding survival and recovery after water stress, two factors are important: 1) limited loss of soil cover and 2) recovery of expansion growth and of the photosynthetic rate. There are indications that potato is able to regain its photosynthetic rate after drought stress. Chapman and Loomis (1953) found that after “permanently wilted” potato plants were watered, the rate of photosynthesis recovered fully after three days. Bansal and Nagarajan (1987) studied recovery of leaf growth after a period of water stress in eight potato cultivars. They reported that some potato cultivars showed minimum growth reduction under stress and had rapid recovery on re-watering with final increase in the leaf length exceeding that of the unstressed controls. In the second group, moisture stress caused moderate reduction in growth and on recovery; the increase in leaf length was comparable to that of controls. The third group of cultivars was characterised by huge reduction in growth and on re-watering, the final leaf length was less than in the controls.

Indicators of drought stress or drought tolerance

The plants water status can be described by leaf water potential (LWP) indicating the energy level of water in the plant and the relative water content (RWC). Since these parameters are related to leaf growth and photosynthetic rates, they can provide information on drought tolerance (Minhas et al., 2003). When plants are exposed to water stress, reduced water uptake leads to reduction in RWC and LWP, increased stomatal resistance due to closure of stomata and consequently reduction in the rate of photosynthesis. Basu et al. (1998) found that under controlled conditions the photosynthetic rate of potatoes was reduced by 17 % at an LWP of–4 bars and by 50 % at–11 bars. Both passive and active osmotic adjustment of the cell sap allows the plant to maintain photosynthetic rate at a lower LWP. When RWC of a leaf decreases, the osmotic value of the cell sap will increase passively. However, it is not clear whether active osmotic adjustment takes place in potato plant or not. Potatoes exhibit isohydric characteristic with soil water potential (Ψsoil) and stomatal conductance (gs) decreasing under water stress while maintaining leaf water potential (Ψ1) similar to values obtained from non-stressed conditions (Liu et al., 2005) suggesting delay in onset of stress. Predawn Ψ1 and gs can therefore be used as parameters for water stress as they exhibit coherent relationship with growth and yield.
 

Further, direct or indirect indications of drought tolerance can be obtained by measuring photosynthetic rate, water retention of excised leaves, depth and extension of the root system, anatomical structure of the leaf, yield under dry growing conditions in the field, the extent of wilting and recovery after severe drought stress, water use efficiency and transpiration rate (Minhas et al., 2003).

Breeding for drought tolerance in potatoes

It is well known that varieties of many field crops including potato respond differently to water stress with respect to various growth parameters and yield. Successful breeding requires exact information on effective drought tolerance traits, their heritability, genotype x environment interaction and, suitable selection tools for the traits of interest (Anithakumari et al., 2012). The varietal differences in response to stress can be due to variation in one or more of the following factors: leaf longevity, harvest index, root factors, anatomical structure of the leaf, stomatal behaviour, leaf water potential, osmotic potential, relative water content, photosynthetic rate, water retention by the leaves, wilting and recovery after a long water stress period, water use efficiency, yield determinations on a soil profile with varying depths, water use efficiency and transpiration rate (Minhas et al., 2003). All these parameters can serve as a basis for screening potato clones for drought tolerance but most of these parameters are not easy to apply or are not sufficiently reliable so that they could be used to screen a large number of potato clones (Minhas et al., 2003). The most promising method for screening potato clones for drought tolerance seems to be estimating water use efficiency (WUE) and transpiration rate based on isotope ratio of carbon and oxygen by isotope ratio mass spectrometer. This instrument can measure time averaged rate of water use efficiency based on ratio of carbon isotopes and transpiration based on ratio of oxygen isotopes (Minhas et al., 2003). Significant progress in assessing genetic variability in WUE was achieved after the establishment of the physiological links between Carbon isotope (δ13C) and WUE (Farquhar and Richards. 1984). The δ13C measures the ratio of stable carbon isotopes (13C/12C) in the plant dry matter compared to the ratio in the atmosphere (Condon et al., 1990). Carbon isotope (δ13C) discrimination shows a significant positive correlation with plant height and the dry biomass of the plant foliage (Anithakumari et al., 2012). In a water deficit condition, δ13C represents a reliable estimator of stomatal conductance (Condon et al., 2004) and WUE in crops (Turner, 1997) and δ13C has been used as a surrogate for WUE. Although understanding the inheritance of δ13C could be useful for the development of potato cultivars with high WUE (Anithakumari et al., 2012), an understanding of the relationship between transpiration efficiency, WUE and yield under different levels of water stress is important (Obidiegwu et al., 2015). Even if one might expect characters such as enhanced water use efficiency or improved tolerance of given water deficits to be generally adaptive; this is not generally the case as there will usually be some corresponding cost (e.g., higher WUE usually comes at the expense of lower photosynthetic rates while tolerance of water deficits often has a metabolic cost). Indeed the important character is not WUE but the “effective use of available water” (i.e., tailored to the specific growing environment) (Blum, 2009).

A number of approaches have been used to alleviate the problem of drought; nevertheless, plant breeding seems to be the most effective and economical one. There exists genotypic variability among potatoes with respect to drought tolerance with some varieties performing better under drought conditions. These drought-tolerant potato cultivars can produce reasonable yields where grain crops fail especially when drought coincides with flowering and seed set in grains (Iwama and Yamaguchi, 2006). The genetic variability that exists within S. tuberosum and its relatives can be exploited to improve drought tolerance (Levy, 1983; Jefferies and MacKerron, 1987). Although modern S. tuberosum are highly susceptible to drought stress, several landraces as well as wild species of potato are adapted to harsh and water-scarce conditions (Vasquez-Robinet et al., 2008; Hijmans and Spooner, 2001). However, there is limited use of native potatoes in conventional breeding programs due to linkage drag and the fact that some of the traits contributing to drought tolerance in native and wild potatoes are associated with low yield (Cabello et al., 2012). Desirable drought phenotypic traits must be genetically associated with high yield under stress, be highly heritable, genetically variable, easy to measure, stable within the measurement period, and must not be associated with a yield penalty under unstressed conditions (Okogbenin et al., 2013). Because potato yields depend on the timing of water stress within the growing period (Spitters and Schapendonk, 1990) and upon climatic and soil conditions (Tourneux et al., 2003), then it is necessary to consider these factors before making recommendations of optimal phenotypes for any specific environments.

Drought tolerance is a genetically complex polygenic trait with multiple pathways implicated. Therefore, effective crop improvement for drought tolerance will require pyramiding of many disparate characters, with different combinations being appropriate for different growing environments (Obidiegwu et al., 2015). However, it is difficult to pyramid drought tolerance related genes in highly heterozygous tetraploid potato cultivars while considering other important economic traits; the situation is aggravated by linkage drag and distortion in segregation between inter-specific hybrids. Hopefully these issues may be circumvented by use of biotechnological approaches.

Most of the quantitative trait loci (QTL) mapping studies in potato have been performed on diploid populations because the efficiency of association mapping is much higher in diploid than in polyploid species. Target traits included leaf senescence (Malosetti et al., 2006), tuberization (Fernandez-Del-Carmen, 2007) and tuber shape, eye depth and flesh colour (Sliwka et al., 2008). Recently, QTL have also been identified in tetraploid populations for traits such as plant height, maturity, crop emergence, tuber size, and tuber quality traits (Bradshaw et al., 2008; Hoop et al., 2010). However, insight into genetics and genes underlying QTL that are related to drought tolerance is still limited in potato (Anithakumari et al.,  2012).

Conclusions

This review discusses traits that have already been identified and are useful in selecting for drought tolerance in potatoes. Drought tolerance or resistance is a complex polygenic trait with multiple pathways implicated.  Breeding for drought tolerance using conventional methods has challenges especially in tetraploid potatoes due to lower efficiency of association mapping in polyploids than in diploids.

References

Alva A.K., Moore A.D., and Collins H.P., 2012, Impact of deficit irrigation on tuber yield and quality of potato cultivars, Journal of Crop Improvement, 26: 211-227

http://dx.doi.org/10.1080/15427528.2011.626891

 

Anithakumari A.M., Karaba N.N., Richard G.F., Visser C., and Gerard D.L., 2012, Genetic dissection of drought tolerance and recovery potential by quantitative trait locus mapping of a diploid potato population, Molecular Breeding, 30: 1413-1429

http://dx.doi.org/10.1007/s11032-012-9728-5

 

Ashraf M., 2009, Biotechnological approach of improving plant salt tolerance using antioxidants as markers, Biotechnological Advances, 27: 84-93

http://dx.doi.org/10.1016/j.biotechadv.2008.09.003

 

Ashraf M., 2010, Inducing drought tolerance in plants: some recent advances, Biotechnological Advances, 28: 169-183

http://dx.doi.org/10.1016/j.biotechadv.2009.11.005

 

Eiasu B.K., Soundy P., Hammes P.S., 2007, Response of potato (Solanum tuberosum) tuber yield components to gel-polymer soil amendments and irrigation regimes, N. Z. J. Crop Hort., 35: 25-31

http://dx.doi.org/10.1080/01140670709510164

 

Bansal K.C. and Nagarajan S., 1987, Reduction of leaf growth by water stress and its recovery in relation to transpiration and stomatal conductance in some potato genotypes, Potato Research, 30: 497-506

http://dx.doi.org/10.1007/BF02361929

 

Basu P.S., Sharma A., and Sukumaran N.P., 1998, Changes in net photosynthesis rate and chlorophyll fluorescence in potato leaves induced by water stress, Photosynthetica, 35: 13-19

http://dx.doi.org/10.1023/A:1006801311105

 

Basu P.S., Sharma A., Garg I.D., and Sukumaran N.P., 1999, Tuber sink modifies photosynthetic response in potato under water stress, Environmental and Experimental Botany, 42: 25-39

http://dx.doi.org/10.1016/S0098-8472(99)00017-9

 

Bhaskar P.B., Wu L., Busse J.S., Whitty B.R., Hamernik A.J., Jansky S.H., Buell C.R., Bethke P.C., and Jiang J., 2010, Suppression of the vacuolar invertase gene prevents cold-induced sweetening in potato, Plant Physiology, 154: 939-948

http://dx.doi.org/10.1104/pp.110.162545

 

Blum A., 2009, Effective use of water (EUW) and not water-use efficiency (WUE) is the target of crop yield improvement under drought stress, Field Crops Research, 112: 119-123

http://dx.doi.org/10.1016/j.fcr.2009.03.009

 

Bradshaw J.E., Hackett C.A., Pande B., Waugh R., and Bryan G.J., 2008, QTL mapping of yield, agronomic and quality traits in tetraploid potato (Solanum tuberosum subsp. tuberosum), Theoretical and Applied Genetics, 116: 193-211

http://dx.doi.org/10.1007/s00122-007-0659-1

 

Bürling K., Cerovic Z.G., Cornic G., Ducruet J.M., Noga G., and Hunsche M., 2013

Cabello R., Chujoy E., Mendiburu F., Bonierbale M. and Monneveux P., 2012, Large scale evaluation of potatoes improved varieties, pre-breeding material and landraces for drought tolerance, American Journal of Potato Research, 89: 400-410

http://dx.doi.org/10.1007/s12230-012-9260-5

 

Chapman H.W., and Loomis W. E., 1953, Photosynthesis in potato under field conditions, Plant Physiology, 28: 703-716

http://dx.doi.org/10.1104/pp.28.4.703

 

Chaves M.M., Pereira J.S., Maroco J., Rodrigues M.L., Ricardo C.P.P., Osorio M., Carvalho L., Faria T., and Pinheiro C., 2002, How plants cope with water stress in the field: Photosynthesis and growth, Annals of Botany, 89: 907-916

http://dx.doi.org/10.1093/aob/mcf105

 

Condon A.G., Richards R.A., Rebetzke G.J., and Farquhar G.D., 2004, Breeding for high water-use efficiency, Journal of Experimental Botany, 55: 2447-2460

http://dx.doi.org/10.1093/jxb/erh277

 

Condon A., Farquhar G., and Richards R., 1990, Genotypic variation in carbon isotope discrimination and transpiration efficiency in wheat, Leaf Gas Exchange and Whole Plant Studies, Functional Plant Biology, 17: 9-22

http://dx.doi.org/10.1071/pp9900009

 

Cooper S.G., Douches D.S., and Grafius E.J., 2009. Combining engineered resistance, avid in, and natural resistance derived from Solanum chacoense Bitter to control Colorado potato beetle (Coleoptera: Chrysomelidae), Journal of Economic Entomology, 102: 1270-1280

http://dx.doi.org/10.1603/029.102.0354

 

Deblonde P.M.K., and Ledent J.F., 2001, Effects of moderate drought conditions on green leaf number, stem height, leaf length and tuber yield of potato cultivars, European Journalof Agronomy, 14: 31-41

http://dx.doi.org/10.1016/S1161-0301(00)00081-2

 

Ekanayake I.J., and Midmore D.J., 1989, Root-pulling resistance of potatoes in a drought environment, American Potato Journal, 66: 519

 

Farquhar G.D., and Richards R.A., 1984, Isotopic composition of plant carbon correlates with water-use efficiency of wheat genotypes, Australian Journal of Plant Physiology, 11: 539-552

http://dx.doi.org/10.1071/PP9840539

 

Fernandez-Del-Carmen A., Celis-Gamboa C., Visser R.G., and Bachem C.W., 2007, Targeted transcript mapping for agronomic traits in potato, Journal of Experimental Botany, 58: 2761-2774

http://dx.doi.org/10.1093/jxb/erm140

 

Finkel T., and Holbrook N.J., 2000, Oxidants, oxidative stress and the biology of ageing, Nature, 408: 239-247

http://dx.doi.org/10.1038/35041687

 

Fluorescence-based sensing of drought-induced stress in the vegetative phase of four contrasting wheat genotypes, Environmental and Experimental, Botany, 89: 51-59

 

GAIN, 2010, Global Agricultural Information Network of the USDA Foreign Agricultural Service, GAIN Report No RS1060

 

Gandergander P.W., and Tanner C.B., 1976, Leaf growth, tuber growth and water potential in potatoes, Crop Science, 16: 534-538

http://dx.doi.org/10.2135/cropsci1976.0011183X001600040025x

 

Heuer B., and Nadler A., 1998, Physiological response of potato plants to soil salinity and water deficit, Plant Science, 137: 43-51

http://dx.doi.org/10.1016/S0168-9452(98)00133-2

 

Hijmans R.J., 2003, The effect of climate change on global potato production, Am. J. Potato Research, 80: 271-280

http://dx.doi.org/10.1007/BF02855363

 

Hijmans R.J., and Spooner D.M., 2001, Geographic distribution of wild potato species, American Journal of Botany, 88: 2101-2112

http://dx.doi.org/10.2307/3558435

 

Hoop B.B., Paulo M.J., Kowitvanich K., Sengers M., Visser R.G.F., Herman J.E., and Eeuwijk F., 2010, Population structure and linkage disequilibrium unravelled in tetraploid potato, Theoretical and Applied Genetics, 121: 1151-1170

http://dx.doi.org/10.1007/s00122-010-1379-5

 

Imlay J.A., 2003, Pathways of oxidative damage, Annual Review of Microbiology, 57: 395-418

http://dx.doi.org/10.1146/annurev.micro.57.030502.090938

 

Iwama K., and Yamaguchi J., 2006, Abiotic Stress, In: Handbook of potato production, improvement and post-harvest management, Gopal J. and Khurana S.M., Paul (Eds.), Food Product Press, New York

 

Jackson R.B., Sperry J.S., and Dawson T.E., 2000, Root water uptake and transport: using physiological processes in global predictions, Trends Plant Science, 5: 482-488

http://dx.doi.org/10.1016/S1360-1385(00)01766-0

 

Jefferies R.A., and MacKerron D.K.L., 1987, Aspects of the physiological basis of cultivar differences in yield under droughted and irrigated conditions, Potato Research, 30: 201-217

http://dx.doi.org/10.1007/BF02357663

 

Jefferies R.A., and Mackerron D.K.L., 1994, Genotypic differences in water use efficiency in potato, In: Heath M.C., Hess T.M., Hocking T.J., MacKerron D.K.L., Stephens W. (eds), Aspects of applied biology 38, efficiency of water use in crop systems, Association of Applied Biologists, Wellesbourne, pp.63-70

 

Jeffery R.A., 1995, Physiology of crop response to drought, pp.61-74, In: Haverkortand A.J., MacKerron D.K.L., (eds), Potato Ecology and Modeling of Crops Under Conditions Limiting Growth, Wageningen Academic Publishers, The Netherlands

 

Jones H.G., 2014, Plants and microclimate: A quantitative approach to environmental plant physiology, Cambridge: Cambridge University Press

 

Jones H.G., and Corlett J.E., 1992, Current topics in drought physiology, Journal of Agricultural Science, 119: 291-296

http://dx.doi.org/10.1017/S0021859600012144

Jovanovic Z., Stikic R., Radovic B.V., Paukovic M., Brocic Z., Matovic G., Rovcanin S., and Mojevic M., 2010, Partial root zone drying increase WUE, N and antioxidant content in field potatoes, European Journal of Agronomy, 33: 124-131

http://dx.doi.org/10.1016/j.eja.2010.04.003

 

Kashyap P.S., and Panda P.K., 2003, Effect of irrigation scheduling on potato crop parameters under water stressed conditions, Agric Water Manage, 59: 49-66

http://dx.doi.org/10.1016/S0378-3774(02)00110-5

 

King B.A. and Stark J.C., 1997, Potato IrrigationManagement, bulletin 789, University of Idaho Cooperative Extension System, Moscow, Idaho, USA

 

Kuhl J.C., Zarka K.A., Coombs J., Kirk W.W., and Douches D.S., 2007, Late blight resistance of RB transgenic potato lines, Journal of the American Society for Horticultural Science, 132: 783-789

 

Kumar D. and Minhas J.S., 1999, Effect of water stress on photosynthesis, productivity and water status in potato, Journal of Indian Potato Association, 26: 7-10

 

Kumar S., Asrey R., and Mandal G., 2007, Effect of differential irrigation regimes on potato (Solanum tuberosum) yield and post-harvest attributes, Indian Journal of Agricultural Science, 77: 366-368

 

Lahlou O., and Ledent J.F., 2005, Root mass and depth, stolons and roots formed on stolons in four cultivars of potato under water stress, European Journal of Agronomy, 22: 159-173

http://dx.doi.org/10.1016/j.eja.2004.02.004

 

Larcher W., 2000, Temperature stress and survival ability of Mediterranean sclerophyllous plants, Plant Biosystems, 134: 279-295

http://dx.doi.org/10.1080/11263500012331350455

 

Levitt J., 1980, Responses of Plants to Environmental Stresses, Vol. 2. Water, Radiation, Salt, and Other Stresses, New York, NY: Academic; Press

 

Levy D., 1983, Varietal differences in the response of potatoes to repeated short periods of water stress in hot climates. 2. Tuber yield and dry matter accumulation and other tuber properties, Potato Research, 26: 315-321

http://dx.doi.org/10.1007/BF02356152

 

Liu F.L., Jensen C.R., Shahanzari A., Andersen M.N., and Jacobsen S.E., 2005, ABA regulated stomatal control and photosynthetic water use efficiency of potato (Solanum tuberosum L.) during progressive soil drying, Plant Science, 168: 831-836

http://dx.doi.org/10.1016/j.plantsci.2004.10.016

 

Lugt C., Bodlaender K.B.A., and Goodijk G., 1964, Observations on the induction of second growth in potato tubers, European Potato Journal, 7: 219-227

http://dx.doi.org/10.1007/BF02368253

 

MacKerron D., and Jefferies R., 1986, The influence of early soil moisture stress on tuber numbers in potato, Potato Research, 29: 299-312

 

MacKerron D.K.L., and Jefferies R.A., 1988, The distribution of tuber sizes in droughted and irrigated crops of potato, I. Observations on the effect of water stress on graded yields from different cultivars. Potato Research, 31: 269-278

http://dx.doi.org/10.1007/BF02365535

 

Mafakheri A., Siosemardeh A., Bahramnejad B., Struik P., and Sohrabi Y., 2010, Effect of drought stress on yield, proline and chlorophyll contents in three chickpea cultivars, Australian Journal of Crop Science, 4: 580-585

 

Malosetti M., Visser R.G.F., Celis-Gamboa C., and van Eeuwijk F.A., 2006, QTL methodology for response curves on the basis of non-linear mixed models, with an illustration to senescence in potato, Theoretical and Applied Genetics, 113: 288-300

http://dx.doi.org/10.1007/s00122-006-0294-2

 

Mane S.P., Robinet C.V., Ulanov A., Schafleitner R., Tincopa L., Gaudin G., Nomberto A., Alvarado C., Solis C., Bolivar L.A., Blas R., Ortega O., Solis J., Panta A., Rivera C., Samolski I., Carbajulca D.H., Bonierbale M., Pati A., Heath L.S., Bohnert H.J. and Grene R., 2008, Molecular and physiological adaptation to prolonged drought stress in the leaves of two Andean potato genotypes, Functional Plant Biology, 35: 669-688

http://dx.doi.org/10.1071/FP07293

 

Maroco J.P., Edwards M.S.B., and Edwards G.E., 2000, Utilization of O-2 in the metabolic optimization of C-4 photosynthesis, Plant Cell Environment, 23: 115-121

http://dx.doi.org/10.1046/j.1365-3040.2000.00531.x

 

Martínez C.A., and Moreno U., 1992, Expresiones fisiológicas de resistencia a sequíaen dos variedades de papa sometidas a estréshídrico, Rev. Brasil. deFisiol. Veget., 4: 33-38

 

Miller D., and Martin M., 1987, Effect of declining or interrupted irrigation on yield and quality of three potato cultivars grown on sandy soil, American Journal of Potato Research, 64: 109-117

http://dx.doi.org/10.1007/BF02854207

 

Minhas J.S. and Bansal K.C., 1991, Tuber yield in relation to water stress at different stages of growth in potato, Journal of Indian Potato Association, 18: 1-8

 

Minhas J.S. and Sukumaran N.P., 1988, Diurnal changes in net photosynthetic rate in potato in two environments, Potato research, 31: 375-378

http://dx.doi.org/10.1007/BF02357871

 

Minhas J.S., Paul Khurana S.M., Sheshshayee M.S., and Udayakumar M., 2003, Potato varieties show genetic variability in water use efficiency based on carbon isotope discrimination, Journal of Indian Potato Association, 30: 193-194

 

Mitra J., 2001, Genetics and genetic improvement of drought resistance in crop plants, Current Science, 80: 758-763

 

Morgan J.M., 1984, Osmoregulation and water stress in higher plants, Annual Review of Plant Physiology and Plant Molecular Biology, 35: 299-319

http://dx.doi.org/10.1146/annurev.pp.35.060184.001503

 

Nicole L.N., Zarka K.A., Coombs J.J., and Douches D.S., 2014, Field assessment of AtCBF1 transgenic potato lines (Solanum tuberosum) for drought tolerance, American Journal of Potato Research

http://dx.doi.org/10.1007/s12230-014-9424-6

 

Obidiegwu J.E., Glenn J.B., Hamlyn G.J., and Prashar A., 2015, Coping with drought: stress and adaptive responses in potato and perspectives for improvement. Frontiers in Plant Science, 6: 542

http://dx.doi.org/10.3389/fpls.2015.00542

 

Okogbenin E., Setter T.L., Ferguson M., Mutegi R., Ceballos H., Olasanmi B., and Fregene M., 2013, Phenotypic approaches to drought in cassava: review. Frontiers in Physiology, 4: 93

 

Onder S., Caliskan M.E., Onder D., and Caliskan S., 2005, Different irrigation methods and water stress effects on potato yield and yield components, Agriculture and Water Management, 73: 73-86

http://dx.doi.org/10.1016/j.agwat.2004.09.023

 

Pinheiro C., and Chaves M.M., 2011, Photosynthesis and drought: can we make metabolic connections from available data? Journal of Experimental Botany, 62: 869-882

http://dx.doi.org/10.1093/jxb/erq340

 

Rampino P., Pataleo S., Gerardi C., Mita G., and Perrotta C., 2006, Drought stress response in wheat: physiological and molecular analysis of resistant and sensitive cultivars. Plant Cell and Environment, 29: 2143-2152

http://dx.doi.org/10.1111/j.1365-3040.2006.01588.x

 

Sanchez-Rodriguez E., Rubio-Wilhelmi M., Cervilla L.M., Blasco B., Rios J.J., Rosales M.A., Romero L., and Ruiz J.M., 2010, Genotypic differences in some physiological parameters symptomatic for oxidative stress under moderate drought in tomato plants, Plant Science, 178: 30-40

http://dx.doi.org/10.1016/j.plantsci.2009.10.001

 

Schafleitner R., Gutierrez R., Espino R., Gaudin A., Perez J., Martinez M., Dominguez A., Tincopa L., Alvarado C., Numberto G., and Bnierbale M., 2007, Field screening for variation of drought tolerance in Solanum tuberosum L. by agronomical, physiological and genetic analysis, Potato Research, 50: 71-85

http://dx.doi.org/10.1007/s11540-007-9030-9

 

Schafleitner R., Gaudin A., Gutierrez Rosales R.O., Alvarado Aliaga C.A., and Bonierbale M., 2007, Proline accumulation and real time PCR expression analysis of genes encoding enzymes of proline metabolism in relation to drought tolerance in Andean potato, ActaPhysiologia(Plant), 29: 19-26

http://dx.doi.org/10.1007/s11738-006-0003-4

 

ScienceDaily, 30 June 2008. "Drought Tolerance In Potatoes."  American Society of Plant Biologists

 

Serraj R., Krishnamurthy L., Kashiwagi J., Kumar J., Chandra S., and Crouch J.H., 2004, Variation in root traits of chickpea (Cicer arietinum L.) grown under terminal drought, Field Crops Research, 88: 115-127

http://dx.doi.org/10.1016/j.fcr.2003.12.001

 

Shaw M.R., Zavaleta E.S., Chiariello N.R., Cleland E.E., Mooney H.A., and Field C.B., 2002, Grassland responses to global environmental changes suppressed by elevated CO2, Science, 298: 1987-1990

 

Sliwka J., Wasilewicz-Flis I., Jakuczun H., and Gebhardt C., 2008, Tagging quantitative trait loci for dormancy, tuber shape, regularity of tuber shape, eye depth and flesh colour in diploid potato originated from six Solanum species, Plant Breeding, 127: 49-55

http://dx.doi.org/10.1111/j.1439-0523.2008.01420.x

 

Spitters C.J.T., and Schapendonk A., 1990, Evaluation of breeding strategies for drought tolerance in potato by means of crop growth simulation, Plant Soil, 123: 193-203

http://dx.doi.org/10.1007/BF00011268

 

Steckel J.R.A. and Gray D., 1979, Drought tolerance in potatoes, Journal of Agricultural Science, Cambridge, 92: 375-381

http://dx.doi.org/10.1017/S0021859600062900

 

Sunkar R., Kapoor A., and Zhu J.K., 2006, Post transcriptional induction of two Cu/Zn superoxide dismutase genes in Arabidopsis is mediated by down regulation of miR398 and important for oxidative stress tolerance, Plant Cell, 18: 2051-2065

http://dx.doi.org/10.1105/tpc.106.041673

 

Susnoschi M., and Shimsi D., 1985, Growth and yield studies of potato development in semi-arid tropics, Potato Research, 28: 161-176

http://dx.doi.org/10.1007/BF02357442

 

Tourneux C., Devaux A., Camacho M.R., Mamani P., and Ledent J.F., 2003, Effect of water shortage on six potato genotypes in the highlands of Bolivia (II): water relations, physiological parameters, Agronomie, 23: 181-190

http://dx.doi.org/10.1051/agro:2002079

 

Tourneux C., A.Devaux,M.R. Camacho, Mamani P., and Ledent J.F., 2003, Effects of water shortage on six potato cultivars in the highlands of Bolivia (I): morphological parameters, growth and yield. Agronomie, 23: 169-179

http://dx.doi.org/10.1051/agro:2002079

 

Turner N.C., 1986, Crop water deficits: a decade of progress. Advances in Agronomy, 39: 1-51

http://dx.doi.org/10.1016/S0065-2113(08)60464-2

 

Turner N.C., 1997, Further progress in crop water relations. Advances in Agronomy, 58: 293-338

http://dx.doi.org/10.1016/S0065-2113(08)60258-8

 

USDA, ARS, National Genetic Resources Program, 2012, Germplasm Resources Information Network-(GRIN), http: // www.ars-grin.gov/cgi-bin/npgs/html/tax_site_acc.pl?NR6%20Solanum%20hjertingii. Accessed 18 Dec 2012

 

Vasquez-Robinet C., Mane S.P., Ulanov A.V., Watkinson J.L., Stromberg V.K., De Koeyer D., Schafleitner R., Willmot, D.B., Bonierbale M., Bohnert H.J., and Grene R., 2008, Physiological and molecular adaptations to drought in Andean potato genotypes, Journal of Experimental Botany, 59: 2109-2123

http://dx.doi.org/10.1093/jxb/ern073

 

Vos J., and Groenwold J., 1989, Characteristics of photosynthesis and conductance of potatocanopies and the effect of cultivar and transient drought, Field Crops Research, 20: 237-250

http://dx.doi.org/10.1016/0378-4290(89)90068-3

 

Batima, Punsalmaa, Brander, Keith, Erda, Lin, Howden, Mark, Kirilenko, Andrei, Morton, John F., Soussana, Jean-François, Schmidhuber, Josef, Tubiello, Francesco, Easterling, William and Aggarwal, Pramod, 2007, Food, fibre and forest products, in: M.L. Parry, O.F. Canziani, J.P. Palutikof, P.J. van der Linden, C.E. Hanson (Eds.), Climate Change 2007: Impacts, Adaptation and Vulnerability, Cambridge University Press, Cambridge, UK, 2007, pp. 273-313

 

Weisz R., Kaminski J., and Smilowitz Z., 1994, Water deficit effects on potato leaf growth and transpiration: utilizing fraction extractable soil water for comparison with other crops, American Journal of Potato Research, 71: 829-840

http://dx.doi.org/10.1007/BF02849378

 

Wilson J.R., Ludlow M.M., Fisher M.J., and Schulze E.D., 1980, Adaptation to water stress of the leaf water relations of four tropical forage species, Australian Journal of Plant Physiology, 7: 207-220

http://dx.doi.org/10.1071/PP9800207

 

Yuan B.Z., Nishiyama S., and Kang Y., 2003, Effects of different irrigation regimes on the growth and yield of drip-irrigated potato, Agricultural Water Management, 63: 153-167

http://dx.doi.org/10.1016/S0378-3774(03)00174-4

 

Zarka K.A., Greyling R., Gazendam I., Olefse D., Felcher K., Bothma G., Brink J., Quemada H., and Douches D.S., 2010, Insertion and characterization of the cry1Ia1 gene in the potato cultivar ‘Spunta’ for resistance to potato tuber moth, Phthorimaea operculella (Zeller), Journal of the American Society for Horticultural Science, 135: 317-324

International Journal of Horticulture
• Volume 6
View Options
. PDF(369KB)
. HTML
Associated material
. Readers' comments
Other articles by authors
. Jane Muthoni
. J.N. Kabira
Related articles
. Breeding
. Drought tolerance
. Potatoes
. Water use efficiency
Tools
. Email to a friend
. Post a comment