Running title - Rice responses to rising temperatures

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14

Title - Rice responses to rising temperatures – challenges, perspectives and future directions

1

Jagadish S V K1 *, Murty M V R1, Quick W P1,2 1 International Rice Research Institute, DAPO Box 7777, Metro Manila, Philippines 2 Animal and Plant Sciences, University of Sheffield, Sheffield S10 2TN, UK *Corresponding author - Jagadish SVK, International Rice Research Institute, DAPO Box 7777, Metro Manila, Philippines; phone: +63 (2) 580-5600-2767; fax: +63 (2) 580-5699; 845-0606, email: [email protected]

This article has been accepted for publication and undergone full peer review but has not been through the copyediting, typesetting, pagination and proofreading process, which may lead to differences between this version and the Version of Record. Please cite this article as doi: 10.1111/pce.12430

1 This article is protected by copyright. All rights reserved.

Abstract

3

Phenotypic plasticity in overcoming heat stress−induced damage across hot tropical rice-growing

4

regions is predominantly governed by relative humidity. Expression of transpiration cooling, an

5

effective heat-avoiding mechanism, will diminish with the transition from fully-flooded paddies

6

to water-saving technologies such as direct-seeded and aerobic rice cultivation, thus further

7

aggravating stress damage. This change can potentially introduce greater sensitivity to previously

8

unaffected developmental stages such as floral meristem (panicle) initiation and spikelet

9

differentiation, and further intensify vulnerability at the known sensitive gametogenesis and

Accepted Article

1 2

10

flowering stages. More than the mean temperature rise, increased variability and a more rapid

11

increase in nighttime temperature compared with the daytime maximum present a greater

12

challenge. This review addresses (i) the importance of vapor pressure deficit under fully flooded

13

paddies and increased vulnerability of rice production to heat stress or intermittent occurrence of

14

combined heat and drought stress under emerging water-saving rice technologies, (ii) the major

15

disconnect with high night temperature response between field and controlled environments in

16

terms of spikelet sterility, (iii) highlight the most important mechanisms that affect key grain

17

quality parameters such as chalk formation under heat stress and, finally, (iv) we model and

18

estimate heat stress−induced spikelet sterility taking South Asia as a case study.

19 20

Key words – flowering, high temperature, relative humidity, rice, spikelet fertility, water-saving

21

technologies

22 23 24

2 This article is protected by copyright. All rights reserved.

Accepted Article

1 2

Introduction

3

Rice (Oryza sativa L.) traditionally has been grown as a fully flooded crop and has been the

4

major source of calories for more than half the world’s population and livelihood of many small

5

and marginal farmers of Asia and increasingly in Africa. With a significant jump in rice

6

production during the 1960s (Green Revolution), followed by stagnation during the late 1990s, it

7

is evident that increasing productivity is essential to keep pace with the demographic demand. In

8

addition, global climatic predictions, of late with greater certainty, indicate increased frequency

9

of heat spikes and warmer nights (IPCC, 2013), exerting additional challenges towards achieving

10

higher crop yields. Considering the global harvested area, climate models predict that, by 2030,

11

16% of the rice-growing area would be exposed to at least 5 days of temperatures above the

12

critical threshold during the reproductive period, with a non-linear increase to 27% by 2050

13

(Gourdji et al. 2013). Similarly, a global heat risk map for 2071–2100, with 1971−2000 as a

14

reference, involving short episodes of heat stress coinciding with the reproductive period,

15

resulted in more than 120 million ha of suitable wetland rice area to be under threat (Teixeira et

16

al. 2013). Publications by Teixeira et al. (2013), Battisti & Naylor 2009 and Challinor et al.

17

(2014) support the urgent need to enhance research efforts towards developing tolerant varieties

18

and suitable crop management practices as adaptation strategies to minimize predicted damage.

19

Further, rice yields having opposing sensitivity to daytime maximum and nighttime minimum

20

temperature across South and Southeast Asia have been documented (Welch et al. 2010), with

21

the latter accounting for a larger proportion of losses under field conditions (Peng et al. 2004;

22

Welch et al. 2010; Nagarajan et al. 2010). Hence, increasing night (minimum) temperatures can

23

potentially have a global impact compared with maximum day temperature, with the latter

3 This article is protected by copyright. All rights reserved.

posing a threat only to those locations that are already close to their optimal growing temperature

2

(Prasad et al. 2006). In addition, a disproportional increase in night temperatures would reduce

3

the diurnal amplitude, bringing with it a suite of negative impacts on crop production (Bueno et

4

al. 2012).

Accepted Article

1

5

Rice is extremely sensitive to heat stress (>35 oC), particularly during the gametogenesis

6

(Jagadish et al. 2013) and flowering (Prasad et al. 2006, Jagadish et al. 2007, 2008, 2010) stages,

7

while information on other developmental stages such as early floral meristem growth, spikelet

8

differentiation and grain filling are limited. The ability of rice to be highly productive even under

9

temperatures >40 oC (higher than the defined critical threshold) across hot tropical countries such

10

as Pakistan and Senegal is largely driven by a sufficient and timely availability of irrigation

11

water, complemented by low relative humidity (RH). This helps plants to maintain their tissue

12

temperature well below the critical threshold due to efficient transpiration-mediated cooling

13

(Weerakoon et al. 2008; Wassmann et al. 2009; Julia & Dingkuhn, 2013). A 6 oC lower panicle

14

temperature under well-irrigated arid climates of Australia (Matsui et al. 2007) and 4 oC higher

15

panicle temperature under hot and humid conditions in China (Tian et al. 2010) substantiate the

16

critical role played by RH when dealing with heat stress. The recent finding of high night

17

temperature (HNT) under field conditions affecting grain quality through a reduced non-

18

structural carbohydrate pool size (Shi et al., 2013) poses questions about source-sink dynamics

19

and proportional damage caused by increased dark respiration. This would add to the challenges

20

of increased stress during early grain development, resulting in less rice production and lower

21

grain quality, leading to a significant reduction in economic benefits (Lyman et al., 2013).

22 23

In addition to the projected increase in temperature, increased uncertainty in precipitation

patterns resulting in fewer rainy days along with the rapid expansion of urbanization and

4 This article is protected by copyright. All rights reserved.

industrialization in developing countries adds to the challenges faced by flooded rice production.

2

Water-saving rice technologies such as direct-seeded rice and aerobic rice (Bouman et al. 2005,

3

2007) are considered as potential options to sustain rice production, with alternate wetting and

4

drying feasible only in areas with sufficient and timely availability of water, a luxury that most

5

rice-growing regions do not possess. The transition from water-sufficient scenarios to water-

6

saving methods can potentially change the entire cropping system dynamics such as increased

7

weed pressure (Kumar & Ladha, 2011; Mahajan et al. 2013), nematode severity (Kreye et al.

8

2009a, b) and other soil-related problems (Nie et al. 2007, 2009). This shift could expose

9

previously unaffected developmental stages such as floral meristem initiation (panicle initiation)

Accepted Article

1

10

to damaging levels of heat stress, a key developmental phase that determines overall sink size

11

and, hence, potential yield. In addition, the already known heat-sensitive reproductive stages

12

such as gametogenesis and flowering will come under a higher intensity of exposure to heat

13

stress. It is almost certain that tropical rice-growing regions will persist in future climate

14

scenarios, with a higher evaporative demand wherein temperature-driven drought stress would

15

feature as a built-in phenomenon. This needs to be addressed if we are to avoid catastrophic

16

damage, especially when combined heat and drought stress coincide with the key developmental

17

stages described above. A recent study documents the role of extreme heat in increasing demand

18

for soil water to sustain carbon assimilation, and consequently reducing water availability during

19

the later stages of growth due to enhanced evapo-transpiration (Lobell et al., 2013). Insufficient

20

water availability then leads to a significant increase in tissue temperature that exacerbates the

21

problem (Rizhsky et al., 2002; Jagadish et al., 2011).

22

Hence, the objectives of this review are to examine:

5 This article is protected by copyright. All rights reserved.

(i)

Accepted Article

1

The importance of vapor pressure deficit under fully flooded paddies and

2

increased vulnerability of rice production to heat stress or intermittent

3

occurrence of combined heat and drought stress under emerging water-saving

4

rice technologies.

5

(ii)

6 7

(iii)

8 9

10

(iv)

The major disconnect with high night temperature response between field and controlled environments in terms of spikelet sterility. Highlight the most important mechanisms that affect key grain quality parameters such as chalk formation under heat stress. Modeling of the heat stress−induced reduction in spikelet fertility, taking South Asia as a case study.

11

Past, present and future dimensions of the impact of heat stress on rice

12

Prior to the late 1990s, very little attention was given to increasing temperature and in particular

13

the negative impact on rice yield and grain quality. There has been a series of IPCC global

14

climate change estimates over the last decade, coupled with an improvement in the precision of

15

projections on temperature increases in the future (IPCC, 2013). There is also significant

16

evidence that an increase in high temperature will result in major crop losses even under current

17

climates (Kadam et al. 2014 and references within). These data have essentially elevated the

18

importance of addressing the effects of temperature within a future climate change scenario. It is

19

only recently that significant rice yield losses, induced by heat stress have been documented in

20

the major rice-producing regions of China (Li et al. 2004), Japan (Hasegawa et al. 2009), the

21

Philippines (Peng et al. 2004) and across South and Southeast Asia (Welch et al. 2010). In

22

addition, mapping exercises have identified the regional vulnerability of rice when grown under

23

conditions closer to the critical heat stress threshold in tropical and subtropical regions of South

6 This article is protected by copyright. All rights reserved.

and Southeast Asia (Wassmann et al. 2009) and at a global scale (Teixeira et al. 2013). Future

2

temperature predictions have been shown to have a much greater impact on the minimum night

3

temperature compared with the effect on the maximum day temperature, hence reducing the

4

diurnal temperature amplitude (Vose et al. 2005). The frequency of occurrence of temperatures

5

beyond defined thresholds was geographically mapped to identify vulnerable rice-growing

6

regions with high day temperature (HDT) stress, HNT stress or a combination of both (Laborte et

7

al. 2012). These data indicate large variability in the regional occurrence of stress and the change

8

in diurnal temperature amplitude. Currently, the available literature related to the impacts of heat

9

stress, the mechanisms that lead to damage (Jagadish et al. 2010, 2013, Prasad et al. 2006) and

Accepted Article

1

10

the subsequent economic losses (Lyman et al. 2013) is well documented for flooded rice

11

conditions. A significant strain on the availability of labour and water has already started to

12

emerge as a major reason for a possible shift to use less water and labour for rice production,

13

thus requiring alternative cropping technologies such as direct-seeded rice and aerobic rice.

14

Under flooded rice systems, gametogenesis and flowering are identified as the two most

15

sensitive stages to HDT stress (Yoshida et al. 1981; Jagadish et al. 2007, 2010, 2013). Earlier

16

developmental stages, such as floral meristem initiation and spikelet differentiation, are better

17

protected because of the buffering layer of floodwater. In fact, the process of floral meristem

18

initiation and its early development that also determines total spikelet number is completed under

19

water. Higher ambient temperatures, even up to 39 oC under controlled-environment conditions,

20

did not result in a reduction in either spikelet number or spikelet fertility across five different rice

21

cultivars (Fig. 1). Data obtained from thermocouples indicated that the increase in water and soil

22

temperature was significantly lower than the prevailing ambient air temperature and that the

23

increase in water temperature may be well below the critical threshold (which is unknown)

7 This article is protected by copyright. All rights reserved.

affecting the initiating floral meristem (Fig. 2A, Jagadish et al. UnPub). Considering future

2

climate scenarios, in which a larger proportion of rice cultivation would be more mechanized and

3

under direct-seeded or aerobic conditions, the luxury of the water layer protecting initial

4

meristem growth could be challenged by higher soil and tissue temperatures. A study involving

5

fully flooded transplanted paddy and direct-seeded rice (DSR) recorded up to 25% shorter plants

6

in a DSR system. The internal competition from a much higher seeding density with DSR leads

7

to a significant delay in canopy cover, thus exposing the soil to increased temperature for longer

8

durations. This would lead to a similar condition termed “noise” that is observed when taking

9

thermal images to quantify crop canopy temperature under conditions of much hotter soil

Accepted Article

1

10

temperature (see Fig. 5 in Munns et al. 2010). Hence, the increase in temperature just above or at

11

the ground surface, close to the initiating panicle, could potentially expose this sensitive stage to

12

much greater temperatures than in flooded conditions, thereby increasing its vulnerability. Under

13

DSR, even a 10 kPa soil water tension (not considered a stress – Kumar & Ladha, 2011)

14

maintained throughout the crop growth period resulted in a 53% reduction in spikelet number.

15

This caused a serious reduction in overall sink size among main tiller panicles even under a non-

16

stress (10 kPa) condition but recorded no reduction in grain-filling percentage (Supplementary

17

Figure S1; Quinnones et al. UnPub). However, the proportion of the decline in sink size that can

18

be compensated for by a greater number of plants per unit area, under direct-seeded conditions,

19

needs to be quantified. Recently, higher temperatures, which increase vapor pressure deficit

20

(conditions prevailing with combined high temperature and low relative humidity) and hence

21

transpiration, were shown to advance severe drought stress in maize (Lobell et al. 2013).

22

Imposing drought stress during pre-anthesis resulted in a significant reduction in both seed-set

23

and abortion in primary rachis branches and more strongly in secondary rachis branches in rice

8 This article is protected by copyright. All rights reserved.

(Kato et al. 2008). Hence, the sensitive floral meristem and spikelet differentiation stage already

2

vulnerable with the transition from fully flooded conditions to non stress (10 kPa) direct-seeded

3

system (Supplementary Figure S1), when exposed to higher soil temperatures, or combined

4

drought and heat stress, could emerge as a new vulnerable developmental stage under future

5

climates. This warrants detailed investigation to ascertain the level of damage caused and the

6

appropriate measures to overcome such damage.

Accepted Article

1

7

Flowering is identified as the most sensitive development stage to heat stress, with pollen

8

viability being the major yield-determining factor (Yoshida et al. 1981; Jagadish et al. 2010).

9

The crucial role of transpiration cooling overcoming heat stress−induced spikelet sterility has

10

been highlighted earlier (Weerakoon et al. 2008; Wassmann et al. 2009; Julia & Dingkuhn,

11

2013). A transition towards water-saving technologies would automatically deprive rice plants of

12

this plasticity to respond to high temperature stress either partially or completely depending on

13

the availability of water. This would increase the canopy or panicle tissue temperature by about 2

14

o

15

flowering is a novel strategy identified and exploited in the flooded rice system to overcome the

16

damaging effects of high temperature stress by advancing the peak flowering time towards dawn

17

when the temperature is cooler (Ishimaru et al. 2010). An extension of this trait into DSR

18

cultivars could partially reduce the impact of high temperatures occurring during the late

19

morning and near noon. On the other hand, an increase in panicle and overall canopy

20

temperature would hasten crop senescence, thus decreasing both crop duration and grain filling

21

as shown in wheat (Lobell et al. 2012). Grain-filling duration in tropical rice varieties is already

22

shorter than in temperate cultivars and, with a further stress-induced decrease in grain-filling

23

duration, this could potentially lead to a significant decrease in grain quantity and quality due to

C, even under non-stress (10 kPa) conditions (Fig. 2b; Quinones et al. UnPub). Early morning

9 This article is protected by copyright. All rights reserved.

increased chalk formation (Lyman et al. 2013). In summary, the otherwise well-protected floral

2

meristem initiation under flooded rice cultivation could emerge as another highly vulnerable

3

stage under the proposed use of water-saving technologies. In addition, the transition could cause

4

greater damage to seed-set and early grain filling because of higher canopy temperatures. These

5

changes present additional challenges to overcome, if we are to sustain rice yield and quality

6

under the predicted hotter climates of the future.

7

Mechanisms driving HDT- and HNT-induced rice yield and quality losses

8

At the global (Vose et al. 2005), country (Zhou et al. 2004; Rao et al. 2014) and farm level

9

(Peng et al. 2004), a much greater increase in the minimum night temperature over the maximum

10

day temperature has been documented and supported by the recent IPCC (2013) report. Research

11

focused on addressing the impact of HDT stress has made significant progress in using heat-

12

tolerant donors such as rice landrace N22, developing mapping populations, identifying QTLs

13

and developing heat-tolerant NILs that are currently used to enhance tolerance of HDT stress

14

(Jagadish et al. 2010, Ye et al. 2012). Mechanistically, the major damage caused by HDT to rice

15

either at anthesis or gametogenesis is a reduction in pollen viability and pollen tube growth,

16

translating into lower spikelet fertility (Jagadish et al. 2010, 2013). The recent literature from

17

controlled-environment studies with HNT treatments also records stress impacts on rice and the

18

conclusions drawn indicate a similar phenomenon of decreased pollen viability, leading to

19

spikelet sterility, as the major cause (Mohammed & Tarpley, 2009 a, b, 2010, 2011, 2013; Cheng

20

et al. 2009). Other associated traits have also been reported such as increased respiratory losses,

21

increased oxidative damage and decreased membrane stability (Mohammed & Tarpley, 2009a).

22

It has to be noted that all of these results are generated from controlled environments. HNT

23

resulting in yield losses under field conditions is documented to be around 23 or 24 oC (Peng et

Accepted Article

1

10 This article is protected by copyright. All rights reserved.

al. 2004; Nagarajan et al. 2010) and 28 oC (Shi et al., 2013), whereas this is measured to be ≥32

2

o

3

has been a tendency to use thresholds for examining HNT responses by imposing temperature

4

treatments that are closer to HDT thresholds (35 oC; Jagadish et al. 2007, 2008). This is mainly

5

due to the lack of a well-defined critical HNT threshold, which has resulted in finding similar

6

physiological mechanisms that are susceptible to HDT stress, i.e. lower spikelet fertility and

7

yield losses. Two reasons why these conclusions need more thorough investigation is that almost

8

all studies have restricted their analysis to individual cultivars and they also use very high night

9

temperatures. The observed effects could also have more to do with extremely low diurnal

Accepted Article

1

10 11

C in controlled chambers (Mohammed & Tarpley, 2009 a, b, 2010; Cheng et al. 2009). There

temperature amplitude than with HNT per se. Unlike the previously published literature, a study carried out under unique field-based

12

HNT tents exposing rice plants to HNT from panicle initiation to maturity recorded no losses in

13

spikelet fertility in both stress-tolerant and susceptible cultivars (Shi et al. 2013). Shi et al (2013)

14

demonstrated that the heat treatment caused a significant decline in overall biomass and reduced

15

non-structural carbohydrate (NSC) content in plant tissues, including the panicle, and resulted in

16

decreased grain weight through reduced grain width, ultimately resulting in yield and grain

17

quality losses. This study clearly indicates the differential responses of rice plants exposed to

18

HDT and HNT, with a different chain of processes leading to HNT damage under field

19

conditions. To further substantiate this claim, we have taken the data recorded from four

20

different growth seasons, imposing higher night temperatures (reaching 29 oC), including a range

21

of inbreds and hybrids tested under the same field-based tents. The data show no loss of spikelet

22

fertility under HNT in field-grown rice (Fig. 3, Shi et al. 2013; Zhang Y et al. 2013). Therefore,

23

we urge caution and awareness of the limitations when investigating HNT stress effects and

11 This article is protected by copyright. All rights reserved.

drawing conclusions from experiments carried out under controlled-environment conditions. One

2

proposed strategy is to identify genetic resources that maintain a greater NSC pool in the stem

3

and panicle and a higher biomass in spite of HNT to serve as good genetic donors to overcome

4

HNT-induced yield losses. Predictions of significant damage that can be caused by a rapid

5

increase in global night temperature indicate an urgent need to identify and validate a critical

6

temperature threshold to facilitate thorough investigation of HNT-induced rice yield and grain

7

quality losses under field conditions.

Accepted Article

1

8

Two recent reviews indicate the critical HDT thresholds for different growth stages of

9

flooded rice crops (Shah et al. 2011; Sanchez et al. 2014). However, RH actually drives these

10

thresholds under field conditions and this has been largely overlooked in many studies (see

11

above reviews) and with regional or global estimates of heat stress (Gourdji et al. 2013; Teixeira

12

et al. 2013), thus potentially overestimating the damage caused. The interaction between

13

increasing temperature and the prevailing RH determines canopy temperature (Weerakoon et al.

14

2008) and has been recently estimated in rice across multiple rice-growing regions of the world

15

(Yoshimoto et al. 2011; Julia & Dingkuhn, 2013). These studies highlight a serious discrepancy

16

in model estimations of damage when they neglect the effects of RH (White et al. 2011).

17

Recently, two such studies modeled the impact of heat stress and spikelet sterility in rice

18

(Nguyen et al. 2013; van Oort et al. 2014), with the former considering the flowering date of

19

panicle and the time of day for anthesis, but considering only the air temperature. The latter

20

study included panicle temperature, which accounts for the dynamics between temperature and

21

RH, and also considered an arid and a humid climate. van Oort et al. 2014 showed that ignoring

22

transpiration-cooling led to an overestimation of panicle sterility by 14 to 72%. Recent

23

experimental and modeling efforts show that it is important to consider arid and humid

12 This article is protected by copyright. All rights reserved.

conditions in combination with higher temperatures for impact assessment studies and for

2

strategic breeding efforts (Zhao & Fitzgerald, 2013).

Accepted Article

1

3

Another aspect of rice production that has received little emphasis is the impact of HDT

4

or HNT (or a combination of both) on grain quality, which is equally important in terms of total

5

economic returns, particularly in the whole-grain-consuming rice markets. Recently, the non-

6

inclusion of grain quality in heat stress impact was shown to seriously underestimate milling

7

outcomes, particularly head rice yield, and subsequent economic revenue (Lyman et al. 2013).

8

Chalkiness is the major grain quality component that is increased by heat stress and hence is the

9

focus of this section, although other aspects of grain quality such as moisture content, fissuring,

10

cooking quality and palatability could be affected by higher temperatures (Fitzgerald et al. 2009;

11

Fitzgerald & Resurreccion, 2009). Three routes are identified that could lead to increased chalk

12

formation under heat stress: (i) source–sink relationship − lower source potential to meet sink

13

demand or a reduction in grain-filling duration, which in both cases limits the availability or

14

supply of sufficient assimilates for complete seed filling; (ii) starch metabolism enzymes in the

15

sink; and (iii) hormonal imbalance, in particular the ratio of ABA and ethylene. High

16

temperature during grain filling increases the rate of filling but the increase is found to be

17

insufficient to compensate for the decrease in grain-filling duration (Kobata and Uemuki, 2004),

18

resulting in increased chalkiness (Fitzgerald & Resurreccion, 2009). Similarly, under field

19

conditions, rice exposed to HNT resulted in reduced NSC content of the panicles in a

20

temperature-susceptible rice cultivar, leading to decreased grain width and increased chalkiness

21

(Shi et al. 2013). Possible explanations are earlier degradation of the nuclear epidermal

22

membrane, loose packaging of amyloplasts and poor water displacement from the gaps created

23

by small and disorganized amyloplasts across both HDT and HNT (Ishimaru et al. 2009; Zakari

13 This article is protected by copyright. All rights reserved.

et al. 2002; Song et al. 2013). With a close association between grain filling and the availability

2

of assimilates, the negative impact of increased dark respiration, particularly under HNT, leading

3

to poor grain filling and quality can’t be ruled out. Attempts have been made to examine this

4

aspect, but mostly using controlled-chamber conditions (Mohammed & Tarpley 2009a;

5

Mohammed et al. 2013; Glaubitz et al. 2014). To date, there has not been a systematic analysis

6

of HNT and the contribution of dark respiration to reducing NSC under realistic field conditions.

7

This continues to be an intriguing hypothesis. The knowledge gap identified above is a major

8

bottleneck preventing accurate quantification of the proportion of yield or quality loss that can be

9

accounted for by dark respiration in the field.

Accepted Article

1

10

During grain filling, sugar supply may not be the only limiting factor, as starch

11

metabolism enzymes such as sucrose synthase (SuSy), adenosine diphosphate glucose

12

pyrophosphorylase (AGPase), starch synthase (StSase) and starch branching enzyme (SBE) are

13

identified to be equally important in the successful conversion of sugars to starch (Yang &

14

Zhang, 2010). Additionally, a high ABA-to-ethylene ratio is equally important during grain

15

filling; external application of ABA has been shown to significantly stimulate grain filling

16

(Zhang ZX et al. 2012). High temperature during grain filling is documented to reduce the

17

expression of the sucrose transporter gene OsSUT1 and starch synthesis−related genes SuSy2,

18

AGPS2b and BEIIb and granule-bound starch synthase in grains, with OsSUT1 reduced by about

19

60% in the grains and flag-leaf tissue (Phan et al. 2013). A panicle clipping approach resulted in

20

the up-regulation of OsSUT1 in response to increased assimilate supply, resulting in increased

21

grain weight. A similar strategy employed by lowering planting density reduced yield losses in

22

rice even under high temperature (Kobata & Uemuki, 2004), indicating a strong link between

23

assimilate supply and enzymatic efficiency during grain filling. Contrasting views on the role of

14 This article is protected by copyright. All rights reserved.

α-amylase have been presented, with Hakata et al. (2012) recording a high negative correlation

2

between chalky grains and α-amylase. They indicated that high temperatures induced activation

3

of α-amylase as a crucial trigger for grain chalkiness. On the other hand, Ishimaru et al. (2010)

4

proposed that changes in water distribution led to increased chalkiness under high temperature

5

rather than starch degradation by α-amylase. A comprehensive atlas of the metabolomic and

6

transcriptomic changes that occur during grain filling when exposed to high temperatures

7

indicated either the down-regulation of sucrose import/degradation and starch biosynthesis or up-

8

regulation of starch degradation to be the major bottleneck for efficient grain filling. The existing

9

literature indicates an important role for starch metabolism (sugar transport and starch

Accepted Article

1

10

accumulation efficiency) in controlling chalkiness through sustained grain filling under high

11

temperatures.

12

Does pollen viability alone determine heat stress−induced spikelet sterility in rice?

13

In response to increasing day temperatures, rice has three adaptive mechanisms, with one being

14

heat avoidance through efficient transpiration-mediated cooling as highlighted above, and this is

15

dependent on external factors. On the other hand, heat escape and true tolerance are inherent or

16

can be genetically altered to increase adaptive resilience. Oryza species show a very wide

17

variability in flowering time of day, for which some wild rice can flower as early as 6 AM (O.

18

officinalis) or as late as 5 PM (O. australiensis), and a few can flower during the night (Sheehy et

19

al. 2007). Most cultivated rice has its peak flowering occurring between 10 AM and 12 noon

20

(Prasad et al. 2006). The phenomenon of early-morning flowering (EMF) has been exploited for

21

the first time in rice (Ishimaru et al. 2010), wherein genetic alteration allows peak flowering to

22

occur closer to dawn to help escape flowering from late-morning and early-afternoon heat.

23

Although temperatures above the critical threshold occurring an hour after flowering have been

15 This article is protected by copyright. All rights reserved.

shown not to affect fertility (Jagadish et al. 2007), post-fertilization processes and early embryo

2

growth can be potentially vulnerable under future warmer climates (Shi et al. 2014). Hence,

3

options related to late-evening flowering (LEF) are another alternative strategy (Fig. 4), wherein

4

the post-fertilization and early embryo formation phase would automatically occur under the

5

much cooler night temperatures. Both EMF and LEF can only partially overcome heat damage as

6

other sensitive developmental stages such as gametogenesis would still be equally vulnerable.

7

An answer to the above would be to target true heat tolerance: the plants’ ability to set

Accepted Article

1

8

seed in spite of their peak flowering or other sensitive developmental stage coinciding with

9

temperatures higher than the critical stress threshold, achieved by virtue of their resilient

10

reproductive physiology. This resilience has been largely associated with pollen viability and the

11

ability of pollen to germinate and fertilize. The role of the ovary is considered to be minimal

12

based on cross-fertilization studies by Yoshida et al. (1981). In order to quantify the level of

13

tolerance, both in vitro and in vivo pollen viability tests have been employed (Prasad et al. 2006;

14

Jagadish et al. 2010). Using 25 diverse rice cultivars, pollen viability was estimated using pollen

15

collected from the same set of plants, exposed to defined heat stress conditions (Fig. 5; Jagadish

16

et al. UnPub). In vitro pollen viability estimated using an iodine starch stain (iodine potassium

17

iodide) showed a poor correlation between pollen viability and spikelet fertility (r = 0.17) under

18

control conditions and was even weaker (r=0.05) in response to heat stress (Fig. 5 A and B). On

19

the other hand, using aniline blue to measure in vivo pollen viability by recording the percentage

20

of pollen germination on the stigmatic surface resulted in a doubling of the correlation

21

coefficient under control conditions (r = 0.34) and the correlation further increased in strength

22

under heat stress (r = 0.68) (Fig. 5 C and D). Interestingly, the overall variability captured by in

23

vivo pollen viability alone (R2 = 0.46, Fig 5 D) still left a lot of room for other aspects of pollen

16 This article is protected by copyright. All rights reserved.

development to be involved, such as pollen-tube growth rate (Jagadish et al. 2010), fertilization

2

and early embryo abortion. The simple and fast iodine stain is frequently used to estimate in vitro

3

pollen viability, but the poor correlation with fertility suggests that other techniques may be

4

better suited to estimating pollen viability. Starch stains have been widely used as it is crucial

5

that sufficient starch be available in the mature pollen to fuel subsequent fertilization events (De

6

Storme & Geelen 2014). This allows the use of iodine stains to measure pollen viability.

7

However, the role of pollen exine and intine, in particular the pollen coat, which comprises the

8

lipids and the proteins filling the exine cavities, is also essential for pollen viability (Suzuki et al.

9

2008), and hence viability is not just determined by starch content. Recently, using sorghum

Accepted Article

1

10

(Sorghum bicolor) pollen, a significant reduction in phospholipids accompanied by an increase

11

in reactive oxygen species accounted for a decrease in pollen activity under HNT (Vara Prasad &

12

Djanaguiraman, 2011). Similar approaches targeting pollen lipid and protein compositional

13

changes under heat stress conditions in contrasting rice cultivars would be an interesting future

14

research direction. Progress achieved in this area would allow for a comparative analysis of

15

pollen from diverse rice cultivars and other hardy species to identify mechanisms for heat

16

tolerance and genetic markers to assist plant breeders. Such markers would be more robust than

17

iodine stains and also less cumbersome than aniline blue staining, and could help establish a

18

high-throughput phenotyping protocol to identify candidates with heat-tolerant pollen.

19

The early developmental stages of pollen, including tetrad formation and early

20

microspore stages, are also susceptible to heat stress (Jagadish et al. 2013). The impact of

21

different abiotic stresses on early pollen development has been relatively well studied in

22

sorghum and wheat (Jain et al. 2007, 2010; Ji et al. 2010, 2011; Oliver et al. 2007), but not in

23

rice, except for Jagadish et al. (2013). Although a single study (Yoshida et al. 1981) indicates the

17 This article is protected by copyright. All rights reserved.

minimal role of the female reproductive organ in heat stress damage at anthesis, its vulnerability

2

during the most sensitive gamete formation stage has not been tested. Many reviews highlight

3

this as a challenge and a major knowledge gap in addressing abiotic stress−induced reproductive-

4

stage damage (Hedhly 2011; Hedhly et al. 2009).

5

High temperature impact under future warmer climate

6

Heat stress has been documented to have a negative interaction with drought and salinity stress

7

(Mittler, 2006). In general, increased [CO2] concentration has a beneficial effect because of

8

increased photosynthesis (Shimono et al. 2013). However, increased biomass accumulation was

9

not effective in ameliorating the impact of HDT stress−induced spikelet sterility (Matsui et al.

Accepted Article

1

10

1997; Madan et al. 2012). Madan et al. (2012) tested three contrasting rice cultivars for high

11

temperature and elevated CO2 combinations. The authors exposed the flowering stage to 5 days

12

of heat stress while the plants were grown under elevated CO2 throughout the crop growth

13

period. This experimental setup mimics a possible realistic scenario under gradually increasing

14

atmospheric CO2 concentration accompanied by a predicted increase in frequency of short

15

extreme heat spikes. None of the three cultivars benefited from elevated CO2 in terms of

16

lowering spikelet sterility when temperatures exceeded the critical threshold. Hence, elevated

17

CO2 is unable to compensate for heat stress damage during the flowering stage.

18

The impact of heat stress under current and future warming scenarios (1 to 3 oC,

19

following IPCC, 2013) was mapped taking rice-growing regions of South Asia as a case study.

20

Crop simulation models are increasingly being used to assess the impacts of climate change

21

(Lobell & Gourdji, 2012). The amount of detailed information available will increase the

22

accuracy of simulations that involve physiological processes. As an example, a complex

23

simulation model that will track the diurnal variation of temperature will need hourly

18 This article is protected by copyright. All rights reserved.

temperature as an input. It is often impractical to use such a model for spatial impact studies as

2

the input requirement for such high-resolution data is not available. Hence, the most appropriate

3

simulation model should capture the summary effects of temperature on spikelet sterility by

4

using coarse-resolution data of climate variables at a daily time step. To explore the impact of

5

future elevated temperature on spikelet sterility and its spatial variability, an upgraded version of

6

rice crop growth simulation model ORYZA2000 (Bouman et al. 2001) that simulates growth at a

7

daily time interval was used in our approach. The model simulates temperature responses to

8

spikelet fertility according to Horie (1993). The relationship from Horie (1993) was derived

9

across elevated and ambient CO2 concentrations and it shows that CO2 concentration has no

Accepted Article

1

10

effect on the temperature and spikelet fertility relationship (Bouman et al. 2001), similar to the

11

recent experimental evidence provided above. Using this model, we have explored the spatial

12

impact of elevated temperatures across rice-growing environments of South Asia.

13

Standard input parameters describing crop growth and development were used as inputs

14

to the model. A spatial soil dataset derived from the WISE soil database at 5 arc-minute

15

resolution (Batjes, 2006) was used as an input for the simulations. The simulations were carried

16

out on irrigated rice area grids having more than 75 hectares of rice area at 5 arc-minute

17

resolution using the MIRCA2000 database (Portmann et al. 2010). Daily solar radiation,

18

maximum and minimum temperatures, wind speed and vapor pressure from the NASA POWER

19

dataset (http://power.larc.nasa.gov) were downscaled to 15 arc minutes and bias corrected

20

(Sparks, A. UnPub) and combined with daily rainfall data derived from Tropical Rainfall

21

Measurement Mission (TRIMM ) (http://trmm.gsfc.nasa.gov). The simulations were carried out

22

under unlimited water and potential nitrogen supply for a 5-year period from 2006 to 2010 on the

23

irrigated rice grids. The most recent IPCC report explores future climate change weather outputs

19 This article is protected by copyright. All rights reserved.

from several GCMs (Global Climate Models) and different representative concentration

2

pathways (RCPs). It is shown that RCP 8.5 in three GCMs predicts an increase in global mean

3

temperature (GMT) by 1 oC by the early 2020s, by 2 oC by the mid-2040s and by 3 oC by the

4

mid-2060s (Warszawski et al. 2013). Keeping this in mind, four different scenarios were created

5

for simulations on irrigated rice grids across South Asia. A set of simulations with the original

6

weather (2006 to 2010) was the first scenario. An increase of 1 degree (scenario 2), 2 degrees

7

(scenario 3) and 3 degrees (scenario 4) centigrade to the daily maximum temperature represented

8

the other three scenarios. The CO2 fertilization effect was not considered. For all these scenarios

9

and for each rice grid, simulations were carried out once every 5 days during the planting

Accepted Article

1

10

window for the main rice season using a planting window database (IRRI, UnPub). From these

11

simulations, the best planting date was selected based on the average highest yields of five years

12

for each planting date for each grid from scenario 1. The five-year average spikelet sterility for

13

the same planting dates from all the scenarios for each grid was chosen for the comparisons. The

14

percentage spikelet sterility in the simulations was calculated based on the number of spikelets

15

and number of grains.

16

Simulations under current weather quantify variability in spikelet sterility across the rice

17

grids (Fig. 6 and Table. 1). Some 39.5% of the rice area showed less than 5% spikelet sterility

18

and another 50% of the rice area has sterility ranging between 5 and 15% under the current

19

scenario (Fig 6A and Table 1). In general, sink size is in excess under potential conditions and up

20

to 30% of the spikelets could remain unfilled (Sheehy et al. 2001). Under scenario 2 (1 oC

21

increase), the area recording sterility between 5 and 15% increased to 62% (Table 1). The

22

percent area that recorded more than 30% sterility increased significantly with each degree

23

increase in temperature, i.e. 0.3% of the area under current conditions to as high as 24.4% with a

20 This article is protected by copyright. All rights reserved.

3-degree rise in temperature (Fig. 6 A and C; Table 1). Flowering in cultivated rice usually takes

2

place during the daytime. However, the time of day of flowering varies from early morning to

3

midday depending on the cultivar and environment as detailed earlier. The relationship built in

4

the model used was at best an approximation and a best possible representation to be used

5

regionally bearing in mind the available input data at this scale.

Accepted Article

1

6

Increasing temperature will affect several plant growth processes and spikelet sterility is

7

one of them. Among these effects, an increase in nighttime temperature is shown to decrease rice

8

yield (Peng et al. 2004; Welch et al. 2010). Currently, few models have the capability to simulate

9

the effects of increased temperature on different crop growth and development processes.

10

ORYZA2000 is one of the few that has a temperature response function to simulate spikelet

11

sterility. However, other aspects such as response to increased minimum temperature and the

12

interactions of increased temperature and RH (vapor pressure difference) need to be addressed to

13

reduce uncertainties in simulating future impacts of climate change. Process-based understanding

14

of these issues is a precursor to incorporating such functionalities into simulation models.

15

Conclusions and future perspectives

16

Solutions to overcome current challenges faced with increasing temperature-induced yield losses

17

have advanced significantly, but examining the complex issues surrounding grain quality losses

18

continues to be a major challenge. Additional challenges that could emerge with the transition

19

from fully flooded rice cultivation to water-saving technologies need greater emphasis to ensure

20

that the advantage gained under fully flooded conditions facilitates the transition with minimum

21

damage under a future warmer and drier climate. To ensure sustained adoption of water-saving

22

technologies under future hotter climates, rice cultivars with enhanced tolerance of heat and

23

combined heat and drought stress during the floral meristem stage will be crucial to complement

21 This article is protected by copyright. All rights reserved.

the progress achieved in overcoming the damage across other sensitive developmental stages

2

such as flowering. With a more recent increase in research interest in addressing HNT impacts

3

on rice, caution needs to be exercised in imposing the right levels of stress and targeting traits

4

that can overcome the damage under realistic field conditions. In addition, the dynamic change in

5

temperature amplitude due to differential day and night temperature increase could potentially

6

affect crops differently than the known mechanisms of damage induced by either HDT or HNT.

7

The emergence of tolerant cultivars such as N22, which can tolerate both HDT and HNT at

8

flowering and gametogenesis stages, and heat-escaping strategy by employing EMF or LEF,

9

provides excellent opportunities to breed heat escape, tolerance or a combination of both

Accepted Article

1

10

strategies to induce greater resilience in rice to increasing temperatures. The quest to identify

11

such novel donors that can tolerate heat stress across sensitive developmental stages and

12

different environmental conditions needs to be intensified to provide sufficient options to

13

mitigate the impact of heat stress across hot-dry and hot-humid locations. In principle, the rice-

14

growing areas currently in the hot-dry regions with an advantage from transpiration cooling

15

could well be much more greatly affected than hot-humid regions with reduced irrigation

16

availability or the transition to water-saving technologies. To overcome damage at the flowering

17

stage, phenotyping protocols have been standardized for quantifying pollen viability, but a more

18

reliable and high-throughput technique or markers that can potentially allow assessment of large

19

genetic panels have yet to be identified. In comparison to the impact of heat stress on yield, the

20

mechanism leading to grain quality losses is more complex and requires intensified efforts to

21

continue producing high-quality grain. An interesting relationship between altered assimilate

22

supply and the expression of starch metabolic enzymes opens up an opportunity to identify

23

mechanisms regulating panicle senescence that could extend grain-filling duration. A high-

22 This article is protected by copyright. All rights reserved.

throughput technique such as chlorophyll fluorescence imaging that could quantify the active

2

grain-filling duration would be a way to identify new germplasm with longer grain-filling

3

duration under stress. This extension of grain-filling duration would help counteract the direct

4

impact of heat stress and indirectly postpone the enzymatic trigger that stops active starch

5

metabolism in filling grains. Interestingly, the critical role of sugars/carbohydrates and invertases

6

in overcoming the negative impact of abiotic stress (cold, drought and heat stress) during

7

gametogenesis and early pollen development has been documented across cereals (Ji et al. 2010;

8

Jain et al. 2010; Nguyen et al. 2010). On the other hand, a higher CO2 resulting in increased

9

biomass and assimilate supply fails to reduce heat stress−induced spikelet sterility losses at the

10

anthesis stage (Madan et al. 2012), indicating a missing link to stress response in reproductive

11

organs across their developmental stages. Finally, the major factor driving climate change is

12

increased CO2 and it is projected to continue to increase gradually, but research indicates a lack

13

of amelioration of heat stress−induced yield loss under elevated CO2 conditions. However,

14

progress achieved through breeding efforts to safeguard sensitive reproductive processes from

15

heat stress would allow better use of the additional biomass accumulated from gradually

16

increasing [CO2]. Indeed, this is an intriguing hypothesis that could partially address the

17

persisting challenge to increase rice productivity in spite of a projected warming climate.

Accepted Article

1

18

23 This article is protected by copyright. All rights reserved.

Accepted Article

1 2

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43

References˜ Batjes N.H. (2006) ISRIC-WISE derived soil properties on a 5 by 5 arc-minutes global grid. Report 2006/02 (available through www.isric.org), ISRIC – World Soil Information, Wageningen. Battisti D.S. & Naylor R.L. (2009) Historical warnings of future food insecurity with unprecedented seasonal heat. Science 323, 240−244. Bouman B.A.M., Kropff M.J., Tuong T.P., Wopereis M.C.S., ten Berge H.F.M. & Van Laar H.H. (2001) ORYZA2000: Modeling lowland rice. International Rice Research Institute, Los Baños, Philippines and Wageningen University and Research Centre, Wageningen, The Netherlands. Bouman B.A.M., Peng S., Castañeda A.R. & Visperas R.M. (2005) Yield and water use of irrigated tropical aerobic rice systems. Agricultural Water Management 74, 87–105. Bouman B.A.M., Feng L., Tuong T.P., Lu G., Wang H. & Feng Y. (2007) Exploring options to grow rice using less water in northern China using a modelling approach. II. Quantifying yield, water balance components, and water productivity. Agricultural Water Management 88, 23−33. Bueno A.C.R., Prudente D.A., Machado E.C. & Ribeiro R.V. (2012) Daily temperature amplitude affects the vegetative growth and carbon metabolism of orange trees in a rootstock-dependent manner. Journal of Plant Growth Regulation 31, 309–319. Challinor A.J., Watson J., Lobell D.B., Howden S.M., Smith D.R. & Chhetri N. (2014) A metaanalysis of crop yield under climate change and adaptation. Nature Climate Change DOI: 10.1038/NCLIMATE2153 Cheng W., Sakai H., Yagi K. & Hasegawa T. (2009) Interactions of elevated [CO 2] and night temperature on rice growth and yield. Agricultural and Forest Meteorology 149, 51−58. De Storme N. & Geelen D. (2014) The impact of environmental stress on male reproductive development in plants: biological processes and molecular mechanisms. Plant, Cell & Environment 37, 1−18. Fitzgerald M.A., McCouch S.R. & Hall R.D. (2009) Not just a grain of rice: the quest for quality. Trends in Plant Science 14, 133−139. Fitzgerald M.A. & Resurreccion A.P. (2009) Maintaining the yield of edible rice in a warming world. Functional Plant Biology 36, 1037–1045.

24 This article is protected by copyright. All rights reserved.

Glaubitz U., Li X., Köhl K.I., van Dongen J.T., Hincha D.K. & Zuther E. (2014) Differential physiological responses of different rice (Oryza sativa) cultivars to elevated night temperature during vegetative growth. Functional Plant Biology 41, 437-448.

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26

Gourdji S.M., Sibley A.M. & Lobell D.B. (2013) Global crop exposure to critical high temperatures in the reproductive period: historical trends and future projections. Environmental Research Letters 8, doi: 10.1088/1748-9326/8/2/024041. Hakata M., Kuroda M., Miyashita T., Yamaguchi T., Kojima M., Sakakibara H., …, Yamakawa H. (2012) Suppression of α-amylase genes improves quality of rice grain ripened under high temperature. Plant Biotechnology Journal 10, 1110–1117. Hasegawa T., Kuwagata T., Nishimori M., Ishigooka Y., Murakami M., Yoshimoto M., …, Matsuzaki H. (2009) Recent warming trends and rice growth and yield in Japan. MARCO Symposium on Crop Production under Heat Stress: Monitoring, Impact Assessment and Adaptation. Tsukuba, Japan. Hedhly A. (2011) Sensitivity of flowering plant gametophytes to temperature fluctuations. Environmental and Experimental Botany 74, 9–16. Hedhly A., Hormaza J.I. & Herrero M. (2009) Global warming and sexual plant reproduction. Trends in Plant Science 14, 30–36. Horie T. (1993) Predicting the effect of climate variation and elevated CO2 on rice yield in Japan. Journal of Agricultural Meteorology 48, 567–574. IPCC. (2013) Working Group I Contribution to the IPCC Fifth Assessment Report on Climate Change 2013: The Physical Science Basis, Summary for Policymakers. www.climatechange2013.org/images/report/WG1AR5_SPM_FINAL.pdf

27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

Ishimaru T., Hirabayashi H., Ida M, Takai T., San-Oh Y.A., Yoshinaga S., …, Kondo M. (2010) A genetic resource for early-morning flowering trait of wild rice Oryza officinalis to mitigate high temperature-induced spikelet sterility at anthesis. Annals of Botany 106, 515−520. Ishimaru T., Horigane A.K., Ida M., Iwasawa N., San-oh Y.A., Nakazono M., …, Yoshida M. (2009) Formation of grain chalkiness and changes inwater distribution in developing rice caryopses grown under high-temperature stress. Journal of Cereal Science 50, 166–174. Jagadish K.S.V., Cairns J.E., Kumar A., Somayanda I.M. & Craufurd P.Q. (2011) Does susceptibility to heat stress confound screening for drought tolerance in rice? Functional Plant Biology 38, 261−269. Jagadish S.V.K., Craufurd P.Q. & Wheeler T.R. (2007) High temperature stress and spikelet fertility in rice (Oryza sativa L.). Journal of Experimental Botany 58, 1627–1635.

25 This article is protected by copyright. All rights reserved.

Jagadish S.V.K., Craufurd P.Q. & Wheeler T.R. (2008) Phenotyping parents of mapping populations of rice for heat tolerance during anthesis. Crop Science 48, 1140–1146.

Accepted Article

1 2 3

Jagadish K.S.V., Craufurd P.Q., Shi W. & Oane R. (2013) A phenotypic marker for quantifying heat stress impact during microsporogenesis in rice (Oryza sativa L.). Functional Plant Biology 41, 48−55.

4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

Jagadish S.V.K., Muthurajan R., Oane R., Wheeler T.R., Heuer S., Bennett J. & Craufurd P.Q. (2010) Physiological and proteomic approaches to address heat tolerance during anthesis in rice (Oryza sativa L.). Journal of Experimental Botany 61, 143–156. Jain M., Chourey P.S., Boote K.J. & Allen Jr. L.H. (2010) Short-term high temperature growth conditions during vegetative-to-reproductive phase transition irreversibly compromise cell wall invertase-mediated sucrose catalysis and microspore meiosis in grain sorghum (Sorghum bicolor). Journal of Plant Physiology 167, 578–582. Jain M., Vara Prasad P.V., Boote K.J., Hartwell Jr. A.L. & Chourey P.S. (2007) Effects of season-long high temperature growth conditions on sugar-to-starch metabolism in developing microspores of grain sorghum (Sorghum bicolor L. Moench). Planta 227, 67– 79. Ji X., Dong B., Shiran B., Talbot M.J., Edlington J.E., Hughes T., …, Dolferus R. (2011) Control of abscisic acid catabolism and abscisic acid homeostasis is important for reproductive stage stress tolerance in cereals. Plant Physiology 156, 647–662. Ji X., Shiran B., Wan J., Lewis D.C., Jenkins C.L.D., Condon A.G., …, Dolferus R. (2010) Importance of pre-anthesis anther sink strength for maintenance of grain number during reproductive stage water stress in wheat. Plant, Cell and Environment 33, 926–942. Julia C. & Dingkuhn M. (2013) Predicting temperature induced sterility of rice spikelets requires simulation of crop-generated microclimate. European Journal of Agronomy 49, 50–60. Kadam N.N., Xiao G., Melgar R.J., Bahuguna R.N., Quinones C., Tamilselvan A., …, Jagadish SVK. (2014) Agronomic and physiological responses to high temperature, drought and elevated CO2 interaction in cereals. Advances in Agronomy In press. Kato Y., Kamoshita A. & Yamagishi J. (2008) Preflowering abortion reduces spikelet number in upland rice (Oryza sativa L.) under water stress. Crop Science 48, 2389–2395. Kobata T. & Uemuki N. (2004) High temperatures during the grain-filling period do not reduce the potential grain dry matter increase of rice. Agronomy Journal 96, 406–414. Kreye C., Bouman B.A.M., Reversat G., Fernandez L., Vera Cruz C., Elazegui F., …, Llorca L. (2009a) Biotic and abiotic causes of yield failure in tropical aerobic rice. Field Crops Research 112, 97–106.

26 This article is protected by copyright. All rights reserved.

Kreye C., Bouman B.A.M., Castañeda A.R., Lampayan R.M., Faronilo J.E., Lactaoen A.T. & Fernandez L. (2009b) Possible causes of yield failure in tropical aerobic rice. Field Crops Research 111, 197–206.

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45

Kumar V. & Ladha J.K. (2011) Direct seeding of rice: recent developments and future research needs. Advances in Agronomy 111, 297–413. Laborte A., Nelson A., Jagadish K., Aunario J., Sparks A., Ye C. & Redoña E. (2012) Rice feels the heat. Rice Today 11, 30−31. Li C-Y., Peng C-H., Zhao Q-B., Xie P. & Chen W. (2004) Characteristic analysis of the abnormal high temperature in 2003’s midsummer in Wuhan City. Journal of Central China Normal University (Natural Sciences) 38, 379−381. Lobell D.B. & Gourdji S.M. (2012) The influence of climate change on global crop productivity. Plant Physiology 160, 1686–1697. Lobell D.B., Sibley A. & Ortiz-Monasterio J.I. (2012) Extreme heat effects on wheat senescence in India. Nature Climate Change 2, 186−189. Lobell D.B., Hammer G.L., McLean G., Messina C., Roberts M.J. & Schlenker W. (2013) The critical role of extreme heat for maize production in the United States. Nature Climate Change 3, 497−501. Lyman N.B., Jagadish K.S.V., Nalley L.L., Dixon B.L. & Siebenmorgen T. (2013) Neglecting rice milling yield and quality underestimates economic losses from high-temperature stress. PloS ONE 8, e72157. Madan P., Jagadish S.V.K., Craufurd P.Q., Fitzgerald M., Lafarge T. & Wheeler T.R. (2012) Effect of elevated CO2 and high temperature on seed-set and grain quality of rice. Journal of Experimental Botany 63, 3843−3852. Mahajan G., Chauhan B.S. & Gill M.S. (2013) Dry-seeded rice culture in Punjab State of India: lessons learned from farmers. Field Crops Research 144, 89–99. Matsui T., Kobayasi K., Yoshimoto M. & Hasegawa T. (2007) Stability of rice pollination in the field under hot and dry conditions in the Riverina region of New South Wales, Australia. Plant Production Science 10, 57–63. Matsui T., Namuco O.S., Ziska L.H. & Horie T. (1997) Effects of high temperature and CO2 concentration on spikelet sterility in indica rice. Field Crops Research 51, 213−219. Mittler R. (2006) Abiotic stress, the field environment and stress combination. Trends in Plant Science 11, 15−19.

27 This article is protected by copyright. All rights reserved.

Mohammed R., Cothren J.T. & Tarpley L. (2013) High night temperature and abscisic acid affect rice productivity through altered photosynthesis, respiration and spikelet fertility. Crop Science 53, 2603–2612. Mohammed A-R. & Tarpley L. (2009a) Impact of high nighttime temperature on respiration, membrane stability, antioxidant capacity, and yield of rice plants. Crop Science 49, 313−322.

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Mohammed A.R. & Tarpley L. (2009b) High nighttime temperatures affect rice productivity through altered pollen germination and spikelet fertility. Agricultural and Forest Meteorology 149, 999−1008. Mohammed A.R. & Tarpley L. (2010) Effects of high night temperature and spikelet position on yield-related parameters of rice (Oryza sativa L.) plants. European Journal of Agronomy 33, 117−123. Mohammed A.R. & Tarpley L. (2011) Effects of night temperature, spikelet position and salicylic acid on yield and yield-related parameters of rice (Oryza sativa L.) plants. Journal of Agronomy and Crop Science 197, 40−49. Munns R., James R.A., Sirault X.R.R., Furbank R.T. & Jones H.G. (2010) New phenotyping methods for screening wheat and barley for beneficial responses to water deficit. Journal of Experimental Botany 61, 3499–3507. Nagarajan S., Jagadish S.V.K., Hari Prasad A.S., Thomar A.K., Anand A., Madan P. & Agarwal P.K. (2010) Local climate affects growth, yield and grain quality of aromatic and nonaromatic rice in northwestern India. Agriculture, Ecosystems & Environment 138, 274−281. Nguyen G.N., Hailstones D.L., Wilkes M. & Sutton B.G. (2010). Role of carbohydrate metabolism in drought-induced male sterility in rice anthers. Journal of Agronomy and Crop Science 196, 346–357. Nguyen D-N., Lee K-J., Kim D-I., Anh N.T. & Lee B-W. (2013) Modeling and validation of high-temperature induced spikelet sterility in rice. Field Crops Research 156, 293–302. Nie L., Peng S., Bouman B.A.M., Huang J., Cui K., Visperas R.M. & Park H-K. (2007) Alleviating soil sickness caused by aerobic monocropping: responses of aerobic rice to soil oven-heating. Plant Soil 300, 185–195. Nie L., Peng S., Bouman B.A.M., Huang J., Cui K, Visperas R.M. & Xiang J. (2009) Alleviating soil sickness caused by aerobic monocropping: responses of aerobic rice to various nitrogen sources. Soil Science and Plant Nutrition 55, 150–159. Oliver S.N., Dennis E.S. & Dolferus R. (2007) ABA regulates apoplastic sugar transport and is a potential signal for cold-induced pollen sterility in rice. Plant & Cell Physiology 48, 1319–1330.

28 This article is protected by copyright. All rights reserved.

Peng S., Huang J., Sheehy J.E., Laza R.C., Visperas R.M., Zhong X., …, Cassman K.G. (2004) Rice yields decline with higher night temperature from global warming. Proceedings of the National Academy of Sciences of the United States of America 101, 9971−9975.

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

Phan T.T.T., Ishibashi Y., Miyazaki M., Tran H.T., Okamura K., Tanaka S., …, Iwaya-Inoue M. (2013) High temperature-induced repression of the rice sucrose transporter (OsSUT1) and starch synthesis-related genes in sink and source organs at milky ripening stage causes chalky grains. Journal of Agronomy and Crop Science 199, 178–188 Portmann F., Siebert S. & Döll P. (2010) MIRCA2000—Global monthly irrigated and rainfed crop areas around the year 2000: a new high-resolution data set for agricultural and hydrological modeling. Global Biogeochemical Cycles 24 GB1011, doi:10.1029/2008GB003435 Prasad P.V.V., Boote K.J., Allen Jr. L.H., Sheehy J.E. & Thomas J.M.G. (2006) Species, ecotype and cultivar differences in spikelet fertility and harvest index of rice in response to high temperature stress. Field Crops Research 95, 398–411. Rang Z.W., Jagadish S.V.K., Zhou Q.M., Craufurd P.Q. & Heuer S. (2011) Effect of heat and drought stress on pollen germination and spikelet fertility in rice. Environmental and Experimental Botany 70, 58-65. Rao B.B., Chowdary P.S., Sandeep V.M., Rao V.U.M. & Venkateswarlu B. (2014) Rising minimum temperature trends over India in recent decades: implications for agricultural production. Global and Planetary Change 117, 1–8. Rizhsky L., Liang H. & Mittler R. (2002) The combined effect of drought stress and heat shock on gene expression in tobacco. Plant Physiology 130, 1143–1151. Sanchez B., Rasmussen A. & Porter J.R. (2014) Temperatures and the growth and development of maize and rice: a review. Global Change Biology 20, 408−417. Shah F., Huang J., Cui K., Nie L., Shah T., Chen C. & Wang K. (2011) Impact of hightemperature stress on rice plant and its traits related to tolerance. The Journal of Agricultural Science 149, 545−556. Sheehy J.E., Dionora M.J.A. & Mitchell P.L. (2001) Spikelet numbers, sink size and potential yield in rice. Field Crops Research 71, 77−85. Sheehy J.E., Mabilangan A.E., Dionora M.J.A. & Pablico P.P. (2007) Time of day of flowering in wild species of the genus Oryza. International Rice Research Notes 32, 12−13. Shi W., Muthurajan R., Rahman H., Selvam J., Peng S., Zou Y. & Jagadish K.S.V. (2013) Source–sink dynamics and proteomic reprogramming under elevated night temperature and their impact on rice yield and grain quality. New Phytologist 197, 825−837.

29 This article is protected by copyright. All rights reserved.

Shi W., Ishimaru T., Gannaban R.B., Oane W. & Jagadish S.V.K. (2014) Popular rice (Oryza sativa L.) cultivars show contrasting responses to heat stress at gametogenesis and anthesis. Crop Science doi:10.2135/cropsci2014.01.0054.

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

Shimono H., Nakamura H., Hasegawa T. & Okada M. (2013) Lower responsiveness of canopy evapotranspiration rate than of leaf stomatal conductance to open-air CO2 elevation in rice. Global Change Biology 19, 2444−2453. Song X., Du Y., Song X. & Zhao Q. (2013) Effect of high night temperature during grain filling on amyloplast development and grain quality in japonica rice. Cereal Chemistry 90, 114– 119. Suzuki T., Masaoka K., Nishi M., Nakamura K. & Ishiguro S. (2008) Identification of kaonashi mutants showing abnormal pollen exine structure in Arabidopsis thaliana. Plant and Cell Physiology 49, 1465–1477. Teixeira E.I., Fischer G., van Velthuizen, H., Walter C. & Ewert F. (2013). Global hot-spots of heat stress on agricultural crops due to climate change. Agricultural and Forest Meteorology 170, 206−215. Tian X., Matsui T., Li S., Yoshimoto M., Kobayasi K. & Hasegawa T. (2010) Heat-induced floret sterility of hybrid rice (Oryza sativa L.) cultivars under humid and low wind conditions in the field of Jianghan Basin, China. Plant Production Science 13, 243–251. van Oort P.A.J., Saito K., Zwart S.J. & Shrestha S. (2014) A simple model for simulating heat induced sterility in rice as a function of flowering time and transpirational cooling. Field Crops Research 156, 303–312. Vara Prasad P.V. & Djanaguiraman M. 2011. High night temperature decreases leaf photosynthesis and pollen function in grain sorghum. Functional Plant Biology 38, 993– 1003. Vose R.S., Easterling D.R. & Gleason B. (2005) Maximum and minimum temperature trends for the globe: an update through 2004. Geophysical Research Letters 32, L23822. Warszawski L., Frieler K., Huber V., Piontek F., Serdeczny O. & Schewe J. (2013) The intersectoral impact model intercomparison project (ISI-MIP): project framework. Proceedings of the National Academy of Sciences of the United States of America 111, 3228−3232. Wassmann R., Jagadish S.V.K., Sumfleth K., Pathak H., Howell G., Ismail A., …, Heuer S. (2009) Regional vulnerability of climate change impacts on Asian rice production and scope for adaptation. Advances in Agronomy 102, 93–133. Weerakoon W.M.W., Maruyama A. & Ohba K. (2008) Impact of humidity on temperatureinduced grain sterility in rice (Oryza sativa L). Journal of Agronomy and Crop Science 194, 135–140.

30 This article is protected by copyright. All rights reserved.

Welch J.R., Vincent J.R., Auffhammer M., Moya P.F., Dobermann A. & Dawe D. (2010) Rice yields in tropical/subtropical Asia exhibit large but opposing sensitivities to minimum and maximum temperatures. Proceedings of the National Academy of Sciences of the United States of America 107, 14562−14567.

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

White J.W., Kimball B.A., Wall G.W., Ottman M.J. & Hunt L.A. (2011) Responses of time of anthesis and maturity to sowing dates and infrared warming in spring wheat. Field Crops Research 124, 213–222. Yang J.C. & Zhang J.H. (2010) Grain-filling problem in ‘super’ rice. Journal of Experimental Botany 61, 1–4. Ye C., Argayoso M.A., Redoña E.D., Sierra S.N., Laza M.A., Dilla C.J., …, Hernandez J.E. (2012) Mapping QTL for heat tolerance at flowering stage in rice using SNP markers. Plant Breeding 131, 33–41. Yoshida S., Satake T. & Mackill D. (1981) High temperature stress. IRRI Research Paper Series No. 67, 1−15. Yoshimoto M., Fukuoka M., Hasegawa T., Utsumi M., Ishigooka Y. & Kuwagata T. (2011) Integrated micrometeorology model for panicle and temperature (IM2PACT) for rice heat stress studies under climate change. Journal of Agricultural Meteorology 67, 233–247. Zakari S., Matsuda T., Tajima S. & Nitta Y. (2002) Effect of high temperature at ripening stage on the reserve accumulation in seed in some rice cultivars. Plant Production Science 5, 160−168. Zhang Y., Tang Q., Peng S., Zou Y., Chen S., Shi W. & Laza M.R.C. (2013) Effects of high night temperature on yield and agronomic traits of irrigated rice under field chamber system conditions. Australian Journal of Crop Science 7, 7−13. Zhang Z.X., Chen J., Lin S., Li Z., Cheng R.H., Fang C.X., …, Lin W.X. (2012) Proteomic and phosphoproteomic determination of ABA’s effects on grain-filling of Oryza sativa L. inferior spikelets. Plant Science 185, 259–273. Zhao X. & Fitzgerald M. (2013) Climate change: implications for the yield of edible rice. PLoS ONE 8(6): e66218. Zhou L., Dickinson R.E., Tian Y., Fang J., Kaufmann R.K., Tucker C.J. & Myneni R.B. (2004) Evidence for a significant urbanization effect on climate in China. Proceedings of the National Academy of Sciences of the United States of America 101, 9540-9544.

31 This article is protected by copyright. All rights reserved.

Figure legends

3

Fig. 1 Rice plants maintained under fully flooded conditions exposed to control (30 oC) and heat

4

stress (39 oC) for consecutive 4 days (6 hours of stress [0900 to 1500] on each day following

5

Jagadish et al. 2010) coinciding with the panicle initiation stage resulted in no reduction in

6

spikelet fertility (A) or sink size (B). Grey striped bars are data obtained from independent

7

experiments and data for number of spikelets in N22 (6264) is not available. Numbers in

8

parentheses after the cultivar are the IRRI Genebank accession numbers. Bars indicate ±SE

9

(Jagadish et al. UnPub).

Accepted Article

1 2

10

Fig. 2 Soil, water and air temperature recorded using thermocouples under control (30 oC) and

11

high temperatures (39 oC) for consecutive 4 days (6 hours of stress [0900 to 1500] on each day

12

following Jagadish et al. 2010) coinciding with panicle initiation stage under fully flooded pots

13

in controlled environment chambers (A). Soil, air and panicle temperature under fully flooded

14

(puddled) transplanted rice (PTR) and direct-seeded rice (DSR) at 10 kPa (B), recorded using

15

water-proof temperature pendants (Onset HOBO data loggers, Utah, USA) for 15 days

16

coinciding with flowering. Panicle temperature on ten independent panicles was recorded under

17

PTR and 10 kPa DSR conditions using a thermal camera (NEC Avio Infrared Technologies Co.

18

Ltd., Tokyo, Japan). Bars indicate SE.

19

Fig. 3 High night temperature (HNT) exposure under controlled environments (A) and field

20

conditions (B) shows contrasting responses to stress impact on spikelet fertility, with only the

21

former recording a significant decline in fertility. Published literature using environmentally

22

controlled growth chambers (control night temperature ranging between 22 and 27 oC and HNT

23

from 30 to 32 oC − Mohammed and Tarpley, 2011 and 2010; Cheng et al. 2009; Mohammed et

24

al. 2013) indicates HNT inducing a significant reduction in spikelet fertility. Under field 32

This article is protected by copyright. All rights reserved.

conditions (see supplementary Fig. S1 in Shi et al. 2013), exposure to optimum night

2

temperature between 22 and 23 oC and HNT ranging between 28 and 29 oC using two contrasting

3

cultivars (Shi et al. 2013) from a preliminary screen of 36 cultivars over four different seasons

4

exposed to day/night temperature of 22/27 oC (Zhang Y et al. 2013) indicated no reduction in

5

spikelet fertility. Similarly, diverse genotypes, across four seasons (two dry seasons (DS) and

6

two wet seasons (WS) at IRRI), involving HNT and different nitrogen levels (150 and 250 N

7

during the DS and 75 and 125 N during the WS) and interactions (Shi et al. UnPub. data

8

indicated by blue and white bars in panel B), including different indica cultivars and hybrids,

9

validate the hypothesis that HNT does not directly lead to a yield reduction through a decline in

Accepted Article

1

10

spikelet fertility. Average day temperatures across all the field experiments including Zhang Y et

11

al. (2013) and Shi et al. (2013) varied between 27 and 30 oC. Day temperatures ranged between

12

32 and 33.9 oC in Cheng et al. (2009) and Mohammed and Tarpley (2013), with Mohammed and

13

Tarpley 2010 and 2011 recording temperatures ranging between 31.6 and 32.8 oC.

14

Fig. 4 Time of day of flowering as an adaptive mechanism to mitigate high-temperature-induced

15

reduction in spikelet fertility. Data presented for Oryza sativa Koshihikari and Koshihikari early-

16

morning flowering (EMF) line were extracted from Ishimaru et al. (2010). The flowering pattern

17

of O. australiensis indicating the late-evening flowering (LEF) was obtained by recording the

18

flowering pattern over 5 flowering days (Quinones et al. UnPub) and supports the finding of

19

Sheehy et al. (2007).

20

Fig. 5 Pollen viability estimated using iodine potassium iodide (IKI) correlates poorly with

21

spikelet fertility under control (A; Y=0.43x + 43.42, R2=0.03, n = 31 entries) and more so under

22

heat stress (B; Y=0.02x + 84.93, R2=0.003, n=31). In vivo pollen viability measured by number

23

of pollen germinated on the stigma is strongly correlated with spikelet fertility under control (C;

33 This article is protected by copyright. All rights reserved.

Y=0.22x + 83.51, R2=0.12, n=32) and with higher significance under heat stress (D; Y=0.23x +

2

2.15, R2=0.46, n=32). Black circles (Jagadish et al. UnPub) are data obtained simultaneously

3

from the same set of plants following the same crop management practices and exposed to 39 oC

4

following Jagadish et al. (2010) and white circles are data extracted from published literature

5

(Prasad et al. 2006; Rang et al. 2011; Jagadish et al. 2010).

6

Fig. 6 Spatial variability of percent spikelet sterility across South Asia with current (2006 to

7

2010) temperature as baseline (A), and , 2 oC (B) and 3 oC (C) increase in maximum

8

temperature.

Accepted Article

1

34 This article is protected by copyright. All rights reserved.

Accepted Article

1 2

Table 1 Percent rice area across South Asia and spikelet sterility induced under current and three

3

future scenarios, that is with 1, 2 and 3 oC increase in maximum temperature over the current

4

baseline (2006 to 2010) and coinciding with the critical flowering stage. Numbers in the table

5

indicate percent rice area.

6

Temperature

Spikelet sterility (%)

Rice responses to rising temperatures--challenges, perspectives and future directions.

Phenotypic plasticity in overcoming heat stress-induced damage across hot tropical rice-growing regions is predominantly governed by relative humidity...
971KB Sizes 0 Downloads 6 Views