You are on page 1of 59

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/265852219

Advances in Agronomy

Article in Advances in Agronomy · January 2014

CITATIONS READS

5 6,642

6 authors, including:

Bhagirath S Chauhan Gulshan Mahajan


The University of Queensland Punjab Agricultural University
472 PUBLICATIONS 16,622 CITATIONS 128 PUBLICATIONS 5,696 CITATIONS

SEE PROFILE SEE PROFILE

Dr Ramanjit Kaur Harpreet Singh


Indian Council of Agricultural Research Punjab Agricultural University
135 PUBLICATIONS 807 CITATIONS 50 PUBLICATIONS 276 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Manjit S Kang on 02 November 2018.

The user has requested enhancement of the downloaded file.


Provided for non-commercial research and educational use only.
Not for reproduction, distribution or commercial use.

This chapter was originally published in the book Advances in Agronomy, Vol. 123,
published by Elsevier, and the attached copy is provided by Elsevier for the author's
benefit and for the benefit of the author's institution, for non-commercial research and
educational use including without limitation use in instruction at your institution,
sending it to specific colleagues who know you, and providing a copy to your
institution’s administrator.

All other uses, reproduction and distribution, including without limitation commercial
reprints, selling or licensing copies or access, or posting on open internet sites, your
personal or institution’s website or repository, are prohibited. For exceptions,
permission may be sought for such use through Elsevier's permissions site at:
http://www.elsevier.com/locate/permissionusematerial

From: Bhagirath Singh Chauhan, Prabhjyot-Kaur, Gulshan Mahajan,


Ramanjit Kaur Randhawa, Harpreet Singh, Manjit S. Kang, Global Warming
and its Possible Impact on Agriculture in India. In Donald L. Sparks, editor:
Advances in Agronomy, Vol. 123,
Burlington: Academic Press, 2014, pp. 65-121.
ISBN: 978-0-12-420225-2
© Copyright 2014 Elsevier Inc.
Academic Press
Author's personal copy

CHAPTER TWO

Global Warming and Its Possible


Impact on Agriculture in India
Bhagirath Singh Chauhan*, Prabhjyot-Kaur†, Gulshan Mahajan†,
Ramanjit Kaur Randhawa{, Harpreet Singh†, Manjit S. Kang}
*Crop and Environmental Sciences Division, International Rice Research Institute, Los Baños, Philippines

Punjab Agricultural University, Ludhiana, Punjab, India
{
Indian Agricultural Research Institute, New Delhi, India
}
Department of Plant Pathology, Kansas State University, Manhattan, Kansas, USA

Contents
1. Introduction 66
2. Greenhouse Effect and Global Warming 67
3. Agents of Global Warming 68
3.1 Carbon dioxide 69
3.2 Methane 69
3.3 Nitrous oxide 69
3.4 Water vapor 70
3.5 Ozone (O3) 70
4. Evidence for Climate Change and Impacts on Agriculture 70
5. Projected Climate Change in India 72
6. Impact of Global Warming on Agriculture and the Food Supply 75
6.1 Effect of elevated concentrations of CO2 on crop growth 76
6.2 Effect of ozone on plants 80
6.3 Effect of increasing temperature on crop growth 81
6.4 Interactive effects of changing climatic factors on crop production 86
6.5 Effect of climate change on the quality of produce 90
6.6 Agricultural surfaces and climate change 91
6.7 Soil erosion and soil fertility 93
6.8 Potential effects of climate change on pests 97
7. Key Adaptation and Mitigation Strategies to Reduce the Effects of Climate Change 103
7.1 Crop-based approaches 105
7.2 Crops and cultivars that fit into new cropping systems and seasons 106
7.3 Cultivars suitable for high temperature, drought, inland salinity, and
submergence tolerance 107
7.4 Cultivars that respond to high CO2 concentration 109
7.5 Mitigation of the impact of climate change 109
7.6 Other strategies 111
7.7 Policy issues for managing climate change 112

Advances in Agronomy, Volume 123 # 2014 Elsevier Inc. 65


ISSN 0065-2113 All rights reserved.
http://dx.doi.org/10.1016/B978-0-12-420225-2.00002-9
Author's personal copy

66 Bhagirath Singh Chauhan et al.

8. Conclusions 113
Acknowledgment 114
References 114

Abstract
Progress has been significant in climate science and the direct and indirect influences of
climate on agricultural productivity. With the likely growth of the world's population
toward 10 billion by 2050, demand for food crops will grow faster than demand for
other crops. The prospective climate change is global warming (with associated
changes in hydrologic regimes and other climatic variables) induced by the increasing
concentration of radiatively active greenhouse gases. Climate models project that
global surface air temperatures may increase by 4.0–5.8  C in the next few decades.
These increases in temperature will probably offset the likely benefits of increasing
atmospheric concentrations of carbon dioxide on crop plants. Climate change would
create new environmental conditions over space and time and in the intensity and fre-
quency of weather and climate processes. Therefore, climate change has the potential
to influence the productivity of agriculture significantly. Climate variability has also
become a reality in India. The increase in mean temperature by 0.3–0.6  C per decade
since the 1860s across India indicates significant warming due to climate change. This
warming trend is comparable to global mean increases in temperature in the past
100 years. It is projected that rainfall patterns in India would change with the western
and central areas witnessing as many as 15 more dry days each year, whereas the north-
ern and northwestern areas could have 5 to 10 more days of rainfall annually. Thus, dry
areas are expected to get drier and wet areas wetter. It is projected that India's popu-
lation could reach 1.4 billion by 2025 and may exceed China's in the 2040s. If agricultural
production is adversely affected by climate change, livelihood and food security in India
would be at risk. Because the livelihood system in India is based on agriculture, climate
change could cause increased crop failure and more frequent incidences of pests.
Therefore, future challenges will be more complex and demanding. This chapter
focuses on the variability of climate change and its probabilistic effects on agricultural
productivity and adaptation and mitigation strategies that can help in managing the
adverse effect of climate change on agricultural productivity, in particular for India.

1. INTRODUCTION
Climate is the synthesis of weather conditions in a given area, charac-
terized by long-term statistics (mean values, variances, probabilities of
extreme values, etc.) for the meteorologic elements in that area (World
Meteorological Organization, 1992). Generally, the quantities measured
are surface variables, such as temperature, precipitation, and wind. More
broadly, the “climate” is the description of the state of the climate system
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 67

(IPCC, 1995). Climate variability reflects deviations in climate statistics


across a given period (a specific month, season, or year) from the long-term
climate statistics relating to the corresponding calendar period (World
Meteorological Organization, 1992).
Climate affects human life on Earth. It regulates food production and
water resources and influences energy use, disease transmission, and other
aspects of human health and well-being (National Research Council
(United States), 2010). Earth’s atmosphere is mainly composed of nitrogen
(N) and oxygen (O2), but these gases have little or no influence on radiation
coming from the sun or that emitted by Earth’s surface. The so-called green-
house gases (GHGs), which include water vapor, carbon dioxide (CO2),
methane (CH4), nitrous oxide (N2O), and chlorofluorocarbons (CFCs),
however, absorb and reemit infrared radiation emitted by the surface of
the Earth and trap heat in the atmosphere. This amplified warming keeps
Earth’s surface warmer (about 33  C) than it would be without the presence
of GHGs (National Research Council (United States), 2010).
Then, what is climate change? According to the World Meteorological
Organization (1992), climate change represents a significant change, that is, a
change with important economic, environmental, and social effects, in the
mean values of a meteorologic element, such as temperature and amount of
precipitation during a certain period, for which the means are computed
across a decade or longer (World Meteorological Organization, 1992). In
the IPCC usage, climate change occurs because of internal changes within
the climate system or in the interaction among its components or because of
changes in external forcing, either for natural reasons or because of human
activities (IPCC, 1995). Projections of future climate change reported by the
IPCC generally consider the influence on the climate of only anthropogenic
increases in GHGs and other human-related factors (IPCC, 1995). Agricul-
ture accounted for 10–12% of the total global anthropogenic emissions of
GHGs (IPCC, Working Group III, 2007).

2. GREENHOUSE EFFECT AND GLOBAL WARMING


According to the Free Dictionary by Farlex, greenhouse effect is the
phenomenon whereby Earth’s atmosphere traps solar radiation, caused by
the presence in the atmosphere of GHGs, which allow incoming sunlight
to pass through but absorb heat radiated back from Earth’s surface. Earth’s
atmosphere balances the absorption of solar radiation with emission of long-
wave radiation (infrared) to space. Not only does Earth’s surface primarily
Author's personal copy

68 Bhagirath Singh Chauhan et al.

Table 2.1 Relative contribution of greenhouse gases to anthropogenic greenhouse


effect in the United States (EIA, 2008)
Carbon equivalent
Greenhouse gases Million metric tons Percentage
Energy-related carbon dioxide 5735.5 81.3
Methane 737.4 10.5
Nitrous oxide 300.3 4.3
Carbon dioxide from other sources 103.8 1.5
Hydrofluorocarbons (HFCs), perfluorocarbons 175.6 2.5
(PFCs), and sulfur hexafluoride (SF6)

absorb most of the short-wave radiation from the sun, but also it reradiates
some of this radiation as long-wave radiation. The atmosphere is more effi-
cient at absorbing long-wave radiation, which is then emitted both back to
space and downward toward the Earth. This downward emission heats Earth
further. This further warming by reradiated long-wave radiation from the
atmosphere is known as the greenhouse effect. The amount of long-wave
radiation that is absorbed and then reradiated downward is a function of
the constituents of the atmosphere. Certain gases in the atmosphere are
particularly good at absorbing long-wave radiation and are known as GHGs.
These include water vapor, CO2, CH4, N2O, and CFCs. (Table 2.1). If the
makeup of the atmosphere changes and the result is an increase in concen-
trations of GHGs, then more of the infrared radiation from Earth will be
absorbed by the atmosphere and then reradiated back to Earth. This changes
the radiative forcing of the climate system and results in increased temper-
ature of Earth’s surface, which affects crop growth and production. Global
warming results from the increase in greenhouse effect in the atmosphere. It
is a blanket of gases that wraps around the Earth and holds the heat in. CO2 is
the most common gas that causes global warming. The more the global tem-
perature increases, the more the climate changes.

3. AGENTS OF GLOBAL WARMING


Human and industrial activities are mainly responsible for the rise in
the concentration of GHGs in the atmosphere. CO2, the most abundant
GHG, is mainly increasing because of fossil fuel combustion. Similarly,
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 69

industrial processes cause CFC emissions. The increased agricultural activ-


ities and organic waste management are presumed to be contributing to the
buildup of CH4 and N2O in the atmosphere (Hundal and Abrol, 1991). The
GHGs produced by various activities are considered as the agents of climate
change; a brief description of these agents is given in the succeeding text.

3.1. Carbon dioxide


There is a large fixation of CO2 in agriculture, but its estimates are generally
not available because of the continuous consumption of its products by
human beings and other secondary consumers. In India, fixation of CO2
is assumed to be important because almost 190 million hectares of land is
being used for farming. The estimated dry matter production from agricul-
ture in India is almost 800 million t year1 (Khan et al., 2009). This is equiv-
alent to fixation of 320 Tg of C or 1000 Tg of CO2 per annum. Only a part
of this is retained over time, while the rest is released back to the atmosphere.

3.2. Methane
The total annual output of CH4 into the atmosphere from all sources in the
world is estimated to be 535 Tg year1 (Khan et al., 2009). India’s total con-
tribution to global CH4 emissions from all sources is only 18.5 Tg year1.
The increase in annual load of CH4 in the atmosphere is much less than that
of CO2, but its higher absorption accounts for a major contribution to global
warming. Agriculture, mainly continuously flooded rice (Oryza sativa L.)
fields and ruminant animals, is the major (68%) source of CH4 emissions.
Global annual CH4 emissions from rice paddies are less than 13 Tg year1
and the contribution of Indian paddies to this is estimated to be only 4.2 Tg
year1 (Bhattacharya and Mitra, 1998; Sinha et al., 1998). Low CH4 emis-
sions from rice fields in India are mainly because the soils of the major rice-
growing areas have very low organic carbon and are not continuously
flooded.

3.3. Nitrous oxide


N2O, which is present in the atmosphere at a very low concentration
(310 ppbv), is increasing at 0.22  0.02% per year (Battle et al., 1996;
Machida et al., 1995; Mosier et al., 1998). But, in spite of its low concen-
tration and less rapid rise, N2O is becoming important because of its longer
lifetime (150 years) and greater global warming potential than CO2 (about
300 times more than CO2). Both fertilized and unfertilized soils contribute
Author's personal copy

70 Bhagirath Singh Chauhan et al.

to the release of this gas. The estimates of total N2O released from Indian
agriculture are low because of relatively low native soil fertility and lower
amounts of fertilizer used than in many developed countries (Khan
et al., 2009).
The major contributor to global warming is the energy sector (which
includes fossil fuel burning), which is responsible for 61% of the total con-
tribution toward global warming. Agriculture and its allied activities con-
tribute to 28% of global warming, followed by the industrial sector (8%),
wastes (2%), and land-use changes (1%).

3.4. Water vapor


Water vapor is the most abundant and most important GHG because of its
contribution to the natural greenhouse effect (National Research Council,
United States, 2010). Its concentration in the lower atmosphere is controlled
by the rate of evaporation and precipitation. It is not considered a climate
forcing agent.

3.5. Ozone (O3)


Ozone is found in its highest concentrations in the stratosphere that extends
from 15 to 50 km in height. It is produced by the dissociation of O2 by ultra-
violet light. It absorbs harmful ultraviolet radiation. However, the use of
aerosols, man-made halogenated gases, and CFCs has destroyed the ozone
layer in the stratosphere, contributing to global warming. The Montreal
Protocol, signed in 1987, has been ratified by 196 countries now. The ozone
layer has started to recover, but it may take decades for complete recovery
(Ebi et al., 2008). Near the Earth’s surface, however, ozone is regarded as a
pollutant, causing damage to plants, animals, and humans, and it is the main
component of smog (National Research Council, United States, 2010).

4. EVIDENCE FOR CLIMATE CHANGE AND IMPACTS


ON AGRICULTURE
During the past few decades, climate change, caused by global
warming, has received worldwide attention (Baer and Risbey, 2009).
The most significant change is the rise in the atmospheric temperature cau-
sed by increased concentrations of GHGs in the atmosphere. Between 1000
and 1750 AD, CO2, CH4, and N2O concentrations were 280  6 ppm,
700  60 ppb, and 270  10 ppb, respectively (IPCC, 2007a). More
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 71

recently, however, these values have increased to 369 ppm, 1750 ppb, and
316 ppb, respectively.
The global mean annual temperature at the end of the twentieth century
was almost 0.7  C above that recorded at the end of the nineteenth century
and it is likely to increase further by 1.8–6.4  C by AD 2100, with a best
estimate of 1.8–4.0  C (IPCC, 2001, 2007a,b). The decade 1990–2000
was the warmest in the last 300 years and was 0.5  C warmer than the mean
temperature of 1961–1990. Warmer summers have included record hot
spells and high sunshine hours, and the warm winters have reduced the
number of frosts. The quantity of rainfall and its distribution are also greatly
affected by climate change and these are expected to increase the problems of
flooding and soil erosion. Moreover, the sea level has risen and snow cover is
also gradually decreasing due to glacier meltdown, especially near the poles,
and arable land is decreasing near coastal regions due to inundation.
The Panel on Advancing the Science of Climate Change, set up by the
National Research Council (United States) (2010), arrived at several conclu-
sions, whose important ones were (1) that several different research groups
had shown that the planet’s mean temperature was 0.8  C higher during the
first decade of the twenty-first century than during the first decade of the
twentieth century; (2) that much of the warming during the past several
decades could be attributed to anthropogenic activities that released into
the atmosphere CO2 and other GHGs that trap heat and burning of fossil
fuels (coal, oil, and natural gas) for energy was the largest contributor to cli-
mate change; and (3) that agriculture, forest clearing, and certain industrial
activities also made significant contributions to climate change.
The United Nations Environment Programme (UNEP) and the WMO
established the IPCC in 1988 to periodically assess the state of the global
environment. A report of the IPCC during 2001 projected that the global
mean temperature above Earth’s surface would rise 1.4–5.8  C during the
next 100 years (IPCC, 2001).
Climate change is a well-recognized, significant, man-made, global
environmental challenge and agriculture is strongly influenced by it
(Hillel and Rosenzweig, 2011; Kang and Banga, 2013). A team of experts
from the Food and Agriculture Organization (FAO) concluded that each
1  C rise in mean temperature would cause annual wheat (Triticum
aestivum L.) yield losses in India of about 6 million tons (US$ 1.5 billion
at current prices), and, when losses of all other crops were taken into con-
sideration, farmers were projected to lose US$ 20 billion each year (FAO,
Food and Agriculture Organization, 2008; Swaminathan, 2012). Such losses
Author's personal copy

72 Bhagirath Singh Chauhan et al.

can happen as temperature can reduce crop duration; increase crop-


respiration rates; alter photosynthate movement from source to sink; affect
the survival and distribution of pest populations, thus developing a new
equilibrium between crops and pests; hasten nutrient mineralization in soils;
decrease fertilizer-use efficiencies; and increase evaporation (Kumar et al.,
2011). Indirectly, there may be considerable effects on land use that can
be attributed to snowmelt, availability of irrigation, frequency and intensity
of inter- and intraseasonal droughts and floods, and availability of energy
(Sharma and Chauhan, 2011). All of these changes can have tremendous
effects on agricultural production and hence on the food security of a region.
More flooding, droughts, and forest fires; decreases in agricultural pro-
ductivity; and the displacement of thousands of coastal residents by sea-level
rise and intense tropical cyclones are the likely consequences of climate
change in Asia. A rise in mean temperature of 2  C above the normal could
mean that small islands, such as Tuvalu in the Pacific Ocean and Maldives,
Lakshadweep, and Andaman and Nicobar in the Indian Ocean, could be
submerged (Swaminathan, 2012).
Global warming is mainly brought about by rapid industrialization, com-
bustion of fossil fuels, increased agricultural operations, deforestation, and the
increased number of vehicles (National Research Council, United States,
2010). The driving force behind these factors is the ever-increasing human
population. The global share of various countries in CO2 emissions is given
in Table 2.2 (Sathaye et al., 2006), and per capita emissions of C by residents
of various countries are shown in Table 2.3 (Sathaye et al., 2006). The United
States is the major contributor to CO2 emission; the US share was about one-
fourth of the total global emissions, whereas India’s contribution was about 4%
(Table 2.2). Per capita production of C was also highest in the United States
(5.4 t of C per person), followed by Canada (5.2 t of C per person) (Table 2.3).
India’s per capita C emissions were only 0.26 t of C per person.

5. PROJECTED CLIMATE CHANGE IN INDIA


Analysis of a representative rainfall series across the past 176 years for
India as a whole did not suggest any significant trend in rainfall change
(Sontakke, 1990). However, Rao (2007) analyzed the rainfall data of
1140 meteorologic stations in India, which showed a negative trend in rain-
fall among the stations situated in the southern states of India, southern pen-
insular areas, central India, and parts of the north and northeastern regions.
Positive trends in rainfall were observed for Gujarat, Maharashtra, coastal
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 73

Table 2.2 Global share of some countries in CO2 emissions (Sathaye et al., 2006)
CO2 emissions (%)
Country 1990 2003
Canada 2.19 2.39
China 10.41 14.07
France 1.80 1.63
Germany 4.24 3.35
India 2.63 4.07
Italy 1.91 1.85
Japan 5.54 4.79
Russia 9.67 6.38
United Kingdom 2.76 2.24
United States 23.04 23.06
Rest of the world 38.61 36.17

Table 2.3 Per capita carbon emissions from energy for the year 2003
(Sathaye et al., 2006)
Country Carbon emissions (t person1)
Canada 5.19
China 0.78
France 1.86
Germany 2.78
India 0.26
Italy 2.21
Japan 2.58
Russia 3.06
United Kingdom 2.60
United States 5.44
Author's personal copy

74 Bhagirath Singh Chauhan et al.

Andhra Pradesh, and Odisha. However, the parts of the country covering
eastern Uttar Pradesh, eastern Madhya Pradesh, the west coast, and greater
parts of northwest India did not show any changes. Among the rainfed dis-
tricts, 40% of the stations showed a negative trend, 48% showed a positive
trend, and 12% showed no changes in rainfall.
Based on observations from 73 stations, an analysis of the mean annual
surface air temperature across India for 1901–1988 showed a significant
warming of about 0.4  C per 100 years (Hingane et al., 1985). This warming
trend was comparable to the global mean temperature change of 0.5  C in
the last 100 years. Later on, using the all-India mean surface air temperature
for 1901–2000 from a network of 31 well-distributed, representative sta-
tions, the trends in mean annual temperatures across the country were deter-
mined (Rupa Kumar et al., 2002). Warming trends were observed during
four seasons (winter, premonsoon, monsoon, and postmonsoon) with a
higher rate of temperature increase during winter (0.04  C per decade)
and postmonsoon seasons (0.05  C per decade) compared with the
premonsoon (0.02  C per decade) and monsoon seasons (0.01  C per
decade). The warming across the Indian subcontinent was mainly contrib-
uted by the postmonsoon and winter seasons. The monsoon temperatures
did not show a significant trend in most parts of India, except for a significant
negative trend across northwest India (De and Mukhopadhyay, 1998). The
diurnal temperature range has also decreased, with nighttime temperature
increasing at twice the rate of the daytime maximum temperature (Sen
Roy and Balling, 2005).
In a regional study in Punjab, Prabhjyot-Kaur and Hundal (2010)
reported gradual increases in minimum temperature across a recent
30-year period. The maximum temperature, however, showed no signifi-
cant trend at most locations in the state. A 5-year mean of annual rainfall-
variability analysis in Punjab revealed that rainfall decreased significantly
during the past five decades at the rate of 5.5, 3.4, 7.1, 4.3, and 5.5 mm
year1 at Amritsar, Gurdaspur, Ferozepur, Bathinda, and Sangrur, respec-
tively, and rainfall increased significantly at 1.4 mm year1 during the past
108 years at Ludhiana (Prabhjyot-Kaur et al., 2011). However, no significant
trend in rainfall variability was observed in Kapurthala, Jalandhar,
Hoshiarpur, Rupnagar, Patiala, and Faridkot districts. The IPCC compiled
data on the magnitude of change in temperature, rainfall, and CO2 for dif-
ferent parts of the world, according to which CO2 was expected to increase
to 397–416 ppm by 2050 and to 605–755 ppm by 2070 (Watson et al.,
1998). By the end of the twenty-first century, there could be a change in
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 75

Table 2.4 Changes in temperature and rainfall projected for India (Lal et al., 2001)
Temperature change ( C) Rainfall change (%)
Year Season Lowest Highest Lowest Highest
2020 Annual 1.00 1.41 2.16 5.97
Rabi 1.08 1.54 1.95 4.36
Kharif 0.87 1.17 1.81 5.10
2050 Annual 2.23 2.87 5.36 9.34
Rabi 2.54 3.18 9.22 3.82
Kharif 1.81 2.37 7.18 10.52
2080 Annual 3.53 5.55 7.48 9.90
Rabi 4.14 6.31 24.83 4.50
Kharif 2.91 4.62 10.10 15.18

precipitation of 5–25% across India, with more reductions in winter than


summer rainfall (Prabhakar and Shaw, 2008).
There is considerable uncertainty in the magnitude of change in rainfall
and temperature predicted for India. The relative increase in temperature has
been projected to be less in kharif than in the rabi season; rabi rainfall would
have a large uncertainty, but kharif rainfall is likely to increase by as much as
10% (Lal et al., 2001) (Table 2.4). Because of the large spatial and temporal
variability in weather factors in a region, the availability of more detailed
scenarios for different agroclimatic zones is desirable. There is also increasing
consensus that climatic variability will increase in the future, leading to more
frequent extremes of weather in the form of erratic monsoons and increased
frequency and intensity of drought and flooding.

6. IMPACT OF GLOBAL WARMING ON AGRICULTURE


AND THE FOOD SUPPLY
Agriculture is both a victim and an abettor of climate change (Kang
and Banga, 2013). Climate change plays a significant role in a nation’s food
security and economy, especially in a developing country such as India. For
example, Killman (2008) wrote (p. iii), “Climate change will affect all four
dimensions of food security: food availability, food accessibility, food utili-
zation and food systems stability. It will have an impact on human health,
livelihood assets, food production and distribution channels, as well as
Author's personal copy

76 Bhagirath Singh Chauhan et al.

changing purchasing power and market flows.” All agricultural commodities


are sensitive to climate change or climate variability (Kumar et al., 2011).
The rising temperature and CO2 and uncertainties in rainfall associated
with global climate change have serious direct and indirect consequences for
crop production and food security (Sinha and Swaminathan, 1991). It is
therefore important to evaluate the direct and indirect consequences of
global warming on different crops contributing to food security. Future agri-
cultural strategies thus have to formulate a holistic approach in the coming
decades on productivity, sustainability, profitability, stability, and equity in
Indian agriculture. Dar and Gowda (2013) suggested that improved crop,
soil, and water management practices and stress-tolerant varieties should
overcome the detrimental impacts of climate change, which in turn would
lead to improved food security, livelihoods, and environmental security.
They further pointed out that negative effects of climate change on food
security can be counteracted by broad-based economic growth, particularly
improved agricultural productivity, and robust international trade in agricul-
tural products that can offset regional shortages and agricultural productivity
investments.
The effects of changes in temperature, CO2 concentrations, and precip-
itation on crop productivity have been studied broadly using crop simulation
models (Parry et al., 2004). The combined effects of climate change may
have implications for dryland and irrigated crop yields. However, the effect
on production is expected to vary by crop and location and by the magni-
tude of warming and the direction and magnitude of precipitation change
(Adams et al., 1998). Projections by the IPCC and a few other global studies
are that, unless we adapt, there is a probability of a 10–40% loss in crop pro-
duction in India by 2080–2100 as a result of global warming (IPCC, 2007a;
Parry et al., 2004; Rosenzweig and Parry, 1994) despite the beneficial aspects
of increased CO2 concentrations (Kumar et al., 2011).

6.1. Effect of elevated concentrations of CO2 on crop growth


CO2 is essential for photosynthesis and hence for plant growth. An increase
in atmospheric CO2 concentration affects crop production through altering
photosynthetic and transpiration rates. It is therefore important to assess the
combined effects of elevated atmospheric CO2 concentration and climate
change on the productivity of a region’s dominant crops (Haskett et al.,
1997). The direct effects of increased concentrations of CO2 are normally
beneficial to vegetation, especially for C3 plants, as elevated concentrations
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 77

enhance assimilation rates and increase stomatal resistance, which result in a


decline in transpiration and improved water-use efficiency in crops
(Farquhar, 1997).
Several simulation studies have been carried out to predict the likely
effects of elevated CO2 on crop yield. In northwestern India, for example,
yields of rice and wheat increased by 15% and 28%, respectively, at elevated
(doubled) CO2 concentrations (Lal et al., 1998). Similarly, Hundal and
Prabhjyot-Kaur (2007) reported a gradual increase in rice and wheat grain
yield from elevated CO2. The effects of elevated CO2 on simulated grain
yield of wheat under optimal and suboptimal (stressed) moisture conditions
are presented in Table 2.5 (adapted from Pandey et al., 2007). With the grad-
ual increase in CO2 concentration from 440 to 660 ppm, yield increased
from 21% to 68% under optimal conditions, whereas, under suboptimal
conditions, similar responses were observed with slightly lower magnitudes
(19–57%). Thus, under climate change, CO2 enhancement may increase
crop productivity. Hundal and Prabhjyot-Kaur (1996) reported that, if all
other climatic variables were kept constant (normal), leaf area index
(LAI), biomass yield, and grain yield of rice increased with elevated CO2.
With an increase in CO2 from 330 to 600 ppm, an increase as high as
11% in LAI was reported, in addition to an 8% increase in biomass yield
and 9% in grain yield (Hundal and Prabhjyot-Kaur, 1996) (Table 2.6).
Rice grown under elevated CO2 had significantly more grains and grain
yield per unit land area than under ambient CO2 and open-field conditions
(De Costa et al., 2006) (Table 2.7). In the open-field treatment, rice was
grown under normal atmospheric conditions to detect whether the presence
of a chamber had any effect on the crop. The percentage of filled grains was
also significantly higher under the elevated CO2 concentration than under

Table 2.5 Change in simulated wheat yield due to varying CO2 concentration under
optimal and suboptimal (stressed) moisture conditions (Pandey et al., 2007)
Change (%) from
Simulated grain yield base suboptimal and
(kg ha1) optimal yield
CO2 concentration (ppm) Suboptimal Optimal Suboptimal Optimal
330 (base value) 3112 3837 – –
440 3695 4630 19 21
550 4327 5687 39 48
660 4876 6465 57 68
Author's personal copy

78 Bhagirath Singh Chauhan et al.

Table 2.6 Effect of CO2 increase on growth and yield of rice (Hundal and Prabhjyot-
Kaur, 1996)
Deviation from normal (%)
Parameters 330 ppm (normal) 400 ppm 500 ppm 600 ppm
Maximum leaf area index 5.22 þ 1.9 þ 8.5 þ 11.1
1
Biomass yield (kg ha ) 12495 þ 1.1 þ 6.1 þ 7.7
Grain yield (kg ha1) 7563 þ 1.5 þ 6.6 þ 8.7

Table 2.7 Number of grains, percent of filled grains, grain weight, grain yield, and
harvest index of rice grown under elevated (570 ppm) and ambient (370 ppm) CO2 in
open-top chambers and in open-field conditions (De Costa et al., 2006)
Treatments
Parameters Elevated CO2 Open-field Ambient CO2
2
Grains (no. m ) 47007 (11) 41793 42336
Filled grains (%) 82.9 (9) 72.1 76.0
Grain weight (mg) 24.9 (2) 24.0 24.5
2
Grain yield (g m ) 871 (24) 723 783
Harvest index 0.47 0.41 0.45
The values in parentheses represent percentage increase over ambient CO2 concentration.

the ambient CO2 concentration. Individual grain weight and harvest index,
however, did not differ significantly between the elevated and ambient CO2
treatments. Grain yield under elevated CO2 was 24% higher than under
ambient CO2 concentration. Panicle dry weight in the elevated CO2 treat-
ment was significantly higher than that under the ambient CO2 and open-
field treatments throughout the grain-filling period (De Costa et al., 2006)
(Table 2.8). This was attributable to the crop’s significantly higher panicle
growth rate during the early grain-filling period, that is, 54–67 days after
transplanting. The partitioning coefficient in the elevated CO2 treatment
did not exceed that of the ambient and open-field treatments during the
early and late grain-filling periods. In cotton (Gossypium spp.), an increase
in CO2 from subambient to ambient and then to elevated concentrations
resulted in a significant increase in total dry matter production by cotton
plants (Reddy et al., 2004) (Table 2.9). This response was mainly attributable
to increased boll dry weight and lint dry weight per boll.
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 79

Table 2.8 Variation of partitioning coefficient and panicle dry weight at different times
after rice transplanting under elevated (570 ppm) and ambient (370 ppm) CO2 in open-
top chambers and in open-field conditions (De Costa et al., 2006)
Days after transplanting Elevated CO2 Ambient CO2 Open-field
Partitioning coefficient
54 0.14a 0.12b 0.15a
67 0.30a 0.33a 0.34a
94 0.49b 0.51a 0.47b
2
Panicle dry weight (g m )
54 165a 118b 167a
67 693a 428b 418b
94 1036a 884b 833b
Partitioning coefficient was calculated as ratio between panicle dry weight and total dry weight.
For each time (days after transplanting), horizontal means with the same letters are not significantly dif-
ferent at p ¼ 0.05.

Table 2.9 Effect of different CO2 concentrations on total dry matter, boll dry weight, lint
dry weight, and seed dry weight of cotton (Reddy et al., 2004)
CO2 Total dry weight Boll dry Lint dry weight Seed dry weight
concentration (g plant1) weight (g) (g boll1) (g boll1)
Subambient 165c 5.6b 1.8b 2.7b
(180 ppm)
Ambient 233b 5.8a 1.8ab 2.8ab
(360 ppm)
Elevated 309a 5.9a 1.8a 2.9a
(720 ppm)
Means with the same letters within the same column are not significantly different at p ¼ 0.05.

The growth and yield response of black gram (Vigna mungo) to CO2 con-
centrations (550 and 700 ppm) was investigated and compared with ambient
CO2 concentration (365 ppm) using open-top chambers (Vanaja et al.,
2007). The growth parameters (root and shoot length, leaf area, and root
volume) were significantly greater at 700 ppm CO2 than at 550 ppm. Com-
pared to the ambient (chamber) control, the increase in total biomass at 700
and 550 ppm CO2 was 65% and 39%, respectively (Table 2.10). The
increase in seed yield at 700 ppm (129%) was attributable to an increase
Author's personal copy

80 Bhagirath Singh Chauhan et al.

Table 2.10 Yield parameters of black gram (of 10 plants) under 365, 550, and 700 ppm
CO2 (Vanaja et al., 2007)
Parameter values Increase (%)
550 vs. 700 vs. 700 vs.
Parameters 365 ppm 550 ppm 700 ppm 365 ppm 365 ppm 550 ppm
Pods (#) 158 187 239 18 51 28
Pod weight (g) 35.6 66.5 78.2 87 120 18
Seed weight (g) 17.4 32.9 39.9 89 129 21
100-seed weight (g) 2.6 2.7 3.9 2 51 48
Harvest index (%) 28.5 38.7 39.5 36 38 2
Total biomass (g) 61.1 84.9 101.1 39 65 19

in the number of pods per plant and 100-seed weight, whereas the increase
in total seed yield at 550 ppm (89%) was caused by a higher number of pods
per plant and seeds per pod. The harvest index, a very important parameter
in pulses for breaking the yield barrier, increased to 36% and 38% at 550 and
700 ppm, respectively (Table 2.10).
In rice, mean biomass increment, leaf area duration, and net assimilation
rate increased with increasing CO2 concentrations (Baker et al., 1990)
(Table 2.11). In the same study, net assimilation rate decreased and leaf area
duration increased in rice with the progression of growth stages. Grain yield
increased by nearly 32% when the CO2 concentration increased from 330 to
660 mmol CO2 mol1 air (Baker et al., 1990; Table 2.12). The number of
panicles per plant was mainly responsible for the observed differences in
grain yield among the CO2 concentrations (Table 2.12). The number of
filled grains per panicle was the most variable yield component, whereas
individual grain weight was stable across different CO2 concentrations.
Therefore, it was concluded that grain yield depended mainly on the num-
ber of panicles on rice plants.

6.2. Effect of ozone on plants


Ozone is likely to have adverse effects on plant growth. Necrotrophic
pathogens can colonize plants that are weakened by O3 at an accelerated
rate, while obligate biotroph infections might be reduced (Manning, 1995).
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 81

Table 2.11 Effect of CO2 enrichment on mean biomass increment (DW), leaf area
duration (LAD), and net assimilation rate (NAR) in rice in controlled-environment
chambers (Baker et al., 1990)
NAR
LAD (g m2 LAD NAR
DW (g) (m day ) day1) DW (g) (m2 day1) (g m2 day1)
2 1
CO2 concentration
(ppm) 19–44 days after sowing 44–71 days after sowing
160 1.2 0.37 4.9 1.3 1.0 1.4
250 2.1 0.50 6.0 2.2 1.2 1.8
330 2.5 0.55 6.1 2.7 1.2 2.2
500 2.5 0.51 6.5 3.9 1.2 3.2
660 3.2 0.65 6.7 2.0 1.4 1.5
900 3.9 0.76 7.1 3.8 1.6 2.4
Standard error 0.4 0.07 1.1 1.2 0.1 0.9

Table 2.12 Mean grain yield and components of yield of rice in controlled-environment
chambers (Baker et al., 1990)
CO2 concentration Grain yield Panicles Filled grains 1000-grain
(ppm) (g plant1) (no. plant1) (no. panicle1) weight (g)
160 1.4 3.6 22 17.4
250 1.3 4.6 17 18.2
330 1.9 5.4 19 17.9
500 3.0 7.3 23 18.1
660 2.8 6.0 25 18.4
900 3.3 6.4 28 18.1
Standard error 0.6 0.9 5.5 0.83

6.3. Effect of increasing temperature on crop growth


Increases in temperature increase crop-respiration rates; reduce crop dura-
tion, the number of grains formed, and crop yield; inhibit sucrose assimila-
tion in grains; affect the survival and distribution of pest populations; hasten
nutrient mineralization in soil; decrease fertilizer-use efficiency; and increase
evaporation. In a simulation study, an increase in temperature by 2  C
Author's personal copy

82 Bhagirath Singh Chauhan et al.

brought about a 3–10% decrease in grain/seed yield of kharif crops, such as


rice, groundnut (Arachis hypogaea L.), and soybean (Glycine max L.), and a
29% decrease in grain yield of rabi crops such as wheat (Prabhjyot-Kaur
and Hundal, 2006). Pandey et al. (2007) simulated grain yield of wheat
under incremental units of maximum temperature (1–3  C) using the
CERES-wheat model and found a gradual decrease in yield from 3546 to
2646 kg ha1 (8% to 31% less than the base yield) under optimal moisture
conditions (Table 2.13). Similarly, under suboptimal conditions, yield
declined from 2841 to 2398 kg ha1 (9% to 23% less than the base yield).
The reduction in wheat yield with an increase in maximum temperature
was mainly attributable to a reduction in the duration of anthesis and in grain
filling with a rise in ambient temperature, and vice versa (Aggarwal and
Kalra, 1994).
An increase in temperature from the greenhouse effect would decrease
cereal and groundnut production; however, the impact of changes in tem-
perature would vary with the type of crop and the direction of the change
occurring. For example, if all other climatic variables were kept constant, a
temperature increase of 0.5, 1.0, 2.0, and 3.0  C compared with the nor-
mal temperature would advance the maturity of wheat by 3, 6, 12, and
17 days, respectively (Prabhjyot-Kaur and Hundal, 2010) (Table 2.14).
On the other hand, with a temperature rise of up to 1  C above normal,
heading of rice was not affected, but a further increase in temperature to
3  C prolonged heading and maturity by 4 and 5 days, respectively, com-
pared with the normal temperature. Flowering in soybean was delayed up

Table 2.13 Change in simulated wheat yield due to varying temperature under optimal
and suboptimal moisture conditions (Pandey et al., 2007)
Change in simulated
Simulated grain yield grain yield relative to
(kg ha1) base yield (%)
Change in maximum temperature

relative to base temperature ( C) Suboptimal Optimal Suboptimal Optimal
þ3 2398 2646 23 31
þ2 2668 3091 14 19
þ1 2841 3546 9 8
1 3190 4206 3 10
2 3358 4485 8 17
3 3641 4817 17 26
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 83

to 4 days and its maturity was delayed by 2 days (Table 2.14). Deviation
(increase) in temperature from the normal temperature greatly influenced
flowering and maturity in gram. With an increase in temperature of 3  C
compared with the normal temperature, for example, flowering and matu-
rity in chickpea (Cicer arietinum L.) advanced by 23 and 24 days, respec-
tively (Table 2.14). A study analyzed the relationship between the yield
of rice and minimum temperature across the range of 22.1–23.7  C using
a quadratic equation and reported that rice yield declined by 10% with each
1  C rise in minimum temperature and yield declined by 15% with each
1  C rise in mean temperature (Peng et al., 2004).
Reddy et al. (1992) reported that cotton plants grew faster at 30/22  C
(maximum/minimum temperatures) than at either higher or lower temper-
atures. However, the plants at 35/27  C had more boll weight than those
grown at 30/22  C and they were more advanced in fruiting-structure for-
mation (Table 2.15). The time required to produce the first square was only
2 days longer at 40/32  C than at 30/22  C. At 20/12 and 25/17  C, squares
abscised at a very early stage, whereas the maximum number of squares was

Table 2.14 Effect of temperature increase on phenology of crops (Prabhjyot-Kaur and


Hundal, 2010)
Deviation from normal temperature (days)
Crop and phenological stage Normala þ0.5  C þ1.0  C þ2.0  C þ3.0  C
Rice
Heading 223 0 0 þ1 þ4
Maturity 263 þ1 þ1 þ1 þ5
Wheat
Anthesis 41 3 6 12 16
Maturity 82 3 6 12 17
Soybean
Flowering 239 þ1 þ2 þ3 þ4
Maturity 294 þ1 þ1 þ2 þ2
Chickpea
Flowering 08 4 7 19 23
Maturity 99 5 8 16 24
a
Julian day (calendar day).
Author's personal copy

84 Bhagirath Singh Chauhan et al.

Table 2.15 Effect of temperature on days to appearance of first flower bud (square),
days to first flower, and biomass of different parts of cotton seedlings at 56 days after
emergence (Reddy et al., 1992)
Day/night temperature ( C)
Parameters 20/12 25/17 30/22 35/27 40/32
Days to first square 38 33 27 24 29
(day)
a a b
Days to first flower 53 43
(day)
Stem biomass 3.6  0.3 18.0  1.9 33.9  3.7 31.1 17.2  2.1
(g plant1) 11.2
Leaf biomass 9.0  0.7 22.7  2.0 33.7  2.0 31.5  9.6 19.8  2.1
(g plant1)
Root biomass 1.3 2.9 6.5 6.2 4.5
(g plant1)
Boll biomass a a
1.3  0.4 4.1  0.6 b

(g plant1)
Square biomass 0.04  0.01 0.61  0.1 2.27  0.3 2.70  0.3 b

(g plant1)
Total biomass 13.5  0.9 44.2  4.0 77.7  6.9 75.3 41.5  4.2
(g plant1) 21.7
a
Squares were abscised at very early stage.
b
Flowers were not formed at final harvest due to slow growth.

obtained at 30/22  C. Fruit branches were four times greater at 30/22  C


than at 20/12  C, whereas maximum vegetative branches were produced
at low temperatures (20/12  C). Bolls and squares were produced at 30/
22  C, whereas a 10% boll and square loss was observed at 35/27  C during
the early reproductive period.
Hundal and Prabhjyot-Kaur (2007) studied the effect of temperature on
the growth and yield of rice and wheat. An increase in temperature
decreased the growth and yield of rice and wheat, but a decrease in temper-
ature increased their growth and yield. Both the decrease in yield and the
increase in yield were more for wheat than for rice. Compared with normal
conditions, a temperature increase of 1.0–2.0  C caused a decrease of 4–9%
in the maximum LAI in rice (Fig. 2.1) and a decrease of 18–29% in wheat
(Fig. 2.2). Similarly, biomass yield decreased by 2–5% in rice (Fig. 2.1) and
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 85

30
Deviation in LAI and yield from normal (%)
25
Grain yield
Biomass yield
20
Max. LAI

15

10

0
−3 −1 Normal 1 3
−5

−10

−15
Deviation in temperature from normal (°C)
Figure 2.1 Effect of temperature change on growth and yield of rice using CERES-wheat
model (Hundal and Prabhjyot-Kaur, 2007).

50
Deviation in LAI and yield from normal (%)

40
Grain yield
30 Biomass yield
Max. LAI
20

10

0
−3 −1 Normal 1 3
−10

−20

−30

−40

−50
Deviation in temperature from normal (°C)
Figure 2.2 Effect of temperature change on growth and yield of wheat using CERES-
wheat model (Hundal and Prabhjyot-Kaur, 2007).
Author's personal copy

86 Bhagirath Singh Chauhan et al.

by 14–23% in wheat (Fig. 2.2), and grain yield decreased by 3–10% in rice
(Fig. 2.1) and by 10–18% in wheat (Fig. 2.2). A decrease in temperature by
1.0–2.0  C increased the simulated maximum LAI by 3–5% in rice and by
12–28% in wheat, biomass yield increased by 4–10% in rice and by 9–16% in
wheat, and grain yield increased by 8–15% in rice and by 7% in wheat vis-à-
vis the normal conditions (Figs. 2.1 and 2.2).
A simulation study was conducted using the CERES-wheat model to
assess the effect of an intraseasonal increase in temperature on the yield of
wheat sown on different dates (Prabhjyot-Kaur and Hundal, 2010). The
simulation results revealed that, in general, an increase in temperature from
February to mid-March severely affected the yield of early-, normal-, and
late-sown wheat (Table 2.16). The temperature increase mostly affected
the yield of the early (October)-sown crop beginning in the fourth week
of January to the first fortnight of March, the timely (November)-sown crop
during February and March, the late (fourth week of November)-sown
crop during March, and the very late (December)-sown crop during
March and the first week of April (Table 2.16). A maximum of about a
17% decrease in grain yield occurred in the early-sown crop if the temper-
ature increased by 6  C in the first fortnight of February. This was mainly
because of a reduction in grain-filling period caused by an increase in
temperature.

6.4. Interactive effects of changing climatic factors on crop


production
The ultimate productivity of crops is determined by the interactions of cul-
tivars, soil constituents, water, temperature, day length, etc. Temperature,
solar radiation, and water directly affect the physiological processes involved
in grain development and indirectly affect grain yield by influencing the
incidence of diseases and insects (Yoshida and Parao, 1976). Rice grain yield
was positively correlated with mean solar radiation and negatively correlated
with daily mean temperature during the reproductive stage (Yoshida and
Parao, 1976). Relatively low temperature and high solar radiation during
the reproductive stage had a positive effect on the number of spikelets
and hence increased grain yield. Solar radiation had a positive influence
on grain filling during the ripening period.
The simulation results indicated that warm climate with decreasing radi-
ation would affect the growth and yield of cereal crops. However, the harm-
ful effects of increasing temperature on growth and yield could be
counterbalanced to some extent by the increasing CO2 concentrations
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 87

Table 2.16 Effect of intraseasonal temperature increase (maximum and minimum) from
normal on grain yield (% deviation from normal) of wheat sown on different dates using
the CERES-wheat model (Prabhjyot-Kaur and Hundal, 2010)
Temperature increase from normal ( C)
Time period Date of sowing þ1.0 þ2.0 þ3.0 þ4.0 þ5.0 þ6.0
First fortnight of Early sown 3.4 3.7 7.6 11.5 13.0 17.2
February (October 28)
Normal sown þ1.7 1.6 1.8 3.9 7.7 7.3
(November 8)
Normal sown 0.5 2.7 1.5 2.0 1.3 1.9
(November 15)
Normal sown þ0.5 þ2.4 þ2.8 þ4.7 þ4.9 þ7.0
(November 25)
Late sown þ0.7 þ0.6 þ0.6 þ3.4 þ3.6 þ3.7
(December 2)
Second fortnight Early sown 2.4 2.8 5.2 8.1 10.9 13.8
of February (October 28)
Normal sown 0.4 4.1 5.1 9.9 14.2 16.4
(November 8)
Normal sown 2.0 5.8 6.0 8.7 9.7 14.2
(November 15)
Normal sown þ2.5 þ1.1 þ3.4 0.6 2.6 3.3
(November 25)
Late sown 0.5 0.4 1.7 2.3 3.1 3.6
(December 2)
First fortnight of Early sown 2.3 4.6 6.8 13.8 8.2 10.4
March (October 28)
Normal sown 2.7 3.3 6.0 9.5 9.5 13.0
(November 8)
Normal sown 4.8 9.3 10.1 14.2 16.0 20.8
(November 15)
Normal sown 0.5 5.4 6.7 3.3 16.0 19.4
(November 25)
Late sown 2.3 1.6 6.8 7.6 12.5 17.7
(December 2)
Continued
Author's personal copy

88 Bhagirath Singh Chauhan et al.

Table 2.16 Effect of intraseasonal temperature increase (maximum and minimum) from
normal on grain yield (% deviation from normal) of wheat sown on different dates using
the CERES-wheat model (Prabhjyot-Kaur and Hundal, 2010)—cont'd
Temperature increase from normal ( C)
Time period Date of sowing þ1.0 þ2.0 þ3.0 þ4.0 þ5.0 þ6.0
Second fortnight Normal sown þ1.1 1.5 0.5 0.1 1.9 1.5
of March (November 8)
Normal sown 2.5 1.6 4.3 6.9 5.9 8.1
(November 15)
Normal sown 0.1 4.7 5.6 9.2 10.1 11.2
(November 25)
Late sown 5.5 6.6 12.3 14.5 19.1 21.4
(December 2)

(Hundal and Prabhjyot-Kaur, 1996). A study reported that in India, the


adverse effects of a 1–2  C rise in temperature could be absorbed with a
5–10% increase in precipitation (Abrol et al., 1991). A grain yield increase
of 20–30% might be possible on about 70% of the area in the rice–wheat
cropping system in India. In northern India, warming could offset some
losses in yield by early pod setting in winter grain legumes, such as chickpea
and lentil.
Mahi (1996) found that the maximum LAI, biomass, and grain yield of
wheat and rice declined when radiation decreased by 10% relative to normal
radiation but increased when radiation was enhanced by 10%. The simula-
tion results suggested that the growth and yield of wheat and rice would
be influenced by increasing temperature. The adverse effects generated
by a high-temperature scenario might be lessened to some extent by a
decrease in radiation amounts. In Punjab, India, there are indications that
the amount of radiation is likely to decrease. As a result, the production
of wheat and rice could be adversely affected, depending upon the degree
of change in radiation amount in the coming years. The past increase in
CO2 experienced to date and the projections of its increase in the future will
no doubt counterbalance the negative effects of a rise in temperature on crop
productivity.
Increased CO2 concentrations could result in greater growth and grain
yield of rice and compensate for the yield reductions caused by warmer tem-
peratures (Hundal and Prabhjyot-Kaur, 1996). Under all the scenarios of
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 89

increased CO2 concentrations (i.e., 400, 500, and 600 ppm), maximum
LAI, biomass, and grain yield of rice were favorably affected (Tables 2.17
and 2.18). The interactive effect of enhanced temperature and CO2 revealed
that adverse effects caused by an increase in temperature of up to 0.5  C
could be compensated for by concentrations of CO2 above 400 ppm
(Hundal and Prabhjyot-Kaur, 1996). A further increase in temperature of
up to 1.0  C did not decrease maximum LAI, biomass, and grain yield when
CO2 was nearly doubled (600 ppm) relative to the normal temperature. But,
in scenarios with nearly doubled CO2 concentrations of 600 ppm, temper-
ature increases of more than 1.0  C above normal reduced the maximum
LAI, biomass, and grain yield of rice (Tables 2.17 and 2.18).

Table 2.17 Effect of CO2 and temperature on the deviation of leaf area index of rice
(Hundal and Prabhjyot-Kaur, 1996)
Deviation of leaf area index from normal (%)
Temperature ( C) 330 ppm (normal) 400 ppm 500 ppm 600 ppm
Normal [5.22]a þ1.9 þ8.5 þ11.0
þ 0.5 5.5 1.9 þ2.5 þ6.6
þ 1.0 9.3 6.1 4.0 þ1.7
þ 1.5 9.8 9.1 5.7 1.7
þ 2.0 12.3 11.9 7.8 5.5
a
Leaf area index at normal CO2 concentration and temperature.

Table 2.18 Effect of CO2 and temperature on the deviation of rice biomass yield from
normal (Hundal and Prabhjyot-Kaur, 1996)
Deviation of rice yield from normal (%)
Temperature ( C) 330 ppm (normal) 400 ppm 500 ppm 600 ppm
Normal [12495] a
þ1.1 þ6.1 þ7.7
þ 0.5 3.5 1.4 þ2.2 þ4.5
þ 1.0 6.0 1.4 þ2.2 þ4.5
þ 1.5 7.2 6.8 5.0 2.0
þ 2.0 7.3 7.1 4.0 2.6
a 1
Biomass yield (kg ha ) at normal CO2 and temperature.
Author's personal copy

90 Bhagirath Singh Chauhan et al.

Das et al. (2007) conducted a preliminary study to test crop simulation


model ORYZA 2000. The model has been used to investigate the impact of
climate change (with changing temperature and CO2 concentrations) on
rice yields. The model predicted direct changes in yield by 10%, 46%,
and 72% for temperature changes of þ1, þ2, and þ3  C, respectively.
The data demonstrated that even with an up to þ1  C warmer climate than
the normal, production might increase by about 10% in an atmosphere with
doubled CO2 concentration. However, a further increase in temperature
could negate the effect of increased CO2 concentration.
The results of the simulation study for the interactive effects of increasing
temperature and CO2 concentrations revealed that the adverse effects of
increased temperature on the growth and yield of rice were counterbalanced
to some extent by the favorable effect of increasing CO2 (Hundal and
Prabhjyot-Kaur, 2007) (Table 2.19). Under enhanced CO2 concentration
of 600 ppm, a temperature increase of 2  C compared with normal reduced
maximum LAI by 5.5%, biomass by 2.6%, and grain yield by 2.8%. With a
temperature increase of 1  C over normal, a CO2 concentration of
>500 ppm was able to nullify the negative deviations in growth and yield,
but, when the temperature increased by 2  C, 600 ppm CO2 was needed to
nullify the adverse effect of temperature (Table 2.19).

6.5. Effect of climate change on the quality of produce


According to the IPCC’s Third Assessment Report, the significance of the
impact of climate change on grain and forage quality emerges from new
research. In a previous study, the amylose content in rice grain (a major
determinant of cooking quality) increased under elevated CO2 (Conroy

Table 2.19 Effect of increasing temperature and CO2 on change (%) in maximum leaf
area index (LAI), biomass yield, and grain yield of rice (Hundal and Prabhjyot-Kaur, 2007)
Temperature change Temperature change from
from normal (þ1  C) normal (þ2  C)
CO2 concentration Biomass Grain
(ppm) LAI yield yield LAI Biomass yield Grain yield
330 9.3 6.0 6.6 12.3 7.3 7.5
400 6.1 4.0 4.3 11.9 7.1 7.2
500 4.0 2.9 2.8 7.8 4.0 4.4
600 þ0.8 þ0.8 þ0.5 5.5 2.6 2.8
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 91

et al., 1994). In another study, cooked rice grains from plants grown in high-
CO2 environments were firmer than those from plants grown in ambient
CO2 environments; however, the concentrations of iron and zinc, which
are important for human nutrition, were lower (Seneweera and Conroy,
1997). Moreover, the protein content of the grains decreased with com-
bined increases in temperature and CO2 concentration (Ziska et al.,
1997). Studies have shown that higher CO2 concentrations led to reduced
plant uptake of nitrogen (N) and trace elements, such as zinc, resulting in
crops with lower nutritional value (Taub and Wang, 2008). This would pri-
marily impact people in poorer countries, who are less able to compensate by
eating more food and have less varied diets (Kaur and Rajni, 2012). Reduced
N content in plants used for grazing may also reduce animal productivity
(e.g., sheep depend on microbes in their gut to digest plants, which in turn
depend on N intake).
In a study, the protein content of soybean grain decreased with increases
in CO2 concentration; however, because of increased grain yield, the total
quantity of nutrients accumulated in grain per hectare still increased with
high CO2 concentrations (Mulchi et al., 1992). In the same study, increases
in CO2 concentrations from 360 to 510 ppm increased grain oil from 20.4 to
22.3% and decreased grain protein content. The N content of plants is likely
to decrease with elevated CO2, implying reduced protein.
At the International Rice Research Institute, rice (“IR72”) was grown
under three different CO2 concentrations (ambient, ambient þ200, and
ambient þ300 mL L1 CO2) and two different air temperatures (ambient
and ambient þ4  C) using open-top field chambers (Ziska et al., 1997).
Increasing both CO2 and air temperature reduced grain quality (e.g., protein
content). The combined effects of CO2 and temperature suggested that, in
warmer regions (i.e., >34  C) where rice is grown, quantitative and qual-
itative changes in rice supply could occur if both CO2 and air temperature
continued to increase.

6.6. Agricultural surfaces and climate change


Climate change might increase the amount of arable land near the poles by
decreasing the amount of frozen land. As per IPCC Assessment Report (AR)
4 (Bindoff et al., 2007), the reduction in area in the ice sheets of Greenland
and Antarctica contributed greatly to sea-level rise from 1993 to 2004
(Table 2.20). Although the impacts of sea-level rise are local in nature,
the causes are global and can be attributed to nonlinearly coupled
Author's personal copy

92 Bhagirath Singh Chauhan et al.

Table 2.20 Observed rate of global sea-level rise and estimated contributions from
different sources (Bindoff et al., 2007)
Rate of sea-level rise
(mm per year)
Source of sea-level rise 1961–2003 1993–2003
Thermal expansion 0.42  0.12 1.60  0.50
Glaciers and ice caps 0.50  0.18 0.77  0.22
Greenland ice sheet 0.05  0.12 0.21  0.07
Antarctica ice sheet 0.14  0.41 0.21  0.35
Sum of individual climate contributions to sea-level rise 1.1  0.5 2.8  0.7
Observed total sea-level rise a
1.8  0.5 3.1  0.7
Difference (observed minus sum of estimated climate 0.7  0.7 0.3  1.0
contributions)
a
Data prior to 1993 are from tide gauges and after 1993 are from satellite altimetry.

components of the Earth system. Sea levels are expected to rise by another
1 m by 2100 though this projection is disputed (Bindoff et al., 2007). The
rise in sea level may decrease agricultural land area, particularly in Southeast
Asia. With increasing sea levels, erosion, submergence of shorelines, and
salinity of the water table could affect agriculture by inundating low-lying
areas. Future climatic changes will affect water availability for agriculture.
Apart from monsoon rains, India depends on rivers that emanate from
the Himalayas for water-resource development. As a result of global
warming, the increase in temperature may increase snowmelt and conse-
quently snow cover will decrease. In the short run, snowmelt may increase
water flow in many rivers, which, in turn, may increase the frequency of
floods. In the long run, however, the receding snow line might reduce water
flow in these rivers. In climate-change scenarios, the onset of summer mon-
soon across India may become more uncertain and could be delayed. This
will influence not only rainfed crops but also water storage in irrigated areas.
CO2-induced warming is expected to raise the sea level as a result of
thermal expansion of the oceans and partial melting of glaciers and ice caps,
which, in turn, is expected to affect agriculture, mainly through the inun-
dation of low-lying farmland and increased salinity of coastal groundwater.
The IPCC estimates of sea-level rise above present levels are 9–29 cm by
2030, with a best estimate of 18 cm, and 28–96 cm by 2090, with a best esti-
mate of 58 cm (IPCC, 2007a). Preliminary surveys of the vulnerability of
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 93

land to inundation were made on the basis of existing contoured topo-


graphic maps, in conjunction with knowledge of the local “wave climate”
that varies between different coastlines (Parry and Carter, 1988). On the
basis of the extent of land liable to inundation, 27 countries were identified
as being vulnerable to sea-level rise.
The most severe effects of climate change on agriculture are likely to be
from flooding. Southeast Asia, because of the extreme vulnerability of sev-
eral large and heavily populated deltaic regions, would be most affected.
According to more than 20-year-old projections, a 1.5 m sea-level rise
would cause submergence of about 15% of all land (and about one-fifth
of all farmland) and render another 6% of the land more prone to frequent
flooding (UNEP, 1989). Altogether, 21% of agricultural production could
be lost. Estimates revealed that, in Egypt, 17% of national agricultural pro-
duction and 20% of all farmland, especially the most productive farmland,
would be lost as a result of a 1.5 m sea-level rise. Island nations, particularly
low-lying coral atolls, would suffer the most. In the Indian Ocean, 50% of
the land area of the Maldives would be submerged as a result of a 2 m rise in
sea level. In addition to direct farmland loss from flooding, agriculture would
experience increased costs from saltwater intrusion into surface water and
groundwater in coastal regions. Deeper tidal penetration would likely
increase the risk of flooding, and recharge of aquifers with seawater would
need to be prevented.
In addition, indirect impacts of flooding would relocate both farming
populations and production to other regions, which is a serious concern.
In Bangladesh, for example, as a result of the farmland loss from an estimated
1.5 m sea-level rise, about one-fifth of the nation’s population would be dis-
placed. It is important to emphasize, however, that the IPCC estimates of
sea-level rise are much lower (about 0.5 m by 2090 under the “business-
as-usual” scenario) than 1.5 m (UNEP, 1989).

6.7. Soil erosion and soil fertility


Climatic changes are expected to affect soils in many ways. Global warming
is likely to cause soil degradation, which could influence soil fertility.
Because the ratio of carbon (C) to N is a constant, doubling of C should lead
to storage of extra N in soils as nitrates, thus providing higher fertilizing ele-
ments for plants and enhancing crop yields (Blanco-Canqui and Lal, 2010).
The average need for N could decrease and provide an opportunity for
changing costly fertilization strategies. Climatic extremes (e.g., flooding)
Author's personal copy

94 Bhagirath Singh Chauhan et al.

would probably enhance the risk of erosion. The possible evolution of


organic matter in the soil is a highly contested issue. An increase in temper-
ature would increase the mineral production rate and lessen soil organic mat-
ter content (Salinger, 1989) (Table 2.20).
High temperatures could increase the rate of microbial decomposition of
organic matter, thus adversely affecting soil fertility in the long run, but
increases in root biomass resulting from higher rates of photosynthesis could
offset these effects (Buol et al., 1990). High temperatures could accelerate
the cycling of nutrients in the soil and more rapid root formation could pro-
mote more N fixation (Hillel and Rosenzweig, 1989). However, these ben-
efits could be minor compared with the adverse effects of changes in rainfall.
Increased rainfall in regions that are already moist, for example, could lead to
increased leaching of minerals, especially nitrates. In the Leningrad region of
Russia, an estimated one-third increase in rainfall would reduce soil produc-
tivity by more than 20% (Pitovranov et al., 1988). Large increases in fertilizer
applications would be necessary to restore productivity (Pitovranov et al.,
1988). Decreased rainfall, particularly during summers, could have a dra-
matic effect on the soil through the increased frequency of dry spells, leading
to increased proneness to wind erosion. Susceptibility to wind erosion,
however, depends in part on the cohesiveness of the soil, which is affected
by precipitation effectiveness, and wind velocity.
An experiment was conducted with free-air CO2 enrichment (FACE) in
paddy fields at Wuxi, China, to study the effects of elevated CO2 on the
availability of soil N and phosphorus (P) (Ma et al., 2007). Soil-available
N decreased with elevated CO2 by 47% in low N and by 29% in normal
N status at the rice tillering stage (Table 2.21). In this study, elevated
CO2 caused a significant increase in root biomass, which led to higher
N uptake by the rice plants. The enhanced C input by FACE increased soil
microbial use of N and this could be the reason for reduced soil-available N.
The results of the study also reported decreased extractable soil-available
N due to elevated CO2 during the early period of rice growth. Therefore,
N mineralization was increased and N uptake was decreased by elevated
CO2 concentration during the later growth stages. In the same study,
P uptake in rice was significantly increased by elevated CO2 under both
low N and normal N rates (Ma et al., 2007). Elevated CO2 caused a decrease
in soil-available P by 32% in low N and by 30% in normal N rates at the
jointing stage, but an increase of 22% and 21% at the heading stage and
an increase of 34% and 31% at the ripening stage were observed under
low N and normal N rates, respectively (Ma et al., 2007) (Table 2.22).
Table 2.21 Effects of elevated CO2 on soil-available N in 0–15 cm soil depth at four rice growth stages (Ma et al., 2007)
Content of soil-available N at different growth stages after rice transplanting
—D27— —D49— —D76— —D123—
Fertilizer treatment CO2 level mg kg1 IR mg kg1 IR mg kg1 IR mg kg1 IR
Low N F 17.6  0.6 47 17.9  0.8 10 10.7  2.2 0 9.8  2.3 10
A 33.4  5.1 16.4  3.7 10.7  2.2 10.8  2.5
Normal N F 36.3  5.3 29 16.6  2.1 3 10.8  0.7 6 11.5  2.8 11
A 51.0 10.3 16.1  1.8 11.5  2.1 10.3 2.4
Author's personal copy

Abbreviations: A, ambient CO2; F, elevated CO2 (200 m mol mol1 higher than ambient).
Increasing rate (IR) ¼ (F  A)/A  100.
D27, D49, D76, and D123 are the days after rice transplanting and equal to the tillering, jointing, heading, and ripening stages, respectively.
Table 2.22 Effects of elevated CO2 on soil-available P in 0–15 cm soil depth at four rice growth stages (Ma et al., 2007)
Content of soil-available P at different growth stages after rice transplanting
—D27— —D49— —D76— —D123—
Fertilizer treatment CO2 level mg kg1 IR mg kg1 IR mg kg1 IR mg kg1 IR
Low N F 10.4  2.3 11.3 3.3  1.0 32.0 5.1  1.0 22.4 5.3  0.9 33.8
A 9.3  1.8 4.8  0.3 4.2  0.9 4.0  1.4
Normal N F 10.2  1.3 5.8 4.0  1.3 29.6 5.1  1.8 20.8 5.4  1.6 30.7
A 10.9  2.3 5.7  0.8 4.3  1.8 4.1  1.0
Abbreviations: A, ambient CO2; F, elevated CO2 (200 m mol mol1 higher than ambient).
Increasing rate (IR) ¼ (F  A)/A  100.
Author's personal copy

D27, D49, D76, and D123 are the days after rice transplanting and equal to the tillering, jointing, heading, and ripening stages, respectively.
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 97

Elevated atmospheric CO2 enhanced soil P mineralization and availability of


soil P. This response was mainly because of higher root biomass. Elevated
CO2 was shown to increase phosphatase activity in the rhizosphere and exu-
dates at elevated CO2 increased the availability of soil P. Besides acid phos-
phatase activity, exudation of citrate was also important in releasing
inorganic P from aluminum–iron complexes in soils of highly P-limited sys-
tems (Ma et al., 2007).

6.8. Potential effects of climate change on pests


Like crop plants, weeds would undergo the same acceleration of growth
cycle and also benefit from carbonaceous fertilization. Under high temper-
ature, weeds with a C4 photosynthetic pathway have a competitive advan-
tage over C3 crop plants (Yin and Struik, 2008). Most of the weeds in rice are
of a C4 type and elevated CO2 may give a competitive advantage to rice (a
C3 crop) over C4 weeds (Fuhrer, 2003; Patterson, 1995). In crop–weed
competition studies, in which the photosynthetic pathway was the same,
weed growth was favored as CO2 increased. For example, the infestation
of Phalaris minor Retz. (C3) in wheat (C3) would worsen with an increase
in CO2 concentration (Mahajan et al., 2012). Because of CO2 enrichment,
the wheat crop could gain greater biomass than P. minor under well-irrigated
conditions. Under water stress, however, P. minor may have an advantage
over wheat at elevated CO2 concentration (Naidu and Varshney, 2011).
In a study in the United States, weedy rice (Oryza sativa L.) responded more
strongly than cultivated rice to rising CO2 concentration with greater com-
petitive ability (Ziska et al., 2010). These results suggest that weedy rice may
become a more problematic weed in the future in India.
In a study around three decades ago (Patterson and Flint, 1980), biomass
production in four plant species responded quite differently to CO2 concen-
trations (Table 2.23). In corn (Zea mays L.), 12 days after planting, plant bio-
mass was significantly greater at 600 and 1000 ppm CO2 concentrations than
at 350 ppm. Twenty-four days after planting, however, biomass did not dif-
fer significantly among the three CO2 concentrations. Forty-five days after
planting, biomass was significantly greater at 350 ppm than at 1000 ppm. In
Rottboellia cochinchinensis (Lour.) W.D. Clayton, biomass was significantly less
at 350 ppm CO2 than at 600 ppm CO2 at all three harvests. In soybean,
increasing CO2 concentration from 350 to 1000 ppm increased the biomass
of 45-day-old plants by 72%. In Abutilon theophrasti Medic., biomass was sig-
nificantly greater at 600 and 1000 ppm CO2 than at 350 ppm CO2 and this
Author's personal copy

98 Bhagirath Singh Chauhan et al.

Table 2.23 Effect of CO2 concentration on dry biomass of corn, Rottboellia


cochinchinensis, soybean, and Abutilon theophrasti harvested at 12, 24, and 45 days after
planting (Patterson and Flint, 1980)
Biomass (g plant1)
Species Days after planting 350 ppm CO2 600 ppm CO2 1000 ppm CO2
Corn 12 0.63b 0.73a 0.76a
24 6.96a 6.24a 6.63a
45 91.29a 89.49ab 80.08b
R. cochinchinensis 12 0.08b 0.16a 0.15a
24 2.10b 3.82a 3.50a
45 39.25b 47.47a 38.62b
Soybean 12 0.34c 0.52b 0.61a
24 3.60c 4.68b 6.38a
45 50.55c 62.19b 87.09a
A. theophrasti 12 0.08c 0.27a 0.21b
24 1.94b 3.73a 3.65a
45 35.4c 47.96b 54.34a
Different letters within a row are significantly different at 0.05 level (according to Duncan’s multiple
range test).

was true at all three harvests. Such responses of biomass production to CO2
concentrations indicate that the effects of CO2 also depend on the age or
growth stage of the plant.
In another study, soybean was grown at ambient and elevated CO2 con-
centrations (þ250 mL L1 CO2 above ambient concentration) with and
without the presence of a C3 weed (Chenopodium album L.) and a C4 weed
(Amaranthus retroflexus L.), to evaluate the impact of rising atmospheric CO2
on crop production losses caused by weeds (Ziska and Goins, 2006).
A significant reduction in soybean seed yield was observed with both weed
species relative to their weed-free control at each CO2 concentration. Inter-
ference from C. album caused a reduction in soybean seed yield relative to
the weed-free conditions; the reduction increased from 28% to 39% as CO2
concentration increased. There was a 65% increase in the mean biomass of
C. album at the enhanced CO2 concentration. Conversely, with A. retroflexus
interference, soybean seed yield losses decreased with increasing CO2 from
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 99

45% to 30%, with no change in the mean biomass of A. retroflexus. In a weed-


free environment, elevated CO2 resulted in a significant increase in vegeta-
tive biomass (33%) and seed yield (24%) for soybean compared with the
ambient CO2 concentration. Interestingly, the presence of both weeds
negated the ability of soybean to respond, vegetatively or reproductively,
to enhanced CO2 concentration. The results from this study suggested that
rising CO2 concentrations could alter current yield losses associated with
competition from weeds and weed control would be very crucial in realizing
any potential increase in the economic yield of agronomic crops (e.g.,
soybean) as atmospheric CO2 concentration increases.
Climate changes may also necessitate the adaptation of agronomic prac-
tices, which in turn influence weed growth. Weed management operations,
for example, chemical and mechanical, could be influenced by climate
change. Because of sudden changes in climate, environmental stress on crops
would increase and as a result the crop could become more vulnerable to
attack by insects and pathogens and less competitive with weeds. Climate
change increases the importance of conservation tillage practices and the
adoption of these practices requires knowledge of local conditions and an
understanding of the overall system dynamics. For instance, zero tillage as
a component of conservation agriculture in wheat and inappropriate fertil-
izer application in dry-seeded rice can increase weed infestation, which in
turn increases herbicide-use and/or reduces fertilizer-use efficiency. Tem-
perature changes may cause an expansion of weeds, with some species mov-
ing to higher latitudes and altitudes (Mahajan et al., 2012). Irrigation water in
northwestern India is increasingly becoming scarce and many resource-
conserving technologies are recommended to conserve irrigation water:
for example, zero tillage in wheat, bed planting in rice and wheat, and
dry-seeded rice. This will have consequences for weed abundance and com-
position. Hardy weeds, such as Rumex spp., may increase in wheat because of
increased adoption of zero tillage in wheat. Flooding is commonly the pri-
mary cultural means to suppress weeds in rice as water depths of a few cen-
timeters can suppress germination and emergence of a majority of weeds in
rice (Chauhan, 2012a; Chauhan and Johnson, 2008, 2009a, b, 2010). Alter-
nate wetting and drying in puddled and in dry-seeded rice-production sys-
tems may encourage weeds, such as Leptochloa chinensis (L.) Ness., Eleusine
indica (L.) Gaertn., Panicum repens L., Eclipta prostrata (L.) L., Eleocharis
spp., and Cyperus esculentus L. (Mahajan et al., 2009b). A dwindling supply
of irrigation water makes it difficult to maintain ponding in rice for effective
weed control. Under these circumstances, where farmers are not aware of
Author's personal copy

100 Bhagirath Singh Chauhan et al.

alternative weed control, yield losses are to be expected. Therefore, strate-


gies for weed management need to be changed according to the environ-
mental conditions.
The risk of herbicide application for weed control may increase as a result
of environmental change. Problems of resistant weeds, herbicide toxicity,
and poor weed control with herbicide application may increase in the near
future. Herbicide efficacy could be affected by elevated CO2, which has
been shown to increase the tolerance of weeds of herbicide (Ziska and
Teasdale, 2000). Changes in temperature and CO2 concentration may alter
transpiration, the number of leaf stomata, or leaf thickness, which, in turn,
may affect the absorption and translocation of herbicides. In C3 plants, an
increased concentration of leaf starch under elevated CO2 may reduce her-
bicide efficacy. Higher CO2 concentration may stimulate belowground
growth relative to aboveground growth and may favor rhizome and tuber
growth of perennial weeds (Ziska, 2003). Such information suggests that
the problem of perennial weeds in rice and wheat may increase in the near
future. These weeds, however, could be controlled through integrated
approaches that combine preventive, cultural, and chemical control mea-
sures. Integrated weed management strategies need to be developed that tar-
get weed invasion, recruitment, and reproduction. Such strategies may
include a combination of optimal fertilizer schedule, summer plowing, crop
rotation, land preparation, plant geometry modification, stale-seedbed tech-
nique, and the use of weed-competitive cultivars (Chauhan, 2012b;
Chauhan et al., 2006, 2012a,b). Knowledge of weed ecology and biology
could be used as a tool for effective weed management in futuristic
climate-change scenarios. Timely efforts to fill research gaps in management
and weed interactions are needed for the sustainability of the rice–wheat
cropping system in India. There are compelling reasons for expecting cli-
mate change to alter weed management; therefore, integrated novel
approaches must be developed to assist farmers in coping with the challenges
of weed control.
Global warming would increase rainfall in some areas, which would lead
to an increase in atmospheric humidity and the duration of the wet season
(Kaur and Rajni, 2012). Combined with higher temperatures, these condi-
tions could favor the development of fungal diseases. Similarly, because of
higher temperatures and humidity, incidences of insects and disease vectors
could increase.
Elad and Pertot (2013) had discussed the impacts of climate change on
plant pathogens and plant diseases. They explained that climate change
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 101

would affect the optimal conditions for infection, host specificity, and mech-
anisms of plant infection. They also pointed out that both pathogens and
host plants would be affected by the changing climate. The authors
suggested that climate-induced changes would affect the measures farmers
use to effectively control diseases and the viability of particular cropping sys-
tems in particular regions.
Sharma et al. (2007) reported that an increase in total rainfall
(69–260 mm) resulted in an epidemic of bacterial leaf blight (Table 2.24).
Plant diseases have a direct influence on crop productivity; however, limited
information is available on the impact of climate change on plant diseases
(Agrios, 2005; Elad and Pertot, 2013). The risk of yield losses from plant
diseases is likely to increase in the wake of climate change; however, such
production losses are rarely considered in climate assessments (Anderson
et al., 2004; Reilly et al., 2001). Climate change may have a direct influence
on the temporal and spatial distribution of plant diseases. Plant pathogens are
strongly affected by environment; therefore, the survival, rate of multiplica-
tion, vigor, sporulation, direction, distance of dispersal of inocula, rate of
spore germination, and penetration of pathogens can be affected in the
climate-change scenario (Kang et al., 2010). Research evidence has revealed
that elevated CO2 increased disease incidence or severity in some cases
(Eastburn et al., 2010; McElrone et al., 2005; Shin and Yun, 2010), but
decreased it in other cases (Hibberd et al., 1996; McElrone et al., 2010;
Pangga et al., 2004).

Table 2.24 Comparative performance of weather variables in the month of August


during some epidemic and nonepidemic years for bacterial leaf blight (BLB) of rice
in Punjab, India (Sharma et al., 2007)
BLB epidemic year BLB nonepidemic year
Weather parameters 1976 1985 1991 1992 1999 1979 1988 1993 2002
Rainfall (mm) 260 245 107 251 69 50 70 65 25
Rainy days (no.) 13 10 10 12 4 5 7 3 5
a
Temperature 30 16 18 20 15 8 18 2 7
b
Humidity 15 17 10 19 10 9 2 2 8
Sunshine (h) 6.6 7.0 7.6 5.9 7.9 9.7 8.3 10.7 6.3
Disease severity (%) 42 47 40 34 38 12 9 8 2
a
No. of days with temperature between 25 and 30  C.
b
Number of days with RH > 80%.
Author's personal copy

102 Bhagirath Singh Chauhan et al.

The implications of climate change relative to insects, crop protection,


and food security have recently been highlighted by Sharma (2013). He
pointed out that changes in geographic range and insect abundance would
be expected to increase the extent of crop losses, which would affect crop
production and food security. Global warming is expected to affect host–
plant resistance, biopesticides, natural enemies, and the synthetic chemicals
now used for integrated pest management (Sharma, 2013). The author
argued that climate change would cause increased problems with insect-
transmitted diseases, particularly in developing countries, where the need
to increase and sustain food production is most critical.
Abiotic factors may have a direct effect on insect–pest population
dynamics. These factors may influence developmental rate, fecundity, sur-
vival, and voltinism (Bale et al., 2002). The rise in temperature associated
with global warming may decrease the survival rate of brown planthopper
and leaffolder in rice (Heong et al., 1995). The authors concluded that rising
temperature could change pest population dynamics of the rice ecosystem.
In another study, elevated CO2 (570 ppm) exhibited a positive effect on
brown planthopper multiplication and more than doubled its population
(45 hoppers hill1) at peak incidence compared with the ambient CO2
(380 ppm) during the kharif season of 2010 (Prasannakumar et al., 2012).
Drought and rainfall play significant roles in soil insects’ abundance
(Srivastava et al., 2010; Staley et al., 2007). Global warming could alter vol-
tinism and this could be reflected in a change in geographic distribution
(Tobin et al., 2008). Elevated CO2 showed some impact on pest population
abundance by altering the nutritional value of plants and this could alter
insect abundance and increase the rate of herbivory (Dermody et al.,
2008; Rao et al., 2009). In a nutshell, climate change might alter the pop-
ulation dynamics of insects differently in different agroecosystems and
agroclimatic zones of India; therefore, more research is needed to under-
stand these issues. The effect of climate change may be more on temperate
insects; it could expand their range. Research is needed to systematically
document major and minor pests by investigating metabolic alteration in
insects in response to changing environment, developing prediction models,
and studying evolutionary changes under modified environment
(Karuppaiah and Sujayanad, 2012). In conclusion, changes in the spectrum
of weeds, diseases, insects, natural enemies, and antagonists; high risk of
invasion by exotic and migrant pathogens and insects; extension of geo-
graphic range; noxious abundance of several species at higher altitudes;
increased overwintering; changes in morphology and reproduction; altered
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 103

development; increased number of generations; loss of resistance in cultivars


containing temperature-sensitive genes; extension of the crop-development
season and its effect on pest synchrony; change in interspecific interactions at
different trophic levels; and decreased pesticide efficacy are some of the
impacts of climate change. Better understanding the direct and indirect
effects of climate change is crucial for developing pest-management
programs. The use of historical crop, climate, and pest-management data
vis-à-vis current conditions provides ample and immediate scope for under-
standing the effects of climate change and planning for adaptive, integrated
pest-management strategies. Another approach toward understanding of the
potential direct effects is to conduct studies under controlled conditions for
knowing how intrinsic population growth is related to temperature and
identifying relationships among temperature, phenology, and population
growth rates by the use of appropriate models. Assessing the changing pest
scenario, mapping regions vulnerable to pest risk, and evolving curative and
preventive pest-management strategies for climate stress should be empha-
sized among many approaches.

7. KEY ADAPTATION AND MITIGATION STRATEGIES TO


REDUCE THE EFFECTS OF CLIMATE CHANGE
Climate change, involving frequent changes in temperature, precipi-
tation, and sea level and increased impact of GHGs, is bound to affect agri-
cultural production. Potential changes in temperature and precipitation may
have a strong influence on Indian agriculture. There have been trepidations
about a possible increase in the El Niño current, which is thought to be a
major factor contributing to drought in India. These climate-change issues
call for greater understanding of crop–climate interactions and for develop-
ing crop–weather models to devise efficient agricultural production strate-
gies (Lal et al., 2001).
India has hardly any scope for future horizontal expansion to meet the
increasing demand for food, fodder, fiber, fuel, and other products. Scope
exists only for vertical expansion. To overcome the problem of climate
change, agronomic management practices could play a significant role.
Agronomists will have to be in the forefront of the research agenda and must
develop cropping/farming systems and agronomic management practices
that would harmonize high production with ecological safety. They must
play a key role on a research team in formulating crop production technol-
ogy practices in the wake of climate change. Agronomists have to decide
Author's personal copy

104 Bhagirath Singh Chauhan et al.

which and how much each of the recommendations made by plant breeders,
soil scientists, entomologists, plant pathologists, etc., will make a technically
viable, socially acceptable, economically profitable, and environmentally
sound package for a particular crop in a cropping system. They must caution
against practices that can harm the system. Starting from the basics of the
soil–plant–water–atmospheric system to develop cultivation practices for
high productivity and fitting them into farming systems through multi-
disciplinary collaborative research, or a systems approach, is required.
Climatic variations are present throughout the world, but the impacts of
climate change are the most devastating in developing countries, such as
India, that have fewer resources than developed countries to cope with these
adverse affects. Sustainable food security is therefore difficult to achieve in
developing countries, especially in India, because of the ever-increasing
human population; higher demand for, and intensification of, resource
use; and increased per capita consumption (Rosenzweig and Parry, 1994).
With the emerging threats from climate change, there are many uncer-
tainties as agriculture is sensitive to short-term changes in weather and to
seasonal, annual, and long-term climatic variations. Variations in meteoro-
logic parameters, in combination with other parameters, such as soil char-
acteristics, cultivars, and pests, have a paramount influence on agricultural
productivity (Pathak and Wassmann, 2009). The Fourth Assessment Report
(AR4) of the IPCC suggested that increasing trends of GHGs in the Earth’s
atmosphere could accelerate in the future, as a consequence of which, the
best estimates of increases in mean global surface temperature are likely to
be in the range of 1.8–4  C (IPCC, 2007a). Globally, mean precipitation
is projected to increase with great deviances regionally (Meehl et al.,
2007). It is therefore imperative to chart adaptation and mitigation strategies
to counter the effects of climate change on agricultural commodities. Mit-
igation and adaptation are measured on temporal and spatial scales on which
they are effective. Mitigation strategies aim at reducing GHG emissions into
the atmosphere, and adaptation strategies aim at enabling the plants to per-
form optimally under adverse climatic conditions through cultural and
genetic manipulations.
The benefits of mitigation activities will be evident in several decades
because of the longer duration of GHGs in the atmosphere, whereas the
effects of adaptation measures should be seen immediately or in the near
future (Kumar and Parikh, 2001; Lal, 2011). The effects or benefits of mit-
igation strategies are both global and local, whereas the effects of adaptation
strategies are local or regional. The purpose of mitigation and adaptation
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 105

measures is to attempt a gradual reversal of the impact of climate change and


sustain development under the inescapable effects of climate change.

7.1. Crop-based approaches


Adjustment in sowing dates is a simple yet powerful tool for adapting to the
effects of potential global warming. Krishnan et al. (2007) demonstrated
potential outcomes by adjusting the sowing time of rice at two sites
(Cuttack and Jorhat in India) by simulating crop growth under different
climate-change scenarios. Improved agronomic practices, such as altering
planting dates, helped in minimizing the effect of high temperature respon-
sible for yield instability in rice and wheat. Manipulation of planting dates
helped in reducing yield instability by keeping flowering from coinciding
with the hottest growing season (Mahajan et al., 2009a).
On several occasions in the last decade, South Asia witnessed adverse
effects of climatic variations, that is, terminal heat stress, on wheat produc-
tivity. For example, despite favorable weather conditions during the winter
of 2009–2010, an abrupt rise in night temperature during the grain-filling
stage in wheat adversely affected wheat productivity in the Indo-Gangetic
Plains (IGP) and other northern states of India (Gupta et al., 2010). The
results indicated that terminal heat stress on wheat in Punjab in
2009–2010 led to a mean yield penalty of 5.8% compared with the previous
year. However, yield losses reached 20% in a few districts of Punjab and
other transects of the IGP. The magnitude of losses varied depending on
planting time, cultivars, and other management practices.
Agronomic management of crops, such as method of sowing, can be an
effective adaptation strategy under the climate-change scenario. Bed plant-
ing of crops has proved successful in the wake of climate change as it results
in increased water-use efficiency, reduced waterlogging, better access
for interrow cultivation, weed control, banding of fertilizers, better stand
establishment, less crop lodging, and reduced seeding rates (Bhardwaj
et al., 2009; Chauhan et al., 2012a). In irrigated areas, zero tillage in wheat
cultivation has successfully reduced the demand for water and other
resources (e.g., diesel and herbicides) and zero-till systems are now consid-
ered a viable option to combat climate change. Intercropping is a time-tested
practice in the wake of climate change. If one crop fails because of flooding
or drought, the second crop provides minimum assured returns for liveli-
hood security (Mittal and Singh, 1989). Large numbers of recommendations
have been made for different zones on crop substitution and cultivar
Author's personal copy

106 Bhagirath Singh Chauhan et al.

replacement in the case of delayed onset of monsoon (Joshi and Kar, 2009).
In this approach, short-duration crops and cultivars replace long-duration
ones. In the case of early-season drought, possible options are replanting
with a crop and cultivar(s) suitable for late sowing and transplanting by rais-
ing seedlings or taking seedlings from other fields. For midseason drought,
viable options are forming dead furrows at convenient intervals (3–4 m) well
in advance of anticipated drought (within a month after seeding or imme-
diately after intercultivation) to minimize runoff and store rainwater and
thinning the plant population either within rows or by removing alternate
rows in a sole crop or removing more sensitive crop in intercropping and
harvesting the crop for fodder and allowing the stubbles to grow for grain
(ratooning) as in the case of sorghum and pearl millet. For late-season
drought, however, options are limited. A crop on relatively deep soil can
be removed and a short-duration rabi pulse crop can be sown on stored soil
moisture with subsequent rain. In the case of sorghum and pearl millet,
ratooning appears to be ideal even at the time of late-season drought, espe-
cially in deep black soils (Venkateswarlu and Shanker, 2009).

7.2. Crops and cultivars that fit into new cropping systems
and seasons
With climate change, the IGP could continue to be the major contributor of
food grains despite the scarcity of irrigation water, provided new cultivars
are grown with judiciously selected cultivation schedules. The area has three
cropping seasons: rabi or winter season from October/November to March/
April, zaid or summer season from March/April to June/July, and kharif or
rainy season from July/August to October/November. Different cropping
systems could be practiced with the use of suitable cultivars for high yields
in this era of climate change, for example, dry-seeded rice/pigeon
pea/soybean/urad bean/cotton in kharif, potato/rape/mustard–wheat/
chickpea/lentil in rabi, and mung bean/soybean in zaid. For high profitabil-
ity for farmers, essential-oil crops such as menthol mint and medicinal crops
could be substitutes for mung bean/soybean/cotton in the zaid season as
suitable cultivars of these crops are already available. For strategic reasons,
a water-guzzling crop like sugarcane should continue to be grown in the
Himalayan terai/foothill region. Again, for augmenting the liquid biofuel
supply, crops such as mustard can be cultivated in irrigated areas for high
yield. In this direction, a mustard crop could be grown twice in the rabi sea-
son if suitable short-duration, early-maturing cultivars become available in
the future. The cotton crop yields lint and oil; therefore, it could be grown
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 107

on additional land during the kharif season. Potato should assume the role of a
staple food as it is a rich source of carbohydrates. India will need to produce
120 million tons of wheat, 25 million t of pulses, and 100 million t of oil
seeds (50 million t for biofuel purposes) by 2025 (Kumar, 2006). In the
climate-change scenario, the IGP should become a major supplier of these
commodities. Crop breeding programs with the objective of developing
climate-resilient (temperature- and drought-tolerant), high-yielding culti-
vars of the identified crops should be given high priority, so that the desired
kinds of cultivars become available in the wake of climate change.
A combination of conventional and molecular marker-assisted, muta-
tional, and transgenic breeding approaches will be required to develop
the desired kinds of crop cultivars. Crop-based coordinated programs
need to begin as early as possible to develop climate-resilient cultivars.
Recently, the Indian Agricultural Research Institute (IARI), New Delhi,
released an early-maturing basmati (fragrant) rice and a wheat cultivar suit-
able for late planting (Swaminathan and Kesavan, 2012). It appears that the
desired kinds of cultivars can also be selected in some of the ongoing breed-
ing programs. There will be a need for identifying areas where climate-
change conditions already exist or are mimicked (e.g., Rajasthan, Madhya
Pradesh, and Uttar Pradesh border areas in the IGP) and/or setting up suit-
able environmental chambers for the purpose of screening large segregating
populations to make selections (Rupakumar et al., 2006). Climate-change
issues need to be converted from a “problem” into an “opportunity.”

7.3. Cultivars suitable for high temperature, drought, inland


salinity, and submergence tolerance
Cultivars that fit into an erratic rainfall season and are drought- and
submergence-tolerant, have high fertilizer- and radiation-use efficiency,
and can tolerate coastal salinity and seawater inundation are needed. Germ-
plasm of wild relatives and local land races could be used for developing
climate-resilient crop cultivars. There is a need to revisit germplasm that
has tolerance of heat and cold stresses but has not been used in the past
because of low yield potential. Breeding cultivars that are tolerant of
high-temperature stress should receive utmost importance. Recently, for
example, Tao et al. (2008) identified rice hybrid Guodao 6 as heat-tolerant.
Genetic improvement of heat-tolerant genotypes, especially in pulses, by
identifying and validating markers for high-temperature tolerance with high
yield potential is one of the key technological processes that could be a
significant approach for adapting to climate change. Inclusion of abiotic
Author's personal copy

108 Bhagirath Singh Chauhan et al.

stress-tolerant cultivars in cropping systems and the development and use of


new cultivars with increased resilience to drought and flooding and
increased resistance to heat shock is important for climate adaptation.
Improving the adoption and dissemination of short-duration crop cultivars
can enhance the ability of farmers to cope with variable climatic conditions.
One of the most predictable issues of CO2-induced global warming is the
melting of the ice caps and glaciers and the addition of excess water to the
sea, thereby leading to the submergence of coastal areas. Further, the coastal
areas of the world are most fertile and densely populated. Therefore, breed-
ing of crops and cultivars that can be grown in saline soils could be an effi-
cient adaptation strategy to offset the loss of agricultural land. Genetically
modified rice cultivars possessing genes for salinity tolerance have been bred
by transferring genes from the mangrove species Avicennia marina (Sultana
et al., 2012). Similarly, for developing drought-tolerant strains of rice and
other crops, Prosopis juliflora L. is being used as a donor of a gene for drought
tolerance (Swaminathan and Kesavan, 2012). With the change in climatic
scenario, corn could emerge as an alternative crop in the rice–wheat
cropping system, replacing rice. Temperate germplasm, being highly pro-
ductive, could be used for the introgression of desirable genes into promising
tropical/subtropical backgrounds for the development of inbreds suitable for
different agroclimatic conditions.
Using FR13A, one of the submergence-tolerant donors, improved rice
cultivars with submergence tolerance have been developed (e.g., Swarna-
Sub1 rice, the first submergence-tolerant product using marker-assisted
backcrossing to introduce a major QTL locus, SUB1). Cultivars with the
SUB1 trait tolerate complete submergence at the seedling stage
(Iftekharuddaula et al., 2011; Septiningsih et al., 2009; Xu et al., 2006);
however, they are susceptible to anaerobic germination. Introgression of
SUB1 did not improve germination percentage under flooded conditions.
In this direction, Nipponbare (a japonica cultivar), possessing characteristics
amenable to underwater seedling establishment, exhibited greater seedling
vigor under submergence because of rapid shoot elongation and was even
found to be better than the internationally recognized submergence-tolerant
cultivar FR13A (Vu et al., 2010). Commendable work is going on in this
direction at IRRI and elsewhere to face flood-like situations in the future
at the germination stage of the crop.
Drought occurrence is likely to be more pronounced in rainfed areas.
Although the occurrence of drought is not as sudden as other weather haz-
ards, its effect can be equally devastating, especially in developing nations. In
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 109

India, significant progress has been made in the genetic dissection of


flowering time, inflorescence architecture, temperature, and drought toler-
ance in certain model plant systems and in comparative genomics in crop
plants. The Central Research Institute for Dryland Agriculture (CRIDA)
in Hyderabad has developed a sorghum cultivar (transgenic), “SPV 462,”
that possesses tolerance of water deficit and salt stress. The germination
potential of these transgenic seeds was several times higher when challenged
with salt and water stresses and the plants had a robust root system (root bio-
mass and root length) (Maheswari et al., 2010). The scientific approach
toward drought mitigation involves the pinpointing of drought-prone areas
and the scientific management of such vulnerable areas with drought-
tolerant cultivars. A holistic approach toward the management of water,
land, crop, and other natural resources, coupled with drought-tolerant cul-
tivars, can go a long way toward alleviating the adverse impacts of drought
(Sharma et al., 2010).

7.4. Cultivars that respond to high CO2 concentration


As projected by various models, the future climate will be rich in CO2 con-
centration and it is also clear that the vegetation will be positively benefited
by increased CO2 concentration (Farquhar, 1997). This beneficial effect will
be more pronounced for C3 plants, such as wheat, rice, barley, oats, peanut,
cotton, sugar beet, tobacco, spinach, soybean, and most trees. In all of these
plants, the elevated concentrations of CO2 will lead to higher assimilation
rates and an increase in stomatal resistance, resulting in a decline in transpi-
ration rate and improved water-use efficiency in crops. Hence, plant
breeders need to develop cultivars that are able to benefit from the high
CO2 fertilization effect. IRRI, in collaboration with several countries, is
working to alter the photosynthesis of rice from the C3 to C4 pathway by
introducing cloned genes from C4 species (e.g., corn and sorghum).

7.5. Mitigation of the impact of climate change


Many mitigation technologies are available that can help cope with the
challenges of climate change with desirable results. Agriculture, forestry,
and fisheries/aquaculture have great potential for mitigating GHG emis-
sions. According to the IPCC, the global technical mitigation potential
for agriculture will be between 5500 and 6000 million t CO2 equivalents
per year by 2030, 89% of which is assumed to be from sequestration of C in
the soil. Rice is a staple food for a large population in India and the crop
Author's personal copy

110 Bhagirath Singh Chauhan et al.

occupies the largest area in India. The maximum emission of CH4 is from
the rice-growing areas. CH4 emissions from rice fields can be restricted
with the adoption of improved agricultural practices. Resource-
conserving technologies, such as zero tillage in wheat and dry-seeded rice,
could play a major role in this direction. Zero-till systems have a direct
mitigation effect as they convert GHGs such as CO2 into O2 in the atmo-
sphere and C enriches soil organic matter. In dry-seeded rice, because of
minimum anaerobic conditions, improved root growth and diversity of
aerobic soil organisms may help in mitigating climate change. Research
has shown that yields similar to those in puddled-transplanted rice can
be achieved with alternate wetting and drying (Mahajan et al., 2011).
However, alternate wetting and drying may lead to emissions of N2O,
which has greater global warming potential than CH4 does. However, this
problem could be reduced by adopting integrated nutrient-management
practices, which can help in mitigating climate change. Integrated nutrient
management involves, in general, a combination of organic, inorganic, and
biofertilizers in proportions that will keep the soil capable of producing at
an accelerated rate without suffering physical, chemical, and biological
damage. The advantages of integrated nutrient management are increased
N-use efficiency and increased yield. The application of urease, hydroqui-
none, and nitrification inhibitors, dicyandiamide together with urea, is an
effective technology for reducing NO2 and CH4 emissions from rice fields.
The use of neem-coated urea is another simple and cost-effective technol-
ogy. Improved management of livestock and their diet could also assist in
the mitigation of GHGs. The use of improved food additives, substitution
of low-digestibility feeds with high-digestibility ones, concentrate feeding,
substituting fibrous concentrate with starchy concentrate, supplementa-
tion with molasses, and changing microflora of rumen could help in reduc-
ing CH4 emissions (Aggarwal, 2008). Efficiency of energy use in
agriculture could be improved by using better-designed machinery
(e.g., the Happy Seeder, a drill for dry seeding, zero-till drill, and bed
planter) that could increase fuel-use efficiency and help in the commercial-
ization of wind and solar power potential, and the use of a laser leveler
could also lead to mitigation (Chauhan et al., 2012a; Lal, 2011). Changing
land use by increasing the area under biofuel-producing crops and
agroforestry could help in mitigating GHG emissions, but this has to be
considered with the goal of increasing food production for national secu-
rity (Venkateswarlu et al., 2011).
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 111

7.6. Other strategies


Many options exist in soil-, water-, and nutrient-management technologies,
which can contribute to both adaptation and mitigation. The addition of
crop residues and manure to arable soils improves the soil water-holding
capacity (Benbi et al., 2011). Soil C sequestration is a useful strategy in
the wake of climate change. Although it has limitations in tropical areas
because of high temperature, a substantial quantity of C can be sequestered
with the adoption of improved agricultural practices. There are two types of
C sequestration, soil C sequestration and sequestration into vegetation.
Tree-based systems can sequester substantial quantities of C into biomass
in a short period of time (Lal, 2004). The total potential of soil
C sequestration in India is 39–49 Tg year1. This amount includes the
potential of the restoration of degraded soils and ecosystems, which is esti-
mated at 7 –10 Tg C year1. The potential of the adoption of improved
agronomic practices on agricultural soil is 6–7 Tg C year1. In addition,
there is also the potential of soil inorganic-carbon sequestration estimated
at 21.8–25.6 Tg C year1 (Lal, 2011). By providing shelter and shade, as
in agroforestry systems, the effects of extremely high temperatures may
decline (Cannell et al., 1996). The planting of multipurpose trees on
degraded lands helps in C sequestration. The agroforestry system protects
farmers from climatic variability and helps in reducing the atmospheric load
of GHGs. In India, much of the research done on in situ moisture conser-
vation, energy efficiency in agriculture, and the use of poor-quality water
relates to rainfed agriculture. The watershed approach may prove successful
in rainfed areas and help in both adaptation and mitigation. For example, soil
and water conservation work, farm ponds, and check dams help moderate
rainwater runoff and minimize flooding during high-intensity rainfalls. Lal
(2004) estimated that, by arresting water and wind erosion, 3–4.6 Tg year1
of C could be sequestered. The withdrawal of groundwater from deeper
layers demands more energy and leads to GHG emissions in agriculture
(Hira, 2009). If surface storage of rainwater in dugout ponds is encouraged,
dependence on withdrawing groundwater might decrease. The conjunctive
use of surface water and groundwater is an important strategy to mitigate
climate change, especially where groundwater is not suitable for agriculture.
Innovative approaches in groundwater sharing may also contribute to an
equitable distribution of water and reduced energy use in pumping. Biochar
is another novel approach to sequester C in terrestrial ecosystems; several
associated products are in the process of being manufactured. India could
Author's personal copy

112 Bhagirath Singh Chauhan et al.

produce almost 310 million t of biochar annually, whose application might


offset about 50% of C emissions (292 Tg C year1) from fossil fuels (Lal,
2005). The rice–wheat cropping system in the IGP of India produces sub-
stantial quantities of crop residues, and, if these residues can be pyrolyzed,
50% of the C in biomass could be returned to the soil as biochar, thereby
increasing soil fertility and crop yield by sequestering C (Lal, 2011).
In situ water harvesting can help increase the rainwater-use efficiency of
crops through water-conserving techniques such as compartment bunding,
ridges and furrows, strip cropping, mulching, and vegetative barriers to soil
moisture.
The diversification of sensitive agricultural production systems (e.g.,
rainfed agriculture) into less sensitive agricultural microenterprises (small-
scale vegetable and fruit production, livestock rearing, bee keeping, etc.)
can enhance adaptation to the short- and medium-term impacts of climate
change (Aggarwal, 2008). Diversifying income through other farming activ-
ities, such as livestock raising, tree farming, pond aquaculture, agroforestry,
and silvicultural practices, is a viable and effective approach to combating
climate change.

7.7. Policy issues for managing climate change


In addition to the use of technological strategies in overcoming climate
change-related impacts on crop production, there must be a solid policy
structure and strong political determination on the part of governments
to effectively tackle climate change. A sound policy framework should
examine the issues of redesigning the social sector with a focus on vulnerable
areas/populations, liability during extreme weather events, the introduction
of new credit instruments with deferred repayment, and weather insurance
as a major vehicle to transfer risk. The role of community institutions and the
private sector in relation to agriculture should be a policy matter. Policy
initiatives toward access to banking; microcredit; insurance service before,
during, and after a disaster event; and access to communication and infor-
mation services are needed in the envisaged climate-change scenario. The
establishment of an early warning system for emerging climatic risks, such
as drought, flood, heat and cold waves, and pest outbreaks, is desired. In
India and globally, climate change is now on the political and public agenda.
In India, particular attention is being paid to the impact of climate change on
agriculture, because it could have serious implications for the food security
of the country. Most scientists and policymakers now acknowledge that
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 113

climate change will have far-reaching effects on livelihood and food security
unless significant steps are taken to deal with it effectively now. The key pol-
icy issues are (1) establishment of a “Green Research Fund” for strengthen-
ing research on adaptation, mitigation, and impact assessment; (2) facilitating
greater adoption of scientific and economic pricing policies for water, land,
energy, and other natural resources; (3) providing financial incentives for
improved land management; (4) ensuring food and livelihood security;
(5) establishing seed banks in highly variable and unpredictable environ-
ments, etc. In India, governments should invest more in water storage
and efficient water-use technologies. Investment should be made in technol-
ogies that allow aquifer recharge and microirrigation to increase the efficient
use of available water. More investment should be made in funding projects
on cultivars with tolerance of adverse climate and on land-use systems to
ensure adequate food production in the face of climate change.

8. CONCLUSIONS
Climate change is a reality. Elevated CO2 concentration may increase
crop growth and yield due to increased photosynthesis, decreased photores-
piration, and decreased stomatal conductance. The increase in temperature,
however, may decrease grain yields of rice and wheat due to the shorter
duration of crop growth. The protein content of legume grains may decrease
with increased CO2 concentration. Elevated CO2 concentration may
increase the availability of soil N and P because of increased mineralization
and activity of phosphatase enzyme in the rhizosphere. C3 plants are likely to
compete even more vigorously than now against C4 crops and vice versa.
Increased temperature along with humidity may increase the occurrence
of insects and diseases. Because of the complexity of crop–environment
interactions, a multidisciplinary approach to the problem is required in
which plant breeders, crop physiologists, agrometeorologists, and agrono-
mists need to interact to find long-term solutions in sustaining agricultural
production. There is a need for strategic research to enhance the resilience of
Indian agriculture, including crops, natural resource management, horticul-
ture, livestock, and fisheries, for the development and application of
improved production and risk management technologies. In addition, there
is a need for technology demonstration of existing management practices for
enhancing the resilience of crops and livestock to climate change. Capacity
building of scientists and other stakeholders in agricultural research on climate
resilience may also help in developing solutions for climate change.
Author's personal copy

114 Bhagirath Singh Chauhan et al.

ACKNOWLEDGMENT
We would like to acknowledge the help of Bill Hardy in providing comments on the
manuscript.

REFERENCES
Abrol, Y.P., Bagga, N., Chakravarty, N.V.K., Wattal, P.N., 1991. Impact of rise in tem-
perature on the productivity of wheat in India. In: Abrol, Y.P., Gnanam, A.,
Govindjee, D.R. Ort, Teramura, A.H., Wattal, P.N. (Eds.), Proceedings of the Indo-
US Workshop on Impact of Global Climate Change on Photosynthesis and Plant
Productivity. Oxford and IBH Publishing, New Delhi, pp. 787–798.
Adams, R.M., Hurd, B.H., Lenhart, S., Leary, N., 1998. Effects of global climate change on
agriculture: an interpretative review. Climate Res. 11, 19–30.
Aggarwal, P.K., 2008. Global climate change and Indian agriculture: impact, adaptation and
mitigation. Indian J. Agric. Sci. 78, 911–919.
Aggarwal, P.K., Kalra, N., 1994. Analyzing the limitations set by climatic factors, genotype
and water and nitrogen availability on productivity of wheat II. Climatically potential
yields and management strategies. Field Crop. Res. 38, 93–103.
Agrios, G.N., 2005. Plant Pathology. Elsevier, San Diego, CA.
Anderson, P.K., Cunningham, A.A., Patel, N.G., Morales, F.J., Epstein, P.R., Daszak, P.,
2004. Emerging infectious diseases of plants: pathogen pollution, climate change and
agrotechnology drivers. Trends Ecol. Evol. 19, 535–544.
Baer, P., Risbey, J.S., 2009. Uncertainty and assessment of issues posed by urgent climate
change: an editorial comment. Clim. Change 92, 31–36.
Baker, J.T., Allen, L.H., Boote, K.J., 1990. Growth and yield responses of rice to carbon
dioxide concentration. J. Agric. Sci. 115, 313–320.
Bale, J., Masters, G., Hodkinson, I., Awmack, C., Jnbezemer, M., Brown, V.K.,
Butterfield, J., Buse, A., Coulson, J.C., Farrar, J., Good, J.G., Harrington, R.,
Hartley, S., Jones, T.H., Lindroth, L., Press, M., Mrnioudis, I., Watt, A.,
Whittaker, A., 2002. Herbivory in global climate change research: direct effects of rising
temperature on insect herbivores. Glob. Chang. Biol. 8, 1–16.
Battle, M., Bender, M., Sowers, T., Tans, P.P., Butler, J.H., Elkins, J.W., Ellis, J.T.,
Conway, T., Zhang, N., Lang, P., Clarke, A.D., 1996. Atmospheric gas concentrations
over the past century measured in air from firn at the south pole. Nature 383, 231–235.
Benbi, D.K., Manchanda, J.S., Gossal, S.K., Walia, S.S., Toor, A.S., Minhas, P.S., 2011. Soil
Health Issues for Sustaining Agriculture in Punjab. Directorate of Research, Punjab
Agricultural University, Ludhiana, Punjab, Research Bulletin 4/2011.
Bhardwaj, V., Yadav, V., Chauhan, B.S., 2009. Effect of nitrogen application timings and
varieties on growth and yield of wheat grown on raised beds. Arch. Agron. Soil. Sci.
56, 211–222.
Bhattacharya, S., Mitra, A.P., 1998. Greenhouse gas emissions in India for the base year 1990.
Global Change 11, 30–39.
Bindoff, N.L., Willebrand, J., Artale, V., Cazenave, A., Gregory, J., Gulev, S., Hanawa, K.,
LeQuere, C., Levitus, S., Nojiri, Y., Shum, C.K., Talley, L.D., Unikrishnan, A., 2007.
The physical basis: contribution of working group I to the fourth assessment report of the
IPCC. Cambridge University Press, Cambridge, UK.
Blanco-Canqui, H., Lal, R., 2010. Soil and water conservation. In: Principles of Soil Con-
servation and Management. Springer, Netherlands, pp. 1–19.
Buol, S.W., Sanchez, P.A., Kimble, J.M., Weed, S.B., 1990. Predicted impact of climatic
warming on soil properties and use. Am. Soc. Agron. Spec. Pub. 53, 71–82.
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 115

Cannell, M.G.R., van Noordwijk, M., Ong, C.K., 1996. The central agroforestry hypoth-
esis: the trees must acquire resources that the crop would not otherwise acquire. Agro-
forestry Syst. 34, 27–31.
Ebi, K.L., Sussman, F.G., Wilbanks, T.J., 2008. Analyses of the effects of global change on
human health and welfare and human systems. In: Gamble, J.L. (Ed.), A report by the US
Climate Change Science Program and the Subcommittee on Global Change Research.
U.S. Environmental Protection Agency, Washington, DC.
Chauhan, B.S., 2012a. Can knowledge in seed ecology contribute to improved weed man-
agement in direct-seeded rice? Curr. Sci. 103, 486–489.
Chauhan, B.S., 2012b. Weed ecology and weed management strategies for dry-seeded rice in
Asia. Weed Technol. 26, 1–13.
Chauhan, B.S., Johnson, D.E., 2008. Germination ecology of Chinese sprangletop
(Leptochloa chinensis) in the Philippines. Weed Sci. 56, 820–825.
Chauhan, B.S., Johnson, D.E., 2009a. Ecological studies on Cyperus difformis, C. iria and
Fimbristylis miliacea: three troublesome annual sedge weeds of rice. Ann. Appl. Biol.
155, 103–112.
Chauhan, B.S., Johnson, D.E., 2009b. Ludwigia hyssopifolia emergence and growth as affected
by light, burial depth and water management. Crop Prot. 28, 887–890.
Chauhan, B.S., Johnson, D.E., 2010. The role of seed ecology in improving weed manage-
ment strategies in the tropics. Adv. Agron. 105, 221–262.
Chauhan, B.S., Gill, G., Preston, C., 2006. Tillage system effects on weed ecology, herbicide
activity and persistence: a review. Aust. J. Exp. Agric. 46, 1557–1570.
Chauhan, B.S., Mahajan, G., Sardana, V., Timsina, J., Jat, M.L., 2012a. Productivity
and sustainability of the rice-wheat cropping system in the Indo-Gangetic Plains of
the Indian subcontinent: problems, opportunities, and strategies. Adv. Agron. 117,
315–369.
Chauhan, B.S., Singh, R.G., Mahajan, G., 2012b. Ecology and management of weeds under
conservation agriculture: a review. Crop Prot. 38, 57–65.
Conroy, J.P., Seneweera, S., Basra, A., Rogers, G., Nissen-Wooley, B., 1994. Influence of
rising atmospheric CO2 concentrations and temperature on growth, yield, and grain
quality of cereal crops. Aust. J. Plant Physiol. 21, 741–758.
Dar, W.D., Gowda, C.L.L., 2013. Declining agricultural productivity and global food secu-
rity. In: Kang, M.S., Banga, S.S. (Eds.), Combating Climate Change: An Agricultural
Perspective. CRC Press, Boca Raton, FL, pp. 1–10.
Das, L., Lohar, D., Sadhukhan, L., Khan, S.A., Saha, A., Sarkar, S., 2007. Evaluation of the
performance of ORYZA2000 and assessing the impact of climate change on rice pro-
duction in Gangetic West Bengal. J. Agrometeorol. 9, 1–10.
De Costa, W.A.J.M., Weerakoon, W.M.W., Herath, H.M.L.K., Amaratunga, K.S.P.,
Abeywardena, R.M.I., 2006. Physiology of yield determination of rice under elevated
carbon dioxide at high temperatures in a subhumid tropical climate. Field Crop. Res. 96,
336–347.
De, U.S., Mukhopadhyay, R.K., 1998. Severe heat wave over Indian subcontinent in 1998
in a perspective of global climate. Curr. Sci. 75, 1308–1311.
Dermody, O., O’Neill, B.F., Zangrel, A.R., Berenbaum, Y.M.R., Delucia, E.H., 2008.
Effects of elevated CO2 and O3 on leaf damage and insect abundance in soybean eco-
system. Anthropod-Plant Interact. 2, 125–135.
Eastburn, D.M., Degennaro, M.M., Delucia, E.H., Dermody, O., McElrone, A.J., 2010.
Elevated atmospheric carbon dioxide and ozone alter soybean diseases at SoyFACE.
Glob. Chang. Biol. 16, 320–330.
EIA, 2008. Emissions of Greenhouse Gases in the United States 2008 (www.eia.gov/oiaf/
1605/ggrpt/pdf/0573(2008).pdf ).
Author's personal copy

116 Bhagirath Singh Chauhan et al.

Elad, Y., Pertot, I., 2013. Climate change impacts on plant pathogens and plant disease. In:
Kang, M.S., Banga, S.S. (Eds.), Combating Climate Change: An Agricultural Perspec-
tive. CRC Press, Boca Raton, FL, pp. 183–211.
FAO (Food and Agriculture Organization), 2008. Climate Change and Food Security:
A Framework Document. FAO of the United Nations, Rome, 107 pp.
Farquhar, G.D., 1997. Carbon dioxide and vegetation. Science 278, 1411.
Fuhrer, J., 2003. Agroecosystem responses to combinations of elevated CO2, ozone, and
global climate change. Agric. Ecosyst. Environ. 97, 1–20.
Gupta, R., Gopal, R., Jat, M.L., Jat, R.K., Sidhu, H.S., Minhas, P.S., Malik, R.K.,
Gupta, R., Gopal, R., Jat, M.L., Jat, R.K., Sidhu, H.S., Minhas, P.S., Malik, R.K.,
2010. Wheat productivity in Indo-Gangetic plains of India during 2010: terminal heat
effects and mitigation strategies. PACA Newsletter .
Haskett, J.D., Pachepsky, Y.A., Acock, B., 1997. Increase of CO2 and climate change effects
on Iowa soybean yield, stimulated using GLYSIM. Agron. J. 89, 167–176.
Heong, K.L., Song, Y.H., Pimsamarn, S., Zhang, R., Bae, S.D., 1995. Global warming and
rice arthropod communities. In: Peng, S., Ingram, K.T., Neue, H.U., Ziska, L.H. (Eds.),
Climate Change and Rice. Springer, Berlin, pp. 327–335.
Hibberd, J.M., Whitbread, R., Farrar, J.F., 1996. Effect of elevated concentrations of CO2 on
infection of barley by Erysiphe graminis. Physiol. Mol. Plant Pathol. 48, 37–53.
Hillel, D., Rosenzweig, C., 1989. The Greenhouse Effect and Its Implications Regarding
Global Agriculture. Massachusetts Agricultural Experiment Station, Amherst, MA,
Research Bulletin No. 724.
Hillel, D., Rosenzweig, C., 2011. Climate change and agroecosystems: key issues. In:
Hillel, D., Rosenzweig, C. (Eds.), Handbook of Climate Change and Agroecosystems:
Impacts, Adaptation and Mitigation. Imperial College Press, London, pp. 1–5.
Hingane, L.S., Rupa Kumar, K., Ramana Murty, B.V., 1985. Long term trends of surface air
temperature in India. J. Climatol. 5, 521–528.
Hira, G.S., 2009. Water management in northern states and the food security of India. J. Crop
Improv. 23, 136–157.
Hundal, S.S., Abrol, I.P., 1991. Perspectives of greenhouse gases in climatic change and plant
productivity. In: Abrol, I.P. et al., (Ed.), Indo-US Workshop on Impact of Global Cli-
matic Changes on Photosynthesis and Plant Productivity, New Delhi, India,
pp. 767–779.
Hundal, S.S., Prabhjyot-Kaur, 1996. Climatic change and its impacts on crop productivity in
Punjab, India. In: Abrol, Y.P., Gadgil, S., Pant, G.B. (Eds.), Climatic Variability and
Agriculture. Narosa Publishing House, New Delhi, India, pp. 377–393.
Hundal, S.S., Prabhjyot-Kaur, 2007. Climatic variability and its impact on cereal productiv-
ity in Indian Punjab. Curr. Sci. 92, 506–512.
Iftekharuddaula, K.M., Newaz, M.A., Salam, M.A., Ahmed, H.U., Mahbub, M.A.A.,
Septiningsih, E.M., Collard, B.C.Y., Sanchez, D.L., Pamplona, A.M., Mackill, D.J.,
2011. Rapid and high-precision marker assisted backcrossing to introgress the SUB1
QTL into BR11, the rainfed lowland rice mega variety of Bangladesh. Euphytica
178, 83–97.
IPCC, 1995. Climate Change: A Glossary by the Intergovernmental Panel on Climate
Change. IPPCC, Geneva, Switzerland.
IPCC, 2001. Climate Change 2001: The Scientific Basis. Contributions of Working Group
I to the Third Assessment Report of the Intergovernmental Panel on Climate Change.
Cambridge University Press, Cambridge.
IPCC, 2007a. Contribution of working groups I, II, and III to the fourth assessment report of
the intergovernmental panel on climate change. In: Pachauri, R.K., Reisinger, A. (Eds.),
Climate Change 2007: Synthesis Report. IPCC, Geneva, Switzerland.
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 117

IPCC, 2007b. Summary for policymakers. In: Parry, M.L., Canziani, O.F., Palutikof, J.P.,
van der Linden, P.J., Hanson, C.E. (Eds.), Climate Change 2007: Impacts, Adaptation
and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change. Cambridge University Press,
Cambridge, pp. 7–22.
IPCC. Working Group III, 2007. Mitigation of climate change. In: Metz, B.,
Davidson, O.R., Bosch, P.R., Dave, R., Meyer, L.A. (Eds.), 2007. Fourth Assessment
Report. Cambridge University Press, New York.
Joshi, N.L., Kar, M., 2009. Contingency crop planning for dryland areas in relation to climate
change. Indian J. Agron. 54, 237–243.
Kang, M.S., Banga, S.S., 2013. Global agriculture and climate change: a perspective. In:
Kang, M.S., Banga, S.S. (Eds.), Combating Climate Change: An Agricultural Perspec-
tive. CRC Press, Boca Raton, FL, pp. 11–25.
Kang, W.S., Yun, S.C., Park, E.W., 2010. Nonlinear regression analysis to determine infec-
tion models of Colletotrichum acutatum causing anthracnose of chili pepper using logistic
equation. Plant Pathol. J. 26, 17–24.
Karuppaiah, V., Sujayanad, G.K., 2012. Impact of climate change on population dynamics of
insect pests. World J. Agric. Sci. 8, 240–246.
Kaur, R., Rajni, 2012. Climate change and its possible impacts on agriculture in India. Indian
Farming 62, 10–15, 18.
Khan, S.A., Kumar, S., Hussain, M.Z., Kalra, N., 2009. Climate change, climate variability and
Indian agriculture: impacts, vulnerability and adaptation strategies. In: Singh, S.N. (Ed.),
Climate Change and Crop. Springer-Verlag, Berlin, Heidelberg, pp. 19–38.
Killman, W., 2008. Foreword. In: Climate Change and Food Security: A Framework Doc-
ument. FAO of the United Nations, Rome, p. iii.
Krishnan, P., Swain, D.K., Bhaskar, B.C., Nayak, S.K., Dash, R.N., 2007. Impact of ele-
vated CO2 and temperature on rice yield and methods of adaptation as evaluated by crop
simulation studies. Agr. Ecosyst. Environ. 122, 233–242.
Kumar, S., 2006. Climate change and crop breeding objectives in the twenty first century.
Curr. Sci. 90, 1053–1054.
Kumar, K.S.K., Parikh, J., 2001. Indian agriculture and climate sensitivity. Glob. Environ.
Chang. 11, 147–154.
Kumar, S.N., Aggarwal, P.K., Rani, S., Jain, S., Saxena, R., Chauhan, N., 2011. Impact of
climate change on crop productivity in Western Ghats, coastal and north eastern regions
of India. Curr. Sci. 101, 332–341.
Lal, R., 2004. Soil carbon sequestration impacts on global climate change and food security.
Science 304, 1623–1627.
Lal, M., 2005. Climate change—implications for India’s water resources. In: Mirza, M.M.Q.,
Ahmad, Q.K., Balkema, A.A. (Eds.), Climate Change and Water Resources in South
Asia. Taylor and Francis Group, UK, pp. 15–193.
Lal, M., 2011. Implications of climate change in sustained agricultural productivity in South
Asia. Regional Environ. Change 11, 79–94.
Lal, M., Singh, K.K., Srinivasan, G., Rathore, L.S., Naidu, D., Tripathi, C.N., 1998.
Growth and yield response of soybean in Madhya Pradesh, India to climatic variability
and change. Agr. Forest. Meteorol. 89, 101–114.
Lal, M., Nozawa, T., Emori, S., Harasawa, H., Takahashi, K., Kimoto, M., Abe-Ouchi, A.,
Nakajima, T., Numaguti, A., 2001. Future climate change: implications for Indian sum-
mer monsoon and its variability. Curr. Sci. 81, 1196–1207.
Ma, H.L., Zhu, J.G., Liu, G., Xie, Z.B., Wang, Y.L., Yang, L.X., Zeng, Q., 2007. Avail-
ability of soil nitrogen and phosphorous in a typical rice–wheat rotation system under
elevated atmospheric CO2. Field Crop. Res. 100, 44–51.
Author's personal copy

118 Bhagirath Singh Chauhan et al.

Machida, T., Nakazawa, T., Fujii Yaola, S., Watanabe, O., 1995. Increase in the atmospheric
nitrous oxide concentration during the last 250 years. Geophys. Res. Lett. 22,
2921–2924.
Mahajan, G., Bharaj, T.S., Timsina, J., 2009a. Yield and water productivity of rice as affected
by time of transplanting in Punjab, India. Agric. Water Manage. 96, 525–535.
Mahajan, G., Chauhan, B.S., Johnson, D.E., 2009b. Weed management in aerobic rice in
Northwestern Indo-Gangetic Plains. J. Crop Improv. 23, 366–382.
Mahajan, G., Timsina, J., Singh, K., 2011. Performance and water use efficiency of rice rel-
ative to establishment methods in northwestern Indo-Gangetic Plains. J. Crop Improv.
25, 597–617.
Mahajan, G., Singh, S., Chauhan, B.S., 2012. Impact of climate change on weeds in the rice–
wheat cropping system. Curr. Sci. 102, 1254–1255.
Maheswari, M., Varalaxmi, Y., Vijayalashmi, A., Yadav, S.K., Sharmila, P.,
Venkateswarlu, B., Pardha Saradhi, P., 2010. Metabolic engineering using mtlD gene
enhances tolerance to water deficit and salinity in sorghum. Biol. Plant. 54, 647–652.
Mahi, G.S., 1996. Effect of Climate Change on Simulated Wheat and Rice Yield Under
Punjab Conditions. Punjab Agricultural University, Ludhiana, India.
Manning, W.J., 1995. Climate change: potential effects of increased atmospheric carbon
dioxide (CO2), ozone (O3), and ultraviolet-B (UV-B) radiation on plant diseases. Envi-
ron. Pollut. 88, 219–245.
McElrone, A.J., Reid, C.D., Hoye, K.A., Hart, E., Jackson, R.B., 2005. Elevated CO2
reduces disease incidence and severity of a red maple fungal pathogen via changes in host
physiology and leaf chemistry. Glob. Chang. Biol. 11, 1828–1836.
McElrone, A.J., Hamilton, J.G., Krafnick, A.J., Aldea, M., Knepp, R.G., De Lucia, E.H.,
2010. Combined effects of elevated CO2 and natural climatic variation on leaf spot dis-
eases of redbud and sweetgum trees. Environ. Pollut. 158, 108–114.
Meehl, G.A., Stocker, T.F., Collins, W.D., Friedlingstein, P., Gaye, A.T., Gregory, J.M.,
Kitoh, A., Knutti, R., Murphy, J.M., Noda, A., Raper, S.C.B., Watterson, I.G.,
Weaver, A.J., Zhao, Z.-C., 2007. Global climate projections. In: Solomon, S.,
Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M.,
Miller, H.L. (Eds.), Climate Change 2007: The Physical Science Basis. Contribution
of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel
on Climate Change. Cambridge University Press, Cambridge, UK and New York,
USA, pp. 747–846.
Mittal, S.P., Singh, P., 1989. Intercropping field crops between rows of Leucaena leucocephala
under rainfed conditions in northern India. Agroforestry Syst. 8, 165–172.
Mosier, A.R., Duxbury, J.M., Freney, J.R., Hienemeyer, O., Mnami, K., 1998. Assessing
and mitigating nitrous oxide emissions from agricultural soils. Climate Change 40, 7–38.
Mulchi, C.L., Slaughter, L., Saleem, M., Lee, E.H., Pausch, R., Rowland, R., 1992. Growth
and physiological characteristics of soybean in open-top chambers in response to ozone
and increased atmospheric CO2. Agric. Ecosyst. Environ. 38, 107–118.
Naidu, V.S., Varshney, J.G., 2011. Interactive effect of elevated CO2, drought and weed
competition on carbon isotope discrimination (13C) in wheat (Triticum aestivum) leaves.
Indian J. Agric. Sci. 81, 1026–1029.
National Research Council (US), 2010. Advancing the science of climate change. America’s
Climate Choices: Panel on Advancing the Science of Climate Change, The National
Academies Press, Washington, DC.
Pandey, V., Patel, H.R., Patel, V.J., 2007. Impact assessment of climate change on wheat
yield in Gujarat using CERES-wheat model. J. Agrometeorol. 9, 149–157.
Pangga, I.B., Chakraborty, S., Yates, D., 2004. Canopy size and induced resistance in
Stylosanthes scabra determine anthracnose severity at high CO2. Phytopathology 94,
221–227.
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 119

Parry, M.L., Carter, T.R., 1988. The assessments of the effects of climatic variations on agri-
culture: aims, methods and summary of results. In: Konijn, N.T. (Ed.), The Impact of
Climatic Variations on Agriculture, Vol. 1, Assessments in Cool Temperate and Cold
Regions. Kluwer, Dordrecht, The Netherlands.
Parry, M.L., Rosenzweig, C., Iglesias, A., Livermore, M., Fischer, G., 2004. Effects of cli-
mate change on global food production under SRES emissions and socio-economic sce-
narios. Glob. Environ. Chang. 14, 53–67.
Pathak, H., Wassmann, R., 2009. Quantitative evaluation of climatic variability and risks for
wheat yield in India. Climate Change 93, 157–175.
Patterson, D.T., 1995. Weeds in a changing climate. Weed Sci. 43, 685–701.
Patterson, D.T., Flint, E.P., 1980. Potential effects of global atmospheric CO2 enrichment on
the growth and competitiveness of C3 and C4 weed and crop plants. Weed Sci. 28, 71–75.
Peng, S., Huang, J., Sheehy, J.E., Laza, R.C., Visperas, R.M., Zhong, X., Centeno, G.S.,
Khush, G.S., Cassman, K.G., 2004. Rice yields decline with higher night temperature
from global warming. Proc. Natl. Acad. Sci. U. S. A. 101, 9971–9975.
Pitovranov, S.E., Lakimets, V., Kiselev, V.I., Sirotenko, O.D., 1988. The effects of climatic
variations on agriculture in the subarctic zone of the USSR. In: Parry, M.L.,
Carter, T.R., Konijn, N.T. (Eds.), The Impact of Climatic Variations on Agriculture,
Volume 1, Assessments in Cool Temperate and Cold Regions. Kluwer, Dordrecht,
The Netherlands, pp. 617–722.
Prabhakar, S.V.R.K., Shaw, R., 2008. Climate change adaptation implications for drought
risk mitigation: a perspective for India. Climate Change 88, 113–130.
Prabhjyot-Kaur, Hundal, S.S., 2006. Effect of possible futuristic climate change scenarios on
productivity of some kharif and rabi crops in the central agroclimatic zone of Punjab.
J. Agric. Phys. 6, 21–27.
Prabhjyot-Kaur, Hundal, S.S., 2010. Global climate change vis-à-vis crop productivity. In:
Jha, M.K., Jha, M.K. (Eds.), Natural and Anthropogenic Disasters: Vulnerability, Pre-
paredness and Mitigation. Capital Publishing Company and Springer, New Delhi and
The Netherlands, pp. 413–431.
Prabhjyot-Kaur, Singh, H., Mukherjee, J., Bal, S.K., 2011. Changing rainfall scenarios in
Punjab: a variability trend analysis. In: Proceedings of International Conference on “Pre-
paring Agriculture for Climate Change”. PAU, Ludhiana, India, p. 411.
Prasannakumar, N.R., Chander, S., Pal, M., 2012. Assessment of impact of climate change
with reference to elevated CO2 on rice brown planthopper, Nilaparvata lugens (Stal.) and
crop yield. Curr. Sci. 103, 1201–1205.
Rao, G.G.S.N., 2007. Consolidated report (2004–07) on impact, adaptation, and vulnera-
bility of Indian agriculture to climate change. Network Project on Climate Change.
Indian Council of Agricultural Research, New Delhi, India.
Rao, M.S., Srinivas, K., Vanaja, M., Rao, G.G.S.N., Venkateswarlu, B., Ramakrishna, Y.S.,
2009. Host plant (Ricinus communis Linn.) mediated effects of elevated CO2 on growth
performance of two insect folivores. Curr. Sci. 97, 7–10.
Reddy, K.R., Reddy, V.R., Hodges, H.F., 1992. Temperature effects on early season cotton
growth and development. Agron. J. 84, 229–237.
Reddy, K.R., Koti, S., Davidonis, G.H., Reddy, V.R., 2004. Interactive effects of carbon
dioxide and nitrogen nutrition on cotton growth, development, yield, and fiber quality.
Agron. J. 96, 1148–1157.
Reilly, J., et al., 2001. Agriculture: the potential consequences of climate variability and change
for the United States. In: US National Assessment of the Potential Consequences of Cli-
mate Variability and Change. Press Syndicate of the University of Cambridge, New
York.
Rosenzweig, C., Parry, M.L., 1994. Potential impact of climate change on world food sup-
ply. Nature 367, 133–138.
Author's personal copy

120 Bhagirath Singh Chauhan et al.

Rupa Kumar, K., Krishna Kumar, K., Pant, G.B., Srinivisan, G., 2002. Climate change: the
Indian scenario. In: Background Paper Prepared by FICCI, International Conference on
Science and Technology Capacity Building for Climate Change. New Delhi, India,
pp. 5–17.
Rupakumar, K., Sahai, A.K., Krishna, K., Patwardhan, S.K., Mishra, P.K., Revadekar, J.V.,
Kamala, K., Pant, G.B., 2006. High-resolution climate change scenarios for India for the
21st century. Curr. Sci. 90, 334–345.
Salinger, M.J., 1989. The effects of greenhouse gas warming on forestry and agriculture. Draft
Report for WMO Commission of Agrometeorology.
Sathaye, J., Shukla, P.R., Ravindranath, N.H., 2006. Climate change, sustainable develop-
ment and India: global and national concerns. Curr. Sci. 90, 314–325.
Sen Roy, S., Balling, R.C., 2005. Analysis of trends in maximum and minimum temperature,
diurnal temperature range, and cloud cover over India. Geophys. Res. Lett. 32, L12702.
Seneweera, S.P., Conroy, J.P., 1997. Growth, grain yield and quality of rice (Oryza sativa L.) in
response to elevated CO2 and phosphorus nutrition. Soil Sci. Plant Nutr. 43, 1131–1136.
Septiningsih, E.M., Pamplona, A.M., Sanchez, D.L., Neeraja, C.N., Vergara, G.V.,
Heuer, S., Ismail, A.M., Mackill, D.J., 2009. Development of submergence tolerant rice
cultivars: the Sub1 locus and beyond. Ann. Bot. 103, 151–160.
Sharma, H.C., 2013. Climate change effects on insects: implications for crop protection and
food security. In: Kang, M.S., Banga, S.S. (Eds.), Combating Climate Change: An Agri-
cultural Perspective. CRC Press, Boca Raton, pp. 213–236.
Sharma, S.K., Chauhan, R., 2011. Climate change research initiative: Indian network for
climate change assessment. Curr. Sci. 101, 308–311.
Sharma, V.K., Thind, T.S., Singh, P.P., Mohan, C., Arora, J.K., Raj, P., 2007. Disease-
weather relationships and forecasting of bacterial leaf blight of rice. Plant Dis. Res.
22, 52–56.
Sharma, B.R., Rao, K.V., Vittal, K.P.R., Ramakrishna, Y.S., Amarasinghe, U., 2010. Esti-
mating the potential of rainfed agriculture in India: prospects for water productivity
improvements. Agric. Water Manage. 97, 23–30.
Shin, J.W., Yun, S.C., 2010. CO2 and temperature effects on the incidence of four major
chili pepper diseases. Plant Pathol. J. 26, 178–184.
Sinha, S.K., Swaminathan, M.S., 1991. Deforestation, climate change and sustainable nutri-
tion security: a case study of India. Climate Change 19, 201–209.
Sinha, S.K., Singh, G.B., Rai, M., 1998. Decline in Crop Productivity in Haryana and Pun-
jab: Myth or Reality? Indian Council of Agricultural Research, New Delhi, India.
Sontakke, N.A., 1990. Indian Summer Monsoon Rainfall Variability During the Longest
Instrumental Period 1813–1988. M.Sc. Thesis, University of Poona, Pune, India.
Srivastava, C.P., Joshi, N., Trivedi, T.P., 2010. Forecasting of Helicoverpa armigera population
and impact of climate change. Indian J. Agric. Sci. 80, 3–10.
Staley, J.T., Hodgson, C.J., Mortimer, S.R., Morecroft, M.D., Masters, G.J., Brown, V.K.,
Taylor, M.E., 2007. Effects of summer rainfall manipulations on the abundance and ver-
tical distribution of herbivorous soil macro-invertebrates. Eur. J. Soil Biol. 43, 189–198.
Sultana, S., Khew, C.-Y., Morshed, M.M., Namasivayam, P., Napis, S., Ho, C.L., 2012.
Overexpression of monodehydroascorbate reductase from a mangrove plant
(AeMDHAR) confers salt tolerance on rice. J. Plant Physiol. 169, 311–318.
Swaminathan, M.S., 2012. Remember your Humanity: Pathway to Sustainable Food Secu-
rity. New India Publishing Agency, New Delhi.
Swaminathan, M.S., Kesavan, P.C., 2012. Agricultural research in an era of climate change.
Agric. Res. 1, 3–11.
Tao, L.X., Tan, H.J., Wang, X., Cao, L.Y., Song, J., Cheng, S.H., 2008. Effects of high-
temperature stress on flowering and grain-setting characteristics of Guodao 6. Acta
Agron. Sin. 34, 609–674.
Author's personal copy

Global Warming and Its Possible Impact on Agriculture in India 121

Taub, D.R., Wang, X.Z., 2008. Why are nitrogen concentrations in plant tissues lower
under elevated CO2? A critical examination of the hypotheses. J. Integr. Plant Biol.
50, 1365–1374.
Tobin, P.C., Nagarkatti, S., Loeab, G., Saunders, M.C., 2008. Historical and projected inter-
actions between climate change and insect voltinism in a multivoltine species. Glob.
Chang. Biol. 14, 951–957.
UNEP, 1989. Criteria for assessing vulnerability to sea level rise: a global inventory to high
risk areas. United Nations Environment Programme and the Government of the
Netherlands. Draft Report.
Vanaja, M., Raghuram, R., Lakshmi, N.J., Maheswari, M., Vagheera, P., Ratnakumar, P.,
Jyothi, M., Yadav, S.K., Venkateswarlu, B., 2007. Effect of elevated atmospheric CO2
concentrations on growth and yield of blackgram (Vigna mungo L. Hepper): a rainfed
pulse crop. Plant Soil Environ. 53, 81–88.
Venkateswarlu, B., Shanker, A.K., 2009. Climate change and agriculture: adaptation and
mitigation strategies. Indian J. Agron. 54, 226–230.
Venkateswarlu, B., Singh, A.K., Prasad, Y.G., Ravindra Chary, G., Srinivasarao, C.,
Rao, K.V., Ramana, D.B.V., Rao, V.U.M., 2011. District Level Contingency Plans
for Weather Aberrations in India. Central Research Institute for Dryland Agriculture,
Natural Resource Management Division, Indian Council of Agricultural Research,
Hyderabad, India.
Vu, H.T.T., Manangkil, O., Mori, N., Yoshida, S., Nakamura, C., 2010. Post-germination
seedling vigor under submergence and submergence-induced SUB1A gene expression in
indica and japonica rice (Oryza sativa L.). Aust. J. Crop Sci. 4, 264–272.
Watson, R.T., Zinyowera, M.C., Moss, R.H. (Eds.), 1998. The Regional Impacts of Cli-
mate Change: An Assessment of Vulnerability. Cambridge University Press, Cambridge,
pp. 517, IPCC II report.
World Meteorological Organization, 1992. International Meteorological Vocabulary. sec-
ond ed., WMO, Geneva, Switzerland.
Xu, K., Xu, X., Fukao, T., Canlas, P., Rodriguez, R.M., Heue, S., Ismail, A.M., Bailey-
Serres, J., Ronald, P.C., Mackill, D., 2006. Sub1A is an ethylene-response-factor-like
gene that confers submergence tolerance to rice. Nature 442, 705–708.
Yin, X., Struik, P.C., 2008. Applying modelling experiences from the past to shape system
biology: the need to converge crop physiology and functional genomics. New Phytol.
179, 629–642.
Yoshida, S., Parao, P.T., 1976. Climatic influence on yield and yield components of low land
rice in tropics. In: Proc. Symposium on Climate and Rice. International Rice Research
Institute, Manila, Philippines, pp. 471–494.
Ziska, L.H., 2003. Evaluation of the growth response of six invasive species to past, present
and future atmospheric carbon dioxide. J. Exp. Bot. 54, 395–404.
Ziska, L.H., Goins, E.W., 2006. Elevated atmospheric carbon dioxide and weed populations
in glyphosate treated soybean. Crop. Sci. 46, 1354–1359.
Ziska, L.H., Teasdale, J.R., 2000. Sustained growth and increased tolerance to glyphosate
observed in a C3 perennial weed, quackgrass (Elytrigia repens), grown at elevated carbon
dioxide. Aust. J. Plant Physiol. 27, 159–166.
Ziska, L.H., Namuco, O., Moya, T., Quilang, J., 1997. Growth and yield response of field-
grown tropical rice to increasing carbon dioxide and air temperature. Agron. J. 89,
45–53.
Ziska, L.H., Tomecek, M.B., Gealy, D.R., 2010. Competitive interactions between culti-
vated and red rice as a function of recent and projected increases in atmospheric carbon
dioxide. Agron. J. 102, 118–123.

View publication stats

You might also like