You are on page 1of 52

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

net/publication/241749068

Management of Soil Salinity and Alkalinity Problems in India

Article  in  Journal of Crop Production · October 2008


DOI: 10.1300/J144v07n01_07

CITATIONS READS
16 2,715

2 authors, including:

P. S. Minhas
National Institute of Abiotic Stress Management
215 PUBLICATIONS   2,722 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

ICAR-NIASM View project

Use of poor quality water in agriculture View project

All content following this page was uploaded by P. S. Minhas on 07 August 2014.

The user has requested enhancement of the downloaded file.


This article was downloaded by: [National Institute of Abiotic Stress Management]
On: 24 September 2013, At: 21:59
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Journal of Crop Production


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/wzcp20

Management of Soil Salinity and Alkalinity Problems in


India
a b
P. S. Minhas & O. P. Sharma
a
Coordinating Unit, Central Soil Salinity Research Institute, Karnal, 132001, India
b
College of Agriculture, Indore, 452001, India
Published online: 15 Oct 2008.

To cite this article: P. S. Minhas & O. P. Sharma (2003) Management of Soil Salinity and Alkalinity Problems in India, Journal of
Crop Production, 7:1-2, 181-230

To link to this article: http://dx.doi.org/10.1300/J144v07n01_07

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Management of Soil Salinity
and Alkalinity Problems in India
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

P. S. Minhas
O. P. Sharma

SUMMARY. Like the problem of salt-affected soils afflicting many


other countries with arid and semi-arid climates, these lands occupy
about 8.6 M ha in India out of which 3.5 M ha are in canal commands,
commonly referred to as man made or wet deserts. Starting from the
early systematic work of Dr. Leather about a century ago, real impetus to
development and application of reclamation technologies has occurred
in the post independence period, especially the seventies. Salt-affected
soils are grouped into saline and alkali/sodic soils on the considerations
of soil management and crop responses. Out of the 3.4 M ha of alkali
lands in Indo-Gangetic Plains, about 1.0 M ha has been reclaimed by the
hydro-chemical technology. Application of amendments like gypsum,
equivalent to 50% gypsum requirement, to the surface 15-cm soil only
was enough to start cropping with rice. Light and frequent irrigation are
ideal for upland crops. Application of higher doses of nitrogen in splits to
compensate for volatilization losses, organic matter additions through
green manure to increase reclaiming action, skipping phosphorus appli-
cation in the initial years and the application of zinc to each crop were
some of the emerging recommendations. Application of low-grade py-

P. S. Minhas is affiliated with the Coordinating Unit, Central Soil Salinity Research
Institute, Karnal 132001, India. O. P. Sharma is affiliated with the College of Agricul-
ture, Indore 452001, India.
[Haworth co-indexing entry note]: “Management of Soil Salinity and Alkalinity Problems in India.”
Minhas, P. S., and O. P. Sharma. Co-published simultaneously in Journal of Crop Production (Food Products
Press, an imprint of The Haworth Press, Inc.) Vol. 7, No. 1/2 (#13/14), 2003, pp. 181-230; and: Crop Produc-
tion in Saline Environments: Global and Integrative Perspectives (ed: Sham S. Goyal, Surinder K. Sharma,
and D. William Rains) Food Products Press, an imprint of The Haworth Press, Inc., 2003, pp. 181-230. Single
or multiple copies of this article are available for a fee from The Haworth Document Delivery Service
[1-800-HAWORTH, 9:00 a.m. - 5:00 p.m. (EST). E-mail address: docdelivery@haworthpress.com].

http://www.haworthpress.com/store/product.asp?sku=J144
 2003 by The Haworth Press, Inc. All rights reserved.
10.1300/J144v07n01_07 181
182 CROP PRODUCTION IN SALINE ENVIRONMENTS

rites also gave encouraging results in calcareous soils. Feasibility of


drainage has been successfully demonstrated in 1980s for the quicker re-
habilitation of saline and waterlogged soils. Provision for subsurface
drains at 1.5-2.0 m depth and 50-75 m spacing in alluvial soils and 12-24
m interval in vertisol was sufficient for facilitating growing crops within
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

2-3 years on lands lying barren for considerable period. However, tech-
nological, economic, social, and political considerations are the major
bottlenecks for large-scale amelioration of saline soils, in addition to the
safer disposal of drainage waters. Strategies worked out for successful
crop production on saline soils after drainage include initial leaching
management (minimizing water requirement for leaching by synchroniz-
ing with monsoon rains, etc.), proper selection of crops/cultivars, irrigation
(method, controlled frequency for enhancing water-table contribution,
canal water use during the initial stages including presowing for con-
junctive use with drainage waters) and cultural practices (furrow plant-
ing, increasing seed rate and fertility management). Some of the future
issues for combating salinity and also preventing further land degrada-
tions in India are also highlighted. [Article copies available for a fee from
The Haworth Document Delivery Service: 1-800-HAWORTH. E-mail address:
<docdelivery@haworthpress.com> Website: <http://www.HaworthPress.com>
 2003 by The Haworth Press, Inc. All rights reserved.]

KEYWORDS. Indo-Gangetic Plains, gypsum, rice, land degradation

HISTORIC PERSPECTIVE

The arid and semi-arid climate associated with topographic features and
groundwater hydrology is often responsible for in situ accumulation or trans-
port and deposit of salts in other places and manifestation of saline and alkali
characters in the soils. These soils become unproductive when the accumula-
tion of salts is beyond a certain level so that many crops fail to survive. Such
soils form an important ecological entity in India. Coastal soils, soils receiving
saline seeps and saline springs, basin lands, salty alluvial deposits, and lower
level depressions have been affected by salinity. Thus the problem of contem-
porary salinity predates human civilization.
In India, tank irrigation and well irrigation became prevalent during the
Aryan Civilization Period (1500 BC). It is during this and the Vedic period that
cultivable lands were distinguished between fertile (Urvara) and infertile
(Anurvara) on the basis of productivity. The unproductive barren lands were
called Ushtra and the salt affected lands were called Usara. Medieval scripts
have used Usar as also Kallar and other terms for salt-affected lands.
P. S. Minhas and O. P. Sharma 183

Systematic inventory of such soils appeared in the British Period either in


form of tour diaries, official notes or through public litigation in the middle of
19th century. Earliest reference to salt lands (known as Kallar, Usar or Reh)
was made by Sleeman who passed through Usar plains of south Oudh in
United Province (now Uttar Pradesh) around 1850 (Moreland, 1901). He also
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

gave an indigenous account of reclamation of Usar by flooding for two or


three years, cross plowing, manuring, and irrigating. If these soils are left fal-
low after reclamation then they can revert back to their original status. The
salt-affected soils were in existence before introduction of any canal irrigation
in Punjab of Indus Plains. Jameson (1852) noticed a belt of 16 km ⫻ 4.8 km
having soil salinity problem in Kapurthala state. Later, it was mentioned in his
official notes that the “whole of the country in Punjab lying west of Jamuna is
impregnated with the elements of Reh (local term for soils having dominant
sodium carbonate and bicarbonate salts).” These elements of Reh existed in the
soil and the water in proportions varying from traces to an extent, which is ab-
solutely injurious to all vegetative life (Whitecombe, 1971). Bhargava et al.
(1981) opined that these soils are generally characterized with 20-30 cm de-
pressions. Release of salts in the Shivalik piedmont zone and the plains
through alkali hydrolysis of alumino-silicate minerals and their repeated con-
vergence in depressions lead to the genesis of these sodic soils at the surface,
even though good quality ground water occurs in the substratum.
The British realigned the Western Jamuna Canal (which was constructed in
1356 by Firozshah Tughlak) in Punjab (Haryana) in 1839 and constructed
Eastern Yamuna Canal in 1830, Ganga Canal in 1855 and Lower Ganga Canal
in 1878 for irrigation in Western United Province, Uttar Pradesh (U.P.). An in-
digo planter from Sikandara Rau village in Aligarh lodged early complaints
about development of salinity to British rulers from Munak area of Karnal near
the Western Jamuna Canal in 1855 and in 1876. This led to the formation of
Reh committee in 1877 to investigate the causes of the formation and to sug-
gest probable ways to check the problem and reclamation of such lands. The
members of Reh committee in their report in 1879 could not agree on the the-
ory of alkali formation in soils (Singh, 1998a). However, the committee sug-
gested a series of experiments for reclaiming the lands for profitable cultivation
by (1) removal of salts, (2) drainage, (3) silting, (4) deep cultivation, (5) ma-
nuring, and (6) plowing of green crops. Several experiments were planned and
conducted in Aligarh and Kanpur districts of U.P. on surface and subsurface
drainage, scrapping of Reh, use of farmyard manure (FYM) along with initial
leaching of salts and planting of trees. The results of these experiments were
not supported by soil chemical analysis; hence it was difficult to form an opin-
ion about the application of the results to other areas (Moreland, 1901).
In the peninsular India, Nira Valley Irrigation Canal was opened in 1884 to
irrigate deep black soils on a large scale. The scars of seepage and percolation
184 CROP PRODUCTION IN SALINE ENVIRONMENTS

of water from canals appeared within 5 years of irrigation causing secondary


salinization in the area where existence of salt land had not been recorded in
the neighborhood. It was estimated that about 6-7% of the area was being dam-
aged annually (Mann and Tamhane, 1910).
The extent of damage due to canal irrigation caused anxiety to both the
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

farmers and the government. The first systematic attempt to study the vagaries
of the problems and the causes was initiated by an Imperial Chemist named
A.W. Leather. Leather (1906) recommended application of gypsum along
with heavy manuring for reclamation of Reh. He calculated gypsum require-
ment on the basis of soluble carbonates to be neutralized. The quantity of gyp-
sum to be used was, therefore, very high and the results did not find large-scale
application. Proper scientific approach for reclamation of alkali soils was not
known in this period, as the principles of cation exchange were not evolved.
Experiments for reducing the losses due to waterlogging and poor drainage
were also conducted in Northern and Peninsular India, and these were quite
successful (Leather, 1914; Mackenna, 1918). Nasir (1923) suggested drainage
and use of gypsum followed by flooding to remove excessive salts from such
soils. A comparison of salt lands in Deccan and Sind (Tamhane, 1920) re-
vealed that sodium carbonate, which appears to be the main constituent of
Kallar in alluvial soils, was not noticed in case of black Kallar. Deccan soils
contained sodium sulphate predominantly as compared to Sind soils, which
contained sodium chloride. Tamhane (1920) also listed the limits of salt toler-
ance of various crops. The intensity of secondary salinization varied in the area
depending on topographical situation and distance from canal. The rate of sec-
ondary salinization in the irrigated area around Etah district of U.P. (Gangetic
alluvium), was found to vary from 1 to 208% (Agarwal et al., 1957) in a period
of 40 years (1912 to 1952) within 400 m distance from the Ganga canal.
The work of Gedroiz (1916) on the base exchange of soils led to a better un-
derstanding of the problems of soil salinity and sodicity while de’Sigmond
(1927) and Kelly (1927) established the basic principles of alkali soil recla-
mation. The chemical reactions occurring during the reclamation, formed
soluble sodium salts (Na2SO4) which must be leached into the lower zones
(de’Sigmond, 1927). Kelley and Arany (1928) and de’Sigmond (1932) later
revealed that for replacement of exchangeable sodium with calcium, the
amendments like gypsum, sulphur, sulphuric acid, iron sulphate and alumi-
num sulphate could be useful. During this period, the Indian scientists were
also busy in comparing the efficacy of different amendments for alkali soils.
Tamhane and Krishnan (1930) studied the percolation rate of alkali soils at
Sarkand (Sind-Pakistan) with different calcium salts. Calcium chloride, though
more effective than CaSO4, was considered prohibitive due to high cost. Singh
and Nijhawan (1932), and Puri (1934) also compared CaCl2 and gypsum for
reducing exchangeable sodium. Basu and Tagare (1943) employed sulphur,
gypsum and FYM in black smectitic soils. Talati (1947) conducted field exper-
P. S. Minhas and O. P. Sharma 185

iments on the reclamation and crop management of the alkali soils in Baramati
region of the Bombay State. Agarwal and Mehrotra (1953) used gypsum and
green manure crop dhaincha (Sesbania aculeata) in the reclamation of saline
alkali soils of U. P. Use of gypsum coupled with adequate provision of leach-
ing and flushing should constitute the basic treatments for reclamation of sa-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

line-alkali soils of Indo-Gangetic alluvium (Yadav and Aggarwal, 1959).


According to these workers, paddy should be taken invariably as the first crop
preferably after burying the green manure crop of dhaincha and an alkali toler-
ant crop like barley should follow paddy in rabi (winter) season.
In the second half of the 20th century after independence of India, several
major and medium irrigation projects were rapidly commissioned in different
states and these have resulted in the largest irrigated area in the world. In the
most of irrigation projects, the faulty system of irrigation and improper
on-farm-management practices have aggravated the problem of rise in water
table and consequent degradation of soils through secondary salinization. The
problem is, however, more pronounced in arid and semi-arid regions. Accord-
ing to estimates, every year about 3% of the canal irrigated land in India either
becomes waterlogged or salinized (Singh, 1998a).
After independence, the government of India started the movement of
self-sufficiency in food and agriculture and a serious attempt was made to uti-
lize these salty wastelands for successful crop production. For studying the
problem and reclamation of salt-affected soils, the Indian Council of Agricul-
tural Research, New Delhi, sanctioned the schemes for Punjab (1958-1961),
Rajasthan (1960-1963), Andhra Pradesh (1955-1960), and Madras (1959-1961).
Since these schemes were sponsored for shorter periods, a permanent solution
to the problem could not be achieved. With a broad outlook on this problem,
the Indian Council of Agricultural Research, New Delhi, started All India
Co-Ordinated Research Projects (AICRP), with centers in different State Agri-
cultural Universities (SAU) and established the Central Soil Salinity Research
Institute, Karnal, in 1969. The main objectives were to carry out systematic
studies on the nature, extent and characteristics, reclamation, and management
of salt-affected soils occurring in different geographical and farming situations
in the country so that these problematic soils can be brought to productive
phase. The research work from the Institute and the AICRP centers gave new
dimensions to the solution of problems in Indian context. Different compo-
nents of reclamation technology emerging out of the research efforts are dealt
herein.

NOMENCLATURE AND EXTENT


Though reclamation research and practices have a long history in India,
there have been divergent opinions on various aspects–ranging from nomen-
186 CROP PRODUCTION IN SALINE ENVIRONMENTS

clature to the best means of restoring their productivity. Instead of three cate-
gories of salt-affected soils; saline, alkali, and saline alkali as classified by
Richards (1954) in India soils were classified into former two groups. The ba-
sis was the nature of plant response to the presence of salt and the management
practices desired for reclamation (Bhumbla, 1977; Abrol and Bhumbla, 1978).
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Due to the presence of excess salts (Electrical Conductivity, saturation paste


extract, ECe > 4 dS m⫺1), low exchangeable sodium (Exchangeable Sodium
Percentage, ESP < 15) and saturation paste pH (ⱕ 8.5), saline soils remain
flocculated and have permeability equal to or higher than that of normal soils.
The alkali soils have ECe < 4 dS m⫺1 and ESP > 15 and pH > 8.5. It was later
suggested that distinguishable pH for alkali soils should be 8.2, rather than
8.5 (Abrol et al., 1980; Gupta and Abrol, 1989) suggested. They also pointed
out that these soils contain soluble bicarbonates and carbonates such that
Na/[Cl+SO4] > 1. Moreover, the value of 15 ESP to distinguish alkali soil from
non-alkali soil has been considered too high in alkali smectitic soils (Talati et
al., 1959; Northcote and Skene, 1972; Rao and Sheshachalam, 1976; Gupta and
Verma, 1983). The threshold ESP for these swell-shrink clay soils (Vertisols)
lies between 6-10 and thus ESP value of 8 has been observed to be more appro-
priate for categorizing alkali soils. Plant growth in saline soils is impaired
chiefly because of osmotic effects of excess salts and nutritional imbalances,
whereas in alkali soils it is affected chiefly by the poor physical properties
(Acharya and Abrol, 1998) and toxicity of sodium (Sharma, 1987) and carbon-
ate. High ESP causes dispersion of soil colloids, especially at low electrolytic
concentrations resulting in pore blockage. Thus from the management point of
view, saline soils require leaching of salts to a level below the threshold limits
of crops and alkali soils need amendments like gypsum before being leached.
Despite the simplistic criteria, it is important to recognize limitations. In fact,
the salinity requirements for maintaining stability conditions tend to increase
with increasing ESP, i.e., there is salinity-sodicity continuum instead of single
point (ESP 15 or pH 8.5), which demarcates saline and alkaline conditions and
the guidelines for their reclamation and management. Perhaps due to lack of
data under diverse conditions, the category “saline-alkali” was ignored but
later research efforts showed that salt-affected soils occurring in the geograph-
ical transition zones of saline and alkali soils, heavy black soils, and the soils
where inherent calcium is exhausted due to irrigation with waters having resid-
ual alkalinity or high SAR tend to become alkaline (> 8.5) upon leaching. Thus
these soils are prone to dispersion upon leaching and their permeability be-
comes poor especially during monsoon rains. The crops grown on such soils
show favorable response to gypsum application, and meet the criteria for clas-
sification as “saline-alkali” soils.
Conventional soil-survey methods have led to varying estimates on the ex-
tent of occurrence of salt-affected soils in India. Amongst these, the most re-
P. S. Minhas and O. P. Sharma 187

cent estimate is by Singh (1992) where it is estimated that nearly 8.5 million
hectares of land are affected by the menace of salinity (Table 1). The quantum
of inland waterlogged-saline lands is fluctuating, which is mainly related to
monsoonal rains. Indo-Gangetic Plains, the arid and semi-arid desert region,
the black cotton region, and the coastal regions, all have a fair share of such
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

lands. The geographic distribution of these lands as well as the mode of forma-
tion and characterization varies in these regions. These along with the evolved
reclamation technologies are discussed in the following sections.

Characteristics of Alkali Soils

Alluvial alkali soils: These soils are located mainly in the semi-arid tracts of
Indo-Gangetic Plains of Punjab, Haryana, and Uttar Pradesh. These soils have
developed under impeded drainage due to nearly impermeable clayey sub-
soils usually underlain by hard-indurated caliche (calcium carbonate concre-
tions) pan with a high fluctuating ground water table (Bhargava, 1977). The al-
luvial sodic soils also occur in sporadic patches in arid tracts of Rajasthan
(Mehta et al., 1969) and Gujarat (Dubey et al. 1995).
The salient features of these soils include: very high soil pH (10.8) and high
ESP (as high as 90), soluble salt content (ECe) in upper part (up to 70 cm) of
the pedon ranges from 4 to 24 dSm⫺1 (Table 2, Bhargava, 1977; Acharya and
Abrol, 1978; Yadav, 1981; Tiwari and Sharma, 1989). The principal anions
are CO32⫺ and HCO3⫺ whereas Na+ is the dominant cation. The surface hori-
zon usually has a platy structure and the soils have very low water transmission
properties and are highly dispersed (Acharya and Abrol, 1978, 1991, 1998).
Usually, there is a zone of clay accumulation in the subsurface horizons along
with the presence of argillans. The bulk density of this zone is quite high.
Natric horizon is present in many soils (Bhargava, 1977; Tiwari and Sharma,
1989). The natric horizon generally overlies a zone of calcic or petrocalcic ho-
rizon indurated from a few cm to almost 1 m thick. The zone of CaCO3 accu-
mulation lies within the zone of a fluctuating water table. The subsurface
horizon overlying calcic or petrocalcic horizon usually possesses distinct yel-
lowish-brown iron mottles along with moderate to abundant amounts of
ferro-manganese concretions. The water table may appear nearly just below
the surface during monsoon but recedes to about 5 m depth in subsequent peri-
ods. The ground waters in these areas usually have low salt content comprising
mainly of Na+ and HCO3⫺ ions. A number of hypotheses have been advanced
to account for the origin of alkali salts, e.g., transportation or in situ silicate
mineral weathering, salt springs, remnants of an inland sea. However, none are
able to explain the content and distribution of salts in soil profiles (Dhir, 1998;
Raj-Kumar, 1998).
188 CROP PRODUCTION IN SALINE ENVIRONMENTS

TABLE 1. Extent and distribution of salt-affected soils in India.

State Area Affected Rainfall Soil Characteristics


(‘000 ha) (mm)
Indo-Gangetic alluvial region–alkaline zone
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Punjab 347 500-1000 High pH, medium EC and presence of


Haryana 180 500-1000 CaCO3 at some depth, conspicuous
Uttar Pradesh 1100 500-1000 presence of carbonates in soil solution,
Bihar 104 1000-1400 often presence of restricting layers
Rajasthan 132 500-1000

Indo-Gangetic alluvial region–saline zone


Punjab 173 300-500 Neutral to alkaline pH, medium to high
Haryana 275 300-500 EC, preponderance of chlorides and
Uttar Pradesh 195 450-600 sulfates of sodium, free carbonates
Bihar 504 300-500 absent
Rajasthan 207

Red and grey desert soil region–saline zone


Gujarat 842 500-630 Neutral to alkaline pH, high EC,
Rajasthan 255 160-640 preponderance of chlorides and
sulfates, Mg:Ca often > 1

Medium and deep black soil region–alkaline/saline zone


Gujarat 70 600-1500 Neutral to alkaline pH, medium to high
Alk/Sal EC chlorides and sulfates predominate,
Madhya Pradesh 164/78 presence of smectites, high CEC,
Maharashtra 59/475 calcareous
Tamil Nadu 4/328
Karnataka 76/328
Andhra Pradesh 193/2
Rajasthan 97
(mixed red and black)

Coastal saline soils region


West Bengal 982 1400-1600 Neutral to slightly alkaline pH,
Gujarat 302 300-1650 med-high salt concentration, prone
Orissa 404 1250-1570 to monsoon and tidal flooding
Andhra Pradesh 176 830-1140
Tamil Nadu 204 600-1250
Maharashtra 64 2000-2500
Goa 18 2000-2500

Acid sulphate soils of humid tropics


Kerala 35 2000-3000 Acidic pH, presence of humic pyritic
West Bengal Patches 1500-1600 horizon, preponderance of chlorides
Andaman and Patches 3200 and sulfates
Nicobar Islands
Adapted from: Singh, N.T. (1992)
P. S. Minhas and O. P. Sharma 189

TABLE 2. Some characteristics of salt-affected soils representing different


broad groups.

Depth pH2 ECe CaCO3 Org. C Clay ESP


(cm) (dS m⫺1) (g kg⫺1) (g kg⫺1) (%)
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Aquic Natrustalf (Karnal, Haryana)1


0-10 10.6 22.3 51 1.4 12.5 96
10-48 10.2 6.3 89 1.0 18.9 91
48-76 9.8 4.2 94 0.7 22.7 88
76-104 9.5 2.3 126 0.5 21.2 85
104-163 9.6 1.3 138 0.4 31.8 69

Typic Natrustalf (Akbarpur, Kanpur)2


0-20 10.0 60.5 46 3.0 27.4 75
20-64 9.9 4.7 48 2.8 37.0 74
64-97 9.3 0.9 12 2.1 38.6 33
97-121 8.5 0.6 56 1.7 33.2 18
121-180 8.4 0.4 100 1.6 25.6 10

Typic Haplusterts (Barwaha, Madhya Pradesh)2


0-20 8.5 8.9 75 6.3 42.8 58
20-44 8.5 8.9 115 4.3 53.4 74
44-84 8.6 5.9 145 4.3 44.2 74
84-106 8.6 4.8 190 2.4 39.6 63
106-145 8.6 2.3 190 1.2 37.5 63
145-166 8.5 1.6 200 0-6 35.0 45

Typic Ustochrept (Sampla, Haryana)1


0-19 6.8 65.2 09 - 16.5 3
19-38 7.2 6.9 04 - 17.2 3
38-79 7.3 4.2 18 - 18.2 4
79-119 7.4 3.4 38 - 16.5 3
119-147 7.4 4.0 20 - 17.2 3

Typic Natrargrids (Banaskantha, Gujarat)2


0-10 8.0 68.9 37 3.2 15.8 45
10-23 8.2 44.3 65 3.0 21.9 41
23-46 8.5 32.8 56 3.0 22.1 43
46-71 8.2 29.5 130 2.5 23.3 63
71-87 8.1 25.5 223 2.1 17.2 52
87-112 8.0 20.2 233 2.1 17.6 57
112-140 8.1 30.1 364 1.5 22.5 43
140-170 7.6 32.7 279 1.2 16.1 62
Adapted from: 1Bhargava (1980); 2Dubey et al. (1998)
190 CROP PRODUCTION IN SALINE ENVIRONMENTS

Alkali vertisols: Basu and Tagare (1943) suggested that it is the basic nature
of the parent material, aridity of the climate, proximity to sub-soil water, poor
drainage conditions, and topography that are responsible for the development
of sodic problems in Vertisols and associated soils. Most of the areas under
black soils fall in semi-arid tropics with low leaching intensity and alternate
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

wet and dry seasons. Thus the climatic conditions are quite favorable for
buildup of salts in the root zone to a level detrimental to normal plant growth
particularly under restricted drainage conditions. These soils are potentially
saline/alkali in the compacted sub-surface horizon that can be determined from
physico-chemical analysis of some benchmark pedons as belonging to the or-
der Vertisol (Table 2). The soil texture varies from clay loam to clay through-
out the profile depth. The soils are deep, calcareous and the calcite content
increases with depth. The organic carbon content is low and decreases with
depth. The structure on the surface is weak, fine, sub-angular blocky, whereas
coarse prismatic breaking to moderate, medium sub-angular/angular blocky in
the lower horizons (Bhargava, 1977; Sharma, 1990a). These soils swell on
wetting and crack on drying due to predominance of smectite (Sharma,
1990a). Sodicity (ESP) levels above 10 lead to severe structural degradation
(Gupta and Verma, 1983) due to high degree of clay dispersion. The dispersed
clays clog the pores and induce increased water retention at all suctions.
With increasing ESP, the rate of drying front movement declines and mois-
ture changes in the lower layers are much slower (Acharya and Abrol, 1978;
Gupta and Verma, 1984). With higher water retention deep cracks do not de-
velop in sodic Vertisols (Sharma, 1990a). The effects of soil ESP on the crack-
ing pattern (Table 3) indicate that above an ESP of 10 the soil cracks do not
develop and would not qualify as a Vertisol (Verma and Sharma, 1998a). The
above physico-chemical characteristics render salt leaching rather difficult un-
less their physical conditions are first improved with the addition of suitable
amendments (Yadav, 1981). The crops grown on alkali soils suffer mainly on
account of poor aeration of the root zone, reduced moisture availability, crust-
ing on drying, hindrance in infiltration of irrigation water, poor nutrient avail-
ability, and Na+ toxicity.

RECLAMATION OF ALKALI SOILS

Chemical Reclamation

The replacement of exchangeable Na+ with Ca2+ requires the application of


amendments, which can either supply soluble calcium ions directly or increase
its solubility from the soil constituents.
P. S. Minhas and O. P. Sharma 191

TABLE 3. Cracking behavior of black clay soils as influenced by alkalinity


levels.

Soil ESP Depth of crack (cm) Width of crack (cm) No. of flakes (m⫺2)
6 >90 5-6 -
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

10 50 2-3 2-5
15 30 1-2 5-10
22 10 0.5-1.0 10-30
38 <2 0.2-0.3 30-50
58 0.1-0.2 0.05-0.10 80-100
> 60 Negligible Absent Nil
Adapted from: Verma and Sharma (1998)

Gypsum

As evidenced from the historical background, gypsum was considered as


the most popular, cheap, and easily available amendment in India in the past
and still remains to be the first choice in the Northwestern and Central India.
The resources of mined gypsum are estimated to be more than 1,000 million
Mg, which occur mainly in the arid regions of Rajasthan and Gujarat. The in-
dustrial byproduct phosphogypsum available annually from phosphatic fertil-
izer industry is around 1.5 million Mg. The annual availability of gypsum from
mined and industrial sources is around 2.5 million Mg, which is sufficient to
meet the demand for agricultural use (Singh, 1998b). Considerable research
information is available on the use of gypsum for amending alkali soils in the
country (Kanwar et al., 1965; Singh et al., 1969; Bhumbla and Abrol, 1972;
Abrol et al., 1973). The results reported earlier from different parts of the coun-
try have been critically examined in several review articles (Singh et al., 1981,
Yadav, 1981, Gupta and Abrol, 1989, Singh, 1998b; Sharma et al., 1998).
Application rate and frequency: For the alluvial illitic soils, an application
of 25-50 percent of the estimated gypsum requirement (GR), as determined by
Schoonover’s method, is sufficient for obtaining good returns from the first
rice crop (Singh et al., 1969; Abrol et al., 1973). Obviously, near normal yields
of first rice crop after the application of gypsum can be obtained even in soils
having an ESP of 55 (Verma and Abrol, 1980). However, for wheat crop, ap-
plication of gypsum at a minimum of 50% GR is necessary (Abrol and
Bhumbla, 1979). Abrol et al. (1973) reported a graphical relationship between
the soil pH and the gypsum requirement for use in field situations.
Due to resource constraints, most of the farmers owning alkali lands are un-
able to apply the recommended doses of gypsum (50% GR). The quantity of
gypsum is still higher in case of black smectitic alkali soils having higher CEC,
192 CROP PRODUCTION IN SALINE ENVIRONMENTS

as compared with alluvial soils with similar ESP. Hence, the sole application
of the required quantity of gypsum becomes fairly expensive. Experiments
were conducted on the possibilities of splitting the amendment cost over the
years. The sole application of gypsum at 50% GR proved better than applied in
two split applications (25% in first year + 25% in second year) in alkali loam
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

soil (pH 9.9-10.3, ESP 73) of Gangetic alluvium plains (Tiwari and Sharma,
1989). The yields, decrease in pH, ECe, and ESP were highest (Table 4) under
split application (50% GR in first year + 25% GR in second year) and were
higher than a sole application of 75% GR in the first year. However, sole appli-
cation of gypsum at 75% GR in the first year produced the maximum cumula-

TABLE 4. Effect of frequency of gypsum application on grain yield (Mg ha⫺1) of


rice and wheat in alkali loam and black clay soil.

Gyp. (%GR*) Rice Wheat Rice Wheat Rice Wheat Total


1st 2nd 3rd yr 1981 1981-82 1982 1982-83 1983 1983-84
Loamy alkali soil (Kanpur, U.P.)1
_ _ _ 0.28 0.26 1.23 0.32 0.80 0.53 3.43
25 _ _ 2.19 1.57 2.10 1.41 1.01 1.48 10.40
50 _ _ 2.66 1.97 3.06 1.77 2.25 1.88 10.38
25 25 _ 2.07 1.46 2.98 1.14 2.43 2.00 12.78
75 _ _ 2.47 2.04 2.84 1.91 2.30 2.86 13.76
25 50 _ 2.00 1.63 3.25 2.01 2.50 2.17 13.56
50 25 _ 2.60 2.08 3.03 2.12 2.44 2.21 14.49
25 25 25 2.15 1.52 2.92 1.80 2.77 2.27 13.38
LSD (p = 0.05) 0.21 0.13 0.26 0.33 0.20 0.13
Black alkali clay soil (Barwaha, M.P.)2
1983 1983-84 1984 1984-85 1985 1985-86 Total
_ _ _ 0.16 0.41 0.07 0.57 0.29 0.47 1.99
25 _ _ 1.13 0.84 0.61 0.95 0.85 0.91 5.29
50 _ _ 5.51 1.53 0.93 1.55 2.00 1.46 12.97
25 25 _ 1.08 0.86 0.84 1.38 1.47 1.27 6.90
75 _ _ 3.65 2.34 1.35 2.43 2.75 2.16 14.68
25 50 _ 1.10 0.85 1.15 1.85 2.30 1.84 9.10
50 25 _ 2.60 1.78 1.22 2.02 2.35 1.87 11.83
25 25 25 1.32 0.86 0.79 1.55 2.67 1.57 8.75
LSD (p = 0.05) 0.13 0.14 0.06 0.10 0.13 0.07
*Gypsum requirement
Adapted from: 1Tiwari and Sharma (1989); 2Sharma (1990)
P. S. Minhas and O. P. Sharma 193

tive yield of rice and wheat over a period of three years (1983-84 to 1985-86)
in the alkali black clay soil (Sharma, 1990). Gupta et al. (1988) observed that
gypsum use efficiency was maximum when gypsum was applied at 25% GR.
Gypsum at 50 and 75% GR gave maximal efficiency after third and fourth
cropping seasons suggesting that much of the gypsum remained undissolved
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

during the first year. These results indicate that in loamy soil of Gangetic allu-
vium and in smectitic clay soil, gypsum at 50% GR and 75% GR must be ap-
plied, respectively, in the first year to achieve maximum financial returns.
Methods of gypsum application: Apart from dose, the depth of mixing the
amendment is as important since it influences the ability of applied gypsum to
provide soluble Ca2+ to displace Na+ and maintain infiltration rates. Elgably
(1971) reported that gypsum mixed with surface 15 cm soil removed ex-
changeable sodium more effectively than did gypsum applied to surface. For
soils containing soluble carbonates, mixing limited quantities of gypsum in
shallower depths was more beneficial than mixing in greater depths (Khosla et
al., 1973). Mixing gypsum in greater depths diluted it and resulted in a smaller
decrease in ESP throughout the depth. It is also likely that a fraction of gypsum
may have been used in neutralizing soluble CO32⫺ at the expense of exchange-
able Na+. Surface application of gypsum is as effective as an admixture with
the surface soil when reclamation is initiated with rice as the first crop (Hira,
1998). The rice seedlings could be transplanted after surface application of
gypsum on the pre-flooded fields (Singh et al., 1981).
Experiments in highly sodic alluvial soils show that shallow mixing of gyp-
sum up to 10 cm is superior to its deeper (20 cm) application (Tiwari and
Sharma, 1989). Nevertheless, the depth of reclamation seldom exceeds 20 cm
depth in black alkali soils and this restricts root proliferation and thus choice of
crops. The depth of reclamation can be increased either by mixing the gypsum
to deeper layers or through subsoil placement, but this involved additional
quantities of gypsum as well as labor. When gypsum was applied either in
strips of 60 cm wide and 30 cm deep or 50 cm wide and 35 cm deep leaving 30
cm and 40 cm strips in between, it increased the cotton yield and root penetra-
tion depth as compared with the application on entire surface (Gupta et al.,
1993). Thus, the reclamation depth can be increased for widely spaced row
crops if a given amount of amendment can be applied to narrow strips meant
for planting.
Particle size: The mined gypsum is obtained in lumps, which are then
ground and packed before sale. Thus, agricultural grade gypsum involves cost
in grinding and sieving. Finer fractions are very reactive in the beginning but
their efficiency is lowered by relatively greater precipitation of dissolved cal-
cium in the presence of soluble carbonates often present in alkali soils (Abrol
et al., 1979; Hira and Singh, 1980). Fineness of gypsum increases its solubility
and may vary between 0.42 and 2% for gypsum sizes of 0.5 to 2.0 mm and 0.1
194 CROP PRODUCTION IN SALINE ENVIRONMENTS

mm, respectively, in alkali soils (Hira and Singh, 1980). The fineness of gyp-
sum also strongly affects the water conductance of soils containing Na2CO3
(Chawla and Abrol, 1982). The finest gypsum (< 0.125 mm) had the highest
initial hydraulic conductivity that decreased sharply with time. Treatment with
coarser particles (< 2.00 mm) had a lower initial hydraulic conductivity that
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

was maintained or decreased with time (Figure 1). This was attributed to precip-
itation of dissolved Ca2+ to CaCO3. Coarser particles, on the other hand, give
lower initial electrolyte levels, which are maintained or increased with time.
Thus it is desirable to have a mixture of various particle sizes with 2 mm upper
limit, to obtain the dual benefit of rapid dissolution of fine gypsum followed by
sustained release of Ca2+ from coarser particles (Chawla and Abrol, 1982).
Dissolution of gypsum: The solubility of gypsum in water is 0.26% at 25°C.
To dissolve 9 to 11.2 Mg ha⫺1 of applied agricultural grade gypsum (85%
passing through 100 mesh sieve), an application of 92 to 122 cm of irrigation
water was recommended (Richards, 1954). Based on solubility of gypsum and

FIGURE 1. Hydraulic conductivity of soil as affected by gypsum grades: (I) < 2


mm, (II) < 1 mm, (III) < 0.6 mm, (IV) < 0.25 mm, and (V) < 0.125 mm. Adapted
from Chawla and Abrol (1982).

Grade
0.6 V

IV

II
K (cm min⫺1 ⫻ 102)

0.4

0.2

0
8 16 24
Cumulative leachate (PV)
P. S. Minhas and O. P. Sharma 195

sodium exchange equation, Dutt et al. (1972) computed that 52 to 72 cm of


water was required to dissolve 16.5 to 24.0 Mg ha⫺1 of applied gypsum. Oster
and Halvorson (1972), however, showed that the solubility of gypsum in-
creased more than 10 fold after mixing it in an alkali soil. Similar observations
were also made in alluvial sodic soils (Abrol et al., 1979; Hira and Singh,
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

1980; Hira et al., 1981). In soil systems, the quantity of gypsum dissolved in-
creases linearly with increasing ESP of the soil (Abrol et al., 1979). At the
highest ESP (78.9), approximately 7.44 g L⫺1 gypsum dissolved, which is
nearly 3 times the solubility of gypsum in water.
Apart from factors, such as composition of soil solution, temperature, water
flow velocity, and fineness of gypsum, the sodium saturation of soil is also an
important factor that affects the dissolution of gypsum. During the Na+-Ca2+
exchange, the soil exchange complex acts as a sink for Ca2+ released from gyp-
sum (solid phase). The solid phase gypsum thus continues to dissolve until the
solution phase is saturated or the ion activity product of Ca2+ and SO42⫺ some-
times equals the Ksp of gypsum. As Ca2+, released by gypsum dissolution is
used up to replace exchangeable Na+, more of it will be released from the solid
to the solution phase which results in increasing the amount of gypsum dis-
solved per unit of water.
The mean solubility of gypsum (Y) increased linearly with an increase in
the ESP of soil (Hira and Singh, 1980) and could be described by the relation:

Y = 0.0186 ESP + 0.18 (r = 0.98)

in an alluvial sodic soil. At the same level of ESP, the dissolution of gypsum
was the maximal in the soil with CO32⫺as the dominant anion, followed by
HCO3⫺ and Cl⫺ (Arora et al. 1981). Hira and Singh (1980) observed a linear
relationship between the agricultural grade (fineness < 0.26 mm) gypsum
dissolved (mg mL⫺1) (Y); and the gypsum requirement (X) of the soil (Mg
ha⫺1) as:

Y = 0.46 X + 2.1 (r = 0.99)

in alkali soils belonging to seven different soil series of Punjab. However, the
increase in GR of soil did not affect the amount of water needed for complete
dissolution of gypsum. About 4 cm of water was sufficient to dissolve all the
agricultural grade gypsum (< 0.26 mm), provided it equaled the GR of soil.
Hira et al. (1981) also observed a linear relationship between the exchangeable
Na+ and the dissolution rate of gypsum of different size grades, ranging from
0.1 to 2.0 mm. Gupta et al. (1985) observed that the water requirement for dis-
solution of gypsum are lower than those observed by Hira et al. (1981). Gupta
and Ranade (1987) also evaluated the applicability of model proposed by Hira
196 CROP PRODUCTION IN SALINE ENVIRONMENTS

et al. (1981) for alkali Vertisols. Good agreement was obtained between exper-
imental and predicted water requirement for gypsum dissolution at higher ESP
(> 30) levels. However, it underestimated the water requirement at lower ESP
(< 30) levels. The water requirement for dissolution of gypsum applied
through 30 cm depth of a sodic black clay soil is about 35 cm for commonly
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

used gypsum size grade irrespective of ESP.

Gypsum vs. Acids/Acidulants

Gypsum and eight different amendments at 3 levels of application (33, 66,


and 100% of GR) were tested by Chand et al. (1977) during 1970-72 in a field
experiment on alluvial alkali (pH1:2 9.9 to 10.5, ESP 88-93) soil (Natric
Ustalf). When applied as chemically equivalent to gypsum, aluminium sul-
phate, and sulphuric acid were equally effective in improving the soil proper-
ties and the yield of barley. Relatively lower yields were obtained in case of
nitric acid and ferrous sulphate. The lowest yields were obtained in case of
FYM and pressmud. The initial experiments in All India Coordinated Re-
search Project trials (1970-75) consisted of comparing the effect of gypsum
with acids (sulphuric acid) and acidulants, viz., aluminum sulfate on sodic
loamy (pHs 9.1-9.8, ECe 4-6 dSm⫺1, ESP 61-80) and black smectitic clay
(pHs 8.7, ECe 4.6 dSm⫺1 and ESP 50) soils. At equivalent amendment doses
(80% GR), sulphuric acid was as effective as gypsum in terms of improving
the soil chemical properties in a period of five years, though it was more effec-
tive than gypsum in the initial years of reclamation of a loamy soil (Sharma
and Mehrotra, 1971). Sulphuric acid was also more effective in improving hy-
draulic properties of the soils. The saturated K improved to 5.5 and 7.1 cm h⫺1
in gypsum- and sulphuric acid-treated alluvial soils, respectively, from an ini-
tial K of 1.7 cm h⫺1 (Tiwari and Sharma, 1989). The infiltration rates for gyp-
sum and sulphuric acid treated soils were 0.39 and 0.50 cm h⫺1, respectively.
The efficacy of gypsum was evaluated with various amendments in alkali
black clay soils (Verma et al., 1985; Sharma et al., 1986a). For reducing ESP
of the soils, the amendments followed the order: Al2 (SO4)3 > H2SO4 > gyp-
sum > pyrites. The dissolution of native CaCO3 was much higher with acid and
acidulants. When all the amendments were applied at equivalent basis, these
had variable effect on the physical properties. Gypsum followed by Al2(SO4)3
was more effective in increasing saturated K and decreasing clay dispersion
than other amendments (Verma et al., 1985).
The overall results favor the use of acids and acidulants for their effective-
ness in reducing soil ESP but the handling hazards and exorbitant prices re-
strict the practical use of these amendments. Apart from improving the physical
and chemical properties, the amendments also led to a decrease in the wa-
ter-soluble boron content that often becomes injurious to crop plants. Pathak et
P. S. Minhas and O. P. Sharma 197

al. (1981) observed that application of amendments at 60% GR proved better


than smaller doses in reducing boron toxicity.

Gypsum vs. Pyrites


Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

The efficacy and cost effectiveness of pyrites and gypsum, for reclamation
of alkali soils, were tested using rice and wheat crops. When the two amend-
ments were applied on equivalent basis in alluvial alkali soils, either the pyrites
were inferior (Verma and Abrol, 1980) or as effective as gypsum (Pathak et al.,
1978; Singh et al., 1981; Tiwari et al., 1985; Tiwari and Sharma, 1989). Gyp-
sum proved to be superior to pyrites in black alkali soils (Table 5). Pyrites had
92% effectiveness as compared to gypsum (Sharma and Gupta, 1986) and the
dissolution of native CaCO3 was much higher under gypsum. Sharma (1990b)
concluded that the pyrite dose should be 1.3 times the laboratory estimated GR
for lowering the soil ESP to a level of 10 in black alkali soils.
Techniques to improve efficiency of pyrites: The main cause for low effi-
ciency of pyrites is their non-oxidation to release soluble sulfur. Several work-
ers have discussed the issues relating to the oxidation problems, particle size

TABLE 5. Influence of graded doses of gypsum and pyrites on crop yield and
soil properties of an alkali alluvium and black clay soil.

Amendment dose Grain yield ESP


(% GR) (Mg ha⫺1) (0-15 cm)
Rice (1977) Wheat (1977-78) Rice (1978) After wheat
Gyp. Pyr. Gyp. Pyr. Gyp. Pyr. Gyp. Pyr.
Alkali loam soil1
0 1.34 1.34 0.51 0.51 3.23 3.23 59 59
25 2.82 2.40 1.00 0.92 4.79 4.37 29 40
50 3.55 3.22 1.41 1.22 5.48 5.22 22 29
75 3.69 3.50 1.68 1.47 5.45 5.52 40 24
LSD (p = 0.05) 0.23 0.10 0.83 30
Alkali clay soil2
Rice (1983) Wheat (1983-84) Wheat (1984-85)
0 0.20 0.20 0.52 0.52 0.47 0.47 44 44
25 1.32 1.18 0.90 0.85 1.09 0.90 30 35
50 2.60 2.29 1.78 1.70 1.55 1.32 20 25
75 3.64 3.18 2.36 2.22 2.43 2.41 17 20
LSD (p = 0.05) 0.26 0.23 0.56
Source: 1Tiwari and Sharma (1989): 2Sharma (1990)
198 CROP PRODUCTION IN SALINE ENVIRONMENTS

effects, and the dose and method of application of pyrites in the reclamation of
calcareous alkali soils (Singh et al., 1978; 1981; Sharma, 1990b; Singh, 1998b;
Sharma et al., 1998). It is evident that soil moisture plays a vital role during the
oxidation of pyrites. The reaction is given below:
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

2FeS2 + 7O2 + 2H2O = 2FeSO4 + 2H2SO4

The above reaction cannot occur in a dry soil. To improve efficiency, it is rec-
ommended that the pyrites (< 5 mm, 22.5% S) be surface applied and the soil
be irrigated (5 cm) to bring the surface soil to field capacity (Verma and Gupta,
1984). Oxidation period of 7 days is allowed, followed by submergence (5 cm
water) for the next 7 days. Soluble salts are flushed to surface drains, followed
by another such cycle. With each submergence and flushing cycle, the dissolu-
tion of Ca2+ occurs and the soil ESP exhibits a decreasing trend (Figure 2).
Thus, for black sodic soils with very poor infiltration, the operation of repeated
flooding and flushing should be resorted to (Sharma et al., 1986b). Surface ap-
plication of pyrites (< 5 mm, 22.5% S) at 50% soil moisture about 7 days be-
fore ponding of water for leaching (Table 6) was beneficial in improving
rice-wheat yields as well as ameliorating the soil (Tiwari and Sharma, 1989).
This confirmed the earlier results of Pathak et al. (1985). The efficiency of py-
rites was considerably improved when it had higher soluble S content. Higher
soluble S resulted in increasing the rice yields as well as reducing the pH and
the soil ESP (Singh et al., 1981). These results indicate that the level of pyrites
should be based on its soluble S content rather than its total S content.
Quantity of water required for leaching pyrites reaction product: In case of
pyrites, water is required for leaching/flushing of soluble salts from the root
zone after allowing 7/10 days for oxidation of pyrites with the soil at field ca-
pacity. Verma and Gupta (1984) observed that most of the decrease in soil ESP
due to pyrites occurred within the first 120 hrs of incubation and maximum re-
covery of soluble Na+ in leachate is obtained when about 1.5-2.0 pore volumes
of water passed through the soil column. Leaching with about 2.0 pore vol-
umes of water reduced the ECe, water soluble Na+, and ESP significantly de-
pending on the dose of amendment applied (Gupta and Sharma, 1987).

Sulphur

Sulphur is biologically oxidized in the soil to form H2SO4, which reacts


with native CaCO3 of the soil to release soluble Ca2+. The oxidation of S in al-
kali soils is likely to be very slow (Singh, 1998b). Basu and Tagare (1943) ob-
tained highest yield of sugarcane in black alkali clay soils of Nira Canal
Command with a combination of S and FYM. Talati (1947) found that sulphur
gave slightly higher sugarcane yield on black alkali soils but gypsum also
proved successful when compared on cost basis. Gidnawar et al. (1972) ob-
P. S. Minhas and O. P. Sharma 199

FIGURE 2. Effect of flooding and pyrites levels (•–P0, ∆–P25, 䊊–P50, and
䊐–P75) on the water soluble Ca2+ (Ca2+e) and soil ESP at different periods of
sampling. Adapted from Sharma et al. (1998).

10
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Ca e (me L⫺1)

6
2+

P0
2
P25

P50
70

P75
ESP

50

30

0 7 14 21 134 292
Days after application

served that annual application of 300 kg S ha⫺1 plus 5 cartloads of FYM per
year for 8 years reduced pH and salt content and increased the infiltration rate
of saline alkali soils of Karnataka when cotton and wheat were grown in alter-
nate years under rainfed conditions. In an experiment conducted at Kamma,
sulphur proved as one of the least effective amongst acid formers (Singh,
1998b). Sulphur is costlier as compared to gypsum and is imported by India
and its low efficiency restricts its use as an amendment.

Flyash

The thermal power plants in the country produce about 10 million Mg


flyash annually. It is composed of relatively larger amounts of Si, Al, Fe, and
Ca along with smaller amounts of Mg, Ti, Na, and K compounds and traces of
200 CROP PRODUCTION IN SALINE ENVIRONMENTS

TABLE 6. Influence of size, time, and method of pyrite application on mean


grain yield (1982-83 to 1986-87) and soil properties (1986-87).

Treatment Yield (Mg ha⫺1) ECe ESP Hydraulic


(dS m⫺1) conductivity
Rice Wheat
(cm hr⫺1)
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Size (mm)
< 10 2.57 1.02 4.0 37.3 0.10
<5 2.79 1.26 3.1 28.1 0.14
<3 2.72 1.19 2.6 27.6 0.16
Time of application (days before transplanting)
21 2.61 1.04 3.3 30.9 0.10
14 2.77 1.19 2.5 28.3 0.11
7 3.15 1.43 1.3 19.1 0.14
0 2.33 0.96 5.9 45.6 0.13
Method of application
Surface applied 2.79 1.21 2.8 28.4 0.12
Mixed (0-15 cm) 2.60 1.09 3.7 33.4 0.09
Initial properties 13.5 81.3 0.01
Adapted from: Tiwari and Sharma (1989)

several other elements. The reaction of flyash varies from basic to acidic de-
pending upon the source of coal. Assuming that flyash with acidic reaction can
act as an amendment, the effectiveness of flyash was compared with pyrites in
ameliorating an alluvial alkali loam soil (Tiwari and Sharma, 1989). Applica-
tions of flyash decreased soil pH and ESP, and compared favorably with py-
rites in increasing yields of rice and wheat. But due to limited data, the use of
flyash cannot be recommended on a larger scale at this point.

Biological Reclamation

The reclamation of alkali soils through chemical amendments is a relatively


high cost technology. Therefore, efforts have been made to reduce reclamation
cost through exploiting inherent calcium of these soils through the use of exog-
enously applied, as well as in situ produced organic materials and exploiting
the alkalinity tolerance of crop plants like rice. These are discussed as below.

Use of Organic Materials and Crops

Barren alkali soils, devoid of any vegetation, have been observed to contain
little oxidizable carbon for support of microbial activity needed for production
P. S. Minhas and O. P. Sharma 201

of CO2 and mobilization of calcite. Incorporation of organic matter in the form


of plant residues, farm manure and compost helps in improving the hydraulic
properties thereby facilitating leaching of salts and mobilizing native Ca2+ in
calcareous alkali soils through decomposition products and increasing micro-
biological activity (Singh, 1998b). As early as 1850, Jameson advocated the
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

use of organic matter (FYM) as an amendment for salt affected soils in the
Oudh state of United Province (Moreland, 1901). Dhar (1956) reclaimed
“bad” alkali soils with a mixture of amendments like sunhemp, Sesbania
aculeata (dhaincha), clover straw, dilute molasses, etc., mixed with calcium
phosphate or basic slag. Uppal (1955) reported the reclamation of an alluvial
alkali soil by growing sunhemp and sesbania. Singh (1974) noted the superior-
ity of sesbania to barley. Aggarwal and Ramamoorthy (1974) showed that
decomposition of proteinacious materials like castor cake and semi-protein-
acious materials like sesbania under anaerobic conditions released Ca2+ from
calcite present in alkali soils. Cellulosic materials like wheat straw were inef-
fective. The effectiveness of the amendments also decreased with an increase
in ESP of the soils. Distillery waste (spent wash) has also been tried for recla-
mation of calcareous alkali soils (Singh et al., 1981).
According to Yadav and Agarwal (1959), paddy rice should be preferably
taken as the first crop after burying the green manure crop of dhaincha.
Chhabra and Abrol (1977) reported that the rice culture reduced the ESP from
93.3 to 28.6 in the absence of any amendment while the pH2 declined from
10.3 to 8.9 (Table 7). Gupta et al. (1988) observed that rice root activity influ-
enced chemical reclamation of alkali soils through CO2 production and mobi-
lization of calcite in soils. There are alternative possibilities of reclaiming
alkali soils by growing tolerant varieties of rice after adding smaller (about
25% GR) doses of gypsum (Singh, 1994; Singh et al., 1998; Mishra, 1994).

TABLE 7. Soil properties as affected by rice culture.

Original soil After the experiment


pH2 ESP Without rice With rice Excess Na+
removed by
pH2 ESP pH2 ESP
rice culture
(meq)
10.3 93.3 9.6 68.6 8.9 28.6 163
9.5 46.0 8.9 26.3 8.3 1.2 80
9.0 29.9 8.4 9.5 8.2 0.6 19
8.4 10.5 8.1 1.8 7.2 0.2 3
Adapted from: Chhabra and Abrol (1977)
202 CROP PRODUCTION IN SALINE ENVIRONMENTS

Mendiratta et al. (1972) observed that the rate of gypsum dissolution is en-
hanced by incorporation of FYM/green manuring. The yields of berseem, sug-
arcane, and rice increased significantly by incorporating gypsum with FYM in
comparison to their sole treatment in alkali soils of Indo-Gangetic Plains (Dar-
gan, 1979). The experiments by Chand et al. (1977) indicated that the lowest
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

yields were obtained in case of FYM and pressmud. The pressmud obtained
from sulphitation process was found to be superior than that obtained from car-
bonation process. Application of organic residues like wheat straw, groundnut
husk, and FYM was not beneficial in terms of rice and wheat yields as well as
changes in soil properties of sodic black soils. Nevertheless, when FYM was
incorporated along with pyrites, a marked reduction in ESP and an improve-
ment in crop yields were observed (Sharma, 1988). Similarly in alluvial alkali
soils, application of FYM at 10 Mg ha⫺1 with pyrites, significantly improved
rice and wheat yields over the sole application of either FYM or pyrite (Tiwari
and Sharma, 1989). Increase in yield with FYM alone was even lesser than the
minimum dose of pyrite applied at 25% GR. The magnitude of ESP reduction
and improvement in physical properties also showed the synergistic effects of
the combined use of inorganic and organic amendments. Therefore, it may be
concluded that FYM or any other organic residue alone may not be an useful
amendment but its synergetic effects should be exploited for reducing the
doses of inorganic amendments during the reclamation of alkali soils.

Amelioration Through Agro-Forestry Interventions

Some of the alkali lands with poor resource endowments, such as the com-
munity lands or those owned by marginal farmers, can be bio-ameliorated
through alternate uses such as growing of trees and grasses. Planting methods
such as auger-bore holes for piercing through the subsoil calcareous layer can
be advantageously combined with small doses of gypsum and tolerant tree
species such as Prosopis juliflora and Acacia nilotica. The most important
benefit of growing trees is enrichment of alkali lands through litter fall. In ad-
dition to the recycling of nutrients, the decomposition of litter leads to evolu-
tion of CO2 that helps to mobilize the inherent calcium. The litter fall from tree
plantations depend upon their growth habits, age, density of plantation and
N-fixing capacity of trees, etc. Gill et al. (1987) observed that litter production
by A. nilotica and E. tereticornis plantations on a highly alkali soil was
2,537-5,746 and 1,027-1,125 kg ha⫺1, respectively. The winter months ac-
counted for 40-55% of the total litter fall that was composed of about 75-82%
foliage. The accumulation of N, P, K, and S was substantially more in litter of
Acacia whereas that of Ca, Mg, Na, Fe, and Mn was more in Eucalyptus. The
annual litter yields of P. juliflora and C. equisetifolia was 7,483-9,967 and
6,957-8,769 kg ha⫺1, respectively, and the litter from the latter was more ligni-
P. S. Minhas and O. P. Sharma 203

fied and hence decomposed at a slower rate. Gill et al. (1987) reported that de-
crease in pH, as a consequence of replacement of sodium with mobilized
calcium, was more in surface layers and with A. nilotica than with E. tereticornis
plantation. There was three- and two-fold increase in organic carbon of surface
0.15 m soil in a span of 5 years under A. nilotica and E. tereticornis, respec-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

tively. Plant growth for 20 years of P. juliflora, A. nilotica, E. tereticornis, A.


lebbeck, and T. arjuna, on a highly alkali soil (pH 10.2-10.5) reduced the pH of
surface 1.2 m soil to 8.1, 9.03, 9.1, 8.67, and 8.15, respectively (Singh and Gill,
1992). The organic carbon of the profile increased several-folds, being the
highest under P. juliflora and the lowest under E. tereticornis (Table 8). The
improvement in organic carbon and reduced pH of alkali soils in turn led to im-
provement in their infiltration characteristics and water retention properties
(Gill et al., 1987; Singh et al., 1993). Improvement of sub soil permeability due
to root growth of trees helps in conserving the soil moisture and improving the
water-holding capacity of subsoil layers (Singh et al., 1993). Thus, in general
tree plantations can reclaim the alkali soils to levels that make it possible to
grow normal crops after completion of one life cycle of trees (8-10 years).
Some of the grasses like Leptochloa fusca (Kallar grass) are not only well
adopted to highly alkali conditions but have reclaiming effects through their
profuse and deep root system (Malik et al., 1986; Ashok-Kumar and Abrol,
1986). Alkali lands get sufficiently reclaimed within 3-4 years of grass growth,
making it possible to grow rice and wheat without any amendments. The
ameliorative effect as brought about by the growth of grasses, particularly of
Kallar grass, was studied on different crops with different treatment combina-
tions. The results revealed that after growing grasses for 1 year, satisfactory
yields of rice were obtained, whereas good yields of wheat were achieved after

TABLE 8. Comparative amelioration of an alkali soil by growing tree planta-


tions.

Tree species Soil property (0-15 cm depth) after 20 years


pH2 EC Org. C Avg. P Avg. K
(dS m⫺1) (%) (kg ha⫺1) (kg ha⫺1)
Original soil 10.2 1.18 0.22 28.0 278.0
Acacia nilotica 8.4 0.25 0.85 59.0 498.6
Eucalyptus tereticornis 8.5 0.44 0.66 33.3 358.9
Prosopis juliflora 7.5 0.51 0.93 110.5 701.5
Terminalia arjuna 7.9 0.32 0.86 67.8 409.8
Albizia lebbek 7.9 0.32 0.62 42.6 387.0
Adapted from: Singh and Gill (1992)
204 CROP PRODUCTION IN SALINE ENVIRONMENTS

2 years (Ashok-Kumar and Abrol, 1986). Furthermore, in the fourth and fifth
years the yields of rice and wheat were higher in the grass-grown plots even
than those in gypsum-applied plots. After growing Kallar and Para grasses for
3 year and followed by rice-wheat rotation, sensitive forage crops like Persian
clover, Egyptian clover, teosinte, and maize could also be grown successfully
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

(Ashok-Kumar, 1990).

Reclamation Technique for Rainfed Areas

Ample water is required for the reclamation of alkali soils for their leach-
ing/flushing after the incorporation of chemical amendments as well as for
growing the rice crops. However, sufficient water supplies are not available in
many areas where alkali Vertisols exist. For normal soils, the raised and
sunken bed system has been suggested for improving drainage and water stor-
age under high rainfall situations (Gupta et al., 1978). Moreover, in alkali
Vertisols with high clay contents, the effectiveness of gypsum is limited to a
leaching depth of 15-20 cm due to poor water transmission properties, high
bulk density, and high runoff potential. Under such conditions, upland crops
suffer heavily on account of poor drainage, impedance to root penetration, and
low moisture intake. The raised and sunken bed system was also tested in 1:1
configuration with 3 bed widths (4.5, 6.0, and 7.5 m) to generate a suitable
technology for areas having black alkali clay soil with a modest annual rainfall
(< 800 mm) and devoid of irrigation facilities (Verma and Sharma, 1998a).
Amendment was applied to a depth of 15 cm followed by shifting the soil to the
raised beds and using the 15-30 cm deep sunken bed for rice culture. Forty to
fifty milimeter of stored rainwater, collected just after the onset of monsoon,
was surface applied in order to minimize salt injury to rice crop. Since the rice
crop is fairly alkali tolerant it was grown in the sunken beds. An alkali tolerant
variety of cotton was planted on the raised beds. A strip of natural grasses oc-
cupying 10% of the area on both sides of raised beds was provided as vegeta-
tive barrier to minimize soil and nutrient loss due to runoff. Water balance
analysis showed that 62-70% of the rainwater could be conserved and that
2.5-4.0 Mg ha⫺1 of paddy yields and 0.24-0.49 Mg ha⫺1 of cotton yields could
be obtained from the area without any supplemental irrigation.

CULTURAL PRACTICES DURING RECLAMATION


OF ALKALI SOILS

Crop Rotations

The choice for crops to be grown on alkali soils is restricted due to adverse
physical and chemical conditions. Growing tolerant crops as per the sodicity of
P. S. Minhas and O. P. Sharma 205

soils can ensure reasonable returns during the initial phases of reclamation.
Rice is recommended as the kharif crop as it can tolerate both water stagnation
and higher ESP in soils. Rice growing for 3 years has been observed to lower
the soil pH to levels (about 9.0) such that it becomes possible to profitably
grow the lesser tolerant crops like wheat, berseem, and mustard (Singh et al.,
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

1998). Out of the four crop rotations tried, paddy-wheat-dhaincha (Sesbania


aculeata) and paddy-berseem in the loamy alluvial soils (Tiwari and Sharma,
1989) and paddy-berseem followed by paddy-wheat in the black alluvial soils
(Jain et al., 1985) were observed to be the most remunerative crop rotations.
After 8 years of cropping, soil pHs, ECe, and ESP of black alkali soils were
8.1, 3.8 dS m⫺1 and 31.7, whereas those of alluvial alkali soils were 8.0, 1.8 dS
m⫺1, and 19.0, respectively. Later experiments on loamy sodic soils (Tiwari
and Sharma, 1989) revealed that dhaincha-rice-wheat followed by rice-
wheat are the most remunerative crop rotations. Dhaincha-wheat followed
by dhaincha-barley was observed to be more remunerative crop rotation as
compared to fallow-wheat and fallow-barley in the sandy loam sodic soils of
Rajasthan. The rotations like dhaincha-mustard and fallow-mustard gave the
least returns.

Irrigation Management

Presence of excessive exchangeable sodium in alkali soils modifies the soil


water relations such that their water holding capacity is reduced considerably.
It is a common observation that due to their poor infiltration characteristics,
water stagnates on alkali soils following a rainfall or an irrigation event. Dur-
ing drying, the surface few cm dries quickly whereas there may not be any
change in water contents of layers below 15 cm depth (Acharya and Abrol,
1978). Due to these properties, rice is the principle crop grown on these soils.
Experiments by Singhandhupe and Rajput (1988) showed that upto three irri-
gations could be saved without affecting rice yields if irrigations were applied
a day after disappearance of ponded water. Gaul et al. (1973) recommended
that irrigation frequency for rabi crops should be increased on alkali soils as
compared to normal soils. Sharma et al. (1985) reported that first irrigation to
wheat on sodic soils (ESP 25-40) should be delayed to 30 days after planting as
compared to the recommended 22-25 days for normal soils. Wheat grown on a
black alkali soils (ESP 35 to 65) that received 30 cm of irrigation water either
through flooding or sprinkler, in response to a Cumulative Pan Evaporation
(CPE) of 3.6 cm, produced better yields and higher water use efficiency
(Verma et al., 1993). The production function fitted to predict wheat yields
from soil ESP and depth of irrigation (Diw, cm) was:
206 CROP PRODUCTION IN SALINE ENVIRONMENTS

Yield (Mg ha⫺1) = 0.043 ESP + 0.61 Diw ⫺ 0.32 ESP ⫻ Diw
+ 0.36 (Diw)2 ⫺ 1.358 (R2 = 0.93)

Based on the above relationship, under the conditions of 2.5 cm CPE, appli-
cations of 27 to 30 cm irrigation water through sprinkler system were com-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

puted to be enough for economic yield of wheat. Increased depth of applied


water was more beneficial at lower ESP of soils. Thus irrigation should be ap-
plied more frequently with smaller quantities to upland crops grown on alkali
lands (Acharya and Abrol, 1991).

Nutrient Management

Black alkali soils are known to be deficient in both the available nitrogen as
well as phosphorus whereas alluvial alkali soils are deficient in nitrogen but
are medium to high in available phosphorus. In general, the organic matter
content is low in both soil types.

Nitrogen

The efficiency of nitrogen fertilizer is generally low in alkali soils partly


due to volatilization losses as ammonia and partly due to sub-optimal plant
growth. Nitant (1974) reported that complete hydrolysis of urea in a sodic soil
with pH 10.3 was delayed by 4 days as compared to the normal soil. Bhardwaj
and Abrol (1978) observed that nearly 32 to 52% of the applied nitrogen was
lost through volatilization in sodic soils. Nitrogen losses may also depend
upon the quantity of fertilizer applied at a given time, the cation exchange ca-
pacity of the soil, and the soil pH. To compensate the losses of nitrogen in al-
kali soils, it is recommended that crops should be fertilized with 20% extra
nitrogen than recommended for normal soils (Rao and Batra, 1983; Tiwari and
Sharma, 1989). In loamy alkali soils, a three-year study revealed that increased
N doses of upto 200 kg ha⫺1 resulted in significant increase in wheat and
paddy yields (Sharma et al. 1998). However, in black alkali soils, wheat,
paddy, cotton, and barley responded only upto 120 kg N ha⫺1. Among differ-
ent sources, ammonium sulphate proved to be a better source of nitrogen than
other common N fertilizers (Pathak et al., 1975; Khaddar et al., 1987). Due to a
rapid enzymatic hydrolysis of urea in alkali soils, the released NH3 tends to es-
cape into the atmosphere. Such losses could be checked to a large extent by us-
ing neem-coated urea (NCU, 35.4% N) or phospho-gypsum urea (Urea-G,
36.8% N). Deshmukh and Tiwari (1996) recorded maximum grain yield of
paddy in alkali Vertisols by the use of Urea-G.
Singh and Swarup (1990) reported that two equal applications of the recom-
mended dose of urea, with the basal application turned under at the time of
puddling, gave higher rice yields than three split doses applied after puddling
P. S. Minhas and O. P. Sharma 207

of a mildly alkali soil. In order to have higher efficiency, split applications of


nitrogen in rice should be adopted in black alkali soils (Khaddar et al., 1987).
Symbiotic N fixation is also retarded in salt affected soils due to sensitivity of
microbes to high alkalinity and retarded growth of host plants. Rhizobium
could survive and multiply in sodic soils upto pH 10.0 but the host plant is sen-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

sitive at this sodicity (Rao, 1998). It is limited by the pH effect per se or due to
reduced energy source caused by reduction in the growth of host plants.
Hence, an extra 15-20% N application is recommended in alkali soils (Dargan
and Gaul, 1974).

Phosphorus

Sodic alluvial soils of the Indo-Gangetic Plains of Punjab and Haryana have
high extractable P status (Singh and Nijhawan, 1932; Chhabra and Abrol,
1983). However, in Uttar Pradesh, the alkali soils are not always high in avail-
able P. It is well known that the availability of P depends upon the pH and
CaCO3 content of the soil. In the presence of Na2CO3, Ca3(PO4)2 undergoes
double decomposition forming soluble Na3PO4 and CaCO3, thereby sharply
increasing Olsen’s extractable P (Chhabra et al., 1981). The sodium phosphate
form is highly mobile and moves alongwith water streams towards plant roots
and is also prone to leaching. Upto 60 mg P kg⫺1 soil could leach out of soil
columns when sodic soils were leached after surface application of gypsum
(Chhabra et al., 1981). In nature, the leaching of soluble P is prevented partly
by the poor permeability of sodic soils. Chhabra (1985) observed substantial
leaching losses of soluble P in the initial years of reclamation and cropping.
There was buildup of NaHCO3-extractable P in P-fertilized plots while P sta-
tus of the control plots showed a constant decline. The physiological availabil-
ity of phosphates in alkali soils may be a limiting factor for plant growth.
The salt affected soils are calcareous in nature and in such soils the applied
P is rendered insoluble within few hours of application as P ions precipitate on
CaCO3 particles. In alkali soils amended with gypsum, octa-calcium phos-
phate is reported to be a dominant form of P. Thus the available P contents of
the alkali soils after gypsum applications show a decreasing trend. The type of
amendment used for reclamation of alkali soils also plays an important role in
determining the responses to applied P (Sharma et al., 1998). Experiments
with rice-wheat on alkali soils amended with gypsum and pyrites showed
better responses to applied P especially in wheat when soil was amended with
pyrites (Table 9). The magnitude of decrease in available P varied with the na-
ture of amendment used. An increase in the rate of applied gypsum decreased
the water soluble P greatly, but it had little effect on the strongly adsorbed P
fraction (NaOH-P) or Ca-bound P (HCl-P) (Swarup, 1994).
208 CROP PRODUCTION IN SALINE ENVIRONMENTS

TABLE 9. Effect of phosphorus application on mean yields of rice and wheat


(1991-94) in gypsum and pyrites amended alkali soil.

Amendment Dose Crop yield (Mg ha⫺1) at applied P (kg ha⫺1)


Rice Wheat
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

(% GR) Nil 0.45 Nil 0.45


Control - 1.60 1.62 0.84 1.14
Gypsum 50 2.99 3.10 2.26 2.42
100 3.60 3.98 2.94 3.35
Pyrites 50 2.77 2.87 1.83 2.25
100 3.30 3.72 2.47 2.86
LSD (p = 0.05) 0.43 0.29
Adapted from: Tiwari and Sharma (1989)

In long-term experiments at Karnal, Haryana, it was observed that the sur-


face 15 cm soil lost 24 kg ha⫺1 of soluble P through leaching in the initial years
of cropping. Due to inherent high soluble P content of the soils, rice and wheat
crops grown in rotation did not respond to P application during the initial 3-5
years of cropping. It has been observed that rice crop may suffer once the
Olsen’s P of the surface soil falls below 7.5 kg ha⫺1 (Singh, 1998). Wheat may
not show any decline in yield for the following few years. Therefore, fertiliza-
tion of P should be made after testing the soil for available P. Contrary to this,
studies conducted in Gangetic alluvial plains (Kanpur, U.P.) indicate that al-
kali soils are not always high in available P (Sharma et al., 1998).
Potassium
The clay fraction of Indo-Gangetic alluvium is dominated by illite that acts
as a good reserve of K+ in these soils; likewise basaltic black alkali soils also
have high available K+ due to the presence of potassium feldspar in the mineral
makeup of the soil. However, some alluvial alkali soils have very poor avail-
able soil K2O (72.5 mg kg⫺1) (Sharma et al., 1998). These soils responded
well to the applications of potassium with increased yields of rice and wheat.
The rate of K+ release and its uptake are not ordinarily affected in salt affected
soils. Since the water-soluble and ammonium acetate soluble K+ fraction is at a
high to very high level in these soils, no responses to K+ applications have been
observed. However, it is necessary to maintain a high K+ supply in order to
regulate Na+: K+ balance in plants.
Calcium
The soluble and exchangeable form of Ca2+ becomes deficient in both high
(sodic soils) and low pH (acid) soils and requires supplementation through
P. S. Minhas and O. P. Sharma 209

amendments. Absolute Ca2+ concentration and Na+/Ca2+ ratio are important


indices of plant response to sodic conditions and ultimate yields. Amendments
like gypsum are direct source of Ca2+ but acid (H2SO4) and acid formers (py-
rites) generally mobilize Ca2+ of the precipitated CaCO3 thereby ensuring ade-
quate supply of Ca2+ in calcareous soils.
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Zinc

Zinc precipitates as hydroxide and carbonate in alkali soils. Due to high pH,
presence of CaCO3, high soluble P, and low organic matter content, sodic soils
often contain less than 0.6 mg kg⫺1 DTPA extractable Zn (Katyal et al., 1980).
Zinc deficiency in plants is invariably related to soil reaction rather than to to-
tal zinc content of soils. Singh et al. (1980) attributed high extractability of
added zinc with increase in ESP to the formation of soluble sodium zincate.
The addition of amendments decreased the availability of added zinc. Besides
being an essential element, Zn also plays an ameliorative effect in alkali soils
by enhancing absorption of Ca2+ and K+ and thereby widening the Ca2+/Na+
and K+/Na+ ratios in plants. These physiological attributes interact and con-
tribute towards increased crop yields in salt affected soils.
Rice, though fairly tolerant to sodicity, is sensitive to zinc deficiency, which
may appear 15 to 20 days after transplanting causing stunted growth with
rusty-brown spots and ultimately severe yield reductions. In the soils amended
with 10-15 Mg ha⫺1 gypsum, addition of 10-20 kg ha⫺1 of zinc sulphate was
enough to meet the Zn requirements of the soil (Singh et al., 1987; Table 10).
Accordingly, application of 20-40 kg Zn ha⫺1 is recommended during recla-
mation (Singh, 1998b). Even high level of pyrites with fertilizers, in the ab-
sence of zinc, did not give increased production that was obtained with an
additional application of ZnSO4 at 50 kg ha⫺1 (Tiwari et al., 1985). In black al-
kali soils, applications 50 kg ha⫺1 of ZnSO4 had a positive residual effect on
the succeeding wheat crop (Sharma et al., 1998).

Other Cultural Practices

Due to poor infiltration vis-a-vis water stagnation problems, the optimal


crop stand especially that of upland kharif crop is very difficult to obtain on al-
kali soils. Therefore, alternate methods of planting to offset mortality of young
seedlings due to waterlogging effects have been devised. Planting on the side
of ridges within the capillary fringe, just above the irrigation water in furrow,
gave the maximum yield in alluvial as well as black soils (Sharma et al., 1998).
The other alternatives are increasing seeding rate and increasing number of
plants per hill in case of transplanted crops (Singh et al., 1998).
210 CROP PRODUCTION IN SALINE ENVIRONMENTS

TABLE 10. Effect of gypsum and Zn on rice yield (Mg ha⫺1) in an alkali soil.

ZnSO4 Gypsum applied (Mg ha⫺1)


(Kg ha⫺1)
0 2.5 5.0 10.0 Mean
0 0.12 0.98 2.14 2.65 1.48
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

10 0.49 1.87 3.06 3.77 2.30


20 0.58 2.00 3.14 3.85 2.39
30 0.68 1.75 2.90 3.92 2.34
40 1. 05 2.02 3.29 3.89 2.56
LSD (p = 0.05) Gypsum levels 0.36 Zn levels 0.28 Gyp. ⫻ Zn
NS
Adapted from: Singh et al. (1987)

PRACTICES TO RECLAIM WATERLOGGED SALINE SOILS


Characteristics and Extent
Saline soils are usually formed as a result of the salt stored in the soil profile
and/or groundwater being mobilized by extra water provided by human activi-
ties such as irrigation. The extra water raises the water table or increases pres-
sures of confined aquifers to create an upward movement to the water table.
When the water table comes closer to the soil surface, water is evaporated, leav-
ing salts behind and causing land salinization. In India, the problematic areas
with high water table cover 2.6 million ha while 3.4 million ha suffer from sur-
face water stagnation. The characteristic features of some of the typical saline
pedons are included in Table 2. A wide range of options are available for control
of salinity and waterlogging in canal command areas but technical, economic,
social, and political considerations are the major factors which influence their
large scale implementation as most of the options involve community actions.
Provision for Drainage
The prime requirement for successful crop production in saline soils is that of
reversing the flux of water to enhance salt leaching. To achieve this and to main-
tain water table depth deep enough to provide adequate root development and
aeration, the provision of natural or artificial drainage is considered essential.
The major drainage techniques being pursued in India include both the surface
and subsurface (vertical, horizontal) drainage. These are summarized below.
Surface Drainage
For removing excessive surface water, land shaping is done in such a way
that water flows over the field surface to furrows or ditches and ultimately into
P. S. Minhas and O. P. Sharma 211

the waterways. In shallow water table areas, surface drains are created deep
enough to intercept ground water, which will then enter the main drain. The
rate of groundwater flow into the drain depends on soil permeability and depth
of water table. The deeper a drain, the greater is the width of adjacent land af-
fected by the draw down of the groundwater. The use of the surface drains to
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

control water table suffers from the disadvantages of loss of land, hindrance to
farming operations, heavy maintenance costs in terms of weed control and
overland instability of banks. However, properly designed surface drains can
provide adequate drainage in agricultural farms to reduce ponding problems
during monsoon and check rise in water table. A 3-tier system of optimum
rainwater management was suggested (Gupta and Narayana, 1972; Narayana,
1979). The main features of this system are:

• Part of the rainfall should be collected in the cropland till such time and
in an amount that will not be harmful to the crops, e.g., the experiments
showed that excess rainwater up to 15 cm could be stored in bunded rice
fields.
• After storage in the cropland, the excess water from the catchment
should be collected in the dugout ponds located in the lower regions of
the catchment. The stored water can be utilized for irrigation during the
dry spells.
• The remaining excess water is then led into the regional drainage system.
Sub-Surface Drainage

Two man made systems of sub-surface drainage, viz., vertical and horizon-
tal are in vogue. Vertical drainage is mainly achieved by pumping out ground-
water through the tubewells and the horizontal drainage involves engineering
structures laid parallel to the ground surface.
In other words, these structures work like tubewells laid horizontally. Both
types of drainage systems aim at lowering the water table in response to re-
charge caused by rainfall, irrigation, leaching water, etc.

Vertical Drainage

The success of a vertical drainage system is associated with the presence of


favorable unconfined aquifers within 50 m depth below the ground surface and
with the quality of groundwater. Such favorable conditions prevailed in the al-
kali soil areas of Haryana, Punjab, and Uttar Pradesh, and have led to large-
scale installations of private and government tubewells. As a consequence, the
rising water table trends have been reversed. However, the vertical drainage
has not been feasible in saline areas due to their poor aquifer yields and saline
ground waters. Thus, the water tables are rising at alarming rates leading to
212 CROP PRODUCTION IN SALINE ENVIRONMENTS

secondary salinization. For such areas, a multiple well point system has been
devised to lower water table by skimming fresh water floating over saline wa-
ter (Shakya et al., 1995). The system consists of a number of well points ar-
ranged in a line, interconnected to each other through a horizontal pipe line
(lateral) buried at 70 to 100 cm below the ground level which is pumped by a
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

centrally located sump. In the latter case, it is known as a multiple well-point


syphon system. Similarly, Singh et al. (1994) have tried installation of open
wells placed at 150-300 m grid for skimmimg of good quality water. But both
these systems need to be tested on a large scale in some canal command areas.

Horizontal Drainage

Horizontal subsurface drainage is the technique of controlling the water ta-


ble and salinization by installation of horizontal drains at a certain depth (about
1.5 to 2.5m) below the surface. The pattern of drains allows the land to drain
into the collectors that carry the water away from the land. The drains are in-
stalled at an appropriate depth and are spaced at intervals designed to ensure
that the water table in the intervening space does not rise above a given depth.
Drainage design parameters are related mainly to the hydraulic conductivity of
the soil and the drainage criterion, which specifies the required discharge and
the hydraulic head. Many useful practical recommendations have emerged from
recent installations at different locations in the country (Rao et al., 1990; 1994;
Rao and Gupta, 1998). These recommendations are summarized as follows:

• The depth of lateral drains can vary between 1.5-2.0 m depending upon
the soil type, nature of crops to be grown and the outlet conditions. Initial
and operational costs will escalate with deeper drains, in addition to the
trench caving problems during the laying process.
• The lateral drain spacing of 50-75 m in alluvial soils and 12-24 m in
black soils has been observed to be sufficient to facilitate growing of
crops within 2-3 years on lands that have been fallow for considerable
time due to salinity problems.
• For many years, tile and concrete pipes have been predominantly used
for subsurface drainage. Corrugated plastic drain pipes with small perfo-
rations have now become available. The corrugated form makes these
pipes more resistant to deformation. They are manufactured in diameters
ranging from 50 to 200 mm, and are delivered in coils of different
lengths. These have proved to be so successful that they are gradually re-
placing the tile and concrete drain pipes. Corrugated plastic pipes are
made of polyvinylchloride (PVC), polyethylene (PE), and polypropy-
lene (PP). Preference for one of these materials should be based on eco-
nomic grounds (Kumbhare and Rao, 1994).
P. S. Minhas and O. P. Sharma 213

• Envelope materials were originally intended to protect the drain pipes


against soil particle invasion, but they must also facilitate water inflow
by creating a more permeable environment surrounding the pipe. The
materials used can be classified as granular or synthetic. Granular mate-
rials such as sand and fine gravel are still widely used. Nowadays syn-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

thetic envelope materials are replacing granular materials. They can


easily be wrapped around the drain pipes and do not decay once the drain
pipes are installed (Rao et al., 1994; Rao and Gupta, 1998). After the in-
troduction of corrugated plastic drain pipes, techniques have been devel-
oped to pre-wrap these pipes with an envelope material in the factory.
• When a seep is fed by a distinct water source as is common in black soil
areas, the use of interceptor drains can cut off this excess ground water
before it reaches the low lying problem area. Interceptor drains are usu-
ally a single tube or open drain placed between the water source and the
problem area. Accurate location of the source of excess groundwater and
proper placement of the drain are critical to the success of this form of
drainage. These drains have been observed to be quite effective for con-
trol of salinity in black soils those are usually underlain with bedrock and
sandwiched in between by a highly permeable murrum layer (AICRP-
Saline Waters, 1998).

MANAGEMENT OF SALINE SOILS AFTER DRAINAGE

For sustainable crop production in saline soils after providing the necessary
drainage measures, adoption of specific system of management is advocated
based upon the soil, crop, and climatic factors at the site (Minhas, 1998b).
Nevertheless, it may be pointed out that crop production on saline-water-
logged soils is generally more costly per unit area of land, whereas crop yields
are usually low. Hence the profit margins are also less whereas risk of crop
failures may still continue even after suitable drainage measures have been
provided. In fact, not much success has been achieved towards raising of
kharif crops at many locations installed with sub-surface tile drainage. This is
mainly due to inundation of these lands with floodwaters during the monsoon
season.

Leaching Management

The first requisite for crop production in saline soils is to lower salinity to
acceptable limits, which is accomplished through the process of leaching.
Leaching may be one time displacement of initial salts by ponding of low sa-
linity waters at the soil surface, i.e., reclamative leaching or maintaining salt
balances by meeting out leaching requirements during cropping season. The
214 CROP PRODUCTION IN SALINE ENVIRONMENTS

extent of leaching required during reclamation depends upon initial salinity,


salt tolerance of crops to be grown, and depth of water table. Minhas (1998a)
has compiled some of the general recommendations on leaching management
of Indian soils . These include:
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

• Regarding the depth of soil to be reclaimed, it has been suggested that


salts from the entire depth of the root zone should be removed when at-
tempting reclamative leaching. However, bringing down the salinity of
surface 45-60 cm of soil below the threshold limits should suffice while
rest of leaching can be taken care of during cropping in the monsoon sea-
son.
• The quantity of water required depends mainly on soil texture and depth
of soil to be reclaimed. Removal of 80% salts would require 1.5, 1.1, and
0.4 cm water per cm soil depth in fine, medium, and coarse textured soils,
respectively (Minhas, 1998a). In other words, if the surface 50 cm soil is
to be leached, water requirement of respective soils would be 75, 55, and
22 cm (Figure 3).
• Intermittent leaching of clay soils may be followed to improve leaching
efficiency whereas less benefit is expected from this approach for leach-
ing the saline soils of alluvial plains. This is because these are usually
light textured soils (loamy sand-sandy loam) and high evaporative de-
mands prevail except in winters (Minhas and Khosla, 1986).
• Incorporation of the monsoon rains into reclamative leaching plans is
also of utmost importance for reducing drainage volumes. Therefore to
synchronize leaching with monsoon rains and to get benefit of deeper
water table, salt leaching should preferably be executed during summer.
Computations by Minhas and Gupta (1992) showed that for the removal
of 80% of the accumulated salts, 1.85, 0.95, and 0.76 cm of rain water per
cm of soil depth would be required in fine, medium, and coarse textured
soils, respectively. In other words, only 30-55% of the total rains during
monsoons are expected to contribute towards salt leaching.
• Addition of amendments for waterlogged saline soils is not recom-
mended. However, in saline-sodic soils, pH increases and soil clays be-
come vulnerable to dispersion and movement during leaching of salts
(Minhas, 1995). Thus, addition of gypsum to prevent surface sealing and
enhance infiltrability can be beneficial during leaching in such dispersible
soils.
• As the requirement of fresh water for leaching is governed by initial sa-
linity, water requirements can be reduced if initially the soil is leached
with saline waters. Estimates by Gupta and Pandey (1983) show that a
saving of 15-40% in fresh water can be achieved with this sequence of
water applications.
P. S. Minhas and O. P. Sharma 215

• In saline soils installed with tile drainage, salt leaching from the area near
the drain will be the maximum and decreases from the area towards mid-
way between the drains. Making ridges parallel to the drains and avoid-
ing ponding near the drains or applying the water intermittently to keep
the water table as low as possible can achieve the uniformity in distribu-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

tion of infiltration. The sequential ponding of entire land, central half,


and central quarter for 12.5 days each for achieving leaching uniformity
in an alluvial sandy loam soil has been recommended (Rao and Leeds-
Harrison, 1990).

FIGURE 3. Leaching curves for different textured soils. Adapted from Minhas
(1998a)

1.0

0.8
[C ⫺ Ceq]/[Co ⫺ Ceq]

0.6 Sandy loam

Loam–silt loam

Clay–clay loam

0.4

0.2

0
1 2 3
DW/DS
216 CROP PRODUCTION IN SALINE ENVIRONMENTS

Crop Management

Tolerance to salinity varies a great deal, almost 10 fold, amongst the crop
plants and their genotypes. The inter- and intrageneric variations in salt toler-
ance of plants can be exploited for selecting crops or varieties that produce sat-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

isfactorily under a given root-zone salinity. However, the cultivation of high


water requiring crops like sugarcane and rice should be avoided as these aggra-
vate the salinity problems. Minhas and Gupta (1992) have compiled the salt
tolerance limits of important agricultural crops of the different agro-ecological
zones of India, which provides a basis for the selection of appropriate crops to
be grown in a rotation. The most common crops for saline areas in alluvial
tracts are usually wheat, mustard, and barley during rabi and cotton, sorghum
(fodder), and pearl millet during kharif. Singh and Sharma (1991) reported that
pearl millet-wheat; pearl millet-barley; pearl millet-mustard; sorghum (fod-
der)-wheat, and sorghum (fodder)-mustard cropping sequences were more re-
munerative in saline soils. Cotton based cropping systems were not of much
benefit since the yields of the following rabi crops were low. In water scarce
areas, mustard can replace wheat since the water requirements of the former
are lower.
During the initial stages, the interacting zone of roots is limited to surface
few inches where most salts concentrate due to the evaporation of water from
soils. Hence, germination and early seedling establishment are the most criti-
cal stages for most crops in saline soils (Minhas and Gupta, 1992). The other
critical periods for crops are during the transition to reproductive develop-
ment, i.e., flowering. Otherwise tolerance to salinity increases with an ad-
vancement of age of crops. Under non-steady state conditions such as existing
during monsoon climate, wheat responses to initially variable salinity profiles
superimposed by various patterns of salinization have been reported by Minhas
and Gupta (1993). Although the total salt with which wheat roots interacted
during growth period was kept constant, three fold variations in its yields were
observed. Independent estimates of response to salinity that existed down to
rooting depth at different stages of wheat showed ECe50 (ECe for 50% yield
reduction) to increase from 9.1 until crown rooting to 13.2 dS m⫺1 at dough
stage. It was thus implied that salt tolerance at critical stages of crop plants
changes in response to salinity and modes of salinization. Initial distribution of
salinity needs to be considered for effective description of crop responses to
salinity.

Use of Saline/Drainage Waters

On farm irrigation management of saline soils should involve the irrigation


schedules in order to minimize the number and depth of irrigations, eliminate
salinity build up, and assure optimal crop production. The ultimate goal should
P. S. Minhas and O. P. Sharma 217

be to control soil salinity by ensuring downward flow of water through the root
zone. Substantial contributions to the seasonal evapotranspiration can come
from the shallow water tables, e.g., wheat yields can be sustained with only
one irrigation under non-saline shallow (1.2 m) water table condition while
50-70% of its water requirement is met when water table was saline (Minhas
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

and Gupta, 1992). Thus, maximizing the crop water use from shallow ground
waters can reduce the volume of drainage effluents but generally such a prac-
tice would simultaneously lead to salt accumulations in rooting zone.

Combined Use of Canal and Saline/Drainage Water Supplies

When canal water supplies are either unassured or are in short supply, such
that the farmers are forced to pump saline ground or drainage waters to meet
the crop water requirements. The waters from these two sources can be applied
either separately or mixed. Mixing of waters to an acceptable quality also re-
sults in improving the stream size and thus the uniformity in irrigation espe-
cially for the surface method practiced on sandy soils. Allocation of the two
waters separately, if available on demand, can be done either to different
fields, seasons or crop growth stages so that higher salinity water is not applied
to sensitive crops or growth stages.

Pre-Irrigation

Primary objectives of presowing irrigation are the creation of optimal soil


moisture conditions to facilitate tillage and seedbed preparation and to re-
charge the projected root zone with water for germination and later ET needs
of the crops. In saline soils, these should further include the leaching of soluble
salts below the seeding zone, as the germination and seedling establishment
are most critical and the failure of crop at this stage cannot be rectified later.
Plants are also known to generally tolerate salinity better with advancement of
age. Experiments have shown that crops like mungbean (green gram), sor-
ghum, and mustard can tolerate higher salinity once the non-saline water was
substituted for presowing irrigation to leach out the salts from the seeding
zone. This substitution increased the germination, crop growth, and yields
greatly and also resulted in better utilization of saline soil water even from the
lower soil depths (Minhas and Gupta, 1992). The presowing irrigation as-
sumes a more critical role for the success of kharif crops. In saline areas prone
to waterlogging, the kharif crops fail to establish properly owing to high salin-
ity and temperatures if sown before the onset of monsoon and due to excess
water if sown after the onset of monsoon. Experiments under such situations
have shown potential of raising a successful crop like sorghum, if it could be
established in the pre-monsoon season after leaching the accumulated salts
even with saline water (ECe > ECiw) but followed by small additions of
218 CROP PRODUCTION IN SALINE ENVIRONMENTS

non-saline waters. The crops established in such a manner could withstand the
excess water from the subsequent monsoon rains.

Cyclic Uses of Canal and Saline Waters


Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

This strategy involves the substitution of non-saline (canal) waters for sa-
line waters at the most sensitive growth stages of the crops grown. Saline wa-
ters are used at other stages such that the deleterious effects of the resultant soil
salinity build up on crops can be minimized. To obviate the critical growth pe-
riods from the detrimental effects of canal failure or inadequate water supplies,
different modes of saline (ECiw 6-8 dS m⫺1) and canal water applications un-
der various cropping intensities (100-300%) have been tried in network exper-
iments (AICRP-Saline Water, 1998). Results of these trials indicate that crop
yields were higher when cropping intensities were low. But in general, the
crop responses to various modes of saline and canal water usage did not vary.
The general conclusions drawn were: (1) Yields could be maintained close to
those obtained under Best Available Water (BAW) by delayed substitution of
saline water, i.e., after two initial irrigations with BAW. The next best way is
to alternate irrigations with BAW and saline water and (2) Irrigation with sa-
line waters should not be at the critical growth stages, especially at sowing be-
cause high salinity hinders germination and seedling establishment.

Mixing vs. Cyclic Use of Multi-Quality Waters

The two options of utilizing the multi-quality waters have been discussed
above. If it is presumed that the pre-requisite facilities for blending exist and
different qualities of waters are simultaneously available on demand, the ques-
tion arises as to which option should be followed. Analysis of experiments
conducted over the years with different crops by Minhas and Gupta (1992)
showed benefits of cyclic use mode over mixing in terms of relative yields at
the same level of weighted average salinity. The advantage from different cy-
clic use modes followed the order; (2 canal:1 saline) > (1 canal:1 saline) >
(1 canal:2 saline) waters. Differences between the observed and estimated
yields were greater at low relative yields, indicating increased benefits from
cyclic use at higher ECiw. Thus it provides a useful evidence that multi-salin-
ity waters should be used cyclically and the use of canal waters at early stages
and saline waters should be delayed to later stages. Recent experiments where
saline (ECiw 9-12 dS m⫺1) and canal waters were combined for cotton-wheat
and pearl millet-mustard rotations also support the suitability of cyclic use
strategy described above (Table 11; Minhas et al., 1998).
P. S. Minhas and O. P. Sharma 219

TABLE 11. Effect of various cyclic use and mixing modes of irrigation with
canal and saline waters on wheat and mustard yields (Mg ha⫺1).

Treatment Yield (Mg ha⫺1)


Cotton Wheat Pearl millet Mustard
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Canal water, CW 1.63 4.88 3.15 2.07


1 CW:1 SW 1.23 4.72 2.96 1.96
2 CW:2 SW 1.28 4.62 -- --
1 SW:1 CW 0.76 4.02 -- --
2 SW:1 CW -- -- 2.91 1.41
1 CW:rest SW 0.98 4.05 2.99 1.88
1 CW:1 CW:rest SW 0.72 4.08 2.80 1.67
Mixing(1:1 CW:SW) 1.04 4.37 2.80 1.81
Saline water, SW 0.46 3.59 2.91 1.18
LSD (p = 0.05) 0.32 0.35 NS 0.36
ECiw = 12 dS m⫺1for pearl millet-mustard and 9-12 dS m⫺1 for cotton-wheat rotation
Adapted from: Minhas et al. (1998)

FERTILIZER MANAGEMENT
Enhanced fertility can alleviate the adverse effects of salts to only some ex-
tent but cannot overcome them, e.g., increasing the level of phosphorus over
the recommended dose mitigates the adverse affects of salinity especially
when chlorides are the dominant anions in saline soils (Minhas and Gupta,
1992; Manchanda, 1998). However, additional doses of nitrogenous fertilizers
(25% extra) are recommended to compensate for volatilization/leaching losses
but not to alleviate salinity stress. Incorporation of organic/green manures
have been shown to have advantages in saline soils in terms of improving nitro-
gen use efficiency (NUE) because it serves as temporary bounding agent for the
ammonical pool of nitrogen and reduce volatilization losses. The role of organic
materials in reducing the volatilization losses and enhance the NUE under saline
environment is indicated by the increased response to N-fertilizers. Long-term
experiments have indicated that a combination of organic and inorganic sources
saved N by 50% in rabi and 25% in kharif (AICRP-Saline Waters, 1998).

ECONOMICS OF RECLAMATION OF SALT-AFFECTED SOILS


It is estimated that about 1.0 million ha of alkali lands have been reclaimed
in the states of Punjab, Haryana, and Uttar Pradesh by following hydro-chemi-
220 CROP PRODUCTION IN SALINE ENVIRONMENTS

cal technology (ICAR, 2000). This alone is leading to additional food grain
production valued at Rs. 10 billion per annum, additional employment genera-
tion of 75 million man months per annum and environment improvement like
reduction in flood hazards, increased groundwater recharge and increased fuel
wood production. This has been facilitated mainly by the governmental poli-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

cies, organized efforts of State Land Reclamation and Developmental Corpo-


rations and credit by nationalized banks. At 50 and 75 percent subsidy on
gypsum, the benefit cost ratio has been estimated to be 3.9 and 4.6, respec-
tively, with a pay back period of 2.0 and 2.6 years (Tyagi, 1998). Similarly the
cost of installation of drainage system in a hectare area with a drain spacing of
75 m has been estimated to be about Rs. 21,000 at 1994-95 prices. Its cost-ben-
efit ratio is 1.76 and internal rate of return is 20%, which is more than the inter-
est being offered by the commercial banks on their term deposits. Based on the
6-8 years rotation of tree plantations, the net annual income from afforestation
of alkali and saline soils has been worked out to be Rs. 8,175 and Rs. 3,587 per
ha, respectively.

CONCLUDING REMARKS
Various problems associated with salt affected soils have been researched
for long in different soil and agro-climatic situations of India, which have
yielded valuable concepts and economically viable technologies to reclaim sa-
line and alkali lands. Techniques evolved have been successfully implemented
to reclaim large areas, yet the rehabilitation of such lands, especially the saline
waterlogged soils, is not progressing at the expected pace. This is mainly be-
cause the resource-poor farmers of the salt affected areas have difficulties to
secure additional inputs in terms of amendments, assured water supplies and
drainage infrastructure for reclamation of these lands. In fact, a major success
in reclamation has been achieved with alkali soils underlain with good quality
water but the technologies for rainfed areas and the alkali soils underlain with
poor quality waters are missing and need to be developed (CSSRI, 1997). Also
the quantum of amendments being finite, their cost is rising and thus methods
to improve their efficiency are required. A better understanding of gypsum dis-
solution, inherent calcite in soils, cation exchange phenomenon, and the mod-
eling to predict long-term desodication/re-sodication can be of help in this
direction. Feasibility of using some alternate amendments like fly ash, press-
mud, dairy and paper effluent should also be tested. Also missing until now are
the studies on the fate of reaction products of alkali soil reclamation, which
upon joining the sub-soil water, will impact groundwater quality. Low input
use efficiency in alkali lands further necessitates the optimization of chemical
fertilizers/organic residues/manure so as to obtain nutrient release in syn-
chrony with the plant needs for enhancing fertilizer use efficiency. Bio-ame-
P. S. Minhas and O. P. Sharma 221

lioration of alkali lands has been shown to be a feasible but a slow process.
Thus it needs to be combined with chemical technology to get the benefits of
their synergies. With liberalization of economy and WTO implications, em-
phasis is being given for production of high value crops. Thus crop diversifica-
tion into vegetable, horticulture and other cash crops should become a priority
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

in future on reclaimed alkali soils.


Soil salinity and waterlogging are issues that need to be considered in a
much broader perspective. Quantitative models describing the quantity and
quality of return flows need to be developed for predicting long-term changes
in canal command areas and provision of drainage is essential in sensitive ar-
eas. The technical feasibility of drainage is now well established for saline wa-
terlogged soils, but the major impediment in its expansion has been the
demand for community efforts and public funding. Moreover, such systems
have to be made cost effective in terms of water table control and reductions in
volume of drainage effluent to be disposed. Thus efforts should be directed to-
wards introduction of controlled drainage systems in association with vertical
drainage wherever feasible. Drainage systems also require assured water sup-
plies for maintaining salt balances, which are not always available. In such ar-
eas especially with sufficient monsoon rainfall, harvesting of rainwater and its
reuse along with drainage water may prove to be cost effective and an environ-
mentally safe solution.
It may also be stated here that all the salt affected soils cannot be economi-
cally reclaimed for crop production. Such lands can be put to alternate land use
systems such as for growing salt tolerant trees, grasses, and other halophytic
plants. The tree based land use systems, because of their higher evapotran-
spiration demands, can also contribute towards minimization of secondary
salinization through reduction in underground water table. However, the choice
of species to be planted, the sites to be planted and planting density required as
well as how to effectively carry on the plantation of salt affected soils, requires
careful planning, testing, and execution. More efforts should efforts should be
directed towards biodiversity and further exploitation of environmentally and
economically useful plants, e.g., halophytes with aromatic and medicinal
value. In other words, biosaline agriculture should be promoted on salt af-
fected soils.

REFERENCES
Abrol, I. P. and Bhumbla, D. R. (1978). Some comments on terminology relating to
salt-affected soils. In: Proc. Dryland-Saline-Seep Control, Edmonton, Canada,
June 1978. pp. 6-27.
Abrol, I. P. and Bhumbla, D. R. (1979). Crop response to differential gypsum applica-
tion in a highly sodic soil and the tolerance of several crops to exchangeable sodium
under field conditions. Soil Sci. 127: 79-86.
222 CROP PRODUCTION IN SALINE ENVIRONMENTS

Abrol, I. P. Dargan, K. S. and Bhumbla, D. R. (1973). Reclaiming Alkali Soils. Bull. 2,


Div. Soil Sci. & Agron. CSSRI, Karnal. 58 p.
Abrol, I. P., Chhabra, R. and Gupta, R. K. (1980). A fresh look at the diagnostic criteria
for sodic soils. Proc. Int. Symp. Salt Affected Soils. CSSRI, Karnal. pp. 142-146.
Abrol, I. P., Dahiya, I. S. and Bhumbla, D. R. (1975). On the method of determining
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

gypsum requirement of soils. Soil Sci. 120: 30-36.


Abrol, I. P., Gupta, R. K. and Singh, S. B. (1979) Note on the solubility of gypsum and
sodic soil reclamation. J. Indian Soc. Soil Sci. 27: 482-483.
Acharya, C. L. and Abrol, I. P. (1978). Exchangeable sodium and soil water behaviour
under field conditions. Soil Sci. 125: 310-319.
Acharya, C. L. and Abrol, I. P. (1991). Soil Water Relationa and Root Water Extrac-
tion in Alkali Soils. Tech. Bull. 16, CSSRI, Karnal. 96 p.
Acharya, C. L. and Abrol, I. P. (1998). Soil-Water relationships. In: Agricultural Salin-
ity Management in India. N. K. Tyagi and P. S. Minhas (Eds.), Central Soil Salinity
Research Institute, Karnal. pp. 79-91.
Aggarwal, R. P. and Ramamoorthy, B. (1974). Effect of different organic materials on
the availability of calcium in calcareous sodic soils. Ann. Arid Zone15: 117-120.
Agarwal, R. R. and Mehrotra, C. L. (1953). Soil Survey and Soil Work in U.P.III. U.P.
Dep. Agric. Supdt. Printing & Stationary, Allahabad, India.
Agarwal, R. R., Mehrotra, C. L. and Gupta, C. P. (1957). Spread and intensity of soil al-
kalinity with canal irrigation in Gangetic alluvium of Uttar Pradesh. Indian J. Agric.
Sci. 27: 363-373.
AICRP-Saline Waters (1998). Annual Progress Reports 1972 to 1998, AICRP on
Management of Salt Affected Soils and Use of Saline Water in Agriculture, CSSRI,
Karnal.
Arora, Y., Chaudhary, M. R. and Singh, N. T. (1981). Dissolution rate of gypsum in
sodic soils. J. Indian Soc. Soil Sci. 29: 361-365.
Ashok-Kumar (1990). Effect of gypsum compared with that of grasses on the yield of
forage crops on a highly sodic soil. Exp. Agric. 26: 185-188.
Ashok-Kumar and Abrol, I. P. (1986). Grasses in Alkali Soils. Bul. 11, CSSRI, Karnal.
95 p.
Basu, J. K. and Tagare, V. D. (1943). Soils of the Deccan Canals IV. The alkali soils,
their nature and management. Indian J. Agric. Sci. 13: 157-181.
Bhardwaj, K. K. R. and Abrol, I. P. (1978). Nitrogen management in alkali soils. Natl.
Symp. Nitrogen Assimilation and Crop Productivity, Hisar. pp. 83-96.
Bhargava, G. P. (1977). Classification of salt affected soils–some problems. Proc.
Indo-Hungarian Symposium on Management of Salt Affected Soils. CSSRI, Karnal.
pp. 31-55.
Bhargava, G. P. (1989). Salt Affected Soils of India. Oxford & IBH Pvt. Ltd., New
Delhi.
Bhargava, G. P., Pal, D. K., Kapoor, B. S. and Goswami, S. C. (1981). Characteristics
and genesis of some sodic soils in the Indo-Gangetic alluvial plains of Haryana and
Uttar Pradesh. J. Indian Soc. Soil Sci. 29: 61-70.
Bhumbla, D. R. (1977). Alkali and saline soils of India. Proc. Indo-Hungarian Sympo-
sium on Management of Salt Affected Soils. CSSRI, Karnal. pp. 14-19.
P. S. Minhas and O. P. Sharma 223

Bhumbla, D. R. and Abrol, I. P. (1972). Effect of application of different levels of gyp-


sum on yield of rice, wheat and barley grown in a saline-sodic soil. International
Symposium on New Developments in the Field of Salt Affected Soils, Cairo, Egypt.
pp. 537-574.
Chand, M., Abrol, I. P. and Bhumbla, D. R. (1977). A comparison of eight amend-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

ments on soil pproperties and crop growth in a highly sodic soil. Indian J. Agric.
Sci. 47: 348-354.
Chawla, K. L. and Abrol, I. P. (1982). Effect of gypsum fineness on the reclamation of
sodic soils. Agric. Water Mgmt. 5: 41-50.
Chhabra, R. (1985). Crop response to phosphorus and potassium fertilisation of a sodic
soil. Agron. J. 77: 669.
Chhabra, R. and Abrol, I. P. (1977). Reclaiming effect of rice grown in sodic soils. Soil
Sci. 124: 49-55.
Chhabra, R. and Abrol, I. P. (1983). Principles governing fertilization in salt affected
soils. Fert. Ind. A. Rev. II: 147-159.
Chhabra, R., Abrol, I. P. and Singh, M. V. (1981). Dynamics of phosphorus during rec-
lamation of sodic soils. Soil Sci. 132: 319-324.
CSSRI (1997). Vision-2020, CSSRI Perspective Plan, Indian Council of Agricultural
Research, CSSRI, Karnal, 95 p.
Dargan, K. S. and Gaul, B. L. (1974). Optimum nitrogen levels of dwarf varieties of
paddy in semi reclaimed soil. Fert. News. 19: 24-25.
Dargan, K. S. (1979). Agronomic and cultural practices for crop production in salt af-
fected soils. Curr. Agric. 3: 1-20.
Dargan, K. S., Abrol, I. P. and Chand, M. (1972). Gypsum–FYM a good combination
in saline alkali soil. Indian Fmg. 22: 27-28.
Dubey, D. D., Sharma, O. P. and Ghosh, D. K. (1998). Characterization and delinea-
tion of salt affected soils. In: 25 Years of Research on Management of Salt-Affected
Soils and Use of Saline Water in Agriculture. P. S. Minhas, O. P. Sharm and S. G.
Patil (Eds). CSSRI, Karnal. pp. 5-52.
Dutt, G. R., Terkeltoub, R. W. and Ranschkolb, R. S. (1972). Prediction of gypsum and
leaching requirements for sodium affected soils. Soil Sci. 114: 93-103.
de’Sigmond, A. A. J. (1927). The chemical characteristics of soil leachings. Proc. Ist
Int. Cong. Soil Sci.1: 60-90.
de’Sigmond, A. A. J. (1932). The reclamation of alkali soils in Hungary. Imp. Bur. Soi
Sci. Tech. Comm. No. 23: 5-26.
Deshmukh, S. C. and Tiwari, S. C. (1996). Efficiency of slow release nitrogen fertilis-
ers in rice (Oryza sativa) on partially reclaimed sodic Vertisol. Indian J. Agron. 41:
544-547.
Dhar, N. R. (1956). Calcium phosphates and their importance in nitrogen fixation and
alkali soil reclamation. Proc. Natl. Acad. Sci. (India), 25A: 211-226.
Dhir, R. P. (1998). Origion and accumulation of salts in soils and ground waters. In:
Agricultural Salinity Management in India. N. K. Tyagi and P. S. Minhas (Eds.),
Central Soil Salinity Research Institute, Karnal. pp. 21-40.
Dubey, D. D., Sharma, O. P., Sethi, M. and Gupta, R. K. (1995). Salt Affected Soils of
Gujarat-Extent, Nature and Management. Central Soil Salinity Research Institute,
Karnal. 95 p.
224 CROP PRODUCTION IN SALINE ENVIRONMENTS

Elgabaly, M. M. (1971). Reclamation and management of salt affected soils. In: Salin-
ity Seminar, Baghdad. pp. 50-79. Irrigation and Drainage Paper 7, FAO, Rome.
Gaul, B. L., Abrol, I. P. and Dargan, K. S. (1973). Frequent and light irrigation best for
saline-sodic soils. Indian Fmg. 23: 14-15.
Gedroiz, K. K. (1916). The absorbing capacity of the soil and soil zeolitic bases. Zhur.
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Opytn. Agron. 17: 485.


Gidnawar, V. S., Gumaste, S. K. and Krishnamurthy, K. (1972). Effects of amend-
ments on the properties of saline alkali soils. Bull. 14, Res. Ser. Univ. Agric. Sci.,
Hebbal. pp. 306-312.
Gill, H. S., Abrol, I. P. and Samra, J. S. (1987). Nutrient recycling through litter pro-
duction in young plantations of Acacia nilotica and Eucalyptus tereticornis in a
highly alkaline soil. Forest Ecol. Mgmt. 22: 57-69.
Gupta, R. K. and Abrol, I. P. (1989). Salt affected soils, their reclamation and manage-
ment for crop production. Adv. Soil Sci. 11: 232-292.
Gupta, R. K. and Ranade, D. H. (1987). Water requirement for dissolution of gypsum
in a sodic clayey soil. J. Indian Soc. Soil Sci. 35: 474-479.
Gupta, R. K., Ranade, D. H. and Paradkar, V. K. (1993). Root growth pattern of cotton
in sodic soils as affected by the pattern of gypsum incorporation. Indian J. Agric.
Engg. 3: 58-62.
Gupta, R. K. and Rao, D. L. N. (1994). Potetiial of wastelands in sequestering carbon
by reforestation. Curr. Sci. 66: 378-380.
Gupta, R. K. and Sharma, O. P. (1987). Effect of post oxidation leaching on reclama-
tion of sodic clay soils with sedimentary pyrites. J. Indian Soc. Soil Sci. 35:
188-193.
Gupta, R. K., Singh, C. P. and Abrol, I. P. (1985). Dissolution of gypsum in alkali soils.
Soil Sci. 140: 382-386.
Gupta, R. K., Sharma, S. G., Tembe, G. P. and Tomar, S. S. (1978). A new approach in
farming system for problem soil of Central Madhya Pradesh. JNKVV Res. J. 12:
73-79.
Gupta, R. K., Sharma, O. P. and Dubey, S. K. (1988). Effect of dose and frequency of
gypsum application on soil properties of sodic Vertisols and crop performance. In-
dian J. Agric. Sci. 58: 449-453.
Gupta, R. K., Singh, R. R. and Abrol, I. P. (1988). Influence of simultaneous changes in
sodicity and pH on hydraulic conductivity of an alkali soil under rice culture. Soil
Sci. 146: 395-402.
Gupta, R. K. and Verma, S. K. (1983). Water behaviour of black clay soil as influenced
by degree of sodicity. Curr. Agric. 7: 114-121.
Gupta, R. K. and Verma, S. K. (1984). Influence of soil ESP and electrolyte concentra-
tion of permeating water on hydraulic properties of a Vertisol. Z. Pflanzen Ernahr.
Bodenk. 147: 77-84.
Gupta, S. K. and Narayana, V. V. D. (1972). Surface drainage methods for saline alkali
soils. Symposium on Waterlogging: Causes and Measures for Its Prevention. CBIP,
publ. 118. Vol. II, 129-134.
Gupta, S. K. and Pandey, R. S. (1983). A mathematical model for evaluating leaching
of saline soils when saline and fresh waters are used in conjunction. Indian J. Agric.
Sci. 53: 701-716.
P. S. Minhas and O. P. Sharma 225

Hira, G. S. (1998). Water requirement during reclamation of alkali soils. In: Agricul-
tural Salinity Management in India. N. K. Tyagi and P. S. Minhas (Eds.), Central
Soil Salinity Research Institute, Karnal. pp. 175-184.
Hira, G. S., Bajwa, M. S. and Singh, N. T. (1981). Prediction of water requirement for
gypsum dissolution in sodic soils. Soil Sci. 131: 353-358.
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Hira, G. S. and Singh, N. T. (1980). Irrigation water requirement for dissolution of gyp-
sum in sodic soil. J. Soil Sci. Soc. Am. 45: 930-933.
ICAR. (2000). Vision 2020, Natural Resource Management. Division of Natural Re-
source Management, Indian Council of Agricultural Research, Krishi Bhavan, New
Delhi. 46 p.
Jaggi, T. N. (1978). Use of Amjhore pyrites as alkali soil amendment. Proc. Symp. Use
of Sedimentary Pyrites in Reclamation of Alkali Soils. FCI-PPCL Dept. of Agri.
U.P., Lucknow. pp. 37-43.
Jain S. C., Gupta Ram, K., Sharma, O. P. and Paradkar, V. K. (1985). Agronomic ma-
nipulation in saline sodic soils for economic biological yield. Curr. Sci. 54:
424-425.
Jameson, Dr. (1852). On the physical aspect of Punjab: Its agriculture and botany. J.
Agric. Hort. Soc., India. 8: 132-138.
Kanwar, J. S., Bhumbla, D. R. and Singh, N. T. (1965). Studies on the reclamation of
saline sodic soils in the Punjab. Indian J. Agric. Sci. 35: 45-51.
Katyal, J. C., Randhawa, N. S. and Sharma, B. D. (1980). Annual Report. All India Co-
ordinated Project on Micronutrients in Soils and Plants. ICAR, New Delhi.
Kelley, W. P. and Arany, A. (1928). The chemical effect of gypsum, sulphur, iron sul-
phate and alum on alkali soils. Hilgardia. 3: 393-420.
Kelley, W. P. (1927). A general discussion on the chemical and physical properties of
alkali soils. Proc.1st Int. Cong. Soil Sci. 5: 483-489.
Khaddar, V. K., Ray, N. and Singh, D. (1987). Transformation of different nitrogenous
fertilizers in predominantly sodic soil and their efficacy in wet land rice culture.
Fertil. News. 32: 23-35.
Khosla, B. K., Dargan, K. S., Abrol, I. P. and Bhumbla, D. R. (1973). Effect of depth of
mixing gypsum on soil properties and yield of barley, rice and wheat grown on a sa-
line sodic soil. Indian J. Agric. Sci. 43: 1024-1031.
Kumbare, P. S. and Rao, K. V. G. K. (1994). Performance of filter and drain pipe mate-
rial. In: Proceedings of Symposium on Land Drainage for Salinity Control in Arid
and Semi-Arid Regions. DRI, Cairo, Egypt pp. 120-132.
Leather, J. W. (1906). Ann. Rep. Imp. Dep. Agric. 1904-5. Government Central Press,
Calcutta.
Leather, J. W. (1914). Investigations on Usar Land in the United Provinces. Govt.
Press United Provinces. 87 p.
Mackenna, J. (1918). Report on Progress of Agriculture in India. 1916-17. Govern-
ment Central Press, Calcutta.
Malik, K. A., Aslam, Z. and Naqvi, N. (1986). Kallar Grass–A Plant for Saline Lands.
Nuclear Institute for Agriculture and Biology, Faisalabad, Pakistan. 93 p.
Manchanda, H. R. (1998). Management of saline irrigation. In: Agricultural Salinity
Management in India. N. K. Tyagi and P. S. Minhas (Eds.), Central Soil Salinity
Research Institute, Karnal. pp. 407-430.
226 CROP PRODUCTION IN SALINE ENVIRONMENTS

Mann, H. H. and Tamhane, U. A. (1910). The salt lands of Nira Valley. Bull. No. 39
Dept. of Agric., Bombay.
Mehta, K. M., Mathur, C. M., Shankaranarayana, H. S. and Moghe, V. B. (1969). Sa-
line-Alkali Soils in Rajasthan–Their Nature, Extent and Management. Res. Mono-
graph–1. Department of Agriculture, Govt. of Rajasthan, Jaipur.
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Mendiratta, R. S., Darra, B. L., Singh, H. and Singh, Y. P. (1972). Effect of somer cul-
tural, chemical and manurial treatments on the chemical characteristics of saline
sodic soils under different crop rotations. Indian J. Agric. Res. 6: 81-88.
Minhas, P. S. (1995). Saline water management for irrigation in India. Agric. Water
Mgmt. 30: 1-24.
Minhas, P. S. (1998a). Leaching for salinity control. In: Agricultural Salinity Manage-
ment in India. N. K. Tyagi and P. S. Minhas (Eds.), Central Soil Salinity Research
Institute, Karnal. pp. 161-174.
Minhas, P. S. (1998b). Crop production in saline soils. In: Agricultural Salinity Man-
agement in India. N. K. Tyagi and P. S. Minhas (Eds.), Central Soil Salinity Re-
search Institute, Karnal. pp. 325-350.
Minhas, P. S. and Gupta, R. K. (1992). Quality of Irrigation Water–Assessment and
Management. Publication and Information Section, Indian Council of Agricultural
Research, New Delhi. 123 p.
Minhas, P. S. and Gupta, R. K. (1993). Conjunctive use of saline and non-saline wa-
ters. I. Response of wheat to initial salinity profiles and salinisation pattrens. Agri.
Water Mgmt. 23: 125-131.
Minhas, P. S. and Khosla, P. K. (1986). Solute displacement in a silt loam soil as af-
fected by the method of water application under different evaporation rates. Agric.
Water Mgmt. 12: 64-74.
Minhas, P. S., Patil, S. G., Chauhan, C. P. S. and Hebbara, M. (1998). Management of
poor quality waters. In: 25 Years of Research on Management of Salt-Affected Soils
and Use of Saline Water in Agriculture. P. S. Minhas, O. P. Sharma and S. G. Patil
(Eds.). CSSRI, Karnal. pp. 150-171.
Mishra, B. (1994). Breeding for salt tolerance in crops. In: Salinity Management for
Sustainable Agriculture, D. L. N. Rao, N. T. Singh, R. K. Gupta. and N. K. Tyagi
(Eds.), Central Soil Salinity Research Institute, Karnal. pp. 226-259.
Moreland, W. H. (1901). An account of the attempts what have been made to utilise the
upland barren lands (Usar) of North-western Provinces and Oudh for profitable
purposes. Agric. Ledger. 8: 415-462.
Narayanan, V. V. D. (1979). Rain water management for lowland rice cultivation. J.
Irrig. Drain. Proc. Amer. Soc. Civil. Engg. 105: 87-89.
Nasir, S. M. (1923). Some observations on barren soils of Lower Bari Doab Colony in
the Punjab. Bull. Imp. Agric. Res. Inst., Pusa No. 145. 11.
Nitant, H. C. (1974). Urea transformations in salt-affected and normal soils. J. Ind. Soc.
Soil Sci. 22: 234-239.
Northcote, K. H. and Skene, J. K. M. (1972). Australian Soils with Saline Properties.
Soils Publ. 27, CSIRO, Canberra, Australia.
Oster, J. D. and Halvorson, A. D. (1972). Saline seep chemistry. In: Proc. Dryland-Sa-
line-Seep Control–11th Int. Cong. Soil Sci. Edmonton, Canada. pp. 27-29.
P. S. Minhas and O. P. Sharma 227

Pathak, A. N., Dixit, V. K., Sharma, D. N. and Sharma, M. L. (1985). Studies on some
physico-chemical aspects of pyrites as amendment on reclamation of alkali soil and
crop yield. Proc. Sym. Reclamation of Salt Affected Soils, Lucknow, Sept. 15-16,
pp. 106-117.
Pathak, A. N., Sharma, D. N. and Sharma, M. L. (1981). Water soluble B concentration
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

as affected by different chemical amendments and its impact on B content of paddy


and barley plants in salt affected soils. Agrokem Talajt. 30 (Suppl.): 198-204.
Pathak, A. N., Sharma, D. N. and Sharma, M. L. (1978). Studies on the performance of
gypsum and pyrites in amelioration of salt affected soils. Proc. Symp. Use of Sedi-
mentary Pyrites in Reclamation of Alkali Soils. FCI-PPCL Deptt. of Agri. U.P.,
Lucknow. pp. 45-53.
Pathak, A. N., Tiwari, K. N. and Upadhyay, R. L. (1975). Studies on Fe and Zn nutri-
tion of rice in saline alkali soil of different moisture regimes. Indian J. Agric. Sci.
48: 335-339.
Puri, A. N. (1934). The relation between exchangeable sodium and crop yield in
Punjab soils and a new method of characterising alkali soils. Res. Publ., Punjab
Irrig. Res. Inst. 4 (5): 1-4.
Raj-Kumar (1998). Mineralogy and mineral transformations. In: Agricultural Salinity
Management in India. N. K. Tyagi and P. S. Minhas (Eds.), Central Soil Salinity
Research Institute, Karnal. pp. 107-124.
Rao, D. L. N. (1998). Microbiological processes. In: Agricultural Salinity Manage-
ment in India. N. K. Tyagi and P. S. Minhas (Eds.), Central Soil Salinity Research
Institute, Karnal. pp. 125-144.
Rao, D. L. N. and Batra, L. (1983). Ammonia volatilisation from applied nitrogen in al-
kali soils. Plant & Soil. 70: 219-228.
Rao, K. V. G. K. and Gupta, S. K. (1998). Drainage of alluvial soils. In: Agricultural
Salinity Management in India. N. K. Tyagi and P. S. Minhas (Eds.), CSSRI, Karnal,
pp. 279-296 .
Rao, K. V. G. K. and Leeds-Harrison, P. B. (1990). Water management strategies for
improving leaching in subsurface draiange. Proc. Natl. Seminar on Management of
Irrigation Systems. CSSRI, Karnal.
Rao, K. V. G. K., Aggarwal, M. C., Singh, O. P. and Oosterbaan, R. J. (1994). Recla-
mation and management of waterlogged saline soils. CSSRI, Karnal.
Rao, K. V. G. K., Kumbhare, P. S., Kamra, S. K. and Oosterbaan, R. J. (1990). Reclama-
tion of waterlogged alluvial saline soils in India by subsurface drainage. In: Proceed-
ings of Symposium on Land Drainage for Salinity Control in Arid and Semi-Arid
Regions. DRI, Cairo, Egypt.
Rao, M. S. R. M. and Sheshachalam, N. (1976). Improvement of intake rate problems
in black soil. Mysore J. Agric. Sci. 10: 52-58.
Richards, L. A. (Ed.) (1954). Diagnosis and Improvement of Saline and Alkali Soils.
USDA Handbk. No. 60, USDA, Washington, DC.
Shakya, S. K., Gupta, P. K. and Singh, S. R. (1995). Multiple Well Point System for Ir-
rigation/Drainage. Bull. P. 5. Punjab. Agric. Univ., Ludhiana.
Sharma, D. N. and Mehrotra, C. L. (1971). Effect of acidulating amendments on recla-
mation of saline and alkali soils. Proc. of All India Symposium on the Soil Salinity,
Kanpur. pp. 137-177.
228 CROP PRODUCTION IN SALINE ENVIRONMENTS

Sharma, O. P. (1988). Ameliorative effect of organic residues and pyrites on sodic


Vertisols. Trans. Int. Workshop Swell Shrink Soils, Nagpur. Oxford & IBH Co. Pvt.
Ltd. New Delhi, pp. 222-225.
Sharma, O. P. (1990). Prospectives of reclamation and management of sodic Vertisols.
Proc. Indo-Pak Workshop on Soil Salinity and Water Management. Pakistan Agri-
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

cultural Research Council, Islamabad, Pakistan, pp. 91-105.


Sharma, O. P. (1990a). Physico-chemical properties and classification of sodic Vertisols.
Proc. Indo-Pak Workshop on Soil Salinity and Water Management. Pakistan Agri-
cultural Research Council, Islamabad, Pakistan, pp. 39-54.
Sharma, O. P. and Gupta, R. K. (1986). Comparative performance of gypsum and py-
rites in sodic Vertisols. Indian J. Agric. Sci. 56: 423-429.
Sharma, O. P., Gupta, R. K. and Chaturvedi, R. S. (1986a). Response of sodic vertisol
to different reclamation agents. Agrokem. Talajt. 35: 472-478.
Sharma, O. P., Gupta, R. K. and Dubey, D. D. (1986b). Amelioration of a sodic clay
soil with pyrites. J. Indian Soc. Soil Sci. 34: 155-159.
Sharma, O. P., Sharma, D. N. and Minhas, P. S. (1998). Reclamation and management
of alkali soils. In: 25 Years of Research on Management of Salt Affected Soils and
Use of Saline Water in Agriculture. P. S. Minhas, O. P. Sharma and S. G. Patil
(Eds.). Central Soil Salinity Research Institute, Karnal, pp. 64-84.
Sharma, D. K., Singh, K. N. and Chhillar, R. K. (1985). Effect of time of first irrigation
on yield of wheat in alkali soil. Indian J. Agron. 30: 459-463.
Sharma, S. K. (1987). Mechanisms of tolerance in wheat genotypes differing in sodicity
tolerance. Plant Physiol. & Biochem. 14: 87-94.
Singh, D. and Nijhawan, S. D. (1932). A study of the physico-chemical changes ac-
companying the process of reclamation of alkali soils. Indian J. Agric. Sci. 2: 1-18.
Singh, G. and Gill, H. S. (1992). Ameliorative effect of tree species on characteristics
of sodic soils at Karnal. Indian J. Agric. Sci. 62: 142-146.
Singh, G. B., Singh, N. T. and Tomar, O. S. (1993). Agro-Forestry in Salt-Affected
Soils. Tech. Bull. 17, CSSRI, Karnal. 65 p.
Singh, K. N. (1994). Crop and agronomic management. In: Salinity Management for
Sustainable Agriculture. D. L. N. Rao, N. T. Singh, R. K. Gupta and N. K. Tyagi
(Eds.). Central Soil Salinity Research Institute, Karnal, pp. 124-144.
Singh, K. N., Ashok-Kumar and Sharma, D. K. (1998). Management of problem soils.
In: Fifty Years of Agronomic Research in India. R. L. Yadav, P. Singh, R. Prasad
and I. P. S. Ahlawat (Eds.), Indian Society of Agronomy, New Delhi. pp. 255-270.
Singh, K. N. and Sharma, D. P. (1991). Crop sequences for saline soils provided with
subsurface drainage. Annu. Rep. 1991-92, CSSRI, Karnal.
Singh, M. V., Chhabra, R. and Abrol, I. P. (1987). Interaction between application of
gypsum and zinc sulphate on the yield and chemical composition of rice grown on
an alkali soil. J. Agric. Sci. Camb. 108: 275-279.
Singh, N. T. (1974). Physico-chemical changes in sodic soils incubated at saturation.
Plant & Soil. 40: 303-311.
Singh, N. T. (1992). Salt affected soils of India. In: Land and Soils. T. N. Khusooo and
B. L. Deekshatalu (Eds.), Har-Anand Publ., New Delhi, pp. 65-102.
Singh, N. T. (1998a). Historical perspective. In: Agricultural Salinity Management in
India. N. K. Tyagi and P. S. Minhas (Eds.), Central Soil Salinity Research Institute,
Karnal, pp. 261-278.
P. S. Minhas and O. P. Sharma 229

Singh, N. T. (1998b). Reclamation and management of alkali soils. In: Agricultural Sa-
linity Management in India. N. K. Tyagi and P. S. Minhas (Eds.), Central Soil Salin-
ity Research Institute, Karnal. pp. 261-278.
Singh, N. T. and Swarup, A. (1990). On reclaimed alkali soil urea before puddling ben-
efits rice. Indian Fmg. 40 (9): 50.
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Singh, N. T., Bhumbla, D. R. and Kanwar, J. S. (1969). Effect of gypsum alone and in
combination with plant nutrients on crop yield and amelioration of saline-alkali
soil. Indian J. Agric. Sci. 39: 1-9.
Singh, N. T., Brar, G. S. and Singh, B. (1980). Composition of soluble salts in salt af-
fected soils of the Punjab State. Proc. Int. Symp. Salt Affected Soils, Karnal.
pp. 7-19.
Singh, N. T., Hira, G. S. and Bajwa, M. S. (1981). Use of amendments in reclamation
of alkali soils in India. Agrokem. Talajt. 30 (suppl.): 158-177.
Singh, N. T., Hira, G. S. and Singh, R. (1978). Comparative effects of pyrites and gyp-
sum on physico-chemical properties of a sodic soil. Proc. Symp. Use of Sedimen-
tary Pyrites in Reclamation of Alkali Soils FCI-PPCL Dept. of Agri., U.P.,
Lucknow, pp. 69-80.
Singh, O. P. (1998). Salinity and waterlogging Problems in irrigation commands. In:
Agricultural Salinity Management in India. N. K. Tyagi and P. S. Minhas (Eds.),
Central Soil Salinity Research Institute, Karnal, pp. 61-76.
Singh, R., Singh, J. and Aggarwal, M. C. (1994). Management of shallow saline water
table through open wells. J. Agric. Engng. 21: 71-78.
Singhandhupe, R. B. and Rajput, R. K. (1988). Response of rice to irrigation and nitro-
gen source and rate in an alkali soil. In: Nat. Seminar on Management of Irrigation
System. CSSRI, Karnal, pp. 149-255.
Swarup, A. (1994). Chemistry of salt affected soils and fertility management. In: Salin-
ity Management for Sustainable Agriculture. D. L. N. Rao, N. T. Singh, R. K. Gupta
and N. K. Tyagi (Eds.). Central Soil Salinity Research Institute, Karnal, pp. 18-40.
Talati, N. R., Mehta, K. M. and Mathur, C. M. (1959). Importance of soil test in devel-
oping canal irrigation with reference to Chambal command area of Rajasthan. J. In-
dian Soc. Soil Sci. 7: 177-186.
Talati, R. P. (1947). Field experiments on the reclamation of salt lands in Baramati of
Bombay Deccan. Indian J. Agric. Sci. 17: 153-174.
Tamhane, V. A. (1920). Comparison of salt lands in the Deccan and Sind. Agric. J. In-
dia 15: 410-417.
Tamhane, V. A. and Krishnan, V. G. (1930). Leaching of alkali soils at Sarkand-Sind
with different calcium salts. Trans. 2nd Int. Cong. Soil Sci. 5: 352.
Tiwari, K. N., Dwivedi, B. S., Upadhyay, G. P. and Pathak, A. N. (1985). Sedimentary
iron pyrites as amendment for sodic soils. Proc. Seminar on Reclamation of Salt Af-
fected Soils, Lucknow, Sept. 15-16, pp. 77-92.
Tiwari, K. N. and Sharma, D. N. (1989). Soil Salinity Research. AICRP on Manage-
ment of Salt Affected Soils, C. S. Azad Univ. of Agric. and Technology, Kanpur.
89 p.
Tyagi, N. K. (1998). Management of salt affected soils. In: Fifty Years of Natural Re-
source Management Research in India. G. B. Singh and B. R. Sharma (Eds.).
ICAR, New Delhi, pp. 363-401.
230 CROP PRODUCTION IN SALINE ENVIRONMENTS

Uppal, H. L. (1955). Green manuring with special reference to Sesbania aculeata for
treatment of alkali soils. Indian J. Agric. Sci. 25: 211-235.
Verma, K. S. and Abrol, I. P. (1980). A comparative study of the effect of gypsum and
pyrites on soil properties and the yields of rice and wheat grown in highly sodic
soils. Int. Symp. on Salt Affected Soils, Karnal, pp. 330-338.
Downloaded by [National Institute of Abiotic Stress Management] at 21:59 24 September 2013

Verma, S. K. and Gupta, R. K. (1984). Effectiveness of pyrites in reclaiming sodic clay


soil under laboratory conditions. Z. Pfflanzenernaehr. Bodenk. 147: 680-686.
Verma, S. K., Gupta, R. K. and Dubey, S. K. (1985). Comparative effectiveness of
some chemical amendments in reclaiming a sodic Vertisol. J. Indian Soc. Soil Sci.
33: 948-950.
Verma, S. K. and Sharma, O. P. (1998a). Reclamation of sodic Vertisols under rainfed
conditions through rain water conservation practices. In: Nat. Seminar on Techno-
logical Advances in Agriculture and Livestock Production in M. P. held at Jabalpur.
Verma, S. K. and Sharma, O. P. (1998b). Soil water relationship and salt movement in
black alkali soil. In: 25 Years of Research on Management of Salt Affected Soils and
Use of Saline Water in Agriculture. P. S. Minhas, O. P. Sharma and S. G. Patil
(Eds.). Central Soil Salinity Research Institute, Karnal, pp. 53-63.
Verma, S. K., Shrivastava, N. C. and Tiwari, S. C. (1993). Effect of soil salinity and
depth of irrigation water on Wheat in black clay soil. J. Indian Soc. Soil Sci. 41:
725-729.
Whitecomb, E. (1971). Agarian Conditions in Northern India. Vol. 1: The United
Provinces under British Rule 1896-1900. Thompson Press (India) Ltd., New Delhi.
Yadav, J. S. P. (1981). Salt affected soils and their management in India. Agrokem
Talajat. 30 (Suppl.): 47-62.
Yadav, J. S. P. and Agarwal, R. R. (1959). Dynamics of soil changes in the reclamation
of saline alkali soils of the Indo-Gangetic alluvium. J. Indian Soc. Soil Sci. 7:
213-222.

View publication stats

You might also like