You are on page 1of 55

C H A P T E R S E V E N

Ameliorating Soil Acidity of Tropical


Oxisols by Liming For Sustainable
Crop Production
N. K. Fageria* and V. C. Baligar†

Contents
1. Introduction 346
2. Distribution and Characteristics of Oxisols 350
3. Beneficial Effects of Liming 352
3.1. Neutralizing soil acidity and improving supply of calcium
and magnesium 353
3.2. Reducing phosphorus immobilization 355
3.3. Improving activities of beneficial microorganisms 356
3.4. Reducing solubility and leaching of heavy metals 358
3.5. Improving soil structure 360
3.6. Improving nutrient use efficiency 360
3.7. Controling plant diseases 361
3.8. Mitigating nitrous oxide emission from soils 362
4. Disadvantages of Overliming 363
5. Factors Affecting Lime Requirements 364
5.1. Quality of liming material 365
5.2. Soil texture 366
5.3. Soil fertility 367
5.4. Crop rotation 368
5.5. Use of organic manures 369
5.6. Conservation tillage 370
5.7. Crop species and genotypes within species 371
5.8. Interaction of lime with other nutrients 372
6. Criteria to Determine Liming Material Quantity 375
6.1. Soil pH 376
6.2. Base saturation 377
6.3. Exchangeable aluminum, calcium, and magnesium levels 379

* National Rice and Bean Research Center of EMBRAPA, Caixa Postal 179, Santo Antônio de Goiás,
GO, CEP. 75375-000, Brazil
{
USDA-ARS-Sustainable Perennial Crops Lab, Beltsville, Maryland 20705-2350

Advances in Agronomy, Volume 99 # 2008 Elsevier Inc.


ISSN 0065-2113, DOI: 10.1016/S0065-2113(08)00407-0 All rights reserved.

345
346 N. K. Fageria and V. C. Baligar

6.4. Aluminum saturation 381


6.5. Crop responses 383
7. Methods, Frequency, Depth, and Timing of Lime Application 384
8. Conclusions 386
Acknowledgment 388
References 389

Abstract
The greatest potential for expanding the world’s agricultural frontier lies in the
savanna regions of the tropics, which are dominated by Oxisols. Soil acidity and
low native fertility, however, are major constraints for crop production on
tropical Oxisols. Soil acidification is an ongoing natural process which can be
enhanced by human activities or can be controlled by appropriate soil manage-
ment practices. Acidity produces complex interactions of plant growth-limiting
factors involving physical, chemical, and biological properties of soil. Soil
erosion and low water-holding capacity are major physical constraints for
growing crops on tropical Oxisols. Calcium, magnesium, and phosphorous
deficiencies or unavailabilities and aluminum toxicity are considered major
chemical constraints that limit plant growth on Oxisols. Among biological
properties, activities of beneficial microorganisms are adversely affected by
soil acidity, which has profound effects on the decomposition of organic matter,
nutrient mineralization, and immobilization, uptake, and utilization by plants,
and consequently on crop yields. Liming is a dominant and effective practice to
overcome these constraints and improve crop production on acid soils. Lime is
called the foundation of crop production or ‘‘workhorse’’ in acid soils. Lime
requirement for crops grown on acid soils is determined by the quality of liming
material, status of soil fertility, crop species and cultivar within species, crop
management practices, and economic considerations. Soil pH, base saturation,
and aluminum saturation are important acidity indices which are used as a basis
for determination of liming rates for reducing plant constraints on acid soils.
In addition, crop responses to lime rate are vital tools for making liming
recommendations for crops grown on acid soils. The objective of this chapter
is to provide a comprehensive and updated review of lime requirements for
improved annual crop production on Oxisols. Experimental data are provided,
especially for Brazilian Oxisols, to make this review as practical as possible for
improving crop production.

1. Introduction
Soil acidity is one of the most yield-limiting factors for crop production.
Land area affected by acidity is estimated at 4 billion ha, representing 30%
of the total ice-free land area of the world (Sumner and Noble, 2003).
About 16.7% of Africa, 6.1% of Australia and New Zealand, 9.9% of
Ameliorating Soil Acidity 347

Europe, 26.4% of Asia, and 40.9% of America have acid soils (Von Uexkull
and Mutert, 1995). They cover a significant part of at least 48 developing
countries located mainly in tropical areas, being more frequent in Oxisols
and Ultisols in South America and in Oxisols in Africa (Narro et al., 2001).
In tropical South America, 85% of the soils are acidic, and 850 million ha
of this area is underutilized (Fageria and Baligar, 2001). Acid and low
fertility Oxisols and Ultisols cover about 43% of the tropics (Sanchez and
Logan, 1992). Most of the central part of Brazil is tropical savanna, known as
the cerrado, and covers about 205 million ha or 23% of the country. Most of
the soils in this region are Oxisols (46%), Ultisols (15%), and Entisols (15%),
with low natural soil fertility, high aluminum saturation, and high P fixation
capacity (Fageria and Stone, 1999). Although low fertility is characteristic of
acid soils, these vast areas have a large proportion of favorable topography for
agriculture, adequate temperatures for plant growth throughout the year,
sufficient moisture availability year round in 70% of the region, and for 6–9
months in the remaining 30% of the region (Narro et al., 2001). When the
chemical constraints are eliminated by liming and using adequate amounts
of fertilizers, the productivity of Oxisols and Ultisols is among the highest in
the world (Sanchez and Salinas, 1981).
Theoretically, soil acidity is quantified on the basis of hydrogen (Hþ) and
aluminum (Al3þ) concentrations of soils. For crop production, however,
soil acidity is a complex of numerous factors involving nutrient/element
deficiencies and toxicities, low activities of beneficial microorganisms, and
reduced plant root growth which limits absorption of nutrients and water
(Fageria and Baligar, 2003a). In addition, acid soils have low water-holding
capacity and are subject to compaction and water erosion (Fageria and
Baligar, 2003a). The situation is further complicated by various interactions
among these factors (Foy, 1992). The components of the soil acidity complex
have been thoroughly discussed in various publications (Foy, 1984, 1992;
Kamprath and Foy, 1985; Tang and Rengel, 2003).
Soils become acidic for several reasons. The most common source of
hydrogen is the reaction of aluminum ions with water. The equation for this
reaction in very acid soils (pH<4.0) is:

Al3þ þ H2 O , AlðOHÞ2þ þ Hþ
The species of aluminum ions present vary with pH. Potassium chloride
extracted Al and Al saturation has an inverse relationship with pH (Chartres
et al., 1990; Kariuki et al., 2007). Increased soil acidity causes solubilization
of Al, which is the primary source of toxicity to plants at pH below 5.5
(Bohn et al., 2001; Carson and Dixon, 1979; Ernani et al., 2002; Kariuki et al.,
2007; Parker et al., 1989). The forms of aluminum are mostly exchangeable
Al3þ under very acidic conditions (pH<4.5) to aluminum-hydroxyl ions at
higher pH (4.5–6.5) (Carson and Dixon, 1979). In general, the net positive
348 N. K. Fageria and V. C. Baligar

charge of the hydroxyl aluminum species decreases as the pH increases and


then becomes negative in the alkaline pH range. The species of aluminum
ions generates hydrogen ions through a series of hydrolysis reactions shown
below (Lindsay, 1979):

Al3þ þ H2 O , AlðOHÞ2þ þ Hþ
Al3þ þ 2H2 O , AlðOHÞ þ
2 þ 2H
þ

Al3þ þ 3H2 O , AlðOHÞ 03 þ 3Hþ

Al3þ þ 4H2 O , AlðOHÞ 


4 þ 4H
þ

þ
Al3þ þ 5H2 O , AlðOHÞ 2
5 þ 5H

The exchangeable Al3þ precipitates as insoluble Al hydroxyl species as


pH increases and is reported to decrease 1000-fold for each unit increase in
pH (Lindsay, 1979). However, at pH values greater than 6.5, Al becomes
increasingly soluble as negatively charged aluminates form (Haynes, 1984).
The Al(OH)2þ species is of minor importance and exists over only a narrow
pH range. The Al3þ ion is predominant below pH 4.7, Al(OH)þ2 between
pH 4.7 and pH 6.5, Al(OH)30 between pH 6.5 and pH 8.0, Al(OH)4
above pH 8.0, and Al (OH)52 species occurs at pH values above those
usually found in soils (Bohn et al., 2001).
Soils become acidic due to the parent material being acidic and naturally
low in the basic cations such as Ca2þ, Mg2þ, Kþ, and Naþ or due to leaching
of these elements down the soil profile by excess rains. This situation is
common in high rainfall areas, where precipitation exceeds evaporation, and
leads to leaching. Soil acidity may also be produced by long-time use of
ammonium fertilizers, removal of cations in the harvested portion of crops
and leaching process, and release of organic acids in decomposition of
crop residues and added organic wastes (Fageria et al., 1990; Sparks, 2003).
Use of adequate amounts of nitrogen fertilizer is fundamental for higher yield
of crops under all ecosystems. Urea and ammonium sulfate are dominant
nitrogen carriers used for crop production around the world. The acidification
of soils by using the ammonium form of nitrogen fertilizers can be explained
by the following equation:

NHþ 
4 þ 2O2 , NO3 þ H2 O þ 2H
þ

The oxidation of NH4þ in the above equation is known as nitrification


and hetrotrophic and autrotrophic bacterias can carry it out. The most
important autrotrophic genera of bacteria are Nitrosomonas and Nitrobacter.
Use of legume crops continuously or in rotation can increase soil acidity.
Ameliorating Soil Acidity 349

In Australia and New Zealand, continuous cultivation of legume crops


decreased the pH of agricultural soils (Bolan and Hedley, 2003). Legume-
based pastures also increased soil acidification (Loss et al., 1993; Williams,
1980). Williams (1980) reported that even the normal growth of clover
pasture for 50 years decreased the pH of an Australian soil from 6.0 to 5.0 at
a depth of 30 cm. Legumes also increase soil acidification in arable cropping
systems (Burle et al., 1997). The reason for generating acidity is associated
with higher absorption of basic cations by these crops and the release of Hþ
ions by the root of legume crops to maintain ionic balance (Bolan and
Hedley, 2003). According to Bolan and Hedley (2003), for different legume
species about 0.2–0.7 mol of Hþ were released per mol N2 fixed. These
authors also state that the amount of Hþ ions released during N2 fixation is
really a function of carbon assimilation and hence depends mainly on the
form and amount of amino acids and organic acids synthesized within the
plants. Soil acidification is also caused by the release of protons (Hþ) during
the transformation and cycling of carbon, nitrogen, and sulfur in the soil–
plant–animal system (Bolan and Hedley, 2003; Robson, 1989). Soil acidifi-
cation is caused by acid precipitation, the result of industrial pollution (Foy,
1984; Ulrich et al., 1980).
For a long time, acid soils have been considered less suitable for produc-
tive agriculture. However, the generation of modern technology, through
intensive research, has brought forth a new reality of increased productivity
of grains and other food, fiber and feed crops, pastures, and energy products
on these soils. Borlaug and Dowswell (1997) concluded that acid lands are no
longer a marginal agriculture frontier, but the most extensive agriculture
frontier of the world, providing hope for adequate food supply and a better
quality of life for millions of people, especially in the tropics. Within the last
few decades, significant advances have been made in management of acid
soils of the tropics for improving pasture for cattle raising and increasing
productivity of annual and plantation crops (Fageria and Baligar, 2003a;
Sumner and Noble, 2003). However, there still remains much to be done
to develop technologies that are economically viable, environmentally
sound, and socially acceptable.
For increase in food supply on acid soils, sustainable cropping systems are
essential for agronomic, economic, and environmental reasons. Sustainable
crop production is defined as practice that over the long term enhances
environmental quality and the resource base on which agriculture depends,
provides for basic human food and fiber needs, is economically viable, is
social acceptable, and improves the quality of life for farmers and society as a
whole (White et al., 1994). Adequately limed soils enhance the sustainability
of cropping systems because of higher crop yields, lower cost of production,
and reduced environmental pollution. World population reached more
than 6 billion in 2007, and is expected to increase to 8 billion by 2025
and over 9 billion by 2050. Over half of the world population currently lives
350 N. K. Fageria and V. C. Baligar

in regions dominated by acid soils (Yang et al., 2004). The greatest potential
for expanding the world’s agricultural frontier lies in the tropical savanna
regions dominated by acid, infertile soils classified mainly as Oxisols and
Ultiosls (Sanchez and Salinas, 1981). Furthermore, about 95% of the current
population growth has taken place in tropical regions; therefore, a continuing
increase in food production is required to meet this demand (Hartemink,
2002). Liming is the most important and most effective practice to ameliorate
soil acidity constraints for optimal crop production (Haynes, 1984). The
practice of well-planned and execution of liming under these situations is
fundamental for increasing crop yield on acid soils. The magnitude of the soil
acidity problem of Oxisols in many areas around the world, and the potential
that these soils offer in increasing the production of food and fiber, provides a
focus for the objectives of this review, that is, to review the nature, causes,
and management of Oxisols acidity in order that crop productivity on these
acid soils might be improved for the benefit of humankind.

2. Distribution and Characteristics of Oxisols


The theory concerning the effects of parent materials on soil formation
originated in Russia during the late 19th century. According to this theory,
soil formation is a function of climatic conditions, biotic activities, topogra-
phy, and time span (West et al., 1998). Soil profile development processes
are known as pedogenic and involve additions, losses, translocations, and
transformations of soil materials (Brady and Weil, 2002). Simonson (1959)
outlined a generalized theory of soil genesis in which he proposed to
consider soil formation as consisting of two overlapping steps: the accumu-
lation of parent materials and the differentiation of horizons in the solum.
The former is largely a geochemical or geogenesis process, whereas the
latter is a pedogenesis process (Van Wambeke, 1991; West et al., 1998).
Information on the extent and distribution of Oxisols in various parts of
the world has increased in recent years. However, soil survey is still incom-
plete due to lack of infrastructure and low population density in some areas
where Oxisols occur (Buol and Eswaran, 2000). Table 1 shows approximate
areas of Oxisols at global level and in the tropical regions compared with
other soil orders. In the tropics, Oxisols occupy a much larger area than do
other soil orders. Acid and low fertility soils meeting such stereotypic
concept of ‘‘tropical soils’’ are mainly classified as Oxisols and Ultisols.
Because of recent global glaciaton, these soils cover only 7% of the temperate
region but 43% of the tropics (Sanchez and Logan, 1992). The largest
concentration of Oxisols occurs in the South American savannas, the eastern
Amazon, and parts of central Africa (Sanchez and Salinas, 1981). Distribution
Ameliorating Soil Acidity 351

Table 1 Approximate distribution of soil orders globally and in tropical regions

Percentage Area in the Percentage


Global area of total tropics of total
Soil order (million ha) global area (million ha) tropical area
Alfisols 1790 12.3 559 15.2
Andisols 144 1.0 43 1.2
Aridisols 2276 15.6 87 2.4
Entisols 2730 18.7 574 15.6
Gelisols 1769 12.1 – –
Histosols 240 1.6 36 0.9
Inceptisols 1547 10.6 532 14.4
Mollisols 1100 7.5 74 2.0
Oxisols 840 5.8 833 23.0
Spodosols 478 3.3 20 0.5
Ultisols 1347 9.3 749 20.4
Vertisols 311 2.2 163 4.4
Source: Sanchez and Logan (1992), Buol and Eswaran (2000), and Brady and Weil (2002).

Table 2 Distribution of oxisols and ultisols in South America

Total area of
Country Area (106 ha) the country (%)
Brazil 572.71 68
Colombia 67.45 57
Peru 56.01 44
Venezuela 51.64 58
Bolivia 39.54 57
Guiana 12.25 62
Surinam 11.43 62
Paraguay 9.55 24
Equador 8.61 23
French Guiana 8.61 94
Chile 1.37 2
Argentina 1.28 0.4
Source: Cochrane (1978).

of Oxisols and Ultisols in South American countries is shown in Table 2.


Extensive areas of Southeast Asia are covered by highly weathered Oxisols
and Ultisols (Ismail et al., 1993; Supriyo et al., 1992; Van Wambeke, 1991;
Von Uexkull and Mutert, 1995). Sanchez and Salinas (1981) reported that
about 15 million ha of Oxisols is found in tropical Asia. Baligar and Ahlrichs
352 N. K. Fageria and V. C. Baligar

(1998), Baligar et al. (2004), Van Wambeke (1991), and Von Uexkull
and Mutert (1995) have reported extensively on the distribution, nature,
and properties of acid soils.
Oxisols are also referred to as latosols, laterite, Ferralsols, and red earths.
Oxisols are defined as mineral soils having an oxic horizon which is
characterized by the virtual absence of weathering primary minerals or 2:1
layer silicate clays, the presence of 1:1 layer silicate clays and highly insoluble
minerals such as quartz sand, the presence of hydrated oxides of iron and
aluminum, and the absence of water-dispersible clay (Soil Science Society of
America, 1997). Oxisols are highly weathered acidic soils, having low basic
cation and effective cation exchange capacity (ECEC). The cation exchange
capacity (CEC) refers to the value obtained with 1 M NH4OAc (pH 7), and
ECEC is the sum of exchangeable cations (Al3þ, Ca2þ, Mg2þ, Kþ, and
Naþ) (Sumner and Noble, 2003). These soils have good physical properties
(except low water-holding capacity and susceptibility to erosion) and have
uniform distribution of clay with increasing depth. On average, these soils
have clay contents of more than 300 g kg1. However, the clays are of low
activity and have a limited capacity to hold basic cations. These soils are very
low in plant-available phosphorus and due to high concentration of iron and
aluminum oxides having high P fixation capacity; selected chemical proper-
ties of Brazilian Oxisols are presented in Tables 3 and 4.

3. Beneficial Effects of Liming


Modern agriculture production requires the implementation of effi-
cient, sustainable, and environmentally sound management practices. In this
context, liming is an important practice to achieve optimum yields of all crops
grown on acid soils. Liming is the most widely used long-term method of soil

Table 3 Selected chemical properties of Oxisols of the cerrado region of Brazil

Organic Mg
pH in matter P K Ca (cmolc
Value H2O (g kg1) (mg kg1) (mg kg1) (cmolc kg1) kg1)
Minimum 4.8 8 0.6 12 0.54 0.22
Maximum 6.3 31 14.8 73 3.42 2.42
Mean 5.5 21.5 4.12 31.6 1.77 0.84
S.D. 0.28 5.1 2.93 13.7 0.70 0.40
Source: Compiled from Fageria and Breseghello (2004).
Values are averages of 43 soil samples collected from the State of Mato Grosso.
Ameliorating Soil Acidity 353

Table 4 Selected soil acidity indices of Oxisols of cerrado region of Brazil

Base Al Ca Mg K
saturation saturation Saturation saturation saturation
Value (%) ((%) (%) (%) (%)
Minimum 9 0 5 2 0.25
Maximum 60 26 34 25 1.64
Mean 28 8 18 9 0.82
SD 10.78 6.33 7.36 4.03 0.31
Source: Compiled from Fageria and Breseghello (2004).

acidity amelioration, and its success is well documented (Conyers et al., 1991;
Haynes, 1982; Kaitibie et al., 2002; Scott et al., 2001). Application of lime at
an appropriate rate brings several chemical and biological changes in the soil,
which are beneficial or helpful in improving crop yields on acid soils.
Adequate liming eliminates soil acidity and toxicity of Al, Mn, and H;
improves soil structure (aeration); improves availabilities of Ca, P, Mo, and
Mg, pH, and N2 fixation; and reduces the availabilities of Mn, Zn, Cu, and Fe
and leaching loss of cations. These beneficial effects are discussed in the
following section.

3.1. Neutralizing soil acidity and improving supply of calcium


and magnesium
Liming raises soil pH, base saturation, and Ca and Mg contents, and reduces
aluminum concentration in Brazilian Oxisols (Fageria, 2000, 2001a; Fageria
and Stone, 2004). The changes in these chemical properties with the use of
dolomitic lime [CaMg (CO3)2] can be explained on the basis of following
equation (Fageria and Baligar, 2005a):

CaMgðCO3 Þ2 þ 2Hþ , 2HCO 


3 þ Ca

þ Mg2þ
2HCO þ
3 þ 2H , 2CO2 þ 2H2 O

CaMgðCO3 Þ2 þ 4Hþ , Ca2þ þ Mg2þ þ 2CO2 þ 2H2 O

The above equations show that acidity-neutralizing reactions of lime occur


in two steps. In the first step, Ca and Mg react with H on the exchange
complex and H is replaced by Ca2þ and Mg2þ on the exchange sites (nega-
tively charged particles of clay or organic matter), forming HCO 3 . In the
second step, HCO 3 reacts with Hþ to form CO and H O to increase pH.
2 2
354 N. K. Fageria and V. C. Baligar

Soil moisture and temperature and quantity and quality of liming material
mainly determine the reaction rate of lime. To get maximum benefits from
liming or for improving crop yields, liming materials should be applied in
advance of crop sowing and thoroughly mixed into the soil to enhance its
reaction with soil exchange acidity. Selected soil chemical properties influenced
by applied lime in a Brazilian Oxisols are presented in Table 5.
Highly weathered tropical soils such as Oxisols have very low levels of
exchangeable Ca, and crops grown on such soils exhibit Ca deficiency
when exchangeable Ca is <1 cmol/kg (Cregan et al., 1989; Kamprath,
1984). Application of limestone (calcium carbonate) and dolomitic lime
(Ca and Mg bicarbonate) increases soil-exchangeable Ca and Ca and Mg,
respectively. Plant growth improvement in acid soil is not due to addition of
basic cations (Ca, Mg), but because of increasing pH reduces toxicity of
phytotoxic levels of Al and Mn. Increasing concentration of Ca2þ, Kþ, and
Hþ in the soil decreases plant Mg uptake due to competitive inhibition
(Clark, 1984; Marschner, 1995).

Table 5 Influence of liming on selected soil chemical properties of Oxisols at 0–10


and 10–20 cm depth

Lime rate (Mg ha1)


Soil property 0 12 24 F test CV (%)
0–10 cm depth
pH (1:2.5 soil water) 5.4c 6.7b 7.1a ** 2.0
Base saturation (%) 28.9c 72.0b 84.9a ** 19
H þ Al (cmolc kg1) 6.5a 2.0b 1.0c ** 14
Acidity saturation (%) 71.1a 27.8b 15.1c ** 14
Ca (cmolc kg1) 1.8c 3.6b 4.3a ** 9
Mg (cmolc kg1) 0.6b 1.3a 1.3a ** 12
10–20 cm depth
pH 5.3c 6.2b 6.5a ** 3.0
Base saturation (%) 22.9c 51.2b 61.7a ** 24
H þ Al (cmolc kg1) 6.6a 3.8b 2.9c ** 13
Acidity saturation (%) 77.1a 49.0b 38.3c ** 12
Ca (cmolc kg1) 1.4c 2.7b 3.3a ** 15
Mg (cmolc kg1) 0.4b 1.0a 1.1a ** 15
** Significant at the 1% probability level, respectively. Means followed by the same letter in the same
line under different lime treatments are not statistically significant at the 5% probability level by
Tukey’s test.
Source: Fageria (2006).
Ameliorating Soil Acidity 355

3.2. Reducing phosphorus immobilization


Oxisols are naturally deficient in total and plant-available phosphorus and
significant portions of applied P are immobilized due to either precipitation
of P as insoluble Fe/Al-phosphates or chemisorption to Fe/Al-oxides and
clay minerals (Nurlaeny et al., 1996). Smyth and Cravo (1992) reported that
Oxisols are notorious for P immobilization because they have higher iron
oxide contents in their surface horizons than do any other kind of soil. The
P fixation capacity in Oxisols is directly related to the surface area, and clay
contents of the soil material, and inversely related to SiO2/R2O3 ratios
(Curi and Camargo, 1988).
Bolan et al. (1999) reported that in variable charge soils, a decrease in pH
increases the anion exchange capacity, thereby increasing the retention of P.
Hence, improving crop yields on these soils requires high rates of P
application (Fageria, 1989; Sanchez and Salinas, 1981). Reports regarding
the effects of liming on P availability in highly weathered acid soils are in
conflict (Friesen et al., 1980a; Haynes, 1984). Liming can increase, decrease,
or have no effect on P availability (Anjos and Rowell, 1987; Curtin and
Syers, 2001; Fageria, 1984; Haynes, 1982; Mahler and McDole, 1985).
However, in a recent study, Fageria and Santos (2008) reported a linear
increase in Mehlich 1 extractable P with increasing soil pH in the range of
5.3–6.9 (average of 0–10 and 10–20 cm soil depth) in Brazilian Oxisols
(Fig. 1). Mansell et al. (1984) and Edmeades and Perrott (2004) reported that
in acid soils of New Zealand, primary benefit of liming occurs through an
increase in the availability of P by decreasing P adsorption and stimulating
the mineralization of organic P. Fageria (1984) also reported that in Brazi-
lian Oxisols there was a quadratic increase in the Mehlich 1 extractable P in
the pH range of 5.0–6.5, and thereafter it was decreased. Increase in
availability of P in the pH range of 5.0 to 6.5 was associated with release
of P ions from Al and Fe oxides, which were responsible for P fixation
(Fageria, 1989). At higher pH (>6.5), the reduction of extractable P was
associated with precipitation of P as Ca phosphate (Naidu et al., 1990).
These increases in extractable P or liberation of this element in the pH range
of 5.0–6.5 and reduction in the higher pH range (>6.5) can be explained by
the following equations:

AlPO4 ðP fixedÞ þ 3OH , AlðOHÞ3 þ PO 3


4 ðP releasedÞ

CaðH2 PO4 Þ2 ðsoluble PÞ þ 2Ca2þ , Ca3 ðPO4 Þ2 ðinsoluble PÞ þ 4Hþ

The liming acid soils result in the release of P for plant uptake; this effect
is often referred to as ‘‘P spring effect’’ of lime (Bolan et al., 2003). Bolan et al.
(2003) reported that in soils high in exchangeable and soluble Al, liming
might increase plant P uptake by decreasing Al, rather than by increasing
356 N. K. Fageria and V. C. Baligar

Average of two soil depths


40

30

20
Y = - 1.5594 + 4.9585X
10 R 2 = 0.5417**
Mehlich-1 extractable P (mg kg-1)

0
10–20 cm Y = - 5.3937 + 3.9288X
30 R 2 = 0.4032**
20

10

0
0–10 cm
60

50

40

30

20 Y = - 3.5026 + 5.7293X
R 2 = 0.4462**
10

0
5.0 5.5 6.0 6.5 7.0 7.5
Soil pH in H2O

Figure 1 Relationship between soil pH and Mehlich 1 extractable soil phosphorus


(Fageria and Santos, 2008).

P availability per se. This may be due to improved root growth where Al
toxicity is alleviated, allowing a greater volume of soil to be explored (Friesen
et al., 1980b).

3.3. Improving activities of beneficial microorganisms


Soil microbiological properties can serve as soil quality indicators because
soil microorganisms are the second most important (after plants) biological
agents in the agricultural ecosystem (Fageria, 2002; Yakovchenko et al.,
1996). Soil microorganisms provide the primary driving force for many
chemical and biochemical processes and thus affect nutrient cycling, soil
fertility, and carbon cycling (He et al., 2003). Plant roots and rhizosphere are
colonized by many plant-beneficial microorganisms such as symbiotic and
nonsymbiotic dinitrogen (N2) fixing bacteria, plant growth promoting
Ameliorating Soil Acidity 357

rhizobacteria, saprophytic microorganisms, biocontrol agents, and mycorrhi-


zae and free living fungi. Soil acidity restricts the activities of these beneficial
microorganisms, except fungi, which grow well over a wide range of soil pH
(Brady and Weil, 2002). Enhancing the activities of beneficial microbes such as
rhizobia, diazotropic bacteria, and mycorrhizae in the rhizosphere has
improved root growth by the fixation of atmospheric nitrogen, suppressing
pathogens, and producing phytohormones, enhancing root surface area
to facilitate uptake of less mobile nutrients such as P and micronutrients
and mobilizing and solubilizing unavailable nutrients (Baligar and Fageria,
1999).
Literature provides ample evidence that low soil pH adversely affects
activities of rhizobium, including a loss of its ability to fix nitrogen (Angle,
1998). Mulder et al. (1977) showed that low soil pH reduced the activity and
their ability to multiply. Holdings and Lowe (1971) further demonstrated
that low soil pH increased the number of ineffective rhizobia in soil. Angle
(1998) reported that soil pH decline below 5.5 reduced rhizobial populations
and rhizobia that survive such a pH lack the capacity to fix atmospheric
nitrogen. Ibekwe et al. (1995) showed that plants grown in an unamended
control soil with low pH often exhibited low rates of nitrogen fixation.
Ibekwe et al. (1995) also reported high rates of nitrogen fixation when high
concentrations of heavy metals were present, but soil pH was nearly neutral.
Franco and Munns (1982) found that decreasing the pH of nutrient solutions
from 5.5 to 5.0 decreased the number of nodules formed by common bean.
Bacterias are divided into three groups based on their tolerance to soil
acidity. The first group is known as acidophiles (grow well under acidic
conditions), the second group is known as neutrophiles (grow well under
neutral pH), and the third group is known as alkaliphiles (grow well
under alkaline conditions) (Glenn et al., 1997). Soil contains all these groups
of bacteria. Most soil bacterias, however, including the nitrogen fixing rhizo-
bium, belong to the neutrophiles group (Glenn et al., 1997). Therefore, acidic
pH ranges are detrimental to bacterial activities.
Lime ameliorates the harmful effects of soil acidity (Cregan et al., 1989).
Studies on bacteria suggest that the success of liming may be due not only to
an effect on the soil pH but also to a direct effect on the bacteria themselves
(Reeve et al., 1993). Like most neutrophilic bacteria, rhizobia appear to
maintain an intracellular pH between 7.2 and 7.5, even when the external
environment is acidic (Kashket, 1985). However, differences exist between
and among acid-tolerant and acid-sensitive strains (Bhandhari and Nicholas,
1985; Graham et al., 1994). Strain tolerance to lower pH has been reported
for rhizobia (Brockwell et al., 1995) and arbuscular mycorrhizal fungi
(Habte, 1995; Siqueira and Moreira, 1997). Muchovej et al. (1986) reported
that liming of a Brazilian Oxisols improved nodule formation in soybean.
Haynes and Swift (1988) reported that liming increased microbial biomass
and enzyme activity in acid soils.
358 N. K. Fageria and V. C. Baligar

Nurlaeny et al. (1996) found that liming increased shoot dry weight, total
root length, and mycorrhizal colonization of roots in soybean and corn
grown on tropical acid soils. These authors also reported that mycorrhizal
colonization improved P uptake and plant growth. Furthermore, coloniza-
tion of the roots with arbuscular mycorrhizal fungi can increase plant uptake
of P and other nutrients with low mobility, such as Zn and Cu (Marschner
and Dell, 1994). Uptake of P by mycorrhizal plants is usually from the same
labile soil P pool from which the roots of nonmycorrhizal plants take up P
(Morel and Plenchette, 1994). However, the external hyphae can absorb
and translocate P to the host from soil outside the root depletion zone of
nonmycorrhizal roots (Johansen et al., 1993). Thereby, under conditions
where P and other nutrients’ diffusion in soil is slow and root density not
very high, more immobile nutrients are spatially available to mycorrhizal
plants than to nonmycorrhizal plants (Nurlaeny et al., 1996). Furthermore,
mycorrhizal plants may utilize organic soil P due to surface phosphatase
activity of hyphae (Tarafdar and Marschner, 1994, 1995), enhanced activity
of P-solubilizing bacteria in the mycorrhizosphere (Linderman, 1992;
Tarafdar and Marschner, 1995), or reduced immobilization of labile P in
organic matter (Joner and Jakobsen, 1994). Additionally, soil pH affects
arbuscular mycorrhizal colonization (Mamo and Killham, 1987; Wang et al.,
1985), species distribution (Porter et al., 1987), and effectiveness of the
mycorrhizal symbiosis with plant species (Hayman and Tavares, 1985).

3.4. Reducing solubility and leaching of heavy metals


Heavy metals are those metals having densities >5.0 Mg m3 (Soil Science
Society of America, 1997). In soils, these include the elements Cu, Zn, Fe,
Mn, Cd, Co, Cr, Hg, Ni, and Pb. Higher concentrations of these heavy
metals in soil solution can lead to uptake by crop plants in quantities that are
harmful for human and animal health; leaching can also occur and contami-
nate ground water (Epstein and Bloom, 2005; Hall, 2002). Soil properties
such as organic matter content, clay type, redox status, and soil pH are
considered the major factors determining the bioavailability of heavy metals
in soil (Huang and Chen, 2003; McLean, 1976; Wagner, 1993).
Increasing pH by application of lime to acid soils reduces the solubility of
most heavy metals (Fageria et al., 2002; Lindsay, 1979; Mortvedt, 2000).
In addition, higher soil pH increases the adsorption affinity of iron oxides,
organic matter, and other adsorptive surfaces (Sauve et al., 2000). This
practice can reduce the leaching of heavy metals to ground water as well
as their absorption by plants and consequently improve soil and water
quality and subsequently human health. Data in Tables 6 and 7 show the
influence of increasing base saturation on soil-extractable Mn, Fe, Zn, and
Cu (heavy metals) after harvest of common bean crop grown on Brazilian
Oxisols. These data show that with exception of Cu, concentrations of
Ameliorating Soil Acidity 359

Table 6 Influence of base saturation on DTPA extractable soil Mn and Fe after harvest
of a common bean crop grown on Brazilian Oxisols

Base saturation Mn Base saturation Fe


(%) (mg kg1) (%) (mg kg1)
23.4 8.1a 28.5 33.1a
54.4 6.4b 65.8 21.1b
67.5 5.5b 75.3 20.7b
Means in the same column followed by same letter are not significantly different at 5% probability level
by Tukey’s test.
Source: Fageria and Santos (2005).

Table 7 Influence of base saturation on DTPA extractable soil Zn and Cu after harvest
of common bean crop grown on Brazilian Oxisols

Base saturation Zn Base saturation Cu


(%) (mg kg1) (%) (mg kg1)
26.2 5.5a 23.4 3.7a
60.1 5.4a 56.7 3.7a
69.8 4.8b 66.5 3.5a
Means in the same column followed by same letter are not significantly different at 5% probability level
by Tukey’s test.
Source: Fageria and Santos (2005).

these microelements in the extractant were significantly reduced with


increasing soil base saturation due to liming.
Increasing soil pH with lime can significantly affect the adsorption of
heavy metals in soils (Adriano, 1986; McBride, 1994). Adsorption of Pb,
Cd, Ni, Cu, and Zn is significantly decreased with increasing soil pH (Basta
and Tabatabai, 1992; Elliott et al., 1986; Harter, 1983). Ribeiro et al. (2001)
reported that among several soil amendments (dolomitic lime, gypsum,
vermicompost, sawdust, and solomax), dolomitic lime was most effective
in reducing bioavailabilities of these heavy metals: Zn, Cd, Cu, and Pb.
Most metals are relatively more mobile under acidic and oxidizing condi-
tions and are strongly retained under alkaline and reducing conditions
(Huang and Chen, 2003). Brummer and Herms (1983) reported that Pb,
Cd, Hg, Co, Cu, and Zn are more soluble at pH 4.0–5.0 than in a pH range
of 5.0–7.0. However, under acidic conditions, As, Se, and Mo are less
soluble because of formation of these elements into anionic forms (Huang
and Chen, 2003).
The risk of food-chain contamination by toxic substances and elements
has been a major concern of both producers and consumers (Treder and
360 N. K. Fageria and V. C. Baligar

Cieslinski, 2005). The concentrations of heavy metals in plants and their


distribution to various parts of the plant are the result of the combined
influence of soil properties and biological factors. However, translocation of
heavy metals from the soil solids to the soil solution, and thus their avail-
ability to plants, depends upon several factors. Soil properties such as organic
matter content, clay type, redox potential, and soil pH are considered the
major factors that determine the bioavailability of heavy metals in soil
(Treder and Cieslinski, 2005). Hence, liming certainly helps in reducing
availability of heavy metals to crop plants.

3.5. Improving soil structure


The clustering of soil particles (sand, silt, and clay) into aggregates or peds
and their arrangement into various patterns resulted in what is termed soil
structure. From the agronomic standpoint, soil structure affects plant
growth through its influence on infiltration, percolation, and retention of
water, soil aeration, and mechanical impedance to root growth. Its general
role in soil–water relations can be evaluated in terms of the extent of soil
aggregation, aggregate stability, and pore size distribution. These soil char-
acteristics change with tillage practices and cropping systems. The major
binding agents responsible for aggregate formation are the silicate clays,
oxides of iron and aluminum, and organic matter and its biological decom-
position products. The Oxisols and Ultisols, characterized by dominant
quantities of iron and aluminum oxides, have a high degree of aggregation
and the aggregates are quite stable. The red color of these soils is attributed
to iron oxide minerals.
Calcium in liming materials helps in the formation of soil aggregates,
hence improving soil structure (Chan and Heenan, 1998). The lime-induced
improvement in aggregate stability manifested through the effect of liming
on dispersion and flocculation of soil particles (Bolan et al., 2003). Liming is
often recommended for the successful colonization of earthworm in pasture
soils. The lime-induced increase in earthworm activity may influence soil
structure and macroporosity through the release of polysaccharide and the
burrowing activity of earthworm (Springett and Syers, 1984).

3.6. Improving nutrient use efficiency


Improving nutrient use efficiency is becoming increasingly important in
modern crop production due to rising costs associated with fertilizer inputs
and growing concern about environmental pollution. Nutrient use efficiency
is defined in several ways in the literature. The most common definitions are
known as agronomic efficiency, physiological efficiency, agrophysiological
efficiency, apparent recovery efficiency, and utilization efficiency (Baligar et al.,
2001; Fageria and Baligar, 2003b, 2005a; Fageria et al., 1997). Most of these
Ameliorating Soil Acidity 361

definitions deal with nutrient uptake and utilization by plants in dry matter
production. In a simple way, efficiency is output of economic produce divided
by fertilizer input. This means that a crop species or genotypes of the same
species producing higher dry matter yield with low nutrient application rate or
accumulation are called efficient plant species or genotypes. According to the
Soil Science Society of America (1997), a nutrient-efficient plant is one
that absorbs, translocates, or utilizes more of a specific nutrient than another
plant under conditions of relatively low nutrient availability in the soil or
growth medium. Fischer (1998) reported that rice cultivars released in 1965
produced less than 40 kg of grain per kilogram of N taken up by the plants.
However, cultivars released in 1995 produced almost 55 kg of grain per
kilogram of N take-up, resulting in a more than 35% increase in N efficiency
(Fischer, 1998).
Nutrient use efficiency (recovery efficiency) seldom exceeds 50% in most
grain production systems (Fageria and Baligar, 2005a; Westerman et al.,
2000). Worldwide, N recovery efficiency for cereal production [rice,
wheat, sorghum, millet, barley, corn, oat, and rye is 33% (Raun and
Johnson, 1999]. The low nutrient use efficiency is associated with the use
of low input crop management technology. Such management practices
include use of low rate of fertilizers and lime, water deficiency, inadequate
control of insects, diseases, and weeds, and planting of inefficient plant species
or genotypes of the same species (Fageria, 1992). Low nutrient use efficiency
in crop production systems is undesirable, both economically and
environmentally.
Soil acidity is also responsible for low nutrient use efficiency by crop
plants. Fageria et al. (2004) reported that liming of Oxisols improved the use
efficiency of P, Zn, Cu, Fe, and Mn by upland rice genotypes. Efficiency of
these nutrients was higher under a pH of 6.4 than with pH 4.5. The
improvement in efficiency of these nutrients was associated with decreasing
soil acidity, improving their availability, and enhanced root system
(Fageria et al., 2004).

3.7. Controling plant diseases


Mineral nutrition plays an important role in controling plant diseases.
Healthy plants provided by adequate essential nutrients in appropriate
balance generally have fewer diseases compared with nutrient-deficient
plants (Fageria et al., 1997). When evaluating the effects of mineral nutrition
on plant diseases, in addition to optimal rate of nutrient application an
appropriate nutrient balance is very important. Indiscriminate excess supply
of a nutrient will create imbalance with other nutrients and chances of plant
infestation with diseases will increase. Furthermore, there is still lack of
systematic research data to determine the effects of mineral nutrition
on plant diseases, involving soil scientists and plant pathologists. Plant
362 N. K. Fageria and V. C. Baligar

pathologists without collaboration of soil scientists have conducted most of


the studies in this discipline, and invariably, with very inappropriate fertil-
izer rates (excess or deficient). In addition, most of the research data related
to nutrient–disease interactions are collected under controlled growth
conditions and field studies are lacking.
However, existing information suggests that in some cases liming decreases
diseases and in others the inverse is true. It has been known for nearly 100
years that amending the soil with CaCO3 would provide a significant measure
of clubroot (Plasmodiphora brassicae Wor.) control in crucifers (Engelhard,
1989). Conversely, many diseases of potato (Solanum tuberosum L.), such as
common scab (Streptomyces scabies), powdery scab (Spongospora subterranea),
black scurf (Rhizoctonia solani), and tuber blight (Phytophthera infestans),
are favored at higher pH compared to lower pH (Haynes, 1984).
Calcium has been implicated in plant resistance to several plant patho-
gens, including Erwinia phytophthora, R. solani,Sclerotium rolfsii, and Fusarium
oxysporum (Kiraly, 1976). Haynes (1984) reported that calcium forms rigid
linkages with pectic chains and thus promotes the resistance of plant cell
walls to enzymatic degradation by pathogens.
Information is lacking on effects of lime on damage by specific insects on
plants. Overall, from the available data, it appears that variable levels of
macro- and micronutrients in plants have positive to negative or no effects
on insect damage to crop plants (Fageria and Scriber, 2002). Lime improves
the availability of Ca, Mg, Mo, and P and reduces the availability of Mn,
Zn, Cu, and Fe. It would be interesting to evaluate how such changes in
availabilities of these nutrients in soil affect insect damage in plant.

3.8. Mitigating nitrous oxide emission from soils


Nitrous oxide (N2O) is globally important, due to its role as a greenhouse gas,
and once oxidized to NOx it can catalyze stratospheric ozone destruction.
Nitrous oxide is a potent greenhouse gas with much greater global warming
potential than CO2 (Izaurralde et al., 2004). Nitrous oxide concentration in
the atmosphere has increased since preindustrial times and agricultural lands
are the main anthropogenic sources (Clough et al., 2004; Perez et al., 2001).
The concentration of N2O in the atmosphere, estimated at 2.68102 mL
L1 around 1750, has increased by about 17% as a result of human alterations
of the global N cycle (IPCC, 2001). Global annual N2O emissions from
agricultural soils have been estimated to range between 1.9 and 4.2 Tg N,
with about half arising from anthropogenic sources (IPCC, 2001).
Since soil pH has a potential effect on N2O production pathways, and
the reduction of N2O to N2, it has been suggested that liming may provide
an option for the mitigation of N2O emission from agricultural soils
(Stevens et al., 1998). Clough et al. (2003, 2004) reported that liming
has been promoted as a mitigation option for lowering soil N2O emissions
Ameliorating Soil Acidity 363

when soil moisture content is maintained at field capacity. Nitrous oxide


forms in soils primarily during the process of denitrification (Robertson and
Tiedje, 1987) and to a lesser extent during nitrification (Tortoso and
Hutchinson, 1990).

4. Disadvantages of Overliming
Adequate lime rate is essential not only for maximum economic return
but also for maintaining appropriate nutrient balance for crop production.
Overliming effects are more pronounced on coarse-textured soils than on
fine-textured soils. Fine-textured soils have high buffer capacity; hence,
changes in chemical properties due to liming are not as pronounced as for
coarse-textured soils. Overliming can create deficiencies of micronutrients
like Fe, Mn, Zn, Cu, and B. These nutrients are usually adsorbed onto
sesquioxide soil surfaces. Table 8 summarizes important changes in micro-
nutrient concentrations as influenced by soil pH and consequent acquisition
by plants. Table 9 shows acquisition of Zn, Cu, Fe, and Mn by upland rice
grown at various soil pH levels on Brazilian Oxisols. Uptake of these
micronutrients has been significantly decreased at higher pH levels. Iron
deficiency in upland rice in Brazilian Oxisols is commonly observed with
pH higher than 6.5 (Fageria, 2000; Fageria et al., 1994, 2002). Similarly,
shoot dry matter, grain yield, and number of panicles in upland rice were
significantly decreased with increasing pH in the range of 4.6–6.8 in
Brazilian Oxisols (Table 10).
The reaction responsible for the reduced solubility of Fe with increasing
pH is well understood. It results in the precipitation of Fe (OH)3 as the
concentration of OH ions is increased as indicated by the following reaction:

Fe3þ þ 3OH , FeðOHÞ3 ðprecipitationÞ


The Fe(OH)3 is chemically equivalent to the hydrated oxide, Fe2O3
3H2O. Acidification shifts the equilibrium, causing a greater release of Fe3þ as
a solution ion.
In addition to deficiency of some micronutrients, overliming can also
create problems of Al toxicity. Aluminum solubility is minimal in the pH
range of 5.5–6.5 (Haynes, 1984). At higher pH values, Al becomes increas-
ingly soluble as the negatively charged aluminate form. Farina et al. (1980)
reported increased uptake of Al by corn (Zea mays L.) at near neutral pH
values and such Al uptake decreased corn yield. This type of detrimental
effect of Al can occur at pH higher than 7.0 in water and liming of Oxisols
around pH 6.5 in water can produce maximum yields of most food crops
(Fageria et al., 1997).
364 N. K. Fageria and V. C. Baligar

Table 8 Influence of soil pH on micronutrient concentrations in soil and plant uptake

Micronutrient Influence on concentration/uptake


Zinc Zinc solubility is highly soil pH dependent and decreases
100-fold for each increase in pH, and uptake by plant
decreases as a consequence.
Iron Ferric (Fe3þ) and ferrous (Fe2þ) activities in soil solution
decrease 1000-fold and 100-fold, respectively, for each unit
increase in soil pH. In most oxidized soils, uptake of Fe by
crop plants decreases with increasing soil pH.
Manganese The principal ionic Mn species in soil solution is Mn2þ, and
concentrations decrease 100-fold for each unit increase in
soil pH. In extremely acidic soils, Mn2þ solubility can be
sufficiently high to induce toxicity problems in sensitive
crop species.
Copper Solubility of Cu2þ is very soil pH dependent and decreases
100-fold for each unit increase in pH. Plant uptake also
decreases.
Boron Sorption of B to Fe and Al oxides is pH dependent and is
highest at pH 6.0–9.0. Bioaavailability of B is higher
between pH 5.5 and 7.5, decreasing below and above this
range mainly due to pH-dependent reactions.
Molybdenum Above soil pH 4.2, MoO42 is dominant. Concentration of
these species increases with increasing soil pH and plant
uptake also increases. Water-soluble Mo increases sixfold as
pH increases from 4.7 to 7.5. Replacement of adsorbed Mo
by OH is responsible for increases in water-soluble Mo as
soil pH increases.
Chlorine Chloride is bound tightly by most soils in mildly acidic to
neutral pH soils and becomes negligible to pH 7.0.
Appreciable amounts can be adsorbed with increasing soil
acidity.
Source: Adriano (1986), Fageria et al. (1997, 2002), Mortvedt (2000), and Tisdale et al. (1985).

5. Factors Affecting Lime Requirements


Lime requirement is defined as the amount of liming material,
as calcium carbonate equivalent, required to change a volume of soil to
a specific state with respect to pH or soluble Al content (Soil Science Society
of America, 1997). However, in economic terms, lime requirement can be
defined as the quantity of liming material required to produce maximum
economic yield of crops cultivated on acid soils. Quantity of lime required to
Ameliorating Soil Acidity 365

Table 9 Influence of soil pH on acquisition of Zn, Cu, Fe, and Mn by upland rice grown
in an Oxisol of Brazil

Zn Cu Fe Mn
Soil pH (mg plant1) (mg plant1) (mg plant1) (mg plant1)
4.6 1090 75 4540 11,160
5.7 300 105 1860 5010
6.2 242 78 1980 4310
6.4 262 64 1630 3610
6.6 163 61 1660 2760
6.8 142 51 1570 2360
R2 0.98** 0.89* 0.97** 0.99**
*,** Significant at 5 and 1% probability levels, respectively.
Source: Fageria (2000).

Table 10 Influence of soil pH on shoot dry weight, grain yield, and panicle number of
upland rice grown on Oxisols

Shoot dry Grain yield Panicle number


pH in H2O weight (g plant1) (g plant1) (plant1)
4.6 15.9 11.0 5.0
5.7 14.9 11.5 4.6
6.2 12.2 11.7 4.5
6.4 10.6 9.5 3.9
6.6 7.8 6.8 3.4
6.8 6.0 5.2 3.3
R2 0.99** 0.91** 0.94**
** Significant at the 1% probability level.
Source: Fageria (2000).

produce maximum economic yields of crops grown on acid soils is deter-


mined by soil properties, liming material quality, management practices,
cropping systems, crop species or genotypes within species, calcium and
magnesium interaction with other nutrients, and economic considerations.

5.1. Quality of liming material


Quality of liming material is very important in correcting soil acidity.
Chemical analysis of the liming material gives its composition. A liming
material containing both calcium and magnesium is desirable for correcting
soil acidity and improving contents of Ca and Mg of the soil. Two impor-
tant characteristics that determine lime material quality are its neutralizing
366 N. K. Fageria and V. C. Baligar

power or reactivity and fineness. The chemical effectiveness of agricultural


limestone is measured by its CaCO3 equivalence. If neutralizing value is
lower than CaCO3, a higher quantity of liming material is required and
vice a versa. The neutralizing power or value of a liming material is defined
as the acid-neutralizing capacity of the material by weight in relation to
CaCO3. A dolomitic limestone has a higher neutralizing capacity than a
calcitic limestone because of the lower atomic weight of Mg. Pure dolomite
has a CaCO3 equivalence of 1.08 (Barber, 1984). Table 11 shows various
liming materials and their CaCO3 equivalent values (neutralizing values).
The degree of fineness indicates the speed with which lime materials will
neutralize soil acidity. Fineness is measured by the proportion of processed
agricultural lime which passes through a sieve with an opening of a particu-
lar size. A 60-mesh sieve, which is the standard for comparisons of lime
finesse and efficiency rating of 100%, is assigned (Caudle, 1991).

5.2. Soil texture


The relative proportions of sand, silt, and clay in a given soil determine soil
texture, a basic physical property of the soil that remains unchanged by
cultural and management practices. The sand group includes all soils whose

Table 11 Liming materials, their composition, and neutralizing power

Commercial Chemical Neutralizing


name formula value (%) Characteristics
Dolomitic CaMg(CO3)2 95–109 Contains 78–120
lime g kg1 of Mg and
180–210 g kg1
of Ca
Calcitic lime CaCO3 100 Contains 284–320
g kg1 of Ca
Dolomite lime MgCO3 100–120 Contains 36–72
g kg1 of Mg
Burned lime CaO 179 Fast reacting and
difficult to handle
Slaked lime Ca(OH)2 136 Fast reacting and
difficult to handle
Basic slag CaSiO3 86 Byproduct of pig-iron
industry, also
contains 1–7% P
Wood ash Variable 30–70 Caustic and water
soluble
Source: Fageria (1989), Bolan et al., (2003), Brady and Weil (2002), and Caudle (1991).
Ameliorating Soil Acidity 367

sand content by weight is 85% or more, the silt group has a silt content of
80% or more, and the clay group includes those whose clay content is 40%
or more (Soil Science Society of America, 1997). Soil texture determines
the buffering capacity of a soil, which refers to the ability of the solid phase
soil materials to resist changes in ion concentrations in the solution phase.
For liming purposes, the resistance of the soil solution to changes in pH is a
main component of soil buffer power. Oxisols contain predominantly iron
and aluminum oxide minerals and kaolinite and have characteristically low
to moderately low cation exchange capacity, but they also have high
buffering capacity. Hence, Oxisols require large amount of liming materials
to raise soil pH to a desired level for maximum crop yields. Maximum
legume grain yield (common bean, soybean) is achieved at a pH of about 6.5
in Brazilian Oxisols (Fageria, 2001b, 2006; Fageria and Stone, 2004). Data
in Fig. 2 show that to raise a pH from 5.3 to 6.5 in Oxisols, a lime rate of
about 7 Mg ha1 is needed.

5.3. Soil fertility


Soil fertility has a significant influence on the quantity of lime required to
correct soil acidity to produce maximum economic crop yields. Quantity of
nutrient present in the soil is defined in term of soil fertility, and, generally,
soil analysis is used as a criterion to make fertilizer recommendations for

7.0

6.5
Soil pH in H2O

6.0
Y = 5.2809 + 0.2181X - 0.0061X 2
R 2 = 0.9656**
5.5

5.0

0 3 6 9 12 15 18
Lime applied (Mg ha-1)

Figure 2 Relationship between lime rate and soil pH in Brazilian Oxisols.


368 N. K. Fageria and V. C. Baligar

Table 12 Influence of liming on base saturation and exchangeable Caþ and Mg2þ at
0–20 cm depth in an Oxisols

Lime rate Base saturation Ca2þ Mg


(Mg ha1) (%) (cmolc kg) (cmolc kg1)
0 40 1.9 1.0
4 44 2.3 1.1
8 51 3.0 1.2
12 53 3.1 1.3
16 56 3.3 1.3
20 66 3.8 1.4
R2 0.80** 0.72** 0.23*
*,** Significant at the 5 and 1% probability level, respectively.
Source: Fageria (2001a,b).

field crops (Fageria and Baligar, 2005b). High fertility soils in terms of
exchangeable Ca2þ, Mg2þ, and Kþ require less lime than do those with
lower soil fertility. When Ca2þ, Mg2þ, and Kþ contents are higher, a lower
lime rate is required, because of higher levels of these basic cations in the
soil, meaning relatively higher base saturation and higher pH than with
lower levels of these cations (Table 12).

5.4. Crop rotation


Crop rotation is defined as a planned sequence of crops growing in a regularly
recurring succession on the same area of land, as contrasted to continuous
culture of one crop or growing a variable sequence of crops (Soil Science
Society of America, 1997). This is a very old system of growing crops and was
practiced during the Han dynasty of China more than 3000 years ago (Karlen
et al., 1994; MacRae and Meheys, 1985). Romans recognized the benefits of
alternating leguminous crops with cereals more than 2000 years ago (Karlen
et al., 1994; Robson et al., 2002). Continuous cultivation of a given crop on
the same land may reduce productivity. This decline in productivity is
associated with a decrease in soil fertility, infestation by weeds, insects, and
diseases, loss of soil by erosion, reduction in soil biological diversity, and
buildup of allelopathy (Fageria, 2002). Ability of legumes to fix atmospheric
nitrogen discovered in the late 19th century was a major reason for crop
rotation (cereals with legumes) that became popular in the early 20th century
(Karlen et al., 1994).
Crop rotation has several benefits. Diversifying crops in rotation can
improve crop yields, a response known as the rotation effect (Anderson,
2003). Phosphorus deficiency or unavailability is a major yield-limiting
Ameliorating Soil Acidity 369

factor in Oxisols, and in the Central Great Plains of the United States, crop
rotations have been reported to improve P-use efficiency (Bowman and
Halvorson, 1997). However, there are some disadvantages in adopting crop
rotation. One such disadvantage is soil acidification, a serious form of land
degradation associated with crop rotation (Coventry et al., 2003). In crop
rotations, the N cycle and C cycle contribute most to the acid input
(Coventry and Slattery, 1991; Helyar et al., 1997). Acidification of soil
can also result from increasing addition of crop residues due to crop rotation
(Poss et al., 1995). It is reported in several studies in southern Australia that
in the wheat-based crop rotation, the rates of acidification can be 0.16–3.6
Hþ ha1 year1 for pastures (Loss et al., 1993; Ridley et al., 1990) and
1.0–7.5 Hþ ha1 year1 for cereal–legume rotations (Coventry and Slattery,
1991; Dolling, 1995; Moody and Aitken, 1997). The rate of acidification in
intensive cropping systems is also associated with removal of large amounts of
calcium and magnesium in the grains. In addition, acidification in the tropical
rainforest such as the cerrado region of Brazil is also associated with the
replacement of perennial vegetation with shallow-rooted annual pastures
and crops. With this change, more water has drained through the soil profile,
which leads to removal of bases (Coventry et al., 2003). The rates of acidifi-
cation are generally more pronounced in higher rainfall areas and soil pH
in the range of 4.8–5.5 (CaCl2) is likely to be more susceptible to rapid
acidification (Coventry et al., 2003; Haynes, 1983).

5.5. Use of organic manures


Organic manures are products from the processing of animal or vegetable
substances that contain reasonable amount of plant nutrients to be of value
as fertilizers. Brosius et al. (1998) reported that plant- and animal-based
organic byproducts may substitute for commercial fertilizers and enhance
chemical and biological attributes of soil quality in agricultural production
systems. Use of organic manures can also affect lime requirement of a crop.
Organic matter increases the soil’s ability to hold and make available
essential plant nutrients and to resist the natural tendency of soil to become
acidic (Cole et al., 1987). Furthermore, addition of organic manures to acid
soils has been shown to increase soil pH, decrease Al saturation, and thereby
improve conditions for plant growth (Alter and Mitchell, 1992; Reis and
Rodella, 2002; Wong and Swift, 2003). Miyazawa et al. (1993) reported that
crop residues of wheat and corn and 20 plant species utilized as green
manure increased soil pH and decreased Al content of the soil. Several
mechanisms have been proposed for reducing acidity by organic manures.
These mechanisms include specific adsorption of organic anions on hydrous
Fe and Al surfaces and the corresponding release of hydroxyl ions which
increase soil pH (Hue, 1992; Wong and Swift, 2003). Adsorption of Al by
organic matter sites and the subsequent isolation of the inorganic phase to
370 N. K. Fageria and V. C. Baligar

maintain the equilibrium Al activity in soil solution have been proposed to


increase soil pH (Wong and Swift, 2003; Wong et al., 1998). Chelating
agents released by decomposing organic matter may detoxify Al ions.
Plant roots decay in the soil and form into soil organic matter. Active roots
also release organic acids such as citrate, malate, and tartrate. These organic
acids react strongly with Al and convert it into less toxic organically bound
forms (Yang et al., 2000). The organic Al affinities or stability constants are in
the order of citrate > tartrate > malate (Hue et al., 1986). The decrease in
Al-activity by addition of organic matter has been reported by Kochain
(1995). The functional groups involved in metal complexation by organic
matter are COOH and OH (Wong and Swift, 2003). Surface application or
surface incorporation of organic matter also decreased phytotoxic subsoil
Al3þ activities because dissolved organic matter (DOM) that leached into
the subsoil formed nontoxic Al–DOM complexes (Hue, 1992; Hue and
Licudine, 1999; Liu and Hue, 1996; Willert and Stehouwer, 2003). The
combined application of CaCO3 and organic matter in lime-stabilized bio-
solids decreased subsoil acidity and increased subsoil Ca saturation, compared
with CaCO3 alone (Brown et al., 1997; Tan et al., 1985; Tester, 1990; Willert
and Stehouwer, 2003). This effect was attributed to increases in Ca mobility
caused by Ca–DOM complexes (Willert and Stehouwer, 2003).
Additional benefits of organic matter addition to acid soils are improving
nutrient cycling and availability to plants through direct additions as well as
through modification in soils’ physical and biological properties. A comple-
mentary use of organic manures and chemical fertilizers has proven to be the
best soil fertility management strategy in the tropics (Fageria and Baligar,
2005a; Makinde and Agboola, 2002). Enhanced soil organic matter increases
soil aggregation and water-holding capacity, provides source of nutrients, and
reduces P fixation, toxicities of Al and Mn, and leaching of nutrients (Baligar
and Fageria, 1999). Build-up of organic matter through additions of crop and
animal residues increases the population and species diversity of microorgan-
isms and their associated enzyme activities and respiration rates (Kirchner
et al., 1993; Weil et al., 1993). The use of organic compost may result in a soil
that has greater capacity to resist the spread of plant pathogenic organisms.
The improvement in overall soil quality may produce more vigorous
growing and high yielding crops (Brosius et al., 1998).

5.6. Conservation tillage


The mechanical manipulation of the soil profile for crop production is
known as tillage. Conservation tillage is minimum manipulation of the
soil profile for crop production and 30% or more soil surface is covered
with crop residues. In conservation tillage, weeds are controlled by herbi-
cides and this practice helps in conservation of water and nutrients, and
reduces soil erosion and labor cost. Conservation tillage reduces oxidation
Ameliorating Soil Acidity 371

of organic matter or conserves soil organic matter content and consequently


decreases the adverse effects of soil acidity. Additional research is needed to
assess the long-term conservation tillage effects on soil acidity.

5.7. Crop species and genotypes within species


Crop species and genotypes within species differ significantly in relation to
their tolerance to soil acidity (Baligar and Fageria, 1997; Devine, 1976;
Fageria and Baligar, 2003b; Fageria et al., 1989, 2004; Foy, 1984; Garvin
and Carver, 2003; Reid, 1976; Sanchez and Salinas, 1981; Yang et al.,
2000). Hence, lime requirements also vary from species to species and
among cultivars within species. Many of the plant species tolerant to acidity
have their center of origin in acid soil regions, suggesting that adaptation to
soil constraints is part of the evolution processes (Foy, 1984; Sanchez and
Salinas, 1981). A typical example of this evolution is the acid soil tolerance
of Brazilian upland rice cultivars. In Brazilian Oxisols, upland rice grows
very well without liming, when other essential nutrients are supplied in
adequate amount and water is not a limiting factor (Fageria, 2000, 2001a).
Experimental results obtained on Brazilian Oxisols with upland rice are
good examples of crop acidity tolerance evaluation. Fageria et al. (2004)
reported that grain yield and yield components of 20 upland rice genotypes
were significantly decreased at low soil acidity (limed to pH 6.4) as compared
with high soil acidity (without lime, pH 4.5), demonstrating the tolerance of
upland rice genotypes to soil acidity. In Table 13, data are presented showing
grain yield and panicle number of six upland rice genotypes at two acidity
levels. These authors also reported that grain yield gave significant negative
correlations with soil pH, Ca saturation, and base saturation. Further, grain

Table 13 Grain yield and panicle number of six upland rice genotypes at two soil
acidity levels in Brazilian Oxisols

Grain yield (g pot1) Panicle number (pot1)


High acidity Low acidity High acidity Low acidity
Genotype (pH 4.5) (pH 6.4) (pH 4.5) (pH 6.4)
CRO97505 74.3 52.0 38.0 28.3
CNAs8983 55.2 42.9 29.0 25.7
Primavera 53.0 47.2 25.0 21.7
Canastra 51.6 38.9 32.0 26.3
Bonança 48.8 36.5 26.3 20.7
Carisma 50.8 17.5 43.3 17.7
Average 66.7 47.0 38.7 28.1
Source: Compiled from Fageria et al. (2004).
372 N. K. Fageria and V. C. Baligar

yield had significant positive correlations with soil Al and H þ Al, confirming
that upland rice genotypes were tolerant to soil acidity. Fageria (1989)
reported stimulation of growth of Brazilian rice cultivars at 10 mg Al3þ L1
in nutrient solution compared with a control (no Al) treatment. Okada and
Fischer (2001) suggested that the mechanism for the genotype difference of
upland rice for tolerance to soil acidity is due to the relationship between
regulation of cell elongation and legend-bound Ca at the root apoplast.
A substantial number of plant species of economic importance are
generally regarded as tolerant to acid soil conditions of the tropics
(Sanchez and Salinas, 1981). In addition, there are cultivars within crop
species that are tolerant to soil acidity (Fageria et al., 2004; Garvin and
Carver, 2003; Yang et al., 2004). Yang et al. (2005) reported significant
differences among genotypes of rye, triticale, wheat, and buckwheat to Al
toxicity. These crop species or cultivars within these species can be planted
on tropical acid soils in combination with reduced rates of lime input.
Combination of legume–grass pasture and agroforestry system of man-
agement are the other important soil acidity management components
useful in tropical ecosystems. For example, Pueraria phaseoloides is used as
understory for rubber, Gmelina arborea or Dalbergia nigra, plantations in
Brazil, presumably supplying nitrogen to the tree crops (Sanchez and
Salinas, 1981) A detailed discussion of combination of legume–grass pasture
and agroforestry in tropical America is given by Sanchez and Salinas (1981).
These authors reported that when an acid-tolerant legume or legume–grass
pasture is grown under young tree crops, the soil is better protected, soil
erosion is significantly reduced, and nutrient cycling is enhanced. Some
important annual food crops, cover or green manure crops pasture species,
and plantation crops tolerant to tropical acid soils are listed in Table 14. Acid
soil tolerant crops are useful to establish low input management systems.

5.8. Interaction of lime with other nutrients


Recognition of the importance of nutrient balance in crop production is an
indirect reflection of the contribution of interactions to yield. The highest
yields are obtained where nutrients and other growth factors are in a
favorable state of balance. As one moves away from this state of balance,
nutrient antagonisms are reflected in reduced yields (Fageria et al., 1997).
Nutrient interactions can occur at the root surface or within the plant and
can be classified into two major categories. In the first category are interac-
tions which occur between ions because the ions are able to form a chemical
bond. Interactions in this case are due to formation of precipitates or
complexes either in soil or in the plant. For example, this type of interaction
occurs where the liming of acid soils decreases the concentration of almost
all micronutrients in soil solution except molybdenum. Such reduction in
ion concentration in soil solution decreases the uptake. Increasing soil pH
Ameliorating Soil Acidity 373

Table 14 Some important crop species, pasture species, and plantation crops tolerant
to soil acidity in the tropics

Annual crop species Pasture species Plantation crops


Rice Brachiaria Banana
Peanut Andropogon Oil palm
Cowpea Panicum Rubber
Potato Digitaria Coconut
Cassava Napiergrass Cashewnut
Pigeon pea Jaraguagrass Coffee
Millet Centrosema Guarana
Kudzu Stylosanthes Tea
Mucuna Leucaena
Crotolaria Brazilian nut
Eucalyptus
Papaya
Source: Sanchez and Salinas (1981), Caudle (1991), Fageria et al. (1997), and Brady and Weil (2002).

due to lime will have a more marked effect on Zn than Cu uptake, mainly
because Cu is more complexed and protected from precipitation by soluble
organic matter (Robson and Pitman, 1983). The second form of interaction
is between ions whose chemical properties are sufficiently similar that they
compete for site of adsorption (soil components, cell walls), absorption,
transport, and function on plant root surfaces or within plant tissues. Such
interactions are more common between nutrients of similar size, charge,
geometry of coordination, and electronic configuration (Robson and
Pitman, 1983).
Generally, three types of interactions occur among essential nutrients in
plants, and these interactions are known as synergistic, antagonistic, and
neutral. If upon addition of two nutrients, an increase in plant growth or
yield that is greater than that achieved by adding only one occurs, the
interaction is synergistic or positive. Similarly, if adding the two nutrients
together produced less plant growth or yield as compared to individual
ones, the interaction is negative or antagonistic. When there is no significant
change in plant growth or yield with the addition of two nutrients, there is
no interaction (Sumner and Farina, 1986). Data in Table 15 show that lime
and P fertilization have positive as well as negative interactions depending
on crop species. In the case of upland rice and wheat, shoot dry weight
yields were higher at zero level of lime compared with 2 and 4 g lime kg1
of soil. Hence, interaction between lime and P in this case was negative.
Whereas, shoot dry matter yield of common bean and corn increased with
increasing lime as well as P rates, indicating that interaction between lime
and P in this case was synergistic. An increasing response to applied P with
374 N. K. Fageria and V. C. Baligar

Table 15 Dry matter yield of shoots of upland rice, wheat, common bean, and corn as
affected by lime and phosphorus applied to a Brazilian Oxisols

Common
Lime rate P rate Rice Wheat bean Corn
(g kg1) (mg kg1) (g pot1) (g pot1) (g pot1) (g pot1)
0 0 0.72 0.17 1.25 1.10
0 50 15.08 5.83 8.30 4.93
0 175 18.63 8.40 10.60 8.73
2 0 0.73 0.20 1.30 1.43
2 50 15.23 5.20 9.00 7.13
2 175 13.23 6.50 10.60 11.47
4 0 0.33 0.20 1.70 1.10
4 50 10.20 4.90 10.50 6.60
4 175 13.20 5.97 12.00 9.93
Source: Compiled from Fageria et al. (1995).

increasing rates of added lime has been attributed to either an improved rate
of supply of P by the soil or an improved ability of the plant to absorb P
when Al toxicity has been eliminated (Friesen et al., 1980b). Liming also
improves the microbiological activities of acid soils, which, in turn, can
increase dinitrogen fixation by legumes and liberate N and other nutrients
from incorporated organic materials (Fageria et al., 1995).
Nutrient interaction is also evaluated by studying the influence of
increasing nutrient concentrations on the uptake of other nutrients and
corresponding plant growth. When plant uptake of a given nutrient
decreased and corresponding plant growth also decreased, interaction is
negative. However, if plant growth and nutrient concentration increased
with increasing nutrient supply, the interaction is positive.
Liming of acid soils is mainly to supply Ca and Mg to plants and
neutralize phytotoxic levels of soil Al3þ. Hence, interaction of Ca2þ,
Mg2þ, and Al3þ with other elements is an important aspect for nutrient
interaction discussion. Although absolute Ca requirements for plant growth
and metabolism are low, it has great significance in providing balance for
levels of other plant nutrients, maintaining membrane integrity, and reduc-
ing potential for toxicity of other elements (Wilkinson et al., 2000). Potas-
sium, Al3þ, Mg2þ, Mn2þ, and Hþ and heavy metals can reduce Ca uptake
by binding at the exterior surface of plasma membrane which increases the
Ca requirement (Marschner, 1995). Fertilization with NO3 generally
enhances Ca and Mg concentrations in plants driven by the need for
cation–anion balance (Wilkinson et al., 2000). Lime and P interactions are
mainly associated with Al toxicity, which limits root growth and prolifera-
tion, and nutrient uptake. Aluminum absorbed by roots can also precipitate
Ameliorating Soil Acidity 375

root-absorbed P and hinder its subsequent translocation to plant tops (Foy,


1983). Mora et al. (2005) reported that the most important Al detoxifying
mechanisms in ryegrass were apparently physiological Al-PO4 precipitation
inside the root and chemical AlSOþ 4 complex formation in the nutrient
solution.
The results of solution culture experiments have shown that negative
effects of Al on root growth can be reduced by increasing Ca2þ concentra-
tion in the growth medium (Nichol et al., 1993; Zysset et al., 1996). Baligar
et al. (1992) also reported that Ca ameliorates Al toxicity in wheat plants.
Huang and Grunes (1992) reported that adequate supply of Ca and Mg might
also ameliorate Al toxicity in wheat plants. Aluminum toxicity continues to
be associated with P nutrition of plants (Foy, 1984). Santana and Braga
(1977) found that P concentrations in rice tops decreased with increasing
Al saturation of the soil. Helyar (1978) concluded that Al toxicity effects
were largely associated with Al interference with P metabolism and with Al
binding to pectins in root cell wall, stopping root elongation. Fageria and
Baligar (1999) studied the interactions between calcium and P, Mg, Zn, Cu,
Mn, Fe, and B in common bean and found that uptake of P was quadratically
increased, whereas uptake of Mg, Zn, Cu, Mn, Fe, and B significantly
decreased as soil Ca increased from 4.9 to 12.5 cmolc kg1 of soil.
Calcium  Al3þ interactions are important in acid soils (Foy, 1984).
Below pH 5.5, Ca2þ  Al3þ antagonism is probably the most important
factor affecting Ca uptake by plants. Lance and Pearson (1969) reported that
reduced Ca uptake was the first externally observed symptom of Al damage
on cotton seedlings. Franco and Munns (1982) reported that increasing
Ca2þ concentrations from 8 to 80 mg L1 decreased Al toxicity in bean
plants. Simpson et al. (1977) attributed poor root growth of alfalfa to Al3þ 
Ca2þ interaction. Aluminum  Feþ3 interactions are frequently reported in
the literature (Foy, 1984). Alam (1981) reported Al-induced Fe deficiency
in oat and postulated that Al interfered with the reduction of Fe3þ to Fe2þ
within the plant, a process essential for normal Fe metabolism and utiliza-
tion. Clark et al. (1981) found that Fe-deficiency chlorosis was a common
symptom of Al toxicity in sorghum.

6. Criteria to Determine Liming


Material Quantity
Use of adequate lime rate to correct soil acidity and production of
maximum yield of a crop species is an important consideration for econom-
ical and ecological reasons. Quantity of liming material required is deter-
mined on the basis of soil pH, base saturation, and aluminum saturation
adjustment at appropriate levels. Appropriate levels of these acidity indices
376 N. K. Fageria and V. C. Baligar

vary with soil type, soil fertility, plant species, and crop genotypes within
species (Fageria and Baligar, 2003a). In addition, crop response
curves related to lime rate and yield is another criterion that can be used
to define lime requirement for any given crop species. Crop response curves
to lime levels should be determined for each crop species under different
agroecological regions to make liming recommendations effective and
economical.

6.1. Soil pH
Soil pH or hydrogen ion activity is the most common acidity index used
in soil testing program for assessing lime requirements of crops grown on acid
soils. Weaver et al. (2004) reported that soil pH buffering capacity, since it
varies spatially within crop production fields, may be used to define sampling
zones to assess lime requirement, or for modeling changes in soil pH when
acid-forming fertilizers or manures are added to a field. The pH is defined as
the negative logarithm of the hydrogen ion concentration or activity. It is
determined by means of a glass, quinhydrone, or other suitable electrode or
indicator at a specified soil to solution ratio in a specified solution, usually
distilled water, 0.01 M CaCl2, or 1 M KCl (Soil Science Society of America,
1997). In soil testing laboratories of Brazil, a soil to solution ratio of 1:2.5 is
commonly used to determine soil pH (EMBRAPA, 1997).
The pH measured in the soil solution represents the active acidity of the
soil. Hydrogen and aluminum ions adsorbed by the soil, as well as other soil
constituents that generate hydrogen ions, constitute the reserve acidity. The
active acidity is neutralized by the addition of lime where more hydrogen ions
from the reserve pool go into solution. This results in the resistance of soil to
changes in the pH of the soil solution. This is termed ‘‘buffering capacity.’’
The pH of acid soils is lower than 7.0, that of neutral soils is equal to 7.0,
and that of alkaline soils is above 7.0. At neutrality, the concentrations of
OH ions and Hþ ions are equal, since these ions are derived in equal
quantities from the ionization of water. The pH of most agricultural soils is
in the range of 4.0–9.0 (Fageria et al., 1990, 1997). Soil acidity is classified
into several groups based on soil pH. Slightly acid soils have a pH range of
6.0–7.0, moderately acid soils pH range from 5.5 to 6.0, strongly acid soils
pH from 4.5 to 5.0, and extremely acid soils have a pH range below 4.5
(Fageria and Gheyi, 1999). These acidity classifications are arbitrary, and
care should be taken when defining adequate pH for crop yields, particu-
larly the extractant used could make a difference. Soil pH in water is higher
than that in CaCl2 solution.
Soil pH measurements not only indicate the acidity level of a soil but also
be used as an initial basis for the prediction of the chemical behavior of soils,
particularly in relation to nutrient availability and the presence of toxic
elements. Most plant-essential nutrients in soil reach maximal or near-maximal
Ameliorating Soil Acidity 377

availability in the pH range 6.0–7.0, and decrease both above and below this
range (McLean, 1973). Optimal pH values for annual crops cultivated on
Brazilian Oxisols are given in Table 16. A curve showing the relationship
between lime rate and pH change of a Brazilian Oxisol is presented in Fig. 2.
Oxisols with application of lime 0–18 Mg ha1 increase pH significantly and
quadratically. Figure 3 shows the relationship between pH of an Oxisol and
shoots dry weight and grain yield of common bean. Maximum shoot dry
weight and grain yields were obtained at a pH of 6.4. Improvement in grain
yield was associated with increasing numbers of pods, grains per pod, and
weight of 100 seeds when pH was increased from 5.3 to 6.4 (Fig. 4).

6.2. Base saturation


Base saturation is another important chemical property of soils used as a
criterion for liming recommendations. Base saturation is defined as the
proportion of the CEC occupied by exchangeable bases. It is calculated as
follows (Fageria et al., 2007):

Table 16 Optimal soil pH for important crop species grown on Brazilian Oxisols

Type of Plant part Soil pH in


Crop species experiment measured H2O (1:2.5) Reference
Wheat Greenhouse Shoot dry 6.0 Fageria et al.
weight (1997)
Common Field Grain yield 6.6 Fageria and Stone
bean (2004)
Upland rice Field Grain yield 5.6 Fageria and
Baligar (2001)
Common Field Grain yield 6.2 Fageria and
bean Baligar (2001)
Corn Field Grain yield 6.4 Fageria (2001b)
Soybean Field Grain yield 6.8 Fageria (2001b)
Upland rice Field Grain yield 5.5 Fageria and Stone
(1999)
Upland rice Greenhouse Grain yield 5.4 Fageria (2000)
Upland rice Greenhouse Grain yield 5.9 Fageria et al.
(1990)
Corn Field Grain yield 6.2 Gonzales-Erico
et al. (1979)
Soybean Field Grain yield 6.4 Raij and Quaggio
(1997)
378 N. K. Fageria and V. C. Baligar

Shoot Grain

Shoot dry weight and grain yield (kg ha−1)


a a
3000

2000 a a

1000

0
5.3 6.4 6.8 5.3 6.4 6.8
Soil pH in H2O

Figure 3 Relationship between soil pH and shoot dry weight and grain yield of
common bean grown on a Brazilian Oxisol (Fageria and Santos, 2005).

a a
a a 4 c b a
300
Number of grains (pod−1)

Weight of 100 grains (g)

b b
Number of pods (m−2)

3 30
200
2 20

100
1 10

0
5.3 6.4 6.8 5.3 6.4 6.8 5.3 6.4 6.8
Soil pH in H2O

Figure 4 Relationship between soil pH and yield components of common bean grown
on Brazilian Oxisols (Fageria and Santos, 2005).

P
ðCa; Mg; K; NaÞ
Base saturationð%Þ ¼  100
CEC
where CEC is the sum of Ca, Mg, K, Na, H, and Al expressed in cmolc
kg1.
Ameliorating Soil Acidity 379

In Brazil, Naþ is generally not determined because of a very low level of


this element in Brazilian Oxisols (Raij, 1991). Hence, Na is not considered
in the calculation of CEC or base saturation.
For crop production, base saturation levels in soil may be grouped into
very low (lower than 25%), low (25–50%), medium (50–75%), and high
(>75%) (Fageria and Gheyi, 1999). Very low and low base saturation means
a predominance of adsorbed hydrogen and aluminum on the exchange
complex. Deficiencies of calcium, magnesium, and potassium are likely to
occur in soils with low CEC and very low to low percent base saturation.
Quantity of lime required by the base saturation method is calculated by
using the following formula (Fageria et al., 1990):
 
1 CECðB2  B1 Þ
Lime rateðMg ha Þ ¼  df
TRNP

where CEC: cation exchange capacity or total exchangeable cations (Ca2þ,


Mg2þ, Kþ, Hþ þ Al3þ) in cmolc kg1, B2: desired optimum base satura-
tion, B1: existing base saturation, TRNP: total relative neutralizing power
of liming material, and df: depth factor, 1 for 20 cm depth and 1.5 for 30 cm
depth.
For Brazilian Oxisols, the desired optimum base saturation for most of
the cereals is in the range of 50–60%, and for legumes it is in the range of
60–70% (Fageria et al., 1990). However, there may be exceptions, like
upland rice, which is very tolerant to soil acidity and can produce good
yield at base saturation lower than 50%. Specific optimal base saturation
values for important annual crops grown on Brazilian Oxisols are given in
Table 17. Nature of the soil alters the optimum base saturation required by
any given crop species. A relationship between lime rate and base saturation
in a Brazilian Oxisol is given in Fig. 5. Bean yield was having significant
quadratic response in relation to base saturation (Fig. 6). Maximum yield
was obtained with base saturation of 73% at 0–10 cm soil depth, with base
saturation of 62% at 10–20 cm soil depth and at 67% base saturation when
averaged across two soil depths. Hence, at topsoil layer, higher base saturation
was required compared with that at lower soil layer (Fageria, 2008).

6.3. Exchangeable aluminum, calcium, and magnesium levels


Aluminum has long been recognized as a toxic element for plant growth
(Cronan and Grigal, 1995; Foy, 1984). In soil–plant systems, plant-available
Al is determined by soil extraction procedure to predict the risk of Al
toxicity and the need for liming (Thomas and Hargrove, 1984). In addition,
Ca and Mg contents of the Oxisols are important in determining growth of
plants. Hence, exchangeable aluminum, calcium, and magnesium contents
380 N. K. Fageria and V. C. Baligar

Table 17 Optimal base saturation for important annual crops grown on Brazilian
Oxisols

Type of Plant part Base


Crop species experiment measured saturation (%) Reference
Common Field Grain yield 60 Fageria and Santos
bean (2005)
Common Field Grain yield 69 Fageria and Stone
bean (2004)
Upland rice Field Grain yield 40 Fageria and
Baligar (2001)
Common Field Grain yield 70 Lopes et al. (1991)
bean
Corn Field Grain yield 59 Fageria (2001a)
Soybean Field Grain yield 63 Fageria (2001a)
Upland rice Field Grain yield 50 Lopes et al. (1991)
Upland rice Field Grain yield 30 Sousa et al. (1996)
Common Field Grain yield 71 Fageria and Stone
bean (2004)
Corn Field Grain yield 60 Raij et al. (1985)
Wheat Field Grain yield 60 Lopes et al. (1991)
Soybean Field Grain yield 60 Raij et al. (1985)
Cotton Field Grain yield 60 Raij et al. (1985)
Sugarcane Field Cane yield 50 Raij et al. (1985)
Soybean Field Grain yield 61 Gallo et al. (1986)

of the soil are taken into account to determine the rate of lime required for
a crop grown on an Oxisol. The equation used for lime rate determination
is (Fageria et al., 1990; Raij, 1991):

Lime rateðMg ha1 Þ ¼ ð2  Al3þ Þ þ ½2  ðCa2þ þ Mgþ Þ

where values of Al3þ, Ca2þ, and Mg2þ are expressed in cmolc kg1.
If values of Ca2þ and Mg2þ cations are more than 2 cmolc kg1, only Al
multiplied by factor 2 is considered. This criterion was originally suggested
by Kamprath (1970) for tropical soils and is still largely used for liming
recommendation for Brazilian acid soils (Paula et al., 1987; Raij, 1991; Raij
and Quaggio, 1997). Alvarez and Ribeiro (1999) recommended that the
factor used to multiply Al should be varied according to soil texture. These
authors suggested that in sandy soil with clay content of 0–15%, the factor
0–1 should be used; for medium-texture soils with clay content of 15–35%,
a factor 1–2 should be used; for clayey soil with clay content of 35–60%,
Ameliorating Soil Acidity 381

100

80
Base saturation (%)

60

40

Y = 20.0658 + 8.1171X - 0.2493X 2


20 R 2 = 0.9635**

0 3 6 9 12 18
Lime rate (Mg ha-1)

Figure 5 Relationship between lime rate and base saturation of Brazilian Oxisols.

a factor 2–3 should be used; and for heavy clayey soil with clay content of
60–100%, a factor of 3–4 should be used.

6.4. Aluminum saturation


Crops grown in soils with acceptable levels of basic cations do not show Al
toxicity symptoms even when the levels of KCl-extractable Al are consid-
ered high (Kariuki et al., 2007). Hence, the mere presence of Al in the soil is
not an indicator of Al toxicity ( Johnson et al., 1997). A more reliable
measure of the potential for Al toxicity is Al saturation (Kariuki et al., 2007).
It has been widely reported in the literature that differences in Al tolerance
are found among plant species and cultivars within species (Fageria and
Baligar, 2003a; Foy, 1992; Kochain, 1995; Okada and Fischer, 2001; Yang
et al., 2004). It is evident, therefore, that crop tolerance to Al should be taken
into account in estimating the amounts of lime needed to correct Al toxicity.
Cochrane et al. (1980) suggested that crop aluminum tolerance should be
considered along with levels of exchangeable Ca and Mg, in determination of
lime requirement. Aluminum saturation, or the proportion of aluminum
among the cations, is calculated by using the following formula:
 
Al
Al saturationð%Þ ¼  100
ECEC
382 N. K. Fageria and V. C. Baligar

Average of two
soil depths
3000

2000

1000 Y = 1248.41 + 56.6337X - 0.4203X 2


R 2 = 0.7736**

0
10–20 cm
Grain yield (kg ha-1)

3000

2000

Y = 1347.17 + 58.6230X - 0.4696X2


1000
R2 = 0.7391**

0
0–10 cm

3000

2000

1000 Y = 1173.82 + 54.2814X - 0.3709X2


R2 = 0.7956**

0 20 40 60 80
Base saturation (%)

Figure 6 Influence of base saturation on grain yield of dry bean (Fageria, 2008).

where ECEC is in cmolc kg1, which is the sum of exchangeable Al3þ,


Ca2þ, Mg2þ, and Kþ in cmolc kg1.
After determining Al saturation, the following formula is used to calcu-
late lime rate (Cochrane et al., 1980):
 3þ 
Al  TASðAl3þ þ Ca2þ þ Mg2þ Þ
1
Lime rateðMg ha Þ ¼ 1:8
100
Ameliorating Soil Acidity 383

where 1 M KCL extracts Al, Ca, and Mg and concentrations are expressed
in cmolc kg1 and TAS is target Al saturation, which varied from crop
species to species.
Critical Al saturation values for important plant species are given in
Table 18. These values can be used as a reference guide to calculate
the lime rate for different crop species. This approach is very useful where
lime is difficult to obtain and rather costly and Al-tolerant cultivars are
available.

6.5. Crop responses


Methods discussed earlier for lime rate determination provide reference
guides for determination of lime rates for correcting soil acidity-related
constraints for crops. The best criterion, however, for determining lime
rate is actual testing of crop responses to applied lime rates. Crop responses
to liming are determined by soil, climate, plant species, and cultivar within
Table 18 Critical soil aluminum saturation for important field crops at 90–95% of
maximum yield

Critical Al
Crop Type of soil saturation (%)
Cassava Oxisols/Ultisols 80
Upland rice Oxisols/Ultisols 70
Cowpea Oxisols/Ultisols 55
Cowpea Oxisols 42
Peanut Oxisols/Ultisols 65
Peanut Oxisols 54
Soybean Oxisols 19
Soybean Oxisols 27
Soybean Oxisols/Ultisols 15
Soybean Not given <20
Corn Oxisols 19
Corn Oxisols/Ultisols 29
Corn Oxisols/Ultisols 25
Corn Oxisols 28
Mungbean Oxisols/Ultisols 15
Mungbean Oxisols/Ultisols 5
Coffee Oxisols/Ultisols 60
Sorghum Oxisols/Ultisols 20
Common bean Oxisols/Ultisols 10
Common bean Oxisols/Ultisols 8–10
Common bean Oxisols/Ultisols 23
Cotton Not given <10
Source: Compiled from various sources by Fageria et al. (1997).
384 N. K. Fageria and V. C. Baligar

species and socioeconomic condition of growers and their interactions.


Hence, crop responses to liming should be determined in different agro-
ecological regions.
Over the years in different region of Brazil, liming experiments have been
conducted on Oxisols with various annual crops. Lime rate to achieve 90% of
the maximum grain yield potential is considered an economic lime rate. Raij
and Quaggio (1997) reported that in the State of Sao Paulo, depending on
grain yield, economic optimum lime rate required for soybean is around 3.6–
5.5 Mg ha1. These authors also reported economic optimum lime rate of
6 Mg ha1 for cotton and 3.8 Mg ha1 for sugarcane. Fageria (2001a)
reported that for the State of Goiás, economic optimum lime rate was
5 Mg ha1 for common bean, 8 Mg ha1 for corn, and 9 Mg ha1 for
soybean. Caires et al. (2005) reported that in Brazilian Oxisols (Hapludox),
the maximum economic yield of wheat was obtained with application of
4 Mg ha1of lime. Figure 7 shows the relationship between lime rate and
grain yield of soybean grown on Brazilian Oxisols in the State of Tocantins.
Maximum grain yield was achieved at 12 Mg ha1 of lime. However, 90% of
the maximum yield (economical rate) was obtained with the application of
5.3 Mg ha1.

7. Methods, Frequency, Depth, and Timing


of Lime Application
Methods, frequency, depth, and timing of liming are important
practices in improving liming efficiency and crop yields on acid soils. Liming
material is applied in large quantity to bring about the desired chemical

3500
Grain yield (kg ha-1)

3000

2000
Y = 2205.7970 + 187.5253X - 7.7712X 2
R2 = 0.8688**
1000

0 3 6 9 12 15 18
Lime rate (Mg ha-1)

Figure 7 Relationship between lime rate and soybean grain yield grown on Brazilian
Oxisols.
Ameliorating Soil Acidity 385

changes in acid soils. Hence, the best method is applying it as broadcast as


uniformly as possible and mixing thoroughly through the soil profile. Lime
broadcasting machines are available for uniform application of liming
materials.
Liming frequency is mainly determined by intensity of cropping, crop
species planted, and levels of Ca2þ, Mg2þ, Al, and pH in a soil after each
harvest. The effect of lime is long lasting but not permanent. After several
crops, Ca2þ and Mg2þ move downward and beyond the reach of roots.
These elements are taken up by crops and, to some extent, are lost through
soil erosion. Acid-forming fertilizers and decomposing organic matter lower
the soil pH and release more aluminum to the soil solution and cation
exchange sites on soil particles. When values of exchangeable Ca2þ, Mg2þ,
and pH fall below optimum levels for a given crop species, liming should be
repeated. This means that soil samples should be taken periodically to
determine changes in soil chemical properties and to decide liming fre-
quency. Effects of lime do last longer than those of most other amendments;
however, it is rarely necessary to lime more frequently than every 3 years
(Caudle, 1991). Lime study on Oxisols of the cerrado region of Brazil
showed that after 4 years, with eight crops grown (two crops each year) in
rotation (rice, common bean, corn, and soybean), the soil maintained levels
of soil pH and Ca2þ and Mg2þ contents to maintain yields of these crops
at an optimal level (Table 19) (Fageria, 2001a). The residual effect of coarse
Table 19 Values of soil pH and Ca2þ and Mg2þ contents after harvest of eight crops
(upland rice, common bean, corn, and soybean) grown in rotation for 4 years on a
Brazilian Oxisol at two soil depths

Ca2þ
Lime rate (Mg ha1) pH in H2O (cmolc kg1) Mg2þ (cmolc kg1)
0–20 cm soil depth
0 5.6 1.9 1.0
4 6.0 2.3 1.1
8 6.2 3.0 1.2
12 6.4 3.1 1.2
16 6.5 3.3 1.3
20 6.8 3.8 1.4
20–40 cm soil depth
0 5.5 1.7 0.9
4 5.9 1.9 1.0
8 6.1 2.3 1.1
12 6.2 2.4 1.1
16 6.3 2.6 1.2
20 6.7 3.3 1.3
Source: Fageria (2001a).
386 N. K. Fageria and V. C. Baligar

lime material is greater than with finer lime material because large lime
particles react slowly with soil acidity and tend to remain in the soil longer.
Liming material should be mixed thoroughly in the soil as deeply as
possible to improve crop-rooting systems in acid soils. However, with
currently available machinery, it is generally mixed to a depth of 20–
30 cm. A depth greater than 30 cm required more powers and became
more costly in terms of labor and energy. Timing of lime application is
important in achieving desirable results. Lime should be applied as far in
advance as possible of planting of crop to allow it to react with soil colloids
and to bring about significant changes in soil chemical properties. Soil
moisture and temperature are determining factors for lime to react with
soil colloids. In Brazil, most of the published work shows that liming should
be done 3 months in advance of sowing a crop. However, other studies
carried at the National Rice and Bean Research Center have shown that in
Brazilian Oxisols significant chemical changes can take place 4–6 weeks
after applying liming materials if soil has sufficient moisture (Fageria, 1984,
2001a). Hence, to obtain desirable results, it is not necessary to wait for a
longer period of time after applying lime.

8. Conclusions
The increasing acidification of agricultural soils, as a result of natural
processes, industrial pollution, and agricultural practices, is adversely affect-
ing crop production worldwide on significant levels. Highly weathered acid
soils cover about two-thirds of tropical America and are of great ecological
and economical importance. Soil order Oxisols make up the largest propor-
tion of acid soils of this region. Soils of this order are acidic with low
fertility. Most soil acidity problems can be managed successfully, without
economic and environmental disruption. Among management factors for
acid soils, perhaps most important is supplying sufficient lime initially to
correct the majority of growth-limiting factors. Hence, adequate liming and
fertilization can reduce the adverse effects of soil acidity and improve crop
yields. The central part of Brazil, locally known as the cerrado region,
contains predominantly Oxisosls. Modern agricultural practices, including
liming and fertilization, could help to achieve economic yields on these
acidic and infertile soils. On these soils, the amount of lime needed for
maximum yield varies with crop species and cultivar within species. Exper-
imental findings have shown that maximum economic yield for most annual
crops on Oxisols can be obtained with application of 4–6 Mg ha1 of lime.
Initially, liming acid soils not only improves crop yields but also helps to
maintain environmental quality and consequently improves animal and
human health. Overliming, however, can significantly reduce the
Ameliorating Soil Acidity 387

Table 20 Common and scientific names of plant species cited in text

Common name Scientific name


Crop species
Corn Zea mays L.
Wheat Triticum aestivum L.
Soybean Glycimne max (L.) Merr.
Rice Oryza sativa L.
Barley Hordeum vulgare L.
Peanut Arachis hypogaea L.
Cowpea Vigna unguiculata (L.) Walp.
Potato Solanum tuberosum L.
Cassava Manihot esculenta Crantz
Pigeon pea Cajanus cajan (L.) Millspaugh
Millet Pennisetum glaucum (L.) R. Br.
Kudzu Pueraria phaseoloides (Roxb.) Benth.
Graybean Mucuna cinerecum L.
Triticale X Triticosecale Wittmack
Buckwheat Fygopyrum esculentum Moench
Rye Secale cereale L.
Crotolaria Crotolaria breviflora
Pasture species
(grasses and legumes)
Gambagrass or Andropogon gayanus Kunth
Carimagua
Surinamgrass Brachiaria decumbens Stapf.
Pangolagrass Digitaria decumbens Stent
Jaraguagrass Hyparrhenia rufa (Nees) Stapf.
Paragrass Brachiaria mutica (Forsk) Stafp.
Guineagrass Panicum maximum Jacq.
Bahiagrass Paspalum notatum Fluegge
Napiergrass Pennisetum purpureum Schumach
Desmodium Desmodium ovalifolium Guillemin & Perrottet
Brazilian stylo Stylosanthes guianiensis (Aubl.) Scv.
Pueraria Pueraria phaseoloides Roxb.
Zornia Zornia latifolia
Plantation crops
Banana Musa paradisiaca L.
Cashew Anacardium occidentale
Coconut Cocos nucifera
Brazilian nut Bertholletia excelsa
Eucalyptus Eucalyptus grandiflora (Gaertn.) Hochr.
Coffee Coffea arabica L.
Rubber Hevea brasiliensis
(continued)
388 N. K. Fageria and V. C. Baligar

Table 20 (continued)

Common name Scientific name


Oil palm Elaeis guineensis
Guarana Paullinia cupana
Tea Cammellia Sinensis (L.) Ktze.
Papaya Carica papaya

bioavailability of micronutrients (Zn, Cu, Fe, Mn, B), which decreases with
increasing pH (Fageria et al., 2002). This can produce plant nutrient
deficiencies, particularly that of Fe.
Planting acid-tolerant crop species and genotypes within species is another
most cost-effective and environmentally sound approach to improve crop
production on acid soils. Some high soil acidity-tolerant crop species like
millet, rice, cassava, and cowpea could be produced with minimum input of
lime. In addition, some pasture grass and legume species have high tolerance
to soil acidity and these could be used in developing sustainable cropping
system for Oxisols. Some plantation crops also have high tolerance to soil
acidity and could be used effectively in acid ecosystem cropping systems,
along with annual crops. Low to medium acidity-tolerant crop species such as
soybean, sorghum, and peanut are suitable crops for low input farming
systems. High acidity-tolerant grasses, such as brachiaria and andropogon
and plantation crops such as rubber and oil palm, are suitable crops for
extremely acidic soils. In addition, soil acidity can be managed successfully
by adopting practices that maximize the biological activities in the soil, such as
nutrient mineralization and cycling, biological nitrogen fixation, and mycor-
rhizal symbiosis.
In areas where limestone quarries and transportation systems are well
developed, lime is almost always inexpensive and typically provides a very
high rate of return on investment. However, in some areas, transportation
costs have made lime very costly. In any case, lime should be viewed as an
investment in improving the production potentials of acid soils. Because of
the long-lasting residual effects of lime applications, the cost of liming may
be amortized over several crops. Common and scientific names of crop
species cited in the text are listed in Table 20.

ACKNOWLEDGMENT
We thank Dr. C. D. Foy for his critical review and valuable suggestions for improving this
manuscript.
Ameliorating Soil Acidity 389

REFERENCES
Adriano, D. C. (1986). ‘‘Trace Elements in the Terrestrial Environment.’’ Springer-Verlag,
New York.
Alam, S. M. (1981). Influence of aluminum on plant growth and mineral nutrition of barley.
Commun. Soil Sci. Plant Anal. 12, 121–138.
Alter, D., and Mitchell, A. (1992). Use of vermicompost extract as an aluminum inhibitor in
aqueous solutions. Commun. Soil Sci. Plant Anal. 23, 231–240.
Alvarez, V. H., and Ribeiro, A. C. (1999). Liming. In ‘‘Recommendations for Using
Amendments and Fertilizers, 5th Approximation’’ (A. C. Ribeiro, P. T. G. Guimarães,
and V. H. Alvarez, Eds.), pp. 43–60. Soil Fertility Commission for State of Minas Gerais,
Viçosa, Brazil.
Anderson, R. (2003). An ecological approach to strengthen weed management in the
semiarid Great Plains. Adv. Agron. 80, 33–62.
Angle, J. S. (1998). Impact of biosolids and co-utilization wastes on rhizobia, nitrogen
fixation and growth of legumes. In ‘‘Beneficial Co-Utilization of Agricultural, Municipal
and Industrial by-Products’’ (S. Brown, J. S. Angle, and L. Jacobs, Eds.), pp. 235–245.
Kluwer Academic Publishers, Dordrecht.
Anjos, J. T., and Rowell, D. L. (1987). The effect of lime on phosphorus adsorption and
barley growth in three acid soils. Plant Soil 103, 75–82.
Baligar, V. C., and Ahlrichs, J. L. (1998). Nature and distribution of acid soils in the world.
In ‘‘Proceeding of Workshop Develops a Strategy for Collaborative Research and
Dissemination of Technology’’ (R. E. Schaffert, Ed.), pp. 1–11. Purdue University,
The Rockfeller Foundation and National Maize and Sorghum Research Center,
Embrapa, Brazil/West Lafayette Indiana.
Baligar, V. C., and Fageria, N. K. (1997). Nutrient use efficiency in acid soils: Nutrient
management and plant use efficiency. In ‘‘Plant–Soil Interaction at Low pH: Sustainable
Agriculture and Forestry Production’’ (A. C. Moniz, A. M. C. Furlani, R. E. Schaffert,
N. K. Fageria, C. A. Rosolem, and H. Cantarella, Eds.), pp. 75–95. Brazilian Society of
Soil Science, Campinas, Brazil.
Baligar, V. C., and Fageria, N. K. (1999). Plant nutrient efficiency: Towards the second
paradigm. In ‘‘Soil Fertility, Soil Biology and Plant Nutrition Interrelationships’’
( J. Q. Siqueira, F. M. S. Moreira, A. S. Lopes, L. R. G. Guilherme, V. Faquin,
A. E. Furtini Neto, and J. G. Carvallo, Eds.), pp. 183–204. Brazilian Soc. Soil Sci. and
University of Lavras, Lavras, MG Brazil.
Baligar, V. C., Wright, R. J., and Ritchey, K. D. (1992). Soil acidity effects on wheat
seedling growth. J. Plant Nutr. 15, 845–856.
Baligar, V. C., Fageria, N. K., and He, H. (2001). Nutrient use efficiency in plants. Commun.
Soil Sci. Plant Anal. 32, 921–950.
Baligar, V. C., Fageria, N. K., Eswaran, H., Wilson, M. J., and He, Z. (2004). Nature and
properties of red soils of the world. In ‘‘The Red Soils of China: Their Nature,
Management and Utilization’’ (M. J. Wilson, Z. He, and X. Yang, Eds.), pp. 171–218.
Kluwer Academic Publishers, Dordrecht.
Barber, S. A. (1984). Liming materials and practices. In ‘‘Soil Acidity and Liming’’
(F. Adams, Ed.), 2nd Ed. pp. 171–209. ASA-CSSA-SSSA, Madison, Wisconsin.
Basta, N. T., and Tabatabai, M. A. (1992). Effect of cropping systems on adsorption of metals
by soils. II. Effect of pH. Soil Sci. 153, 195–204.
Bhandhari, B., and Nicholas, D. J. D. (1985). Proton motive force in washed cells of
Rhizobium japonicum and bacteroids from Glycine max. J. Bacteriol. 164, 1383–1385.
Bohn, H., McNeal, B. L., and O’Connor, G. A. (2001). ‘‘Soil Chemistry.’’ 3rd ed. John
Wiley & Sons, New York.
390 N. K. Fageria and V. C. Baligar

Bolan, N. S., and Hedley, M. J. (2003). Role of carbon, nitrogen, and sulfur cycles in soil
acidification. In ‘‘Handbook of Soil Acidity’’ (Z. Rengel, Ed.), pp. 29–56. Marcel
Dekker, New York.
Bolan, N. S., Naidu, R., Syers, J. K., and Tillman, R. W. (1999). Surface charge and solute
interactions in soils. Adv. Agron. 67, 88–141.
Bolan, N. S., Adriano, D. C., and Curtin, D. (2003). Soil acidification and liming interac-
tions with nutrient and heavy metal transformation and bioavailability. Adv. Agron. 78,
215–272.
Borlaug, N. E., and Dowswell, C. R. (1997). The acid lands: One of agricultures last
frontiers. In ‘‘Plant Soil Interactions at Low pH: Sustainable Agriculture and Forestry
Production’’ (A. C. Moniz, A. M. C. Furlani, R. E. Schaffert, N. K. Fageria,
C. A. Rosolem, and H. Cantarella, Eds.), pp. 5–15. Brazilian Soil Science Society,
Campinas, São Paulo, Brazil.
Bowman, R. A., and Halvorson, A. D. (1997). Crop rotation and tillage effects on phos-
phorus distribution in the Central Great Plains. Soil Sci. Soc. Am. J. 61, 1418–1422.
Brady, N. C., and Weil, R. R. (2002). ‘‘The Nature and Properties of Soils.’’ 13th ed.
Prentice Hall, Upper Saddle River, New Jersey.
Brockwell, J., Bottomley, P. J., and Janice, E. T. (1995). Manipulation of rhizobia microflora
for improving legume productivity and soil fertility: A critical assessment. Plant Soil 174,
143–180.
Brosius, M. R., Evanylo, G. K., Bulluck, L. R., and Ristaino, J. B. (1998). Comparison of
commercial fertilizer and organic by products on soil chemical and biological properties
and vegetable yields. In ‘‘Beneficial Co-Utilization of Agricultural, Municipal and
Industrial by-Products’’ (S. Brown, J. S. Angle, and L. Jacobs, Eds.), pp. 195–202.
Kluwer Academic Publishers, Dordrecht.
Brown, S. R., Chaney, R., and Angle, J. S. (1997). Subsurface liming and metal movement
in soils amended with lime-stabilized biosolids. J. Environ. Qual. 26, 724–732.
Brummer, G. W., and Herms, U. (1983). Influence of soil reaction and organic matter on
solubility of heavy metals in soils. In ‘‘Effects of Accumulation of Air Pollutants in Forest
Ecosystems’’ (B. Ulrich and J. Pankrath, Eds.), pp. 233–243. Reidel, Dordrecht.
Buol, S. W., and Eswaran, H. (2000). Oxisols. Adv. Agron. 68, 151–195.
Burle, M. L., Mielniczuk, J., and Focchi, S. (1997). Effect of cropping systems on soil
chemical characteristics, with emphasis on soil acidification. Plant Soil 190, 309–316.
Caires, E. F., Alleoni, L. R. F., Cambri, M. A., and Barth, G. (2005). Surface application of
lime for crop production under no-tillage system. Agron. J. 97, 791–798.
Carson, C. D., and Dixon, J. B. (1979). Acidity. In The Encyclopedia of Soil Science, part 1’’
(R. W. Fairbridge and C. W. Finkl, Eds.), pp. 1–3. Dowden, Hutchinson & Ross Inc.,
Stroudsburg, Pennsylvania.
Caudle, N. (1991). ‘‘Managing Soil Acidity.’’ North Carolina State University, Tropsoils
Publication, Raleigh, North Caroline.
Chan, K. Y., and Heenan, D. P. (1998). Effect of lime (CaCO3) application on soil structural
stability of a red earth. Aust. J. Soil Res. 36, 73–86.
Chartres, C. J., Cumming, R. W., Bettie, J. A., Bowman, G. M., and Wood, J. T. (1990).
Acidification of soils on a transect from plains to slopes, southwestern New South Wales.
Aust. J. Soil Res. 28, 539–548.
Clark, R. B. (1984). Physiological aspects of calcium and magnesium, and molybdenum
deficiencies in plants. In ‘‘Soil Acidity and Liming’’ (F. Adams, Ed.), 2nd Ed. pp. 99–170.
ASA-CSSA-SSSA, Madison, Wisconsin.
Clark, R. B., Pier, P. A., Knudsen, D., and Maranville, J. W. (1981). Effect of trace element
deficiencies and excesses on mineral nutrients in sorghum. J. Plant Nutr. 3, 357–374.
Clough, T. J., Sherlock, R. R., and Kelliher, F. M. (2003). Can liming mitigate N2O fluxes
from a urine-amended soil? Aust. J. Soil Res. 41, 439–457.
Ameliorating Soil Acidity 391

Clough, T. J., Kelliher, F. M., Sherlock, R. R., and Ford, C. D. (2004). Lime and soil
moisture effects on nitrous oxide emissions from a urine patch. Soil Sci. Soc. Am. J. 68,
1600–1609.
Cochrane, T. T. (1978). An ongoing appraisal of the savanna ecosystems of tropical America
for beef cattle production. In ‘‘Pasture Production on Acid Soils of the Tropics’’
(P. A. Sanchez and L. F. Teggas, Eds.), pp. 1–12. CIAT, Cali, Colombia.
Cochrane, T. T., Salinas, J. G., and Sanchez, P. A. (1980). An equation for liming acid
mineral soils to compensate crop aluminum tolerance. Trop. Agric. 57, 133–140.
Cole, C. V., Williams, J., Shaffer, M., and Hanson, J. (1987). Nutrient and organic matter
dynamics as components of agricultural production systems models. In ‘‘Soil Fertility and
Organic Matter as Critical Components of Production Systems’’ (R. F. Follett, Ed.),
pp. 147–166. ASA-CSSSA-SSSA, Madison, Wisconsin.
Conyers, M. K., Poile, G. J., and Cullis, B. R. (1991). Lime responses by barley as related to
available soil aluminum and manganese. Aust. J. Agric. Res. 42, 379–390.
Coventry, D. R., and Slattery, W. J. (1991). Acidification of soil associated with lupins
grown in a crop rotation in northeastern Victoria. Aust. J. Agric. Res. 42, 391–397.
Coventry, D. R., Farhoodi, A., and Xu, R. (2003). Managing soil acidification through crop
rotations in southern Australia. In ‘‘Handbook of Soil Acidity’’ (Z. Rengel, Ed.),
pp. 407–429. Marcel Dekker, New York.
Cregan, D. D., Hirth, J. R., and Conyers, M. K. (1989). Amelioration of soil acidity by
liming and other amendments. In ‘‘Soil Acidity and Plant Growth’’ (A. D. Robson, Ed.),
pp. 205–264. Academic Press, Sydney.
Cronan, C. S., and Grigal, D. F. (1995). Use of calcium aluminum ratios as indicators of
stress in forest ecosystems. J. Environ. Qual. 24, 209–226.
Curi, N., and Camargo, O. A. (1988). Phosphorus adsorption characteristics of Brazilian
Oxisols. In ‘‘Proceedings of the 8th International Soil Classification Workshop’’ Part 1.
(F. H. Beinroth, M. N. Camargo, and H. Eswarn, Eds.), pp. 56–63. Soil Management
Support Services, U. S. Department of Agriculture, Washington, DC.
Curtin, D., and Syers, J. K. (2001). Lime-induced changes in indices of phosphate availability.
Soil Sci. Soc. Am. J. 65, 147–152.
Devine, T. E. (1976). Aluminum and manganese toxicities in legumes. In ‘‘Proceeding of
Workshop on Plant Adaptation to Mineral Stress in Problem Soils’’ (M. J. Wright, Ed.),
pp. 65–72. Cornell Universty, Ithaca, New York.
Dolling, P. J. (1995). Effect of lupines and location on soil acidification rates. Aust. J. Exp.
Agric. 35, 753–763.
Edmeades, D. C., and Perrott, K. W. (2004). The calcium requirements of pastures in New
Zealand: A review. N. Z. J. Agric. Res. 47, 11–21.
Elliott, H. A., Liberati, M. A., and Huand, C. P. (1986). Competitive adsorption of heavy
metals by soils. J. Environ. Quality 15, 214–219.
EMBRAPA (Empresa Brasileira de Pesquisa Agropecuaria). (1997). ‘‘Manual for Methods of
Soil Analysis.’’ 2nd ed. National Soil Research Center, Rio de Janeiro, Brazil.
Engelhard, A. W. (1989). Historical highlights and prospects for the future. In ‘‘Soil Born Plant
Pathogens: Management o Diseases with Macro and Microelements’’ (A. W. Engelhard,
Ed.), pp. 9–17. The American Phytopathological Society, St. Paul, Minnesota.
Epstein, E., and Bloom, A. J. (2005). ‘‘Mineral Nutrition of Plants: Principles and Perspec-
tives.’’ 2nd ed. Sinauer Associates, Inc. Publishers, Sunderland, Massachusetts.
Ernani, P. R., Bayer, C., and Maestri, L. (2002). Corn yield as affected by liming and tillage
system on an acid Brazilian Oxisol. Agron. J. 94, 305–309.
Fageria, N. K. (1984). Response of rice cultivars to liming in cerrado soil. Pesq. Agropec. Bras.
19, 883–889.
Fageria, N. K. (1989). ‘‘Tropical Soils and Physiological Aspects of Crops.’’ EMBRAPA-
CNPAF, Brasilia, Goiânia, Brazil.
392 N. K. Fageria and V. C. Baligar

Fageria, N. K. (1992). ‘‘Maximizing Crop Yields.’’ Marcel Dekker, New York.


Fageria, N. K. (2000). Upland rice response to soil acidity in cerrado soil. Pesq. Agropec. Bras.
35, 2303–2307.
Fageria, N. K. (2001a). Effect of liming on upland rice, common bean, corn, and soybean
production in cerrado soil. Pesq. Agropec. Bras. 36, 1419–1424.
Fageria, N. K. (2001b). Response of upland rice, dry bean, corn and soybean to base
saturation in cerrado soil. Rev. Bras. Eng. Agri. Amb. 5, 416–424.
Fageria, N. K. (2002). Soil quality vs. environmentally based agricultural management
practices. Commun. Soil Sci. Plant Anal. 33, 2301–2329.
Fageria, N. K. (2006). Liming and copper fertilization in dry bean production on an Oxisol
in no-tillage system. J. Plant Nutr. 29, 1219–1228.
Fageria, N. K. (2008). Optimum soil acidity indices for dry bean production on an Oxisol in
no-tillage system. Commun. Soil Sci. Plant Anal. 39(5/6), 845–857.
Fageria, N. K., and Baligar, V. C. (1999). Growth and nutrient concentrations of common
bean, lowland rice, corn, soybean and wheat at different soil pH on an Inceptisol. J. Plant
Nutr. 22, 1495–1507.
Fageria, N. K., and Baligar, V. C. (2001). Improving nutrient use efficiency of annual crops
in Brazilian acid soils for sustainable crop production. Commun. Soil Sci. Plant Anal. 32,
1303–1319.
Fageria, N. K., and Baligar, V. C. (2003a). Fertility management of tropical acid soils for
sustainable crop production. In ‘‘Handbook of soil acidity’’ (Z. Rengel, Ed.),
pp. 359–385. Marcel Dekker, New York.
Fageria, N. K., and Baligar, V. C. (2003b). Methodology for evaluation of lowland rice
genotypes for nitrogen use efficiency. J. Plant Nutr. 26, 1315–1333.
Fageria, N. K., and Baligar, V. C. (2005a). Enhancing nitrogen use efficiency in crop plants.
Adv. Agron. 88, 97–185.
Fageria, N. K., and Baligar, V. C. (2005b). Nutrient availability. In ‘‘Encyclopedia of Soils in
the Environment’’ (D. Hillel, Ed.), pp. 63–71. Elsevier, San Diego, California.
Fageria, N. K., and Breseghello, F. (2004). Nutritional diagnostic in upland rice production
in some municipalities of State of Mato Grosso, Brazil. J. Plant Nutr. 27, 15–28.
Fageria, N. K., and Gheyi, H. R. (1999). ‘‘Efficient Crop Production.’’ Federal University of
Paraiba, Campina Grande, Brazil, Paraiba.
Fageria, N. K., and Santos, A. B. (2005). In ‘‘Influence of base saturation and micronutrient
rates on their concentration in the soil and bean productivity in cerrado soil in no-tillage
system’’. Paper presented at the VIII National Bean Congress, Goiânia, Brazil, pp. 18–20
October 2005.
Fageria, N. K., and Santos, A. B. (2008). Influence of pH on productivity, nutrient use
efficiency by dry bean, and soil phosphorus availability in a no-tillage sysetm. Commun.
Soil Sci. Plant Anal. 39, 1016-1025.
Fageria, N. K., and Scriber, J. M. (2002). The role of essential nutrients and minerals in insect
resistance in crop plants. In ‘‘Insect and Plant Defense Dynamics’’
(T. N. Ananthakrishnan, Ed.), pp. 23–54. Science Publisher, Ensfield (NH), USA.
Fageria, N. K., and Stone, L. F. (1999). ‘‘Acidity Management of Cerrado and Varzea Soils
of Brazil.’’ Santo Antônio de Goiás, Brazil.
Fageria, N. K., and Stone, L. F. (2004). Yield of common bean in no-tillage system with
application of lime and zinc. Pesq. Agropec. Bras. 39, 73–78.
Fageria, N. K., Baligar, V. C., and Wright, R. J. (1989). The effect of aluminum on growth
and uptake of Al and P by rice. Pesq. Agropec. Bras. 24, 677–682.
Fageria, N. K., Baligar, V. C., and Edwards, D. G. (1990). Soil-plant nutrient relationships at
low pH stress. In ‘‘Crops as Enhancers of Nutrient Use’’ (V. C. Baligar and
R. R. Duncan, Eds.), pp. 475–507. Academic Press, San Diego, California.
Fageria, N. K., Guimarâes, C. M., and Portes, T. A. (1994). Iron deficiency in upland rice.
Lav. Arrozeira 47, 3–7.
Ameliorating Soil Acidity 393

Fageria, N. K., Zimmermann, F. J. P., and Baligar, V. C. (1995). Lime and phosphorus
interactions on growth and nutrient uptake by upland rice, wheat, common bean, and
corn in an Oxisol. J. Plant Nutr. 18, 2519–2532.
Fageria, N. K., Baligar, V. C., and Jones, C. A. (1997). ‘‘Growth and Mineral Nutrition of
Field Crops.’’ Marcel Dekker, New York 2nd Ed.
Fageria, N. K., Baligar, V. C., and Clark, R. B. (2002). Micronutrients in crop production.
Adv. Agron. 77, 185–268.
Fageria, N. K., Castro, E. M., and Baligar, V. C. (2004). Response of upland rice genotypes
to soil acidity. In ‘‘The Red Soils of China: Their Nature, Management and Utilization’’
(M. J. Wilson, Z. He, and X. Yang, Eds.), pp. 219–237. Kluwer Academic Publishers,
Dordrecht.
Fageria, N. K., Baligar, V. C., and Zobel, R. W. (2007). Yield, nutrient uptake, and soil
chemical properties as influenced by liming and boron application in common bean in a
no-tillage system. Commun. Soil. Sci. Plant Anal. 38, 1637–1653.
Farina, M. P. W., Sumner, M. E., Plank, C. O., and Letzsch, W. S. (1980). Effect of pH on
soil magnesium and its absorption by corn. Commun. Soil Sci. Plant Anal. 11, 981–992.
Fischer, K. S. (1998). Toward increasing nutrient-use efficiency in rice cropping systems:
The next generation of technology. Field Crops Res. 56, 1–6.
Foy, C. D. (1983). Plant adaptation to mineral stress in problem soils. Iowa State J. Res. 57,
339–354.
Foy, C. D. (1984). Physiological effects of hydrogen, aluminum and manganese toxicity in
acid soils. In ‘‘Soil Acidity and Liming’’ (F. Adams, Ed.), 2nd Ed. pp. 57–97. ASA-CSSA-
SSSA, Madison, Wisconsin.
Foy, C. D. (1992). Soil chemical factors limiting plant root growth. Adv. Soil Sci. 19, 97–149.
Franco, A. A., and Munns, D. N. (1982). Acidity and aluminum restraints on nodulation,
nitrogen fixation, and growth of Phaseolus vulgaris in nutrient solution. Soil Sci. Soc. Am. J.
46, 296–301.
Friesen, D. K., Juo, A. S. R., and Miller, M. H. (1980a). Liming and lime phosphorus-zinc
interactions in two Nigerian Ultisols. I. Interactions in the soil. Soil Sci. Soc. Am. J. 44,
1221–1226.
Friesen, D. K., Miller, M. H., and Juo, A. S. R. (1980b). Lime and lime-phosphate-zinc
interactions in two Nigerian Ultisols. II. Effects on maize root and shoot growth. Soil Sci.
Soc. Am. J. 44, 1227–1232.
Gallo, P. B., Mascarenhas, H. A. A., Quaggio, J. A., and Bataglia, O. C. (1986). Differential
responses of soybean and sorghum to liming. Rev. Bras. Ci. Solo 10, 253–258.
Garvin, D. F., and Carver, B. F. (2003). Role of genotype in tolerance to acidity and
aluminum toxicity. In ‘‘Handbook of Soil Acidity’’ (Z. Rengel, Ed.), pp. 387–406.
Marcel Dekker, New York.
Glenn, A. R., Tiwari, R. P., Reeve, W. G., and Dilworth, M. J. (1997). The response of
root nodule bacteria to acid stress. In ‘‘Plant Soil Interactions at Low pH: Sustainable
Agriculture and Forestry Production’’ (A. C. Moniz, A. M. C. Furlani, R. E. Schaffert,
N. K. Fageria, C. A. Rosolem, and H. Cantarella, Eds.), pp. 123–138. Brazilian Soil
Science Society, Campinas, São Paulo, Brazil.
Gonzales-Erico, E., Kamprath, E. J., Naderman, G. C., and Soares, W. V. (1979). Effect of
depth of lime incorporation on the growth of corn on an Oxisol of Central Brazil. Soil
Sci. Soc. Am. J. 43, 1155–1158.
Graham, P. H., Draeger, K. J., Ferrey, M. L., Conroy, M. J., Hammer, B. E., Martinez, E.,
Aarons, S. R., and Quinto, C. (1994). Acid pH tolerance in strains of Rhizobium and
Bradyrhizobium, and initial studies on the basis for acid tolerance of Rhizobium tropici
UMR1899. Can. J. Microbiol. 40, 198–207.
Habte, M. (1995). Soil acidity as a constraint to the application of vesicular-arbuscular
mycorrhizal technology. In ‘‘Mycorrhiza’’ (A. Varma and B. Hock, Eds.),
pp. 593–605. Springer-Verlag, New York.
394 N. K. Fageria and V. C. Baligar

Hall, J. L. (2002). Cellular mechanisms for heavy metal detoxification and tolerance. J. Exp.
Bot. 53, 1–11.
Hartemink, A. E. (2002). Soil science in tropical and temperate regions—some differences
and similarities. Adv. Agron. 77, 269–292.
Harter, R. D. (1983). Effect of soil pH on adsorption of lead, copper, zinc, and nickel. Soil
Sci. Soc. Am. J. 47, 47–51.
Hayman, D. S., and Tavares, M. (1985). Plant growth responses to vesicular-arbuscular
mycorrhiza. XV. Influence of soil pH on the symbiotic efficiency of different endo-
phytes. New Phytol. 100, 367–377.
Haynes, R. J. (1982). Effects of liming on phosphate availability in acid soils. A critical
review. Plant Soil 68, 289–308.
Haynes, R. J. (1983). Soil acidification induced by leguminous crops. Grass Forage Sci. 38,
1–11.
Haynes, R. J. (1984). Lime and phosphate in the soil–plant system. Adv. Agron. 37, 249–315.
Haynes, R. J., and Swift, R. S. (1988). Effects of lime and phosphate additions on changes in
enzyme-activities, microbial biomass and levels of extractable nitrogen, sulfur and phos-
phorus in an acid soil. Biol. Fertil. Soils 6, 153–158.
He, Z., Yang, X., Baligar, V. C., and Calvert, D. V. (2003). Microbiological and biochemical
indexing systems for assessing quality of acid soils. Adv. Agron. 78, 89–138.
Helyar, K. R. (1978). Effects of aluminum and manganese toxicity on legume growth.
In ‘‘Mineral Nutrition of Legumes in Tropical and Subtropical Soils’’ (C. S. Andrew and
E. J. Kamprath, Eds.), pp. 207–231. CSIRO, Melbourne, Australia.
Helyar, K. R., Cillis, B. R., Furniss, K., Kohn, G. D., and Taylor, A. C. (1997). Changes in
the acidity and fertility of a red earth soil under wheat-annual pasture rotations. Aust. J.
Agric. Res. 48, 561–586.
Holdings, A. J., and Lowe, J. F. (1971). Some effects of acidity and heavy metals on the
rhizobium-leguminous plant association. Plant Soil Special Volume, 153–166.
Huang, J. W., and Chen, J. (2003). Role of pH in phytoremediation of contaminated soils.
In ‘‘Handbook of Soil Acidity’’ (Z. Rengel, Ed.), pp. 49–472. Marcel Dekker, New
York.
Huang, J. W., and Grunes, D. L. (1992). Potassium/magnesium ratio effects on aluminum
tolerance and mineral composition of wheat forage. Agron. J. 84, 643–650.
Hue, N. V. (1992). Correcting soil acidity of a highly weathered Ultisol with chicken
manure and sewage sludge. Commun. Soil Sci. Plant Anal. 23, 241–264.
Hue, N. V., and Licudine, D. L. (1999). Amelioration of subsoil acidity through surface
application of organic manures. J. Environ. Qual. 28, 623–632.
Hue, N. V., Craddock, G. R., and Adams, F. (1986). Effect of organic acids on aluminum
toxicity in subsoil. Soil Sci. Soc. Am. J. 50, 28–34.
Ibekwe, M. A., Angle, J. S., Chaney, R. L., and Van Berkum, P. (1995). Sewage sludge and
heavy metal effects on nodulation and nitrogen fixation of legumes. Environ. Sci. Technol.
24, 11199–11204.
IPCC. (2001). ‘‘Climate change: The scientific basis. Contribution of Workshop Group I to
Third Assessment Report of the Intergovernmental Panel on Climate Change.’’
Cambridge University Press, Cambridge.
Ismail, H., Shamshuddin, J., and Omar, S. R. S. (1993). Alleviation of soil acidity in Ultisol
and Oxisol for corn growth. Plant Soil 151, 55–65.
Izaurralde, R. C., Lemke, R. L., Goddard, T. W., McConkey, B., and Zhang, Z. (2004).
Nitrous oxide emissions from agriculture topsequences in Alberta and Saskatchewan. Soil
Sci. Soc. Am. J. 68, 1285–1294.
Johansen, A., Jakobsen, I., and Jensen, E. S. (1993). External hyphae of vesicular-arbuscular
mycorrhizal fungi associated with Trifolium subterraneum L.: Hyphal transport of 32P
and 15N. New Phytol. 124, 61–68.
Ameliorating Soil Acidity 395

Johnson, J. P., Carver, B. F., and Baligar, V. C. (1997). Productivity in great Plains acid soils
of wheat genotypes selected for Al tolerance. Plant Soil 188, 101–106.
Joner, E. J., and Jakobsen, I. (1994). Contribution by two arbuscular mycorrhizal fungi to
P uptake by cucumber (Cucumis sativus L.) from 32P-labeled organic matter during
mineralization in soil. Plant Soil 163, 203–209.
Kaitibie, S., Epplin, F. M., Krenzer, E. G., Jr., and Zhang, H. (2002). Economics of lime and
phosphorus application for dual-purpose winter wheat production in low-pH soils.
Agron. J. 94, 1139–1145.
Kamprath, E. J. (1970). Exchangeable aluminum as a criterion for liming leached mineral
soils. Soil Sci. Soc. Am. Proc. 34, 252–254.
Kamprath, E. J. (1984). Crop responses to lime on soils in the tropics. In ‘‘Soil Acidity and
Liming’’ (F. Adams, Ed.), 2nd Ed. pp. 349–366. ASA-CSSA-SSSA, Madison, Wisconsin.
Kamprath, E. J., and Foy, C. D. (1985). Lime-fertilize-plant interactions in acid soils.
In ‘‘Fertilizer Technology and Use’’ (O. P. Engelstad, Ed.), 3rd Ed. pp. 91–151. Soil
Science Society of America, Madison, Wisconsin.
Kariuki, S. K., Zhang, H., Schroder, J. L., Edwards, J., Payton, M., Carver, B. F.,
Raun, W. R., and Krenzer, E. G. (2007). Hard red winter wheat cultivar responses to
a pH and aluminum concentration gradient. Agron. J. 99, 88–98.
Karlen, D. L., Varvel, G. E., Bullock, D. G., and Cruse, R. M. (1994). Crop rotations for the
21st century. Adv. Agron. 53, 1–45.
Kashket, E. (1985). The proton motive force in bacteria: A critical assessment of methods.
Annu. Rev. Microbiol. 39, 219–242.
Kiraly, Z. (1976). Plant disease resistance as influenced by biochemical effects on nutrients in
fertilizers. In ‘‘Fertilizer Use and Plant Health’’, pp. 33–46. International Potash Institute,
Bern, Switzerland.
Kirchner, M. J., Wollum, A. G., and King, L. D. (1993). Soil microbial populations and
activities in reduced chemical input agroecosystems. Soil Sci. Soc. Am. J. 57, 1289–1295.
Kochain, L. V. (1995). Cellular mechanisms of aluminum toxicity and resistance in plants.
Annu. Rev. Plant Physiol. Plant Mol. Biol. 46, 237–260.
Lance, J. C., and Pearson, R. W. (1969). Effect of low concentrations of aluminum on
growth and water and nutrient uptake by cotton roots. Soil Sci. Soc. Am. Proc. 33, 95–98.
Linderman, R. G. (1992). Vesicular-arbuscular mycorrhizae and soil microbial interactions.
In ‘‘Mycorrhizae in Sustainable Agriculture’’ (G. J. Bethlenfalvay and R. G. Linderman,
Eds.), pp. 45–70. ASA-CSSSA-SSSA, Madison, Wisconsin.
Lindsay, W. L. (1979). ‘‘Chemical Equilibria in Soils.’’ John Wiley & Sons, New York.
Liu, J., and Hue, N. V. (1996). Ameliorating subsoil acidity by surface application of calcium
fulvates derived from common organic materials. Biol. Fertil. Soils 21, 264–270.
Lopes, A. S., Silva, M. C., and Guilherme, L. R. G. (1991). ‘‘Soil Acidity and Liming.’’
Technical Bulletin 1, National Association for Diffusion of Fertilizers and Agricultural
Amendments, São Paulo, Brazil.
Loss, S. P., Ritchie, G. S. P., and Robson, A. D. (1993). Effect of lupines and pasture on soil
acidification and fertility in Western Australia. Aust. J. Exp. Agric. 33, 457–464.
MacRae, R. J., and Meheys, G. R. (1985). The effect of green manuring on the physical
properties of temperate area soils. Adv. Soil Sci. 3, 71–91.
Mahler, R. L., and McDole, R. E. (1985). The influence of lime and phosphorus on crop
production in northern Idaho. Commun. Soil Sci. Plant Anal. 16, 485–499.
Makinde, E. A., and Agboola, A. A. (2002). Soil nutrient changes with fertilizer type in
cassava based cropping system. J. Plant Nutr. 25, 2303–2313.
Mamo, T., and Killham, K. S. (1987). Effect of soil liming and vesicular-arbuscular mycor-
rhizal inoculation on the growth and micronutrient content of the teff plant. Plant Soil
102, 257–259.
396 N. K. Fageria and V. C. Baligar

Mansell, G. P., Pringle, R. M., Edmeades, D. C., and Shannon, P. W. (1984). Effects of lime
on pasture production on soils in the North Island of New Zealand. III. Interaction of
lime with phosphorus. N. Z. J. Agric. Res. 27, 363–369.
Marschner, H. (1995). ‘‘Mineral Nutrition of Higher Plants.’’ 2nd ed. Academic Press,
New York.
Marschner, H., and Dell, B. (1994). Nutrient uptake in mycorrhizal symbiosis. Plant Soil
159, 89–102.
McBride, M. B. (1994). ‘‘Environmental Chemistry of Soils.’’ Oxford University Press,
New York.
McLean, E. O. (1973). Testing soils for pH and lime requirement. In ‘‘Soil Testing and Plant
Analysis’’ (L. M. Walsh and J. D. Beaton, Eds.), pp. 77–95. Soil Science Society of
America, Madison, Wisconsin.
McLean, A. J. (1976). Cadmium in different species and its availability in soils as influenced
by organic matter and addition of lime, P, Cd, and Zn. Can. J. Soil Sci. 56, 129–138.
Miyazawa, M., Pavan, M. A., and Calegari, A. (1993). The effect of plant materials on soil
acidity. Rev. Bras. Ci. Solo. 17, 411–416.
Moody, P. W., and Aitken, R. L. (1997). Soil acidification under some tropical agricultural
systems. I. Rates of acidification and contributing factors. Aust. J. Soil Res. 35, 63–173.
Mora, M. L., Demanet, R., Vistoso, E., and Gallardo, F. (2005). Influence of sulfate
concentration in mineral solution on ryegrass grown at different pH and aluminum
levels. J. Plant Nutr. 28, 1117–1132.
Morel, C., and Plenchette, C. (1994). Is the isotopically exchangeable phosphate of a loamy
soil the plant available P? Plant Soil 158, 287–297.
Mortvedt, J. J. (2000). Bioavailability of micronutrients. In ‘‘Handbook of Soil Science’’
(M. E. Sumner, Ed.), pp. 71–87. CRC Press, Boca Raton, Florida.
Muchovej, R. M. C., Borges, A. C., Novias, R. F., and Thiebaut, J. T. L. (1986). Effect of
liming levels and Ca-Mg ratios on yield, nitrogen content and nodulation of soybean
grown in acid cerrado. J. Soil Sci. 37, 235–240.
Mulder, E. G., Lie, T. A., and Houwers, A. (1977). The importance of legumes under
temperate conditions. In ‘‘Treatise on Dinitrogen Fixation, IV Agronomy and Ecology’’
(R. W. F. Hardy and H. A. Gibson, Eds.), pp. 221–242. John Wiley & Sons, New York.
Naidu, R., Syers, J. K., Tillman, R. W., and Kirkman, J. H. (1990). Effect of liming and
added phosphate on charge characteristics of acid soils. J. Soil Sci. 41, 157–164.
Narro, L., Pandey, S., Leon, C. D., Salazar, F., and Arias, M. P. (2001). Implication of soil-
acidity tolerant maize cultivars to increase production in developing countries. In ‘‘Plant
Nutrient Acquisition: New Perspectives’’ (N. Ae, J. Arihara, K. Okada, and
A. Srinivasan, Eds.), pp. 447–463. Springer, Tokyo.
Nichol, B. E., Oliveira, L. A., Glass, A. D., and Siddiqi, M. Y. (1993). The effect of
aluminum on the influx of calcium, potassium, ammonium, nitrate and phosphate in
an aluminum sensitive cultivar of barley (Hordeum vulgare L.). Plant Physiol. 101,
1263–1266.
Nurlaeny, N., Marschner, H., and George, E. (1996). Effects of liming and mycorrhizal
colonization on soil phosphate depletion and phosphate uptake by maize (Zea mays L.)
and soybean (Glycine max L.) grown in two tropical acid soils. Plant Soil 181, 275–285.
Okada, K., and Fischer, A. J. (2001). Adaptation mechanisms of upland rice genotypes to
highly weathered acid soils of South American savannas. In ‘‘Plant Nutrient Acquisition:
New Perspectives’’ (N. Ae, J. Arihara, K. Okada, and A. Srinivasan, Eds.), pp. 185–200.
Springer, Tokyo.
Parker, D. R., Kinraide, T. B., and Zelazny, I. W. (1989). On the phytotoxicity of
polynuclear hydroxyl-aluminum complexes. Soil Sci. Soc. Am. J. 53, 789–796.
Paula, M. B., Nogueira, F. D., Andrade, H., and Pitts, J. E. (1987). Effect of liming on dry
matter yield of wheat in pots of low humic gley soil. Plant Soil 97, 85–91.
Ameliorating Soil Acidity 397

Perez, T., Trumbore, S. E., Tyler, S. C., Matson, P. A., Ortiz-Monasterio, I., Rahn, T., and
Griffith, D. W. T. (2001). Identifying the agricultural imprint on the global N2O budget
using stable isotopes. J. Geophy. Res. Atmosp. 106, 9869–9878.
Porter, W. M., Robson, A. D., and Abbott, L. K. (1987). Factors controlling the distribution
vesicular-arbuscular mycorrhizal fungi in relation to soil pH. J. Appl. Ecol. 24, 663–672.
Poss, R., Smith, C. J., Dunin, F. X., and Angus, J. F. (1995). Rate of soil acidification under
wheat in a semi-arid environment. Plant Soil 177, 85–100.
Raij, B. V. (1991). ‘‘Soil Fertility and Fertilization.’’ Agronomy Editor Ceres, São Paulo.
Raij, B. V., and Quaggio, J. A. (1997). Methods used for diagnosis and correction of soil
acidity in Brazil: An overview. In ‘‘Plant Soil Interactions at Low pH: Sustainable
Agriculture and Forestry Production’’ (A. C. Moniz, A. M. C. Furlani, R. E. Schaffert,
N. K. Fageria, C. A. Rosolem, and H. Cantarella, Eds.), pp. 205–214. Brazilian Soil
Science Society, Campinas, São Paulo, Brazil.
Raij, B. V., Silva, N. M., Bataglia, O. C., Quaggio, J. A., Hiroce, R., Cantarella, H.,
Bellinazzi, R., Jr., Dechen, A. R., and Trani, P. E. (1985). ‘‘Fertilizer and lime recom-
mendations for the State of São Paulo, Brazil. Tec. Bulletin 100, Campinas.’’ Agronomy
Institute, Brazil.
Raun, W. R., and Johnson, G. V. (1999). Improving nitrogen use efficiency for cereal
production. Agron. J. 91, 357–363.
Reeve, W. G., Tiwari, R. P., Dilworth, M. J., and Glenn, A. R. (1993). Calcium affects the
growth and survival of Rhizobium meliloti. Soil Biol. Biochem. 25, 581–586.
Reid, D. A. (1976). Aluminum and manganese toxicities in the cereal grains. In ‘‘Proceeding
of Workshop on Plant Adaptation to Mineral Stress in Problem Soils’’ (M. J. Wright,
Ed.), pp. 55–64. Cornell University, Ithaca, New York.
Reis, T. C., and Rodella, A. A. (2002). Dynamics o organic matter degradation and pH
variation of soil under different temperatures. Rev. Bras. Ci. Solo 26, 619–626.
Ribeiro, M. R., Siqueira, J. O., Curi, N., and Simão, J. B. P. (2001). Fractioning and
bioavailability of heavy metals in contaminated soil incubated with organic and inorganic
materials. Rev. Bras. Ci. Solo 25, 495–507.
Ridley, A. M., Helyar, K. R., and Slattery, W. J. (1990). Soil acidification under subterra-
nean clover (Trifolium subterraneum L.) pastures in north-eastern Victoria. Aust. J. Exp.
Agric. 30, 195–201.
Robertson, G. P., and Tiedje, J. M. (1987). Nitrous oxide sources in aerobic soils: Nitrifica-
tion, denitrification and other biological processes. Soil Biol. Biochem. 19, 187–193.
Robson, A. D. (1989). ‘‘Soil Acidity and Plant Growth.’’ Academic Press, Sydney.
Robson, A. D., and Pitman, J. B. (1983). Interactions between nutrients in higher plants.
In ‘‘Inorganic Plant Nutrition: Encyclopedia of Plant Physiology’’ (A. Lauchli and
R. L. Bieleski, Eds.), Vol. 1, pp. 147–180. Springer-Verlag, New York.
Robson, M. C., Fowler, S. M., Lampkin, N. H., Leifert, C., Leitch, M., Robinson, A. D.,
Watson, C. A., and Litterick, A. M. (2002). The agronomic and economic potential of
break crops for ley/arable rotations in temperate organic agriculture. Adv. Agron. 77,
369–427.
Sanchez, P. A., and Logan, T. J. (1992). Myth and science about the chemistry and fertility of
soils in the tropics. In ‘‘Myths and Science of the Soils of the Tropics’’ (R. Lal and
P. A. Sanchez, Eds.), Soil Science Society of America, Madison, Wisconsin.
Sanchez, P. A., and Salinas, J. G. (1981). Low-input technology for managing Oxisols and
Ultisols in tropical America. Adv. Agron. 34, 280–406.
Santana, M. B., and Braga, J. M. (1977). Aluminum phosphorus interactions of acidic soils in
southern Bahia. Rev. Ceres 24, 200–211.
Sauve, S., Martinez, C. E., McBridge, M., and Hendershot, W. (2000). Adsorption of free
lead (Pb2þ) by pedogenic oxides, ferrihydrite, and leaf compost. Soil Sci. Soc. Am. J. 64,
595–599.
398 N. K. Fageria and V. C. Baligar

Scott, B. J., Fisher, J. A., and Cullins, B. R. (2001). Aluminum tolerance and lime increase
wheat yield on the acidic soils of central and southern New South Wales. Aust. J. Exp.
Agric. 41, 523–532.
Simonson, R. W. (1959). Outline of a generalized theory of soil genesis. Soil Sci. Soc. Am.
Proc. 23, 152–156.
Simpson, J. R., Pinkerton, A., and Lazdovokis, J. (1977). Effects of subsoil calcium on the
root growth of some lucerne genotypes (Medicago sativa L.). Aust. J. Agric. Res. 29,
629–638.
Siqueira, J. O., and Moreira, F. M. S. (1997). Microbial populations and activities in highly
weathered acidic soils: Highlights of the Brazilian research. In ‘‘Plant Soil Interactions at Low
pH: Sustainable Agriculture and Forestry Production’’ (A. C. Moniz, A. M. C. Furlani,
R. E. Schaffert, N. K. Fageria, C. A. Rosolem, and H. Cantarella, Eds.), pp. 139–156.
Brazilian Soil Science Society, Campinas, São Paulo, Brazil.
Smyth, T. J., and Cravo, M. S. (1992). Aluminum and calcium constraints to continuous
crop production in a Brazilian Oxisol. Agron. J. 84, 843–850.
Soil Science Society of America. (1997). ‘‘Glossary of Soil Science Terms.’’ Soil Science
Society of America, Madison.
Sousa, D. M. G., Miranda, L. N., and Lobato, E. (1996). ‘‘Evaluation Methods of Lime
Requirements in Cerrado Soils.’’ Cerrado Center of EMBRAPA, Planaltina, Brazil.
Sparks, D. L. (2003). ‘‘Environmental Soil Chemistry.’’ 2nd Ed. Academic Press, San Diego,
California.
Springett, J. A., and Syers, J. K. (1984). Effect of pH and calcium content of soil on
earthworm cast production in the laboratory. Soil Biol. Biochem. 16, 185–189.
Stevens, R. J., Laughlin, R. J., and Malone, J. P. (1998). Soil pH affects process reducing
nitrate to nitrous oxide and DI-nitrogen. Soil Biol. Biochem. 30, 1119–1126.
Sumner, M. E., and Farina, M. P. W. (1986). Phosphorus interactions with other nutrients
and lime in field cropping systems. Adv. Soil Sci. 5, 201–236.
Sumner, M. E., and Noble, A. D. (2003). Soil acidification: The world story. In ‘‘Handbook
of Soil Acidity’’ (Z. Rengel, Ed.), pp. 1–28. Marcel Dekker, New York.
Supriyo, H., Matsue, N., and Yoshinaga, N. (1992). Chemistry and mineralogy of some soils
from Indonesia. Soil Sci. Plant Nutr. 38, 217–225.
Tan, K. H., Edwards, J. H., and Bennett, O. L. (1985). Effect of sewage sludge on
mobilization of surface applied calcium in a greenville soil. Soil Sci. 139, 262–268.
Tang, C., and Rengel, Z. (2003). Role of plant cation/anion uptake ratio in soil acidifica-
tion. In ‘‘Handbook of Soil Acidity’’ (Z. Rengel, Ed.), pp. 57–81. Marcel Dekker,
New York.
Tarafdar, J. C., and Marschner, H. (1994). Phosphate activity in the rhizosphere and hypo-
sphere VA mycorrhizal wheat supplied with inorganic and organic phosphorus. Soil Biol.
Biochem. 26, 387–395.
Tarafdar, J. C., and Marschner, H. (1995). Dual inoculation with Aspergillus fumigatus and
Glomus mosseae enhances biomass production and nutrient uptake in wheat supplied with
organic phosphorus as Na-phytate. Plant Soil 173, 97–102.
Tester, C. F. (1990). Organic amendments effects on physical and chemical properties of
a sandy soil. Soil Sci. Soc. Am. J. 54, 827–831.
Thomas, G. W., and Hargrove, W. L. (1984). The chemistry of soil acidity. In ‘‘Soil Acidity
and Liming’’ (F. Adams, Ed.), 2nd Ed. pp. 3–56. ASA-CSSA-SSSA, Madison, Wisconsin.
Tisdale, S. L., Nelson, W. L., and Beaton, J. D. (1985). ‘‘Soil Fertility and Fertilizers.’’ 4th
ed. MacMillion, New York.
Tortoso, A. C., and Hutchinson, G. L. (1990). Contributions of autotrophic and heterotro-
phic nitrifiers to soil NO and N2O emissions. Appl. Environ. Microbiol. 56, 1799–1805.
Treder, W., and Cieslinski, G. (2005). Effect of silicon application on cadmium uptake and
distribution in strawberry plants grown on contaminated soils. J. Plant Nutr. 28, 917–929.
Ameliorating Soil Acidity 399

Ulrich, B., Mayer, R., and Khanna, P. K. (1980). Chemical changes due to acid precipitation
in a loess derived soil in central Europe. Soil Sci. 130, 193–199.
Van Wambeke, A. (1991). ‘‘Soils of the Tropics: Properties and Appraisal.’’ McGraw-Hill,
New York.
Von Uexkull, H., and Mutert, E. (1995). Global extent, development and economic impact
of acid soils. Plant Soil 171, 1–15.
Wagner, G. J. (1993). Accumulation of cadmium in crop plants and its consequence to
human health. Adv. Agron. 51, 173–212.
Wang, G. M., Stribley, D. P., Tinker, P. B., and Walker, C. (1985). Soil pH and vesicular-
arbuscular mycorrhizas. In ‘‘Ecological Interactions in Soil’’ (A. H. Fitter, Ed.),
pp. 219–224. Blackwell, Oxford.
Weaver, A. R., Kissel, D. E., Chen, F., West, L. T., Adkins, W., Rickman, D., and
Luvall, J. C. (2004). Mapping soil pH buffering capacity of selected fields in the coastal
plain. Soil Sci. Soc. Am. J. 68, 662–668.
Weil, R. R., Lowell, K. A., and Shade, H. M. (1993). Effects of intensity of agronomic
practices on a soil ecosystem. Am. J. Alternate Agric. 8, 5–14.
West, L. T., Beinroth, F. H., Sumner, M. E., and Kang, B. T. (1998). Ultisols: Character-
istics and impacts on society. Adv. Agron. 63, 179–236.
Westerman, R. L., Raun, W. R., and Johnson, G. V. (2000). Nutrient and water use
efficiency. In ‘‘Handbook of Soil Science’’ (M. E. Sumner, Ed.), pp. 175–189. CRC
Press, Boca Raron, Florida.
White, D. C., Braden, J. B., and Hornbaker, R. H. (1994). Economics of sustainable agricul-
ture. In ‘‘Sustainable Agriculture’’ (J. L. Hatfield and D. L. Karlen, Eds.), pp. 229–260. CRC
Press, Boca Raton, Florida.
Wilkinson, S. R., Grunes, D. L., and Sumner, M. E. (2000). Nutrient interactions in soil and
plant nutrition. In ‘‘Handbook of Soil Science’’ (M. E. Sumner, Ed.), pp. 89–112. CRC
Press, Boca Raton, Florida.
Willert, F. J. V., and Stehouwer, R. C. (2003). Compost, limestone, and gypsum effects on
calcium and aluminum transport in acidic minespoil. Soil Sci. Soc. Am. J. 67, 778–786.
Williams, C. H. (1980). Soil acidification under clover pasture. Aust. J. Exp. Agric. Animal
Husbandry 20, 561–567.
Wong, M. T. F., and Swift, R. S. (2003). Role of organic matter in alleviating soil acidity.
In ‘‘Handbook of Soil Acidity’’ (Z. Rengel, Ed.), pp. 337–358. Marcel Dekker, New
York.
Wong, M. T. F., Nortcliff, S., and Swift, R. S. (1998). Method for determining the acid
ameliorating capacity of plant residue compost, urban waste compost, farmyard manure
and peat applied to tropical soils. Commun. Soil Sci. Plant Anal. 29, 2927–2937.
Yakovchenko, V., Sikora, L. J., and Kaufman, D. D. (1996). A biological based indicator of
soil quality. Biol. Fertil. Soils 21, 245–251.
Yang, Z. M., Sivaguru, M., Horst, W. J., Matsumoto, H., and Yang, Z. M. (2000).
Aluminum tolerance is achieved by exudation of citric acid from roots of soybean
(Glycine max). Physiol. Plantarum 110, 72–77.
Yang, X., Wang, W., Ye, Z., He, Z., and Baigar, V. C. (2004). Physiological and genetic
aspects of crop plant adaptation to elemental stresses in acid soils. In ‘‘The Red Soils of
China: Their Nature, Management and Utilization’’ (M. J. Wilson, Z. He, and X. Yang,
Eds.), pp. 171–218. Kluwer Academic Publishers, Dordrecht.
Yang, J. L., Zheng, S. J., He, Y. F., Tang, C. X., and Zhou, G. D. (2005). Genotypic
differences among plant species in response to aluminum stress. J. Plant Nutr. 28,
949–961.
Zysset, M., Brunner, I., Frey, B., and Blaser, P. (1996). Response of European chestnut to
varying calcium/aluminum ratios. J. Environ. Qual. 25, 702–708.

You might also like