Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Published 1996

Chapter 16

Soil pH and Soil Acidity

G. W. THOMAS, University of Kentucky, Lexington, Kentucky

Soil pH is probably the single most informative measurement that can be made
to determine soil characteristics. At a single glance, pH tells much more about a
soil than merely indicating whether it is acidic or basic. For example, availabili-
ty of essential nutrients and toxicity of other elements can be estimated because
of their known relationship with pH.
The term pH was "invented" by the Swedish scientist Sorensen (1909) in
order to obtain more convenient numbers and the idea quickly caught on.
Gillespie and Hurst (1918) seem to have been among the earliest to determine pH
(or PH, as it was then called) electrometrically using a platinum-palladium black-
hydrogen gas electrode, a calomel reference electrode and a fairly cumbersome
potentiometer and galvanometer system. At that period, it was still much more
common to use colorimetric methods with indicator dyes than the electrometric
method.
This changed rapidly, however. Sharp and Hoagland (1919) used a similar
but less involved method than Gillespie and Hurst (1918) and Healy and Karraker
(1922) used a commercially available platinum-hydrogen gas electrode, poten-
tiometer and galvanometer which had been designed by Clark (1920).
The decade of the 1920s saw the development of the quinhydrone electrode
which was less fragile and much less expensive than the hydrogen-platinum elec-
trode. But, it was the development of the glass electrode in the 1930s that brought
the determination of pH very rapidly to its present importance and convenience.
The Beckman Model G pH meter (circa 1931) was practically indestructible and
could be used as a portable as well as a laboratory instrument. Although it was
cumbersome by today's standards, it was virtually foolproof (except for the con-
stantly failing batteries) and many are still capable of operating if not actually
operating today.
As recently as two decades ago, the use of the small, handheld portable pH
meters then available to determine pH in the field was a very imprecise and haz-
ardous undertaking because both electrodes and meters were subject to sudden
failures but this has changed rather abruptly in the last few years. Microcircuitry
and plastic have contributed to rugged pH meters and electrodes that withstand

Copyright © 1996 Soil Science Society of America and American Society of Agronomy, 677 S.
Segoe Rd., Madison, WI 53711, USA. Methods of Soil Analysis. Part 3. Chemical Methods-SSSA
Book Series no. 5.

475
476 THOMAS

considerable abuse and which, in terms of real cost, are very inexpensive com-
pared to the pH meters and electrodes of just a few years ago.
The value of pH as an indicator of soil conditions and the quality of the
equipment now available at reasonable cost make the acquisition of pH data a
fairly simple chore at present. This chapter discusses the methods used, their
strengths and weaknesses and some aspects of interpretation of the data.

DEFINITION OF pH
The concept of pH was derived from the ion product of water, which dis-
sociates very slightly

H20 ... H+ + OH- [1]

Kw = [W][OH] =1 x 10-14
at 25°C temperature where H+ and OH- in brackets are activities. When [H+] and
[OW] are equal, each has an activity of (10- 14 )1/2 or (10-7). pH was defined by
Sorensen (1909) as the negative logarithm to base 10 of the hydrogen ion con-
centration, but is now defined in terms of hydrogen ion activity

1
pH = or - log [W] [2]
log [W]

Thus, the pH of pure water would be -log of 1 x 10-7 or 7. As defined, any solu-
tion with a pH below seven is considered acidic, and one with a pH greater than
seven is defined as basic.
Because only dissociated hydrogen ions affect pH, the type of acid that is
present is important in interpreting pH values. Acids are roughly divided into
strong and weak based on their degree of dissociation. Strong acids are almost
completely dissociated (e.g., hydrochloric acid) while weak acids are slightly
dissociated (e.g., acetic acid). In the case of weak acids, the total amount of acid
in solution can be calculated from the pH and the dissociation constant using the
relationship

[3a]

or in a form that is simpler to use

[A-]
PH - pK. = log-- [3b]
a [HA]

(where Ka is the dissociation constant for the acid, [A-] is the activity of dissoci-
ated anion and [HA] is the activity of undissociated acid in solution. Because in
pure acid systems (A-) = (H+), where ( ) refers to concentrations, the negative
log of a weak acid concentration is given by
SOIL pH & SOIL ACTIVITY 477

pHA = 2pH - pKa [4]

At a concentration of 1 M, pHA = 0 and pH = 1/2 pKa' so that pH also can be


used to determine the value of pKa in a molar concentration of acid.
In mixed acid-salt systems (buffers), Eq. [3b] can be used to estimate de-
gree of dissociation. A modification of that equation, the Henderson-Hasselbalch
equation, has been used to estimate degree of dissociation in organic matter
(Hargrove & Thomas, 1981) or other pH-dependent exchangers in soil

a
pH = pKap = ~log - - [5]
1- a

(where Kap is the apparent dissociation constant, a is the degree of dissociation,


and ~ is a constant which, for organic matter, and other polyacids generally has
a value of around two). This equation gives the pKap at half neutralization (a =
0.5) since the log portion disappears and pH = pKap.

SIGNIFICANCE OF SOIL pH VALUES

When used with soils, pH can aid the investigator greatly in determining
major soil characteristics. Although not adequate as a lime requirement indica-
tor, pH can be used to make a rough estimate of lime requirement and the rela-
tive availability of both P and many of the minor elements (Zn, Cu, B, Mn, Fe,
and Mo, for example). In a more specific way, particular pH values in water can
be used to predict rather well the dominant cations on soil cation exchangers at
the time of soil sampling and analysis. These characteristic pH values are
described below.

Presence of Free Acids


Generally, soil pH values in the neighborhood of two to three indicate the
presence of free mineral acid, usually H2S04 • A pH value much below 4.0 is
impossible to achieve with Al3+ -saturation so that acidity 10 to 100 times more
intense (pH 2-3) indicates not only the presence of H+ but also a continuing
source of them. A H-saturated soil, prepared from H2S04 will not persist but will
rather quickly form an AI-soil due to dissolution of the clay minerals by
exchangeable W (Coleman & Craig, 1961). Therefore, a soil with pH of two to
three always is found to have a bountiful supply of free acid. The usual source of
this acid is pyritic minerals which, upon oxidation, form the H2S04 . Likely occa-
sions where this excess of acid is found are in mine spoils and recently drained
marine sediments, both of which are frequently well-supplied with pyrites.
The presence of free acid indicates three serious problems: (i) plants will
not grow, (ii) the soil minerals are slowly (or rapidly) being dissolved and (iii)
there will be a high cost of ameliorating soil acidity using lime. In commercial
agriculture this cost most often cannot be met but, for land restoration, where it
must be paid, it can equal the value of the land itself. For example 100 Mg of
478 THOMAS

CaC03 per hectare is more typical than not of the lime required to control acid-
ity in these cases.

Presence of Aluminum Ions

At pH values of four to five, the presence of exchangeable, trivalent Al will


be encountered in mineral soils and, at times even in certain organic soils. The
hydrolysis of both exchangeable and solution Al in Al-saturated montmorillonite
and hectorite,

gives a minimum pH value of 3.84 in montmorillonite (Frink & Peech, 1963) and
as the Al3+ becomes small, the pH rises to 4.89 in hectorite (hence the range from
pH 4-5).
Values of pH in this range practically always indicate trouble ahead for
most crop plants because of the negative effects that Al has, especially on plant
root growth (Foy, 1984). Nevertheless, the case is not nearly so serious as that of
free H+ where no plant growth is possible and where costs are very high. In most
cases, the problems of exchangeable Al3+ can be resolved with a few megagrams
of CaC03, a treatment that usually lasts several years and produces economic
increases in crop yields.

Presence of Hydroxy-Aluminum

At a pH of 5.5 and above, exchangeable Al3+ is no longer present. Instead,


Al chemistry is dominated by a complex mixture of hydroxy-Al ions, many of
them highly polymerized and virtually nonexchangeable (Rich, 1960).
The smaller particles have a high net positive charge and are true solutes
(Hsu, 1989). These may be involved directly in suppression of root growth (Alva
et aI., 1986). The larger polycations have little direct effect on plant growth but
may strongly affect both P and K availability (Rich, 1964).
In many soils, the major part of the "buffering" region involved in practi-
cal liming is controlled by the hydroxy-Al ions adsorbed on both clay minerals
and on organic matter. From the .standpoint of acidity control in soils, these
hydroxy-Al compounds are very well buffered, indeed, (Coleman & Thomas,
1964) and while resisting the effect of liming materials, they also resist equally
well, the tendency of acidifying agents (such as NH4- containing fertilizers) to
reacidify the soil. They act about equally well as "sinks" for Wand OH- and, of
course, are a major source of the so-called pH-dependent charge (Thomas &
Hargrove, 1984).

Presence of Calcium Carbonate

At the other end of the scale, pH is invaluable as an indicator of excess


CaC03 in the soil. Although the effect of the partial pressure of CO2 on pH of
SOIL pH & SOIL ACTIVITY 479

calcareous soils is strong, normally soils with pH values of 7.6 to 8.3 are found
to be calcareous. Excess CaC03 in the soil immediately indicates that no money
need be spent on lime, that soil acidity will not be a problem and, unfortunately,
that minor elements such as Zn and Fe could prove to be troublesome. On the
whole, the conditions are good but there are some potential problems.
Knowledge of whether a soil is calcareous or not can be especially valu-
able in areas where both calcareous and noncalcareous soils are interspersed on
the landscape. Both soil treatments and expectations will be likely to vary
according to the presence or absence of CaC03.

Presence of Sodium Carbonate


When pH values stray towards nine, it can be inferred that CaC03 is no
longer controlling the system and that Na2C03 is assuming control with dire con-
sequences for the long-term health of the soil. When Na2C03 becomes dominant,
not only is Na an important cation on the soil exchangers but Ca ceases to be very
important because of its precipitation as CaC03. In a calcareous soil at normal
pH values of 8.3 or less, Ca in solution is relatively abundant but, as the pH rises
towards nine, CaC03 becomes so insoluble that Na2C03' not CaC03, buffers the
soil. Being a soluble salt, the system is swamped with Na at the expense of Ca.
When the concentrations of the normal salts of Na, such as NaCI or Na2S04 are
relatively low, the combination of exchangeable Na and Na2C03 with low Ca
causes dispersal of the clay and of the organic matter; the result is called black
alkali soils. Fortunately, these soils usually exist as only portions of the land-
scape but, even so, their appearance is a warning of further degradation of the
soil in the future.

FACTORS WHICH AFFECT SOIL pH

The Effect of Dilution


The ratio of water to soil in the suspension has the effect of increasing pH
as the ratio increases. However, the effect is not as straightforward as might be
expected. For example, increasing water content 10 times will not give one pH
unit increase as it would in a solution of acid. Instead, the increase in pH will be
only about 0.4 as shown in Fig. 16-1 from Davis (1943). What is the reason for
the relative insensitivity of suspension pH to dilution?
It appears that, even though pH is a measure of the H+ activity in solution,
in a suspension, the dilution of the soil sample tends to increase the dissociation
of H+ from the soil surface and, in addition, to increase the amount of hydroly-
sis from whatever forms of AI might be present. Both effects help to "buffer" the
solution and, therefore, to maintain the pH at a relatively stable value over a wide
range of dilution in acid soils. In soils with higher pH values, hydrolysis of basic
cations tends to maintain a stable pH with dilution. In a practical way, this effect
leads to the conclusion that water/soil ratios are not a highly important factor to
consider in interpreting soil suspension pH values as long as interpretation is
based on a consistent measurement and experience with that measurement.
480 THOMAS

0.60

0.50

0.40
1
I
<l
Q. 0.30 1 -----
_/--- - /--1-------- ----------·l

0.20

0.10

,
0.00
1: 1 1 :5 1:10

Soil:Water Ratio
Fig. 16-1. The effect of dilution on pH values of California soils showing standard deviation at each
dilution (data from Davis, 1943).

Salt Content

A major factor influencing pH of soils is the salt content of the soil solu-
tion. These salts may be a natural part of the soil such as NaCI, Na2S04 or
Ca(N03)2, or they may be added in the form of fertilizers. In any event, the usual
tendency of the salts is to lower the pH value of the soil progressively as the salt
concentration increases. In acid soils the effect apparently is due to both dis-
placement of AI3+ from the exchange complex and to increased hydrolysis of
various kinds of AI species in the presence of salt (Ragland & Coleman, 1960).
In calcareous soils, an effect of about the same magnitude also is observed
which rather obviously cannot be explained by the same mechanisms. Moore and
Loeppert (1987) relate this reduction in pH to the displacement of Ca2+ from
exchange sites by the salt. At a constant partial pressure of CO 2 of 60.6 Pa
(0.0006 atm), their equation for the pH of a salt-affected calcareous soil is

pH = 6.48 - 0.54 log [Ca2+]


where [Ca2+] is the activity of Ca in the soil solution.
It is worth pointing out also that in tropical soils, there is frequently a ten-
dency for pH values to rise in the presence of salts (Van Raij & Peech, 1972).
This behavior is an indication that salt is releasing more OH- than H+ from the
soil or, in other words, the soil has more sources of positive than of negative
charge. This behavior offers one possibility for the use of both salt and water pH
values for the chemical characterization of the soil. This characterization has
value for predicting the behavior of both sulfate and phosphate as well as the
cations that may be present.
There are two practical problems which have to do with the effect of salt
on pH: the first problem is the seasonal effect first noted by Baver (1927), and
the second problem is the interpretation (or reinterpretation) of pH values taken
SOIL pH & SOIL ACfIVITY 481

7
y-O.8269x-O.2936

r'=O.7S24 •
6


•• •
•• •
U 5
• •
~
c:
•• • •
•• •
J:
Cl. 4 •• •
•••••••• ••
•• •• • •• •••
3 •

2
4 5 6 7 8

pH in Hp

Fig. 16-2. A comparison of pH in water and in 1 M KCI for Nigerian soils (data from Okusami et
al.,1987).

in 0.01 M CaClz or M KCI vs. those taken in water. The seasonal effect is basi-
cally a reflection of the loss, formation or accretion of salts during various times
of the year. Under humid conditions, the soil is most nearly free of salts in early
spring due to winter leaching and lack of nitrification. At spring planting season,
the soil is likely to be high in salts due to rapid nitrification and to application f
fertilizer. As crops grow, salts are taken up gradually, lowering the salt content.
But, soon after harvest, the salt level in the soil can be quite high because of the
water deficit, absence of plant uptake, nitrification and mineralization of nutri-
ents in crop residues. Baver (1927) showed that the lowest soil pH occurred dur-
ing June, was nearly equaled in September and that the highest soil pH occurred
in April, generally confirming the scenario given above.
Attempting to deal with problems of variable salt by addition of salts does
not necessarily resolve the problem. As shown in Fig. 16-2, (Okusami et aI.,
1987) the addition of 1 M KCllowers pH about one unit at pH 4 and more than
two units at pH 8, giving still another set of numbers of calibrate mentally with
observed soil conditions.
The approach of Schofield and Taylor (1955) which has been widely used
in England and Canada (Turner et aI., 1963) is to determine pH in 0.01 M CaClz
The resulting difference pH - 1/2pCa has been called the lime potential because
it corresponds to the mean activity of Ca(OH)z. It is somewhat, but not entirely
invariable with naturally changing salt concentration in soils.
Webster and Harward (1959), using both water pH and pH - 1/2 p (Ca +
Mg), showed that the ,.2 (,.2 = proportion of variation explained by the simple
482 THOMAS

regression equation) for the relationship between water pH and percentage Ca +


Mg saturation was slightly better than the ,.z for a similar relationship between
pH - 1/2 P (Ca + Mg) and percentage Ca + Mg saturation. Perhaps more impor-
tantly, their values of l/2p(Ca + Mg) obtained using 0.001 M CaCl2showed prac-
tically no variation over a range of base saturations from 15 to 150%, good evi-
dence that Ca2+ from CaCl2 essentially covered up any soil differences. This is a
clear indication that the extra trouble involved in determining pH in 0.001 M
CaCl2 was not worth the effort.
In summary, even with the problems involved, a soil-water suspension
gives values which can be used with considerable confidence, if not precision,
for making practical field decisions. The use of salts of one kind or another to
correct for variable salts in the soil does not, in general, lead to better interpreta-
tion.

Carbon Dioxide Content

The effect of CO2 on the pH of calcareous soils is very large and, given the
variability of the partial pressure of CO2 in the soil atmosphere, it can be con-
sidered the single largest factor affecting the measured pH of calcareous soils.
Bradfield (1942) published data on the pH of CaC03 over a range of Pc~ and
Whitney and Gardner (1943) determined pH on a number of western U.S. soils
as affected by Pc~. Their data for soils are plotted on a graph drawn from
Bradfield's data for CaC03 (Fig. 16-3) and the agreement is excellent. Two soils
dominated by Na2C03 were not plotted because the pH values were so much
higher, but even they had the same slope as the other soils. Figure 16-3 shows
clearly that the reaction

is the major controlling factor which determines pH in calcareous soils.


Naturally, in soils which are dried and ground prior to determining the pH, the
effect of CO2 should not be so important but where pH is determined on a fresh
sample or in situ, the effect can be very large.

Suspension Effect

The reduction in pH (and the occasional rise) that occurs when the elec-
trodes are placed in the soil suspension rather than the supernatant solution above
the suspension has been called the suspension effect. The interpretations of the
suspension effect basically have been two: first, Marshall (1964) believed that H+
near the clay surface dissociated sufficiently to affect pH when the electrode was
placed near them. In other words, he believed that the pH was truly different in
the soil paste as compared to the supernatant solution. Second, according to
Coleman et aI. (1951), the suspension effect is primarily caused by the effect of
the electrical charge of the soil on the mobilities of K+ and Cl- from the calomel
electrode, rather than the glass electrode. For example, a strongly negative soil
would tend to promote K+ mobility and impede that of Cl whereas, a positively
SOIL pH & SOIL ACTIVITY 483

9.0 .---------------------------------------~

.,...
8.5

8.0

7.5

:r:a. 7.0 •

6.5

6.0

5.5
5.0 L - -_ _ _ _ _ _ _ _"---_ _ _ _ _...l.-_ _ _ _ _ _- - ' -_ _ _ _ _ _ _ _--'

-4.0 -3.0 -2.0 -1.0 0.0

log Pco,
Fig. 16-3_ The effects of PC02 on pH of calcareous soils (curve from Bradfield, 1942; points from
Whitney & Gardner, 1943).

charged soil (an oxisol horizon, for example) would have exactly the opposite
effect. A soil with very low charge would have very little effect at alL
Although it is generally accepted (Olsen & Robbins, 1971) that the sus-
pension effect is more likely the result of the second explanation (that is, it is
largely spurious), there are some indications that Marshall (1964) was correct
about relative H+ activity on the soil surface as compared with those in the soil
solution. For example, Swoboda and Kunze (1968), based their conclusions on
ionization of weak organic bases, concluded that the surface pH of clays had to
be lower than that of the bulk suspension in order for ionization of the bases to
occur.
Nevertheless, distances a few nanometers from the clay surface have very
little to do with those obtained with a large electrode in a soil suspension.
Perhaps, the emerging field of microelectrodes for pH measurement will shed
further light on the subject.

FUNDAMENTALS OF pH MEASUREMENT

Colorimetric Determinations of pH

Colorimetric methods for pH determination are based on weak acids or


weak bases, the colors of which change with undissociated or dissociated forms.
For example, the dissociation of an indicator weak acid gives

acid alkaline [6]


color color
484 THOMAS

The overall color observed is a function of the relative concentrations of


HI and 1-, which varies with pH according to the following equation

pH = pKHI + log -
W]
+ log fI- [7]
[HI]

where pKHI is the negative logarithm of the dissociation constant of HI and f r- is


the activity coefficient of the indicator ion. The color achieved will vary with the
concentration of solutions used because of its effect on fI-. For example, with
bromcresol green, chlorophenol red and bromothymol blue, pH values in a 0.5 M
NaCI concentration have a positive error of between 0.3 and 0.4 pH units (Peech,
1965).
Another error in colorimetric methods is caused by the use of ethanol to
dissolve some indicators. The dissociation constant is decreased, changing the
pH of color change to higher values. This is especially important when indica-
tors are used directly on the moist soil but not so important in soil suspensions.

Electrometric Measurements
In the vast majority of cases, pH is determined using a glass electrode-
calomel electrode system. This is represented by the following diagram

Ej
HgIHgClz, KCI(sat) I test solution Iglass 10.1M Hel, AgClIAg

where the single bars represent phase boundaries and the double bar represents a
liquid junction potential marked Ej. The observed voltage of the cell, E, is pro-
portional to the H+ activity of the solution on suspension and is given by

(E -E')F
pH - -'----'--- [8]
2.303RT

where R is the gas constant, T, the absolute temperature and F is the Faraday con-
stant. At 25°C, this reduces to

(E -E')
pH - -'----"- [9]
0.0591

The term E' includes the potential of the reference (calomel) electrode, the Ag-
AgCI portion of the glass electrode, the asymmetry potential of the glass elec-
trode and the liquid junction potential, E j • When the pH meter is standardized
with a buffer of known pH, the value of E' has been evaluated, and effectively
canceled out. Therefore,

pH = pHs + (E - Es) [10]


0.0591
SOIL pH & SOIL ACTIVITY 485

where pH~ is the pH of the standard buffer solution, Es is the voltage at pHs, and
E is the voltage of an unknown solution or suspension (Peech, 1965). This
assumes that the junction potential will not change from the standard buffer to a
soil suspension which, of course, it will as has been discussed earlier.
Equation [10] signifies that, in practical terms, the voltage change for 1 pH
unit is 0.0591 V (or 59.1 mY) at 25°C, and all commercial pH meters are cali-
brated on this basis.

ELECTROMETRIC MEASUREMENT OF SOIL pH

Equipment
There is currently such a proliferation of equipment for determining soil
pH that it is impractical to discuss them all. In general, the electrometer itself
varies in quality with price and the cheaper electrometers cannot be used for
readings more precise than ±0.1 pH unit. For field work they are entirely ade-
quate. For more precise laboratory work they probably are not.
Combination electrodes, which usually have a Ag-AgCI reference rather
than the usual calomel reference electrode, do not give results as accurate as
those obtained with the standard two-electrode system. Nevertheless, for field
work the combination electrodes may be preferable because they are easier to
use. Similarly, plastic-encased electrodes are clearly preferable for field use
because they resist breakage. On the other hand, standard glass electrodes are
much easier to clean between samples and should be more useful in laboratory
applications, unless careless operators are a problem. The new "Ross" electrodes
which use a platinum wire and a redox filler solution to replace the traditional
Ag-AgCl and Hg-HgCI 2/KCI systems have the advantage that the reference is
not sensitive to temperature changes and readings can be made much more
quickly. Their disadvantage is that they cost two to three times as much as con-
ventional electrodes.
Although digital readout is now much more common than the traditional
analog and has many advantages, particularly with indecisive operators, the lat-
ter has the advantage of visually determining the approach of pH to an equilibri-
um value. This can be a very useful feature in certain kinds of studies, such as
titrations.
It should be stated, in general, that the move to solid-state instruments has
eliminated many of the problems of the older machines. On the other hand, many
of those problems (shielding, grounding) could be handled by an experienced
operator. When the new machines fail, they tend to fail completely a require a
return trip to the manufacturer.

Standardization of the Meter


Most modem pH meters are equipped with a meter testing program to
check the meter separately. This is done with a shorting plug installed to deter-
mine whether the meter is stable internally. Having done this check-up accord-
ing to the operating manual, the meter and electrodes may be standardized.
486 THOMAS

For soils, a two-buffer standardization, at pH 7 and pH 4, should be per-


formed. This will determine whether the electrodes are working properly and
also will store the slope value (0.0591 V per unit of pH) in the memory of the
electrometer.
In general, with solid-state pH meters, the procedure is to use a buffer of
pH 7.0 and allow the machine to display the number. If the number is 7.0, the
value is approved by pressing the yes button and the electrodes are cleaned and
placed in the pH 4.0 buffer. The second buffer reading should be very close to
4.0. If it is, that number also can be approved by pressing the "yes" button and
the pH meter and electrodes are now ready for use. If the meter cannot be stan-
dardized, it indicates electrode problems which are covered in the next section.
On older pH meters without programmed check sequences and without
automatic temperature compensation (ATC), the procedure for standardization is
as follows: the pH 7 buffer is introduced, the manual temperature control is set
to the temperature of the buffer and the meter is set at 7.0. The electrodes are
rinsed and blotted and the pH 4.0 buffer is introduced. If the buffer value reads
4.0 :t 0.1, the instrument is now ready for use. If not, adjust the reading to pH 4.0
using the temperature compensation knob. Repeat this with the pH 7.0 buffer and
again with the pH 4.0 buffer until both readings are satisfactory. If this cannot be
done, one or both electrodes probably are faulty.

Electrode Problems

Glass used in the H+ -sensitive electrode is characterized by low electrical


resistance, low melting point, and high Na content (Dole, 1941). In addition, the
glass must have the tendency to hydrate easily if it is to be predictably related to
the activity of H+. That the adsorption of water by the pH-sensitive glass was of
overwhelming importance was known as early as the 1920s (Dole, 1941).
The most common problems with electrodes are caused by allowing them
to dry out. In the case of the glass electrode, there is a hydrated layer of glass
which facilitates the movement ofH+ (Westcott, 1978). Hydration does not occur
immediately when the glass electrode is placed in water or dilute acids so that
there will be a delay of some hours as proper hydration occurs. In addition, the
surface of the electrode, upon drying, may become coated with carbonates or
other compounds, rendering it somewhat impermeable to H+. The electrode
should be given alternate 5-min soakings in 0.1 M HCI followed by 0.1 M NaOH
and finally in 0.1 M HCI again prior to use. This treatment will do a good job of
removing any accumulated carbonates which restrict operation of the glass elec-
trode.
If the rejuvenation procedure outlined above does not resolve the problem,
the electrode may be lightly etched in 20% ammonium fluoride solution for 10
to 30 sec and then rinsed in water. This is a "last resort" treatment that cannot be
repeated regularly because the glass is already very thin.
The calomel reference electrode also suffers from drying but the cause of
potential problems is not due to the glass but, rather, the lack of flow of internal
filling solution from inside the electrode to the soil suspension. The electrode
SOIL pH & SOIL ACTIVITY 487

should be full of solution before use and the vent cap should be open so that flow
can occur. For soils measurements, it usually is preferred to have relatively slow
leakage to minimize contamination of the soil suspension. However if leakage is
too slow, the reading will be in error. The asbestos or glass fiber, or ceramic
opening in the reference electrode (all three are used) may be plugged by KCI
crystals (usually from inside the electrode) or by soil particles on the outside. It
is necessary that some measurable flow occur. This can be checked by filling and
cleaning the electrode and checking that slight wetting occurs.

PROCEDURE FOR SOIL pH MEASUREMENT

Equipment and Reagents

1. pH meter equipped with glass and reference electrode, or combination


electrode.
2. 50- or 100-mL beakers.
3. Pipet or automatic pipet of 10 mL.
4. Standard buffers, pH 7 and pH 4.
5. Deionized water.
6. 0.01 M CaCl2 solution.
7. 1 M KCl solution.

Determination-pH in water

1. Weigh out 10 g of air-dry soil in a 50- or 100-mL beaker.


2. Add 10 mL of deionized water to the soil in the beaker and mix well. A
stirring stick, or stirring machine can be used but care should be exer-
cised to minimize contamination. (For large-scale determinations, a
shaking machine can be employed as is done in most soil testing labo-
ratories.)
3. Let stand for 10 min.
4. Swirl the suspension in the beaker and insert the electrodes into the sus-
pension. Electrodes may be placed in the clear supernatant above the
soil, directly in the sedimented soil, or the entire suspension may be
stirred during the pH determination. The important thing is that the
measurements be carried out ina consistent way. In water pH determi-
nation, values taken in the supernatant generally will be slightly higher
than in the stirred suspension. With a salt pH, the differences between
the three techniques practically disappear. For exposed glass electrodes,
it is useful to have a stop of some type so that the bulb will not contact
the bottom of the beaker. McLean (1982) suggested a glass rod, slight-
ly longer than the electrode, mounted on the electrode holder that makes
contact with the beaker before the electrode does.
5. Read pH and record as pHw.
6. Between pH readings, rinse the electrodes with distilled water. Blotting
is not necessary.
488 THOMAS

pH in One-One Hundredth Molar Calcium Chloride

1. Repeat the above procedure (1-4) but use 0.01 M CaCl2 instead of
water to make the soil suspension.
2. Read pH and record as pHeacl2.

pH in One Molar Potassium Chloride

1. Repeat steps 1 to 4 but use 1 M KCI instead of water to make the soil
suspensions.
2. Read pH and record as pHKo .

ALTERNATIVE MEmODS FOR pH MEASUREMENT

Use of Microelectrodes

The possibility of taking microsite pH values in intact soil systems has


been facilitated by the use of microprocedures in plant cells (Felle & BertI,
1986). Conkling and Blanchar (1989) have used homemade microelectrodes for
this purpose and have obtained relatively stable and reproducible results. It is not
yet clear, however, that startling new findings about the variability of pH in a soil
matrix will be forthcoming.
Basically, the technique of Conkling and Blanchar, (1989) is to melt a
small cap of pulled H+ ion-sensitive glass over a pulled glass capillary, to install
a silver wire, coated with AgCI and to fill it with a solution. This requires micro-
scopic observation and manipulation and must be thought of as tedious work.
The reference electrode is a calomel electrode connected to a gel-filled
KCI salt bridge to assure that a complete circuit is made. Readings on suspen-
sions have shown close to 1:1 slopes with conventional electrodes so it is clear
that the microelectrodes have some promise.
It is quite interesting to note that the pioneers of microelectrodes are fol-
lowing the path of the earlier pioneers (Sharp & Hoagland, 1919), making the
equipment as they go.

Use of Test Kits

Test kits have always suffered somewhat by comparison with electrodes


because (i) there are only certain pH values where ionic dyes change, and (ii) the
color of the soil somewhat obscures those colors. Summing up (i) and (ii), it is
clear that using kits is partly an art and depends considerably on the pair of eyes
that is observing the color. Mason and Obenshain (1939) found good agreement
between colorimeters and electrode pH values, but agreement was in the range
of 0.3 to 0.4 pH units which is not very satisfactory for many applications.
Certainly, in many cases, these field kits can be useful in roughtly deter-
mining soil pH categories but they are more time-consuming than a portable pH
meter and, probably, more expensive in the long run, and not as good.
SOIL pH & SOIL ACTIVITY 489

pH Papers

The newer pH papers are reasonably useful for field work is accuracy of
closer than 0.5 pH units is not critical. These papers have maximum practical use
as a first approximation in the field when no pH meter is available. As such they
are helpful in progressive soil mapping programs or in trouble-shooting situa-
tions. Nevertheless, the values usually have to be rechecked using a pH meter
before complete confidence in the results can be attained.

REFERENCES
Alva, AK., D.G. Edwards, C.S. Asher, and EP.C. Blarney. 1986. Relationships between root length
of soybeans and calculated activities of aluminum monomers in nutrient solution. Soil Sci.
Soc. Am. 1. 50:959-962.
Baver, L.D. 1927. Factors affecting the W-ion con~entration of soils. Soil Sci. 23:399-414.
Bradfield, R. 1942. Calcium in the soil: I. Physico-chemical relations. Soil Sci. Soc. Am. Proc.
6:8-15.
Clark, W.M. 1920. The determination of hydrogen ions. Williams and Wilkins, Baltimore, MD.
Coleman, N.T., and D. Craig. 1961. The spontaneous alteration of hydrogen clay. Soil Sci. 91:14-18.
Coleman, N.T., and G.w. Thomas. 1964. Buffer curves of acid clays as affected by the presence of
ferric iron and aluminum. Soil Sci. Soc. Am. Proc. 28:187-190.
Coleman, N.T., D.E. Williams, T.R. Nielsen, and H. Jenny. 1951. On the validity of interpretations
of potentiometrically measured soil pH. Soil Sci. Soc. Am. Proc. 15: 1O~110.
Conkling, B.L., and R.W. Blancher. 1989. Glass microelectrode techniques for in-situ pH measure-
ments. Soil Sci. Soc. Am. J. 53:58-52.
Davis, L.E. 1943. Measurements of pH with the glass electrode as affected by soil moisture. Soil Sci.
56:405-422.
Dole, M. 1941. The glass electrode. John Wiley & Sons, New York.
Felle, H., and A BertI. 1986. The fabrication of H+-selective liquid-membrane micro- electrodes for
use in plant cells. 1. Exp. Bot. 37:141~1428.
Foy, C.D. 1984. Physiological effects of hydrogen, aluminum, and management toxicities in acid
soils. p. 57-97. In F. Adams (ed.) Soil acidity and liming. Agron. Monogr. 12. 2nd ed. ASA,
CSSA, and SSSA, Madison, WI.
Frink, C.R., and M. Peech. 1963. Hydrolysis and exchange reactions of the aluminum ion in hectorite
and montmorillonite suspensions. Soil Sci. Soc. Am. Proc. 27:527-530.
Gillespie, L.J., and L.A Hurst. 1918. Hydrogen ion concentration-Soil type-Common potato scab.
Soil Sci. 6:219-236.
Hargrove, W.L., and G.W. Thomas. 1981. Effect of organic matter on exchangeable aluminum and
plant growth in acid soils. p. 151-166. In R.H. Dowdy (ed.) Chemistry in the soil environ-
ment. ASA Spec. Publ. 40. ASA and SSSA, Madison, WI.
Healy, 0.1., and P.E. Karraker. 1922. The Clark hydrogen-electrode vessel and soil measurements.
Soil Sci. 13:323-328. .
Hsu, P.H. 1989. Aluminum hydroxides and oxy-hydroxides. p. 331-378. In J.B. Dixon and S.B.
Weed (ed.) Minerals in the soil environment. SSSA Book Ser. 1. SSSA, Madison, WI.
Marshall, C.E. 1964. The physical chemistry and mineralogy of soils. Vol. I. John Wiley & Sons, Inc.,
New York.
Mason, D.O., and S.S. Obershain. 1939. A comparison of methods for the determination of soil reac-
tion. Soil Sci. Soc. Am. Proc. 3:129-137.
Mclean, E.O. 1982. Soil pH and lime requirement. p. 199-224. In AL. Page et al. (ed.) Methods of
soil analysis. Part 2. 2nd ed. Agron Monogr. 9. ASA and SSSA, Madison, WI.
Moore, T.1., and R.H. Loeppert. 1987. Significance of potassium chloride pH of calcareous soils. Soil
Sci. Soc. Am. J. 51:908-912.
Okusami, T.A, R.H. Rust, and AS.R. Juo. 1987. Reactive characteristics of certain soils from south
Nigeria. Soil Sci. Soc. Am. J. 5:1256-1262.
Olsen, R.A, and J.E. Robbins. 1971. The cause of the suspension effect in resin-water systems. Soil
Sci. Soc. Am. ProC. 35:26G--265.
490 THOMAS

Peech, M. 1965. Hydrogen-ion activity. p. 914-926. In C.A. Black (ed.) Methods of soil analysis.
Part 2. Agron. Monogr. 9. ASA and SSSA, Madison, WI.
Ragland, J.L., and N.T. Coleman. 1960. The hydrolysis of aluminum salts in clay and soil systems.
Soil Sci. Soc. Am. Proc. 24:457-460.
Rich, c.1. 1960. Aluminum in interlayers of vermiculite. Soil Sci. Soc. Am. Proc. 24:26-32.
Rich, c.1. 1964. Effect of cation size and pH on potassium exchange in Nason soil. Soil Sci.
98:100-106.
Schofield, R.R., and A.W. Taylor. 1955. The measurement of soil pH. Soil Sci. Soc. Am. Proc.
19:164-167.
Sharp, L.T., and D.R. Hoagland. 1919. Notes on recent work concerning acid soils. Soil Sci.
7:197-200.
Sorensen, S.P.L. 1909. Enzyme studies: II. The measurement and importance of the hydrogen ion
concentration in enzyme reaction. Comp!. Rend. Trav. Lab. (Carlsberg). 8:1.
Swoboda, A.R., and G.W. Kunze. 1968. Reactivity of montmorillonite surfaces with weak organic
bases. Soil Sci. Soc. Am. Proc. 32:806-811.
Thomas, G.W., and w.L. Hargrove. 1984. The chemistry of soil acidity. p. 3-56. In F. Adams (ed.)
Soil acidity and liming. Agron. Monogr. 12. 2nd ed. ASA and SSSA, Madison, WI.
Turner, R.c., W.E. Nichol, and J.E. Brydon. 1963. A study of the lime potential. Soil Sci.
95:186-191.
Van Raij, B., and M. Peech. 1972. Electrochemical properties of some oxisols and alfisols of the trop-
ics. Soil Sci. Soc. Am. Proc. 36:587-593.
Webster, G.R., and M.E. Harward. 1959. Hydrogen and calcium ion ratios in dilute equilibrium solu-
tions as related to cation saturation. Soil Sci. Soc. Am. Proc. 23:446-451.
Westcott, C.c. 1978. pH measurements. Acad. Press, New York.
Whitney, R.S., and R. Gardner. 1943. The effect of carbon dioxide on soil reaction. Soil Sci.
55:127-141.

You might also like