Sie sind auf Seite 1von 38

Alkalinity

Sea surface alkalinity (from the GLODAP climatology).

Alkalinity (from Arabic "al-qalī"[1]) is the


capacity of water to resist changes in pH
that would make the water more acidic.[2]
(It should not be confused with basicity
which is an absolute measurement on the
pH scale.) Alkalinity is the strength of a
buffer solution composed of weak acids
and their conjugate bases. It is measured
by titrating the solution with a monoprotic
acid such as HCl until its pH changes
abruptly, or it reaches a known endpoint
where that happens. Alkalinity is
expressed in units of meq/L
(milliequivalents per liter), which
corresponds to the amount of monoprotic
acid added as a titrant in millimoles per
liter.

Although alkalinity is primarily a term


invented by oceanographers, [3] it is also
used by hydrologists to describe
temporary hardness. Moreover, measuring
alkalinity is important in determining a
stream's ability to neutralize acidic
pollution from rainfall or wastewater. It is
one of the best measures of the sensitivity
of the stream to acid inputs.[4] There can
be long-term changes in the alkalinity of
streams and rivers in response to human
disturbances.[5]

History
In 1884, Professor Wilhelm (William)
Dittmar of Anderson College, now the
University of Strathclyde, analysed 77
pristine seawater samples from around
the world brought back by the Challenger
expedition. He found that in seawater the
major ions were in a fixed ratio, confirming
the hypothesis of Johan Georg
Forchhammer, that is now known as the
Principle of Constant Proportions.
However, there was one exception. Dittmar
found that the concentration of calcium
was slightly greater in the deep ocean, and
named this increase alkalinity.

1884 was also the year when Svante


Arrhenius submitted his PhD theses in
which he advocated the existence of ions
in solution, and defined acids as
hydronium ion donors and bases as
hydroxide ions donors. For that work, he
received the Nobel Prize in Chemistry in
1903.

Thus Dittmar's alkalinity is the hydronium


cations which exist to balance electrically
the increase in calcium anions in deep
ocean water, although now the meaning
alkalinity has expanded.

Simplified summary
Alkalinity roughly refers to the amount of
bases in a solution that can be converted
to uncharged species by a strong acid.[6]
The cited author, James Drever, provides
an equation expressed in terms of molar
equivalents, which means the number of
moles of each ion type multiplied by (the
absolute value of) the charge of the ion.
For example, 1 mole of HCO31− in solution
represents 1 molar equivalent, while 1
mole of CO32− is 2 molar equivalents
because twice as many H+ ions would be
necessary to balance the charge. The total
charge of a solution always equals zero.

Quoting from page 52, "Ions such as Na+,


K+, Ca2+, Mg2+, Cl −, SO42−, and NO3− can
be regarded as "conservative" in the sense
that their concentrations are unaffected by
changes in the pH, pressure, or
temperature (within the ranges normally
encountered near the earth's surface and
assuming no precipitation or dissolution
of solid phases, or biological
transformations)."

On the left-hand side of the equation is the


sum of conservative cations minus the
sum of conservative anions. Balancing
this on the right side is the sum of the
anions that could be neutralized by added
H+ ions (non-conservative anions) minus
H+ ions already present, as indicated by
the pH. All numbers are molar equivalents.
This right side term is called total
alkalinity. It is, quoting Drever, "formally
defined as the equivalent sum of the
bases that are titratable with strong acid
(Stumm and Morgan, 1981)".[7] The listing
of ions shown on the right in Drever was
"mHCO3− + 2mCO32− + mB(OH)4− + mH3(SiO)4− +
mHS− + morganic anions + mOH− - mH+". Total
alkalinity is measured by adding a strong
acid until all the anions listed above are
converted to uncharged species. The total
alkalinity is not (much) affected by
temperature, pressure, or pH, though the
values of individual constituents are,
mostly being conversions between HCO3−
and CO32−.
Drever further notes that in most natural
waters, all anions except HCO3− and CO32−
have low concentrations. Thus carbonate
alkalinity, which is equal to mHCO3− +
2mCO32− is also approximately equal to the
total alkalinity.

Detailed description
Alkalinity or AT measures the ability of a
solution to neutralize acids to the
equivalence point of carbonate or
bicarbonate. The alkalinity is equal to the
stoichiometric sum of the bases in
solution. In the natural environment
carbonate alkalinity tends to make up
most of the total alkalinity due to the
common occurrence and dissolution of
carbonate rocks and presence of carbon
dioxide in the atmosphere. Other common
natural components that can contribute to
alkalinity include borate, hydroxide,
phosphate, silicate, dissolved ammonia,
the conjugate bases of some organic
acids, and sulfate. Solutions produced in a
laboratory may contain a virtually limitless
number of bases that contribute to
alkalinity. Alkalinity is usually given in the
unit mEq/L (milliequivalent per liter).
Commercially, as in the swimming pool
industry, alkalinity might also be given in
parts per million of equivalent calcium
carbonate (ppm CaCO3).

Alkalinity is sometimes incorrectly used


interchangeably with basicity. For
example, the addition of CO2 lowers the
pH of a solution. This increase reduces the
basicity; however, the alkalinity remains
unchanged (see example below).For total
alkalinity testing, N/10 H2SO4 is used by
hydrologists along with phenolphthalein
indicator.

Theoretical treatment
In typical groundwater or seawater, the
measured alkalinity is set equal to:
AT = [HCO3−]T + 2[CO32−]T + [B(OH)4−]T +
[OH−]T + 2[PO43−]T + [HPO42−]T +
[SiO(OH)3−]T − [H+]sws − [HSO4−]

(Subscript T indicates the total


concentration of the species in the
solution as measured. This is opposed to
the free concentration, which takes into
account the significant amount of ion pair
interactions that occur in seawater.)

Alkalinity can be measured by titrating a


sample with a strong acid until all the
buffering capacity of the aforementioned
ions above the pH of bicarbonate or
carbonate is consumed. This point is
functionally set to pH 4.5. At this point, all
the bases of interest have been protonated
to the zero level species, hence they no
longer cause alkalinity. In the carbonate
system the bicarbonate ions [HCO3−] and
the carbonate ions [CO32−] have become
converted to carbonic acid [H2CO3] at this
pH. This pH is also called the CO2
equivalence point where the major
component in water is dissolved CO2
which is converted to H2CO3 in an
aqueous solution. There are no strong
acids or bases at this point. Therefore, the
alkalinity is modeled and quantified with
respect to the CO2 equivalence point.
Because the alkalinity is measured with
respect to the CO2 equivalence point, the
dissolution of CO2, although it adds acid
and dissolved inorganic carbon, does not
change the alkalinity. In natural conditions,
the dissolution of basic rocks and addition
of ammonia [NH3] or organic amines leads
to the addition of base to natural waters at
the CO2 equivalence point. The dissolved
base in water increases the pH and titrates
an equivalent amount of CO2 to
bicarbonate ion and carbonate ion. At
equilibrium, the water contains a certain
amount of alkalinity contributed by the
concentration of weak acid anions.
Conversely, the addition of acid converts
weak acid anions to CO2 and continuous
addition of strong acids can cause the
alkalinity to become less than zero.[8] For
example, the following reactions take
place during the addition of acid to a
typical seawater solution:

B(OH)4− + H+ → B(OH)3 + H2O


OH− + H+ → H2O
PO4−3 + 2H+ → H2PO4−
HPO4−2 + H+ → H2PO4−
[SiO(OH)3−] + H+ → [Si(OH)40]

It can be seen from the above protonation


reactions that most bases consume one
proton (H+) to become a neutral species,
thus increasing alkalinity by one per
equivalent. CO3−2 however, will consume
two protons before becoming a zero level
species (CO2), thus it increases alkalinity
by two per mole of CO3−2. [H+] and [HSO4−]
decrease alkalinity, as they act as sources
of protons. They are often represented
collectively as [H+]T.

Alkalinity is typically reported as mg/L as


CaCO3. (The conjunction "as" is
appropriate in this case because the
alkalinity results from a mixture of ions but
is reported "as if" all of this is due to
CaCO3.) This can be converted into
milliEquivalents per Liter (mEq/L) by
dividing by 50 (the approximate MW of
CaCO3/2).
Example problems
Sum of contributing species

The following equations demonstrate the


relative contributions of each component
to the alkalinity of a typical seawater
sample. Contributions are in
μmol.kg−soln−1 and are obtained from A
Handbook of Methods for the analysis of
carbon dioxide parameters in seawater "
[1] ,"(Salinity = 35 g/kg, pH = 8.1,
Temperature = 25 °C).

AT = [HCO3−]T + 2[CO32−]T + [B(OH)4−]T +


[OH−]T + 2[PO43−]T + [HPO42−]T +
[SiO(OH)3−]T − [H+] − [HSO4−] − [HF]
Phosphates and silicate, being nutrients,
are typically negligible. At pH = 8.1 [HSO4−]
and [HF] are also negligible. So,

= [HCO3−]T + 2[CO32−]T + [B(OH)4−]T +


AT
[OH−]T − [H+]
= 1830 + 2 × 270 + 100 + 10 − 0.01
= 2480 μmol.kg−soln−1

Addition of CO2

Addition (or removal) of CO2 to a solution


does not change its alkalinity, since the net
reaction produces the same number of
equivalents of positively contributing
species (H+) as negative contributing
species (HCO3− and/or CO32−). Adding
CO2 to the solution lowers its pH, but does
not affect alkalinity.

At all pH values:

CO2 + H2O ⇌ HCO3− + H+

Only at high (basic) pH values:

HCO3− + H+ ⇌ CO32− + 2H+

Dissolution of carbonate rock

Addition of CO2 to a solution in contact


with a solid can (over time) affect the
alkalinity, especially for carbonate
minerals in contact with groundwater or
seawater . The dissolution (or
precipitation) of carbonate rock has a
strong influence on the alkalinity. This is
because carbonate rock is composed of
CaCO3 and its dissociation will add Ca+2
and CO3−2 into solution. Ca+2 will not
influence alkalinity, but CO3−2 will increase
alkalinity by 2 units. Increased dissolution
of carbonate rock by acidification from
acid rain and mining has contributed to
increased alkalinity concentrations in
some major rivers throughout the Eastern
U.S.[5] The following reaction shows how
acid rain, containing sulfuric acid, can have
the effect of increasing river alkalinity by
increasing the amount of bicarbonate ion:
2CaCO3 + H2SO4 → 2Ca+2 + 2HCO3− +
SO4−2

Another way of writing this is:

CaCO3 + H+ ⇌ Ca+2 + HCO3−

The lower the pH, the higher the


concentration of bicarbonate will be. This
shows how a lower pH can lead to higher
alkalinity if the amount of bicarbonate
produced is greater than the amount of H+
remaining after the reaction. This is the
case since the amount of acid in the
rainwater is low. If this alkaline
groundwater later comes into contact with
the atmosphere, it can lose CO2,
precipitate carbonate, and thereby become
less alkaline again. When carbonate
minerals, water, and the atmosphere are all
in equilibrium, the reversible reaction

CaCO3 + 2H+ ⇌ Ca+2 + CO2 + H2O

shows that pH will be related to calcium


ion concentration, with lower pH going
with higher calcium ion concentration. In
this case, the higher the pH, the more
bicarbonate and carbonate ion there will
be, in contrast to the paradoxical situation
described above, where one does not have
equilibrium with the atmosphere.

Oceanic alkalinity
Processes that increase
alkalinity

There are many methods of alkalinity


generation in the ocean. Perhaps the most
well known is the dissolution of CaCO3
(calcium carbonate, which is a component
of coral reefs) to form Ca2+ and CO32−
(carbonate). The carbonate ion has the
potential to absorb two hydrogen ions.
Therefore, it causes a net increase in
ocean alkalinity. Calcium carbonate
dissolution is an indirect result of ocean
pH lowering. It can cause great damage to
coral reef ecosystems, but has a relatively
low effect on the total alkalinity (AT) in the
ocean.[9] Lowering of pH due to absorption
of CO2 actually raises the alkalinity by
causing dissolution of carbonates.

Anaerobic degradation processes, such as


denitrification and sulfate reduction, have
a much greater impact on oceanic
alkalinity. Denitrification and sulfate
reduction occur in the deep ocean, where
there is an absence of oxygen. Both of
these processes consume hydrogen ions
and releases quasi-inert gases (N2 or H2S),
which eventually escape into the
atmosphere. This consumption of H+
increases the alkalinity. It has been
estimated that anaerobic degradation
could be as much as 60% of the total
oceanic alkalinity.[9]

Processes that decrease


alkalinity

Anaerobic processes generally increase


alkalinity. Conversely, aerobic degradation
can decrease AT. This process occurs in
portions of the ocean where oxygen is
present (surface waters). It results in
dissolved organic matter and the
production of hydrogen ions.[9] An
increase in H+ clearly decreases alkalinity.
However, the dissolved organic matter
may have base functional groups that can
consume these hydrogen ions and negate
their effect on alkalinity. Therefore, aerobic
degradation has a relatively low impact on
the overall oceanic alkalinity.[10]

All of these aforementioned methods are


chemical processes. However, physical
processes can also serve to affect AT. The
melting of polar ice caps is a growing
concern that can serve to decrease
oceanic alkalinity. If the ice were to melt,
then the overall volume of the ocean
would increase. Because alkalinity is a
concentration value (mol/L), increasing
the volume would theoretically serve to
decrease AT. However, the actual effect
would be much more complicated than
this.[11]

Global temporal variability

Researchers have shown oceanic alkalinity


to vary over time. Because AT is calculated
from the ions in the ocean, a change in the
chemical composition would alter
alkalinity. One way this can occur is
through ocean acidification. However,
oceanic alkalinity is relatively stable, so
significant changes can only occur over
long time scales (i.e. hundreds to
thousands of years).[12] As a result,
seasonal and annual variability is generally
very low.[9]

Spatial variability

Researchers have also shown alkalinity to


vary depending on location. Local AT can
be affected by two main mixing patterns:
current and river. Current dominated
mixing occurs close to the shore in areas
with strong water flow. In these areas,
alkalinity trends follow current and have a
segmented relationship with salinity.[13]

River dominated mixing also occurs close


to the shore; it is strongest close to the
mouth of a large river (i.e. the Mississippi
or Amazon). Here, the rivers can act as
either a source or a sink of alkalinity. AT
follows the outflow of the river and has a
linear relationship with salinity. This
mixing pattern is most important in late
winter and spring, because snowmelt
increases the river’s outflow. As the
season progresses into summer, river
processes are less significant, and current
mixing can become the dominant
process.[9]

Oceanic alkalinity also follows general


trends based on latitude and depth. It has
been shown that AT is often inversely
proportional to sea surface temperature
(SST). Therefore, it generally increases
with high latitudes and depths. As a result,
upwelling areas (where water from the
deep ocean is pushed to the surface) also
have higher alkalinity values.[14]

Measurement data sets

Throughout recent history, there have been


many attempts to measure, record, and
study oceanic alkalinity. Some of the larger
data sets are listed below.

GEOSECS (Geochemical Ocean Sections


Study)
TTO/NAS (Transient Tracers in the
Ocean/North Atlantic Study)
JGOFS (Joint Global Ocean Flux Study)
WOCE (World Ocean Circulation
Experiment)
CARINA (Carbon dioxide in the Atlantic
Ocean)

See also
Alkali soils
Base (chemistry)
Biological pump
Dealkalization of water
Global Ocean Data Analysis Project
Ocean acidification
References
1. "the definition of alkali" .
www.dictionary.com. Retrieved
2018-09-30.
2. "What is Alkalinity?" . Water Research
Center. 2014. Retrieved 5 February
2018.
3. Dickson, Andrew G. (1992). "The
development of the alkalinity concept
in marine chemistry". Marine
Chemistry, 40, 1: 49–63.
doi:10.1016/0304-4203(92)90047-E .
4. "Total Alkalinity" . United States
Environment Protection Agency.
Retrieved 6 March 2013.
5. Kaushal, S. S.; Likens, G. E.; Utz, R. M.;
Pace, M. L.; Grese, M.; Yepsen, M.
(2013). "Increased river alkalinization
in the Eastern U.S". Environmental
Science & Technology:
130724203606002.
doi:10.1021/es401046s .
6. Drever, James I. (1988). The
Geochemistry of Natural Waters,
Second Edition. Englewood Cliffs, NJ:
Prentice Hall. pp. 51–58 [52]. ISBN 0-
13-351396-3.
7. Stumm, W. & J.J Morgan (1981).
Aquatic Chemistry, 2n Ed. New York:
Wiley-Interscience. p. 780.
8. Benjamin. Mark M. 2015. Water
Chemistry. 2nd Ed. Long Grove, Illinois:
Waveland Press, Inc.
9. Thomas, H.; Schiettecatte, L.-S.; et al.
Enhanced Ocean Carbon Storage from
Anaerobic Alkalinity Generation in
Coastal Sediments. Biogeosciences
Discussions. 2008, 5, 3575-3591
10. Kim, H.-C., and K. Lee (2009),
Significant contribution of dissolved
organic matter to seawater alkalinity,
Geophys. Res. Lett., 36, L20603,
doi:10.1029/2009GL040271
11. Chen, B.; Cai, W. Using Alkalinity to
Separate the Inputs of Ice-Melting and
River in the Western Arctic Ocean.
Proceedings from the 2010 AGU
Ocean Sciences Meeting, 2010, 22-26.
12. Doney, S. C.; Fabry, V. J.; et al. Ocean
Acidification: The Other CO2 Problem.
Annu. Rev. Mar. Sci., 2009, 69-92.
doi:10.1146/annurev.marine.010908.1
63834
13. Cai, W.-J.; Hu, X. et al. Alkalinity
Distribution in the Western North
Atlantic Ocean Margins. Journal of
Geophysical Research. 2010, 115, 1-
15. doi:10.1029/2009JC005482
14. Millero, F. J.; Lee, K.; Roche, M.
Distribution of alkalinity in the surface
waters of the major oceans. Marine
Chemistry. 1998, 60, 111-130.

External links
Holmes-Farley, Randy. "Chemistry and
the Aquarium: What is Alkalinity? ,"
Advanced Aquarist's Online Magazine.
Alkalinity as it pertains to salt-water
aquariums.
DOE (1994) "[2] ,"Handbook of methods
for the analysis of the various parameters
of the carbon dioxide system in sea
water. Version 2, A. G. Dickson & C.
Goyet, eds. ORNL/CDIAC-74.
GEOSECS data set [3]
JGOFS data set [4]
WOCE data set [5]
CARINA data set [6]

Carbonate system calculators

The following packages calculate the state


of the carbonate system in seawater
(including pH):

CO2SYS , available as a stand-alone


executable, Excel spreadsheet, or
MATLAB script.
seacarb , a R package for Windows,
Mac OS X and Linux (also available
here )
CSYS , a Matlab script

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Alkalinity&oldid=897116432"

Last edited 3 months ago by Interne…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

Das könnte Ihnen auch gefallen