You are on page 1of 28

Reduction potential

This article relies largely or entirely on a single


source.
Learn more

Reduction potential (also known as redox


potential, oxidation / reduction potential,
ORP, pe, ε, or ) is a measure of the
tendency of a chemical species to acquire
electrons and thereby be reduced.
Reduction potential is measured in volts
(V), or millivolts (mV). Each species has its
own intrinsic reduction potential; the more
positive the potential, the greater the
species' affinity for electrons and tendency
to be reduced. ORP is a common
measurement for water quality.[1]

Measurement and
interpretation
In aqueous solutions, reduction potential
is a measure of the tendency of the
solution to either gain or lose electrons
when it is subjected to change by
introduction of a new species. A solution
with a higher (more positive) reduction
potential than the new species will have a
tendency to gain electrons from the new
species (i.e. to be reduced by oxidizing the
new species) and a solution with a lower
(more negative) reduction potential will
have a tendency to lose electrons to the
new species (i.e. to be oxidized by
reducing the new species). Because the
absolute potentials are difficult to
accurately measure, reduction potentials
are defined relative to a reference
electrode. Reduction potentials of
aqueous solutions are determined by
measuring the potential difference
between an inert sensing electrode in
contact with the solution and a stable
reference electrode connected to the
solution by a salt bridge.[2]
The sensing electrode acts as a platform
for electron transfer to or from the
reference half cell. It is typically platinum,
although gold and graphite can be used as
well. The reference half cell consists of a
redox standard of known potential. The
standard hydrogen electrode (SHE) is the
reference from which all standard redox
potentials are determined and has been
assigned an arbitrary half cell potential of
0.0 mV. However, it is fragile and
impractical for routine laboratory use.
Therefore, other more stable reference
electrodes such as silver chloride and
saturated calomel (SCE) are commonly
used because of their more reliable
performance.

Although measurement of the reduction


potential in aqueous solutions is relatively
straightforward, many factors limit its
interpretation, such as effects of solution
temperature and pH, irreversible reactions,
slow electrode kinetics, non-equilibrium,
presence of multiple redox couples,
electrode poisoning, small exchange
currents and inert redox couples.
Consequently, practical measurements
seldom correlate with calculated values.
Nevertheless, reduction potential
measurement has proven useful as an
analytical tool in monitoring changes in a
system rather than determining their
absolute value (e.g. process control and
titrations).

Explanation
Just as the transfer of hydrogen ions
between chemical species determines the
pH of an aqueous solution, the transfer of
electrons between chemical species
determines the reduction potential of an
aqueous solution. Like pH, the reduction
potential represents how strongly
electrons are transferred to or from
species in solution. It does not
characterize the amount of electrons
available for oxidation or reduction, in
much the same way that pH does not
characterize the buffering capacity.

In fact, it is possible to define pe, the


negative logarithm of electron
concentration (-log[e]) in a solution, which
will be directly proportional to the redox
potential.[2][3] Sometimes pe is used as a
unit of reduction potential instead of ,
for example in environmental chemistry.[2]
If we normalize pe of hydrogen to zero, we
will have the relation pe=16.9 at room
temperature. This point of view is useful
for understanding redox potential,
although the transfer of electrons, rather
than the absolute concentration of free
electrons in thermal equilibrium, is how
one usually thinks of redox potential.
Theoretically, however, the two approaches
are equivalent.

Conversely, one could define a potential


corresponding to pH as a potential
difference between a solute and pH
neutral water, separated by porous
membrane (that is permeable to hydrogen
ions). Such potential differences actually
do occur from differences in acidity on
biological membranes. This potential
(where pH neutral water is set to 0 V) is
analogous with redox potential (where
standardized hydrogen solution is set to 0
V), but instead of hydrogen ions, electrons
are transferred across in the redox case.
Both pH and redox potentials are
properties of solutions, not of elements or
chemical compounds per se, and depend
on concentrations, temperature etc.

Standard reduction potential


The standard reduction potential ( ) is
measured under standard conditions: 25
°C, a 1 activity for each ion participating in
the reaction, a partial pressure of 1 bar for
each gas that is part of the reaction, and
metals in their pure state. The standard
reduction potential is defined relative to a
standard hydrogen electrode (SHE)
reference electrode, which is arbitrarily
given a potential of 0.00 V. However,
because these can also be referred to as
"redox potentials", the terms "reduction
potentials" and "oxidation potentials" are
preferred by the IUPAC. The two may be
explicitly distinguished in symbols as
and .

Half cells
The relative reactivities of different half
cells can be compared to predict the
direction of electron flow. A higher
means there is a greater tendency for
reduction to occur, while a lower one
means there is a greater tendency for
oxidation to occur.

Any system or environment that accepts


electrons from a normal hydrogen
electrode is a half cell that is defined as
having a positive redox potential; any
system donating electrons to the hydrogen
electrode is defined as having a negative
redox potential. is measured in
millivolts (mV). A high positive
indicates an environment that favors
oxidation reaction such as free oxygen. A
low negative indicates a strong
reducing environment, such as free
metals.

Sometimes when electrolysis is carried


out in an aqueous solution, water, rather
than the solute, is oxidized or reduced. For
example, if an aqueous solution of NaCl is
electrolyzed, water may be reduced at the
cathode to produce H2(g) and OH− ions,
instead of Na+ being reduced to Na(s), as
occurs in the absence of water. It is the
reduction potential of each species
present that will determine which species
will be oxidized or reduced.
Absolute reduction potentials can be
determined if we find the actual potential
between electrode and electrolyte for any
one reaction. Surface polarization
interferes with measurements, but various
sources give an estimated potential for the
standard hydrogen electrode of 4.4 V to
4.6 V (the electrolyte being positive.)

Half-cell equations can be combined if one


is reversed to an oxidation in a manner
that cancels out the electrons to obtain an
equation without electrons in it.

Nernst equation
The and pH of a solution are related.
For a half cell equation, conventionally
written as reduction (electrons on the left
side):

The half cell standard potential is


given by:

where   is the standard Gibbs free


energy change, n is the number of
electrons involved, and F is Faraday's
constant. The Nernst equation relates pH
and :
 

where curly brackets indicate activities


and exponents are shown in the
conventional manner. This equation is the
equation of a straight line for as a
function of pH with a slope of
  volt (pH has no units.)
This equation predicts lower at higher
pH values. This is observed for reduction
of O2 to OH− and for reduction of H+ to H2.
If H+ were on the opposite side of the
equation from H+, the slope of the line
would be reversed (higher at higher
pH). An example of that would be the
formation of magnetite (Fe3O4) from HFeO
− [4]
2 (aq) :

3 HFeO−2 + H+ = Fe3O4 + 2 H2O + 2 [[e−]]

where
Eh = −1.1819 − 0.0885 log([HFeO−2]3) + 0.0296
. Note that the slope of the line is −1/2 the
−0.05916 value above, since h/n = −1/2.

Biochemistry
Many enzymatic reactions are oxidation-
reduction reactions in which one
compound is oxidized and another
compound is reduced. The ability of an
organism to carry out oxidation-reduction
reactions depends on the oxidation-
reduction state of the environment, or its
reduction potential ( ).

Strictly aerobic microorganisms are


generally active at positive values,
whereas strict anaerobes are generally
active at negative values. Redox
affects the solubility of nutrients,
especially metal ions.[5]

There are organisms that can adjust their


metabolism to their environment, such as
facultative anaerobes. Facultative
anaerobes can be active at positive Eh
values, and at negative Eh values in the
presence of oxygen bearing inorganic
compounds, such as nitrates and sulfates.

Environmental chemistry
In the field of environmental chemistry, the
reduction potential is used to determine if
oxidizing or reducing conditions are
prevalent in water or soil, and to predict
the states of different chemical species in
the water, such as dissolved metals. pe
values in water range from -12 to 25; the
levels where the water itself becomes
reduced or oxidized, respectively.[2]

The reduction potentials in natural


systems often lie comparatively near one
of the boundaries of the stability region of
water. Aerated surface water, rivers, lakes,
oceans, rainwater and acid mine water,
usually have oxidizing conditions (positive
potentials). In places with limitations in air
supply, such as submerged soils, swamps
and marine sediments, reducing
conditions (negative potentials) are the
norm. Intermediate values are rare and
usually a temporary condition found in
systems moving to higher or lower pe
values.[2]

In environmental situations, it is common


to have complex non-equilibrium
conditions between a large number of
species, meaning that it is often not
possible to make accurate and precise
measurements of the reduction potential.
However, it is usually possible to obtain an
approximate value and define the
conditions as being in the oxidizing or
reducing regime.[2]

In the soil there are two main redox


constituents: 1) anorganic redox systems
(mainly ox/red compounds of Fe and Mn)
and measurement in water extracts; 2)
natural soil samples with all microbial and
root components and measurement by
direct method [Husson O. et al.: Practical
improvements in soil redox potential ( )
measurement for characterisation of soil
properties. Application for comparison of
conventional and conservation agriculture
cropping systems. Anal. Chim. Acta 906
(2016): 98-109].

Water quality
Oxidation reduction potential (ORP) can be
used for water system monitoring with the
benefit of a single-value measure of the
disinfection potential, showing the activity
of the disinfectant rather than the applied
dose.[1] For example, E. coli, Salmonella,
Listeria and other pathogens have survival
times of under 30 s when the ORP is above
665 mV, compared against >300 s when it
is below 485 mV.[1]

A study was conducted comparing


traditional parts per million chlorination
reading and ORP in Hennepin County,
Minnesota. The results of this study argue
for the inclusion of ORP above 650mV in
local health codes.[6]

Geology
Eh-pH (Pourbaix) diagrams are commonly
used in mining and geology for
assessment of the stability fields of
minerals and dissolved species. Under
conditions where a mineral (solid) phase is
the most stable form of an element, these
diagrams show that mineral. As with
results from all thermodynamic
(equilibrium) evaluations, these diagrams
should be used with caution. Although the
formation of a mineral or its dissolution
may be predicted to occur under a set of
conditions, the process may be negligible
because its rate is so slow. Under those
circumstances, kinetic evaluations are
necessary. However, the equilibrium
conditions can be used to evaluate the
direction of spontaneous changes and the
magnitude of the driving force behind
them.
See also
Galvanic cell
Electrolytic cell
Electromotive force
Electrochemical potential
Standard electrode potential
Solvated electron
Table of standard electrode
potentials
Oxygen radical absorbance capacity
Redox

Weblinks
Online Calculator Redoxpotential
(„Redox Compensation”)

References
1. Trevor V. Suslow, 2004. Oxidation-
Reduction Potential for Water Disinfection
Monitoring, Control, and Documentation,
University of California Davis,
http://anrcatalog.ucdavis.edu/pdf/8149.pd
f
2. vanLoon, Gary; Duffy, Stephen (2011).
Environmental Chemistry -(*Gary Wallace) a
global perspective (3rd ed.). Oxford
University Press. pp. 235–248. ISBN 978-0-
19-922886-7.
3. 1981 Stumm, W. and Morgan, J. J.
(1981): Aquatic chemistry, 2nd Ed.; John
Wiley & Sons, New York
4. Garrels, R.M.; Christ, C.L. (1990).
Minerals, Solutions, and Equilibria. London:
Jones and Bartlett.
5. Chuan, M.; Liu, G. Shu. J (1996).
"Solubility of heavy metals in a
contaminated soil: Effects of redox
potential and pH". Water, Air, & Soil
Pollution. 90: 543–556.
Bibcode:1996WASP...90..543C .
doi:10.1007/BF00282668 .
6. Bastian, Tiana; Brondum, Jack (2009).
"Do Traditional Measures of Water Quality in
Swimming Pools and Spas Correspond with
Beneficial Oxidation Reduction Potential?" .
Public Health Rep. 124: 255–61.
PMC 2646482 . PMID 19320367 .
This article needs additional citations for
verification.
Learn more

Additional notes
Onishi, j; Kondo W; Uchiyama Y (1960).
"Preliminary report on the oxidation-
reduction potential obtained on surfaces
of gingiva and tongue and in interdental
space". Bull Tokyo Med Dent Univ (7): 161.

External links
Redox potential definition
Large table of potentials (Site broken.
Archived version on the Internet
Archive.)

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Reduction_potential&oldid=879799279"

Last edited 2 months ago by Syrenk…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like