Anda di halaman 1dari 36

The internal structure of atoms

Our concept of the internal structure of atoms has evolved dramatically since
Rutherford first demonstrated the existence of atomic nuclei. However, it will be
sufficient for our purposes to adopt a fairly simple view of atoms.
It is now known that the nucleus contains a large number of different
elementary particles which interact with each other and which are organized
into complex patterns within the nucleus. We need only be concerned with two
of these, the proton and neutron, collectively known as nucleons.
Protons and neutrons are the main building blocks of the nucleus because they
account for its mass and electrical charge. A proton is a particle having a
positive charge that is equal in magnitude but opposite in polarity to the charge
of an electron. Neutrons have a very slightly larger mass than protons and
carry no electrical charge. Extranuclear neutrons are unstable and decay
spontaneously to form protons and electrons with a "half-life" of 10.6 minutes.
The only other components of atoms of interest to us are the electrons that
swarm around the nucleus. Electrons at rest have a very small mass (1/1836.1
that of hydrogen atoms) and a negative electrical charge.
In a neutral atom the number of extranuclear electrons is equal to the number
of protons. The protons in the nucleus of an atom therefore determine how
many electrons that an atom can have when it is electrically neutral. The
number of electrons and their distribution about the nucleus in turn determine
the chemical properties of that atom.
Nuclear systematics
The number of protons (Z) is called the "atomic number" and the number of
neutrons (N) is the "neutron number".

The atomic number Z also indicates the number of extranuclear electrons in a
neutral atom.
The sum of protons and neutrons in the nucleus of an atom is the "mass
number" (A). We can therefore represent the composition of the nuclei of
atoms by means of a simple relationship:

A = Z + N
The composition of atoms is conveniently described by specifying the number
of protons and neutrons that are present in the nucleus.
Another word for atom that is widely used is "nuclide". Now that we have
defined A, Z, and N, we can specify the composition of any nuclide by means of
a shorthand notation consisting of the chemical symbol of the element, the
mass number written as a superscript, and the atomic number written as a
subscript.
For example,
14
6
C identifies the nuclide as an atom of carbon having six protons
(therefore 6 electrons in a neutral atom) and a total of 14 nucleons.
We can calculate that the nucleus of this nuclide contains 14 - 6 = 8 neutrons.
Similarly, is a sodium atom having 11 protons and 23 - 11 = 12 neutrons.
Because it is redundant to specify Z when the chemical symbol is used, the
subscript is generally omitted in informal usage.
In this course, as in the stable isotope business in general we will never use the
proton number subscript. A great deal of information about nuclides can be
shown on a diagram in which each nuclide is represented by a square in the
ordinates of Z and N.
We are now in a position to define several additional terms. Referring to the
chart of the nuclides, we see that each element having a particular atomic
number Z is represented by several atoms arranged in a horizontal row having
different neutron numbers. Such atoms, which have the same Z but different
values of N, are called "isotopes".
Because they have the same Z, isotopes are atoms of the same chemical
element.

They have very similar chemical properties and differ only in their masses.
Nuclides, which occupy vertical columns on the chart of the nuclides, are called
"isotones". They have the same value of N but different values of Z. Isotones
are therefore atoms of different elements.
Isotopes are therefore defined as atoms whose nuclei contain the same
number of protons but a different number of neutrons. The term
isotopes is derived from Greek (meaning equal places) and indicates that
isotopes occupy the same position in the Periodic Table. Isotopes can be
divided into stable and unstable (radioactive) species. The number of stable
isotopes is about 300; while over 1200 unstable ones have been discovered so
far. There will be moremark my words!
The term stable is relative, depending on the detection limits of radioactive
decay times. In the range of atomic numbers from 1 (H) to 83 (Bi), stable
nuclides of all masses except 5 and 8 are known.
Only 21 elements are pure elements, in the sense that they have only one
stable isotope. All other elements are mixtures of at least two isotopes.
In some elements, the different isotope may be present in substantial
proportions. In copper, for example,
63
Cu accounts for 69% and
65
Cu accounts
for 31%. In most cases one isotope is predominant, the others being present
only in trace amounts.
http://www2.bnl.gov/ton/
6-carbon-12
------------------------------------------------------------------------
* Atomic Mass: 12.0000000 +- 0.0000000 amu
* Excess Mass: 0.000 +- 0.000 keV
* Binding Energy: 92161.753 +- 0.014 keV
* Beta Decay Energy: B- -17338.083 +- 1.000 keV
"The 1995 update to the atomic mass evaluation" by G.Audi and A.H.Wapstra, Nuclear Physics A595 vol. 4 p.409-480, December 25,
1995.
------------------------------------------------------------------------
* Atomic Percent Abundance: 98.89%
* Spin: 0+
* Stable Isotope
Possible parent nuclides:
Beta from B-12
Electron capture from N-12
------------------------------------------------------------------------
6-carbon-13
------------------------------------------------------------------------
* Atomic Mass: 13.0033548 +- 0.0000000 amu
* Excess Mass: 3125.011 +- 0.001 keV
* Binding Energy: 97108.065 +- 0.016 keV
* Beta Decay Energy: B- -2220.445 +- 0.270 keV
"The 1995 update to the atomic mass evaluation" by G.Audi and A.H.Wapstra, Nuclear Physics A595 vol. 4 p.409-480, December 25,
1995.
------------------------------------------------------------------------
* Atomic Percent Abundance: 1.11%
* Spin: 1/2-
* Stable Isotope
Possible parent nuclides:
Beta from B-13
Electron capture from N-13

R.R.Kinsey, et al.,The NUDAT/PCNUDAT Program for Nuclear Data,paper submitted to the 9 th International Symposium of Capture-
Gamma_raySpectroscopy and Related Topics, Budapest, Hungary, Octover 1996.Data extracted from NUDAT database (Jan.
14/1999)
The stability of nuclides is characterized by several important rules, two of
which are briefly discussed here.
1) Symmetry rule- in a stable nuclide with low atomic number, the number of
protons is approximately equal to the number of neutrons, or the neutron-
to-proton ratio, N/Z, is approximately equal to unity. In stable nuclei with
more than 20 protons or neutrons, the N/Z ratio is always greater than unity,
with a maximum value of about 1.5 for the heaviest stable nuclei. The
electrostatic Coulomb repulsion of the positively charged protons grows
rapidly with increasing Z. To maintain the stability in the nuclei, electrically
more neutral neutrons than protons are incorporated into the nucleus.
http://www.triumf.ca/safety/rpt/rpt_2/node3.html
In the beginning
In 1931 Harold Urey predicted on theoretical grounds that there should be a
difference in the vapor pressures of the isotopes of hydrogen. His interest in
hydrogen had been aroused by a suggestion of Birge and Menzel that it may
have naturally occurring isotopes.
Urey, working with Murphy and Brickwedde, promptly planned and carried out
an experiment to detect
2
H and
3
H by spectroscopic means in the residual
volume of gas produced by evaporating about 6 liters of liquid hydrogen.
The results immediately confirmed the presence of
2
H but
3
H was not found.
Urey named the newly discovered isotope "deuterium" because it has nearly
twice the mass of hydrogen. The specific reason for this was not yet known
because the existence of neutrons was not established until 1932. In 1934,
Harold Urey won the Nobel Prize for chemistry for his discovery of deuterium.
2) Oddo-Harkins rule- states that nuclides of even atomic numbers are more
abundant than those with odd numbers. The most common of the four possible
combinations is even-even, the least common odd-odd. The same relationship
shows that there are more stable isotopes with even than with odd proton
numbers.
During W.W.II, he used his knowledge of isotope fractionation to develop
methods for the separation of
235
U by gaseous diffusion.
When the war ended, he turned his attention to the possibility that the stable
isotopes of oxygen may be fractionated by natural processes. He suggested
that such fractionation might occur during the formation of calcium carbonate in
the oceans and that the extent of fractionation depends on the temperature. Out
of these ideas has developed the oxygen isotope method of measuring the
temperature of deposition of skeletal calcium carbonate.
The research inspired and led by Harold Urey has evolved into an important
branch of isotope geology that deals with the fractionation of the stable isotopes
by physical and chemical reactions occurring in nature. The group of elements
whose isotopes are especially susceptible to natural isotope fractionation
includes hydrogen, carbon, nitrogen, oxygen, and sulfur.
These are among the most abundant elements in the Earth, and they are
intimately associated with the biosphere, the hydrosphere, and the lithosphere.
Consequently, the study of fractionation of their isotopes provides information
on a great variety of important biological, geological, and meteorological
processes occurring in many different environments.
Beginning with the classic paper of Urey (1947), in which he calculated stable
isotope fractionation factors between species of geochemical interest, there has
been an increasing use of stable isotope variations in natural materials for
studies in the Earth and cosmic sciences.
Equilibrium fractionation factors that measure the distribution of a rare stable
isotope between two species have been determined directly by laboratory
experiments and also by calculations using the methods of statistical
thermodynamics. They have also been inferred from regularities in the stable
isotope ratios of natural materials.
These fractionation factors have been used in geochemistry, meteorology,
oceanography, water and aqueous chemistry, cosmochemistry, paleontology,
and other scientific fields for a variety of purposes.
In his 1947 paper, Harold Urey reviewed theoretical and experimental results
from studies on isotopes of various elements in a publication entitled
"Thermodynamic Properties of Isotopic Substances" wherein he concluded
that isotopes and their compounds have different thermodynamic properties
that vary primarily as a function of temperature.
As a result of the small but significant differences in the thermodynamic
properties of isotopes of the same element, Urey suggested that fractionations
of the isotopes of oxygen and carbon among water, CO
2
, and carbonate shell
material were sufficiently large as to be useful in discerning the CO
2
contents of
ancient atmospheres and the paleotemperatures of the oceans.
Not only did this suggestion generate a plethora of studies concerning possible
paleoclimates of the earth, but it also led to the birth of stable isotope
geochemistry, a discipline which is now an integral part of many studies in the
earth sciencesand my personal favorite...
Light stable isotope geochemistry
1) They have low atomic mass. Isotopic variations have now been found for
heavy elements like Cu, Sn, W, and Fe.
Light stable isotope geochemistry primarily involves accounting for variations in
the isotopic composition of H, C, N, O, Si and S in a variety of natural
substances. These elements are comprised of one very abundant isotope and
one or more minor isotopes the ratios of which vary differentially in natural
substances. The stable isotope ratios of these elements are particularly useful
in the earth sciences for the following reasons:
2) The relative mass difference between the rare (heavy) and abundant
isotope is large. For example compare the values of 12.5 and 8.3 percent for
the pairs
18
O-
16
O and
13
C-
12
C respectively, with the value of only 1.2 percent for
the
87
Sr-
86
Sr. The relative mass difference between D and H is almost 100 per
cent and hydrogen isotope fractionations are, accordingly, about 10 times larger
than those of the other elements of interest.
3) They form chemical bonds that have a high degree of covalent
character. Attesting to the importance of bond type to isotopic fractionation, the
48
Ca/
40
Ca ratio varies little in terrestrial rocks despite the large relative mass
difference (only D-H is larger) between the isotopes.
4) The abundance of the rare isotope is sufficiently high (tenths to a few
percent) to assure the ability to make precise determinations of the
isotopic ratio by mass spectrometry. Depending on the instrument used, the
analytical error of deuterium analyses is up to ten times larger than those of the
other elements because of the low abundance of deuterium (about 160ppm) in
nature.
5) Unlike radiogenic isotopes, the isotopic composition of these elements
in natural substances is not a function of time or the chemical behavior of
the parent element.
6) These elements are significant components of most rocks, minerals
and fluids as well as forming the basis of most forms of life.
7) The distribution of the isotopes of these elements among different
phases varies primarily as a function of temperature, but because of their
low atomic masses and large relative mass differences, mass-dependent
fractionations of isotopes of these elements among phases is much more
pronounced than those of heavier elements.
8) The occurrence of more than one oxidation state as with C, N, and S or
of very different types of bonds as in H-O, C-O or Si-O also enhances the
mass-dependent fractionation of isotopes; the isotopic compositions of most
cations in geological materials do not vary much, not only because of their
small relative mass differences, but also because they tend to occur in a limited
number of oxidation states and in similar atomic environments.
9) The great abundance of these elements in most substances coupled
with the ability to determine precise relative isotope ratios using gas-
source mass spectroscopy allows determination of the isotopic
composition of many geologically relevant materials.
Natural variations in isotopic ratios of terrestrial materials have been reported
for other light elements like Mg and K, but such variations usually turn out to be
laboratory artifacts.
The case of Mg is fairly straightforward. Aside from the fact that its bonds are
dominantly ionic in character, the same atomic environment (an octahedron of
oxygen almost always surrounds magnesium) in nature. SS effect?
Thus with little or no possibility of site preference in magnesium compounds,
conditions are not favorable for isotopic fractionation of this element in nature.
In any event, variations in stable isotope ratios of light elements other than the
seven mentioned are small in terrestrial substances. The reasons for this are
not completely understood and are only loosely discussed in terms of
characteristics such as those noted. These characteristics are only observed
and are not rigorously tied to theoretical principles.
(1) Semi-empirical calculations using spectroscopic data and the
methods of statistical mechanics.

(2) Laboratory calibration studies.

(3) Measurements of natural samples whose formation conditions
are well known or highly constrained.
Essential to the interpretation of natural variations of light stable
isotope ratios is knowledge of the magnitude and temperature
dependence of isotopic fractionation factors between the common
minerals and fluids. These fractionation factors are obtained in
three ways:
Kinetic and Equilibrium isotope effects
Kinetic isotope effects
Kinetic isotope effects are common both in nature and in the laboratory and
their magnitudes are comparable to and sometimes significantly larger than
those of equilibrium isotope effects.
Kinetic isotope effects are normally associated with fast, incomplete, or
unidirectional processes like evaporation, diffusion, and dissociation reactions.
The examples of diffusion and evaporation are explained by the different
translational velocities of isotopic molecules moving through a phase or across
a phase boundary. Kinetic theory tells us that the average kinetic energy (K.E.)
per molecule is the same for all ideal gases at a give temperature.
Consider the isotopic molecules
12
C
16
O and
12
C
18
O that have molecular weights
of 28 and 30, respectively. Solving the expression equating the kinetic energies
(K.E. = 1/2 Mv
2
) of both isotopic species, the ratio of velocities of the light to
heavy isotopic species is (30/28)
1/2
, or 1.034.
That is, regardless of T, the average velocity of
12
O
16
O molecules is 3.4 percent
greater than the average velocity of
12
C
18
O molecules in the same system. This
and other such velocity differences lead to isotopic fractionations in a variety of
ways.
Molecules with lower molecular mass can preferentially diffuse out
of a system and leave the reservoir enriched in the heavy isotope.
In the case of evaporation, the greater average translational
velocities of lighter molecules allows them to break through the
liquid surface preferentially, resulting in an isotopic fractionation
between vapor and liquid.
For example, the o
18
O value of water vapor above the ocean (o
18
O
= 0) is typically around -13, whereas at equilibrium the value
should only be about -9 depending on the temperature of
evaporation.
The magnitude of the isotopic fractionation reduces to the value of
the equilibrium fractionations as the vapor phase approaches
saturation or equilibrium vapor pressure. At that point the rates of
molecular transfer between liquid and vapor and between vapor
and liquid are equal. Condensation, on the other hand, is
dominantly an equilibrium process.
Molecules containing the heavy isotope are more stable and have
higher dissociation energies than those containing the light
isotope. Therefore it is easier to break such bonds as
12
C-H and
32
S-O than to break bonds like
13
C-H and
34
S-O.
If youre a bacterium, which molecule are you going to eat?
Kinetic isotope effects arising from these differences in
dissociation energies can be extremely large in dissociation and
bacterial reactions that occur in nature. While it is very important
to be aware of kinetic isotope effects, they are relatively rare in
high-temperature processes occurring on Earth.
On the other hand, transient processes can occur whereby
differing rates of isotopic exchange between coexisting minerals
themselves or between the minerals and an external fluid can
result in assemblages that are grossly out of isotopic equilibrium.
Such examples are not explained by kinetic isotope effects but
rather by a series of equilibrium isotope exchange reactions that
have not gone to completion.
Background terminology
Isotopes, ratios, deltas and permil ()

Stable isotopes are measured as the ratio of the two most
abundant isotopes of an element.
For oxygen it is the ratio of
18
O, with an abundance of 0.204%, to
16
O which represents 99.796% of oxygen. Therefore the
18
O/
16
O
ratio is about 0.00204.
Fractionation processes will modify this ratio in any given
compound containing oxygen, but these differences are seen in
the 5th or 6th decimal place.

Measuring the absolute isotope ratio or abundance is very difficult
and requires some expensive toys.
Doing this on a regular basis would be heinously difficult and lab
to lab comparisons would be a nightmare. So what labs do is
measure an apparent or relative ratio by gas source mass
spectrometry.
The apparent ratio differs from the true ratio due to operational
variation (called machine error, m). This variation m differs from
lab to lab and even day to day on one machine.
By measuring a known reference on the same mass spec at the
same time, we compare the sample to the reference.
Isotopic values are then expressed as the difference between the
measured ratios of the sample and reference over the measured
ratio of the reference. This allows us to cancel the error (m) and
present the ratio in delta notation:
o
18
O
sample
=
m(
18
O
16
O
)sample m(
18
O
16
O
)reference
m(
18
O
16
O
)reference
The o value may also be defined as follows:

where Rx=(D/H)x, (
13
C/
12
C)x, (
18
O/
16
O)x, (
34
S/
32
S)x, and so forth,
and Rstd is the corresponding ratio in a standard. Note that R is
always written as the ratio of the heavy (rare) isotope to the light
(common) isotope.

x o
=
x R

STD R
STD R
|
\


|
.
|
|
3
10
,
Because fractionation processes do not yield huge variations in
isotope values, o-values are expressed as the parts per thousand
or permil () difference from the reference. This yields the
equation:

o
18
O
sample
=
18
O
16
O
sample
18
O
16
O
ref erence
1






(

(
(
(
(
x1000VSMOW
VSMOW (Vienna Standard Mean Ocean Water) is the reference
used. A o- value that is positive, 20 indicates that the sample
has 20 or 2% more
18
O than the reference. Some people say
"enriched" by 20.

A sample that is lower than the reference by the same amount
would be indicated as o
18
O
sample
= -20VSMOW.
Over the years, stable isotope geochemists have developed a
certain uniformity in the presentation of their data. However, there
are still a few noteworthy differences. Some workers write
o(
18
O/
16
O), o(D/H), and so forth, whereas others write o
18
O and
oD, and so forth, the latter being more common.
In the earlier literature oD values were given in percent, but,
because so many laboratories now report both oD and o
18
O
values for the same substances (primarily water), it has become
standard practice to report both values in per mil to avoid
confusion. o
13
C values reported from most laboratories in the
Soviet Union (now Russia) are given in percent.
The water and carbonate standards have changed to Vienna
Standard Mean Ocean Water (VSMOW) for both oxygen and
hydrogen, a belemnite from the Cretaceous Pee Dee Formation,
South Carolina (VPDB) for carbon (and, sometimes for oxygen in
carbonates).
The standards were called Standard Mean Ocean Water (SMOW)
for both oxygen and hydrogen, a belemnite from the Cretaceous
Pee Dee Formation, South Carolina (PDB) for carbon (and,
sometimes for oxygen in carbonates), air (Air) for nitrogen, troilite
from the Canyon Diablo iron meteorite (CDT) for sulfur, NBS Boric
Acid Standard for boron, and the Caltech Rose Quartz Standard
(NBS-28) for silicon.
Stable isotope standards and measurement
NBS is now the National Institute of Standards and Technology (www.nist.gov)
IAEA United Nations International Atomic Energy Agency (www.iaea.or.at)

ISOGEOCHEM http://www.uvm.edu/~geology/geowww/isogeochem.html
Oxygen-18 and deuterium in water
In 1961 Harmon Craig proposed the Standard Mean Ocean Water (SMOW)
standard which was not really standard seawater, but a calculated seawater
value relative to NBS-1. You can read about the absolute isotope abundances
if youd like, but for our purposes you need to know that the SMOW standard
and today the VSMOW standard are defined as having o
18
O and oD values of
0.0.
For waters that are highly depleted in
18
O relative to VSMOW, a water standard
with much lower values is used. Standard Light Antarctic Precipitation, or SLAP
is defined as having a o
18
O value of -55.50VSMOW and oD value of -
428.0VSMOW.
Measurement of o
18
O and oD values is not quite straightforward. You can't just
pour the water sample into a mass spectrometer. Because water molecules
cling to the guts of machines, o
18
O values are determined by equilibrating water
with CO
2
and then comparing the CO
2
to a standard CO
2
.
For rapid equilibration the pH should be less than 4.5. The o
18
O value of the
water sample is derived by compensating for the equilibrium offset
(fractionation factor). The fractionation factor (o) is generally assumed to be
1.0412 (Friedman and O'Neil, 1977). Therefore the CO
2
in equilibrium with
water is enriched by 41.2. We'll go over the actual process of measuring o
18
O
values in the lab soon.

Measurement of deuterium can take place in three ways, measurement of
elemental hydrogen reduced with uranium or zinc, reduction in a Cr furnace, or
measurement of hydrogen that has been equilibrated with water using a
platinum catalyst.
The fractionation factor, o
The isotope fractionation factor between two substances, A and B, o
A-B
,
is defined as:
(5)

where Ra is the ratio of the heavy (rare) isotope to the light (common) stable
isotope in phase A, such as D/H,
11
B/
10
B,
13
C/
12
C,
18
O/
16
O,
30
Si/
28
Si or
34
S/
32
S. If
the isotopes are randomly distributed over all the positions in substance A and
B, o is related to the equilibrium constant, K, for isotope exchange reactions by

AB o
=
a R
b R
,

o =
1
n
K
where n is the number of atoms exchanged. Rather than determining the
absolute ratios, ca.
18
O/
16
O, in every phase, it is easier and more precise to
measure the difference in absolute ratios between two substances. At
equilibrium, o is related, to the very good approximation that the isotopes are
randomly distributed among all possible sites in the molecule, to the equilibrium
constant K for the isotope exchange reaction between the two substances
(Biegeleisen, 1955).
Different authors report equilibrium fractionation factors as o, lno, 10
3
lno, K,
lnK, c, and A (see appendix 1, ONeil, 1986). For simplicity, isotope exchange
reactions are usually written such that only one atom is exchanged. For
example, the oxygen isotope exchange reaction between CO
2
and water vapor
can be written:

1
2
16
C 2 O
+
2
18
H O

1
2
18
C 2 O
+
2
16
H O
The equilibrium constant for this reaction is

K =
1
2
18
C
2
O ( )
1
2
16
C
2
O ( )
2
16
H O ( )
2
18
H O ( )
This formalism is normally used in the calculation of fractionation factors from
spectroscopic and thermodynamic data. C
18
O
2
means that both oxygen atoms
in the molecule are
18
O. The equilibrium constants for these reactions are
written are equal to the fractionation factor:

K = o =
2 CO
18
O
16
O
|
\

|
.
|
2 H O
18
O
16
O
|
\

|
.
|
Values of o are normally very close to unity, typically 1.00X. Commonly, isotopic
fractionations are discussed in terms of the value of X, in per mil (per mil
fractionations). For example, the sulfur isotope fractionation factor between
ZnS and PbS at 200C is 1.0036. It is accepted parlance to state that at 200C
(1) the sphalerite-galena fractionation is 3.6 (or 3.6 per mil), or (2) sphalerite is
enriched in
34
S by 3.6 per mil relative to galena.
In terms of quantities actually measured in the laboratory (o values), this
expression becomes:

AB o
=
1+
A o
1,000
1+
B o
1,000
=
1,000+
A o
1,000+
B o
,
Values of o are usually close to unity, so that isotopic fractionations are
expressed as per mil fractionations. For example, o for oxygen isotopes
between quartz and water at 200C is 1.0110 so that the fractionation of oxyegn
isotopes between quartz and H
2
O is +11; that is, quartz is enriched in
18
O by
11 relative to water. Similarly, o for oxygen isotopes between water and
quartz at 200C is 0.9890 so that water is depleted in
18
O by 11 relative to
quartz.
As with any equilibrium constant, o is related to the energy of any exchange
reaction as lno. The energies involved in any reaction having an equilibrium
constant close to unity are small, i.e. lno (quartz-water) = 0.011 at 200C. For o
= 1.0X, 1000lno is approximately equal to X so that for the quartz-water
exchange reaction at 200C, 1000lno = +11. Thus, 1000lno is approximately
the per mil fractionation. From equation (6), defining A
A-B
as o
A
-o
B
:

1000lno
AB
=o
A
o
B
= A
AB
(6)
provided o
A-B
is within about 2 percent of unity. Therefore, the difference
between the o values of two coexisting phases is approximately equal to the per
mil fractionation.

10
3
lno and the A value
It is a useful mathematical fact that

3
10
ln(1.00X)~X
For the
34
S example mentioned above where a = 1.0036, 10
3
lno = 3.6. That is,
10
3
lno is the per mil fractionation. This logarithm function has added
theoretical and experimental significance. For perfect gases, lno varies as 1/T
-2

and 1/T
-1
in the high- and low-temperature limits, respectively (Bigeleisen and
Mayer, 1947). In addition, smooth and often linear curves have been found to
be obtained when 10
3
lno is plotted against 1/T
-2
for experimentally determined
fractionation factors between mineral pairs or mineral-water pairs.
The per mil fractionation, 10
3
lno, is then of prime importance in stable isotope
geochemistry. This quantity is very well approximated by the A value:

A
AB
= o
A
o
B
~ 10
3
lno
AB
That is, merely subtracting o values will be an excellent approximation to the
per mil fractionation and identical to it within the limits of analytical error for
values of both As and os, which are less than about 10.

Anda mungkin juga menyukai