Sunteți pe pagina 1din 33

APPENDIX

X-RAY DIFFRACTION AND SAMPLE PREPARATION

The following discussion is intended to summarize some of the

elementary aspects of X-ray diffraction (aRD) and sample treatment as

they are applied to the identification of clay minerals. The reader is

referred to the standard texts cited at the end of this section for a

more comprehensive treatment of the subject. When X-rays strike atoms,

they are more or less scattered in all directions -- each atom may be

regarded as acting like a point source of X-rays. XRD can be defined as

a regular array of atoms (a crystal) mutually reinforcing coherent

scattering. By measuring the geometry of the reinforced X-rays, we may

deduce the nature of the atomic arrangement. The reader is refered to

useful texts on methods by Jackson (1975), Elug and Alexander (1974),

and Wilson (1987), and to the compilation in Tables A-2 and A-3.

Crystallography

Crystals

A crystal may be defined as a solid made of atoms arranged in

a three-dimensional periodic pattern. Noncrystalline solids are said to

be amorphous (such as glass) and do not have a periodic arrangement.

The presence of periodicity is inferred from external morphology


(crystal faces), optical properties, and especially by XRD. The latter

is necessary for fine-grained materials such as clays for which optical

and morphological data may be difficult to obtain. The distinction

between crystalline and amorphous is not always a clear one when it

comes to very fine-grained materials. If the individual grains or

diffracting domains are only a few atomic layers thick, the diffraction

effects will be very minimal and the material would be called amorphous

even though the atomic arrangement is quite periodic; hence the term

sometimes used is X-ray amorphous. Many clay minerals have diffraction

properties that could be described as intermediate between those of

well-crystallized and amorphous materials. For example, they may have

periodicity in only one or two dimensions.

Lattices and Unit Cells

Crystals consist of a motif (pattEuT) which is periodically

repeated to form a 3-D arrangement. The motif can be a single atom, an

atomic group, or a complex molecule. In thinking about crystals, it is

useful to ignore the actual atoms or atomic groups (motif) and replace

them with a set of imaginary points which have a fixed relation in space
to the motif. In other words, each complex motif is replaced by a

point. The points are created by dividing space into three sets of

planes, the planes in each set being parallel and equally spaced. In

the two-dimensional analogy, two-dimensional space is divided by two

sets of lines (Fig. A.1) instead of three sets of planes. Intersections


TWO-DIMENSIONAL INTERATOMIC
SPACINGS

e o
3 s t e

e
Figure A.1

BRAGG'S LAW
n A. = 2d SINE

X-RAY X-RAY
SOURCE DETECTOR

Figure A.2
of lines form the points, or in the 3-D case, intersections of planes

form lines and intersections of these lines form points. The result is

a point lattice -- an array of points in space such that each point has

identical surroundings.

The planes divide space into a series of cells identical in

size, shape, and orientation. These cells are parallelepipeds (opposite

faces parallel and each face a parallelogram) and any one of them

represents the whole lattice and is called a unit cell. The size and

shape of the unit cell may be described by the lengths of its edges (a,

b, c) which are referred to as the crystallographic axes, and by the

angles between them (q48, ); these values constitute the lattice

parameters of the unit cell. Repetition of the unit cell by translation

along the three axial directions will build the entire lattice. More

than one unit cell can be chosen to represent the same lattice. Each

cell may have only one lattice point (1/4 of each of four corners),

which is called a primitive cell; or it may have two, three, or four

points (compound, or nonprimitive cells). Note that, although the

lattice points are commonly drawn at the corners of the cell, we could

just as well place one point in the center of each cell and have no

lattice points at the corners. In other words, lattices and unit cells

are, like the hours of the day, imaginary constructs of man to help him

comprehend the infinite variety of the universe. The unit cell for all

phyllosilicates typically is similar in its a and b dimensions (see Fig.

2.1) and has a variable c dimension.


Miller Indices

Many sets of par, equally spaced planes can be drawn


through a given point lattice (Fig. A-1). In practice, each set can

give rise to intensity maxima (selective reinforcement) in the XRD

measurements (pattern). These same planes typically constitute the

external faces of well-formed crystals. A notation has been developed

for describing each unique set of planes in relation to the

crystallographic axes (which serve as reference axes) and to the unit

cell. The system of notation is called Miller indices under which each

set of planes is designated by three numbers (e.g., 121) each of which

is essentially the reciprocal of the intercept of that plane on the

respective (a, b, c) crystallographic axes. The plane which cuts all

three axes (a, b, c) at their unit lengths (as defined by the unit cell

edges) is designated the unit plane and has a Miller index (M.I.) of

111. A plane which is parallel to the "C" crystallographic axis could

have an M. I . of 120, or 320, etc. The general expression for this plane

would be hk0. A face which is parallel to two axes and cuts the third

could have an M. I . of 010, 100, or 001. A plane which cuts all three

axes at different ratios has a general expression of hk0.

For each set of planes (hkl, etc.), there is a unique

interplane spacing called d. X-ray powder diffraction actually measures

d for all diffracting planes in the crystal, and for each intensity

maximum (peak) there is a specific associated M. I . (hkl, etc.) which


designates the specific set of planes involved. Assigning a Miller

index (set of planes) to each measured d (spacing) is called indexing, a

formerly laborious task now rendered relatively painless by modern

computers. For routine clay-mineral analysis, the spacings between

planes perpendicular to the "C" axis (001 planes) are the most

diagnostic. These are the planes which are parallel to the fundamental

phyllosilicate layer structure.

X-Ray Methods

Properties of X -Rays

X-rays are electromagnetic radiation of the same nature as

light but with a much shorter wavelength. The unit of measurement is

the angstrom () which is equal to 10-8 cm, but the approved

international unit (SI) is now the nanometer 070 which is equal to 10


0
A,. The latter unit has not been widely employed in the U. S., but we

can look forward to its increasing use. X-rays used in diffraction have

wavelengths in the 0.5-2.5 range, whereas visible light is in the

order of 6000 Á-

X-rays are produced when a beam of electrons hits matter of

any sort and the electrons are rapidly decelerated. The energy of the

electrons is transformed into electromagnetic radiation in the X-ray


wavelength region. X-rays are, therefore, produced by TV sets which

must be appropriately shielded by thick glass.

X-Ray Diffraction

The German physicist, Max Von Laue, is credited with the

discovery of X-ray diffraction in 1912, but the actual work was carried

out by two of his graduate students, Friedrich and Knipping. That same

year, two Englishmen (father and son) named Bragg devised the simple

mathematical expression for the optical conditions of diffraction which

bears their name - The Bragg Law. The following year, they used XRD to

solve the details of the crystal structures of several simple compounds.

The existence of X-ray diffraction not only provided a powerful tool for

atomic structure analysis, but also proved that X-rays were

electromagnetic radiation and could be dealt with as wave phenomena.

Scattering. When a beam of monochromatic X-rays strikes an atom, each

electron is induced to 'vibrate." Indeed, if the X-rays have the

correct energy, the electron will be vibrated right out of the atom,

which produces the phenomenon of fluorescence. The simplified net

effect is that each atom, composed of electrons all vibrating at the

same frequency, becomes a point-source of X-rays of the same frequency

as the X-rays in the beam causing the excitation. The new X-rays caused

by the vibrating atoms are called scattered X-rays, and if the

wavelength of the initial and scattered X-rays are the same, then the
phenomenon is called coherent scattering. The efficiency of scattering

by any atom is a direct function of the number of electrons in the atom,

and is designated the atomic scattering factor.

Diffraction. Diffraction results from the existence of certain phase

relations between scattered X-ray waves:

differences in path length lead to differences in phase;

the phase differences result in interference producing a change in

amplitude, but no change in wavelength.

If two rays are exactly one-half wavelength out of phase

(one-half wavelength path difference), then they are completely out of

phase and cancellation will occur. If the rays are exactly one

wavelength out of phase, then no cancellation will take place and

reinforcement will result. Therefore, two rays are completely in phase

and reinforce whenever their path lengths differ by zero or by an

integral number of wavelengths.

A diffracted beam may be defined as composed of a large number

of coherently scattered (all with same wavelength) rays mutually

reinforcing each other. Atoms scatter X-rays in all directions, but

where the atoms are arranged in a regular way (as in a crystal), in same

of these directions the scattered rays will be completely in phase so as


to reinforce each other and produce diffracted beams. Each atom becomes

a point source of coherently scattered X-rays which spread outward with

circular wave fronts. The scattered X-ray wavefronts constructively

recombine in several directions at angles to the oncoming beam to form

the first- and second-order beams.

This effect is analogous to diffraction of visible light from

a line grating, but diffraction of X-rays is from the "atomic grating"

of crystals. By measuring the angle of the diffracted beams and knowing

the wavelength of the X-rays, we can measure the distance between the

atoms in the raw. This is just the opposite of the use of a line

grating (e.g., in a spectrometer) where the line spacing is known and

wavelength is measured as a function of angle. Crystals of known atomic

spacing are used as gratings to analyze X-ray wavelengths

(characteristic spectra) in X-ray spectroscopy, or wavelength dispersive

X-ray fluoresence analysis.

The Bragg Law. We have seen haw diffraction can occur from a raw of

atoms; now we will examine a more general case, still using a

two-dimensional analogy, in which a beam of X-rays impinges at an angle

onto two rows of atoms representing planes in a lattice (Fig. A-2).

Unlike diffraction from a visible-light line grating, which takes place

on the grating surface, X-ray diffraction from crystals takes place

within a 3-D lattice.


For the geometry shown, the only diffracted beam formed is

that shown, which makes an angle O of "reflection" (really diffraction,

but the term reflection is often used; note that 0 is not measured

against the normal to the reflector, as in visible-light optics) which

is equal to the angle of incidence. We will demonstrate the validity of

the last sentence first with respect to one plane, and then for all

atoms in the crystal.

Rays in the top incident beam strike atoms in the first plane

of atoms (really a raw in our two-,dimensional analog) and are scattered

in all directions. The circular wave fronts, so produced, are not shown

but should by dynamically visualized by the reader. Only in the

directions shown on the right are the scattered rays totally in phase

and able to reinforce one another; this is because the difference in

their path length is zero or integral. The rays scattered by all the

atoms in the first plane in a direction parallel to the outgoing beam

are in phase and contribute to the diffracted beam. This will also be

true of all planes separately. The path difference between the upper

and lower rays is 2d sin O. Scattered rays will be completely in phase

and reinforce one another if the path difference is equal to an integral

number of wavelengths ( A = wavelength; "n" is any whole number), that

is if

n = 2d sin O.
This relation was first stated by W. L. Bragg and is known as

the Bragg Law or Bragg Equation. It states the conditions under which

diffraction will take place. The integer n is called the order of

reflection, and is any whole number (1,2, 3...n, such that sin e does

not exceed unity) and is equal to the path difference between rays

scattered from adjacent planes. For fixed values of A and d, there may

therefore be several angles of incidence (e) at which diffraction may

occur, corresponding to n= 1,2,3... Only one of these angles is shown

in Fig. A-2. At all angles of e other than those which satisfy the

Bragg equation, complete cancellation takes place. The diffracted beams

are very weak compared with the primary incident beam. Poorly

crystallized materials, those with extremely small (thin) crystal size,

and mixed-layer clays may give significant diffraction intensity at

angles of G not predicted by the Bragg equation.

Powder Diffraction. Most diffraction work on clay minerals is done with

powdered samples in a diffractometer, and several points need to be

mentioned with regard to this method. A schematic diagram of a

diffractometer is shown in Fig. A.3. The sample (above the word

"diffractometer") which in this case is an oriented aggregate of illite

placed so that the phyllosilicate layers are parallel to the sample

surface (perp. to page). A detector (proportional counter) is mounted

on an arm which pivots around the axis of the instrument (dot). The

sample and counter are mechanically linked so that rotation of the

counter through 2X° is automatically accompanied by rotation of the


rE0110.41111.1101.61..160,411111

CC

CL

Ci
2

A-12
MICA
PREFERENTIALLY ORIENTED

50 o 10° 15° 20° 25° 300 350


17.659/A 8.838 5.901 4.436 3.559 2.976 2.562
TVVO THETA-d SPACING
specimen through X°. This means that the angles of incidence and

reflection (0) will always be equal and will be one-half the diffraction

angle (20).

If we allow the motor to scan the counter and sample assembly

about the axis up to 35° 28, we will get a diffraction pattern as shown

in Fig. A.4. The peaks (or reflections) are all from the same set of

lattice planes, those which are parallel to the illite (mica) cleavage.

The 2:1 structural units are 1QA thick, and the peak at 8.8°20 shows

this value when d is determined with the Bragg Equation, setting n=1 as

it is the largest spacing (1.542i = X Cu Ea radiation).

n Ì`, = 2d sin 8

1 x 1.542 = 2d sin 4.4°


o
d=

This would normally be indexed at the 001 peak, but the 2M muscovite

unit cell is two 16, units thick, so this is correctly the 002 peak (201i

unit cell divided by 2). For simplicity, we will index this as the 001

peak. The peak at about 17.7°20 gives a 5i spacing on solution of the

Bragg equation, and is often referred to as "the mica 5X peak," but

actually it is the second order reflection (n=2) from the fundamental

la lattice spacing. That is


2 X 1.542 i= 2d sin 8.9°
o
d = 101.

0 0
The peak at SA is, therefore, the 002 peak and the peak at 3.333A is the

003 peak. The 004, 005, etc. peaks would be present at higher

28 angles.

Most clays are fine-grained with a paate-like morphology and,

as described in a subsequent section, they are typically prepared for

XRD by orienting all these little plates to form an oriented aggregate

whose diffraction properties much resemble those of a single mica

crystal.

For general powder diffraction work, the sample is prepared as

a packed, randomly oriented fine powder. Under these conditicns,

reflections corresponding to all possible lattice spacings appear on the

diffraction pattern. The way this works is as follows: First, it must

be noted that reflections (diffraction) will be measured by the counter

only for lattice planes which are parallel to the flat surface of the

packed-powder specimen. Other planes may diffract, but the counter will

not be in a proper position to measure them. This is elucidated in Fig.

A-5 in which exaggeratedly large powder grains are shown with two sets

of lattice planes. This sample should replace the illite in Fig. A-3.

Diffraction will never be measured from grains A, C, and G because none


X-RAY TUBE DETECTOR
Et TOWER

DIFF ACTOMETER

/ C DE F

- oZo ?Lanes
00% pt;hes
of their lattice planes are parallel to the surface. When the proper

20 angle is attained, so that the Bragg Equation is satisfied for the

020 d-spacing, grains B, E, and H will diffract. At some other angle,

grains D and F will diffract. It is important that the sample be ground

to a fine (generally <50 ».ni) powder in order to have a statistically

random orientation of all planes. In actuality, most crystals have many

more than just two sets of lattice planes.

Intensity and Geometry of Diffraction. The positions (20) of

reflections are controlled entirely by the dimensions and shape of the

unit cell, but the intensity of individual peaks is controlled by the

types of atoms present and their distribution within the cell. The
presence or absence of heavy atoms, which have many electrons and are

efficient scattering centers, may have a strong effect on the peak

intensities. This is illustrated in Fig. 2.14 and 2.15 in which it is

noted that the 002 peak for glauconite is much weaker (in some specimens

it is ,not there at all) than that of illite. Both have a nearly

identical structural spacing of 105,. The intensity difference is due to

the presence of iron in the octahedral position of the glauconite. In

fact, the ratio of 001/002, or 002/003 has been used to measure the iron

content of micas.
Sample Mounting and Pretreatments

Sample Mounting

Randomly Oriented Mounts. Clays are phyllosilicates with a tabular

habit and good basal cleavage and thus have a tendency to orient with

their 001 planes parallel to the mounting substrate. However, samples

can be mounted in either a randomly or preferentially oriented manner,

though, in practice, neither orientation is perfect. If a clay sample

is randomly oriented, all of the X-ray diffraction peaks for all

minerals will be observed unless peaks are not resolved or are of such

low intensity as not to be detected. The relative peak intensities,

however, may not be the same as theoretically predicted, since the

particles may exhibit some preferential orientation. The advantages to

analyzing a randomly oriented sample are given in Chapter 2. Random

mounts are prepared by packing a sample in the shallow cavity of an

aluminum, plastic, or glass holder. This is done with minimal pressure

and without smearing. Less satisfactorily, small amounts of clay

powder can be sprinkled onto grease or double-stick tape mounted on a

glass slide. Most minerals can be identified and characterized by

scanning the sample from 2°20 to 65°20 (CuKcx radiation). The random

mounting methods are considered to have a law sensitivity for clays but

high precision.
Preferentially Oriented Mounts. Preferentially orienting a clay

involves either sedimenting a suspension onto a flat surface, usually

glass; suction of the suspended material onto a flat, unglazed ceramic

tile, membrane filter, or porous metal surface; smearing a clay paste

onto a glass slide; or application of pressure and/or shear to a dry

powder (Brin and Brindley, 1980). The chosen method should provide a

specimen having the maximum amount of preferred orientation of basal

planes (sheets) parallel to the substrate, and which is sufficiently

thick and homogeneous. The clay cake must also be essentially smooth

and coplanar with the substrate surface to prevent sample displacement

errors which result in shifts of peak positions. To assure the accuracy

of calculated d-spacings, quartz or corundum internal standards may be

added, the former being naturally present in many samples.

According to Gibbs (1965), the smear method, the rapid-suction

method, and the pressure method are the only acceptable mounting methods

for preferentially orienting clays. Methods that involve sedimentation

give inhomogenous specimens because different mineral components have

different settling velocities. For example, clays that contain

typically fine-grained smectite often have this phase concentrated on

the surface layers of specimens prepared by sedimentation. This results

in an overemphasis of this phase during X-ray diffraction analysis.

However, the sedimentation method is the best for giving a high degree

of preferred orientation. The method selected will depend on the aims

of the work. Most clay and associated minerals mounted in a


preferential orientation can be identified by scanning form 2°20 to

35°20 (CuKa radiation). Extension of this range might be needed in

certain instances. Generally, the preferentially oriented mounts give

more sensitive data than do randomly oriented mounts. Mounting may also

require practice, particularly for the smear and suction methods. A

fairly easy, highly recommended method is the suction-on-ceramic-tile

method. Suction onto a "millipore" filter and transfer to a glass slide

is also recommended.

With all mounting methods, it is desirable to obtain an

"infinite thickness" of sample; i.e., the X-rays do not penetrate the

entire sample and diffract off of the substrate. An "infinite

thickness" can generally be achieved by mounting greater than 15 mg/cm2

of sample for analysis to 35°28, or greater than 25 mg/cm2 of sample for

analysis to 65°20.

Sample Pretreatments for X-Ray Diffraction

Once mounted, samples can be subjected to several

pretreatments prior to X-ray diffraction analysis (Fig. A.6). Table A.1

lists minimum diagnostic pretreatments required to qualitatively

identify clay mineral groups. The four common pretreatments (generally

for preferentially oriented samples) are cation saturating/air-drying,

solvating with ethylene glycol or glycerol, heating to 3500 C and/or

550°C, and intercalating (kaolinite) with dimethylsulfoxide (D4S0).


..)4 0.1
Qd
V) 0 t
"4 (U .13 O(1.1
C
Ca,
0 Ena rt
L. 1 in 1))
4C 4-,
14.

°
re

.0 Cu Ac E

I. Cr
C U00m. ,e1

(1)
o
0.1
.0 0
-lc 0 0
C Q.)

E
34
U 0 ,L4
1,3 /.
C U LI1

rta C C
0 0
:a
4
4..,
+I
1.4
4. fa 0 ,t PO
0 D% E "4 7N
--4. it -10. 113
I-4 I 0
C CI) X 4.
0 4...
(4
C
,13 a.; ,-.1 al
U
(..0 Q.) 0
ir)
Z.
A V)
QJ

tC
4.,) U)

-4.4 0 0 1/40
4-1
4J ..0 I C .,
O
%
to
.. L, ,..
I
I,-,
o- 4.,co..
I ) (..,

L
O)
ftz, CL)

'CL)
C....) I4.4 14... E
ea
In I )

4.)
4.)
4.J
C
-H
C QJ 4-4
= +I o.
'rt I

z0L4a ^o
Li) ia
QJ
QJ o
N o +4
-4 in 1.4 4) (...) n". Cri 4.J
u) '0 'Q ea ea 0 13 0 (13 a.) 0
'44
1.4 '.4
4-.
X 0 X rn
'-"RI (r) ca. 14-.
C QJ
4.1 '.4.0
1.4 Cad

o. 14
0
44.4(
4.1
QTI
'CI 41.1

(U OJ C
IC 4J 4C +4
q C (LI VI PC
it 0.) "I
"I Ci,
0U
q tPt "41
ra4
41) U)
"4 C,
(.) "4 PC1
Pt
c C
o e04)
o)

ea
-14.)
-, 4)
4)
lo
q qea0
-,
muce re

U --, 0 Cu
ea QJ ,C3
"'I
=
RI
61 - -,

---,
-4. E
le +4
4-, U
(J
U __,, rt?I LIPt
4 4Q.U in 0
CI) L1

fa ^4 I ,L.
Q..) LI C44 X
14.4
Other pretreatments include acid boiling to dissolve chlorite and

equilibration of the sample at various relative humidities to observe

clay swelling. These will not be discussed here.

Cation Saturating/Air Drying. Samples should be saturated with a

single, known cation prior to any pretreatment. The only exception is

for the study of clays saturated with their natural cations. Saturation

can be accomplished by flocculating a suspension of clay with the salt

of interest, washing the clay with this salt (by centrifuge) or, if

using a rapid-suction mounting technique, suction of a salt solution

through the sample, excess salt being removed by subsequent suction of

distilled water. Clays are typically studied after saturation with Et

or Mg2+, though at times analysis of the clays with their natural

exchange cations is desired. Smectite and vermiculite clays can show

dramatic differences in their diffraction patterns depending on

interlayer cation saturation and relative humidity (MacEwan and Wilson,

1980), though the latter effect may be difficult to assess without a

humidity-controlled diffractometer.

Ethylene Glycol and Glycerol Solvation. Solvation is generally done

after air drying and before heat treating. Not all clays will react to

solvation, but the swelling clays, particularly smectite and some

vermiculite, will expand to specific basal spacings depending somewhat

on their cation saturation (1,1Ewan and Wilson, 1980). Mg2+ saturation

is commonly, though not exclusively, used. This swelling allows


identification of these minerals. The method used for applying the

solvent to the sample depends on haw the clay is mounted. It can be

applied by dripping the solvent manually onto the oriented mount, by

placing the mount in the solvent vapor for an appropriate time, or by

soaking the clay suspension with the solvent and then making the

oriented mount. Ethylene glycol is generally easier to use than

glycerol, but may give different results (vlatEwan and Wilson, 1980).

Heating to 3500C or 550°C. Heating can selectively alter or destroy the

structure of clays, thus aiding in their identification (Jackson, 1979;

Brown and Brindley, 1980). Heating a sample at high temperature usually

requires equilibration of the sample at that temperature for at least

two hours. If cracking or peeling of the clay occurs (typical with

smectite), and the clay peels off as a thin film, it may be stuck back

onto its substrate with double-stick tape, but results are not

satisfactory in all cases. Some expandable clays may rehydrate and

partially re-expand when subjected to the atmosphere after heating,

particularly after 3500C. Heating stages for diffractometers are

available.

DMS0 (Dimethylsulfoxide) Intercalation. This treatment may be used to

expand and differentiate kaolinite from chlorite and serpentine by X-ray

diffraction when two or more of the phases exist in the same sample.

The best procedure is the CsC1-hydrazine-DMS0 method of Calvert (1984).

The method requires grinding the clay sample in CsCl, then reacting the
sample with hot hydrazine and hot DMSO. The method is useful, but not

routinely used.

Quantitative Analysis of Clay Minerals

A "quantitative" clay mineral analysis is typically a

"semi-quantitative" analysis; errors of 50 percent or more are typical.

These errors can be controlled, to a degree, by proper sample

preparation and analysis techniques (Brindley, 1980). X-ray diffraction

techniques using internal standards, measuring or controlling particle

orientation, and measuring peak areAs are much better than techniques

that measure peak height or area alone. Addition of other data, such as

chemical, surface area, thermal, and selective dissolution data, can

greatly improve the X-ray diffraction analysis, though this requires

much more time and sample.

In most instances, a rapid, visualj examination of peak

intensities will indicate relative abundances of clay mineral phases.

It should be remembered that peak areas are more representative of clay

mineral abundance than are peak heights. However, equal amounts of each

phase in an ethylene glycol-solvated sample, the smectite basal

reflection near 171 is often 3 to 5 times more intense (peak area) than

the 10A illite peak, and 2 to 3 times more intense than the 7A kaolinite

or chlorite peak. However, these mineral intensity factors depend on

experimental conditions, particularly on the geometry of the clay mount,


and on the crystallodhemical properties of the clay minerals. The

factors above were calculated from data of Roberts (1974) and Hallberg

(1978). Generally, quantitative data are almost always reported with

more significant figures than are justified.


2)
C -C
... o< ....:

0.0
C
.-vo 6.
E C 11r
--e 0
1:1

od EO -. o<
-,3

.
IJ ,
C ....
N
4.
CV

,
... ..0

ez - u
0
6.. ....
.... ,,,,,
.
..
0.<
rs.
...
rti
In .-.
C ,,.,
.
601
- ...
O

C --
..-.
....
*4
.C-O
=..
,..
....,.

t.-
CJ Ti
C

'
ef,

u..--.
. co O Cl -o C
...-) -8 r., a -5 E -o E
o >, .... X .. C cl, ,-.
(.2. i 0 U.1 -17. cl

0< O< V I,r


.0
O O -2)
E
o . _. r .
"V
0,) ,3 O

..- S.,
c. ,,,,,
-
o -. ,,
O 6.
)...
V
, - -3
co
.0
U
,.... 8 ..x ,.:
ro
...y,
C;
....,

,, ..> , .. o. o .,
.., ,..
(.; +
o
o
, 2-
.->. ,o
4) CU y
6,,
o. -
L
-E , -,o 7,
.. _.,t,
_

=O ,,
,,,, >. !.-, ,.., ,,,, " ti, o ., c, 15' '1

r
>.. ,-
,I,J

2Sc
a.,
73
U U .c-
o< 4c; 45 -O o
-7 cu
- .- o 2 'O 1...
Tr1
c; 0. CL
_. MI
c13
(.4 CL O.
_. T13 ,13

E -Y
lii v
_x 1-
Ri v
V (13 S.
yt. O. 8. 2:
,CI 11
O<
'C
r".1 r'3'..
...*:

I.)
..,...
c: E °- E
....

o<
- i u_ v
...'
- o, uE, -v.,
al c.< . 7..

V -,3
o ao c,,,,o- c ....
CE u. 8 -=-
o o . ...... -5 C , ..., t.,
vi

-.mi , - oo
*2c ro
C o 4.
,...., 'O A >. --,--. cs.,
C
. , ry cl:'
-,:r
' Eo ID .0
>. - vl
C.,
(13 T
)... .... CL.
c
.
3 10c E o 3 -o
ad ,.. b.. .
E f. . C 5 2 -. ,.,.....,
....f.
,... - ,-. ,13

E E
O< O<
+
N
2 ,,, o<
r s.
O<
I,
0<. 0<
..1 N
0'cr: 0 o<
O<
, o<-7
- O<
-
-7'
.#-

-6 r4

2i 11

- -.
4)
-.uV....7.
....

r-.
u

.
ar
7.
.,..
.... -3.
U
a.,
'3
...
I; >.
E 2
,..,,,,
o U O
ro ..... E Tii
> 1.1
U a.

A-26
Summary of Various Clay Mineral Analysis Techniques

Routine Turn Around Minimum


Minimum Time lor 10 Prepar a S ion
Analysis Required Techniques Available Sample Required Samples') Resumed References

Sample preparatiom Fractionation, dialysis, etc. Depends on % clay and Depends on preparas ion -- Jackson, 1979
analyses required, techniques used,
usually lg. usually =-1 day.

Particle size analysis --Quantitative fractionation - lg whole rock -- 1-2 days - dispersion - Jackson, 1979
- Pipette - 2g whole rock - 1 day = dispersion - Day, 1965
- Sedigraply - lg whole rock -41 day - dispersiori
Qualitative clay -- X-ray diffraction (XRD) - 50-100mg whole rock - 1 day Ir ar t unat ion for - Br indley Brown, 1980;
mineral anlaysis or clay clays, crushing la, kson, 1579
for whole rock. Chen, 1977

Quantitat ive clay = XRD - 50-100mg whole rock -- 1 day = Fractionation and -- ioriley, 1980
mineral analysis or clay dr ying for clays, Roher iis, 1974
' crushing lot
whole rock
= Elemental, thermal, - 400-500mg clay or , 3-4 days = Frac. & drying for -Hussey, 1972
surface & charge analyses ".3g whole rock clays, crushing for
whole rock
-- Selective dissolution - Var ies -- Varies - Varies -Jackson, 1979
Elemental and -- Sample dissolution -- 100mg whote rock or -- 2-3 days - Fractionation and -Calvert et al., 1980
thermal analysis clay drying for clays, Jackson, 1972;
crushing for whole Bernas, 1968
rock Roth et al., 1968
Energy dispersive X-ray -- Thin section or rock - 2 days -- Thin section -Goldstein et al., 1981
analysis or microprobe chunk preparat inn may
be required
-- Thermal analyses (DTA, -- 100mg whole rock or -- 2 days t Fractional ion and -Jackson, 1979
TGA, DSC, etc.) clay drying for clays, Van and Ilatek, 1979
crushing for MacKenzie, 1957
whole rock
Photomicrographic - Petrographic microscopy Thin section -- 2 days - Thin section -Heinrich, 1965
analysis preparation Kerr, 1959
-- Scanning electron micro- - Thin section or rock -- 1-2 days -- Thin section -Goldstein et al., 1981
scopy chunk Freparat ion may Mr Kee & Brown, 1977
be required
-- Transmission electron - Ultrathin section or -= )-2 days - Fractionation for -Sudo et al., 1981
microscopy 5mg clay clay, McKee & Blown, 1977
ultra-thin sect ion Gard, 1971
prep.cratoci may
he i equiced

Surface and charge. -- Cation exchange capacity -- 200mg clay or>2g whole -- 3 days -
Fractionation and - lackson, 1979
property analysis (CE C) rock drying for clays, Wor thingt on, 1973
crushing for Alex lades g lj kson, 1966
whole rock Chapman, 1965
-
, 'Surface area analysis by - -
- 500mg clay oe>2g whole - 2 days - Fractionation and - Greenland & Mot t, 1978
adsorption of polar and rock drying for c lays, Aylmore, 1974
nonpolar liquids and gases crushing for whole Carter et al., 1965
rock -

JJ -spical turnaround time for most laboratories after sample preparation fe-cept for 'sample preparation categor y itsel 1).

Table A.2

A-27
Selective Dissolution Techniques for Removal of Cements and Other Selected Phases

Method Reference(s) Comments

Dissolution of carbonates Jackson, 1979; Na0Ac is less damaging than HCI. though
with pH 5 buffered Jurik, 1964 may not work well with coarse-grained
Na0Ac, or HQ carbonares, dolomite, and siderite.
Carbonates in abundance should be removed
prior to particle size fractionation to
improve dispersion. Cations (Ca, Mg, Fe,
etc) dissolved from carbonates during CEC
or elemental analyses can interfere with
these analyses.

Decomposition or organic Gluskoter, 1965 Useful for concentrating clay minerals


matter by low-tempera- in coal or kerogen, with minimal damage
ture ashing the clays.
Decomposition or organic Jackson, 1979; Chlorox may be safer than H202 since the
matter by chemical Anderson, 1963 latter removes Mn02 and alkaline earth
oxidation with H202 or carbonates and phosphates, and may produce
or Chlorox C.J oxalate. Good for decomposition of
colloidal organic matter. This may improve
dispersion prior to particle size fractiona-
tion. Organic matter removal might be
necessary prior to CEC and surface area
determinations due to the high CEC and
surface area values of colloidal organic
matter. Less effective for coalified
organics, kerogen, or oils.
Reduction of iron oxides Mehra & Jackson,: Removes all iron oxides and oxy-hydroxides
by sodium dithionite 1960 other than ilrnenite and magnetite. Does
not strip iron from silicate phases. lron
oxides, if in high concentration, may prevent
good dispersion prior to particle size
fractionation. Iron also fluoresces when
X-irradiated and may produce a high
diffraction background unless a monochrometer
is used.

Dissolution of amorphous Follett et al., 1965; Partial dissolution of crystalline phases


aluminosilicates, alumina, Hasimoto and (particularly kaolinite) also occurs. Should
and silica with 0.5N Jackson, 1960 only be used if cement is in high concentra-
NaOH or 5% Na2CO3 tion and removal is necessary for good
dispersion prior to particle size fractionation.
Dissolution of micas and Kiely and Jackson, Not entirely selective. Calcic feldspars and
resistance of quartz and 1965 feldspars <0.2pm may be dissolved. May be
feldspars with Na2S207 useful to quantitatively approximate
fusion quartz, mica, Na-feldspar and K-feldspar.
Resistance of quartz and Jackson, 1979 Separates Si02 polymorphs from remaining,
cristobalite to H2SiF6 Henderson, 1972 residue after Na25207 fusion.,
dissoiution

Dissolution of 'kaolinite Hasimoto and. Not entirely selective.


and halloysite with 0.5N Jackson, 1960
NaOH after dehydroxy-
lation.
Resistance of anatase and Dolcater et al., May be useful in distinguishing Ti phases
rutile to H2TiF6 dissolu- 1970 from Ti-bearing aluminosilicate phases in
tion clays.

Dissolution of chlorites Jackson, 1979; May be useful in distinguishing chlorites


in hot HC1 F..?%erhoi and from kaolinite by XRD. though not all
P::nningsland, 1978 chlorites (particularly Mg2.-rich chlorites)
will dissolve. Other minerals. especially
iron-bearing minerals and carbonates, may
also dissolve.

Table A.3

A-28
References - Appendix

Alexaides, C. A., and Jackson, M. L., 1965, Quantitative clay


mineralogical analysis of soils and sediments, Clays Clay Miner,
V. 14, p. 35-42

Anderson, J. U., 1963, An improved pretreatment of mineralogical


analysis of samples containing organic matter, Clays Clay Miner,
V. 10, p. 380-387.

Aylmore, L. A. G., 1974, Gas sorption in clay mineral systems, Clay Clay
Miner, v. 22 p. 175-183.

Berna, B., 1968, A new method for decomposition and comprehensive


analysis of silicates by atomic absorption spectrophotometry, Anal.
Chem. v. 49 p. 1682-1686.

Brindley, G. W., 1980, Quantitative X-ray mineral analysis of clays in


Brindley, G. W., and Brown, G. (eds.), Crystal Structures of Clay
Minerals and their X-Ray Identification, Mineralogical Society,
London, p 495.

Brindley, G. W., and Brown, G. (eds.), 1980, Crystal Structures of Clay


Minerals and their X-Ray Identification, Mineralogical Society
London, p. 495.

Brown, G., and Brindley, G. W., 1980, X-ray diffraction procedures for
clay mineral identification, in Brindley, G. W., and Brown, G.
(eds.), Crystal Structures of Clay Minerals and their X-Ray
Identification, Mineralogical Society, London, p 495.

Calvert, C. S., 1984, Simplified, complete CsCl-hydrazine-


dimethylsulfoxide intercalation of kaolinite, Clays and Clay
Minerals, v. 32, p. 125-130.

Calvert, C. S., Bu, S. W., and Weed, S. B., 1980, Mineralogical


characteristics and transformations of a vertical
rock-saprolite-soil sequence in the North Carolina Piedmont: 1.
Profile morphology, chemical composition, and mineralogy, Soil Sci.
Soc. Amer. J. v. 44, p. 1096-1103.

Carter, D. L., Heilman, M. D., and Gonzalez, C. L., 1965, Ethylene


glycol monoethyl either for determining surface area of silicate
minerals, Soil Sci. v. 100, p. 356-360.

Chapman, H. D., 1965, Cation exchange capacity, in Black, C. A. (ed.),


Methods of Soil Analysis, Part 2, American Society of Agronomy,
Madison, Wisconsin, p. 1572.

A-29
Chen, P-Y, 1977, Table of key lines in X-ray powder diffraction patterns
of minerals in clays and associated rocks, Indiana Department of
Natural Research, Geo. Survey Occas., Paper 21, p. 67.

Day, P. R., 1965, Particle fractionation and particle size analysis in


Black, C. A-, (ed.), Methods of Soil Analysis, Part I, American
Society of Agronomy, Madison, Wisconsin, p. 770.

Dolcater, D. L., Syers, J. K., and Jackson, M. L., 1970, Titanium as


free oxide and substituted forms in kaolinites and other soil
minerals. Clays Clay Miner V. 18, p. 71-79.

Elverhoi, A-, and Ronningsland, T. M., 1978, Semiquantitative


calculation of relative amounts of kaolinite and chlorite by X-ray
diffraction. Marine Geology, V. 27, p. M19-M23.

Follett, E. A. C., McHardy, W. J., Mitchell, B. D., and Smith, B. F. L.,


1965, Chemical dissolution techniques in the study of soil clays,
Part I, Clay Miner, v. 6, p. 23-24.

Gard, J. A-, 1971, The electron-optical investigation of clays,


Mineralogical Society, London, p. 383.

Jackson, M. L., 1975, Soil Chemical Analysis - Advanced Course, 2nd.


ed., Published by the author, Madison, Wisconsin

Jackson, M. L., and Abdel-Kader, F. H., 1978, Kaolinite intercalation


procedure for all sizes and types with X-ray diffraction spacing
distinctive from other phyllosilicates, Clays Clay Miner, v. 26,
p. 81-87.

Jurik, P., 1964, Quantitative insoluble residue procedure. Journal of


Sedimentary Petrology, v. 34, p. 666-668.

Kiely, P. V., and Jackson, M. L., 1965, quartz, feldspar, and mica
determination for soils by sodium pyrosulfate fusion. Soil Sci.
Soc. Ameri. Proc. V. 29, p. 154-163.

Kerr, P. F., 1959, Optical Mineralogy, McGraw-Hill, New York, p. 442.

'<lug, H. P., and Alexander, L. E., 1974, X-Ray Diffraction Procedures


for Polycrystalline and Amorphous Materials (2nd. ed.): John Wiley
& Sons, Inc., New York, p. 716.

MacEwan, D. M. C., and Wilson, M. J., 1980, Interlayer and Intercalation


Complexes of Clay Minerals in Brindley, G. W., and Brawn, G.
(eds.), Crystal Structures of Clay Minerals and their X-Ray
Identification, Mineralogical Society, London, p. 495.
MacKenzie, R. C., 1957, The Differential Thermal Investigation of Clays,
Mineralogical Society, London, p. 456.

McKee, T. R., and Brown, J. L., 1977, Preparation of Specimens for


Electron Microscope Examination in Dixon, J. B., and Weed, S. B.
(eds.), Minerals in Soil Environments. Soil Science Society of
America, Madison, Wisconsin, p. 948.

Mehra, O. P., and Jackson, M. L., 1960, Iron oxide removal from soils
and clays by a dithionite-citrate system buffered with sodium
bicarbonate, Clays and Clay Miner, v. 7, p. 317-327.

Roberts, J. M., 1974, X-ray diffraction and chemical techniques for


quantitative soil clay mineral analyses, Ph.D. Thesis, The
Pennsylvania State University.

Roth, C. B., Jackson, M. L., Lo, E. G., and Syers, J. K., 1968,
Ferrous-ferric ratio and CEC Changes on deferration of weathered
micaceous vermiculite, Israel J. Chem. v. 1, p. 261-273.

Sudo, T., Shimoda, S., Yotosumoto, H., and Alta, S., 1981, Electron
Micrographs of Clay Minerals, v. 31 in series Developments in
Sedimentology, Elsevier Sci. Pub. Co., New York, p. 203.

Tan, K. H., and Hajek, B. F., 1977, Thermal analysis of soils in Dixon,
J. B., and Weed, S. B. (eds.), Minerals in Soil Environments, Soil.
Sci. Soc. of America, Madison, Wisconsin, p. 948.

Wilson, M. J. (ed.), 1987, Determinative Methods in Clay Mineralogy,


Blackie & Son (Chapman & Hall, N. Y.), p. 384

Worthington, A. E., 1973, An automated method for the measurement of


cation exchange capacity of rocks. Geophysics v. 38, p. 140-153.
In order to facilitate the most timely publication, SEPM Short Course
Notes are not subjected to the level of review required of an SEPM
Special Publication.
ISBN 0-918985-73-0

S-ar putea să vă placă și