You are on page 1of 12

Chemical Geology - Elsevier Publishing Company, Amsterdam

Printed in The Netherlands


ATOMIC ABSORPTION ANALYSES OF 18 ELEMENTS FROM A SINGLE
DECOMPOSITION OF ALUMINOSILICATE
1
D.E. BUCKLEY and R.E. CRANSTON
Marine Geology, Atlantic Oceanographic Laboratory, Bedford Institute, Dartmouth,
Nova Scotia (Canada)
(Received August 18, 1970)
ABSTRACT
Buckley, D.E. and Cranston, R.E., 1971. Atomic absorption analyses of 18 elements
from a single decomposition. Chern. Geol., 7: 273-284.
The atomic absorption determination of 18 elements in aluminosilicate minerals
and marine sediments is accomplished after a single decomposition method using
hydrofluoric acid and aqua regia. The digested sample is combined with boric acid
to give a 100 ml solution from which all 18 elements including Si, Ti, V, and Al can
be determined directly by means of atomic absorption spectroscopy. Inter-element
interferences and ionization effects are fully compensated by using combined stand-
ards having a matrix similar to that of the samples. Precision and accuracy for
most determinations are comparable to the other more complex analytical techniques
while the speed of analyses is considerably faster than existing methods.
INTRODUCTION
In the field of marine inorganic geochemistry an ever increasing need
exists for good quality chemical analyses of large numbers of samples of
natural sediments. These analyses commonly require determinations of
both major and minor elements from a wide variety of materials. These
range from nearly pure calcium carbonate to clay sediments, the latter
consisting of an assemblage of 4 or 5 principal aluminosilicate minerals.
This paper describes the procedure developed at our laboratories to facili-
tate the rapid and reasonably precise analyses of 18 elements in clay
minerals by means of atomic absorption spectroscopy.
The objective was to analyse small samples (100 mg) of suspended
clay sized sediments. These analyses were required to obtain data on the
major cation constituents of the silicates to assist in mineralogical identifi-
cation and classification. In addition minor element analyses were required
to obtain data on the concentration, absorption and exchange of trace-
transition elements from the marine environment. To carry out these
analyses a single decomposition technique was adapted from that described
by Bernas (1968). Samples were digested in aqua regia and HF in a sealed
teflon bomb. This technique offered several distinct advantages over open
acid digestion methods or fusion methods. Because the sample was decom-
posed in a closed container volatilization losses were eliminated thus allow-
ing the analyses of Si, AI, V, and Ti from the same solution as all other
1 n.1. Contribution 222
Chern. Geol., 7 (1971) 273-284
273
elements. Because no fluxes were used, all alkali and alkaline earth elements
could be analysed. The entire sample preparation is simple requiring no
special skills and is easily adaptable to mass production procedures. The
total decomposition and dilution procedure could be completed in less than
one hour.
Effective reduction of interelement interference was achieved over a
range of concentrations by matching the matrix of standards and samples.
The presence of fluoboric acid formed by the addition of boric acid to the
digested sample had no apparent detrimental effect on the analyses, and
dissolved any insoluble fluoride salts which may have precipitated in the
digestion bomb.
Instrumental settings were selected in such a way that a wide variety
of aluminosilicates could be analysed with the single solution preparation
technique. No dilution beyond the initial 1:1,000 was necessary for comple-
tion of the 18 element analysis.
TECHNIQUE
Reagents and standards
All reagents used were certified reagent grade chemicals. Standard
cation solutions are prepared from high purity metals or salts as noted in
Table I. Water used as solvent or dilutant is prepared by micro-filtration
and ion exchange to obtain a resistivity of 18 MQ-cm (Super Q water -
Millipore Corp.).
A one litre stock standard of each element is prepared to give a cation
concentration of 1,000 J.l g/ml. From these stock standards a combined
standard "A" is prepared by adding the proportionate quantity of stock solu-
tion to a one litre polypropylene bottle thus giving the final concentr;ltions
of each element as shown in Table I. These combined quantities represent
an upper limit of the concentration to be expected from digestion and
1:1,000 dilution of most marine sediments (Chester, 1965, pp.68-77). Three
additional standards are prepared from combined standard "A" by 1:2, 1:10,
and 1:20 dilutions. All four standards are prepared in such a way that all
contain 5 % v Iv HF, 5.6% w Iv H3 B0
3
, and 1% v Iv aqua regia in order that
the standards match the sample solution obtained by the decomposition
method described below.
Sample preparation
Decomposition of all samples as well as the Si0
2
and Ti0
2
standards
is carried out in a teflon-lined bomb similar to that described by Bernas
(1968). A 100 mg sample is placed in the bomb and wetted with 1 ml of aqua
regia. 6 ml of- concentrated HF is added and the bomb sealed tightly. The
bomb is heated at 100
0
C for 30 min for most samples, but has required 60
min for some clay samples. After cooling, the sample solution is washed
into a 125 ml polypropylene bottle containing 5.6 g of H3B03 and 20 ml of
water. The insoluble metal fluorides, present in the digested sample after
treatment with the HF, are dissolved in the boric acid system (Bernas,
274
Chern. Geol., 7 (I971) 273-284
(
"
)

:
:
r

(
1
)

1
3

0

(
1
)

S

.
.
,

-
;
:
:
;

C
D

.
.
,

.
.
'
:
:
!

t
-
:
>

.
.
,

w

I

t
-
:
>

0
0


t
-
:
>

.
.
,

<
.
i
l

T
A
B
L
E

I

E
l
e
m
e
n
t
a
l

s
t
a
n
d
a
r
d
s

M
a
j
o
r

e
l
e
m
e
n
t
s

A
l

C
a

F
e

K

S
o
u
r
c
e

m
e
t
a
l

C
a
C
0
3

m
e
t
a
l

K
C
I

S
o
l
v
e
n
t

H
C
I

H
C
I

H
C
I

w
a
t
e
r

S
t
d
.

A

1
0
0

1
0
0

1
0
0

5
0

(
u
g
/
m
l
.
)

M
i
n
o
r

e
l
e
m
e
n
t
s

C
o

C
r

C
u

L
i

S
o
u
r
c
e

m
e
t
a
l

K
2
C
r
0
4

m
e
t
a
l

L
i
C
I

S
o
l
v
e
n
t

H
N
0
3

w
a
t
e
r

H
N
0
3

w
a
t
e
r

S
t
d
.

A

0
.
2

0
.
5

0
.
2

0
.
1

(
u
g
/
m
l
.
)

1
D
e
c
o
m
p
o
s
i
t
i
o
n

c
a
r
r
i
e
d

o
u
t

i
n

t
e
f
l
o
n

b
o
m
b
.

M
g

M
n

N
a

S
i

T
i

m
e
t
a
l

m
e
t
a
l

N
a
C
I

S
i
0
2

T
i
0
2

H
C
I

H
C
I

w
a
t
e
r

H
F
1

H
F
I

5
0

2

4
0

3
5
0

2
0

N
i

P
b

S
r

V

Z
n

m
e
t
a
l

m
e
t
a
l

S
r
(
N
0
3
)
2

V
2
0
5

m
e
t
a
l

H
N
0
3

H
N
0
3

w
a
t
e
r

H
C
I

H
C
I

0
.
3

0
.
2

1
.
0

0
.
5

0
.
2

1968, pp.1685-1686). Shaking the sample completes the dissolution more
quickly. The sample solution is transferred to a glass volumetric flask to
bring the volume to 100 ml and is then returned to the polypropylene bottle
for storage until analysed. Blank solutions are prepared in the bomb in the
same way as the sample. Bernas (1968) tested the effect of the HBF4 -
H3 B03 system on glass for short periods of time and found no contamination.
We confirmed this finding from use of glass volumetrics for periods of a
few minutes to an hour.
Instrumental analyses
An atomic absorption spectrophotometer (Perkin-Elmer Model 303)
was used with a model 165 recorder. Intensitron lamps (trade mark: Perkin-
Elmer Corp.) were used for all elements except Si, Sr, Pb, Fe, and Ti in
which cases standard hollow cathode lamps were used. Except as noted in
the text and tables of this paper, all instrumental settings were as
recommended in the Perkin-Elmer Analytical Methods Book.
A baseline for each element was determined continuously by aspirating
an aqueous solution containing HF, H3B03 and aqua regia in the same con-
TABLE II
Analytical conditions for atomic absorption spectroscopy
Major
elements
Upper limit
elemental conc.
in rock
Upper limit
oxide cone. in
rock
Sensitivity
(ug/ml/1 %abs)
Burner type
1
Burner position
2
Recorder expo
Minor
elements
Lower limit
cone. in rock
(p.p.m.)
Sensitivity
(ug/ml/1 %abs)
Burner type
1
BUrner position
2
Recorder expo
Al
15%
28%
1.2
N
II
X 1
Co
20
0.15
B
II
X 30
Ca
10%
14%
1.3
N
+
Xl
Cr
20
0.20
B
II
X 30
Fe K Mg
10% 5% 5%
14% 6% 8%
4.0 1.5 0.6
5 cm 5 em N
+ + +
x3 X3 Xl
Cu Li Ni
10
0.07
B
II
X 30
10
0.05
B
II
X 30
20
0.15
B
II
X 30
Mn
0.3%
0.4%
0.06
B
II
x3
Pb
50
0.50
B
II
X 30
Na Si
4% 35%
5% 75%
0.5 4.0
5 em N
+ II
X 1 xl
Sr V
100
0.12
N
II
X 10
100
1.0
N
II
X 30
Ti
3%
5%
2.0
N
II
X 10
Zn
20
0.20
B
II
X 10
15 em = short path air-acetylene burner; N = nitrous oxide-acetylene burner; and
B = Boling (3 slot) burner.
2 I I = burner parallel to light path; and + = burner perpendicular to light path.
276 Chern. Geol., 7 (1971) 273-284
centrations as in the sample and standards. Burner type, burner orientation,
fuel selection and recorder expansion were all chosen to allow the direct
determination of the element from the 100 ml sample solution without any
concentration or dilution steps. The specific instrumental conditions are
given in Table II and the range of concentrations of the elements or oxides
in the undigested sample for which these conditions are appropriate are
also given. Only the lower limits for the common minor elements are given
since a 1,000-fold dilution of these elements is always below the upper limit
of the instrument. Sensitivity is defined as the concentration of metal in
solution that gives a 1% absorption signal for the instrumental conditions
designated in Table ll.
Calculations
Calibration for each set of standards was accomplished by computing
the derivative between each pair of standards. Thus, rather than an integrated
curve, a segmented curve was calculated in which small portions of the
curve were bracketed by two standards. The sample concentrations were
interpolated from within the small ranges. The use of the computer to carry
out this standard-sample comparison has resulted in the removal of much
of the human error and bias that usually influences the final result. The
computer was also used for error analyses and data compilation.
RESULTS
Over a period of several months a record of 20 replicate analyses of
diabase W-1 (U.S. Geological Survey Standard Rock) and 16 analyses of
Illite 35 (American Petroleum Institute Clay Minerals Standard Project 49,
Fithian Illinois) was maintained in our laboratory. A summary of these data
is given in Table III. Our average determination for W-1 is compared willi
the amount present as reported in the summary by Fleischer and Stevens
(1962). The average relative error for the 9 major elements in W -1 is 1. 9 %.
The chemical composition of illite minerals appears to be variable (Grim,
1968, p.580), although no comparative studiesbydifferinganalyticaltech-
niques are known. For this reason only the determinations by our laboratory
are reported.
The coefficients of variation differ considerably from element to
element for the trace metals. These coefficients can be labelled as the
"noise" of the system since this variation is directly dependent on sensitivity
of the element as shown by vanadium, and inversely dependent on low metal
concentration, such as is the case for lithium. The sensitivity / concentration
ratio for each element has been plotted with the coefficients of variation
(Fig.1) for W-1 and Illite 35 replicate analyses (Table III).
Table IV is a compilation of data from a single analysis of each of five
U.S.G.S. rock standards. The elemental concentration found was determined
by means of the computer interpolation method described earlier. The
values tabulated as (a) were calculated on the basis of standard curves,
derived from the four combined standards, plus solutions of G-2 and W-1
treated as known standards containing the amount of metal quoted from
Chern. Geol., 7 (1971) 273-284 277
t
<
>

T
A
B
L
E

I
I
I

-
.
:
,

0
0

R
e
s
u
l
t
s

f
o
r

r
e
p
l
i
c
a
t
e

a
n
a
l
y
s
e
s

M
a
j
o
r

e
l
e
m
e
n
t
s

(
%
)

A
l

C
a

F
e

K

M
g

M
n

N
a

S
i

T
i

W
-
1

P
r
e
s
e
n
t
1

7
.
9
7

7
.
8
3

7
.
7
8

0
.
5
3
1

3
.
9
9

0
.
1
3
2

1
.
5
4

2
4
.
5

0
.
6
4

F
o
u
n
d

8
.
1
7

7
.
6
6

7
.
7
2

0
.
5
4
0

4
.
2
0

0
.
1
3
1

1
.
5
6

2
4
.
6

0
.
6
3

S
t
d
.

d
e
v
.

0
.
2
9

0
.
7
7

0
.
2
7

0
.
0
2

0
.
2
9

0
.
0
0
5

0
.
1
1
4

1
.
2

0
.
0
3

C
o
e
f
f
.

v
a
r
.

3
.
5

1
0
.
0

3
.
5

3
.
7

6
.
9

3
.
8

7
.
3

4
.
8

4
.
8

%

d
i
f
f
e
r
e
n
c
e

+
2
.
5

-
2
.
2

-
0
.
8

+
1
.
9

+
2
.
8

-
0
.
8

+
1
.
3

+
0
.
1

-
1
.
6

I
l
l
i
t
e

3
5

F
o
u
n
d

9
.
8
5

0
.
9
2

3
.
8
9

3
.
6
4

1
.
3
7

0
.
0
2
7
5

0
.
4
1

2
5
.
5

0
.
5
0

S
t
d
.

d
e
v
.

0
.
2
5

0
.
0
7

0
.
1
3

0
.
1
0

0
.
0
7

0
.
0
0
1
4

0
.
0
3

1
.
6

0
.
0
3
5

C
o
e
f
f
.

v
a
r
.

2
.
5

7
.
6

3
.
3

2
.
7

5
.
1

5
.
1

7
.
3

6
.
3

7
.
0

M
i
n
o
r

e
l
e
m
e
n
t
s

(
p
.
p
.
m
.
)

C
o

C
r

C
u

L
i

N
i

P
b

S
r

Y

Z
n

W
-
1

(
1

C
o
n
e
.

r
a
n
g
e
1

:
:
>
'
3
5
-
5
5

9
0
-
1
6
0

8
0
-
1
5
0

1
0
-
1
2

5
5
-
8
8

5
-
1
0

1
5
0
-
3
0
0

1
2
0
-
3
2
0

2
0
-
9
5

C
D

<
5
0

?

F
o
u
n
d

6
8

1
2
8

1
0
5

1
4

8
1

1
6
7

2
4
9

8
8

0

S
t
d
.

d
e
v
.

1
8

8

1
7

3

1
2

1
9

3
4

8

C
D

C
o
e
f
f
.

v
a
r
.

2
6

6

1
6

2
1

1
5

1
1

1
4

9


-
.
:
,

I
l
l
i
t
e

3
5


F
o
u
n
d

4
5

1
1
3

6
0

4
8

8
0

5
3

1
1
8

1
7
8

8
6

'
"

S
t
d
.

d
e
v
.

1
2

1
8

1
5

4

1
7

1
3

1
6

4
3

1
3

-
.
:
,

.
t
:
!

C
o
e
f
f
.

v
a
r
.

2
7

1
6

2
5

8

2
1

2
4

1
4

2
4

1
5

t
<
>

-
.
:
,

1
Y
a
l
u
e
s

f
r
o
m

F
l
e
i
s
c
h
e
r

a
n
d

S
t
e
v
e
n
s

(
1
9
6
2
)
.

w

I

t
<
>

0
0

o
j>
.

28
26
24
22
....
IN;
20
z
0

18
a:
;; 16
"-
Iz' f
Li 0
14
I-
z
!oJ
12
Q
"-
"-
10 !oJ
0
u
8
6 .6. W-I dolO
4
IIIite:# 35 data
2
00 0.001 0.002 0.003 0.604 0.005 0.006
INSTRUMENT SENSITIVITY (ppm/%obs.) I SAMPLE CONCENTRATION (ppm)
Fig.1. Each pair of points represents the variation of the trace element
statistics for two types of silicates. The positive slope of the line indicates
that decreasing concentration results in higher coefficients of variation.
Fleischer and Stevens (1962). The values tabulated as (b) were calculated
on the basis of the four combined standards alone and treating G-2 and W-1
as unknowns. In all cases the data calculated from the six-standard curves
have a lower relative error than those from the four-standard curves.
Calcium values determined in this particular set of analyses have higher
relative errors than are normally obtained. No apparent reason could be
found for these high values since such anomalies did not occur in other
replicate analyses of the same standard rocks. Nickel and chromium values
for GSP-1 roay be high because of contamination from the stainless steel
casing surrounding the teflon bomb.
The data in Table V represent part of the results obtained from
analyses of suspended clay-sized sediments collected from a glacial fiord
in southeast Alaska (Buckley and Loder, 1968). These data are from analyses
of sediments from two sample stations with fractionation of the clay-sized
material into three size groups having average diameters less than O.5J.1.,
O.5-2J.1. and 2-4J.1.. The mineralogy determined by X-ray diffractometry
consists of biotite, chlorite, hornblende, plagioclase feldspar, and quartz.
The variability of the mineral assemblage with the variation in size is
reflected in the major element data. The increase in calcium and sodium
with increasing grain size correlates with a relative increase in the quantity
of plagioclase feldspar. Similarly increasing silicon reflects higher per-
centages of quartz as well as feldspar. The lesser percentages of biotite
and chlorite in the coarser sizes are indicated by decreasing iron and
Chern. Geo!., 7 (1971) 273-284 279
T
A
B
L
E

I
V

"
"

U
.
S
.
G
.
S
.

S
t
a
n
d
a
r
d

R
o
c
k
s

0
0

0

M
a
j
o
r

e
l
e
m
e
n
t
s

(
%
)

A
l

C
a

F
e

K

M
g

M
n

N
a

S
i

T
i

A
G
V
-
1

p
r
e
s
e
n
t
1

9
.
0
0

3
.
5
6

4
.
7
5

2
.
4
1

0
.
9
0

0
.
0
7
6

3
.
2
1

2
7
.
5

0
.
6
5

f
o
u
n
d

a
)

9
.
1
7

3
.
8
6

4
.
5
9

2
.
3
1

0
.
9
6

0
.
0
7
4

3
.
1
5

2
9
.
6

0
.
6
1

b
)

9
.
3
2

4
.
0
1

4
.
5
8

2
.
3
4

0
.
9
6

0
.
0
7
3

3
.
1
0

2
8
.
0

0
.
6
5

B
C
R
-
1

p
r
e
s
e
n
t

7
.
2
2

4
.
9
7

9
.
4
4

1
.
3
9

1
.
9
8

0
.
1
3
6

2
.
4
6

2
5
.
4

1
.
3
4

f
o
u
n
d

a
)

7
.
2
5

5
.
0
7

9
.
0
4

1
.
2
9

2
.
1
0

0
.
1
3
7

2
.
1
7

2
5
.
1

1
.
3
0

b
)

7
.
4
3

5
.
0
9

8
.
9
8

1
.
2
8

2
.
1
0

0
.
1
3
8

2
.
1
6

2
3
.
2

1
.
3
0

G
S
P
-
1

p
r
e
s
e
n
t

8
.
0
0

1
.
4
5

3
.
0
3

4
.
5
6

0
.
5
8

0
.
0
3
4

2
.
1
4

3
1
.
4

0
.
4
2

f
o
u
n
d

a
)

8
.
0
0

1
.
5
8

3
.
0
0

4
.
7
0

0
.
5
5

0
.
0
3
4

1
.
9
4

3
1
.
6

0
.
4
2

b
)

8
.
2
7

2
.
1
1

2
.
9
2

4
.
6
9

0
.
5
5

0
.
0
3
1

1
.
9
4

3
0
.
0

0
.
4
2

G
-
2

p
r
e
s
e
n
t

8
.
1
1

1
.
4
2

1
.
9
3

3
.
7
4

0
.
4
7

0
.
0
2
6

3
.
0
8

3
2
.
3

0
.
3
2

f
o
u
n
d

b
)

8
.
4
5

1
.
9
7

1
.
8
1

3
.
7
1

0
.
4
2

0
.
0
2
5

3
.
0
3

3
0
.
8

0
.
3
0

W
-
1

p
r
e
s
e
n
t
2

7
.
9
1

7
.
8
3

7
.
7
8

0
.
5
3

3
.
9
9

0
.
1
3
2

1
.
5
4

2
4
.
5

0
.
6
4

f
o
u
n
d

b
)

8
.
2
0

8
.
0
4

7
.
6
6

0
.
5
3

4
.
1
1

0
.
1
3
3

1
.
5
1

2
2
.
5

0
.
6
8

M
i
n
o
r

e
l
e
m
e
n
t
s

(
p
.
p
.
m
.
)

C
o

C
r

e
u

L
i

N
i

P
b

S
r

V

Z
n

A
G
V
-
1

r
a
n
g
e
1

1
0
-
3
0

7
-
4
5

5
0
-
8
3

9
-
1
4

1
1
-
3
1

1
8
-
4
8

3
4
8
-
1
0
5
0

7
0
-
1
7
1

6
4
-
3
0
4

n

f
o
u
n
d

a
)

3
1

2
0

4
3

1
0

6
1

<

5
0

6
3
1

1
1
1

7
9

g

b
)

3
2

2
0

4
3

1
1

6
3

<

5
0

6
3
1

1
0
1

7
8

3

B
C
R
-
1

r
a
n
g
e

2
9
-
6
0

8
-
4
5

7
-
3
3

1
0
-
1
9

6
-
3
0

4
-
3
5

2
4
4
-
5
5
0

1
2
0
-
7
0
0

9
4
-
2
7
8

0

f
o
u
n
d

a
)

4
3

4
7

1
3

1
2

2
7

<

5
0

3
1
1

4
7
6

1
1
2

C
D

b
)

4
6

5
6

1
6

1
4

2
7

<

5
0

3
1
9

4
7
6

1
1
2

g
,

G
S
P
-
1

r
a
n
g
e

3
-
2
2

5
-
1
8

8
-
5
4

2
5
-
4
7

3
-
2
5

1
4
-
8
0

1
4
8
-
4
0
0

2
5
-
6
8

5
4
-
3
4
0

"
"
"

f
o
u
n
d

a
)

2
3

5
9

1
8

4
0

3
1

6
2

2
6
8

1
1
1

8
4

-
;
:
:
;

b
)

2
4

5
4

1
9

2
9

3
1

<

5
0

2
7
4

1
0
1

8
3

<
.
0

"
"
"

G
-
2

r
a
n
g
e

2
-
2
1

5
-
2
9

2
-
1
7

2
6
-
6
3

2
-
1
4

1
5
-
4
3

2
3
5
-
6
5
0

2
6
-
6
0

4
2
-
1
3
8

.
.
.
.


f
o
u
n
d

b
)

2
0

2
0

1
6

3
0

2
0

<

5
0

4
7
3

1
0
0

7
3

"
"

r
a
n
g
e
2

"
"
"

W
-
1

3
5
-
5
5

9
0
-
1
6
0

8
0
-
1
5
0

1
0
-
1
2

5
5
-
8
8

5
-
1
0

1
8
0
-
3
0
0

1
2
0
-
3
2
0

2
0
-
9
5

t
.
:
>

I

f
o
u
n
d

b
)

5
5

1
7
4

1
2
0

2
5

6
3

<

5
0

2
0
5

2
0
9

8
0

"
"

0
0

0
1
>
-
1
F
l
a
n
a
g
a
n

(
1
9
6
9
)
.

2
F
l
e
i
s
c
h
e
r

a
n
d

S
t
e
v
e
n
s

(
1
9
6
2
)
.

a
)

c
a
l
c
u
l
a
t
e
d

o
n

b
a
s
i
s

o
f

6

s
t
a
n
d
a
r
d
s
.

b
)

c
a
l
c
u
l
a
t
e
d

o
n

b
a
s
i
s

o
f

4

s
t
a
n
d
a
r
d
s
.

TABLE V
Analyses of glacial marine suspended sediment
Major
elements (%) Al Ca Fe K Mg Mn Na Si Ti
ND-2 surface
0.5Jl 8.78 2.64 9.30 3.75 4.18 0.146 1.53 18.2 0.96
0.5-2Jl 8.53 3.22 8.25 3.31 3.67 0.130 2.01 17.7 0.80
2-4Jl 8.65 3.76 7.19 2.81 3.16 0.122 2.41 22.0 0.75
ND-4F 1 meter
0.5Jl 8.53 2.15 9.72 3.42 3.34 0.155 1.25 18.4 1.12
0.5-2Jl 8.65 2.68 9.24 3.40 4.24 0.152 1.59 18.8 1.01
2-4Jl 8.81 2.97 8.05 3.06 3.86 0.140 2.03 19.0 0.92
Minor
elements Co Cr Cu Li Ni Pb Sr V Zn
ND-2 surface
0.5Jl 59 125 138 61 91 716 252 145 412
0.5-2Jl 50 263 93 49 100 519 300 155 282
2-4Jl 40 230 66 38 131 317 352 144 208
ND-4F 1 meter
0.5Jl 58 216 130 63 68 68 215 175 284
0.5-2Jl 66 136 100 56 72 73 340 175 229
2-4Jl 55 119 75 43 71 104 316 155 220
potassium. The minor element data generally illustrate the tendency of
transition and heavy metals to be concentrated with layered silicates and
minerals having the smallest grain size (largest specific surface area)
because of increased adsorption capacity (Krauskopf, 1967, pp. 592-593).
DISCUSSION
Advantages of the preparation technique
In the preparation of samples, our objective was to choose a decompo-
sition technique that would be applicable to the widest possible variety of
naturally occurring aluminosilicates and that this decomposition would yield
a single sample solution in which all elements could be determined directly.
Langmyhr and Paus (1968a) described several methods for decomposition
of silicon containing materials and a modification of their method number 2
was chosen by Bernas (1968) and by ourselves because of several adv;p1tages
'Fhe HF digestion technique contributes little contamination to the sample
solution if carried out in plastic containers, and eliminates the necessity of
using fluxes which contain high concentrations of alkali salts as well as
trace element contaminants. The addition of 5.6g/100 ml boric acid to
dissolve the fluorides appeared to be an effective means of completing the
decomposition of the aluminosilicates and allowed the precise volumetric
handling of the sample. Although Langmyhr and Paus (1968b) encountered
Chem. Geol., 7 (1971) 273-284 281
difficulty in dissolving the precipitated fluorides formed in the digestion of
clay samples we found no such difficulty. Perhaps this discrepancy is clue
to differences in the concentration of H3B03. We found that certain samples
containing high carbon concentrations yield an insoluble residue after treat-
ment with HF, but that this residue is probably unoxidized carbon and
represents less than 1% of the total weight of the sample. In a few cases we
had difficulty in digesting clay samples in the 30 m ~ period at 100 C
usually required for other silicates. In these cases we found that longer
periods of time for the digestion is more effective than heating at higher
temperatures. The over-all preparation technique including decomposition
of 20 samples and preparation of a complete set of standards can easily be
accomplished by one technician in a day.
Advantages of the analytical technique
In addition to the normal instrumental analyses advantages that atomic
absorption spectroscopy offers over more cumbersome classical methods
of silicate analyses there are some specific advantages to the technique
described here. The principal advantage is that all 18 elements may be
determined from a single sample solution and with the use of a set of only
four or six combined standards. This achievement is based on two factors
as follows: (1) all interelement and reagent interferences are fully compen-
sated over a range of chemical compositions by matching'the matrix of
samples and standards; and (2) the sensitivity of the instrument can be ad-
justed so that major and trace constituents can be analyzed from the same
sample solution.
Van Loon and Parissis (1969) carried out a fairly extensive study of
inter-element interferences in silicate analysis by atomic absorption. Their
results indicated that Fe, Si, AI, Na and Mn could be analyzed at the relative
concentration levels found in most silicates without mutual interference.
They found that interference of potassium by sodium could be overcome by
the addition of sodium to the standards. The interferences of Si and Al with
Mg and Ca determinations was reduced by addition of 10,000 p.p.m. La to
samples and standards, and an unspecified quantity of lithium to the stand-
ards. In the case of the calcium analyses the Si02 level had to be maintained
below 150 p.p.m. Langmyhr and Paus (1968b) state that no interferences are
encountered for Co, Cu, Ni, Pb or Cr analyses.
It should be obvious that if all cation standards are combined in a
single solution in the same proportion as that present in the sample then all
mutual interferences will be compensated and no need exists for addition of
excessive salts or acids to standards or sample solutions. Our data indicate
that this is indeed the case even for aluminosilicate samples having matri-
ces considerably different from those of the combined standard. It should
be noted that no effort is made to match anion matrix of the combined
standards with that of the samples although care is taken to match precisely
the concentration of the reagents in all solutions including the aspiration
fluid.
Our experience with the analyses of the U.S.G.S. rock standards PCC-l
and DTS-l has shown that our standard matrix based on the average compo-
sition of W-l cannot be applied to silicates having low concentrations of Al
282 Chern. Geol., 7 (1971) 273-284
and high concentrations of Mg. Similar experience in the analyses of carbo-
nates has indicated that the matrix of the combined standard must match
within certain limits. These limits are determined by the nature of the
major element composition. The disadvantage of determining this matrix
and its limits in a completely unknown substance is obvious.
Analyses of major constituents in silicate rocks are usually achieved
by diluting the sample solution to a level where the element concentration
is within the optimum range of the instrument at its maximum sensitivity.
This dilution places the concentration of the minor elements below the
detection limit and therefore makes it impossible to analyse both minor and
major elements tn the same solution. We found that a series of dilution
steps may introduce random and systematic errors that are difficult to
evaluate. As a result we chose a single dilution technique which yields a
1:1,000 dilution of the original sample concentration. This leaves the con-
centration of AI, Ca, Fe, K, Mg, Na, and Si above the optimum range for
standard conditions of the instrument. To establish a convenient working
range we found that altering the name path-length by turning the burner
perpendicular to the light source (see Table II) reduced the sensitivity by
as much as 95% and still allowed the element to be determined with suffi-
cient precision.
The precision and accuracy of the minor element analyses were limited
by the sensitivity of the particular element, and by the concentration level.
The signal noise of the absorption from the recorder output may be such
that the analyst may have difficulty in determining the precise peak height.
In the analyses of U.S.G.S. rock standard G-2 the relative detection limit
(Slavin, 1968) was exceeded for the elements Co, Cr, Ni, Pb, and V. These
low concentrations rarely occur in layered silicates or marine sediments.
One of the improvements in the analytical procedure has been the
incorporation of the computer in the calculation stages. The method of cal-
culating the slope of the standard curve between standards has eliminated
random errors caused by manual plotting and interpolation. In addition non-
linearity of the curve as a function of ionization effects has been calculated
more precisely than could be graphically plotted. An example of this
phenomenon was encountered in the analyses of potassium in which case
the standard curve had a steeper slope at the intermediate concentration
levels than at low concentration levels. This phenomenon may have been
caused by the ionization of relatively more potassium atoms at low concen-
trations than at high concentrations. By calculating the slopes exactly as
the combined standards indicated, all interferences were taken into consid-
eration and the interpolated values for the samples should be accurate
provided the sample matrix and standard matrix were matched.
Future extensions of the technique
Work has already begun on analyses of Ba in the U.S.G.S. standards
and we have found that our results are within the range of recommended
values in all cases, without the addition of excess quantities of alkali metal
to suppress ionization. Other trace metals such as Hg, Cd, Mo, and Sn will
be added to future analyses of marine samples after preliminary standard
calibration work is completed.
Chern. Geol., 7 (1971) 273-284 283
ACKNOWLEDGEMENT
The authors wish to thank Drs. ~ Baadsgaard, Department of Chemis-
try of the University of Alberta; D.W. Spencer of Woods Hole Oceanographic
Institution; and J.M. Bewers and B.R. Pelletier of the Atlantic Oceano-
graphic Laboratory, Bedford Institute, for critically reading the manuscript
and offering helpful suggestions in the presentation of the data.
REFERENCES
Bernas, B., 1968. A new method for decomposition and comprehensive analysis of
silicates by atomic absorption spectrometry. Anal. Chem., 40: 1682-1686.
Buckley, D.E. and Loder, T.C., 1968. Particulate organiC-inorganic geochemistry of
a glacial fiord. In: D.C. Burrell and D.W. Hood (Editors), Clay-Inorganic and
Organic-Inorganic Associations in Aquatic Environments, I. University of
Alaska, Institute of Marine Science, College, Alaska, pp.1-53.
Chester, R., 1965. Elemental geochemistry of marine sediments. In: J.P. Riley and
G. Skirrow (Editors), Chemical Oceanography, 2. Academic Press, London,
pp.68-77.
Flanagan, F.J., 1969. U.S. Geological Survey standards - II. First compilation of
data for the new U.S.G.S. rocks. Geochim. Cosmochim. Acta, 33: 81-120.
Fleischer, M. and Stevens, R.E., 1962. Summary of new data on rock samples G-1
and W-l. Geochim. Cosmochim. Acta, 26: 525-543.
Grim, R.E., 1968. Clay Mineralogy. McGraw-Hill, New York, N.Y., 2nd ed., 592 pp.
Krauskopf, K.B., 1967. Introduction to Geochemistry. McGraw-Hill, New York, N.Y.,
706 pp.
Langmyhr, F.J. and Paus, F.E., 1968a. The analysis of inorganic siliceous materials
by atomic absorption spectrophotometry and the hydrofluoric acid decomposi-
tion technique. Anal. Chim. Acta., 43: 397-408.
Langmyhr, F.J. and Paus, F.E., 1968b. Hydrofluoric acid decomposition - atomic
absorption analysis of inorganic siliceous materials. Atomic Absorption News-
letter, 7: 103-106.
Slavin, W., 1968. Atomic Absorption Spectroscopy. Interscience, New York, N.Y.,
307 pp.
Van Loon, J.C. and Parissis, C.M., 1969. Scheme of silicate analysis based on the
lithium metaborate fusion followed by atomic absorption spectrophotometry.
Analyst, 94: 1057-1062.
284
Chem. Geol., 7 (1971) 273-284

You might also like