Vous êtes sur la page 1sur 8

Analytica Chimica Acta 487 (2003) 181188

Simultaneous determination of copper, nickel, cobalt and


zinc using zincon as a metallochromic indicator with
partial least squares
J. Ghasemi a, , Sh. Ahmadi a , K. Torkestani b
a

Department of Chemistry, Razi University, Kermanshah, Iran


b Research Institute of Petroleum Industry, Tehran, Iran

Received 3 December 2002; received in revised form 17 April 2003; accepted 2 May 2003

Abstract
The partial least squares (PLS) applied to the simultaneous determination of the divalent ions of copper, nickel, cobalt and
zinc based on the formation of their complexes with 2-carboxy-2 -hydroxy-5 -sulfoformazyl benzene (zincon). The absorption
spectra were recorded from 515 through 750 nm. The effect of pH on sensitivity and the selectivity was studied in the range
3.010.0 and the pH 8.0 was choused according to net analyte signal (NAS) as a function of pH. The concentration range
for Cu2+ , Ni2+ , Co2+ and Zn2+ in solution calibration sets were 02.6, 04.6, 03.0 and 04.92 ppm, respectively. The root
mean squares differences (RMSD) for copper, nickel, cobalt and zinc were 0.0181, 0.0488, 0.0309 and 0.0463, respectively.
2003 Published by Elsevier Science B.V.
Keywords: Partial least squares; Copper; Nickel; Cobalt; Zinc; Spectrophotometric determination

1. Introduction
Copper, nickel, cobalt and zinc are metals that appear together in many real samples. Several techniques
such as X-ray fluorescence [1], atomic fluorescence
spectrometry [2], polarography [3], chromatography
[4], atomic absorption spectrometry [5,6], etc. have
been used for the simultaneous determination of these
ions in different samples. Among the most widely
used analytical methods are those based on the UV-Vis
spectrophotometry techniques [711], due to the resulting experimental rapidity, simplicity and the wide
application. However, the simultaneous determination
of these ions by the use of the traditional spectropho Corresponding author. Tel.: +98-8317233063;
fax: +98-8318231618.
E-mail address: jahan.ghasemi@tataa.com (J. Ghasemi).

tometry techniques is difficult because, generally, the


absorption spectra overlap in a bright region and the
superimposed curves are not suitable for quantitative
evaluation.
Nowadays quantitative spectrophotometry has been
greatly improved by the use of a variety of multivariate
statistical method; particularly principle component
regression (PCR) and partial least squares regression
(PLS).
PLS regression has been found important in handling regression tasks in case there are many variables.
The theoretical basic for PLS regression is found in
several references [1621]. The basic aspect of PLS
regression is that it suggests that, after decomposition
of X and Y matrices into two new score and loading
matrices using singular value decomposition or principal component analysis, we should maximize the covariance between score vector in X-space and a score

0003-2670/03/$ see front matter 2003 Published by Elsevier Science B.V.


doi:10.1016/S0003-2670(03)00556-7

182

J. Ghasemi et al. / Analytica Chimica Acta 487 (2003) 181188

vector in Y-space or equivalently to maximize the size


of the loading vector in Y-space derived from the score
the vector in X-space.
Since the metallochromic indicators are also known
as acidbase indicators, then, the complex formation
reaction is affected by the variation of pH of the solution. So it is clear to see that the efficiency of the
complex formation and as well as the sensitivity and
selectivity are depends on the established pH value.
The effect of pH on the sensitivity and selectivity was
studied according to the net analyte signal (NAS) for
each component in a first-order system [15,16]. The
net analyte signal is defined in Eq. (1).
NAS = (I Rn R+
n )rn

(1)

where I is the identity matrix, Rn the matrix of pure


spectra of all constituents except the nth analyte, R+
n
the pseudo inverse of Rn and rn is the spectrum of
the analyte. The NAS is a vector and is related to the
regression vector in following equations:
c = Rb + e

(2)

NAS
b=
NAS2

(3)

in which ||NAS||2 designates the square root of the sum


of squares of each element in the vector, b. Sensitivity
and selectivity was calculated by using the following
equations:
SEN =

1
= NAS2
b 2

(4)

SEL =

NAS2
1
=
b2 rn 2
rn 2

(5)

In this work, the simultaneous spectrophotometric


determination of copper, nickel, cobalt and zinc with
zincon using PLS algorithm is reported. This ligand is
a good chromogenic reagent [23].

their nitrate salts. A stock zincon solution (0.001 M) in


water was prepared by dissolving solid reagent samples. Universal buffer solutions (pH 310) were prepared by mixing phosphoric, acetic and boric acids.
0.04 M, of each chemical and sufficient amount of
0.2 N NaOH solutions is poured into 100 ml of the
mixture.
2.2. Apparatus
Electronic absorption measurements were carried
out on a CECIL 9000 spectrophotometer (slit width
0.2 nm and scan rate 300 nm/min) using glass or
quartz cells of 1 cm path length. A Metrohm 692
pH-meter furnished with a combined glass-saturated
calomel electrode was used for pH measurements.
The pH-meter was calibrated with at least two buffer
solutions at pH 2.00 and 9.00.
2.3. Computer hardware and software
All absorption spectra were gathered, digitized and
stored from 515 to 750 nm in steps of 1 nm and then
transferred (in ASCII format) to a Pentium 200 MHz
computer for subsequent manipulation by PLS program. The data treatment was done with MATLAB
for windows (Mathworks, version 6.0). PLS program
(for calibrationprediction and experimental design)
of PLS-Toolbox (Eigenvector company); was used.
2.4. Procedure
Known amounts of the standard solutions of each
cation, 2.0 ml of buffer solution were placed in a 10 ml
volumetric flask and completed to final volume with
deionized water (final pH was 8.0). The final concentration of these solutions varied between 0.02.6,
0.04.6, 0.03.0 and 0.04.92 ppm for Cu, Ni, Co and
Zn, respectively. Finally, the spectra of all prepared
solutions were recorded on spectrophotometer.

2. Experimental
3. Results and discussion
2.1. Reagent
All chemical were of analytical-reagent grade and
deionized water was used throughout. Stock solutions
of copper, nickel, cobalt and zinc were prepared from

Fig. 1 shows the influence of the pH of the medium


on the absorption spectra of metal complexes were
studied over the pH range 3.010.0. Using NAS based
relations [13,14], sensitivity and selectivity for each

J. Ghasemi et al. / Analytica Chimica Acta 487 (2003) 181188

183

Fig. 1. Absorption spectra of the zincon and copper, nickel, cobalt and zinc complexes in different pH-value. Concentration of zincon is
1.168 105 and each ion is 1 ppm, ( ) zincon, (---) Cu, () Ni, (+) Co, () Zn.

analyte at different pH values were calculated Fig. 2.


Therefore, pH 8.0 was selected as the optimum value,
to compromise the sensitivity and selectivity of all four
metal ions.
The effect of zincon concentration was also investigated, a reagent concentration of 1.6 104 M was
chosen because it ensures a sufficient reagent excess.

3.1. One component calibration


In order to find the linear dynamic range of concentration of each cation, one component calibration was
performed for each element. For the preparation of
zincon solution (0.001 M), appropriate amount added
to 25 ml universal buffer [22] in a 100 ml volumetric

184

J. Ghasemi et al. / Analytica Chimica Acta 487 (2003) 181188

flask and diluted to the mark with distilled water. Different volumes of 100 ppm solution of copper were
added to 5.0 ml of zincon solution at pH = 8 in a
10.0 ml volumetric flask and diluted to the mark with
distilled water. After 30 min, the absorbance was read
at the 600 nm. The same procedure was followed for
nickel, cobalt, copper and zinc and the absorbance
of the solutions were read at 665, 656 and 630 nm,
respectively. Linear regression results, line equations
and R2 are shown on the Fig. 3.
3.2. Calibration and test mixtures

Fig. 2. Plot of the sensitivity and selectivity versus pH for the


zincon and four complexes: () zincon, () Cu, () Ni, () Co,
() Zn.

Two sets of standard solutions were prepared. The


calibration set contains 35 standard solutions. The
compositions of the calibration mixtures were selected
according to a (4, 4) simplex lattice design [12,13]. For
model assessment, it was used of seven test mixtures.
The concentration of each cation solution was in the
linear dynamic range of the cation, for the preparation
of each solution, different volumes of four cation solutions (20 ppm) were added to 5.0 ml zincon solution
(pH = 8) in a 10 ml volumetric flask. After 30 min,

Fig. 3. Analytical curve for univariate determination of copper, nickel, cobalt and zinc complexes.

J. Ghasemi et al. / Analytica Chimica Acta 487 (2003) 181188


Table 1
The 35 designed experimental
No.

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35

Concentration (ppm)
Cu

Ni

Co

Zn

2.60
1.95
1.95
1.95
1.30
1.30
1.30
1.30
1.30
1.30
0.65
0.65
0.65
0.65
0.65
0.65
0.65
0.65
0.65
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
1.15

0.00
1.15
0.00
0.00
2.30
1.15
1.15
0.00
0.00
0.00
3.45
2.30
2.30
1.15
1.15
0.00
0.00
0.00
0.00
4.60
3.45
3.45
2.30
2.30
2.30
1.15
1.15
1.15
1.15
0.00
0.00
0.00
0.00
0.00
1.15

0.00
0.00
0.75
0.00
0.00
0.75
0.00
1.50
0.75
0.00
0.00
0.75
0.00
1.50
0.00
3.00
1.50
0.75
0.00
0.00
0.75
0.00
1.50
0.75
0.00
2.25
1.50
0.75
0.00
3.00
2.25
1.50
0.75
0.00
0.75

0.00
0.00
0.00
1.23
0.00
0.00
1.23
0.00
1.23
2.46
0.00
0.00
1.23
0.00
2.46
0.00
1.23
2.46
3.69
0.00
0.00
1.23
0.00
1.23
2.46
0.00
1.23
2.46
3.69
0.00
1.23
2.46
3.69
4.92
1.23

absorption spectra of the mixtures recorded. The


calibration matrix used for the analysis is shown in
Table 1.
3.3. Selection of optimum number of factors
To select the number of factors in PLS algorithm,
in order to model the system without over fitting the
concentration data, a cross-validation method, leaving out one sample at a time, was used [17]. Given
the set of 35 calibration spectra, the PLS calibration
on 34 spectra were performed, and using this calibra-

185

tion the concentration of the compounds in the sample


left out during calibration was predicted. This process
was repeated 35 times until each calibration sample
had been left out once. The predicted concentration
of the compounds in each sample was compared with
the known concentration of the compound in this reference sample and prediction residual error sum of
squares (PRESS) was calculated. The PRESS was calculated in the same manner each time a new factor is
added to the PLS model.
The maximum number of factors used to calculate
the optimum PRESS was selected 18 (half the number of standard plus one). One reasonable choice for
the optimum number of factors would be that number
which yielded the minimum PRESS. However, using
the number of factors (h ) that yields a minimum in
PRESS usually lead to some over fitting. A better criterion for selecting the optimum number of factors involves the comparison of PRESS from model is not
significantly greater than PRESS from the model with
h factors. The F statistic was used to make the significance determination.
Haaland and Thomas [17] empirically determined
that an F-ratio probability of 0.75 is a good choice.
We selected as the optimum the number of factors for
the first PRESS values the F-ratio probability, which
drops below 0.75, Fig. 4. The PRESS values are minimum in the number of 6, 5, 8 and 5 for Cu, Ni, Co
and Zn, respectively, then these numbers of factors
are selected as optimum for the calibration model. In
Fig. 4, the PRESS obtained by optimizing the calibration matrix of the absorbance data with PLS method is
shown.
The results obtained by applying PLS algorithm to
the seven prediction set samples are listed in Table 2.
The plots of these predicted concentrations versus
actual concentration using the optimum model are
shown in Fig. 5. The correlation coefficients are
0.9428, 0.9849, 0.9680 and 0.9918 for Cu, Ni, Co and
Zn, respectively, which again verify the good performance of PLS model in predicting the concentrations
of cations in mixture solutions.
3.4. Statistical parameters
For the constructed model, three parameters were
selected to test the prediction ability of the model for
simultaneous determination of Cu, Ni, Co and Zn. The

186

J. Ghasemi et al. / Analytica Chimica Acta 487 (2003) 181188

Fig. 4. Plot of PRESS vs. number of significant factors: () Ni, () Zn, () Co, () Cu.

RMSD, the square of the correlation coefficient (R2 )


and the relative error of prediction (REP), calculated
for each component as follows:
 n
0.5
1
2
RMSD =
(xi xi )
n
i=1

n
(xi x i )2
R2 = i=1
n
i )2
i=1 (xi x
0.5
 n
100 1 
2
REP(%) =
(xi xi )
x
n

the analyte in the sample i, x i the mean of true concentration in the prediction set and n is the total number
sample used in the prediction sets. The concentration
data, the predicted values and their relative errors of
the prediction set are shown in Table 2. The value of
RMSD, REP, R2 , number of factors and PRESS according the values listed in Table 2 are summarized in
Table 3.
3.5. Effect of foreign ions

i=1

where xi is the true concentration of the analyte in the


sample i, x i represented the estimated concentration of

The interference due to several cations and anions was studied in detail. For these studies different
amounts of the ionic species were added to a mixture of Cu, Ni, Co and Zn containing 0.9, 1.5, 0.58

Table 2
Prediction set composition, predicted values and relative errors
No.

1
2
3
4
5
6
7

Actual value (ppm)

Predicted value (ppm)

Relative error (%)

Cu

Ni

Co

Zn

Cu

Ni

Co

Zn

Cu

Ni

Co

0.66
0.50
0.90
0.60
0.80
0.38
0.00

0.80
1.30
1.50
1.40
1.50
0.00
4.00

0.80
0.90
0.58
0.70
0.60
0.00
0.00

1.50
1.10
0.70
1.14
0.82
4.20
0.64

0.63
0.47
0.82
0.56
0.73
0.37
0.05

0.86
1.31
1.34
1.40
1.64
0.15
4.21

0.88
0.97
0.66
0.76
0.64
0.09
0.11

1.74
1.26
0.70
1.26
0.87
4.13
0.65

4.54
6.00
8.88
6.66
8.75
2.63

7.50
0.76
10.66
0.00
9.33

5.25

11.25
8.88
13.79
8.57
6.66

Zn
16.00
14.54
0.00
10.52
6.09
1.66
1.56

J. Ghasemi et al. / Analytica Chimica Acta 487 (2003) 181188

187

Fig. 5. Plots of predicted concentration vs. actual concentration for four cations in the prediction set.

and 0.7 ppm, respectively. The starting point was a


2000 ppm of interference to metal ions, and if the interference occurred the concentration of interference
was progressively reduced until interference ceased.
The tolerance limits were taken as those concentrations causing changes no greater than %5 in the
concentration of analytes. The tolerance limits are
depicted in Table 4.
3.6. Determination of Cu, Ni, Co and Zn in real
samples
In order to test the applicability and matrix interference of the proposed method to the analysis of
real matrix samples, the method was applied in a

Table 3
Statistical parameters of the test matrix using the PLS model
Cations

No. of
factors

RMSD

R2

REP (%)

Cu2+
Ni2+
Co2+
Zn2+

6
5
8
5

0.018
0.049
0.031
0.045

0.94
0.99
0.97
0.99

3.29
3.25
6.05
3.21

variety of synthetic solutions. Permute, alloy sample


and red brass composition like solution were prepared and analyzed as described under experimental. The results of the prediction are summarized in
Table 5.
It can be seen that good recovery values are obtained
by the PLS method using absorbance data. These
satisfactory results indicate that the method would
be effective for the analysis of samples of similar
complexity.

Table 4
Effect of various ions on the determination of four metals
Concentration
(ppm)

Foreign ions

>2000
1000
500
400
20
10
5
2
1
<1

F , Cl , Br , I , CO3 2 , CH3 COO , NO3


Li(I), Na(I), K(I), Tl(I), Rb(I), Mg(II)
IO4 , BrO3 , S2 O3 2
SCN , Tartarate, ClO4 2 SO4 2
Mo(IV), Sr(II)
Ba(II), Cd(II)
Pb(II)
Ag(I)
Mn(II), Fe(III)
Zr(IV), Al(III), Hg(II)

188

J. Ghasemi et al. / Analytica Chimica Acta 487 (2003) 181188

Table 5
Some industrial alloys analysis
Sample

M1a
M2a
M3a

Predicted value (ppm)

Cu2+

Ni2+

Co2+

(1.48),
(0.64),
(0.91)
Cu2+ (2.03), Ni2+ (0.30), Zn2+ (0.78)
Cu2+ (2.12), Zn2+ (0.94)

Recovery (%)

Cu

Ni

Co

Zn

Cu

Ni

Co

Zn

1.41
1.92
2.06

0.70
0.44
0.21

0.93
0.09
0.00

0.01
0.75
0.92

95.11
94.43
97.22

109.44
110

102.00

95.76
97.35

The number in parenthesis indicates the concentration of element taken for analysis in ppm.
a M1, M2 and M3 are permute, alloy sample and red brass, respectively.

4. Conclusions
The coppernickelcobaltzinc mixture is an extremely difficult complex system due to the high
spectral overlapping observed between the absorption
spectra for these components. PLS modeling was
established, with good prediction ability in the real
matrix samples. Results show that partial least square
is an excellent calibration method to determination of
these metals together without pretreatment in complex
samples.
References
[1]
[2]
[3]
[4]

O.W. Lau, S.Y. Ho, Anal. Chim. Acta 280 (2) (1993) 269.
V. Rigin, Anal. Chim. Acta 283 (2) (1993) 895.
Z. Shen, Z. Wang, Fensi. Huaxue 21 (11) (1993) 1313.
N. Xie, C. Huang, H.D. Fu, Sepu 8 (2) (1990) 114;
N. Xie, C. Huang, H.D. Pu, Chem. Abstr. 113 (1990) 094170.
[5] Z. Shen, F. Nie, Y. Chem, Fensi Huaxane 19 (11) (1991)
1272.

[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]

R.M. Brotheridge, et al., Analyst 123 (19998) 69.


J.H. Kalivas, Anal. Chim. Acta 428 (2001) 31.
J. Ghasemi, A. Niazi, A. Safavi, Anal. Lett. 34 (2001) 1389.
J. Ghasemi, A. Niazi, Microchem. J. 71 (2002) 1.
J. Ghasemi, R. Amini, A. Niazi, Anal. Lett. 35 (2002) 533.
A.M. Garcia Rodriguez, A. Garcia de Torres, J.M. Cano
Pavon, C. Bosch Ojeda, Talanta 47 (1998) 463.
J.A. Corenl, Experiment with Mixture, Wiley, New York,
1981.
I. Duran, M. Arsenio, M. dela Pena, A.E. Mansila, F. Salinas,
Analyst 118 (1993) 807.
K.S. Booksh, B.R. Kowalski, Anal. Chem. 66 (1994) 782A.
A. Lorber, Anal. Chem. 58 (1986) 1167.
H. Martens, T. Naes, Multivariate Calibration, Wiley,
Chichester, 1989.
D.M. Halaand, E.V. Thomas, Anal. Chem. 60 (1988) 1193.
K.R. Beebe, B.R. Kowalski, Anal. Chem. 59 (1987) 1007A.
S. Wold, P. Geladi, K. Esbensen, J. Ochman, J. Chemom. 1
(1987) 41.
S. Wold, et al., Chemom. Intell. Lab. Syst. 58 (2001) 109.
R.W. Gerlach, B.R. Kowalski, H. Wold, Anal. Chim. Acta
112 (1979) 417.
R.M. Rush, J.H. Yoe, Anal. Chem. 26 (1954) 1345.
M. Otto, W. Wegscheider, Anal. Chem. 61 (1989) 1847.

Vous aimerez peut-être aussi