Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Microchemical Journal 164 (2021) 106058

Contents lists available at ScienceDirect

Microchemical Journal
journal homepage: www.elsevier.com/locate/microc

Indirect determination of aluminum (III) in water samples by in-electrode


coulometric titration
Jakub Masac , Jan Lovic , Ernest Beinrohr , Frantisek Cacho *
Institute of Analytical Chemistry, Faculty of Chemical and Food Technology, Slovak University of Technology in Bratislava, Radlinskeho 9, Bratislava 812 37. Slovakia

A R T I C L E I N F O A B S T R A C T

Keywords: A new procedure has been developed for indirect determination of aluminium(III) in water through in-electrode
Aluminum coulometric titration in a porous reticulated vitreous carbon electrode. It was determined through different
Fluoride complexes aluminium fluoride and ferric fluoride complex stabilities. The ferric fluoride complex thereby formed was
Indirect determination
decomposed with Al(III) from the sample and an aliquot of the released Fe(III) was determined electrochemi­
In-electrode coulometric titration
cally. It was measured with an EcaFlow GLP 150 electrochemical flow analyser, while 0.2 mol L− 1 sodium
chloride solution was used as the electrolyte to flush the system. Calibration solutions and samples were modified
in a mixed solution medium of 0.2 mol L− 1 NaCl + 10 mg L− 1F(I-) + 10 mg L− 1Fe(III). All working parameters for
electrochemical determination were optimised. The detection limit and precision were found to be 0.02 mg L− 1
and 1.9% respectively. The linear concentration range was 0.05 – 8.5 mg L− 1 of aluminium. After having been
complexed with sodium cyanide, the iron’s interference effect was suppressed by removing it with an ion-
exchange resin. Among the other influences studied, only the interference from thorium, zirconium and
cerium was significant. The method was applied for the analysis of various water samples. The results were in
good agreement with data from high resolution atomic absorption spectrometry.

1. Introduction (chromatography, extraction, etc.) either to detect trace aluminium


concentrations or to separate and determine the various forms of
Aluminium (Al) is a low density metal and the third most abundant aluminium [6,10–12]. Although these methods offer adequate sensi­
element in the earth’s crust (8%). It is produced industrially from tivity and selectivity, photometric methods for treating samples are
electrolysing molten bauxite and cryolite. Aluminium is used in various often very difficult and time-consuming. Atomic spectrometry methods
industries, with the automotive and aerospace industries taking are problematic because of the high thermal stability of aluminium
advantage of the low density of its compounds. It is often utilised in atoms.
everyday items such as antiperspirants, food foils, kitchen utensils, etc. Direct electrochemical determination of aluminium through its
Despite its high presence in the earth’s crust, aluminium is not part of reduction in water is almost impossible due to the significantly negative
any living tissue. The World Health Organisation (WHO) has indicated a redox potential (− 1.75 V vs. Ag/AgCl), which is close to the potential for
possible link between aluminium and Alzheimer’s disease. The generating hydrogen and reducing alkali metal ions [13]. Anodic
maximum permitted concentration of aluminium in drinking water, stripping voltammetry (ASV) is the most common electrochemical
which does not reduce the water’s sensory quality, is 200 µg L− 1 [1]. method for determination of aluminium. Electrodes with ligand (com­
Spectral methods are most commonly used to determine aluminium plexing agents) treated surfaces are mostly used. The presence of a
in various types of samples. Examples of such methods include graphite ligand forming a complex with aluminium shifts the potential to more
furnace atomic absorption spectrometry (GF-AAS) [2,3] flame atomic positive values, due to the adsorption of the complex on the working
absorption spectrometry (FAAS) [4] inductively coupled plasma atomic electrode. Stripping voltammetry, applied after complexation, is the
emission spectrometry (ICPAES) [5] inductively coupled plasma mass most sensitive method because it allows pre-concentration of the
spectrometry (ICP-MS) [5,6] spectrofluorimetry [6,7] and spectropho­ aluminium on the electrode surface before measurement. The best-
tometry [8,9]. They are often combined with separation techniques known reagents used include Alizarin-S [14,15] Alizarin violet [16]

* Corresponding author.
E-mail address: frantisek.cacho@stuba.sk (F. Cacho).

https://doi.org/10.1016/j.microc.2021.106058
Received 16 September 2020; Received in revised form 23 November 2020; Accepted 25 December 2020
Available online 15 February 2021
0026-265X/© 2021 Elsevier B.V. All rights reserved.
J. Masac et al. Microchemical Journal 164 (2021) 106058

Morin [17,18] and many others. computer automatically calculates the outcome according to Faraday’s
This paper describes a new procedure for indirect electrochemical laws of electrolysis. Individual measurements are recorded by memory
determination of aluminium(III) in water. The determination is possible mapping [21,22]. The graph itself resembles a derived chro­
because of the different stabilities of ferric fluoride and aluminium nopotentiogram in shape, but it is actually recorded by filling the indi­
fluoride complexes and also the different electrochemical properties of vidual channels in the built-in channel counter. Therefore, the displayed
the free ferric ions and the ferric ions bound with the fluoride in the record is not derived from the chronopotentiogram, but from gradually
complex. Here a flow-through laboratory system was used, with in- plotting dt/dE counts depending on the gradual change in the potential.
electrode coulometric titration in a reticulated vitreous carbon Thus, it is primary coulometric thin-film titration and not derivative
electrode. chronopotentiometry, as it might seem. The working electrode (defined
below) used in the system consists of a large number of regularly con­
2. Experimental nected small cavities. These create a thin film of solution above the
surface of the electrode material, with the entire electrode serving as a
The process discussed in this paper is based on the different fluoride single coulometric vessel. Formally, the “counts” have a dimension of
complex stabilities of Al(III) and Fe(III) in solution. A similar principle V− 1, where the measurement always consists of two steps. First, the
was used by Cernanska et al. to determine fluorides in toothpastes [19]. background signal is measured from an analysis of a blank sample before
Electrochemical reduction of Fe(III) on Fe(II) is a well-known process the standard or sample signal in the second step. The background signal
that works on most electrode materials. Adding fluorides to a solution is then subtracted from the sample signal to obtain a clear record with no
containing Fe(III) immediately forms ferric fluoride complexes (pK3 = background influence.
12.06), which are electro-inactive at the reduction potential of free Fe The measurement step involves filling the electrode with the ana­
(III) ions. As fluoride ions are gradually added to the solution, the Fe(III) lysed solution, reducing the Fe(III) to Fe(II) in the electrode’s pores. The
reduction signal decreases to the level of the background signal. If a area of the reduction peak from the obtained record corresponds to the
solution of Al(III) ions is added to such a [Fe(III) + F(I-)] mixture, the time required to reduce Fe(III) in the electrode volume (i.e. the transi­
ferric fluoride complexes decompose and preferential aluminium fluo­ tion time τ from Faraday’s laws of electrolysis). Since the volume is
ride complexes are formed (pK6 = 19.84). The Fe(III) bound in the finite, the size of the peak area corresponds both to the amount of Fe(III)
complex is released into the solution, enabling its reduction on the in the electrode and the concentration of Fe(III) in the analysed solution,
electrode. The detected signal from the reduction of free Fe(III) thus and in this case also to the concentration of Al(III) there, too. Between
increases as the concentration of Al(III) in the analysed solution in­ each measurement, the working electrode and the flow cell are washed
creases. Competing reactions are only possible with thorium and zir­ with the primary electrolyte.
conium ions with their higher pK3 values. However, their occurrence in A commercial compact flow-through electrochemical cell (type 104)
real samples is unlikely. with Pt auxiliary and Ag/AgCl reference electrodes was used (Istran Ltd.,
Slovakia). The cylindrical working electrode was a reticulated vitreous
carbon plug of 100 ppi (pores per inch) porosity (Electrosynthesis Co.
2.1. Instrumentation Inc., Lancaster, New York, USA) of 4 mm and 10 mm in length and
diameter, respectively. All operating parameters required for measure­
All the measurements were performed with an EcaFlow GLP-150 ment are summarised in Table 1. These values were entered into a
electrochemical analyser, manufactured by IstranLtd., Bratislava, control program in the personal computer that controlled the electro­
Slovakia. It is equipped with two solenoid inert valves, a peristaltic chemical analyser. The instrumentation allowed the measurement itself
pump and a computer controlled potentiostat/galvanostat [20]. Fig. 1 to be fully automated. All potential values were reported against the
shows the flow diagram for the flow measuring cell device. All PTFE silver/silver chloride (Ag/AgCl) reference electrode built into the flow
tubes used in the analyser have an internal diameter of 1 mm. The cell [23].
system allows measurements to be made in both calibration and High-resolution continuum source graphite furnace atomic absorp­
calibration-free mode. In calibration mode, results are calculated from tion spectrometry was used as the reference method. A ContraAA HR-CS
the calibration curve. When the calibration-free mode is used, a GFAAS spectrophotometer, MPE 60 auto-sampler and a pyrolytically
coated cuvette with a PIN-platform were all used to determine the
aluminium in the water samples (wavelength: 309.2713 nm, calibration
curve range: 0.5–50 µg L− 1).
The same device was used to monitor the efficiency of the separation
of ferric ions from real water samples (wavelength: 425.0762 nm, cali­
bration curve range: 0.3–20 µg L− 1).

Table 1
Working parameters of the flow-through analyser.
Parameter Value Note

Potential at filling (mV) * 900 Filling the system


Starting potential (mV) * 900 Constant current is applied and potential
monitored
End potential (mV) * 200 Switched to the regeneration potential
Reduction current (µA) − 10
Regeneration potential 500 Regeneration of the electrode
(mV) *
Fig. 1. Block diagram of the flow-through analyser EcaFlow with the Flow- Standby potential (mV) * 100
Flow rate (mL min− 1) 6
through cell: V – solenoid valve, PP – peristaltic pump, 1- working electrode, 2 –
Sample volume (mL) 3
reference electrode, 3 – auxiliary electrode, 4 – membrane, 5 – packing, 6 –
screw, 7 – cell body, 8 – space for reference electrolyte (saturated KCl). *vs. saturated Ag/AgCl reference electrode.

2
J. Masac et al. Microchemical Journal 164 (2021) 106058

2.2. Reagents and solutions the measurement was rapid enough (three minutes per analysis). Table 1
summarises all the parameters for determining Fe(III)- The following
Analytical-reagent grade chemicals were used in all experiments. basic validation parameters were used to determine Fe(III): linear range
The solutions were prepared in demineralized and boiled water of 0.1–10 mg L− 1, LOD = 0.03 mg L− 1, LOQ = 0.1 mg L− 1, relative
(NANOpure, Wilkem Werner GmBH, Germany). standard deviation (RSD) repeatability = 2.3%. The ULA-2 statistical
Mass concentration was used for all the solutions since it is preferred method was used to determine LOD and LOQ [24]. Fig. 2 shows the Fe
over molar concentration. (III) reduction by the measuring device, while the calibration curve for
Aluminium(III) standard: The bulk standard solution of exactly the determination of iron (0.1–10 mg L− 1) is displayed in Fig. 3A. All
1.0000 g L− 1 Al(III) was prepared from aluminium nitrate nonahydrate measurements were made in calibration-free mode.
(.Al(NO3 )3 ⋅9H2 O, ⩾99.997\% trace metals basis, Sigma-Aldrich) in
water. 3.2. Determination of fluorides
Iron(III) standard: The bulk standard solution of exactly 1.0000 g L− 1
Fe(III) was prepared from iron(III) nitrate nonahydrate To determine the most suitable fluoride concentration, a set of so­
(Fe(NO3 )3 ⋅9H2 O, ⩾99.999\% trace metals basis, Sigma-Aldrich) in lutions was prepared with a sodium chloride concentration of 0.2 mol
water. L− 1 + 10 mg L− 1Fe(III) and F(I-) concentrations in the range of 0.1–15
Fluoride standard: Fluoride standard for IC (TraceCERT®, 1000 mg/ mg L− 1. After mixing, the solutions were immediately analysed ac­
L fluoride in water, Sigma-Aldrich) was used. cording to the conditions for determining free Fe(III) in the calibration-
KCN: The bulk solution of 0.0054 mol L− 1 KCN was prepared from free measuring mode. As the concentration of F(I-) increased in solution,
KCN(KCN, BioUltra, ≥98.0%, Sigma-Aldrich) in water. the measured signal of free Fe(III) decreased linearly at a range of
Carrier electrolyte: 0.2 mol L− 1 NaCl was prepared from NaCl(NaCl, 0.1–10 mg of L− 1fluorides. Fluorides do not react in an aqueous medium
ACS grade, Sigma-Aldrich) with Fe(III) ions at a 1:1 ratio by mass, which was why the measured
HNO3 (Suprapur, 65%, Merck KGaA, Darmstadt, Germany) signal did not drop to the zero limit, even in the described case, after the
Preparation of standards for calibration: A solution (designated A1) solution of 10 mg of L− 1 of Fe(III) and 10 mg of L− 1 of fluorine(I-) had
was prepared from 2 mol L− 1 NaCl + 100 mg L− 1Fe(III) + 100 mg L− 1F been prepared. When the concentration of fluoride rose further, the free
(I-). Five millilitres of A1 with the required amount of standard Al(III) Fe(III) signal in the solution continued to gradually decrease. But this
was added into a fifty-millilitre volumetric flask, which was afterward decline showed a non-linear dependence. Fig. 3B displays the wave­
filled with water. forms for the decrease in the concentration of free Fe(III) in the solution,
Samples: 10 samples were analysed of underground, surface, drink­ depending on the added fluorides (up to a concentration of 15 mg L− 1F
ing and waste water gathered across Slovakia. The pH of every sample (I-)). At this point, the described procedure can obviously be used to
was found to be approximately 7. indirectly determine soluble fluorides in water. Unusually, the calibra­
tion dependence would have a negative slope in this case.
3. Result and discussion
3.3. Determination of Al(III)
The first step was to optimise and validate the determination of Fe
(III) using EcaFlow. The optimal concentration of fluoride was deter­ A solution comprised of 0.2 mol L− 1 NaCl + 10 mg L− 1Fe(III) + 10
mined in the second step. The procedure for determining Al(III) was mg L− 1F(I-) was selected as the basic working solution for Al(III) addi­
validated in the third step and the process for removing iron from the tions. In this solution, the concentration of free Fe(III) decreased linearly
real samples was tested. LOD, LOQ, linear range, and the repeatability of after complexation with the fluorides. The first assumption was that the
the determination were defined and several types of water samples were addition of even a minimal amount of free Al(III) would displace the Fe
analysed. The samples were subsequently analysed again using the (III) from the Fe(III)-fluoride complex to form an Al(III)-fluoride com­
reference method. plex. The second assumption was that a gradual linear increase in the
concentration of Al(III) in a solution would cause the concentration of
3.1. Determination of Fe(III) free Fe(III) in the solution to increase linearly, too. To verify these as­
sumptions, a set of solutions was prepared containing 0.2 mol L− 1 NaCl
Free Fe(III) was determined from its reduction to Fe(II) on the sur­ + 10 mg L− 1Fe(III) + 10 mg L− 1F(I-), adding Al(III) in a concentration
face of the working electrode. The effect from the electrolyte’s pH is very
important for the determination. For best results, an electrolyte with a
pH in the range of 6–7 is appropriate (although some authors recom­
mend a pH of 5–7 [19]). A pH lower than the range reduces the
complexation efficiency of fluorides with Fe(III) and Al(III), while the
intensity of the measured free Fe(III) reduction signals drops at a higher
pH (of 8 and more). An aqueous NaCl solution was selected as the
suitable electrolyte because of its neutral pH and sufficient ionic
strength. The concentration of this electrolyte was optimised at the
range of 0.1 – 1 mol L− 1. The maximum signal intensity was observed
from a sodium chloride (NaCl) concentration of 0.2 mol L− 1. The
increased NaCl concentration in the electrolyte no longer increased the
signal, therefore an aqueous solution of NaCl with a concentration of 0.2
mol L− 1 was used as the electrolyte. In addition, the current required to
reduce Fe(III) to Fe(II) on the surface of the working electrode was
optimised to determine Fe(III). The reduction current ranged between
− 5 and 50 µA. As the current shifted toward zero, the intensity of the
measured signals (reduction time, i.e. transient time τ) gradually
increased. A current of − 10 µA was accepted as the optimal value.
Intense signals were obtained at this value while the repeatability of the Fig. 2. Chronopotentiometric signals of a standard with 1 mg L− 1
Fe(III). Pa­
measurements had a satisfactory metrological quality (RSD = 2.3%) and rameters listed in Table 1.

3
J. Masac et al. Microchemical Journal 164 (2021) 106058

Fig. 3. A- Concentration dependence of the Iron(III) signal. Regression line: y = 16.7305x + 0.8025; R2 = 0.9997. B- Dependence of free Fe(III) concentration in a
solution on fluoride concentration. Regression line: y = -13.8957x + 171.0011; R2 = 0.9998. End of the linear decrease in the signal indicated at a concentration of
10 mg L-1F(I-). C- Dependence of the concentration of free Fe(III) in a solution for different concentrations of Al(III) and a different concentration of fluorides. ■
Regression line: y = 101,2466 + 14,04937x, R2 = 0,9997/● Regression line: y = 32,4754 + 15,9965x, R2 = 0,9998/▴ Regression line: y = 1,5541 + 17,6668x, R2
= 0,9998.

ranging from 0.1 to 10 mg L− 1. An electrochemical analyser in 3 mol per litre of sodium chloride. During the recycling, the cyanide
calibration-free mode measured the signal from free Fe (III) in all of the eluate was collected in a KMnO4 solution.
solutions. As Fig. 3C shows, the free Fe(III) signal increased linearly in WARNING: KCN is a lethal poison. Whenever working with it, the
the range of 0.1 – 8.5 mg L− 1 according to the assumptions. For the necessary safety precautions should be taken! The procedure is only
completeness of outcomes, two more identical experiments were per­ applicable to waters with a neutral pH.
formed with fluoride concentrations of 5 mg L− 1 and 15 mg L− 1. In the Preparation of the blank sample: Five millilitres of the A1 solution
first case [10 mg L− 1Fe(III) + 5 mg L− 1F(I -)], a linear increase in the was diluted with water to a volume of 50 ml in a volumetric flask and
signal of free Fe(III) was observed as Al(III) was gradually added to the used as the blank.
solutions. The linear range of the calibration curve was only 0.1 to 4.5 Preparation of samples for HR-CS GFAAS: Concentrated nitric acid
mg L− 1 Al(III). The direction of the calibration curve did not change (50 µL per 50 mL sample) was added to the samples prior to analysis.
compared to the system [10 mg L− 1Fe(III) + 10 mg L− 1F(I-)]. It only
shortened the linear range of the calibration curve. In the second case 3.4. Analytical figures of merit
[10 mg L− 1Fe(III) + 15 mg L− 1F(I-)], the dependence of the signal from
the free Fe(III) in the solution on the concentration of added Al(III) was Table 2 lists the main metrological determination parameters. The
divided into two parts. The dependence was nonlinear at the concen­ limit of detection (LOD) and limit of quantification (LOQ) were calcu­
tration range of 0.1 – 1.5 mg L− 1 Al(III). The linear increase in the signal lated from the calibration dependence using the ULA-2 method [24].
could be observed at the concentration range of 1.5 – 9.5 mg L− 1 Al(III). The repeatability of the measurement was determined from making
In certain circumstances, such an arrangement could be used to deter­ twenty measurements in rapid succession of the same solution at an
mine Al(III) concentrations in a sample, but only to determine higher aluminium concentration of one milligram per litre. The duration of the
concentrations and just at a relatively narrow range of them. These measurement corresponds to the analysis itself (no sample treatment).
experiments confirmed [10 mg L− 1Fe(III) + 10 mg L− 1F (I-)] to be the The effect of selected ions commonly found in water was studied. No
model most suitable for indirect electrochemical determination of Al(III) effect on the determination outcome was detected from ions of Na(I), K
in water. Fig. 3C shows the measured signal waveforms from the Al(III) (I), Ca(II), Mg(II), Zn(II), Cu(II) and Ni(II) even in 100-fold excess. The
concentration plotted for all three monitored systems. The correspond­ anion-exchange resin removed all anions (sulphates, nitrates, nitrites,
ing equations for the calibration lines of the linear areas are also given. chlorides, bromides and bicarbonates) from the samples. Only Th(IV), Zr
All of the experiments were repeated three times. (IV) and Ce(IV) ions increased the analytical signal by 15%, even in a
Preparation of samples for analysis: During determination, it was twofold excess, due to competing reactions. The added fluorides formed
crucial to control the concentration of iron in the sample. All iron was more stable complexes with these ions than with aluminium. Notwith­
first removed from the analysed samples and then an exactly known standing, their occurrence in real water samples is unlikely.
amount of Fe(III) was added. A cyanide process was used to remove the
iron. Iron concentrations in the range of 0.5 – 10 mg per litre are com­ 3.5. Analysis of real samples
mon in water, but concentrations of 50 mg L− 1 have also been reported
[25]. In our samples, the iron concentration was determined by the HR- An analysis of ten real samples verified the method’s accuracy. All of
CS-GFAAS method at the range of 0.5 – 5 mg L− 1. Fifty millilitres of KCN the samples were treated as described to remove the iron. The results
solution at a concentration of 0.0054 mol L− 1 (about 350 mg L− 1, suf­ obtained from the calibration curve and standard addition correlated
ficient for a concentration of 5 mg per litre iron) was added to 50 ml of between each other and with the reference values obtained from HR-CS-
the sample. All the iron in the solution was bound to the [Fe(CN)6 ]3− GFAAS (p = 0.05), according to Student’s t-test for a 95% confidence
complex. The solution then flowed through a column made of Dowex® interval. Table 3 shows the outcome of the analyses and the reference
1-X8 (a strongly basic anion exchanger in Cl− form by Merck). Both the
iron (as an anionic complex) and excess cyanides were removed from the Table 2
solution. Aluminium does not complex with cyanide, so it flowed Metrological data.
through the column unchanged. The central part of the sample eluate Parameter Value
was then collected in a glass beaker. It was then added to a 50-millilitre Linear range (mg L− 1) 0.05–8.5
volumetric flask, to which five millilitres of the A1 solution had been Limit of detection (mg L− 1) 0.02
pipetted. Afterward, the solution was analysed. The HR-CS-GFAAS Limit of quantitation (mg L− 1) 0.05
confirmed the iron in all samples to have been completely removed Repeatability at 1 mg L− 1, n = 20, [RSD %] 2.5
Duration of a measurement (min) 3
from the remaining eluates. The anion-exchange resin was recycled with

4
J. Masac et al. Microchemical Journal 164 (2021) 106058

Table 3 References
Analysis of water samples.
[1] WHO, Aluminium in drinking-water, Background document for development of
Sample Linear calibration Standard addition Reference value
WHO guidelines for drinking-water quality, WHO/HSE/WSH/10.01/13, (2010).
(µg L− 1) (µg L− 1) (µg L− 1) [2] G. Dravecz, L. Bencs, D. Beke, A. Gali, Determination of silicon and aluminum in
1. Drinking water ≤ LOD ≤ LOD 15.3 ± 0.3 silicon carbide nanocrystals by high-resolution continuum source graphite furnace
2. Drinking water ≤ LOD ≤ LOD 9.8 ± 0.2 atomic absorption spectrometry, Talanta 147 (2016) 271–275, https://doi.org/
10.1016/j.talanta.2015.09.067.
3. Waste water 29.0 ± 0.5* 29.2 ± 0.8* 29.2 ± 0.7*
[3] L.A. Pereira, I. Gonçalves de Amorim, J.B. Borba da Silva, Development of
4. Waste water 22.6 ± 0.5* 22.7 ± 0.9* 22.6 ± 0.8*
methodologies to determine aluminum, cadmium, chromium and lead in drinking
5. Waste water*2 134.0 ± 3.0* 133.7 ± 4.6* 133.7 ± 2.4* water by ET AAS using permanent modifiers, Talanta 64 (2) (2004) 395–400,
6. Groundwater 89.1 ± 2.4 88.3 ± 3.0 90.5 ± 1.8 https://doi.org/10.1016/j.talanta.2004.02.026.
7. Groundwater 112.4 ± 3.8 110.9 ± 4 112.9 ± 2.1 [4] A. Safavi, S. Momeni, N. Saghir, Efficient preconcentration and determination of
8. Surface water 240.5 ± 3.6 141.3 ± 4.7 142.0 ± 2.7 traces of aluminum ion using silica-bonded glycerol sorbent, J. Hazard. Mat. 162
9. Surface water 190.1 ± 4.3 189.8 ± 5.5 189.7 ± 2.8 (1) (2009) 333–337, https://doi.org/10.1016/j.jhazmat.2008.05.044.
10. Surface water 304.8 ± 7.5 305.1 ± 7.7 305.1 ± 4.8 [5] M. Frankowski, A. Zioła-Frankowska, I. Kurzyca, K. Novotný, T. Vaculovič,
V. Kanický, M. Siepak, J. Siepak, Determination of aluminium in groundwater
*Dimension is mg L− 1. samples by GF-AAS, ICP-AES, ICP-MS and modelling of inorganic aluminium
*Neutralized waste water from aluminum production. complexes, Environ. Monit. Assess. 182 (1-4) (2011) 71–84, https://doi.org/
10.1007/s10661-010-1859-8.
[6] B. Fairman, A. Sanz-Medel, P. Jones, E.H. Evans, Comparison of fluorimetric and
values. Mean values and confidence intervals (95%) were calculated inductively coupled plasma mass spectrometry detection systems for the
from five measurements. The standard deviations were used to deter­ determination of aluminium species in waters by high-performance liquid
mine confidence intervals. The regression equation of the calibration chromatography, Analyst 123 (1998) 699–703, https://doi.org/10.1039/
A707999K.
graph was τ = 15.997c + 32.475, coefficient of determination R2 = [7] S.M.Z. Al-Kindy, A. Al-Hinai, N.K. Al-Rasbi, F.E.O. Suliman, H.J. Al-Lawati,
0.9998, where τ is the measured transition time in seconds and c denotes Spectrofluorimetric determination of aluminium in water samples using N-((2-
the Al(III) concentrations in mg L− 1. hydroxynaphthalen-1-yl)methylene) acetylhydrazide, J. Taibah Univ. Sci. 9 (4)
(2015) 601–609, https://doi.org/10.1016/j.jtusci.2015.03.009.
[8] P. Norfun, T. Pojanakaroon, S. Liawraungrath, Reverse flow injection
4. Conclusions spectrophotometric for determination of aluminium (III), Talanta 82 (1) (2010)
202–207, https://doi.org/10.1016/j.talanta.2010.04.019.
[9] O.D. Renedo, A.M. Navarro-Cuñado, E. Ventas-Romay, M.A. Alonso-Lomillo,
This paper describes a new alternative method for the indirect Talanta 196 (2019) 131–136, https://doi.org/10.1016/j.talanta.2018.12.048.
electrochemical determination of aluminium in water. The unique [10] M. Frankowski, Simultaneous determination of aluminium, aluminium fluoride
process enables an electrochemically inactive element to be determined complexes and iron in groundwater samples by new HPLC–UVVIS method,
Microchem. J. 101 (2012) 80–86, https://doi.org/10.1016/j.microc.2011.11.002.
in different water samples. It is suitable for all electroanalytical systems. [11] P. Matúš, I. Hagarová, M. Bujdoš, P. Diviš, J. Kubová, Determination of trace
Working electrodes made of reticulated vitreous carbon are adequate for amounts of total dissolved cationic aluminium species in environmental samples by
measurement with no further treatment required. Since these electrodes solid phase extraction using nanometer-sized titanium dioxide and atomic
spectrometry techniques, J. Inorg. Biochem. 103 (11) (2009) 1473–1479, https://
have a long life, it is a very inexpensive analysis. RVC electrodes have
doi.org/10.1016/j.jinorgbio.2009.08.004.
proven quite suitable for the analysis. Notwithstanding, there is no [12] P. Matúš, Evaluation of separation and determination of phytoavailable and
known reason for the process not to work with other carbon electrodes. phytotoxic aluminium species fractions in soil, sediment and water samples by five
The disadvantage of the process is the requirement to control the con­ different methods, J. Inorg. Biochem. 101 (9) (2007) 1214–1223, https://doi.org/
10.1016/j.jinorgbio.2007.06.014.
centration of Fe(III) in the samples. The use of potassium cyanide (KCN) [13] S. Thomas, D. Davey, D. Mulcahy, C.K. Chow, Determination of aluminum by
to completely remove the iron from the original samples was verified. adsorptive cathodic stripping voltammetry with 1,2-dihydroxyanthraquinone-3-
Where necessary, the KCN concentration can be adjusted according to sulfonic acid (DASA): effect of thin mercury film electrode, Electroanalysis 18 (22)
(2006) 2257–2262, https://doi.org/10.1002/(ISSN)1521-410910.1002/elan.v18:
the iron concentration in the samples. Of course, this does not rule out 2210.1002/elan.200503629.
any other suitable method for removing the ions. The determination [14] J. Zuziak, M. Jakubowska, Voltammetric determination of aluminum-Alizarin S
itself is very fast, the solutions easy to prepare and, as confirmed, the complex by renewable silver amalgam electrode in river and waste waters,
J. Electroanal. Chem. 794 (2017) 49–57, https://doi.org/10.1016/j.
process is applicable to different types of water samples. jelechem.2017.04.009.
[15] J. Zuziak, W. Reczyński, B. Baś, M. Jakubowska, Voltammetric determination of
CRediT authorship contribution statement aluminum(III) as Al-Alizarin S complex in tea leaves and infusions, Anal. Biochem.
558 (2018) 69–79, https://doi.org/10.1016/j.ab.2018.08.008.
[16] P. Deng, J. Fei, J. Zhang, Y. Feng Determination of trace aluminum by anodic
Jakub Masac: Data curation, Writing - original draft. Jan Lovic: adsorptive stripping voltammetry using a multi-walled carbon nanotube modified
Data curation. Ernest Beinrohr: Supervision. Frantisek Cacho: Visu­ carbon paste electrode Anal. Lett., 44 (2011), pp. 1521-1535, https://doi.org/
10.1080/00032719.2010.520382.
alization, Writing - review & editing.
[17] O. Domínguez-Renedo, A.M. Navarro-Cuñado, E. Ventas-Romay, M.A. Alonso-
Lomillo, Determination of aluminium using different techniques based on the Al
Declaration of Competing Interest (III)-morin complex, Talanta 196 (2019) 131–136, https://doi.org/10.1016/j.
talanta.2018.12.048.
[18] A. Alonso-Mateos, M.J. Almendral-Parra, Y. Curto-Serrano, F.J. Rodriguez-Martin,
The authors declare that they have no known competing financial Online monitoring of aluminium in drinking water with fluorimetric detection, J.
interests or personal relationships that could have appeared to influence Fluoresc., 18 (2008), pp. 183-192 https://doi.org/10.1007/s10895-007-0262-5.
the work reported in this paper. [19] M. Čerňanská, P. Tomčík, Z. Jánošíková, M. Rievaj, D. Bustin, Indirect
voltammetric detection of fluoride ions in toothpaste on a comb-shaped
interdigitated microelectrode array, Talanta 83 (5) (2011) 1472–1475, https://doi.
Acknowledgements org/10.1016/j.talanta.2010.11.026.
[20] E. Beinrohr, M. Cakrt, J. Dzurov, L. Jurica, J.A.C. Broekaert, Simultaneous
calibrationless determination of zinc, cadmium, lead, and copper by flow- through
This work was supported by the Slovak Research and Development stripping chronopotentiometry, Electroanalysis 11 (15) (1999) 1137–1144,
Agency under the contract No. APVV-15-0355 and by the grant agency https://doi.org/10.1002/(SICI)1521-4109(199911)11:15<1137::AID-
VEGA (project No. 1/0159/20) and by the Competence Center for ELAN1137>3.0.CO;2-Z.
[21] J. Mortensen, E. Ouziel, H.J. Skov, L. Kryger, Multiple- scanning potentiometric
SMART Technologies for Electronics and Informatics Systems and Ser­ stripping analysis, Anal. Chim. Acta 112 (3) (1979) 297–312, https://doi.org/
vices, ITMS 26240220072 and by STU Grant Scheme for Support of 10.1016/S0003-2670(01)83557-1.
Young Researchers (project ALMAS). [22] A. Hu, R.E. Dessy, A. Graneli, Potentiometric stripping with matrix exchange
techniques in flow injection analysis of heavy metals in groundwaters, Anal. Chem.
55 (2) (1983) 320–328, https://doi.org/10.1021/ac00253a031.
[23] E. Beinrohr, M. Németh, P. Tschöpel, G. Tölg, Design and characterization of flow-
through coulometric cells with porous working electrodes made of crushed

5
J. Masac et al. Microchemical Journal 164 (2021) 106058

vitreous carbon, Fresenius J Anal. Chem. 343 (7) (1992) 566–575, https://doi.org/ Quantification: Application to Voltammetric and Stripping Techniques, Pure Appl.
10.1007/BF00324817. Chem. 69 (1997) 297-328 10.1351/pac199769020297.
[24] J. Mocak, A.M. Bond, S. Mitchell, G. Scollary, A Statistical Overview of Standard [25] WHO/SDE/WSH/03.04/08: Iron in Drinking-water.
(IUPAC and ACS) and New Procedures for Determining the Limits of Detection and

You might also like