You are on page 1of 8

Chem. Eng. Technol. 2009, 32, No.

12, 1901–1908 1901

José R. Parga1 Research Article


Víctor Vázquez1
Hector M. Casillas1 Cyanide Detoxification of Mining
Jesús L. Valenzuela2
Wastewaters with TiO2 Nanoparticles
1
Department of Metallurgy and
Materials Science, Institute of
and Its Recovery by Electrocoagulation
Technology of Saltillo, Saltillo
Coahuila, Mexico. Due to the widespread use of cyanide in mining operations, its recovery and de-
2
Department of Chemical struction is important for both the environmental aspects of wastewater and its
Engineering + Metallurgy, treatment, and the economic aspects associated with the high consumption of
University of Sonora, chemicals by the process itself. A photoelectrocatalytic detoxification technique
Hermosillo Sonora, Mexico. with titanium dioxide microelectrodes is one of the most innovative ways for the
treatment of wastewater containing cyanide. However, this technique has a disad-
vantage for industrial application in that the separation of titanium dioxide after
the photocatalytic degradation of cyanide is rather difficult due to the fineness of
the particles, and therefore, the reuse of the titanium dioxide has not been at-
tained for the treatment of cyanide-containing wastewater. To overcome this weak
point, an electrocoagulation (EC) technique is used to recover the titanium diox-
ide from its aqueous suspensions. The results show that photodegradation of
cyanide is 93 % in 30 min using a 450 W halogen lamp. The recovery of anatase
with the EC process is 98 %. The results indicate that this technique has the
potential to serve as a reliable and economical method because sunlight can be
used efficiently as the power source. The Langmuir isotherm is used to obtain the
thermodynamic parameters, i.e., free energy, enthalpy and entropy. The evalua-
tion of these parameters, i.e., DG° = –37 kJ/mol, DH° = –54 kJ/mol and DS° =
0.524 kJ/mol K, indicates the spontaneous and exothermic nature of the adsorp-
tion of the anatase particles on the iron species.

Keywords: Adsorption, Coagulation, Cyanides, Electrochemical methods, Nanoparticles,


Wastewater
Received: April 23, 2009; revised: June 12, 2009; accepted: July 16, 2009
DOI: 10.1002/ceat.200900177

1 Introduction with oxygen (> 7 mg L–1). The overall reaction for the dissolu-
tion of silver may be expressed by the classic Elsner equation
Mexico is the leading silver producer in the world and in ac- [1], Eq. (1):
cordance with records there is an annual production of
2 103 125 ton. The country, as does the rest of the world, uses 4Ag + 8CN– + O2 + 2H2O → 4Ag(CN)2– + 4OH (1)
the traditional method of cyanidation to recover gold and sil-
ver in which effluents with cyanide are generated. This waste is which has the following mechanism, as outlined in Eqs. (2–4):
dangerous for the environment. In the extractive industry, this
method of recovering precious metals is carried out using a 4Ag + 8 CN– → 4 Ag(CN)2– + 4 e– (2)
solution of cyanide of 0.03–0.3 % of NaCN, at a pH greater
than 10 and aeration to keep the pulp or solution saturated O2 + 2H2O + 2e– → 2OH– + H2O2 (3)

H2O2 + 2 e– → 2OH– (4)



Correspondence: Dr. J. R. Parga (jrparga@its.mx), Department of In the above mechanisms, the cyanide ion is the complexing
Metallurgy and Materials Science, Institute Technology of Saltillo, agent or ligand and oxygen is the oxidant [1]. The reaction
V. Carranza 2400, Saltillo Coahuila, CP. 25000, Mexico. with gold is similar. However, the silver cyanide complex is

© 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cet-journal.com


1902 J. R. Parga et al. Chem. Eng. Technol. 2009, 32, No. 12, 1901–1908

weaker than the gold cyanide complex, and stronger cyanide 3 Titanium Dioxide Photocatalyst
solutions and/or longer contact times must be employed for its
dissolution. After extraction and recovery of the precious met- In recent years, photocatalysis by polycrystalline semiconduc-
als, substantial amounts of cyanide are delivered to tailing tors irradiated by near-UV light has been found effective in
ponds, which creates environmental problems due to the tox- oxidizing certain organic and inorganic pollutants to less dan-
icity of cyanide [2]. Cyanide exists in three forms in waste- gerous species under mild reaction conditions. This method
water, i.e., free cyanide as HCN, simple cyanide as NaCN, and was found to be suitable for the oxidation of free and complex
complex cyanide as Fe(CN)63–, Ni(CN)42– , Zn(CN)42–, and cyanides dissolved in wastewater. The process works by expos-
Cu(CN)42–. Total cyanide is the sum of simple and complex cy- ing wastewater to the combined forces of sunlight and semi-
anides but excludes other ligands such as cyanate and thiocya- conductor catalyst. A commonly used catalyst is titanium di-
nate. The United States Environmental Protection Agency oxide [4]. This catalyst may be mixed into water creating a
(EPA) has proposed a limit of 0.2 mg L–1 cyanide in drinking slurry or fixed onto lattice-type structures where the water
water. The German and Swiss regulators have set a limit is of flows through. TiO2 is the semiconductor most used in photo-
0.01 mg L–1 for cyanide for surface water and 0.5 mg L–1 for catalysis, because it is chemically and biologically inert, non-
sewers. In Mexico, the Secretary of the Environment and Natu- poisonous, stable to photochemical and chemical corrosion,
ral Resources (SEMARNAT) [2] has fixed a limit of cyanide in abundant in nature and also possesses a band of energy of
0.2 mg L–1. Due to these considerations, the recovery or de- 3.2 eV that can be excited by UV light of k < 387 nm, which
struction of cyanide is a necessary processing step. Several can be supplied by solar light.
cyanide treatment systems have been developed to reduce the The mechanism of photocatalysis is performed when the
cyanide level for disposal of the industrial effluent to below nanoparticles of TiO2 are illuminated by UV light, in which
< 1.0 mg L–1 [2]. All these methods are based on cyanide re- photons excite the band that contains the valence electrons
covery by acidification and/or chemical oxidation destruction. and these cross the conductance band, leaving spaces or holes
In many cases the process is burdened with high chemicals (h+). A characteristic feature of semiconducting metal oxides is
costs and royalty payments. The process of photocatalytic oxi- the strong oxidation power of their h+ holes. If the TiO2 semi-
dation with titanium dioxide is one of the innovative forms conductor is in an aqueous medium, those spaces will react
for the treatment of waters contaminated with cyanide. TiO2 is with the water molecules (H2O), producing hydroxyl radical
recognized by its high photocatalytic activity, which makes it (.OH) [5], which is capable of oxidizing cyanide to cyanate
the most widely used material in photocatalytic environmental (CNO–). The interrelation of energy and the redox mechanism
applications. Nevertheless, this technique has a disadvantage in the surface of TiO2 is shown in Fig. 1, with and without the
for industrial application in that the separation of titanium di- presence of oxygen.
oxide after the photocatalytic degradation is difficult due to In an aqueous medium with cyanide, the first product of the
the fineness of the particles, and therefore, there has not been photocatalytic oxidation using nanocrystals of TiO2 is CNO–
any significant advance in the reuse of titanium dioxide for the [6]. The chemical reactions that represent this oxidation pro-
treatment of waters contaminated with cyanide. The proposal cess as follows, Eqs. (5)–(7):
in the current work is to use the EC process to recover the tita- – Formation of spaces with ultraviolet light (solar light or arti-
nium dioxide from aqueous cyanide suspension. ficial source), Eq. (5):

2 Cyanide Detoxification
The leading processes currently being applied for cyanide de-
toxification in effluents involve the oxidation of cyanide by
INCO (The International Nickel Company) sulfur dioxide/air
process, alkaline chlorination, ozonization, oxidation with per-
manganate and biological treatment. All these methods have
distinct advantages and disadvantages [3]. However, none of
the abovementioned chemical processes is fully acceptable for
cyanide detoxification and a combination of processes in se-
quence is required to meet the targeted cyanide concentration.
The chemical process may leave an excess of other undesirable
chemicals in the wastewater stream. In addition, the residence
times required for biological processes are typically much
greater than those required for mass transfer. As mentioned
before, in many cases the process involves a high cost of chem-
icals and royalty payments, and in this regard, a promising
alternative to these processes can be found in the cyanide de-
toxification of effluents by the use of photocatalytic detoxifica- Figure 1. Schematic illustration of the interrelation of energy
tion of cyanide and the recovery of anatase by an electrocoagu- and the redox mechanism on the surface of TiO2 particles (a)
lation (EC) process. with, and (b) without the presence of oxygen.

© 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cet-journal.com


Chem. Eng. Technol. 2009, 32, No. 12, 1901–1908 Cyanide Detoxification 1903

TiO2 + hm → h+ + e– (5)

– Reaction of valence band spaces, h, with cyanide


ion, Eq. (6):

CN– + 2H+ + 2OH– → CNO– + H2O (6)

– Finally, oxygen reduction consumes the elec-


trons generated by Eq.(5), as shown in Eq. (7):

O2 + 2e– + 2H2O → H2O2 + 2OH– (7)

In this investigation an EC process was used to


recover the titanium dioxide after the detoxifica-
tion of the cyanide.

4 Electrocoagulation
Fundamentals
The EC process operates on the principle that the
cations produced electrolytically from iron and/or
aluminum anodes enhance the coagulation of con-
taminants from an aqueous medium. Electropho-
retic motion tends to concentrate negatively
Figure 2. Schematic diagram of the electroagulation process.
charged particles in the region of the anode and
positively charged ions in the region of the cath-
ode. The consumable, or sacrificial, metal anodes are used to Fe(OH)+2 + H2O → Fe(OH)2+ + H+ (11)
continuously produce polyvalent metal cations in the vicinity
of the anode. These cations neutralize the negative charge of – Cathode:
the particles carried toward the anodes by electrophoretic mo-
tion, thereby facilitating coagulation. In the flowing EC tech- 4H+ + 4e– → 2H2 (12)
niques, the production of polyvalent cations from the oxida-
tion of the sacrificial anodes (Fe and Al) and the electrolysis
gases (H2 and O2) works in combination to flocculate the coa- Fe(OH)2+ + e– → Fe(OH)2(aq) (13)
gulated materials [7], and the gas bubbles produced by electro-
lysis carry the pollutants to the top of the solution where they Fe(OH)2(aq) → Fe(OH)2(s) (14)
are concentrated, collected and removed. A schematic diagram
of the process is illustrated in Fig. 2.
Fe(OH)2(aq) + H2O → Fe(OH)3– + H+ (15)
Chemical Reactions of the Electrocoagulation Process
The chemical reactions that have been proposed to describe Fe(OH)3– → Fe(OH)3(aq) + e– (16)
the mechanism of EC [8] are outlined as follows1):
– Anode: – Overall reaction:
aM(s) → aM +n –
+ an(e ) (8)
6Fe + (12 + x)H2O → 1⁄2 (12 – x)H2(g)↑ + xFe(OH)3*(6 – x)
Fe(OH)2(s) (17)
The following equations apply when using Fe electrodes:
M(s) = Fe metal electrode
where constant a is a stoichiometric coefficient and n is the
Fe(s) → Fe+3 + 3e– (9) number of electrons. The pH of the medium usually rises as a
result of this electrochemical process and the Fe(OH)n(s)
formed remains in the aqueous stream as a gelatinous suspen-
Fe+3 + H2O → Fe(OH)+2 + H+ (10) sion, which can remove the TiO2 from the water, either by
complexation or by electrostatic attraction followed of coagu-
lation and flotation [9]. Generally, in the EC process, bipolar
electrodes are used [10]. It has been reported that cells with bi-
– polar electrodes, connected in series operating at relatively low
1) List of symbols at the end of the paper. current densities, produce iron or aluminum coagulant more

© 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cet-journal.com


1904 J. R. Parga et al. Chem. Eng. Technol. 2009, 32, No. 12, 1901–1908

effectively, more rapidly and more economically than com- The amount of metal that is dissolved or deposited depends
pared to chemical coagulation. on the quantity of electricity that passes across the electrolytic
solution and the residence time of water in the EC cell.
A simple relation between the current density and the quan-
5 The Langmuir Isotherm tity the substance dissolved (g cm–2 of M) comes from Fara-
day’s law, Eq. (23):
Several adsorption isotherms have proven useful in under-
standing the process of adsorption. The simplest isotherm is W = (D · t · M)/ (n · F) (23)
attributed to a pioneer in the study of surface processes, Lang-
muir, and is called the Langmuir isotherm. If one assumes that where W is the amount of the dissolved electrode (g cm–2), D
(i) adsorption cannot proceed beyond the point at which the is the density of current (A cm–2), t is the time (s), M is the rel-
adsorbates are one layer thick on the surface, i.e., a monolayer ative molar mass of the electrode, n is the electron number in
is formed, (ii) all adsorption sites are equivalent, and (iii) the the redox reaction, and F is the Faraday constant (96 500 cou-
adsorption and desorption rate is independent of the popula- lombs).
tion of neighboring sites [11]. Taking points (i)–(iii) into The Langmuir isotherm can also be written as follows:
account, one can derive a simple formula for an adsorption
N = (NmaxKLCe)/(1 + KLCe) (24)
isotherm. Consider the equilibrium in Eq. (18):
where N is the solid phase adsorbate concentration in equilib-
A + S → A·S (18)
rium (mg g–1), Nmax is the maximum adsorption capacity cor-
responding to complete monolayer coverage on the surface
where A is the free adsorbate, S is the free surface, and A·S is
(mg g–1), Ce is the concentration of adsorbate at equilibrium
the substrate bound to the surface. The rate of adsorption will
(mg L–1) and KL is the Langmuir constant (L mg–1). Eq. (21)
be proportional to the pressure of the gas and the number of
can be rearranged to a linear form, Eq. (22):
vacant sites for adsorption. If the total number of sites on the
surface is N, then the rate of change of the surface coverage Ce/N = 1/(NmaxKL) + Ce/Nmax (25)
due to adsorption is given by Eq. (19):
It was observed that the equilibrium adsorption data fol-
dH/dt = kapN (1 – T) (19) lowed the Langmuir isotherm, Fig. 3. The constants can be
evaluated from the intercepts and the slopes of the linear plots
The rate of change of the coverage due to the adsorbate leav- of Ce/N versus Ce. By following the model of the Langmuir
ing the surface, i.e., desorption, is proportional to the number isotherm, one can obtain the titanium dioxide adsorption ca-
of adsorbed species, Eq. (20): pacity on the iron species.

dH/dt = – kdNT (20)

In these equations, ka and kd are the rate constants for ad-


sorption and desorption, respectively, and p is the pressure of
the adsorbate gas. At equilibrium, the coverage is independent
of time, and thus, the adsorption and desorption rates are
equal. The solution to this condition gives us a relation for T,
Eq. (21):

H = KP/(1 + KP) (21)

where K = ka/kd, which is the usual form for expressing the


Langmuir isotherm [12].
Figure 3. Linear plot of Ce/N versus Ce.

6 Adsorption Study
7 Thermodynamic Calculations of
The amount of titanium dioxide adsorbed per unit mass, Adsorption Data
N (mg g–1), is calculated as Eq. (22):
Various thermodynamic parameters were also calculated by
N = V(C0 – Ce)/W (22) using the following relations:

where C0 and Ce (mg L–1) are the initial and equilibrium liquid DG° = –RTln KL (26)
phase concentrations of titanium dioxide, respectively, V the DH° = –RTlnKL – KL0 (27)
volume of the solution (L) and W is the mass of adsorbent
used (g) which in this case is the Fe dissolved from the EC. DS° = (DH° – DH°)/T (28)

© 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cet-journal.com


Chem. Eng. Technol. 2009, 32, No. 12, 1901–1908 Cyanide Detoxification 1905

where DG° is the change in the Gibbs free energy, DH° is


the change in enthalpy, DS° is the change in entropy, and
KL and KL0 are the Langmuir constants for the final and
initial concentration. R (8.314 J/mol K) is the gas constant
and T (K) is the absolute temperature. The negative value
of the free energy suggests the viability of the process. The
negative value of the enthalpy confirms the exothermic na-
ture of the process and the positive value of the change of
entropy shows the increase of the randomness of the pro-
cess [12].

8 Materials and Methods


The cyanide photocatalytic oxidation tests were performed
in a 400 mL beaker. Catalyst concentrations of 1.25, 1.0,
0.75 and 0.5 g L–1 of titanium dioxide (Degussa P-25) were
used. The initial concentration of cyanide was 400 ppm,
the solution was prepared using NaCN, reagent grade Figure 4. Diagram of the photocatalytic oxidation of cyanide.
(95 %), made by Monterrey chemical products. A 450 W
halogen lamp was used as the UV light source and the tests
were run for 30 min with agitation with the lamp being
placed on top of the beaker for direct irradiation. Cyanide
determinations were made by using the selective ion elec-
trode technique and adjusting the pH to 11 in order to
avoid HCN formation. A diagram of the setup used for the
photocatalytic cyanide detoxification process is shown in
Fig. 4.
EC tests were realized in a 400 mL beaker, with two ver-
tical iron electrodes of 7 × 10 cm and 5 mm separation in
between, the effective volume being 350 mL. The geometry
of the vertical plates allowed the use of the gaseous H2 gen-
erated in the EC process to facilitate the removal of the fer-
ric and ferrous species associated with anatase. A DC po-
tentiometer was used to control the voltage or the current.
The solution was prepared using deionized water and the
conductivity was adjusted by the addition of 1 g of NaCl
per L of water. The pH was adjusted using a 0.13 M NaOH
solution. Solution and solids were separated by filtration. Figure 5. Diagram of the electrocoagulation cell.
The sludge from the EC process was dried for 8 h in an
oven to 80 °C. A diagram of the EC cell is shown in Fig. 5.
A Shimadzu model 6701 atomic absorption spectrometer one can also see that by increasing the concentration of the
was used for the characterization of the aqueous samples. The catalyst, the level of elimination of cyanide diminishes. This is
dried solids were characterized with X-ray diffraction (XRD) due to the fact that an optical saturation happens with an in-
(Phillips model X-PERT difractometer) and scanning electron creased concentration of catalyst, a phenomena that was also
microscopy (SEM) (FEI Quanta 2000, Oxford Instruments). observed by other investigators [5]. The EC technique was
used to recover the titanium dioxide from its aqueous suspen-
sions.
9 Results and Discussion
9.1 Results of the Photocatalytic Technique for 9.2 Electrocoagulation Results
Destruction of Cyanide
The initial conditions and results for the EC tests are given in
A calibration curve was constructed initially for cyanide deter- Tab. 2. The results for the adsorption of anatase particles on
mination, and the results are shown in Fig. 6. The results for the iron species show recoveries of 98, 92 and 96 % of the TiO2
the cyanide oxidation using the photocatalysis technology with microelectrode. In order to determine these recoveries of TiO2,
TiO2 nanoparticles are listed in Tab. 1. From Tab. 1, one can the total sludge was leached with sulfuric acid in order to dis-
see that cyanide oxidation using the photocatalysis technology solve the iron species and the solid remnant was TiO2. The
is a viable option to eliminate this water pollutant. In addition, high efficiency of the EC technology in the recovery of TiO2

© 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cet-journal.com


1906 J. R. Parga et al. Chem. Eng. Technol. 2009, 32, No. 12, 1901–1908

recover the TiO2 particles. The recovery levels of anatase at


pH of 7 with the EC cell (350 cm3) were 98 % and 96 %.

9.4 Scanning Electron Microscope Results

An SEM image of TiO2 nanoparticles on iron oxyhydrox-


ide particles is shown in Fig. 8. The energy dispersive
X-ray analysis (EDAX) of the sample shown on th eright
hand side of Fig. 8 demonstartes shows that the surface of
these particles are coated with a layer of TiO2 nanoparti-
cles.

Figure 6. Calibration curve for the measurement of cyanide.


9.5 Thermodynamics Parameter Results
Table 1. Results of the cyanide oxidation.
The thermodynamic parameters, i.e., change in standard
[TiO2] [CN–]IN [CN–]OUT Elimination of CN– free energy, DG°, enthalpy, DH°, and entropy, DS°, were
(g L–1) (mg L–1) (mg L–1) (%) determined by using Eqs. (19)–(21). The negative value of
free energy, DG°, confirms the viability of the adsorption
0.5 400 28 93
process and spontaneous nature of the TiO2 adsorption on
0.75 400 40 90 the generated species of EC (see Tab. 3). The negative value
1.0 400 52 87 of enthalpy, DH°, indicates the exothermic nature of the
process. The magnitude of the adsorption enthalpy gives
1.25 400 48 88 an idea about the type of adsorption, which is mainly
physical or chemical. Physisorption processes usually have
energies in the range of –(10–40) kJ/mol while higher ad-
allows the later reuse of the TiO2 in the photocatalytic oxida- sorption enthalpies of –(100–500) kJ/mol suggest chemisorp-
tion of the cyanide process. tion [12]. The value of the adsorption enthalpy (–13.397 kJ/
mol) given in Tab. 3 confirms the physisorbed nature of the
processes for titanium dioxide. The positive value of DS°
9.3 X-Ray Diffraction Results shows the increased randomness at the solid/solution interface
during the adsorption process, which suggests that TiO2
A detailed diffraction study was performed on the dried EC replace some water molecules from the solution previously
solid products at an initial pH of 7.2. The sample was separat- adsorbed on the surface of adsorbent. These displaced mole-
ed into two portions using a small permanent magnet. The cules gain more translation entropy than is lost by the adsorb-
curve in Fig. 7a) shows the diffractogram of the ferromagnetic ate ions, thus allowing the prevalence of randomness in the
component (the fraction collected on the permanent magnet). system.
This “magnetic” fraction of the sample demonstrates the dif-
fraction pattern due to either magnetite or maghemite. The
curve in Fig. 7b) shows the diffractogram of the dried EC solid 10 Conclusions
product performed on the pH 7 sample that contained the
TiO2 particles. The curve in Fig. 7c) shows the diffractogram (i) Cyanide destruction with photocatalytic oxidation can be
of the dried EC solid product performed on the pH 7 sample achieved in an economically and environmentally accept-
that contained the TiO2 nanoparticles formed on the iron oxy- able manner.
hydroxide particles. The results of this study suggest that the (ii) Results show that photocatalyzed degradation of cyanide
presence of maghemite and magnetite particles can be used to waste with TiO2 semiconductor particles is very efficient
for the destruction of cyanide at > 90 %.

Table 2. Conditions for EC.

Sample Initial pH Conductivity EC Voltage Current Final Recovery


[TiO2] (lsiemens) Time (V) (A) [TiO2] (%)
(g L–1) (min) (g L–1)

1 0.5 6.5 4.10 30 11.3 0.46 0.01 98


2 0.75 6.7 4.13 30 10.62 0.45 0.06 92
3 1.0 6.6 4.15 38 11.2 0.46 0.04 96

© 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cet-journal.com


Chem. Eng. Technol. 2009, 32, No. 12, 1901–1908 Cyanide Detoxification 1907

Figure 7. X-ray diffractograms for (a) Magnetite, (b–c) Magnetite with anatase.

(iii) Since the EC products contain ferromagnetic materials (iv) The X-ray diffraction technique and scanning electronic
such as magnetite or maghemite, the process seems to microscopy demonstrate that the species formed are of
offer the potential for recovery of the TiO2. Recoveries magnetic type, e.g., magnetite and maghemite, and are
of 96 % and 98 % TiO2 were achieved with the EC pro- adsorbed on the surface of the TiO2 nanoparticles due to
cess. the electrostatic attraction between both metals.

Figure 8. SEM image of iron particles coated with anatase nanoparticles, and EDAX data for same particle system.

© 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cet-journal.com


1908 J. R. Parga et al. Chem. Eng. Technol. 2009, 32, No. 12, 1901–1908

Table 3. Thermodynamics parameters. V [L] volume of the solution


W [g] mass of adsorbent
kcal mol–1 kJ mol–1 D [A cm–2] density of current
DG° –5.946 –29.0184 t [s] time
Nmax [mg g–1] maximum adsorption capacity
DH° –2.745 –13.397
KL [L mg–1] Langmuir constant
–1 –1
kcal mol K kJ mol–1K–1 Ce/N [–] ratio of concentration of
adsorbate at equilibrium and the
DS° 0.0107 0.0524
amount of titanium dioxide
adsorbed per unit mass
DG° [kJ mol–1] standard free energy of
(v) Based on Langmuir isotherm analysis, the thermody- adsorption
namic parameters obtained for the recovery of TiO2 are as DH° [kJ mol–1] standard enthalpy of adsorption
follows: DG° = –29.0184 kJ/mol, DH° = –13.397 kJ/mol DS° [kJ mol–1K–1] standard entropy of adsorption
and DS° = 0.0524 kJ/mol K. The negative value of the free T [K] temperature
energy, DG°, confirms the viability of the adsorption pro- R [J mol–1K–1] Universal gas constant
cess and its spontaneous nature. The adsorption enthalpy
value corresponds to physisorption values. The positive Greek symbol
value of the change of entropy shows the increase of the
randomness of the process. H [–] surface fraction coverage

Acknowledgements References

The authors wish to acknowledge support of this project by [1] M. Jeffrey, M. Ritchie, J. Electrochem. Soc. 2005, 147, 3257.
the National Council of Science and Technology (CONACYT) [2] J. Parga, D. Cocke, J. Desal. 2001, 140, 289.
and the Dirección General de Educación Superior Tecnológica [3] J. Blanco, S. Malato, Ph.D. Thesis, Plataforma Solar de Almer-
(DGEST). ía, Almería, Spain 2004.
[4] E. Pavas, M. Camargo, C. Jones, M.Sc. Thesis, Universidad
The authors have declared no conflict of interest. EAFIT, Mexico 2005.
[5] J. Parga, S. Shukla, D. Cocke, Res. J. Chem. Environ. 2005, 9,
60.
Symbols used [6] V. Augugliaro et al., Catal. Today 1999, 54, 245.
[7] J. Parga, D. Cocke, H. Moreno, J. Hazard. Mater. 2005, B124,
a [–] constant stoichiometric 247.
coefficient [8] H. Moreno et al., Ind. Eng. Chem. Res. 2009, 48 (4), 2275.
n [–] number of electrons [9] M. Kobya, H. Hiz, E. Senturk, J. Hazard. Mater. 2007, 55,
Ms [–] Fe metal electrode 1244.
Ka [–] constant of adsorption [10] I. Tuvert, V. Talanquer, Para Saber, Experimentar y Simular,
Kd [–] constant of desorption Facultad de Quimica, UNAM, 2000, 6, 186.
N [mg g–1] amount of TiO2adsorbed per [11] A. Mittal, A. Malviya, D. Kaur, J. Hazard. Mater. 2007, 148,
unit mass 230.
C0 [mg L–1] initial concentrations of TiO2 [12] A. Gupta, S. Kunda, Sep. Purif. Technol. 2006, 51, 165.
Ce [mg L–1] concentration of TiO2at
equilibrium

© 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cet-journal.com

You might also like