Anda di halaman 1dari 16

Thermochimica Acta 559 (2013) 116

Contents lists available at SciVerse ScienceDirect

Thermochimica Acta
journal homepage: www.elsevier.com/locate/tca

Review

A chemical thermodynamics review applied to V2 O5 chlorination


E.A. Brocchi, R.C.S. Navarro , F.J. Moura
Materials Engineering Department (PUC-Rio), Pontical Catholic University of Rio de Janeiro, Rio de Janeiro, Brazil

a r t i c l e

i n f o

Article history:
Received 24 July 2012
Received in revised form 10 January 2013
Accepted 25 January 2013
Available online 6 March 2013
Keywords:
Chemical thermodynamic
V2 O5
Chlorination

a b s t r a c t
This work is mainly related to the thermodynamic study of the V2 O5 carbon-chlorination. In this context,
three different approaches of diverse complexity level were employed. First, the formation of individual
vanadium chlorides and oxychlorides were considered on the basis of the well known G T diagrams.
It is suggested that these simple constructions can help in understanding the formation sequence of
vanadium chlorinated species. Moreover, the relative stability of the most stable vanadium chloride (VCl4 )
and oxychloride (VOCl3 ) was addressed based on available thermodynamics data. Finally, a gas phase
speciation calculation was performed in order to obtain, simultaneously, the concentration of all possible
vanadium chlorides and oxychlorides as a function of temperature, Cl2 and O2 partial pressures. It is
demonstrated that, although diverse in the complexity level, the three methods considered have a strong
relation with each other, and converge to a better insight into the nature of the chemical equilibrium
states achievable for the reaction system under study. The same approach can be applied to any other
reaction system of technological importance.
2013 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.
4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1.
Reductive chlorination roasting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The system VOCl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Vanadium oxides and chlorides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1.
V2 O5 direct chlorination and the effect of the reducing agent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.2.
Relative stability of VCl4 and VOCl3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Final remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
In general terms the present work can be considered as a contribution to a quantitative study review related to the chemical
equilibrium of metallic oxides chlorination. Three possible ways of
investigating the achievable equilibrium states will be presented.
The rst one is based entirely on the construction of G T diagrams. The second one applies the reactions Gibbs energies for
computing the concentration of specic chlorinated compounds
of interest and enables the study of the relative stability of these
species in a specic temperature range. The third and most general

Corresponding author. Tel.: +55 2135271720.


E-mail addresses: ebrocchi@puc-rio.br (E.A. Brocchi), rogerioncs@gmail.com
(R.C.S. Navarro).
0040-6031/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.tca.2013.01.025

1
2
2
3
3
10
13
15
15
15

one is dened by the minimization of the total Gibbs energy of the


system.
The three mentioned methods are then applied to the study
of the thermodynamic viability of the reaction between V2 O5 and
gaseous Cl2 . The importance of the presence of a reducing agent is
discussed through the construction of proper G T in the presence as in the absence of graphite. The same method is also applied
for the formulation of possible chlorination paths for the formation
of the most stable chlorinated species produced at high P(Cl2 ) and
low P(O2 ) values, conditions usually found in most experiments.
The relative stability of the mentioned chlorides is then studied
based on the independent resolution of chemical equilibrium equations. Finally, the equilibrium states available to the system are
predicted through the minimization of its total Gibbs energy. The
effect of T, P(Cl2 ) and P(O2 ) over the composition of the gas phase in
clearly presented and discussed. Also, calculations of the enthalpy

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

associated with the chlorination reactions are also performed, and


the impact of the amount of Cl2 available over the heat absorbed or
released by the reaction system is explored. Emphasis is given to the
point that the most general strategy must embrace the tendencies
indicated by the implementation of simpler methodologies.
Until recently, the approach applied for chemical equilibrium
studies was almost exclusively based on standard free energy
versus temperature and predominance diagrams. However, nowadays, advances in computational thermodynamics have enabled a
sort of software developments which can perform more complex
calculations. So, it is understood that time has come for a review on
thermodynamics chlorination which can combine its basic aspects
with a now available new kind of approach. Therefore, the generality and depth of the produced information for this work have also
motivated the authors to publish its content in a form of a chapter,
which could well be considered as a part of a thermodynamic book
applied to high temperature process.

1.1. Reductive chlorination roasting


Chlorination roasting has proven to be a very important industrial route and can be applied for different purposes. Firstly, the
chlorination of some important minerals is a possible industrial
process for producing and rening metals of considerable technological importance, such as titanium and zirconium. Also, the same
principle is mentioned as a possible way of recovering rare earth
from concentrates [1] and metals, of considerable economic value,
from different industrial wastes, such as, tailings [2], spent catalysts [3], slags [4] and y ash [5]. The chlorination processes are also
presented as environmentally acceptable [6,7]. In general terms the
chlorination can be described as reaction between a starting material (mineral concentrate or industrial waste) with chlorine in order
to produce some volatile chlorides, which can then be separated by,
for example, selective condensation. The most desired chloride is
puried and then used as a precursor in the production of either the
pure metal (reaction of the chloride with magnesium) or its oxide
(by oxidation of the chloride).
The chlorination reaction has been studied on respect of many
metal oxides [811] as this type of compound is the most common
in the mentioned starting materials. Although some basic thermodynamic data is enclosed in these works, most of them are related
to kinetics aspects of the gassolid reactions. However, it is clear
that the understanding of the equilibrium conditions, as predicted
by classical thermodynamics, of a particular oxide reaction with
chlorine, can give strong support for both the control and optimization of the process. In this context, the impact of industrial
operational variables over the chlorination efciency, such as the
reaction temperature and the reactors atmosphere composition,
can be theoretically appreciated and then quantitatively predicted.
On that sense, some important works have been totally devoted
to the thermodynamics of the chlorination and became classical
references on the subject [1215].
As said before, the approach applied for the study of chemical equilibrium studies was based exclusively on G T and
predominance diagrams. Nowadays, due to the development of
computational thermodynamics, a more detailed calculation is
possible. This approach, together with the one accomplished by
simpler techniques, converge to a better understanding of the intimate nature of the equilibrium states for the reaction system of
interest.
The present chapter focuses on the chemical equilibrium conditions associated with the chlorination of V2 O5 both in the presence
as in the absence of carbon (reducing agent). The impact of carbon over the thermodynamic viability of the reactions is clearly
presented and discussed. Also, the equilibrium conditions are

Fig. 1. Predominance diagram for the system VO.

appreciated through graphical constructions of different complexity level, beginning with the well known G T and predominance
diagrams, and ending with gas phase speciation diagrams, rigorous
calculated through the minimization of the total Gibbs energy of
the system.

2. The system VOCl


Vanadium is a transition metal that can form a variety of oxides.
At ambient temperature and oxygen potential, the form V2 O5 is
the most stable. It is a solid stoichiometric oxide, where vanadium
occupies the +5 oxidation state. By lowering the partial pressure of
O2 , the valence of vanadium varies considerably, making it possible to produce a family of stoichiometric oxides: V2 O4 , V3 O5 , V4 O7 ,
VO, VO2 and V2 O3 . Recently, it has been discovered that vanadium
can also form a variety of non-stoichiometric oxygenated compounds [16], however, to simplify the treatment of the present
review, these phases will not be included in the data-base used for
the following computations. Additionally, it was considered that
the concentration of the oxides in gas phase is low enough to be
neglected. Further, on what touches the computations that follows,
the software Thermocalc was used in all cases, and it will always
be assumed that equilibrium is achieved, or in other words, kinetic
effects can be neglected.
The relative stability of the possible vanadium oxides can be
accessed through construction of a predominance diagram in the
space TP(O2 ) (see Fig. 1). As thermodynamic constraints we have
n(V) (number of moles of vanadium metalit will be supposed that
n(V) = 1), T, P and P(O2 ). The reaction temperature will be varied in
the range between 1073 K and 1500 K and the partial pressure of
O2 in the range between 8.2 104 atm and 1 atm.
The total pressure was xed at 1 atm. It can be seen that for the
temperature range considered and a partial pressure of O2 in the
neighborhood of 1 atm, V2 O5 is formed in the liquid state. Through
lowering the oxygen potential, crystalline vanadium oxides precipitate, VO2 being formed rst, followed by V2 O3 , VO, and nally V.
The horizontal line between elds 5 and 6 indicates the melting
of VO2 , which according to classical thermodynamics must occur at
a xed temperature. Next it will be considered the species formed
by vanadium, chlorine and oxygen.

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Table 1
Physical nature and selected data for vanadium chlorinated compounds.
Chlorinated species

Physical state at 25 C

Equilibrium data

Transition temperature (K)

VCl2

Solid

Psat (T) [17]

Tf = 1088 [18], 1620 [19]


Ts = 1680 [17]
Tb = 1803 [18]

VCl3

Solid

PVCl4 (T) [17]

Td = 1120 [17], 914 [19]


Ts = 1106 [18]

VCl4
VOCl3
VO2 Cl

Liquid
Liquid
Solid

PVOCl3 (T) [20]


PVOCl3 (T) [20]

Tb = 398 [18], 424 [19]


Tb = 694 [18], 400 [19], 399 [20]
Td = 450 [20]

VOCl2

Solid

PVOCl3 (T) [20]

Td = 650 [20]
Ts = 784 [21]

VOCl

Solid

Ts = 1393 [21]

Psat (T), equilibrium vapor pressure as a function of temperature; PVOCl3 (T), equilibrium VOCl3 pressure as a function of temperature; PVCl4 (T), equilibrium VCl4 pressure as
a function of temperature; Td , dismutation/decomposition temperature; Tf , melting temperature; Ts , sublimation temperature; Tb , ebullition temperature.

Table 2
Equilibrium constant for the reaction represented by Eq. (3).

2.1. Vanadium oxides and chlorides


The already identied species formed between vanadium, chlorine and oxygen are: VCl, VCl2 , VCl3 , VCl4 , VOCl, VOCl2 , VOCl3 ,
VO2 Cl. In Table 1 it was included information regarding the physical states at ambient conditions, some references related to phase
equilibrium studies conducted on samples of specic vanadium
chlorinated compounds, as well as some available literature data
regarding specic phase transitions.
Only a few studies were published in literature in relation to
the thermodynamics of vanadium chlorinated phases. In Table 1
some references are given for earlier investigations associated with
measurements of the vapor pressure for the sublimation of VCl2 and
VCl3 , and the boiling of VOCl3 and VCl4 . There are also evidences
for the occurrence of specic thermal dismutation reactions, such
as those of VCl3 [17], VOCl2 and VO2 Cl [20] (Eq. (1)). According to
the mentioned studies, at 1 atm VCl3 should decompose at 1120 K,
VO2 Cl at 450 K and VOCl2 at 650 K. In the case of VCl3 , however, it
is not clear if the two processes (sublimation and dismutation) can
occur simultaneously [21].

Equilibrium constant

1173
1273
1473

1.76257 1013
5.82991 1011
1.0397 1008

2.1.1. V2 O5 direct chlorination and the effect of the reducing


agent
The direct chlorination of V2 O5 is a process, which consists in
the reaction of a V2 O5 sample with gaseous Cl2 .
V2 O5 + Cl2 = Chloride/Oxychloride + O2

(1)

2VOCl2 (s) = VOCl3 (g) + VOCl(s)


Chromatographic measurements conducted recently conrmed
the possible formation of VCl, VCl2 , VCl3 , and VCl4 in the gas phase
[22]. In this study the molar Gibbs energy models for the mentioned chlorides were revised, and new functions proposed. In the
case vanadium oxychlorides, models for the molar Gibbs energies
of gaseous VOCl, VOCl3 , and VOCl2 have already been published
[21]. For gaseous VO2 Cl, on the other hand, no thermodynamic
model exists, indicating the low tendency of this oxychloride to
be stabilized in the gaseous state.
It is worthwhile to mention that, for the calculations conducted
in the present work the SSUB3 Thermocalc data-base [18] was used
for modeling the Gibbs energy of all vanadium chlorides, VCl2 (s,l,g),
VCl3 (s,g) and VCl4 (l,g). On what touches the chloride VCl(g) the
data published by Hildenbrand et al. [22] was considered. In the
case of the Gibbs energies of vanadium metal and the vanadium
oxides V2 O4 (s), VO2 (s,l), V2 O3 (s) and V2 O5 (s,l), the SSUB3 Thermocalc data-base was again employed [18], which also served as
source for the Gibbs energy of VOCl3 (l,g). In the case of the oxides
V4 O7 (s) and V3 O5 (s), as for VO2 Cl(s), the models contained in the
HSC version 6.0 data-base were considered [23]. For VOCl(s,g) and
VOCl2 (s,g), the data published by Hackert et al. [21] was employed.

(2)

For most pyrometallurgical processes, temperature lies usually


between 1173 and 1473 K. Considering the chloride VCl4 , its formation from V2 O5 can be modeled according to Eq. (3). This reaction is
associated with very low thermodynamic driving force in the mentioned temperature range, a fact that is demonstrated by the low
magnitude of its equilibrium constant (Table 2).
V2 O5 + 4Cl2 = 2VCl4 + 2.5O2

VCl3 (s) = VCl2 (s) + VCl4 (g)


3VO2 Cl(s) = VOCl3 (g) + V2 O5 (s)

T (K)

(3)

The very low magnitude of the equilibrium constant is an indicative, that at the standard conditions, the chemical equilibrium
represented by Eq. (3) is almost entirely dislocated in the direction of the decomposition of VCl4 . The equilibrium could then be
shifted in the direction of the formation of the mentioned chloride
if one removes O2 from the reactors atmosphere.
One possible strategy in this direction is to add to the
reaction system some carbon bearing compound [2426]. The
compound decomposes producing graphite, which reacts with oxygen, thereby shifting the chlorination equilibrium in the desired
direction. A simpler route, however, would be to admit carbon as
graphite together with the oxide sample into the reactor. If graphite
is present in excess, the O2 concentration in the reactors atmosphere is maintained at very low values, which are achievable
through the formation of carbon oxides (Eq. (4)).
2C + O2 = 2CO
C + O2 = CO2

(4)

So, for the production of VCl4 in the presence of graphite, the


reaction of C with O2 can lead to the evolution of gaseous CO or
CO2 (Eq. (5)).
V2 O5 (s, l) + 4Cl2 (g) + 2.5C(s) = 2VCl4 (l, g) + 2.5CO2 (g)
V2 O5 (sl) + 4Cl2 (g) + 5C(s) = 2VCl4 (l, g) + 5CO(g)

(5)

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

where the curves corresponding to the formation of CO and CO2


have the same Gibbs energy value. This point is dened by the
temperature, where the Gibbs energy of the Boudouard reaction
(C + CO2 = 2CO) is equal to zero. For each reaction, continuous curves
denote the formation of the most stable species, which are associated with the lowest value of the reaction Gibbs energy.
The equivalence of this point and the intersection associated
with the curves for the formation of VCl4 can be understood, as the
Boudouard reaction can be obtained through a simple linear combination, according to Eq. (6). So, its molar Gibbs energy is equal
to the difference between the molar Gibbs energy of the VCl4 formation with the evolution of CO and the same quantity for the
reaction associated with the CO2 production. Concomitantly, when
the curves for the formation of VCl4 crosses each other, the difference between their molar Gibbs energies is zero, and according to
Eq. (6) the same must happen with the molar Gibbs energy of the
Boudouard reaction.
G1

1) V2 O5 (s, l) + 4Cl2 (g) + 5C(s)2VCl4 (g) + 5CO(g)

G2

Fig. 2. Gr vs. T for the formation of VCl4 in the presence of graphite.

2) V2 O5 (s, l) + 4Cl2 (g) + 2.5C(s)2VCl4 (g) + 2.5CO2 (g)


=

(6)
G3

The effect of the presence of graphite over the G T curves for


the formation of VCl4 can be seen in the diagram of Fig. 2. As a matter of comparison, the plot for the formation of the same species in
the absence of graphite is also shown, together with the curves for
the reactions associated with the formation of CO and CO2 for one
mole of O2 (Eq. (4)). In the temperature range considered, VCl4 is
produced in the gaseous state. It can be readily seen that graphite
strongly reduces the standard molar Gibbs energy of reaction, promoting in this way considerably the thermodynamic driving force
associated with the chlorination process. The presence of graphite
has also an impact over the standard molar reaction enthalpy. The
direct action of Cl2 is associated with an endothermic reaction
(positive linear coefcient), but by adding graphite the processes
become considerably exothermic (negative linear coefcient).
The curves associated with the VCl4 formation in the presence of
the reducing agent cross each other at 973 K, the same temperature

Fig. 3. Gr vs. T the formation of VCl4 -melting of V2 O5 .

3) 2.5C(s) + 2.5CO2 (g)5CO(g)


G3 = G1 G2
Lim (G3 ) =

T 973 K

Lim (G1 G2 ) = 0

T 973 K

The inexion point present on the curves of Fig. 2 is associated


with the melting of V2 O5 . This inexion is better evidenced on the
graphic of Fig. 3. As V2 O5 is a reactant, the curve should experience a
reduction of its inclination at the melting temperature of the oxide.
However, the presence of the mentioned inexion point is much
more evident for the reactions with the lowest variation of number
of moles of gaseous reactants, as is the case for the direct action of
Cl2 , which leads to the evolution of CO2 the variation of number
of moles of gaseous species (ng ) is equal to 0.5, smaller than the
value obtained for the same reaction leading to CO (ng = 3). The
value of ng controls the molar entropy of the reaction. By lowering
the magnitude of ng the value of the reaction entropy reduces,
and the effect of melting of V2 O5 over the standard molar reaction
Gibbs energy becomes more evident.
The presence of graphite affects in the same way the thermodynamic tendency of formation of vanadium oxychlorides. This can
be well illustrated for the synthesis of VOCl3 . Its G T curve is
compared with the one for the formation of VCl4 on the graphic of
Fig. 4. As in the case of VCl4 , VOCl3 is formed as a gas in the temperature range considered. Moreover, the inexion around 954 K
is again associated with the melting of V2 O5 . Further, as the reaction associated with the formation of VCl4 , the formation of VOCl3
has a negative molar reaction enthalpy. So, if during the production
of both chlorinated compounds the gas phase is considered ideal,
the system should transfer heat to its neighborhood (exothermic
reaction).
On what touches the molar reaction entropy, the graphic of Fig. 4
indicates, that the reaction associated with the formation of VCl4
should generate more entropy (more negative angular coefcient
for the entire temperature range). This can be explained by the fact,
that in the case of VCl4 , the value of ng (ng = 3) is higher than the
same value for the formation of VOCl3 (ng = 2). The same tendencies are expected for the formation reactions associated with CO2
evolution, as these result from a simple linear combination between
the reaction equations considered in Fig. 4 and the Boudouard reaction.

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Fig. 4. Gr vs. T for the formation of VOCl3 and VCl4 in the presence of graphite.
Fig. 6. Gr T for reaction paths of Eq. (8).

2.1.1.1. Multiple reaction equilibria. The G T diagram is a valuable tool for suggesting possible reactions paths in the case of
systems, where multiple reactions occur simultaneously. Lets consider rst the formation of VCl4 . In a rst glance, such process could
be thought as the result of three stages. In the rst one, a lower chlorinated compound (VCl) is formed. The precursor then reacts with
Cl2 resulting in higher chlorinated species (Eq. (7)). The G T
plots associated with reactions paths represented by mechanisms
of Eq. (7) were included in Fig. 5.
V2 O5 + Cl2 + 2.5C = 2VCl + 2.5CO2
V2 O5 + Cl2 + 5C = 2VCl + 5CO
VCl + 0.5Cl2 = VCl2
VCl2 + 0.5Cl2 = VCl3

(7)

As said before, in each temperature range, reactions leading


to stable species are represented by continuous curves. This convention will be adopted in all other diagrams included in the
rest of the present article. On the diagram of Fig. 5, two inexion points are evidenced. The rst one around 1000 K is associated
with VCl2 melting. The second more evident one, around 1100 K,
is associated with the sublimation of VCl3 . It can be speculated
that only for temperatures greater than 1600 K the path described
by Eq. (7) could be realized. For lower temperatures, the molar
Gibbs energy of the rst step is higher than the one associated
with the second. In this situation, therefore, another mechanism
must be formulated for the VCl4 appearance. One possibility is
dened by Eq. (8). This time, VCl2 is formed directly from V2 O5 ,
which then reacts to give VCl3 and nally VCl4 . The characteristic

VCl3 + 0.5Cl2 = VCl4

Fig. 5. Gr T for reaction paths of Eq. (7).

Fig. 7. Gr T for reaction paths of Eq. (8).

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Fig. 8. Gr T for reaction paths of Eq. (10).

Fig. 9. Gr T for reaction paths of Eq. (10).

G T curves for the reactions dened in Eq. (8) are presented in


Figs. 6 and 7.

VOCl being rst produced, which then reacts leading to gaseous


VOCl2 , which nally results in gaseous VOCl3 .

V2 O5 + 2Cl2 + 2.5C = 2VCl2 + 2.5CO2

V2 O5 + Cl2 + 1.5C = 2VOCl + 1.5CO2

V2 O5 + 2Cl2 + 5C = 2VCl2 + 5CO


VCl2 + 0.5Cl2 = VCl3

(8)

V2 O5 + Cl2 + 3C = 2VOCl + 3CO


VOCl + 0.5Cl2 = VOCl2

(10)

VCl3 + 0.5Cl2 = VCl4

VOCl2 + 0.5Cl2 = VOCl3

The inexion points have the same meaning as described for


diagram of Fig. 5. It can be seen that the rst step (formation of
VCl2 ) has a much higher thermodynamic tendency as the others.
For temperatures lower than 953 K or higher than 1539 K the mechanism represented by Eq. (8) describes the order of appearance of
the chlorides as P(Cl2 ) gets higher. In these conditions, VCl2 is rst
formed as a solid for temperatures lower than 953 K and liquid for
temperatures higher than 1539 K. In the mentioned temperature
intervals, the second reaction step leads to the formation of VCl3 ,
which can be produced as a solid (T < 953 K) or gas (T > 1539 K),
which then nally reacts resulting in gaseous VCl4 . On the other
hand, for temperatures higher than 953 K and lower than 1539 K,
the step associated with the formation of VCl4 is now the one with
the second lowest standard Gibbs energy. So, in this temperature
range, VCl4 should be formed directly from VCl2 , as suggested by
Eq. (9).

For temperatures lower than 1053 K, VOCl3 should indeed be


formed directly from VOCl (Eq. (11)). Again, to attain thermodynamic consistency for temperatures lower than 1053 K, the Gibbs
energy curves associated with the formation of VOCl2 and VOCl3
according to Eq. (10) must be substituted for the curve associated with reaction represented by Eq. (11). It should be mentioned
indeed, that the reaction equations compared must be written with
the same stoichiometric coefcient for Cl2 , which was set equal to
1/2 (Fig. 9).

0.5VCl2 + 0.5Cl2 = 0.5VCl4

(9)

In order to achieve thermodynamic consistency in the temperature interval between 953 K and 1539 K, the curves associated with
the formation of VCl3 and VCl4 according to Eq. (8) should be substituted for the curve associated with reaction dened by Eq. (9), as
evidenced on the plot of Fig. 7. The inexion point evident in this
diagram around 1088 K is associated with VCl2 melting.
On what touches the synthesis of VOCl3 , a reaction path can be
proposed (Eq. (10)), in that VOCl is formed rst, which then reacts
to give VOCl2 which transforms to VOCl3 . The G T diagrams
associated with these reactions are presented in Fig. 8. The inexion
point at 784 K is related to the sublimation of VOCl2 , and at 1393 K to
the sublimation of VOCl. According to the G T curves presented
in Fig. 8, it can be deduced that the reaction steps will follow the
proposed order only for temperatures higher than 1053 K, gaseous

0.5VOCl + 0.5Cl2 = 0.5VOCl3

(11)

2.1.1.2. Predominance diagrams and G T diagrams. Although


reaction mechanisms lie outside the eld of chemical thermodynamics, if a reliable data-base is available, the stability of the
individual species taking part in a specic mechanism can be judged
based on thermodynamic equilibrium computations. In the case of
the VOCl system, for example, this means that the sequence of
appearance of chlorides and or oxychlorides as predicted by the
predominance diagram at a specic temperature must be consistent as the one deduced by the G T plots analysis for each
reaction taking part in the proposed mechanism. If the formation
of a compound is supported by the G T approach and the same
species is not present in the calculated predominance diagram,
it should be classied as metastable for a short period of time,
the species can be present, but as equilibrium is approached, it
decomposes resulting in the more stable species according to the
predominance diagram. Therefore, if a compound is present in the
predominance diagram, its formation must be supported by the
thermodynamically consistent G T plots.
Considering the formation of VCl4 , the reaction path of Eq. (8) is
valid only for temperatures lower than 953 K or higher than 1539 K.
In the lower temperature range this mechanism suggests that once

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Fig. 10. Predominance diagram for the system VOCl at 900 K.

Fig. 12. Predominance diagram for the system VOCl at 1073 K.

Fig. 11. Predominance diagram for the system VOCl at 1673 K.

Fig. 13. Predominance diagram for the system VOCl at 1273 K.

VCl2 is formed in the solid state, it should be converted to solid


VCl3 with increasing P(Cl2 ). Next, VCl3 reacts further resulting in
gaseous VCl4 .
For temperatures higher than 1539 K, VCl2 must be formed in
the liquid state, being next converted to gaseous VCl3 , which nally
results in gaseous VCl4 . These sequences of vanadium chlorides
formation are exactly the same as observed in the predominance
diagrams calculated at 900 K (Fig. 10) and 1673 K (Fig. 11). Between
953 K and 1539 K the mechanism embedded in Eqs. (8) and (9) suggests that VCl2 reacts directly resulting in gaseous VCl4 . No VCl3 is
formed in this case. This is in fact the chlorination sequence predicted by the predominance diagrams calculated at 1073 K (Fig. 12)
and 1273 K (Fig. 13). In these diagrams, no VCl3 eld is present, and
by starting at the liquid VCl2 eld, the stability area of gaseous VCl4
is nally reached as P(Cl2 ) achieves higher values.

A closer look on the predominance diagrams of Figs. 1013


indicates that, with the exception of gaseous VOCl3 , no other
oxychloride appears (VOCl, VOCl2 , or VO2 Cl), suggesting that the
mentioned species is the only one stable vanadium oxychloride
for temperatures between 900 K and 1673 K. Therefore, considering the present mechanism for the formation of the oxychloride
VOCl3 in the mentioned temperature range (Eqs. (10) and (11)),
the other possible oxychlorides (VOCl and VOCl2 ) could be considered as metastable species. Such result is not a mere consequence
of the fact that during the computations developed in the present
topic the oxychlorides are treated as pure compounds, but are also
contained in the results achieved through speciation calculations
included on topic (Section 2.1.2.2), where besides VCl3 and VCl4 ,
VOCl3 is the only one oxycloride found in the gas phase. Another
possible explanation is the fact that the oxychlorides formation may
not happen by the progressive addition of chlorine atoms, but it

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

would be accomplished through reactions, where one of the already


present vanadium chlorides (VCl4 , VCl3 or VCl2 ) participate. It is
worthwhile to mention that this alternative was addressed on topic
(2.1.2.3), where it was considered that VOCl3 could be produced
through the direct oxidation of one of the chlorides.
The results included on the present topic clearly illustrate how
the intimate relation between G T and predominance diagrams
can help in the reaction mechanisms discussions. The reliability
of the computations is directly linked to the quality of the Gibbs
energy model used for describing the thermodynamic properties
of each considered phase. If a high quality data-base is employed,
it is expected that a consistent equilibrium approach would be
found in both diagrams (G T and predominance diagrams). In
the present work this situation is exemplied for the formation of
vanadium chlorides according to mechanism represented by Eqs.
(8) and (9) and Figs. 6, 7 and 1013.
2.1.1.3. VOCl3 formation. According to the information contained
on the predominance diagrams of Figs. 1013, depending on the
partial pressure of O2 imposed, the formation of VOCl3 can be realized either from the direct reaction of Cl2 with a vanadium oxide
or the oxidation of some vanadium chloride (VCl4 , VCl3 or VCl2 ). In
the rst case the VOCl3 phase eld is achieved as P(Cl2 ) gets higher
for a xed P(O2 ) value, and in the last case the same eld is reached
as P(O2 ) achieves higher values P(Cl2 ) xed.

Fig. 14. Partial pressure of VOCl3 as a function of P(O2 ) at 900 K.

0.5V2 O5 + 1.5Cl2 = VOCl3 + 0.75O2


VO2 + 1.5Cl2 + VOCl3 + 0.5O2
1/3V3 O5 + 1.5Cl2 = VOCl3 + 1/3O2

(12)

1/2V2 O3 + 1.5Cl2 = VOCl3 + 0.5O2


Let us take a look rst on the reaction of vanadium oxides and
Cl2 . At each temperature, depending on the P(O2 ) value imposed,
the valence of vanadium changes, and VOCl3 can be produced
through one of the reactions represented by Eq. (12). The thermodynamic viability of such transformations can be studied based on
the computation of the partial pressure of VOCl3 found in equilibrium at a specic temperature under variable P(O2 ). The calculated
P(VOCl3 ) values can be easily obtained through solving individually
the equilibrium equations associated with reactions represented by
Eq. (12), considering in all cases P(Cl2 ) equal to 1 atm and the vanadium oxides as pure compounds (unity activity). For each P(O2 )
interval, the curve associated with the most stable vanadium oxide
correspond to the lowest partial pressure of VOCl3 . The mentioned
vanadium oxide must have the lowest molar Gibbs energy among
all possible valence states considered. As the oxide is a reactant,
the reduction of its Gibbs energy results in a more positive molar
reaction Gibbs energy, which in the end leads to a lower chemical equilibrium constant. Concomitantly, the calculated P(VOCl3 )
reduces for the same P(O2 ) value imposed. At 900 K, VOCl3 can
be produced from one of the vanadium oxides only for ln P(O2 )
equal or higher than 50.22. At 1273 K, this value becomes 32.42.
These quantities limit the region of physical validity of the curves
drawn of in Figs. 14 and 15. For lower P(O2 ) values, VOCl3 can only
be produced through oxidation of some vanadium chloride (VCl2 ,
VCl3 or VCl4 ), a fact that will be considered in more detail latter in
this section. At both temperatures studied, for ln P(O2 ) < 5, significant partial pressures of VOCl3 are computed, which suggests that
the formation of the mentioned oxychloride is associated with an
expressive thermodynamic driving force.
The present method can also be used for comparing the thermodynamic stability of two chlorinated products. This fact was
explored on topic (Section 2.1.2.1), where the stability of VOCl3 and
VCl4 produced through the chlorination of V2 O5 in the presence of
graphite is discussed. There, the interpretation of the data is the

Fig. 15. Partial pressure of VOCl3 as a function of P(O2 ) at 1273 K.

opposite from the one used for the discussion of the information
contained in Figs. 14 and 15. The most stable vanadium chlorinated compound is associated with the highest vapor pressure
value computed at each condition, as in this case we are speaking from product molecules. If the stability of one of the products is
higher than the other, its Gibbs energy must be lower, reducing the
molar Gibbs energy of the reaction in that it takes part. The equilibrium constant grows, and so the partial pressure of the chlorinated
product. It is interesting to see that the inexion points contained in
the continuous curves (PVOCl3 stability locus) represented on the
diagrams calculated at 900 K (Fig. 14) and 1273 K (Fig. 15) correspond exactly to the P(O2 ) values, where the equilibrium between
two vanadium oxides is established. Such fact can be readily understood, as the linear combination of the chlorination reactions of
two oxides associated with neighboring stability elds, results in
the chemical equation representing the equilibrium between the

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Table 3
P(O2 ) equilibrium value under the presence of an excess of graphite.
T (K)

Ln P(O2 )

900
1273
1673

53.04
42.38
37.17

Fig. 17. Partial pressure of VOCl3 as a function of P(Cl2 ) at 1273 K.

Fig. 16. Partial pressure of VOCl3 as a function of P(Cl2 ) at 900 K.

two mentioned vanadium oxides. This can be illustrated for the


chlorination of V2 O5 and VO2 (Eq. (13)).
(0.5V2 O5 + 1.5Cl2 = VOCl3 + 0.75O2 )
(VO2 + 1.5Cl2 = VOCl3 + 0.5O2 ) =

(13)

0.5V2 O5 = VO2 + 0.25O2


Now let us consider the possible formation of VOCl3 under
much more reductive conditions, as in the case where an excess of
graphite is added to the reaction system under study. If this component is present in excess, P(O2 ) should be limited to very low values.
Under these conditions, according to the predominance diagrams
constructed on topic (Section 2.1.1.2) (Figs. 1013), the mechanism
for VOCl3 formation could be dened by two steps. First a chloride is
produced (VCl2 , VCl3 or VCl4 ), which is then next oxidized to VOCl3
(Eq. (14)).
VCl2 + 0.5Cl2 + 0.5O2 = VOCl3
VCl3 + 0.5O2 = VOCl3

(14)

VCl4 + 0.5O2 = VOCl3 + 0.5Cl2


The thermodynamic tendency associated with the occurrence
of the reactions represented by Eq. (14) can be studied by employing an analogous strategy as used for discussing the formation of
VOCl3 under variable P(O2 ). This time, the partial of any gaseous
chloride (VCl3 or VCl4 ), if present, is xed at 1 atm, and the P(O2 )
value imposed is calculated according to equilibrium between a
gaseous phase containing CO, CO2 and O2 in the presence of an
excess of graphite (Boudouard equilibrium) at each imposed temperature (Table 3). In all cases (Figs. 1618), signicant partial
pressures of VOCl3 are computed in comparison with the low P(O2 )
value imposed, indicating that the proposed reactions are thermodynamically possible. It is also interesting to note that the idea of

Fig. 18. Partial pressure of VOCl3 as a function of P(Cl2 ) at 1673 K.

the occurrence of oxidation reactions as those represented by Eq.


(14) reappears in the discussion of topic (Section 2.1.2.2), as a plausible explanation of the gas phase composition variations as P(O2 )
is systematically varied.
The inexion points present in the curves of Figs. 1618 are
associated with P(Cl2 ) values, which are identical to the partial
pressures related to the equilibrium between two successive chlorinated compounds, as described by the predominance diagrams.
Again, the stability locus is determined by the minimum calculated
partial pressures. The diagrams are lower limited by the horizontal lines, which determine the minimum vapor pressure of Cl2 for
the occurrence of the oxidation reactions considered (1673 K
ln P(Cl2 ) = 7.2, 1273 K ln P(Cl2 ) = 10.4, 900 K ln P(Cl2 ) = 15.2).
For lower partial pressures the oxidation of the most stable chloride
results in a vanadium oxide. Therefore, depending on the partial pressure of oxygen imposed, the formation of VOCl3 can be

10

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Fig. 4). Moreover, according to the data presented in Fig. 19, VCl4 is
the compound with the highest partial pressure for both specied
temperatures while, as temperature gets higher, their concentrations show appreciable reduction. These results will be conrmed
through construction of speciation diagrams for the gas phase, a
task that will be accomplished on topic (Section 2.1.2.2).

Fig. 19. Ln (P(VCl4 )) and Ln (P(VOCl3 )) as a function of Ln (P(CO)).

accomplished either through the oxidation of a vanadium chloride


(lower O2 activity Figs. 1618) or the chlorination of the most stable vanadium oxide under the specied oxygen potential (higher
O2 activity Figs. 14 and 15).
2.1.2. Relative stability of VCl4 and VOCl3
As is evident from the discussion developed on topic (Section
2.1.1), the chlorinated compounds VCl4 and VOCl3 are, respectively,
the most stable vanadium chloride and oxychloride to be found
in the gas phase as the atmosphere becomes concentrated in Cl2 .
The relative stability of these two chlorinated compounds will be
rst accessed on topic (Section 2.1.2.1) by applying the method
introduced by Kang and Zuo [27], which is indeed a very similar
strategy as the one used for the discussion of possible reaction paths
for the formation of VOCl3 on topics (Sections 2.1.1.2 and 2.1.1.3), as
it relies entirely on the computation of G values, and secondly,
on topic (Section 2.1.2.2), through the construction of gas phase
speciation diagrams.
2.1.2.1. Method of Kang and Zuo. The concentrations of VCl4 and
VOCl3 can be directly computed by considering that each chlorinated compound is generated independently. It will be assumed
that the inlet gas is composed of pure Cl2 (P(Cl2 ) = 1 atm). Further, two temperature values were investigated, 1073 K and 1373 K.
At these temperatures, the presence of graphite makes the atmosphere richer in CO, so that for the computations the following
reactions will be considered:
V2 O5 + 4Cl2 + 5C = 2VCl4 + 5CO
V2 O5 + 3Cl2 + 5C = 2VOCl3 + 3CO

(15)

The concentrations of VOCl3 and VCl4 can then be expressed as


a function of P(CO) and temperature according to Eq. (16).
PVCl4 =

PVOCl3 =

5 K ln P
PCO
1
VCl4 =

ln K1
5
+ ln PCO
2
2

3 K ln P
PCO
2
VOCl3 =

3
ln K2
+ ln PCO
2
2

(16)

where K1 and K2 represent, respectively, the equilibrium constants


for the reactions associated with the formation of VCl4 and VOCl3
(Eq. (15)). By applying Eq. (16) the partial pressure of VCl4 and
VOCl3 were computed as a function of P(CO). The results were plotted in Fig. 19. The signicant magnitude of the partial pressure
values computed for VOCl3 and VCl4 is a consequence of the huge
negative standard Gibbs energy of reaction associated with the formation of these species in the temperature range considered (see

2.1.2.2. Gas phase speciation. The construction of speciation diagrams for the gas phase enables the elaboration of a complementary
picture of the chlorination process in question. The word speciation means the concentration of all species in gas. This brings
another level of complexity to the quantitative description of
equilibrium, as the species build a solution, and as so, their concentrations must be determined at the same time. This sort of
information can only arises if one solves the system o equilibrium
equations associated to all possible chemical reactions involving
the species that form the gas. For the present system (VOClC)
this task becomes very tricky, as the number of possible species
present is pretty signicant (e.g. CO, CO2 , O2 , VCl2 , VCl3 , VCl,
VCl4 , VOCl3 , VOCl, VO2 Cl and VOCl2 ), and so the number of possible chemical reactions connecting them. So, we must think in
another route for simultaneously computing the concentration of
the gaseous species produced by our chlorination process. The only
possible way left consists in minimizing the total Gibbs energy of
the system.
The equilibrium state is dened by xing T, P, n(V2 O5 ), P(O2 ) and
P(Cl2 ). The number of moles of V2 O5 is xed at one. If graphite is
present in excess, the partial pressure of O2 is controlled by according to the Boudouard equilibrium, so that, its presence forces P(O2 )
to attain very low values (typically lower than 102 atm). The total
pressure is xed at 1 atm and T varies in the range between 1000 K
and 1473 K. An excess of graphite is desirable, so that the chlorination reactions can achieve a considerable driving force at the
desired conditions. Computationally speaking, this can be done in
two ways. One possibility is to dene an amount of carbon much
greater than the number of moles of V2 O5 . Other possibility, which
has been made accessible through modern computational thermodynamic software, consists in dening the phase solid graphite
as xed with a denite amount. As carbon in graphite is considered
to be pure, the later alternative is equivalent of saying that carbon
is present in the system with a chemical activity equal to one. The
equilibrium compositions (intensive variables) are not a function of
the amount of phases present (size of the system), depending only
of temperature and total pressure. So we are free to choose any
suitable value we desire for the xed amount of carbon present,
such for example zero. This last alternative was implemented in
the computations conducted in the present topic.
In Fig. 20, the number of moles of gas produced was plotted
as a function of P(Cl2 ) for T equal to 1073 K, 1273 K, and 1473 K,
1000 K, 1073 K, 1173 K, and 1373 K. The partial pressure of O2 was
xed at 1.93 1022 atm, and the partial P(Cl2 ) is varied between
3.6 107 atm and 0.61 atm. Each curve represented in Fig. 20 is
dened by three stages. First, for very low values of P(Cl2 ), no gas is
formed. At these conditions VCl2 (l) is present in equilibrium with.
The equilibrium ensemble does not experience any modication
until a critical P(Cl2 ) value is reached, at which a discontinuity can
be evidenced. The gas phase appears and for any P(Cl2 ) higher than
the critical one, the number of moles VCl2 (l) becomes equal to zero.
This condition denes the second stage, where for higher P(Cl2 ) values the gas composition changes accordingly, through forming of
chlorides and oxychlorides. Finally, a P(Cl2 ) value is reached, where
all capacity of the system for forming chlorinated compounds is
exhausted, and the effect of adding more Cl2 is only the dilution of
the chlorinated species formed. As a consequence, the number of
mole of gas phase experiences a signicant elevation.

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Fig. 20. Number of moles of gas as a function of P (Cl2 ).

Table 4
Equilibrium constants at 1073 K for the reactions represented by Eq. (17).
Chemical reaction

Equilibrium constant

VCl2 + Cl2 VCl4


VCl3 + 0.5Cl2 VCl4

1.95 105
2.12 103

At 1073 K, for example, Fig. 21 describes the effect of P(Cl2 ) over


the gas phase composition during the second and third stages. We
see that the mol fraction of VCl4 raises (second stage) and after
achieving a maximum value starts to decrease (third stage). The
concentration variations during the second sage can be ascribed
to the occurrence of reactions represented by Eq. (17), which have
at 1073 K equilibrium constants much higher than unity (Table 4).
The reduction of the mol fraction of VOCl3 can be understood as
a dilution effect, which is motivated by the elevation of the mol
fractions of VCl4 and Cl2 .
VCl2 + Cl2 = VCl4
VCl3 + 0.5Cl2 = VCl4

(17)

11

Besides P(Cl2 ), temperature should also have an effect over the


composition of the gas phase. This was studied as follows. Six temperature values were chosen in the range between 1073 K and
1473 K. Next, for each temperature the critical P(Cl2 ) value (the one
associated with the formation of the rst gas molecules) is identied. The concentrations of the most stable gaseous species at the
computed critical P(Cl2 ) value were then calculated and presented
in Table 5. During the calculations the partial pressure of O2 was
xed at 1.93 1022 atm.
As expected, the mol fraction of CO is greater than the mol fraction of CO2 for the entire temperature range studied. Also, the
chloride VCl4 has the highest concentration at 1073 K, a phase
which occupies a large area of the predominance diagram at this
temperature (Fig. 12). As temperature attains higher values, the
mol fraction of VCl4 and VOCl3 become progressive lower and the
atmosphere more concentrated in VCl2 and VCl3 . So, at 1473 K
the situation is signicant different from the equilibrium state
observed at 1073 K. This behavior is again consistent with the information contained on the predominance diagrams (Figs. 1013)
where can be seen that the stability elds of VCl4 (g) and VOCl3 (g)
shrink while the area representing the phase VCl3 (g) grows, occupying at 1673 K a visible amount of the diagrams space (Fig. 11).
It is worthwhile to mention that a more detailed look on the
results presented in Table 5 seems to incorporate apparent inconsistencies. (i) The minimum partial pressure of Cl2 for the formation
of pure VCl4 (g) at 1073 K (Fig. 12) is higher than the critical pressure
for the formation of the rst gaseous species at this temperature
(Fig. 20). (ii) Measurable amounts of VCl3 (greater or equal to 0.1)
were detected for temperatures higher than 1100 K (Table 5) but no
VCl3 (g) eld was observed in the predominance diagram computed
at 1273 K (Fig. 13). (iii) Also, no eld associated with the formation
of VCl2 (g) could be detected even at 1673 K (Fig. 11) but the speciation computation predicts its presence in measurable amounts at
the last temperature (x(VCl2 ) = 0.14) (Table 5). All these thermodynamic values differences are a consequence of the fact that the pure
molar Gibbs energy of each component is higher than its chemical
potential in the ideal gas solution, the former model being used for
the predominance diagrams construction while the later is applied
to the speciation calculations. Therefore, the driving force for the
formation of the gaseous compounds is reduced accordingly to Eq.
(18) [28].
g

VCl gVCl = RT ln xVCl < 0


3

(18)

Another possible type of computation is to study the effect


of P(O2 ) over the composition of the gas phase. This variable is
restricted by the fact that the amount of graphite is xed. So there

Fig. 21. Concentration of vanadium chlorides and oxychlorides as a function of P(Cl2 ) at 1073 K.

12

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Table 5
Composition of the rst gas formed as a function of temperature.
T (K)

X(CO)

X(CO2 )

X(VOCl3 )

X(VCl2 )

X(VCl3 )

X(VCl4 )

1073
1100
1200
1300
1373
1400
1473

0.16
0.12
4.21 102
1.76 102
1.00 102
8.26 103
5.07 103

3.64 103
1.23 103
3.36 105
1.60 106
2.29 107
1.17 107
2.18 108

1.95 102
1.02 102
1.08 103
1.45 104
3.69 105
2.27 105
6.22 106

1.74 103
2.61 103
9.87 103
2.98 102
5.97 102
7.6 102
0.14

9.27 102
0.12
0.26
0.43
0.56
0.59
0.66

0.72
0.75
0.69
0.52
0.37
0.32
0.19

Table 6
Gas phase speciation as a function of P(O2 ) at 1373 K.
P(O2 ) (atm)

X(CO)

X(CO2 )

X(VOCl3 )

X(VCl2 )

X(VCl3 )

X(VCl4 )

1.30 1024
5.24 1022
1.56 1018

8.22 104
1.65 102
0.902

1.54 109
6.22 107
1.85 103

3.05 106
6.03 105
3.21 104

0.0597
0.0588
5.72 103

0.56
0.55
0.054

0.38
0.37
0.036

is a maximum value of P(O2 ) at each temperature for which the


thermodynamic modeling remains consistent with the Boudouard
equilibrium (Eq. (4)) and the computation can be performed. By xing the temperature at 1373 K, the upper limit for P(O2 ) has shown
to be equal to 1.56 1018 atm and the value of P(Cl2 ) associated
with the appearance of the rst gaseous molecules is identied as
2.05 104 atm. The composition of the gas phase is then computed
by xing P(Cl2 ) at 2.05 104 atm. Three different P(O2 ) levels were
studied, 1.3 1024 atm, 5.24 1022 atm, 1.56 1018 atm. The
results are presented in Table 6.
The mol fractions of CO and CO2 gets higher for higher values
of P(O2 ). This is consistent with the dislocation of the equilibrium
represented by Eq. (4) in the direction of the formation of the two
carbon oxides. Also, the Boudouards equilibrium demands that
at the chosen temperature (1373 K) the atmosphere is more concentrated in CO. This was indeed observed for each equilibrium
state investigated. It is interesting to observe that for P(O2 ) varying
between 5.24 1022 atm and 1.56 1018 atm the mol fraction of
CO and CO2 experience a much higher variation in comparison with
the one observed for lower P(O2 ) values.
In the case of the vanadium chlorides and oxychlorides an interesting trend is evidenced. The concentration of VOCl3 grows and
of VCl2 , VCl3 and VCl4 reduce appreciably for the same O2 partial pressure range. The concentration variations associated with
the vanadium chlorinated compounds is analogous to the variations observed in the concentrations of CO and CO2 . For P(O2 ) lower
than 5.24 1022 atm the variations are much less signicant. To
get a better picture of the trend observed for the chlorides and
oxychlorides, their concentrations were plotted as a function of
P(O2 ), which was varied in the range spanned by the data of Table 4
(Figs. 2224).
The variations depicted in Figs. 2224 are consistent with the
occurrence of reactions represented by Eq. (19). As P(O2 ) achieves
higher values, it reacts with VCl3 , VCl2 and or VCl4 resulting in
VOCl3 . Such phenomena could explain the signicant reduction of
VCl3 , VCl4 and VCl2 concentrations, and the concomitant elevation
of the VOCl3 mol fraction.

Fig. 22. Mol fraction of VCl3 and VCl4 as a function of P (O2 ).

VCl3 + 0.5O2 = VOCl3


VCl2 + O2 + VCl4 = 2VOCl3

(19)

VCl4 + 0.5O2 = VOCl3 + 0.5Cl2


The participation of VCl4 in the second reaction is supported by
the fact that its equilibrium concentration lowering is more sensible
to ln P(O2 ) than observed for VCl3 (Fig. 22). The consumption of
VCl4 by the second reaction is also consistent with the maximum

Fig. 23. Mol fraction of VOCl3 and VCl2 as a function of P (O2 ).

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

13

Fig. 24. Mol fraction of VOCl3 as a function of P (O2 ).


Fig. 25. Total enthalpy as a function of Cl2 partial pressure.
Table 7
Equilibrium constants at 1373 K for reactions represented by Eq. (19).
Chemical reaction

Equilibrium constant

VCl3 + 0.5O2 = VOCl3


VCl2 + O2 + VCl4 = 2VOCl3
VCl4 + 0.5O2 = VOCl3 + 0.5Cl2

8.01 106
1.89 1013
1.01 105

observed in the curve obtained for VOCl3 concentration (Fig. 24).


As less VCl4 is available, less VOCl3 can be produced.
The occurrence of reactions represented by Eq. (19) is supported
by classical thermodynamics, as the equilibrium constant computed at 1373 K assume values appreciably greater than unity (see
Table 7). It is indeed interesting to recognize that the same transformations were considered as plausible reaction paths for the VOCl3
formation based on the oxidation of one of the possible vanadium
chlorides (see Section 2.1.1.3).
2.1.2.3. V2 O5 chlorination enthalpy. For the implementation of an
industry process based on chemical reaction is fundamental to
know the amount of heat generated or absorbed from that. Exothermic processes (heat is released) reach higher temperatures, and
frequently demand engineering solutions for protecting the oven
structure against the tremendous heat generated by the chemical phenomena. In this context, endothermic processes (heat is
absorbed) are easier controlled, but the energy necessary to stimulate the reaction must be continuously supplied, making the energy
investment larger.
The variation of the total enthalpy of the system for the chlorination process in question was calculated as a function of P(Cl2 )
(Fig. 25). The partial pressure of O2 was xed at 1.93 1022 atm
and four temperature levels were studied, 1000 K, 1100 K, 1300 K
and 1700 K. It can be seen that the total enthalpy for the process
conducted at 1000 K reduces with the advent of the chlorination
reactions, indicating that the chlorination process is exothermic.
However, the molar enthalpy magnitude is progressively lower up
to a certain temperature where it is zero. Above that, the molar
reaction enthalpy becomes positive, and Fig. 25 illustrates its value
for 1700 K.
All these facts are in agreement with the results presented in
Table 5. As temperature gets higher, the mol fractions of VCl4 and
VOCl3 reduce and that for VCl3 and VCl2 experience a signicant

Fig. 26. Molar reaction enthalpy for the formation of gaseous VCl3 and VCl2 .

elevation. For some temperature between 1300 K and 1373 K the


mol fractions of VCl4 and VCl3 assume equal values. This point is
related to the condition where the chlorination enthalpy is zero. For
higher temperatures, where x(VCl4 ) < x(VCl3 ) the process becomes
progressively more endothermic.
It is interesting to see that he explained behavior is consistent
with the fact that the global formation reactions of VCl3 and VCl2
from liquid V2 O5 in the temperature 10001800 K are associated
with positive molar reaction enthalpies (Fig. 26) and that of VOCl3
and VCl4 with negative molar reactions enthalpies (Fig. 27).
3. Final remarks
In this chapter three different approaches to the chlorination
equilibrium study of an oxide were presented. The rst two are
based on the construction of G T diagrams (Section 2.1.1) and
on the calculations, rst introduced by Kang and Zuo [27] (Section
2.1.2.1), respectively. Both of them take into consideration that each
chlorinated compound is produced independently. The third one
has its fundamental based on the total Gibbs energy minimization

14

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

Fig. 27. Molar reaction enthalpy for the formation of gaseous VOCl3 and VCl4 .

of the reaction system and the gas phase equilibrium composition is calculated considering that the formed species are produced
simultaneously (topic Section 2.1.2.2).
The method based on the construction of G T was applied on
topic (2.1.2) for studying the thermodynamic viability of the reaction between gaseous Cl2 and V2 O5 . The discussion evidenced that
the chlorination is thermodynamically feasible only in the presence of a reducing agent (graphite in the case of the present work)
and was initially focused on the production of VCl4 and VOCl3 . The
same approach was employed for studying the possible mechanisms associated with the formation of VCl4 and VOCl3 . According
to the results (Sections 2.1.1.12.1.1.3), the synthesis of these two
compounds should be subdivided in different stages, which can
vary in nature, depending on the temperature range considered.
The sequence of formation of vanadium chlorides according
to the predominance diagrams computed in space P(Cl2 )P(O2 )
(Figs. 1013) is perfectly consistent with the ndings associated
with the investigation of possible chlorination paths based on
G T curves, as far as these graphics are plotted with thermodynamic consistence (see examples in Figs. 7 and 9). Also, it is
understood that the interpretation of these two diagrams together
could became a very useful didactic contribution.
According to Figs. 6 and 7, between 953 K and 1539 K, VCl4
should be formed directly from VCl2 . This is indeed observed in the
predominance diagrams computed at 1073 K (Fig. 12) and 1273 K
(Fig. 13). Outside the mentioned range, VCl3 should be produced
prior to VCl4 . This was also conrmed in the diagrams computed at
900 K (Fig. 10) and 1673 K (Fig. 11).
In the case of the vanadium oxychlorides, only the VOCl3 eld
was detected in the predominance diagrams, indicating that this
compound is the only stable vanadium oxychloride to be formed
for the conditions imposed. On the other hand, the discussion based
on G T curves (Figs. 8 and 9) indicates that other oxychlorides
(VOCl and VOCl2 ) could be formed prior to VOCl3 . Therefore, if
the information contained in the thermodynamic data-base used is
considered to be reliable enough, these species should be regarded
as metastable, whose decomposition in the way to thermodynamic equilibrium leads exclusively to the formation of VOCl3 , the
only stable oxychloride according to the calculated predominance
diagrams. Depending on the oxygen potential prevailing in the
reactors atmosphere, however, some other routes for the formation
of VOCl3 can be considered. For higher P(O2 ) values, the results presented on topic (Section 2.1.1.3) clearly indicate that the mentioned
oxychloride can be formed through the direct reaction between
Cl2 and the most stable vanadium oxide under the specied

oxygen potential V2 O5 , VO2 , V2 O3 or V3 O5 (Figs. 14 and 15). On


the other hand, if the atmosphere is made reductive enough (low
oxygen activity), the VOCl3 production should happen through the
oxidation of one of the possible vanadium chlorides VCl2 , VCl3 and
or VCl4 (see Eq. (12)), the stimulated transformation depending on
the P(Cl2 ) value imposed (Figs. 1618). In this latter case (reducing atmosphere) it could also be possible an oxide partial reduction
followed by the lower oxide chlorination.
Generally speaking, for an atmosphere rich in Cl2 , VOCl3 and
VCl4 are the most stable vanadium chloride and oxychloride,
respectively, as their formation reactions are associated with G
signicant negative values (Fig. 4). However, just by looking at the
G plots, it is only possible to estimate the specie in greater concentration (VCl4 ) at chemical equilibrium. A quantitative approach
for the relative stability between VOCl3 and VCl4 was carried out
by the implementation of Kang and Zuo method (Section 2.1.2.1).
The results (Fig. 19) indicated that VCl4 should, in fact, have a
higher concentration in comparison with VOCl3 in the temperature
range between 1073 K and 1373 K. Moreover, the data contained in
Fig. 19 suggest that the tendency of formation of both VCl4 and
VOCl3 should reduce as temperature achieves higher values. This
fact is also in agreement with the gas phase speciation results (see
Table 5).
It can be said that both, the method based on the G T
diagrams construction as well as the Kang and Zuo method, incorporate some simplications and are very easy to implement.
However, they lead to only a supercial knowledge of the true
nature of the equilibrium state achievable. Thanks to the development of computational thermodynamic software, more complex
computations can be conducted. For example, by allowing the
chlorides and oxychlorides to build a gaseous solution, the minimization of the total Gibbs energy of the system results in the
direct computation of the mol fraction of each chlorinated species
present in the gas phase (Section 2.1.2.2). This approach can be
seen as a general improvement in comparison with the usually
employed methods, since all equilibrium equations are solved
simultaneously, with the further advantage that there is no need
to formulate a group of independent reactions that cover all possible chemical interactions among the components, a task that can
become very complex for metals, as in vanadium case, which can
produce a family of chlorides and oxychlorides.
The conclusion that carbon strongly promotes the thermodynamic driving force necessary to chlorination and that VCl4 should
be formed preferentially in relation to VOCl3 are perfectly consistent with the results based on the total Gibbs energy minimization
(Tables 5 and 6). However, by the application of this last method it
was possible to go a little further, through investigation of the effect
of P(Cl2 ) over the chlorination enthalpy (Section 2.1.2.3) and studying the effect of temperature, Cl2 and O2 partial pressures over the
concentrations of vanadium chlorides and oxychlorides in the gas
phase (Section 2.1.2.2).
The predictions associated with the effect of temperature over
the gas phase speciation (Table 5) indicate that the mol fractions
of VCl2 and VCl3 grow signicantly in the range between 1073 K
and 1473 K and, as a result, the concentrations of VOCl3 , VCl4 , CO
and CO2 exhibit a signicant reduction. This nding agrees with the
tendency depicted by the predominance diagrams constructed for
the system VOCl, where the VCl4 (g) and VOCl3 (g) elds shrink
and that of VCl3 (g) grows (Figs. 1013). Also, the calculated mol
fraction of CO is at all temperatures much higher than the mol
fraction of CO2 , a fact that is consistent with the establishment of
the Boudouards equilibrium for temperatures higher than 973 K,
where the concentrations of the two mentioned carbon oxides have
the same magnitude.
The fact that the speciation computation indicates appreciable
amounts of gaseous VCl3 for temperatures higher than 1100 K and

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

of gaseous VCl2 for temperatures higher than 1473 K, apparently


contradicting the information contained in the predominance diagrams of Figs. 11 and 13, is a mere consequence of the fact that on
topic (Section 2.1.2.2) the gaseous chlorides build an ideal gas solution. The chemical potentials of VCl3 and also of VCl2 are lower than
their pure molar Gibbs energies. As a result, the species become
more stable in the gaseous solution in relation to the pure state, and
their mol fractions assume higher values for the same temperature
imposed. The same idea explains why VCl4 is formed in signicant
amounts at 1073 K for a P(Cl2 ) value lower than the one observed in
the predominance diagram of Fig. 12 for the equilibrium between
VCl2 (l) and VCl4 (g).
The effect of adding more Cl2 after all vanadium has been converted to gaseous chlorinated compounds is also consistent with
the expectations. At 1073 K the results indicate that the mol fraction
of VCl4 grows while all other relevant chlorinated species reduces
(Fig. 21). This can be explained by the reaction of VCl2 and VCl3 with
Cl2 resulting in VCl4 , which have a signicant negative driving force
at 1073 K (Table 4).
The study of the impact of varying P(O2 ) over the gas phase composition at 1373 K indicated that the mol fractions of CO and CO2
experience signicant elevation as P(O2 ) becomes higher, a fact
that is also observed in the case of VOCl3 (Table 6). The concentration of all other chlorinated compounds reduces for the same
studied range of P(O2 ). The inuence of the oxygen chemical activity over the gas phase speciation can be explained by a group of
proposed reactions between VCl4 , VCl3 and VCl2 with O2 resulting
in VOCl3 (Eq. (19)). All these reactions have equilibrium constants
much higher than one, indicating an expressive thermodynamic
driving force at the standard conditions for the specied temperature (Table 7). The same reactions were put forward as a possible
method for the production of VOCl3 from one of the stable vanadium chlorides (VCl2 , VCl3 or VCl4 ) when low partial pressures of
O2 are imposed (Section 2.1.1.3).
The conclusions about the exothermic nature of the chlorination
process (Section 2.1.2.3) in the temperature range between 1000 K
and 1300 K and the observation that it becomes progressively more
endothermic as 1700 K is approached (Fig. 25), are perfectly consistent with the fact that the atmosphere becomes progressively
diluted in VCl4 and VOCl3 , whose formations are associated with
negative molar enthalpies (Fig. 27) and becomes richer in VCl2 and
VCl3 , whose molar enthalpy of formation are considerably positive
(Fig. 26).

4. Conclusions
According to the obtained results, the chlorination of V2 O5 in the
presence of carbon is a thermodynamic feasible process in the range
between 900 K and 1600 K. During the process, VCl4 and VOCl3 are,
respectively, the most stable vanadium chloride and oxychloride
formed. In the case of VCl4 it has been observed that its formation
can be explained as a progressive path starting from VCl2 , which is
next converted to VCl3 or directly to VCl4 , depending on the temperature. In relation to VOCl3 two possibilities seems to be present.
The rst one, starts with VOCl, which, depending on the reaction
temperature, can be converted to VOCl3 either directly or passing
through VOCl2 . In the second one, VOCl3 is formed from one of
the vanadium oxides (higher oxygen activity) or chlorides (lower
oxygen activity).
The speciation calculations indicate that the concentration of
the gas phase generated during the chlorination process is very
sensitive to variation in temperature, P(Cl2 ) and P(O2 ). As temperature achieves higher values, the mol fractions of VCl3 and VCl2
grow while VCl4 and VOCl3 reduce. Regarding the effect of P(Cl2 ),
as this parameter achieves higher values, the formation of VCl4 is

15

stimulated as VCl2 and VCl3 reacts with the Cl2 available. The effect
of increasing the oxygen potential P(O2 ) can be seen as a way of
promoting the formation of VOCl3 through the oxidation of one of
the vanadium chlorides already present (VCl2 , VCl4 or VCl3 ). Moreover, the chlorination process is considerably exothermic at 1000 K,
but becomes progressively less exothermic at higher temperatures.
This fact is correlated with the mol fraction reductions of VCl4 and
VOCl3 and the concomitant increasing of those for VCl3 and VCl2 .
Finally, we can conclude that the study of the equilibrium states
achievable through the reaction between a transition metal oxide
and gaseous Cl2 , can be now approached through the implementation of methods of different complexity levels. The most general
one, in which the total Gibbs energy of the reaction system is minimized, enables the construction of a more detailed (quantitative)
picture of the equilibrium states involved. However, as it is evident
from the comparisons explained above, the most general method
must be consistent with the tendencies predicted by simpler calculation procedures (qualitative picture), as the one dened by the
construction of G T diagrams, or the simple solution of individual chemical equilibrium equations. Although the discussion in this
work focuses exclusively on the V2 O5 chlorination, the followed
approach can be applied to any reaction system.

Acknowledgments
The authors are especially grateful for the nancial support
and scholarships provided by CAPES, CNPQ and FAPERJ during the
development of the present work.

References
[1] L. Zhang, et al., Rare earth extraction from bastnaesite concentrate by stepwise
carbochlorinationchemical vapor transport-oxidation, Metall. Mater. Trans. B
35 (2) (2004) 217221.
[2] E. Cecchi, et al., A feasibility study of carbochlorination of chrysotile tailings,
Int. J. Miner. Process. 93 (3/4) (2009) 278283.
[3] I. Gaballah, M. Djona, Recovery of Co, Ni, Mo, and V from unroasted spent
hydrorening catalysts by selective chlorination, Metall. Mater. Trans. B 26 (1)
(1995) 4150.
[4] E.A. Brocchi, F.J. Moura, Chlorination methods applied to recover refractory
metals from tin slags, Miner. Eng. 21 (2) (2008) 150156.
[5] K. Murase, et al., Recovery of vanadium, nickel and magnesium from a y ash of
bitumen-in-water emulsion by chlorination and chemical transport, J. Alloys
Compd. 264 (1/2) (1998) 151156.
[6] D. Neff, Environmentally acceptable chlorination processes, in: Aluminum Cast
House Technology: Theory & Practice, Australasian, Asian, Pacic Conference,
1995, pp. 211225.
[7] D. Mackay, Is chlorine the evil element? Environ. Sci. Eng. (1992) 4952.
[8] G. Micco, A.E. Bohe, H.Y. Sohn, Intrinsic kinetics of chlorination of WO3 particles
with Cl2 gas between 973 K and 1223 K (700 C and 950 C), Metall. Mater. Trans.
B 42 (2) (2011) 316323.
[9] J.P. Gaviria, A.E. Bohe, Carbochlorination of Yttrium oxide, Thermochim. Acta
509 (1/2) (2010) 100110.
[10] M.R. Esquivel, A.E. Boh, D.M. Pasquevich, Carbochlorination of samarium
sesquioxide, Thermochim. Acta 403 (2003) 207218.
[11] M.W. Ojeda, J.B. Rivarola, O.D. Quiroga, Study of the chlorination of molybdenum trioxide mixed with carbon black, Miner. Eng. 15 (2002) 585591.
[12] H.H. Kellog, Thermodynamic relationships in chlorine metallurgy, J. Met. 188
(1950) 862872.
[13] C.C. Patel, G.V. Jere, Some thermodynamical considerations in the chlorination
of ilmetite, Trans. Metall. Soc. AIME 218 (1960) 219225.
[14] R.F. Pilgrim, T.R. Ingraham, Thermodynamics of chlorination of iron, cobalt,
nickel and copper sulphides, Can. Metall. Q. 6 (4) (1967) 333346.
[15] N. Sano, G. Belton, The thermodynamics of chlorination of vanadium pentoxide,
Trans. Jpn. Inst. Met. 21 (9) (1980) 597600.
[16] L. Brewer, B.B. Ebinghaus, The thermodynamics of the solid oxides of vanadium,
Thermochim. Acta 129 (1988) 4955.
[17] R.E. McCarley, J.W. Roddy, The vapor pressures of vanadium(II) chloride,
vanadium(III) chloride, vanadium(II) bromide, and vanadium(III) bromide by
Knudsen effusion, Inorg. Chem. 3 (1) (1964) 6063.
[18] N. Saunders, A.P. Miodownik, Calphad: Calculation of Phase DiagramsA
Comprehensive Guide, Elsevier Science Ltd., The Boulevard, Langford Lane,
Kidlington, Oxford, 1998.
[19] O. Knacke, O. Kubashcewski, K. Hesselmann, Thermochemical Properties of
Inorganic Substances, Springer-Verlag, Berlin, 1991.

16

E.A. Brocchi et al. / Thermochimica Acta 559 (2013) 116

[20] H. Oppermann, Gleichgewichte mit VOCl3 , VO2 Cl, und VOCl2 , Z. Anorg Allg.
Chem. 331 (3/4) (1967) 113224.
[21] A. Hackert, Plies, R. Gruehn, Nachweis und thermochemische Charakterisierung
des Gasphasenmolekuls VOCl, Z. Anorg. Allg. Chem. 622 (1996) 16511657.
[22] D.L. Hildenbrand, et al., Thermochemistry of the gaseous vanadium chlorides
VCl, VCl2 , VCl3 , and VCl4 , J. Phys. Chem. A 112 (2008) 99789982.
[23] HSC version 6.0 data-base.
[24] E. Allain, M. Djona, I. Gaballah, Kinetics of chlorination and carbochlorination
of pure tantalum and niobium pentoxides, Metall. Mater. Trans. B 28 (1997)
223232.

[25] J. Gonzallez, et al., -Ta2 O5 carbochlorination with different types of carbo


canadian, Metall. Q. 41 (1) (2002) 2940.
[26] P.K. Jena, E.A. Brocchi, J. Gonzalez, Kinetics of low-temperature chlorination of
vanadium pentoxide by carbon tetrachloride vapor, Metall. Mater. Trans. B 36
(2) (2005) 195199.
[27] S.X. Kang, Y.Z. Zuo, Chloridizing roasting of complex material containing low
tin and high iron at high temperature, Kunming Metall. Res. Inst. Rep. 89 (3)
(1989).
[28] D. Robert, Thermodynamics in Materials Science, Second ed., Taylor and Francys, NY, USA, 2006.

Anda mungkin juga menyukai