You are on page 1of 7

Hydrometallurgy 105 (2010) 96–102

Contents lists available at ScienceDirect

Hydrometallurgy
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / h yd r o m e t

Thermal analysis and kinetic modeling of manganese oxide ore reduction using
biomass straw as reductant
Yuna Zhao, Guocai Zhu ⁎, Zhuo Cheng
Institute of Nuclear and New Energy Technology, Tsinghua University, Beijing 102201, PR China

a r t i c l e i n f o a b s t r a c t

Article history: In this study biomass straw was used as a reductant and fuel for the reduction of manganese oxide ore at low
Received 14 January 2010 temperature of up to 600 °C. XRD analysis of original and roasted ore was made to identify the reduction
Received in revised form 2 August 2010 mechanism. The reduced products changed with the different amount of straw, and the manganese oxide ore
Accepted 12 August 2010
was successfully reduced to acid-soluble MnO with over 30% biomass straw. Several mixtures with different
Available online 20 August 2010
weight ratios of biomass straw and ore were subjected to thermal analysis, showing that reduction reaction
Keywords:
commenced at approximately 310 °C and was almost completed at 600 °C. Thermal analysis data further
Manganese oxide ore confirmed the individual thermal reaction regions associated with developments of individual manganese
Biomass phases during the heating and were used to calculate the corresponding kinetics of the biomass reduction
Reaction kinetics process. The activation energy (E) was calculated to be 470 kJ/mol at 300–390 °C, 510 kJ/mol at 400–480 °C,
Thermodynamics and 430 kJ/mol at 490–640 °C respectively.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction Recently, the use of biomass as a new reductant used in


manganese ore process has attracted more attention. Tian et al.
Manganese is a strategic element that has several important (2010) utilized corncob as a reductant to extract manganese from
industrial applications such as steel production, carbon–zinc batteries low-grade manganese dioxide ores in sulfuric acid solution, and the
production, fertilizers, as well as colorants for bricks, dyes and leaching efficiency of manganese reached 92.8% at 85 °C.
medicines (Sahoo et al., 2001). The world annual consumption of Our research group found that manganese oxide ore can be
manganese is above 1,500,000 ton and it is destined to increase reduced largely by roasting with biomass straw at temperatures
(Zhang, 2007). Pyrolusite is an important manganese mineral in below 650 °C. Biomass straw is a renewable biomass resource and can
manganese ores which is distributed widely in the world (Harris, be commonly found as process waste in agricultural and food
1997). MnO2 is stable in acid or alkaline oxidizing conditions so the industries. Carbon emissions from biomass utilization have the same
recovery of manganese from manganese oxide ore containing value as carbon fixed through photosynthesis by the plant, thus the
pyrolusite must be carried out under reducing condition (Das et al., overall net carbon emissions from biomass utilization makes no
1982; Sahoo and Rao, 1989; Abbruzzese et al., 1990; Ismail et al., addition to greenhouse gases (Zhang et al., 2005). In 2007, the global
2004). At present, the main reduction technology is reduction total output of biomass straw was about 1.8 billion tons, which is
roasting using coal as reductant and fuel in the pyrolusite treatment equivalent to 0.6 million ton standard coal. It is expected that the
industry (Furlani et al., 2006). However, this traditional technology global total output of straw reaches 2.2 billion ton in 2010 (Zhang and
leads to pollution and greenhouse gas emissions (Acharya and Kar, Zhou, 2007). Therefore, making use of biomass straw to reduce
2003; Gao, 2006). Especially, the rate of increase of the greenhouse manganese oxide ore cannot only decrease the greenhouse gas
gas emissions from manganese oxide ore treatment industry has emissions of the metallurgical industries, but also widen the channels
become one of the fastest growing across all energy utilization sectors of energy to lower the cost, and comply with the requirement of green
(Zhang, 2007; Hariprasad et al., 2007). Therefore, manganese economy.
producers are under increased pressure to transform the existing In this paper, using the manganese oxide ore and biomass straw as
process or develop novel more environmentally sustainable opera- raw material, the thermal reduction process and kinetics of the
tions to meet the growing climate change challenges. reduction reactions were investigated. This initial fundamental study
assessed the potential role of biomass straw in the reduction of
⁎ Corresponding author. manganese oxide ore and provided the theoretical foundation for
E-mail address: zhugc@mail.tsinghua.edu.cn (G. Zhu). industrial scale process conditions.

0304-386X/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.hydromet.2010.08.004
Y. Zhao et al. / Hydrometallurgy 105 (2010) 96–102 97

2. Experimental (a)
6000
2.1. Materials
SiO2
5000 MnO2
Biomass straw was obtained from a local farm in Beijing Mn3O4
Changping, China. Prior to the experiments, it was processed by a

Intensity(Counts.)
4000 Mn2O3
pulverizer to be powder-like with a particle size of 80 μm. The result
Fe2O3
of the preliminary, elemental, and constituent analysis of the biomass
3000 Fe3O4
straw was given in Table 1.
The sample of manganese oxide ore supplied by Mengtai Mineral
Development Corporation (MMDC, China) contained MnO2 34.6%, 2000
Mn2O3 1.34%, Mn3O4 0.98%, Fe2O3 1.21%, Fe3O4 1.78%, Al2O3 2.0%, SiO2
39.2%, Zn 0.02%, Ni 0.01%, and Co 0.01%. The ore was crushed and ground 1000
to a particle size smaller than 0.147 mm (N100 mesh). An X-ray
diffraction analysis of the manganese oxide ore is shown in Fig. 1, and
0
the main phases of the manganese in ore were MnO2, Mn2O3, Mn3O4
and MnO. 10 20 30 40 50 60 70

2.2. Experimental procedure
(b)
Reduction roasting of the manganese oxide ore with biomass 6000
straw was carried out in a muffle electric furnace. The initial
proportion of biomass straw was 10 wt.% of manganese oxide ore. 5000 SiO2
The mixture of ore and straw was first put in a reactor covered with a MnO
Al2Si2O5(OH)4
lip before reduction roasting. After the reduction was completed, the 4000
roasted product was immediately transferred into a leaching vessel Intensity
containing required quantities of sulfuric acid to prevent the reduced 3000
ore from being reoxidized by the oxygen of air. The solution was
stirred continuously, and given amounts of sample was taken from the
2000
leaching vessel at predetermined time intervals. After the sample was
filtrated, manganese in filtrate was estimated volumetrically by EDTA
1000
titration using a thymolphthalexone indicator. The recovery of
manganese was then calculated from the result. The experiments
were repeated with new ore/biomass straw mixtures containing 20, 0
30, 40 and 50% more biomass whilst keeping the mass of ore constant. 10 20 30 40 50 60 70

2.3. Analysis method
Fig. 1. X-ray diffraction analysis of the original ore (a) and the roasted ore with 30% straw (b).

The thermal property measurements of the sample materials were


performed in the TG-DSC equipment from Setaram Instrumentation,
France. The equipment consists of a cylindrical graphite furnace, balance for each node was calculated based on the principle where the
graphite reaction tube, TG-DSC sensor (microbalance and DSC heat accumulated by the node equalled the difference in heat input
transducers), carrier and auxiliary gas inlet, furnace thermocouple and output across the node. Boundary conditions were the tempera-
for temperature control by PID controller, among others using the tures measured at the centre and surface of the sample, zero heat flux
computer-aided thermal analysis technique (Beck and Blackwell, in the centre of the sample, and calculated surface heat flux assuming
1985; Strezov et al., 2003a,b; Kanungo and Mishra, 2002). The sample radiative heat transfer from the cylindrical graphite furnace to the
with 6 mm in height and 5 mm in diameter was packed into the sample:
cylindrical graphite furnace. h i
4 4
Heat was supplied to the sample at a constant heating rate of Q = F1−2 σ ðTξÞ −T ð1Þ
10 °C/min from the graphite cylinder connected to a computer-aided
temperature controller. The maximum temperature was controlled where Q is the heat flux (W/m2), F1–2 the radiation shape factor
according to the result of the experiments in muffle electric furnace. (dimensionless), σ the Stefan Boltzmann constant, ξ the temperature
The specific heat of each sample was estimated by applying an inverse coefficient, and T the sample temperature (K).
numerical technique to the measured temperatures (Strezov et al., The radiation shape factor F1–2 is a function of emissivities of both the
2003a,b). For calculation purposes the sample was divided into a grid sample and cylindrical graphite furnace, as well as their surface areas
pattern with a number of nodes (n) across the radius. The heat and was determined through calibration (Strezov et al., 2003a,b). The
inverse numerical technique detailed by Beck and Blackwell, 1985 was
implemented to accurately resolve the one-dimensional heat conduc-
Table 1
Preliminary, elemental, and constituent analysis of the biomass straw.
tion equation:
 
Preliminary analysis Elemental analysis Heat output Biomass constituent ∂T ∂ ∂T
(%) (%, dry basis) (MJ/kg) (%) ρCp =k r ð2Þ
∂t ∂r ∂r
Moisture 8.0 Carbon 43.57 Cellulose 37.60
Volatiles 78.7 Hydrogen 3.92 15.84 Hemicellulose 21.60 where r is the density (kg/m3), Cp the specific heat (J/kgK), k the
Ash 6.90 Nitrogen 1.43 Lignin 18.40 thermal conductivity (W/mK), T the temperature (K), t the time (s)
Fixed carbon 14.40 Oxygen 44.14
and r the radius (m).
98 Y. Zhao et al. / Hydrometallurgy 105 (2010) 96–102

A computational matrix (Eq. (3)) was applied to determine the 100 200
specific heat of the sample material based on the initial density of the 250
heated sample: (Strezov et al., 2003a,b; Geankoplis, 1983; Beck and 300
80
Blackwell, 1985) 400

Recovery of Mn (%)
500
2πnΔxQðtÞ 60
ρCp = 600
     n−1   
Δx2 π = 4Δt T0t −T0t−1 + Δx2 π = Δt n− 14 Tnt −Tnt−1 + Σ 2πΔx2 i = Δt Tit −Tit−1
i=1

ð3Þ 40

where n is the number of nodes, Tit the temperature expressed in K of


20
the node i for the time t (s) and Q(t) the heat flux expressed in W/m2
for time t, and T(n + 1) = TnΔx / r.
The estimated specific heat had apparent values, meaning the heat 0
evolved during decomposition (ΔH) of the heated sample material was
included, 0 20 40 60 80 100
t (min)

Cp = Cp + ΔH = ΔT ð4Þ
Fig. 2. Effect of roasting temperature and time on recovery of Mn.

Therefore, during an endothermic heat effect, the specific heat


showed increasing values, while during an exothermic reaction the
specific heat values decreased.
resulted in a better overall outcome with a lower total energy cost
This method provides an opportunity for dynamic thermal analysis
per mass of Mn recovered, on the other hand, the domestic straw
by heating the sample with controlled heat flux while continuously
price of 30–50 dollar/ton is cheaper than the coal price of 70–
monitoring boundary temperature conditions within the sample.
100 dollar/ton. The combustion temperature of lignitic coal is about
There was a substantial effort in the past to apply it in a variety of
800 °C while the combustion temperature of straw is about 250 °C,
different thermal studies (Strezov et al., 2003a,b; Geankoplis, 1983). It
so only when temperature reaches 800 °C can carbon monoxide be
was shown that, with controlled heating conditions it is possible to
produced in the conventional process using coal. During the
obtain accurate thermal characterization of the materials.
combustion of straw, the product of CO, CO2 and coke would be
formed. The conversion to CO from CO2 is an endothermic process, so
3. Results and discussions
this conversion is at a higher temperature than the burning point of
straw. From Fig. 3, the recovery of Mn increases from 84.5% to 92.6%
3.1. Effect of weight ratio of manganese oxide ore to biomass straw
when the temperature rises from 250 °C to 600 °C, indicating that
the energy for the conversion to CO from CO2 is supplied from
Initial experiments were carried out by varying the weight ratio of
electrical sources. However, the coal as reductant is burned at least
manganese oxide ore to straw. The amount of biomass straw was
800 °C, and then the reaction of coke and CO2 still needs extra energy
varied from 1 g to 5 g while keeping the amount of manganese oxide
from electrical sources. The total energy is saved using straw as
ore as 10 g, namely, varying the weight ratio of biomass straw to
reductant than the coal.
manganese oxide ore from 1:10 to 5:10. The effect of weight ratio on
leaching recovery of manganese was shown in Table 2.
The recovery rate of Mn increased rapidly with a decrease in the
3.3. Effect of ore particle size
weight ratio, but it became slow with further decrease to 3:10 at
which above 90% recovery of Mn was obtained. This indicated that
The effect of ore particle size was studied by varying size fractions
over 90% of the MnO2 in the ore were reduced in the roasting
from a geometric average particle diameter of 50 μm to 300 μm while
process. Therefore, a weight ratio of 3:10 was used for subsequent
keeping other conditions as: the weight ratio of 10:3 (manganese
tests.
oxide ore /straw), and the roasting temperature at 500 °C for 80 min.
3.2. Effect of roasting temperature and time
120
Reduction of the manganese oxide ore strongly depends on Biomassstraw
roasting temperature, and reduction efficiency can be indicated by 100 ligniticcoal
manganese recovery rate. The effects of roasting temperature and
Recovery of Mn /%

time on the recovery rate of manganese were studied using the


80
weight ratio of 10:3 (manganese oxide ores/straw) and average ore
size of 150 μm. The results obtained were shown in Fig. 2.
60
It was observed that manganese dioxide was virtually not
reduced while the roasting temperature was below 250 °C. By
increasing the temperature from 250 °C to 500 °C, the recovery rate 40
Conditions:
of manganese after 80 min increased from 50.3% to 90.2%. The 10:3 weight ratio
results on using lignitic coal as the conventional reductant under the 20 (Ore/Straw)
same conditions are presented in Fig. 3. It was obvious that straw 1.5h roasting rime
0 150μm ore size
Table 2
Effect of the weight ratio of the ore to biomass straw on recovery rate of Mn. 0 200 400 600 800 1000 1200
Roasting temperature /
Ore: biomass straw (g/g) 10: 1 10: 2 10: 3 10: 4 10: 5
Recovery rate of Mn (%) 33.7 64.5 90.6 92.7 92.4
Fig. 3. Comparison of recovery between straw and lignitic coal.
Y. Zhao et al. / Hydrometallurgy 105 (2010) 96–102 99

100
Recovery of Mn /%

90

80
Conditions:
10:3 weight ratio (Ore/Straw)
500 roasting temperature
70
1.5h roasting rime

60
0 50 100 150 200 250 300 350
Average particle diameter /μm

Fig. 4. Effect of ore particle size on recovery of Mn. Fig. 6. Relationship of specific heat with temperature (ore-biomass straw mixture at
10:1 mass ratio).

For plotting the results (Fig. 4) geometric average particle diameter of


corresponding size fractions was calculated as: calculation of the endothermic reaction of moisture evaporation. This
reaction for the biomass material was estimated to be around 15.8 MJ/
1=2 kg and was applied to back-calculate the amount of moisture absorbed
dp = ðf 1 × f 2 Þ
by the sample and standard latent heat of water evaporation. For
temperatures above the dehydration temperature range, the thermal
Where, dp was the average particle diameter, f1and f2 were the analysis data showed that biomass decomposition commenced at
upper and lower particle sizes of the fraction. around 210 °C with an endothermic reaction, shifting to an exothermic
Fig. 4 showed that with the increase in average diameter from reaction when the temperatures reached 300 °C. From this temperature
75 μm to 275 μm, the recovery rate of manganese decreased from range, up until approximately 350 °C the thermal decomposition was
95.9% to 76.5%. The higher recovery for smaller particle size was due largely exothermic. The thermal region of biomass decomposition is
to an increase in surface area for the reaction with CO. believed to depend on its major constituents, namely cellulose,
hemicellulose and lignin (Strezov et al., 2003a,b; Zhang et al., 2005).
3.4. Thermal analysis For the manganese ore sample, the first endothermic peak also showed
at approximately 100 °C, which was the starting of ore dehydration
Initial TG-DSC experiments were conducted under the non- process. The thermal decomposition of the manganese oxide ore for
isothermal heating rate of 10 °C/min up to 650 °C at a mass ratio 10:1 temperatures above dehydration range, exhibited endothermic reac-
(manganese oxide ore/biomass straw). The result is shown in Fig. 5. tions with peaks at 320, 400 °C and an exothermic trough at 460 °C.
Through Eq. (3), the apparent specific heat during roasting of When manganese oxide ore and biomass straw were mixed at a
manganese ore samples with biomass straw is detailed in Fig. 6. mass ratio of 10:1, the specific heats and consequent heats of
The data revealed the temperature ranges of endothermic heat reactions showed that for temperatures of up to 380 °C the reactions
consumption and exothermic heat release during reaction of these are additive and depend on the thermal behaviour of the individual
samples. For the biomass straw sample the first endothermic peak at material. For temperatures above this range the reactions were
approximately 100 °C identified the presence of moisture. From the
specific heat data, the initial and final temperatures of biomass
dehydration can be detected which can provide a baseline for the

Fig. 5. Thermal analysis of ore-biomass straw mixture at 10:1 mass ratio. Fig. 7. Thermal analysis of ore-biomass straw mixture at 10:3 mass ratio.
100 Y. Zhao et al. / Hydrometallurgy 105 (2010) 96–102

Depending on the excess carbon available in the mixture, the


reduction follows the above stepwise mechanism during heating.
Complete reduction can be achieved only in cases when sufficient
carbon is available in the mixture to maintain the reduction process.
The rate of change in heat of reactions for the 10:3 mixture with
respect to temperature (△H/△T) was showed by subtracting the
specific heat data from a straight line baseline (Fig. 9).
The individual integrals of this curve represent the corresponding
heats of reactions for individual temperature ranges. For temperatures
of up to approximately 390 °C the corresponding reactions were
affected by the intrinsic properties of the biomass straw and manganese
oxide ore material. It appears that there was very little, if any, synergetic
effect between the biomass straw and ore materials at this temperature
range. Reactions associated with decomposition of materials can
generally be predicted and, in this case, primarily depended on the
properties of ore samples. For instance, manganese ore may contain
various amounts of MnO2, kaolinite, or other minerals, which affect the
Fig. 8. Relationship of specific heat with temperature (ore-biomass straw mixture at overall latent heat. These values can be easily determined by the
10:3 mass ratio). mineralogical characterization of the manganese ore. The reduction,
however, is temperature and process dependent and, with careful
associated with manganese oxide ore reduction and they were largely determination of energy requirements and kinetics of the reduction
endothermic but did not reach equilibrium (Fig. 6). reactions, predictable methodologies can be easily applied to model the
For the mass ratio (manganese oxide ore/biomass straw) of 10:3, process variables of possible industrial applications.
the result of TG-DSC experiment and apparent specific heat using a The reactions associated with manganese ore reduction in this case
mass ratio of 10:3 is shown in Figs. 7 and 8, respectively. were detected at temperatures above 390 °C. The rates of the heats of
The reactions associated with manganese oxide ore reduction reactions, however, were largely affected by the stepwise reduction
commenced somewhere around 330 °C with a small exothermic mechanism following biomass carbon gasification. From Fig. 9, three
reaction and continued with two clear endothermic reactions, peaking distinct temperature regions could be divided, which were 300–390 °C,
at 400 °C and 480 °C. The origin of these reactions is the stepwise 400–480 °C, and 490–640 °C. The energy change in each temperature
reduction of manganese oxide ore with reductant CO gas produced region was calculated according to its corresponding integral. It is rather
through the biomass gasification reaction around 350° to produce CO: difficult to associate each temperature region with the individual
reaction listed in Eqs. (7)–(9), as it is expected that the reduction is
C + O2 →CO2 ð5Þ cross-linked and interrelated. For this reason it was more applicable to
appoint overall temperature regions to the corresponding reactions.
CO2 + C→CO ð6Þ
3.5. Kinetic modeling
This is followed by the triggering of the reduction process in a
series of interrelated reactions between 400 and 600° (Ren, 1993; Knowledge of the kinetic behaviour is essential for understanding
Momade and Momade, 1999): and predicting the thermal conversion processes and, in this case, to
be able to model energy requirements in industrial process conditions
MnO2 + CO→Mn2 O3 + CO2 ð7Þ (Beolchini et al., 2001). For this purpose, the measured heats of
reduction reactions were used to determine the activation energies
Mn2 O3 + CO→Mn3 O4 + CO2 ð8Þ associated with the thermal regions of reduction. For any chemical
reaction, enthalpy of system is determined as below:
Mn3 O4 + CO→MnO + CO2 ð9Þ 
H = ∑ n0;B + νB ξ HB ð10Þ
B

Where n0, B is amount of substance B before the reaction started, νB


the stoichiometric coefficient of substance B, and HB the enthalpy of
substance B.
Because the calculation of heat capacity is


∂H
Cp = ð11Þ
∂T p

According to Eqs. (10) and (11), we know that




∂ξ
Cp = ∑ n0;B + νB ξ Cp;B + ΔrHm ð12Þ
B ∂T p

Where Cp is heat
capacity,
ξ extent of reaction, ΔrHm molar
∂ξ
enthalpy change, and change rate of extent of reaction.
∂T p
For chemical reaction in a closed system,
Fig. 9. Rate of change of heat of reactions with respect to temperature (ore-biomass
straw mixture at 10:3 mass ratio).
dG = −SdT + Vdp + ∑BνBμBdξ ð13Þ
Y. Zhao et al. / Hydrometallurgy 105 (2010) 96–102 101

and exponential factor was calculated from the intercept of the plot y0,

which equals:
∂G
∑BνBμB = = −A ð14Þ
∂ξ T;p e y0 βE2
A= ð20Þ
ER−2TR2
The heat capacity of the different temperatures can be calculated
dΔrHmΘ The calculated activation energies and pre-exponential factors for
by Kirchhoff's formula dT
= ΔrCp;m . The data of ΔrHΘ m(T) is the three temperature regions of manganese oxide ore reduction with
Θ Θ
deduced by the known ΔrHm(298K), and K (T) is determined using biomass straw are detailed in Table 3. The origin of these reactions is
Θ the stepwise reduction of manganese oxide ore with reductant CO gas
d lnK Θ ΔrHm
Van't Hoff formula, = , so the heat capacity can be produced through the biomass gasification reaction, and it is expected
dT RT 2

calculated basing on Eq. (12). When the ratio of ore and biomass is that the reduction reaction is cross-linked and interrelated.
10/3, the heat capacity is estimated about 682 kJ in the temperature of
310–360 °C. 3.6. Reaction mechanism
The heats of reaction were integrated with respect to time and
plotted cumulatively as a degree of reaction. The degree of reaction After the biomass straw was decomposed to produce CO gas
was set to 0 before the reaction occurred and 1 when it was through the biomass gasification reaction (Eqs. (5) and (6)), the
completed. The kinetics of each reaction was expressed as a function manganese oxides in the original ore including MnO2, Mn2O3 and
of Arrhenius parameters with reaction model presented as: Mn3O4 are reduced to MnO. From Figs. 2 and 6, it can be seen that at
temperatures below 200 °C, the dehydration process of ore and
dα = dt = kðT Þf ðαÞ ð15Þ biomass is predominant. The reduction of manganese oxides in ore
happen in carbon monoxide atmosphere because of the thermal
where t is the time (s), T the temperature (K), α the extent of decomposition of biomass when temperature was controlled over
conversion (dimensionless), f(α) the reaction model and k(T) the 320 °C. The reduction follows the above-mentioned (Eqs. (7)–(9))
temperature dependent rate constant, which is expressed by the stepwise mechanism during heating. However Eq. (7) constitutes the
Arrhenius equation as: main initial reduction process because the content of the other oxides
in ore (Mn2O3, Mn3O4, Fe2O3 and Fe3O4) reduced by CO is very low.
kðT Þ = A expðE = RT Þ ð16Þ
Complete reduction can be achieved only in cases when sufficient
carbon is available in the mixture to maintain the reduction process.
where A is the pre-exponential or frequency factor (s−1), E the
From Table 2, the recovery rate of Mn is only 33.7% with the biomass
activation energy (J/mol) and R the gas constant (8.314 J/mol K).
weight ratio of 10% because the CO reductant produced is not enough.
The activation energy has been defined as the energy barrier that
Therefore the dosage of biomass and roasting temperature are the key
must be surmounted to enable occurrence of the bond redistribution
factors for the reduction of manganese ore.
steps required to convert reactants into products (Galwey and Brown,
2002). The pre-exponential factor describes the frequency of
4. Technological importance of the use of biomass
occurrence of the reaction situation. Calculating E and A requires
integrating Eq. (15) and, for the first-order of reaction, the equation is
This initial study assessed the potential of biomass straw in reduction
expressed as
of manganese oxide ore. The main significance of this technology using
  biomass in metal smelting was to reduce pollution and CO2 emissions
αdα A T E
∫ = ∫ exp − dT ð17Þ from the metallurgical operations, while at the same time promoting
0 ð1−αÞ β 0 RT
  effective biomass straw utilization to widen the channels of energy for
dT lowering the cost. The results from this work showed that 30% by weight
β stands for heating rate . The right-hand side of Eq. (17) has
dt of biomass straw was sufficient to achieve reduction of manganese
no exact integral. An evaluated simplified expression developed by oxide ore to MnO. It should be highlighted that the results are limited for
Coats and Redfern (1964) was adopted: the selected manganese oxide ore and biomass straw samples only. As
    manganese oxide ore varies considerably depending on the mineral
ART 2 2RT E
− lnð1−αÞ = 1− exp − ð18Þ forms and manganese oxide phases present, the reduction process may
βE E RT differ depending on the manganese oxide ore type and its mineral
compounds. Secondly, biomass straw may contain various impurities
Taking the logarithm of Eq. (18) yields. originating from previous treatments, such as paints or other chemical
solutions, which can react with the manganese oxide compounds.


 
− lnð1−αÞ AR 2RT E Therefore, biomass straw, depending on its origin, may show different
ln = ln 1− − ð19Þ behaviors when applied for processing of minerals and should be
T 2 βE E RT
assessed independently. In this experiment, the selected biomass straw
did not contain any impurities.
By plotting the left-hand side of Eq. (19) against 1/T for each of the
reduction reactions, resulted in approximately straight lines with a 5. Conclusions
slope of −E/R, so activation energy E can be calculated. The pre-
Reduction roasting of manganese oxide ore can be carried out
using biomass straw roasting at the temperature of below 600 °C. At
Table 3
Kinetic parameters of detected temperature ranges of manganese ore reduction with 30% of biomass added, the manganese oxide ore was reduced to
biomass straw at mass ratio 10:3 (ore/biomass straw). predominantly MnO, demonstrating biomass straw as good reductant.
Thermal analysis showed that for temperatures of up to 300 °C the
Temperature range 300–390 °C 400–480 °C 490–640 °C
reactions of ore and 30% biomass depend on the properties and
A (s−1) 3.4E + 20 4.3E + 17 3.6E + 13 decomposition of the initial biomass straw and ore samples, while
E (kJ/mol) 470 510 430
reduction reaction commenced at approximately 390 °C and was
102 Y. Zhao et al. / Hydrometallurgy 105 (2010) 96–102

almost completed at 600 °C. The reduction process was divided into Gao, H.L., 2006. Production and consumption of Mn at home and abroad. China Metal
Bulletin 7, 33–36.
three major temperature events with the first being exothermic while Geankoplis, C.J., 1983. Transport Process and Unit Operations. Allyn and Bacon, Boston.
two higher temperature regions exhibited endothermic heat effects. Hariprasad, D., Dash, B., Ghosh, M.K., Anand, S., 2007. Leaching of manganese ores using
The reaction degree of the three examined thermal regions followed sawdust as a reductant. Minerals Engineering 20, 1293–1295.
Harris, M., 1997. The history production and uses of electrolytic manganese. Metallurgic
the first-order kinetic mechanism. The activation energy (E) was 17 (14), 121–124.
calculated to be 470 kJ/mol at 300–390 °C, 510 kJ/mol at 400–480 °C, Ismail, A.A., Ali, E.A., Ibrahim, I.A., Ahmed, M.S., 2004. A comparative study on acid
and 430 kJ/mol at 490–640 °C respectively. leaching of low grade manganese ore using some industrial wastes as reductants.
Canadian Journal of Chemical Engineering 82, 1296–1300.
Kanungo, S.B., Mishra, S.K., 2002. Reduction of phosphorus content of certain high
Acknowledgement phosphorus manganese ores of India by roasting with sodium chloride followed by
leaching in acid medium I: statistical design of roasting experiments. Transactions
of the Indian Institute of Metals 55 (3), 81–89.
The authors are grateful to the National Natural Science Founda- Momade, F.W.Y., Momade, Zs.G., 1999. A study of the kinetics of reductive leaching of
tion (grant no. 50874067) for financial support. manganese oxide ore in aqueous methanol–sulphuric acid medium. Hydrometal-
lurgy 54, 25–39.
Ren, S.J., 1993. Industrial Minerals Resource Exploitation and Processing Handbook.
References Wuhan Industry University Press, Wuhan, pp. 18–37.
Sahoo, P.K., Rao, K.S., 1989. Sulphating-roasting of low grade manganese ore:
Abbruzzese, C., Duarte, M.Y., Paponetti, B., Toro, L., 1990. Biological and chemical optimisation by factorial design. International Journal of Mineral Processing 25
processing of low-grade manganese ore. Minerals Engineering 3 (3–4), 307–318. (1–2), 147–152.
Acharya, C., Kar, R.N., 2003. Studies on reaction mechanism of bioleaching of Sahoo, R.N., Naik, P.K., Das, S.C., 2001. Leaching of manganese ore using oxalic acid as
manganese ore. Minerals Engineering 16 (10), 1027–1030. reductant in sulphuric acid solution. Hydrometallurgy 62, 157–163.
Beck, J.V., Blackwell, B., 1985. Inverse Heat Conduction-Ill Posed Problems. Wiley, New Strezov, V., Lucas, J.A., Strezov, L., 2003a. Computer aided thermal analysis. Journal of
York. Thermal Analysis and Calorimetry 72, 907–918.
Beolchini, F., Petrangelipapini, L., Toro, I., Trifoni, M., 2001. Acid leaching of Strezov, V., Moghtaderi, B., Lucas, J.A., 2003b. Thermal study of decomposition of
manganiferous ores by sucrose: kinetic modeling and related statistical analysis. selected biomass samples. Journal of Thermal Analysis and Calorimetry 72,
Minerals Engineering 14 (2), 175–184. 1041–1048.
Coats, A.W., Redfern, J.P., 1964. Kinetic parameters from thermogravimetric data. Tian, X.K., Wen, X.X., Yang, C., Liang, Y.J., Pi, Z.B., Wang, Y.X., 2010. Reductive leaching of
Nature 201, 68–69. manganese from low-grade manganese dioxide ores using corncob as reductant in
Das, S.C., Sahoo, P.K., Rao, P.K., 1982. Extraction of manganese from low grade sulfuric acid solution. Hydrometallurgy 100, 157–160.
manganese ores by FeSO4 leaching. Hydrometallurgy 8, 35–47. Zhang, J.S., 2007. Current challenge and chance in China's Mn-industry. China
Furlani, G., Pagnanelli, F., Toro, L., 2006. Reductive acid leaching of manganese dioxide Manganese Industry 1, 6–9.
with glucose: identification of oxidation derivatives of glucose. Hydrometallurgy Zhang, J.T., Zhou, C.Y., 2007. The energy utilizing status and prospects of the crop straw
81, 234–240. resources. Liquor Making 34 (4), 12–15.
Galwey, A.K., Brown, M.E., 2002. Application of the Arrhenius equation to solid state Zhang, M., Yuan, Y.C., Liu, Y.Z., 2005. Research on biomass waste combustion
kinetics. Thermochimica Acta 386, 91–98. technologies. Energy Research and Information 21 (1), 15–19.

You might also like