You are on page 1of 6

Radiation Physics and Chemistry 86 (2013) 90–95

Contents lists available at SciVerse ScienceDirect

Radiation Physics and Chemistry


journal homepage: www.elsevier.com/locate/radphyschem

g-Irradiation effects on the non-isothermal decomposition of strontium


nitrate by model-free and model-fitting methods
Sunil Culas, Jadu Samuel n
Department of Chemistry, Research Centre, Mar Ivanios College, Thiruvananthapuram 695015, Kerala, India

H I G H L I G H T S

c The thermal decomposition of Sr(NO3)2 occurs in a single step reaction.


c Irradiation enhances the thermal decomposition.
c Activation energy decreases on irradiation.
c The Master plot method reveals that D3 model best describes the kinetic process.

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

Article history: The kinetics of the thermal decomposition of untreated and g-irradiated strontium nitrate, Sr(NO3)2
Received 13 July 2012 was studied under non-isothermal conditions at different heating rates (5, 10, 15 and 20 1C min  1) in
Accepted 28 January 2013 nitrogen atmosphere. The data were analysed using both isoconversional and non-isoconversional
Available online 14 February 2013
methods. The activation energies were calculated by Flynn–Wall–Ozawa (FWO), Kissinger–Akahira–
Keywords: Sunose (KAS) and Friedman (FR) methods. The results show that the irradiation enhances the
Strontium nitrate decomposition and the effect increases with the irradiation dose. The activation energy decreases on
Kinetics and mechanism irradiation. The appropriate conversion model for the thermal decomposition process selected by
Effect of g-irradiation means of the master-plot method agrees with three-dimensional diffusion model (D3 mechanism),
Non-isothermal decomposition
g(a)¼ [1  (1  a)1/3]2 for both untreated and irradiated salts at all heating rates.
Isoconversional method
& 2013 Elsevier Ltd. All rights reserved.

1. Introduction smallest for cations exerting the largest field strength. The
sensitivities of potassium and sodium nitrates to decomposition
On irradiation with ionising radiations (Cunningham and Heal, by the radiations from a nuclear reactor (Hennig et al., 1953) were
1958), solid anhydrous nitrates decompose, producing nitrite and in the same order as for ultraviolet light. The higher yield
oxygen. For the barium nitrate, it has been shown that the observed for the potassium salt was attributed (Hennig et al.,
irradiated salt upon solution in water yields a gas presumed to 1953) to the greater space available in the crystal. The stability of
be mainly oxygen and an amount of nitrite ion equivalent to the radiolytic product varies widely with crystal matrix (Golding and
evolved gas (Allen and Ghormley, 1947). These authors suggested Henchman, 1964). Many recent studies on the thermal decom-
that the oxygen must remain trapped in the crystal during position of inorganic solids have included measurements on
irradiation as O atoms or O2 molecules. The existence of other samples that were exposed radiation prior to heating with the
species such as NO23  , NO22  , etc. has also been reported during aim to investigate the effect of ionising radiation on the kinetics
radiolytic decomposition of alkali metal nitrates (Muhammad and and thermal decomposition behaviour of inorganic compounds
Maddock, 1978). Different nitrates vary widely in their sensitivity (Mahfouz et al., 2000; Brown et al., 1980). Measurements of
toward decomposition by radiation (Narayanswamy, 1935). Crys- irradiation effects on the chemical composition, structural,
talline inorganic nitrates decompose (Doigan and Davis, 1952) on optical, electrical and thermal behaviour of various materials give
exposure to ultraviolet light with quantum yields which depend valuable information on the physico-chemical consequences of
on nature of the cation. The yields were related to the field radiation damage in solids (Nair et al., 1983; Diefallah et al., 1985;
strength exerted by the cation on the nitrate ion, and were Nair and Jacob, 1989; James and Samuel, 2005).
The methods proposed for the kinetic study of thermal decom-
position of solids are commonly classified as model-fitting and
n model-free methods. Historically, model-fitting methods are widely
Corresponding author. Tel.: þ91 471 2531053; fax: þ91 471 2530023.
E-mail addresses: prof_jadu@hotmail.com, used because of their ability to directly determine the kinetic triplet
jadu_samuel@yahoo.co.in (J. Samuel). involving activation energy (E), frequency factor (A) and conversion

0969-806X/$ - see front matter & 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.radphyschem.2013.01.042
S. Culas, J. Samuel / Radiation Physics and Chemistry 86 (2013) 90–95 91

function, g(a). It is well established that force-fitting non-isothermal Table 1


60
data to different reaction models results in a widely varying Concentration of radiolytic products generated in Sr(NO3)2 by Co g-rays.
Arrhenius parameters (Vyazovkin, 1997). The kinetic analysis based
Irradiation dose (MGy) 0.5 MGy 1.5 MGy 2.5 MGy
on isoconversional method is frequently referred to as ‘‘model-free’’
because it is possible to obtain the apparent activation energy (E) as NO2 (ppm) 1211 1892 2349
the function of conversion degree (a) and these methods are
considered as the mostly reliable ones. The present paper reports
studies on the thermal decomposition of g-irradiated Sr(NO3)2 by
using non-isothermal multi-heating thermogravimetric (TG) data in 110
terms of model-fitting and model-free isoconversional methods as 3.5
recommended by the International Confederation for Thermal 100
3.0

Derivative mass (%/°C)


Analysis and Calorimetry (ICTAC) Kinetics Committee (Vyazovkin
et al., 2011). This has been carried out with a view to investigate the 90 2.5
effect of g-irradiation on the thermal decomposition and to evaluate

Mass (%)
80 2.0
the kinetic parameters and mechanism of decomposition of both
untreated and irradiated samples. 1.5
70
1.0
60
2. Experimental 0.5
50 0.0
2.1. Irradiation
40 -0.5
AR grade Sr(NO3)2 used in the present investigation, is a colour- 100 200 300 400 500 600 700 800
less crystalline solid. Portions of the crystals were encapsulated Temperature (°C)
under vacuum in glass vials and were exposed to successively
increasing doses of radiation at constant intensity using 60Co g-rays Fig. 1. TG–DTG curves of untreated Sr(NO3)2 at a heating rate of 5 1C min  1.
to different doses between 0.5 and 2.5 MGy at a dose rate of
60
2 kGy h  1. The irradiated samples were also preserved over P2O5 nitrite ions generated in Sr(NO3)2 by various doses of Co g-rays
before thermal decomposition studies. are given in Table 1.

2.2. Estimation of damage 3.2. Thermogravimetric curves

The nitrite was estimated spectrophotometrically by the TG–DTG curves of unirradiated Sr(NO3)2 at heating rate
method developed by Shinn (1941) and modified by Kershaw 5 1C min  1 under nitrogen atmosphere are shown in Fig. 1. TG–
and Chamberlin (1942). The method was shown to be sensitive to DTG curves of other samples of both untreated and irradiated
3 ppm of nitrite. The colour was found permanent up to 70 min Sr(NO3)2 at different heating rates are essentially of the same
after development. pattern and these curves are not included in the present paper.
In the routine measurements, 0.2–0.3 g of the irradiated Duplicate non-isothermal runs were taken for each sample and
nitrate crystals were dissolved in boiled out distilled water to mass-loss temperature relationship was found to be in good
give 100 mL of solution. A suitable aliquot of this solution was agreement in all the runs. The main observations like temperature
transferred to a 50 mL standard flask and 2 mL of a 0.5% solution of inception (Ti), the temperature of completion (Tf) and the peak
of sulphanilamide in 1:1 hydrochloric acid added. After 3 min, temperature of decomposition (Tp) obtained from the thermal
1 mL of a 0.1% solution of N-(1-naphthyl) ethylenediamine curves, for all the samples are summarised in Table 2.
dihydrochloride was added. After mixing the solution was made In order to apply different kinetic methods on the thermal
up to 50 mL and allowed to stand 15 min before measurement of decomposition process of Sr(NO3)2, the dependence of fraction
the absorbance at 540 nm. decomposed (a) versus temperature (T) at different heating rates
for all the samples are plotted (Fig. 2). The sigmoid-shaped curves
are shifted to higher temperatures with an increase of heating
2.3. Thermogravimetric studies rates, but they are shifted to lower temperatures as the irradiation
dose increases.
The thermogravimetric (TG) measurements of Sr(NO3)2 samples
were carried out by a Mettler Toledo TGA/SDTA 851e module
3.3. Estimation of activation energy
apparatus in the temperature range 30–800 1C, under dynamic
atmosphere of nitrogen (30 mL/min). In all experiments 10–12 mg
Initially, the values of activation energy for the decomposition
of the sample was put into platinum crucibles, at four different
of all the samples of Sr(NO3)2 were estimated by using FWO
heating rates 5, 10, 15 and 20 1C min  1. The recorded total % of mass
(Flynn and Wall, 1966; Ozawa, 1965), KAS (Kissinger, 1957;
loss in all cases was 51.3370.05 mg, confirming the complete
Akahira and Sunnose, 1971) and FR (Friedman, 1964) methods.
conversion of Sr(NO3)2 to SrO.
The details of these methods have already been reported
(Vyazovkin et al., 2011; Dogan et al., 2009; Tita et al., 2010).
FWO method is an integral method which is based on the
3. Results and discussion
measurement of the adequate temperature to certain values of
the conversion a, for experiments effectuated to different rates of
3.1. Chemical damage
heating. The equation corresponding to this method is
 
Irradiation produced a yellow coloration of the crystals. The AE E
ln b ¼ ln 5:3311:052 ð1Þ
concentrations of the radiolytic products expressed as damage RgðaÞ RT
92 S. Culas, J. Samuel / Radiation Physics and Chemistry 86 (2013) 90–95

Table 2
Phenomenological data for the thermal decomposition of all samples of Sr(NO3)2 1.0
at different heating rates.

0.8
Sample Sr(NO3)2 b (K min  1) Temperature (K)

0.6

α
Ti Tf Tp

Untreated 5 792.2 914.8 907.3 0.4


10 803.6 927.8 918.3
15 811.0 934.1 925.2 0.2
20 819.6 942.8 920.1

Irrad. 0.5 MGy 5 776.4 899.0 887.2 1.0


10 788.7 9129 907.2
15 796.8 919.9 909.8
0.8
20 804.3 927.5 921.2

0.6

α
Irrad. 1.5 MGy 5 767.4 890.0 880.1
10 781.8 906.0 898.5
0.4
15 789.9 913.1 904.5
20 794.9 918.1 908.7
0.2

Irrad. 2.5 MGy 5 760.0 882.6 841.5


10 774.2 898.4 885.3
15 783.1 906.2 891.5 1.0
20 787.8 911.0 900.2
0.8

0.6

α
where g(a) is integral reaction model, a is extent of conversion, E
is activation energy, A is pre-exponential factor, T is Kelvin 0.4
temperature, b is heating rate and R is gas constant.
The KAS method is another integral isoconversional method 0.2
and is based on the equation:
   
b AR E
ln 2 ln  ð2Þ 1.0
T EgðaÞ RT
5 C min
The derivative method used in this paper is Friedman’s 0.8 10 C min
method, which is probably the most general of the derivative 15 C min
0.6
α

techniques. The equation corresponding to this method is: 20 C min


   
da da   E 0.4
ln ¼ ln b ¼ ln Af ðaÞ  ð3Þ
dt dT RT
0.2
For selected a values from 20% to 90%, straight lines are obtained
by plotting ln b versus 1000/T (Eq. (1)), ln b/T2 versus 1000/T 810 840 870 900 930 960
(Eq. (2)), ln(bda/dT) versus 1000/T (Eq. (3)) for all the samples of Temperature (K)
Sr(NO3)2. From the slope of the straight lines thus obtained, the
Fig. 2. Fractional decomposition (a) versus temperature (T) curves for the non-
values of activation energy, E were calculated and are presented in
isothermal decomposition of untreated and g-irradiated Sr(NO3)2 at different
Table 3. It is evident from this Table that the values of E varies heating rates: (a) unirradiated; (b) 0.5 MGy; (c) 1.5 MGy and (d) 2.5 MGy.
slightly with increase in a. This can also be seen in Fig. 3 which plots
the variation of E with a. The weak variation in the effective
activation energy, E with increase in degree of conversion a slowly but becomes explosive above 600 1C. This disintegration
indicates a single reaction mechanism for the decomposition process appears to be completed over 820 1C. The succeeding horizontal is
(Vyazovkin, 2000). It is also worth to note that there is good due to the oxide SrO. Nair and James (1984) studied the thermal
agreement in the values of activation energy obtained by the KAS decomposition of Sr(NO3)2 under non-isothermal conditions with a
and FWO methods. The values of activation energy estimated by the single heating rate plot (4 1C min  1) and analysed the data by the
Friedman method are higher than the values obtained from the model fitting method. They reported that the decomposition of
other two methods. Because Friedman’s method is the derivative Sr(NO3)2 sets in after melting and the overall decomposition of the
technique based on an intercomparison of the rates of weight loss, salt occurs as.
while FWO and KAS methods are integral methods using the
SrðNO3 Þ2 -SrOþ 2NO2 þ 0:5O2 ð4Þ
approximation of temperature integral. There is also a systematic
error which arises from the numerical differentiation of the experi- They calculated the activation energy for the decomposition of
mental data involved in Friedman method (Srivastava et al., 2010). untreated Sr(NO3)2 using the method of Coats and Redfern (1964)
Since the values of activation energy worked out by the KAS and on the assumption that the reaction is of first order and reported a
FWO method are very close, we chose the average values of two as value of E¼424.1 kJ mol  1. In the present work, the estimated
the value of activation energy to be used in the master-plot method. values of activation energy presented in Table 3 are less than the
It has been reported that Sr(NO3)2 is stable up to 280 1C and reported value of activation energy (Nair and James, 1984). This may
yields a perfectly horizontal level which starts at room tempera- be due to the difference in methods employed for computing the
ture (Duval, 1963). Decomposition then sets in and proceeds very activation energy values. The values of E obtained in the present
S. Culas, J. Samuel / Radiation Physics and Chemistry 86 (2013) 90–95 93

400
380

E (kJ mol-1) 380 360

E (kJ mol-1)
360 340

340 320

FWO FWO
320 KAS 300 KAS
FR FR

300 280
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
α α

360
360

340
340
E (kJ mol-1)

E (kJ mol-1)
320
320
300
300
FWO
FWO 280 KAS
KAS FR
280 FR
260
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
α α

Fig. 3. Dependence of activation energy (E) on the degree of conversion (a) for all samples of Sr(NO3)2 calculated by isoconversional methods: (a) unirradiated;
(b) 0.5 MGy; (c) 1.5 MGy and (d) 2.5 MGy.

Table 3
Values of activation energy (E) in kJ mol  1 obtained by Flynn–Wall–Ozawa (FWO), Kissinger–Akahira–Sunose (KAS) and Friedman (FR) methods for all samples of
Sr(NO3)2.

Model-free methods Sample Fraction decomposed, a

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Mean

FR Untreated 359.1 358.0 358.3 365.1 363.2 361.7 370.2 365.3 362.6
0.5 MGy 335.6 337.6 336.6 344.8 340.8 339.2 348.9 348.8 341.5
1.5 MGy 320.8 322.4 324.4 327.8 327.9 330.1 336.9 334.6 328.1
2.5 MGy 310.3 315.9 315.3 320.6 318.3 319.1 327.9 321.7 318.6

FWO Untreated 335.1 335.8 337.1 343.6 342.6 341.8 349.6 345.6 341.4
0.5 MGy 313.2 316.6 316.7 324.5 321.5 320.6 329.5 329.7 321.5
1.5 MGy 299.6 302.7 305.3 309.0 309.6 311.8 318.3 316.5 309.1
2.5 MGy 289.8 296.4 296.6 301.9 30.4 301.6 309.6 304.5 300.1

KAS Untreated 338.2 338.5 339.7 346.4 345.1 344.2 352.3 348.0 344.1
0.5 MGy 315.4 318.6 318.5 326.6 323.3 322.2 331.4 331.5 323.4
1.5 MGy 301.3 304.1 306.7 310.4 310.8 313.2 319.8 317.8 310.5
2.5 MGy 291.1 297.6 297.6 303.1 301.3 302.4 310.8 305.4 301.2

study would be more reliable as they have been estimated according (Table 3) as has been observed in the case of potassium bromate
to the recommendations of ICTAC Kinetics Committee (Vyazovkin (Samuel et al., 2013).
et al., 2011). Several authors have emphasised the effects of various pre-
treatments such as doping, preheating, particle sizing, etc. on the
3.4. Role of irradiation thermal behaviour of inorganic oxysalts (Abdul Mujeeb et al., 2011;
Das and Patnaik, 2000; Nair and James, 1984). The role of irradiation
From the decomposition data presented in Table 2, it is evident in solid decomposition has been studied extensively (Young, 1966;
that the irradiation lowers the temperature of inception (Ti), James and Samuel, 2003). g-irradiation produces excitation and
temperature of completion (Tf) and peak temperature of decom- ionisation of the anion NO3 and this leads to chemical damage in
position (Tp). The activation energy also decreases on irradiation the crystal matrix. The primary radiolytic fragments produced in the
94 S. Culas, J. Samuel / Radiation Physics and Chemistry 86 (2013) 90–95

Table 4
Expressions for f (a) and g (a) for the most frequently used reaction models to describe the reaction kinetics in heterogeneous solid state systems.

Mechanism Symbols f (a) g (a)

Contracting linear R1 1 a
Contracting sphere R2 2(1  a)1/2 1  (1 a)1/2
Contracting cylinder R3 3(1  a)2/3 1  (1 a)1/3
One-dimentional diffusion D1 ½a a2
Two-dimentional diffusion D2 [ln(1  a)]  1 a þ(1  a)[ln(1  a)]
Three-dimensional diffusion (Jander) D3 3(1  a)2/3/[2(1  (1  a)1/3)] [1 (1  a)1/3]2
Three-dimensional diffusion (Ginstling–Brounshtein) D4 3/2((1  a)  1/3  1) 1  (2a/3)  (1 a)2/3
Three-dimensional diffusion (Zhuravlev) D5 (3/2)(1 þ a)2/3[(1 þ a)1/3  1]  1 [(1 þ a)1/3  1]2
Avrami–Erofeev A2 2(1  a)[  ln(1  a)]1/2 [  ln(1  a )]1/2
Avrami–Erofeev A3 3(1  a)[  ln(1  a)]2/3 [  ln(1  a)]1/3
First-order (Mampel) F1 1a  ln(1  a)
Power law P2 2a1/2 a1/2
Power law P3 3a2/3 a1/3

radiolysis of nitrate are NO2 and O. The radiolytic oxygen becomes


trapped in the crystal (Mohanty and Patnaik, 1989). The presence of 7
these damaged species in the crystal matrix will result in a steady
accumulation of strain. 6
On melting the crystal defects are removed and only the
chemical damage persists (Samuel et al., 2011). In the melt phase, 5

g(α )/g(0.5)
a large proportion of the damage oxygen escapes leaving behind a
portion of NO2 . The NO2 may act as an electron donor and attract 4
oxygen atom from the neighbouring nitrate ion. The resulting
nitrite ion being oxygen deficient may attract oxygen atoms from 3
the neighbouring nitrate ions. In this manner the reaction may
proceed through the liquid phase and a stream of oxygen would 2
flow from the interior to the surface along a line of NO2 ions (Nair
and James, 1985). Thus the NO2 produced by irradiation, cata- 1
lyses the decomposition of the salt. Due to the catalysing effects,
the activation energy is lowered. The fall in the initial horizontal 0
level in the case of g-irradiated samples can be attributed to the 0.0 0.2 0.4 0.6 0.8 1.0
effect of irradiation. α

Fig. 4. Master plots of theoretical g(a)/g(0.5) against a for various reaction models
3.5. Determination of reaction model by using master plot method and experimental data for the thermal decomposition of unirradiated Sr(NO3)2 at
5 1C min  1.

Integral master-plot method (Gotor et al., 2000; Perez-Maqueda


et al., 2002) was used for the determination of the reaction model
Plotting g(a)/g(0.5) against a corresponds to the theoretical
for the thermal decomposition of Sr(NO3)2. Essentially the master
master plots of various kinetic functions (Table 4). Using the
plot method is based on the comparison of theoretical master plot,
predetermined value of E from the isoconversional method, along
which are obtained for a wide range of ideal kinetic models, with
with the temperature measured as a function of a, p(u) can be
the experimental master plot. The integral function of conversion
calculated according to Eq. (6). Then the experimental master
in the solid state non-isothermal decomposition reactions is
plots of p(u)/p(u0.5) against a under various heating rates for both
expressed as:
 Z T     untreated and irradiated samples of Sr(NO3)2 can be drawn. In the
A E AE present work, it is found that the experimental curves of all the
gðaÞ ¼ exp dT ¼ pðuÞ ð5Þ
b T0 RT bR samples at all heating rates overlaps the master-curves of g(D3)/
Ru g(0.5), indicating that thermal decomposition of both untreated
where u¼E/RT. The temperature integral, pðuÞ ¼ 1 ðeu =u2 Þdu
and irradiated samples of Sr(NO3)2 follows D3 reaction mechan-
has no analytical solution and can be expressed by an approxima-
ism. As a representative, we show the master-plot of the uni-
tion. The rational approximation of Doyle (1962) gives sufficiently
rradiated Sr(NO3)2 at 5 1C min  1 in Fig.4.
accurate results
pðuÞ ¼ 0:00484expð1:0516uÞ ð6Þ
4. Conclusions
Using a reference at point a ¼0.5 and according to Eq. (5), one
gets
The present work deals with the kinetic analysis of the non-
 
AE isothermal decomposition of both untreated and g-irradiated
gð0:5Þ ¼ pðu0:5 Þ ð7Þ
bR Sr(NO3)2 in nitrogen atmosphere at different heating rates. Our
results demonstrate that the decomposition of Sr(NO3)2 proceeds in
where u0.5 ¼E/RT0.5
a single step without the formation any intermediate nitrite. The
When Eq. (5) is divided by Eq. (7), the following equation
model-free isoconversional methods (FWO, KAS and FR) as well as
obtained
master-plot method were used for kinetic analysis of TG curves. A
gðaÞ pðuÞ close agreement of the values of activation parameters were
¼ ð8Þ
gð0:5Þ pðu0:5 Þ observed between the KAS and FWO methods. Friedman method
S. Culas, J. Samuel / Radiation Physics and Chemistry 86 (2013) 90–95 95

leads to slightly higher values of activation energy compared with James, C., Samuel, J., 2003. The effect of gamma-irradiation on the thermal
other two methods. The irradiation enhances the thermal decom- decomposition of anhydrous cadmium nitrate. J. Radioanal. Nucl. Chem. 258,
665–668.
position due to the catalysing effects of the fragments produced by James, C., Samuel, J., 2005. Annealing of chemical radiation damage in gamma-
irradiation. The master-plot method revealed that the kinetic irradiated anhydrous lanthanum nitrate. Radiat. Eff. Defect Solids 160,
process of the thermal decomposition of both untreated and 329–335.
Kershaw, N.F., Chamberlin, N.S., 1942. Determination of nitrite, discussion of Shinn
g-irradiated Sr(NO3)2 is best described by Jander diffusion model,
methods applied to examination of water. Ind. Eng. Chem. (Anal. Ed.) 14,
g(a)¼[1 (1 a)1/3]2. 312–313.
Kissinger, H.E., 1957. Reaction kinetics in differential thermal analysis. Anal. Chem.
29, 1702–1706.
Acknowledgements Mahfouz, R.M., Alshehri, S.M., Monshi, M.A.S., Abd El-salam, N.M., 2000. Isother-
mal decomposition of g-irradiated samarium acetate. Radiat. Phys. Chem. 59,
381–385.
The authors are thankful to the University of Kerala for the Mohanty, S.R., Patnaik, D., 1989. Effects of admixtures of potassium bromide on
award of research fellowship and NIIST, and SCTIMST, Thiruva- the thermal decomposition of potassium bromate. J. Therm. Anal. Calorim. 35,
nanthapuram for lending the analytical facilities. 2153–2159.
Muhammad, D., Maddock, A.G., 1978. Radiolysis of the alkali nitrates. J. Chem. Soc.
Faraday Trans. I 74, 919–932.
References Nair, S.M.K., Krishnan, M.S., James, C., 1983. The kinetics of isothermal annealing of
gamma-irradiation damage in crystalline barium nitrate. Radiat. Eff. 71,
87–93.
Abdul Mujeeb, V.M., Aneesh, M.H., Muraleedharan, K., Devi, T.G., Kannan, M.P., Nair, S.M.K., James, C., 1984. The effect of gamma-irradiation on the thermal
2011. Effect of precompression on isothermal decomposition kinetics of decomposition of strontium nitrate. J. Radioanal. Nucl. Chem. Lett. 86,
potassium bromate. J. Therm. Anal. Calorim. 104, 991–997. 311–320.
Akahira, T., Sunnose, T., 1971. Trans joint convention of four electrical institutes, Nair, S.M.K., James, C., 1985. Thermal decomposition of strontium nitrate doped
Paper no. 246, 1969. Res. Rep. Chiba Inst. Technol. 16, 22–31. with strontium nitrite at millimolar concentrations. Thermochim. Acta 87,
Allen, A.D., Ghormley, J.A., 1947. Decomposition of solid barium nitrate by fast 367–371.
electrons. J. Chem. Phys. 15, 208–209. Nair, S.M.K., Jacob, P.D., 1989. The effect of gamma-irradiation on the thermal
Brown, W.E., Dollimore, D., Galwey, A.K., 1980. Comprehensive Chemical Kinetics,
decomposition of magnesium bromate. J. Radioanal. Nucl. Chem. Lett. 2,
22. Elsevier, Amsterdam.
113–125.
Coats, A.W., Redfern, J.P., 1964. Kinetic parameters from thermogravimetric data.
Narayanswamy, L.K., 1935. On the photo-dissociation of single crystals of some
Nature 201, 68–69.
nitrates in polarised light. Trans. Faraday Soc. 31, 1411–1412.
Cunningham, J., Heal, H.G., 1958. The decomposition of solid nitrates by X-rays.
Ozawa, T., 1965. A new method of analysing thermogravimetric data. Bull. Chem.
Trans. Faraday Soc. 54, 1355–1369.
Soc. Jpn. 38, 1881–1886.
Das, B.C., Patnaik, D., 2000. Effect of anion doping on the thermal decomposition of
Perez-Maqueda, L.A., Criado, J.M., Gotor, F.J., Malek, J., 2002. J. Phys. Chem. A 106,
potassium bromate. Therm. Anal. Calorim. 61, 879–888.
2862–2869.
Diefallah, E.M., Baghlaf, A.O., Mahfouz, R.M., 1985. Chemical effects of cobalt
Samuel, J., Culas, S., Surendran, A., 2011. Evaluation of kinetic parameters for the
neutron capture recoils in K3Co(CN)6-K3Fe(CN)6 and K3Co(CN)6-K3Cr(CN)6
thermal decomposition of gamma-irradiated anhydrous calcium nitrate. Asian
mixed crystals. J. Radioanal. Nucl. Chem. 94 (2), 109–120.
Dogan, F., Kaya, I., Bilici, A., 2009. Non-isothermal decomposition kinetics of J. Chem. 23, 1673–1676.
poly(2, 20 -dihydroxy biphenyl). Polym. Bull. 63, 267–282. Samuel, J., Surendran, A., Culas, S., 2013. Kinetic analysis for non-isothermal
Doigan, P., Davis, T.W., 1952. The photolysis of crystalline nitrates. J. Phys. Chem. decomposition of un-irradiated and gamma-irradiated potassium bromate. J.
56, 764–766. Radioanal. Nucl. Chem. 295, 53–61.
Doyle, C.D., 1962. Estimating isothermal life from Thermogravimetric data. J. Appl. Shinn, M.B., 1941. Colorimetric method for determination of nitrite. Ind. Eng.
Polym. Sci. 6, 639–642. Chem. (Anal. Ed.) 13, 33–35.
Duval, C., 1963. Inorganic Thermogravimetric Analysis. Elsevier, Amsterdam Srivastava, S., Zulfequar, M., Kumar, A., 2010. Study of crystallisation kinetics in
pp. 494. glassy Se100-x Bix using isoconversional methods. J. Non-oxide Glasses 2,
Friedman, H.L., 1964. Kinetics of thermal degradation of char-forming plastics 97–106.
from thermogravimetry application to a phenol plastic. J. Poym. Sci. C 6, Tita, B., Fulias, A., Bandur, G., Rusu, G., Tita, D., 2010. Thermal stability of ibuprofen.
183–195. Kinetic study under non-isothermal conditions. Rev. Roum. Chim. 55,
Flynn, J.H., Wall, L.A., 1966. A quick, direct method for the determination of 553–558.
activation energy from thermogravimetric data. J. Polym. Sci. Part B Polym. Vyazovkin, S., 1997. Advanced isoconversional method. J. Therm. Anal. 49,
Lett. 4, 323–328. 1493–1499.
Golding, R.M., Henchman, M., 1964. ESR study of NO2 and NO3 in irradiated lead Vyazovkin, S., 2000. Computational aspects of kinetic analysis. Part C.The ICTAC
nitrate. J. Chem. Phys. 40, 1554–1564. kinetic project-the light at the end of the tunnel. Thermochim. Acta 355,
Gotor, F.J., Criado, J.M., Malek, J., Koga, N., 2000. Kinetic analysis of solid-state 155–163.
reactions: the universality of master plots for analysing isothermal and Vyazovkin, S., Burnham, A.K., Criado, J.M., Perez-aqueda, L.A., Popescu, C., Sbirraz-
nonisothermal experiments. J. Phys. Chem. A 104, 10777–10782. zuoli, N., 2011. ICTAC kinetics committee recommendations for performing
Hennig, G., Lees, R., Matheson, M.S., 1953. The decomposition of nitrate crystals by kinetic computations on thermal analysis data. Thermochim. Acta 520, 1–19.
ionising radiations. J. Chem. Phys. 21, 664–668. Young, D.A., 1966. Decomposition of Solids. Pergamon Press, Oxford.

You might also like