242
Views
0
CrossRef citations to date
0
Altmetric
Research Articles

Boiling and Drying Accident of High-Level Liquid Waste in a Reprocessing Plant: Examination of the Time-Dependent Temperature Increase of the Waste and of the Generation Rates of the Individual Components Released into the Gas Phase

, , ORCID Icon, , &
Pages 958-984 | Received 28 Apr 2023, Accepted 09 Oct 2023, Published online: 12 Feb 2024

Abstract

Using the amount, composition, and decay power density of high-level liquid waste in a storage tank, the temperature change of the waste up to 600°C and the corresponding vapor and gas release rates of H2O, HNO3, NO2, NO, and O2 as a function of time after the loss of cooling function were obtained by the following method. The heat balance equations in and around the tank were derived, and the solution of the waste temperature change was numerically obtained using the vaporization rates of H2O and HNO3 and the generation rate of NOx, which were both obtained from the experiments using the simulated liquid waste. Utilizing the temperature versus time curve obtained from the equation, the release rates of the components described above were obtained as a function of time. This information on the progress of the accident can be used to study the Leak Path Factor of radioactive materials, especially of volatilized Ru, and further, it becomes basic information when considering accident management and suppressing the impact of a disaster.

I. INTRODUCTION

The boiling and drying accident of high-level liquid waste (HLLW) has drawn considerable attention among stakeholders because of high potential radiation exposure to the general public. Accordingly, many experiments and analyses have been conducted.[Citation1–10] In order to assess the radiation exposure, the amount of radioactive aerosols and volatilized 106Ru released into the environment is necessary to be known. To do this, it is necessary to understand the behavior of H2O and HNO3 vaporization and of NO2, NO, and O2 generation, which not only carry the radioactive materials but also affect the Leak Path Factor (LPF),[Citation11] especially of volatilized 106Ru.[Citation9]

Previously, we reported[Citation10] the NO2, NO, and O2 generation rates from the simulated high-level liquid waste (SHLLW) up to 600°C under constant temperature increasing rates of 0.2 and 1°C∙min−1. In this study, using the amount, composition, and decay power density of HLLW in a storage tank, we predict how the wasteFootnotea temperature up to 600°CFootnoteb and also the individual generation rates of H2O and HNO3 vapors and NO2, NO, and O2 gases will change with time, based on the heat balance equation related to the inside and outside of the storage tank and experimental results on the H2O and HNO3 vaporization and NOx generation using SHLLW.

Similar reports on the study of the waste temperature and the gas release rates as a function of time exist so far.[Citation2–4,Citation7] However, we consider that the contents of these reports are not sufficient as shown below.

Ishikawa et al. reported that the temperature increase of the waste up to 140°C could be predicted by extrapolating the molar boiling point elevation model.[Citation4] However, HLLW is a nitric acid solution in which various types of nitrates are dissolved at high concentrations and precipitate with increasing temperature. Therefore, we consider that it is difficult to construct and solve the theoretical model up to 140°C. Indeed, the agreement between the predicted values and the experimental ones is not always good as the temperature approaches 140°C. For the analysis from 100°C to 800°C, Yoshida and Ishikawa[3] used the MELCOR code.[Citation12] But, the code cannot handle nitric acid, so they introduced control functions and multiple state input volumes into the code to model the key phenomena of boiling events such as boiling above 100°C, nitric acid vaporization, and NOx gas generation.[Citation3] However, in a subsequent paper[Citation6] from the same group, the results are said to be qualitative.

To obtain the temperature versus time curve up to 600°C, it is necessary to solve the heat balance equation around the storage tank. The H2O and HNO3 vaporization, nitrate decomposition, and heat transfer to the cell air and cell wall proceed simultaneously. This paper describes an integrated model for simulating time-dependent changes in the waste temperature and the released gas composition, as well as the solution method and the results obtained. There is no other report of such contents.

Note that the values used in the following analysis may not reflect the actual plant values.

II. OUTLINE OF THE MODEL TO BE ANALYZED

One tank is installed in one cell. The tank is filled with HLLW. Cell ventilation air is flowing into the cell but stops due to the accident. The corresponding numerical values are as follows:

  1. The amount of HLLW is 120 m3.

  2. The properties of HLLW are as follows. The amount of 0.4 m3 of HLLW is produced by reprocessing 1 ton of spent fuel (pressurized water reactor fuel, specific power of 38 MW∙ton−1, burnup of 45 000 MWd∙ton−1, and cooling period of 6 years), and its composition is obtained using the ORIGEN-2 code. The decay power of 120 m3 of HLLW becomes 578 kW.

  3. Cell ventilation air is introduced with a flow rate of 57 mol∙s−1 and at 25°C.

  4. All six outer cell walls are surrounded by concrete building walls, and the air in the space and the building walls both have a temperature of 25°C.

The temperature of HLLW rises with time after the cooling function loss and boiling begins. As time passes, the liquid waste dries and solidifies. Therefore, in the following analysis, the model was divided into two stages: the initial stage and the late stage. In the initial stage, the waste remains in the liquid phase, and so, the temperature of the tank wall and inside the tank is considered to be uniform due to a large amount of steam generated by the decay heat. On the other hand, in the late stage, the temperature distribution occurs inside the tank. Based on the experimental results (see Sec. IV.A.2), we assumed that the waste was in a liquid phase up to 130°C and above that was in a solid phase.Footnotec

III. DERIVATION OF THE HEAT BALANCE EQUATION AND ITS SOLUTION METHOD IN THE INITIAL STAGE

III.A. Heat Balance Equation

The various heat transfer processes and related parameters affecting the waste temperature θ(t) (in degrees Celsius) are shown in . Although most of the decay energy is dissipated in the tank, part of the gamma rays is released from the tank contributing to the temperature rise of the cell wall (gamma heating). The decay energy dissipated in the tank is used not only for the temperature rise of the waste and the tank but also for vaporization of H2O and HNO3, decomposition of nitrates, and heat loss out of the tank (convection and heat radiation), and as a result, the waste temperature is determined.

Fig. 1. Heat transfer processes and parameters affecting the waste temperature at the initial stage.

Fig. 1. Heat transfer processes and parameters affecting the waste temperature at the initial stage.

The amount of heat transfer associated with the individual heat transfer processes shown in depends on the following parameters. Variable t (in seconds) means the time after the cooling loss. The value θ (in degrees Celsius) is a function of t, that is, θ(t), and sometimes is described as simply θ. Also, note the following parameters 1 through 8:

  1. Initial storage amount of the waste M0 (m3).

  2. Energy release rate due to the radioactive decay Rh (kW) and the leak rate of gamma-ray energy out of the tank (gamma heating) Rγ(t) (kW); the power of [RhRγ(t)] is used to heat the waste and the tank.

  3. Amount of H2O in the waste MW(t) (mol), its vaporization rate YW(t) (mol∙s−1), latent heat λW(θ) (kJ∙mol−1), and molar specific heat cW (kJ∙mol−1∙°C−1); cW was assumed to be independent of temperature.[Citation13]

  4. Amount of HNO3 in the waste MN(t) (mol), its vaporization rate YN(t) (mol∙s−1), latent heatFootnoted λN(θ) (kJ∙mol−1), and molar specific heat cN (kJ∙mol−1∙°C−1); cN was assumed to be independent of temperature.[Citation13]

  5. Heat absorption rate due to the thermal decomposition of nitrates (salts) RS(t) (kW) and the total specific heat cS(θ) (kJ∙kg−1∙°C−1); the specific heat of each nitrate was assumed to be equal to that of the oxide, and the oxides were assumed to have a constant total mass MS (kg) irrespective of temperature (adequacy of the assumptions is explained in Sec. VI.D).

  6. Mass of the tank including the internal structural material MT (kg) and its specific heat cT(θ) (kJ∙kg−1∙°C−1).

  7. The temperature of the tank is assumed to be the same as that of the waste.

  8. Heat transfer rate from the tank to the cell air by natural convection Rloss1(θ) (kW) and that from the tank to the cell wall by heat radiation Rloss2(θ) (kW), and their sum Rloss(θ) = Rloss1(θ) + Rloss2(θ). The cell ventilation function is assumed to be lost due to the accident.

Using the foregoing parameters, the waste temperature θ(t) is described by the following heat balance equationFootnotee:

(1) Γtdt=RhλwθYwt+λNθYNt+RStRlossθRγt,(1)(1)

where Γ(t) = sum of the heat capacities of the waste (H2O, HNO3, and the nitrates = the oxidesFootnotef ) (kJ∙°C−1) and the tank and is given by the following equation:

(2) Γt=CWMWt+cNMNt+cSθMS+cTθMT.(2)

The right side of EquationEq. (1) shows the net power to increase the temperature of the waste and the tank.

III.B. Solution Method

In order to obtain the solution of EquationEq. (1) up to 130°C, it is necessary to know the time change of the individual parameters 1 through 8 shown in Sec. III.A. In the present study, the relationship between the individual vaporization rates of H2O and HNO3 and the waste temperature was obtained from the experimental results as shown in Sec. IV. The relationship between the heat absorption rate accompanying the decomposition of nitrates and the waste temperature is examined in Sec. V. In this examination, the experimental results of the previous report concerning the NOx generation[Citation10] were used. The heat capacity Γ(t) in EquationEq. (1) is treated in Sec. VI, and the heat transfer from the tank including gamma heating is dealt with in Sec. VII. These results are integrated to solve EquationEq. (1) by the difference method in Sec. VIII. Then, using the result of EquationEq. (1), the generation rates of gaseous components are obtained as a function of time.

IV. HEAT ABSORPTION RATE OF H2O AND HNO3 VAPORIZATION

The content in this section can be used for both stages. The heat absorption rate due to liquid vaporization is the sum of the corresponding individual values of H2O and HNO3. Each heat absorption rate is obtained by multiplying the vaporization rate and the latent heat. The vaporization rates were obtained from the results of experiments using SHLLW as described below. Since H2O and HNO3 are vaporized from nitric acid solution, the interaction between H2O and HNO3 needs to be taken into account to obtain the individual latent heats. This effect is described in Secs. IV.B.3. and IV.C.3.

IV.A. Measurement of the Vaporization Rates from the Waste

IV.A.1. Experimental

Two types of the simulated wastes, SHLLW-J and SHLLW-K, were prepared. The individual compositions are shown in . SHLLW-J simulates HLLW, and SHLLW-K contains only the main 12 compounds. The details are given in the previous report.[Citation9] SHLLW-J was used to examine the vaporization rates of H2O and HNO3 as a function of time, and SHLLW-K was used to examine the effect of the temperature increasing rate on their vaporization ratios versus temperature curves (see Sec. IV.A.2.b). Two runs were carried out using SHLLW-J and five runs using SHLLW-K. Individual experimental conditions are given in .

TABLE I Compositions of the Simulated Solutions

TABLE II List of the Experimental Conditions for Each Run

The apparatus to measure the vaporization rate is shown schematically in . SHLLW-J of 120 mL was charged into a separable flask with an inner diameter of 85 mm and a height of 110 mm (nominal volume of 500 mL). For SHLLW-K, 120 to 200 mL was charged. Two thermocouples were inserted to measure the waste temperature near the bottom and the vapor temperature at the top of the flask. The flask was heated on a hot plate. A 2-mm-thick aluminum plate was placed between the flask and the hot plate so that the heating from the bottom was uniform. In addition, flexible rubber heaters were wound around the side of the separable flask and the lid of the flask. The pipe connecting the steam outlet to the inlet of a Liebig condenser was heated by a ribbon heater. The waste temperature was controlled by the hot plate and the temperatures of the side wall; the lid and the pipe were controlled to be the same as the waste temperature.

Fig. 2. Schematic diagram of the apparatus used for the vaporization experiments.

Fig. 2. Schematic diagram of the apparatus used for the vaporization experiments.

Steam generated in the flask flowed into the condenser, and the condensate was recovered periodically. The temperature of the cooling water was 15°C. The recovery was performed at appropriate intervals in a batch manner from the boiling point to about 280°C (the sampling interval can be seen from the data points in shown in the next section). At temperatures above 120°C where the amount of condensate became small, 10 mL of H2O was injected into the condenser before recovering the condensate through the pipe shown in to wash and collect the condensate attached on the inner wall.

Fig. 3. Time changes of vaporized amounts of H2O and HNO3.

Fig. 3. Time changes of vaporized amounts of H2O and HNO3.

The mass and density of the recovered liquid including the condensate and the washing water were measured for each batch, and then, the liquid was diluted to 1000 to 10 000 times to measure the pH. These results were used to obtain the net amounts of H2O and HNO3 recovered in the condenser. It should be noted that the measured amount of HNO3 is different from the amount vaporized in the flask as described in the next section.

In order to observe the state of the waste, a vertical slit was made in the rubber heater that was wound around the side of the flask. As boiling began and concentration proceeded, it became unclear whether the waste was liquid or solid. At 130°C, no bubbling was observed, and the waste became like a paste. From the previous experiment, the following is known.[Citation10] When the temperature reached 180°C, the waste was more like a solid than a paste, but the waste taken out from the flask was somewhat moist and gave off a strong HNO3 odor. The waste heated to 280°C to 300°C was dry. It was solid with many voids.

IV.A.2. Results and Discussion

IV.A.2.a. Time Change of Vaporized Amounts of H2O and HNO3

We examined the amount of vaporized H2O and HNO3 using SHLLW-J. The initial amount of H2O in the liquid waste obtained was as follows:

  1. Liquid waste density: 1270 kg∙m−3 (measured value of SHLLW-J at room temperature).

  2. Amount of HNO3: 126 kg∙m−3 = 2000 mol∙m−3 (calculated from the concentration of 2 mol∙L−1).

  3. Amount of nitrates (including H2MoO4, SnO2, Sb2O3, and TeO2): 276 kg∙m−3 (calculated from assuming anhydrous nitrates).

  4. Amount of H2O: 868 kg∙m−3 = 48 200 mol∙m−3 (calculated by subtracting the contribution of HNO3 and nitrates from the liquid waste density).

The initial amount of H2O obtained in this way, 48 200 mol∙m−3, became the same as the final vaporized amount shown in , which shows the time-dependent changes of the measured cumulative amounts of H2O (and HNO3) obtained from Runs J1 and J2. In the figure, the cumulative amount is shown as molar quantity per 1 m3 of initial liquid waste, and the boiling start point corresponds to = 0.

On the other hand, the measured cumulative amount of HNO3 at 280°C is about 2600 mol∙m−3 for both runs as shown in , which is different from the initial amount of 2000 mol∙m−3. This difference is explained as follows. Previously, we reported[Citation10] that most of the Zr, Ru, and Pd nitrates in SHLLW-J do not produce NOx but produce HNO3 by 180°C, while HNO3 left in the drying waste produces Nox mainly above 300°C, and that the difference in the amount between them is 100 mol∙m−3. Therefore, the final vaporized amount of HNO3 should be equal to 2100 mol∙m−3. The difference between 2100 mol∙m−3 and the measured 2600 mol∙m−3 should be due to the amount produced from NOx, mainly NO2, in the condenser by the following reactions[Citation14]:

(3) 3NO2+H2O2HNO3+NO,(3)
(4) 2NO2+H2OHNO2+HNO3,(4)
3HNO2HNO3+2NO+H2O,

where H2O comes from the condensate and the washing water injected into the condenser. Formulas (3) and (4) show that H2O is consumed in absorption of NO2 and that the amount of H2O in mol∙m−3 in the initial waste is about 24 times larger than that of HNO3, and so, the consumption due to the NO2 absorption is negligible. The cumulative amount curve of vaporized HNO3 shown in was obtained by using the final value of 2100 mol∙m−3 and correcting the amount of HNO3 produced by NO2 according to its rate of generation.

IV.A.2.b. Effect of the Temperature Increasing Rate on the Vaporization Ratios

In five runs using SHLLW-K, the heating time from 104°C (boiling start temperature) to 300°C was changed variously as shown in , and the effect of the temperature increasing rate on the vaporization ratios of H2O and HNO3 was examined. Individual vaporization ratios were obtained using the corresponding final vaporization amounts as the denominator and the cumulative vaporized amounts at that temperature as the numerator. The final amounts were obtained in the same way as for SHLLW-J described above. The vaporization ratio versus temperature curves for H2O and HNO3 thus obtained are shown in and , respectively. These curves can be approximated by the individual single curves for H2O and HNO3 irrespective of the temperature increasing rate. As shown in Secs. IV.B.1 and IV.C.1, these individual curves are consistent with those obtained from SHLLW-J. By 155°C, almost all H2O and 80% HNO3 are vaporized. The temperature increasing rates up to 155°C used in cover the realistic rates caused by the HLLW’s decay power density of 4.82 kW∙m−3 as seen in of Sec. X.A.

Fig. 4. H2O vaporization ratio versus temperature curves obtained from Runs K1 through K5.

Fig. 4. H2O vaporization ratio versus temperature curves obtained from Runs K1 through K5.

Fig. 5. HNO3 vaporization ratio versus temperature curves obtained from Runs K1 through K5.

Fig. 5. HNO3 vaporization ratio versus temperature curves obtained from Runs K1 through K5.

IV.B. Heat Absorption Rate of H2O Vaporization

The heat absorption rate due to the vaporization is given by the product of the vaporization rate and the latent heat of vaporization. The vaporization rate can be obtained from the above experimental results, and the latent heat can be obtained from the literature.[Citation15]

IV.B.1. Relationship Between the H2O Vaporization Ratio and Temperature

The H2O vaporization ratio ξW(θ), that is, the ratio of the cumulative amount at temperature θ to its final amount, is shown in , which was obtained from the experimental results of Runs J1 and J2. When θ < 104°C (before boiling), ξW(θ) = 0. Based on the results of the figure, the curve for dξW(θ)/dθ was obtained as shown in . Using the figure, the following approximate expressions were obtained. The solid lines in the figure correspond to these expressions:

(5) 104 <θ150C,log10{dξWθ/} =0.0474θ+ 3.952θ> 150C,log10{dξWθ/} =0.00709θ2.094.(5)

Fig. 6. H2O and HNO3 vaporization ratios versus temperature curves obtained from Runs J1 and J2.

Fig. 6. H2O and HNO3 vaporization ratios versus temperature curves obtained from Runs J1 and J2.

Fig. 7. The W/dθ versus temperature curve obtained from Runs J1 and J2.

Fig. 7. The dξW/dθ versus temperature curve obtained from Runs J1 and J2.

Although the data points scatter above 160°C, it is attributed to a small amount of vaporization and is not important since most of the H2O has already been vaporized.

IV.B.2. Vaporization Rate

The H2O vaporization rate YW(t), which is equal to the time derivative of the cumulative amount of H2O vaporization MW(0)ξW(θ), is given as follows:

(6) YWt=MW0dξWθdt=MW0dξWθdt.(6)

The final cumulative amount of vaporized H2O is equal to the initial value of MW(0) as explained in Sec. IV.A.2.a. The value dξW(θ)/dθ in EquationEq. (6) is given by EquationEq. (5).

IV.B.3. Latent Heat of Vaporization

For the latent heat of H2O vaporization λW(θ) (in kJ∙mol−1), the following approximate equations created from data from 50°C to 373.95°C[Citation16] were used:

θ370C,
λWθ=450θ/1006+ 5070θ/100522,600θ/1004+ 50,000θ/100358,100(θ/100)2+28,700(θ/100)+38,000;
370<θ374C,
λWθ = 2.09θ 370 + 8.25.

Since H2O is vaporized from nitric acid solution in the present case, the interaction between H2O and HNO3 needs to be taken into account. This effect was approximated by the heat of dissolution of HNO3 in water and included in the latent heat of HNO3 vaporization (see Sec. IV.C.3).

IV.B.4. Heat Absorption Rate

The product of the vaporization rate YW(t) and its latent heat λW(θ) obtained above is the heat absorption rate.

IV.C. Heat Absorption Rate of HNO3 Vaporization

IV.C.1. Relationship Between the HNO3 Vaporization Ratio and Temperature

The HNO3 vaporization ratio ξN(θ), that is, the ratio of the cumulative amount at temperature θ to its final amount obtained from Runs J1 and J2, is also plotted in . When θ < 104°C (before boiling), ξN(θ) = 0. Based on the results of the figure, the curve for dξN(θ)/dθ was obtained as shown in . Using the figure, the following approximate expressions were obtained. The solid lines in the figure correspond to the following approximate expressions:

(7) 104<θ126C,log10dξNθ/=0.0647θ9.59126<θ170C,log10dξNθ/=0.0277θ+2.055θ>170C,log10dξNθ/=0.00202θ2.313.(7)

Fig. 8. The dξN/dθ versus temperature curve obtained from Runs J1 and J2.

Fig. 8. The dξN/dθ versus temperature curve obtained from Runs J1 and J2.

IV.C.2. Vaporization Rate

The HNO3 vaporization rate YN(t), which is equal to the time derivative of the cumulative amount of HNO3 vaporization MN(0)ξN(θ), is given as follows:

(8) YNt=MN0dξNθdt=MN0dξNθdt,(8)

where dξNθ/ is given by EquationEq. (7).

IV.C.3. Latent Heat of Vaporization

The latent heat of HNO3 vaporization from aqueous nitric acid λN (in kJ∙mol−1) was approximated as the sum of the latent heat of HNO3 vaporization from pure HNO3 and the dissolution heat of HNO3 in water:

λNθ=39.43×1θ+273.15/5201293.15/5200.375+Hsol.

The first term on the right side is the latent heat of HNO3 vaporization,[Citation4] and the second term Hsol (in kJ∙mol−1) is the dissolution heat of HNO3 in water. The latter was calculated by the following approximate EquationEq. (9) created from the data at 25°C[Citation15]:

(9) 0.03 <ω0.265,Hsol=485ω3+308ω214.3ω+ 0.67ω> 0.265,Hsol= 43.5ω2.02,(9)

where ω = molar fraction of HNO3. Hsol does not depend on temperature as is explained below.

The derivative of the dissolution heat with respect to temperature is equal to the difference between the heat capacity of the solution and the sum of the heat capacities of H2O and HNO3. Since the heat capacity of nitric acid solution could be assumed to be equal to the sum of the heat capacities of H2O and HNO3 as is explained in Sec. VI.A, the dissolution heat does not depend on temperature. Note that the values of ω and Hsol change with time as a result of vaporization.

IV.C.4. Heat Absorption Rate

The product of the HNO3 vaporization rate YN(t) and the latent heat of vaporization λN(θ) obtained above is the heat absorption rate accompanying HNO3 vaporization.

V. HEAT ABSORPTION RATE OF THERMAL DECOMPOSITION OF NITRATES

The content in this section can be used for both stages. The total heat absorption rate RS(t) associated with the thermal decomposition of nitrates is equal to the sum of products obtained by multiplying the decomposition heats required for the generation of unit amounts of NO2 and NO by the respective generation rates (see Sec. V.C). In the following, the decomposition heat and the generation rate are examined.

V.A. Heats of Decomposition Generating NO2 and NO

In the waste, various nitrates are included. The decomposition heat is estimated as follows. Increasing the temperature of the waste, the nitrates generate not only NO2 and NO but also HNO3.[Citation10] However, the decomposition heat generating HNO3 is small as is explained in Appendix A, and so, it is not taken into consideration.

First, the decomposition heat generating NO2 is examined taking Nd(NO3)3 as an example:

4NdNO332Nd2O3+12NO2+3O2.

The enthalpy of this reaction Δd1H (in kJ∙mol−1) is obtained as

Δd1H=ΔfHNd2O3/2+3ΔfHNO2ΔfHNdNO33,

where ΔfH(Nd2O3), ΔfH(NO2), and ΔfH(Nd(NO3)3) are their formation enthalpies.

The decomposition reaction of Nd(NO3)3 to generate NO and its enthalpy Δd2H are given as follows:

4NdNO332Nd2O3+12NO+9O2
Δd2H=ΔfHNd2O3/2+3ΔfHNOΔfHNdNO33.

Therefore, if the formation enthalpies of Nd(NO3)3, Nd2O3, NO, and NO2 are given, the enthalpy of the nitrate decomposition is obtained. The temperature dependence of the enthalpy is several percent within the scope of this study,[Citation13] so it was not considered, and the standard enthalpy values were used. In , the standard enthalpies related to the various nitrates and oxides in the waste are shown. In the table, the heat absorption accompanying the individual nitrate decomposition is also shown. The heat absorption shown in units of kJ∙L−1 is obtained by multiplying Δd1H (or Δd2H) by the number of moles of nitrate per 1 L of waste (see ). Standard formation enthalpies of all oxides and part of the nitrates were obtained from the literature.[Citation13,Citation17] The enthalpies of the nitrates that are not given in the literature were estimated as follows.

TABLE III Estimated Heat Absorption Associated with Nitrate Decomposition that Produces NO2 or NO

Many standard formation enthalpies of chlorides are given in the literature.[Citation13] For Mn, Ni, Sr, Ag, Cd, Cs, Ba, and Ce elements, standard formation enthalpies of chlorides and nitrates are given. There is a good correlation between the two, which can be approximated by the following equation:

(10) ΔfHnitrate = 1.17ΔfHchloride.(10)

If the standard formation enthalpy for nitrate was not found in the literature, it was estimated from the standard formation enthalpy of chloride using EquationEq. (10). The values obtained in this way are listed as superscript a (a) in .

The decomposition enthalpies of the individual nitrates generating NO2 or NO were estimated using the standard enthalpies obtained above and are shown in as Δd1H and Δd2H. The values of Δd1H/NO2 and Δd2H/NO show the decomposition heat generating 1 mol of NO2 and NO, respectively.

However, since the standard formation enthalpies of not only nitrate but also chloride of Zr and Ru are not found in the literature, the following estimation was used. Values of Δd1H/NO2 for nitrates excluding alkali and alkaline earth metal nitrates go into the range of 61 to 153 kJ∙mol−1, and so, the average of these values of 120 kJ∙mol−1 was adopted for ZrO(NO3)2 and RuNO(NO3)3. Similarly, 176 kJ∙mol−1 was adopted for Δd2H/NO of Zr and Ru nitrates.

In the following calculation, the generation ratio of NO2, i.e., NO = 80:20, reported in the previous paper[Citation10] was used. The decomposition heat of each nitrate in 1 L of initial waste (in kJ∙L−1) generating NO2 was calculated as the product of the concentration of the nitrate in the waste (in mol∙L−1), the amount of NO2 generated from 1 mol of nitrate (in mol-NO2∙mol−1), and the NO2 generation heat (in kJ mol-NO2−1). However, we reported previously[Citation10] that Zr, Ru, and Pd nitrates mostly decomposed by 180°C but that only 17% of N in these nitrates was released as NOx, and so, 17% of the concentrations of these nitrates given in were used to calculate the above decomposition heat. In , the decomposition heat of each nitrate obtained above is shown as “Heat,” and the amount of NO2 generated from each nitrate is also shown. In addition, we reported that with rising temperature, 0.66 mol of NOx was generated per 1 L of initial waste by the decomposition of HNO3.[Citation10] In , NOx generation from the decomposition of HNO3 is also included.

The amount of NO2 generated from 1 L of initial waste and the corresponding decomposition heat are 1.86 mol∙L−1 and 228 kJ∙L−1, respectively, as shown as “Sum” in the last row in . Therefore, the heat absorption per 1 mol of NO2 generation is obtained as

λNO2=123 kJmol1.

Similarly,

λNO=174 kJmol1.

V.B. NOx Generation Rate

The generation rates of NO2 and NO, expressed as RNO2(t) and RNO(t) (in mol∙s−1), are given by the following Arrhenius-type equations, respectively, which were obtained for a temperature increasing rate of 0.2°C∙min−1 previously.[Citation10] The amount of NOx in the equations is the amount generated from the initial waste of 120 m3:

(11) RNO2t=i=15dMitdt=i=15AiexpEiRθt+273.15Mit(11)(11)
(12) RNOt=i=67dMitdt=i=67AiexpEiRθt+273.15Mit(12)
Mit=Mi0expAi0texpEiRθt +273.15dt.

The value Mi(t) (in moles) shows an amount of NO3 in group i that generates NOx. It is assumed that the decomposition of 1 mol Mi(t) generates 1 mol of NO2 (= 1 to 5) or 1 mol of NO (= 6 and 7). The terms Ai (in inverse seconds), Ei (in kJ∙mol−1), and R (in kJ∙mol−1∙°C−1) are the frequency factor, activation energy, and gas constant, respectively. EquationEquation (11) approximates NO2 generation by five groups and EquationEq. (12) NO generation by two groups. The values for these parameters previously obtained[Citation10] are reprinted in after converting the waste volume 1 to 120 m3 and time minutes to seconds.

TABLE IV List of Constants Used in the Arrhenius-Type Equations

Since 0.25 mol of O2 is generated per 1 mol of NO2 generation and 0.75 mol of O2 per 1 mol of NO, the O2 generation rate RO2(t) is obtained from RNO2(t) and RNO(t) as follows[Citation10]:

RO2t = 0.25RNO2t + 0.75RNOt.

V.C. Heat Absorption Rate

The heat absorption rate RS(t) accompanying the decomposition of nitrate is obtained as the product of the NOx generation rate described in Sec. V.B and the amount of the heat absorption per 1 mol of NOx generation described in Sec. V.A. Therefore, RS(t) is given as follows:

RSt =λNO2RNO2t+ λNORNOt.

VI. HEAT CAPACITY

The content in this section can be used for both stages.

In addition to the vaporization of liquid and the decomposition of nitrates, the heat capacities of the waste and the tank are required as the heat absorption terms. These heat capacities are included in Γ in EquationEq. (1). The total heat capacity Γ is examined below.

VI.A. H2O and HNO3

As shown in EquationEq. (2), the heat capacity of the waste is the sum of the products of the individual amounts of H2O, HNO3, and nitrates in the waste and their corresponding specific heats. The following values were used for the molar specific heats of H2O and HNO3:

cW= 0.076kJmol1C1
cN= 0.110kJmol1C1.

The value of H2O at 100°C and that of HNO3 at 25°C are shown in the literature.[Citation16] Since the temperature dependence on these heat capacities is insignificant,[Citation18] it was not considered.

VI.B. Nitrates

Specific heats of the individual nitrates in the waste were not found in the literature. However, since the dissolved nitrates precipitate and further decompose into oxides with increasing temperature, we assumed that the heat capacity of the nitrate is approximated by that of the oxide even before boiling. Under this assumption, the total amount of the oxides in the initial waste becomes 153 kg∙m−3. The specific heat of 18 oxides lies between 0.27 to 0.79 kJ∙°C−1∙kg−1 (at 25°C),[Citation13,Citation15] and the average value weighted by their concentrations in the initial waste is 0.37 kJ∙°C−1∙kg−1. Using the specific heats of Ba, Cr, Fe, Sr, and Zr oxides at 400 and 800 K,[Citation13] the increase rates per 1°C of these oxides were found to be 0.0028°C−1 to 0.0033°C−1, averaging 0.0029°C−1. Therefore, the average specific heat of the nitrates cS(θ) (in kJ∙kg−1∙°C−1) multiplied by the total mass of the nitrates (= oxides) is approximated as follows:

(13) cSθ×153×120 = 6800 [1+0.0029(θ 25)].(13)

VI.C. Tank

The material of the tank (including the internal structural material) was assumed to be Type 316 stainless steel (SUS-316). Its heat capacity CT(θ) (in kJ∙°C−1) was obtained by the following formula:

(14) CTθ =cTθ ×70,000=15.61θ +34,200,(14)

where cT(θ) = specific heat of SUS-316 at temperatures between 27°C and 727°C[Citation18]; 70 000 kg = mass of the tank.

VI.D. Examination of the Nitrate Heat Capacity Estimation Error

Since some assumptions are included in the derivation of the heat capacity of the nitrates, the effect of the error was examined. Using the initial amount of the waste of M0 = 120 m3, the heat capacities of H2O, HNO3, nitrates, and tank are given as follows:

  1. H2O:cWMWtb = 0.076×48,200×120= 4.40×105kJ C1

  2. HNO3:cNMNtb=0.110×2300×120=3.0 ×104kJ C1

  3. Nitrates: cS(104C)MS = 8.4×103kJC1 (MS=153×120kg) [see EquationEq. (13)]

  4. Tank: cT(104C)MT= 3.6×104kJC1 [see EquationEq. (14)]

  5. Total heat capacity: Γtb=5.14×105kJ C1,

where tb and 104°C are the time and the temperature at the beginning of boiling, respectively.

Since the contribution of the nitrates is only 1.6% before boiling, the effect of the error assuming that the heat capacity of the nitrate is equal to that of the oxide is negligible. At 300°C, at which the waste is almost dry and most of the nitrates convert to oxides, the contribution of the oxides to the total heat capacity becomes 19%, and so, the effect of the heat capacity estimation error of the nitrates (= oxides) is small.

VII. HEAT TRANSFER RATE FROM THE TANK IN THE INITIAL STAGE

The way of handling this part in the late stage is described in Sec. IX.

VII.A. Used Parameters and Values

There are three energy loss paths from the tank: convective heat transfer to the cell air, heat radiation to the cell wall, and gamma-ray release. The following parameters and values were used to determine the heat transfer rate of each path:

Parameter 1. The temperature of the tank is equal to that of the waste, and the initial value is θ(0) = 50°C.

Parameter 2. In the calculation of the convective heat transfer, the following assumptions were used:

a. The tank size is 7-m outside diameter × 4.0-m height, and its surface area ST = 165 m2. The thickness of the tank wall is 0.03 m. The heat transfer coefficients at the top and bottom of the tank are given by the same equation [EquationEq. (B.1)] as that of its vertical side[Citation17] (see Sec. B.II).

b. The inside cell size is an 8.1-m cube, and its surface area SC = 394 m2. The thickness of the concrete cell wall is = 2.0 m.

c. The cell air is heated by natural convective heat transfer on the tank outer wall and cooled on the inner cell wall. Since the heat capacity of air is small and the air is well mixed by convection, this heating and cooling can be regarded as balanced. The air temperature inside the cell, θai(t) (in degrees Celsius), is assumed to be uniform before and after the accident.

Parameter 3. Physical property values used for the calculation of the convective heat transfer are given in Appendix B.

VII.B. Equation for the Heat Transfer Rate

VII.B.1. Convective Heat Transfer

The heat transfer rate from the tank to air Rloss1(θ) is obtained by the following equation:

(15) Rloss1θ=αTSTθt θait,(15)

where αT = convective heat transfer coefficient of the tank wall (kW∙m−2∙°C−1), the details of which are described in Appendix B.

VII.B.2. Heat Radiation

The radiant heat transfer rate from the tank to the cell wall Rloss2(θ) is given by the following equation,[Citation18] which is expressed in absolute temperature θ + 273.15 (in kelvins) to avoid complications in display:

(16) Rloss2θ=εToσSTT4TIW4,(16)

where σ = Stefan-Boltzmann constant (5.67 × 10−11 kW∙m−2∙K−4); T = tank (= waste) temperature; TIW = temperature of the inner cell surface; εTo = emissivity of the tank outer surface[Citation19] of 0.4. The emissivity of concrete[Citation19] is 0.94. In this paper the emissivity larger than 0.9 was approximated as 1, that is, no reflection, in the calculation (see Sec. B.II).

VII.C. Solution Method of the Equations and Calculated Results

To solve EquationEqs. (15) and Equation(16), it is necessary to know the air temperature in the cell and the temperature of the inner cell surface. Since the heat capacity of air is small, the heat transferred from the tank to the air is soon transferred to the cell wall. Therefore, the following equation is obtained:

(17) αTSTθtθait=αCiSCθaitθCt,0,(17)

where αci = convective heat transfer coefficient of the inner cell wall surface (kW∙m−2∙°C−1) given by the same equation as EquationEq. (B.1) in Appendix B; θC(t,0) = inner surface temperature of the cell wall. Using EquationEq. (17), we get

(18) θait=αTSTθt+αciSCθCt,0αTST+αciSC.(18)

The inner surface temperature of the cell θC(t,0) is determined by the balance of the heat transfer from the gas, the heat radiation from the tank, the heat generation inside the cell by gamma rays, and the heat conduction to the inside of the cell wall. Therefore, it is necessary to find the temperature distribution inside the cell by solving the heat equation. The temperature distribution inside the concrete cell wall denoted by θC(t,x) is expressed by the following equation:

(19) θCt,x∂t=αC2θCt,xx2+qγt,xρCCC,(19)

where aC = thermal diffusivity (m2∙s−1); ρC = density (kg∙m−3); CC = specific heat of the cell concrete (kJ∙kg−1∙°C−1) (see Appendix B); qγ(t,x) = heat generation density due to the gamma heating (kW∙m−3) given by EquationEq. (C.1) in Appendix C.

The boundary conditions are

(20) x=0, t>0 : αciθaitθCt,0+Rloss2t,xSC=κCθC∂xx=0(20)(20)
(21)  x=L, t>0: κCθC∂xx=L=αcoθCt,Lθao+FOWσTCt,L4TOW4,(21)

where κC = thermal conductivity of the cell concrete (kW∙m−1∙°C−1); θao = air temperature outside the cell; αco = convective heat transfer coefficient of the outer cell wall surface (kW∙m−2∙°C−1) given by the same equation as EquationEq. (B.1); FOW = configuration factor between the outer cell surfaces and the building wall surfaces facing the outer cell surfaces. The second term on the right side of EquationEq. (21) is the net radiant heat transfer rate from the outer cell wall surfaces to their surrounding concrete building surfaces at TOW = 25 + 273.15 K. The outside air temperature θao was assumed to be constant at 25°C. FOW was set to 0.2 assuming a distance of 8 m between the two surfaces.[Citation19]

These boundary conditions were rearranged as follows in order to apply the numerical calculation algorithm shown in the literature.[Citation19] Substituting EquationEq. (18) into EquationEq. (20) to eliminate θai(t) and substituting EquationEq. (16) for Rloss2, the boundary condition at x = 0 can be rewritten as

(22) x=0, t>0: Λ1θθCt,0=κCθC∂xx=0(22)
Λl=αTαciSTαTST+αciSC+STSCεTσT2+TCt,02T+TCt,0,

where Λ1 = overall heat transfer coefficient including the effect of radiant heat transfer.

The boundary condition at x = L, EquationEq. (21), can also be rewritten as

(23) x=L, t>0: κcθCxx=L=ΛLTCt,LΘL(23)
ΛL=αco+σTCt,L+TowTCt,L2+TOW2ΛLΘL=αcoTao+σTCt,L+TowTCt,L2+TOW2.

EquationEquation (19) was numerically solved under the boundary conditions (22) and (23) and the initial condition, which will be explained later. The 2-m-thick wall was divided into 150 increments (Δx = 0.0133 m), and the time step Δt was set to be 109.7 s, which was derived from the condition that αCΔt/Δx2 = ½, which enables simplicity and stability of numerical integration.[Citation17]

The initial temperature distribution inside the cell wall is the distribution before the accident happens. It was determined as follows. Assuming that the cell ventilation air is continuously fed at 25°C (= θain in the following equation) and the temperature of the tank is maintained at 50°C, the temperature of the cell air θa is given by

θa=caFCaθain+STαTθT+SCαCθCcaFCa+STαT+SCαC,

where ca = molar specific heat of air (kJ∙mol−1∙°C−1); FCa = flow rate of cell ventilation air before the accident (mol∙s−1).

The boundary condition at x = 0 for the time before the accident is

(24) x =0, t>0: Λ0θθCt,0=κcθCxx=0,(24)
Λ0=αcicaFaθainθCt,0/θθCt,0+αciSTαTcaFca+STαT+SCαC+εTσSTT2+TCt,02T+TCt,0SC,

where Λ0 = overall heat transfer coefficient for the time before the accident.

Using EquationEqs. (21), (Equation22), and (Equation24), EquationEq. (19) was solved numerically. Starting with an initial uniform distribution at 35°C, a steady-state distribution before the accident was obtained after a sufficient number of time steps (about 400 h). The temperatures of the cell air, the inside surface of the cell wall, and the outside surface of the cell wall are 34°C, 39°C, and 32°C, respectively. This temperature distribution is used as an initial condition for the waste temperature calculation after the accident.

VIII. CALCULATION OF THE HEAT BALANCE EQUATION IN THE INITIAL STAGE

The time change of the waste temperature is obtained by integrating EquationEq. (1) over time. The amounts of H2O and HNO3, the heat absorption rates due to the liquid vaporization and the nitrate decomposition, and the heat transfer rates are already described in the previous sections as a function of time.

Substituting EquationEqs. (6) and Equation(8) into EquationEq. (1), the following equation is obtained:

(25) Γt+λWMW0dξW/+λNMN0dξN/dt=RhRstRlossRγ,(25)

where Rγ = leak rate of gamma-ray energy out of the tank given by EquationEq. (C.1) shown in Sec. C.II.

Rearranging EquationEq. (25), we get

(26) dt=RhRstRlossθRγtΓt+λWMW0dξWθ/+λNMN0dξNθ/.(26)

Here, variables whose values change with time are expressed as a function of time.

Denoting a time step by Δt, nΔt by tn, and θ(tn) by θn and approximating dθ/dt = (θn+1 − θn)/Δt and the value of the right side of EquationEq. (26) to be equal to the value at t = tn, we get

θn+1=θn+RhRstnRlossθnRγtnΓtn+λWMW0dξWθ/t=tn+λNMN0dξNθ/t=tnΔt.

By repeating this operation, θ can be determined as a function of time.

IX. DERIVATION OF THE HEAT BALANCE EQUATION AND ITS SOLUTION METHOD IN THE LATE STAGE

IX.A. Heat Balance Equation

IX.A.1. Description of the Model

shows the modeled temperature distribution inside and outside the tank. The tank is modeled as a cylinder with a diameter of 7 m, a height of 4 m, and a wall thickness of 0.03 m. The thickness of the debris is 0.32 m (see Sec. C.II). The tank wall is separated into the following two parts. One contacts the debris (bottom and a small portion of the side wall of the tank), and the other does not contact the debris (top and most of the side wall), which are named as the tank bottom and the tank top/side, respectively. The debris is evenly divided into three parts: the upper, the middle, and the bottom. The debris volume was assumed to be a constant 12 m3 above 130°C.

Fig. 9. Parameters on temperature θ and on surface area S around the tank in the late stage.

Fig. 9. Parameters on temperature θ and on surface area S around the tank in the late stage.

The temperature of each part is displayed as follows. The temperatures of the tank bottom and the tank top/side are θTb(t) and θTt(t), respectively. The temperatures of each piece of the debris are denoted as θdu(t), θdm(t), and θdb(t) assuming that the temperature of each part is uniform. We considered θdb(t) = θTb(t) because the heat transfer from the bottom debris to the tank is much larger than that from the tank to the cell. Θg(t) denotes the gas temperature in the tank, and θai(t) denotes the inner cell air temperature. For simplicity, the cell wall temperature θc(t,x) is assumed to be the same regardless of the cell’s top, bottom, left, and right positions.

The surface area of each part is displayed as follows and shown in . S1 is the surface area of each piece of the debris, which is equal to S2 of the tank top. S3 is the surface area of the tank side without contacting the debris. S4 is the area of the side surface of the debris (S4/S3 is the area of the side surface of each piece of the debris). Thus, S1 = S2 = 38.5 m2, S3 = 80.9 m2, and S4 = 7.0 m2. In deriving S3 and S4, the debris thickness of 0.32 m was used. SC is the area of the inner cell surface of 394 m2.

The mass of the tank is 39 200 kg, and that of the structural material inside the tank is 30 800 (= 70 000 – 39 200) kg. Assuming that the structural material was distributed uniformly according to the longitudinal length of the tank, the mass of the tank bottom is 13 300 (= 10 800 + 2500) kg, that of the side part is 47 500 (19 200 + 28 300) kg, and that of the top is 9140 (9140 + 0) kg (the latter figure in parentheses is the contribution of the structural materials).

IX.A.2. Heat Balance Equation for Each Part Inside and Outside the Tank

IX.A.2.a. Equation for Each Piece of the Debris

Bottom Debris

It was assumed that the temperature of the tank bottom wall is equal to that of the bottom debris θdb (in degrees Celsius). Therefore, the heat balance equation that combines the bottom debris and the tank bottom is given as follows:

(27) Cdb+CTbdθdbdt=RhbRSbλWbYWb+λNbYNb+QbRlossb,(27)

where Cdb and CTb = heat capacities of the bottom debris and the tank bottom (kJ∙K−1); Rhb = heat generation rate due to the decay heat in the bottom debris (kW); RSb = heat absorption rate due to the salt (nitrate) decomposition (kW); λWb and λNb = latent heats of vaporization of H2O and HNO3 (kJ∙kg1); YWb and YNb = their vaporization rates (kg∙s−1); Qb = heat input to the bottom debris from its surroundings (kW); Rlossb = sum of the convective heat transfer from the tank bottom to the cell air and the heat radiation to the cell wall (kW). The equation for Qb is given as follows:

Qbt=S1Kdθdmθdb+S4/3Kdθdmθdb+S4/3KdθduθdbSssKssθdbθTt,

where Kd = heat transfer coefficient between two pieces of the debris in contact (W∙m−2∙K−1); Sss = sum of the cross-sectional area of the tank side wall and that of the internal structure (m2); Kss = transfer coefficient between the tank bottom part contacting with the debris to the tank upper part not contacting with the debris (kW∙m−2∙°C−1). The equation for Rlossb is given as follows:

Rlossbt=S1+S4[αlθTtθai+εToσTTb4TCt,04,

where α1 = convective heat transfer coefficient between the tank bottom and the inner cell air; σ = Stefan-Boltzmann constant; εTo = emissivity of the tank outer surface. It is assumed that the proportion of the heat radiation from the cell wall reaching the outer surface of the tank is equal to the surface area ratio. Although S4 is not the bottom surface, it has the same temperature as the bottom surface, so it is included in the equation.

Using the evaporation rates YWb and YNb given in EquationEqs. (6) and Equation(8), EquationEq. (27) can be transformed as follows:

(28) dθdbdt=RhbRSb+QbRlossbCdb+CTb+λWMWb0dξWb/dθdb+λNMNb0dξNb/dθdb.(28)

In this section, suffixes u, m, and b denote the upper, middle, and bottom debris, respectively.

Middle Debris

Similar to EquationEq. (28), the temperature change of the middle debris θdm is derived as follows:

(29) dθdmdt=RhmRSm+QmCdm+λWMWm0dξWm/dθdm+λNMNm0dξNm/dθdm(29)

Qmt=SlKdθdmθdbSlKdθdmθdbS4/3Kdθdmθdb,

where Qm = heat input to the middle debris from its surroundings (kW).

Upper Debris

Similarly, the temperature change of the upper debris θdu is derived as follows:

(30) dθdudt=RhuRSuQuRradCdu+λWMWu0dξWu/dθdu+λNMNu0dξNu/dθdu(30)

Qut=S1KdθdmθduS1αdgθduθgS4/3Kdθduθdb,

where Qu = heat input to the upper debris from its surroundings (kW); Rrad = net radiation heat transfer rate from the upper debris surface to the tank inner surface and is expressed as follows:

Rrad=F12+F13SlσTdu4F21S2+F31S3σTTu4.

In the equation, the emissivity of the upper surface of the debris and the inner surface of the tank are both treated as 1 (see Sec. B.II). F12, F13, F21, and F31 are the configuration factors (suffixes 1, 2, and 3 correspond to the upper debris surface, the upper tank surface, and the side tank surface), and their values are 0.365, 0.635, 0.365, and 0.302, respectively.

IX.A.2.b. Equation for the Gas in the Tank

The gaseous components are H2O vapor, HNO3 vapor, NO2, NO, and O2. The gas temperature θg is described by the following heat balance equation:

Qgcgθdu+S1αdgθduθg=S2+S3αg2θgθTu+Qgcgθg,

where Qg = sum of the gas generation rates from three pieces of the debris (mol∙s−1), each rate being given by equations similar to EquationEqs. (6), Equation(8), Equation(11), and Equation(12); cg = its molar heat capacity; αdg = convective heat transfer coefficient between the upper debris surface and the tank gas; αg2 = convective heat transfer coefficient between the tank top/side surface and the tank gas; ca = specific heat of the air. The first term on the left side of the above equation is the carry-on enthalpy of the gas generated from the debris, and the temperature of the generated gas leaving the debris is assumed to be equal to the temperature of the upper debris. The second term is the heat transfer rate from the debris to the gas in the tank. The first term on the right side is the heat transfer rate from the gas to the top/side wall of the tank, and the second term is the take-out enthalpy of the gas flowing out of the tank.

From the above equation, the gas temperature is described as follows:

(31) θg=Qgcgθdu+S1αdgθdu+S2+S3αg2θTuQgcg+S1αdg+S2+S3αg2.(31)

IX.A.2.c. Equation for the Tank Top/Side Wall

The heat source of the top/side wall is the convective heat transfer from the gas in the tank (the first term on the right side of the following equation), the heat radiation Rrad between the upper debris and the top/side wall, and the conductive heat transfer from the bottom wall to the side wall (the third term on the right side of the following equation). The heat sink Rlosst is the heat release rate from the tank to the cell. Therefore, the following equation is obtained:

CTtdθTtdt=S2+S3αg2θgθTt+Rrad+SSSαSSθdbθTtRlosst

Rlosst=S2+S3α2θTtθai+S2+S3εToσTTt4TCt,04,

where CTt = heat capacity of the tank top/side wall.

IX.A.2.d. Equation for the Cell Air

The following equation is obtained from the conditions described in parameter 2 in Sec. VII.A:

α1S1+S4θTbtθait+α2S2+S3θTutθait=αciSCθaitθCt,0,

where α2 = convective heat transfer coefficient between the tank upper/side and the inner cell air; αci = inner cell wall and the inner cell air. From this, we obtain the following equation, which gives the cell air temperature:

(32) θait=α1S1+S4θ1t+α2S2+S3θ2t+αciSCθCt,0α1S1+S4+α2S2+S3+αciSC.(32)

IX.A.3. Examination of Each Parameter Value in the Heat Balance Equation

IX.A.3.a. Heat Generation Rate Rh of Each Piece of the Debris

The equation

Rhb=Rhm=Rhu= 57810/3=189kW,

where the second suffixes b, m, and u of Rh correspond to the bottom, middle, and upper debris, respectively. In the equation, 10 kW is the gamma-ray release out of the tank [see EquationEq. (C.1)].

IX.A.3.b. Heat Transfer Coefficient Between Two Pieces of the Debris in Contact

The coefficient Kd (in kW∙m−2∙°C−1) is given as follows:

Kd=κd/0.32/3 = 0.0038,

where κd = effective thermal conductivity of debris given as 0.00037 at 150°C to 270°C and 0.00040 at 400°C.[Citation20] We adopted 0.00040 over the entire temperature range.

IX.A.3.c. Heat Transfer Coefficient Between the Tank Bottom to the Tank Top/Side

The coefficient Kss is obtained as follows. If the thermal diffusivity of the material is a (in m2∙s−1) and if the distance is d, the timescale τ (in seconds) required for heat conduction over the distance d (in meters) is given by τd2/a. The duration of the accident targeted in this study is 100 to 200 h, so the timescale of the temperature change would be about 10 h. Using the thermal diffusivity = 4.1 × 10−6 m2∙s−1 of stainless steel (SUS), it is considered appropriate to set the distance scale for heat conduction in SUS to = 0.4 m. Therefore, using the thermal conductivity of SUS, which is 0.0165 kW∙m−1∙K−1,[Citation18] the heat transfer coefficient is calculated as Kss = 0.0165/0.4 = 0.040 kW∙m−2∙°C−1. The sum of the cross-sectional area of the tank side wall and that of the internal structure is 0.66 + 0.98 = 1.64 m2. When the temperature difference between θdb and θTt is 200 K, the heat transfer rate from the bottom surface to the side surface is 1.64 × 0.040 × 200 = 13 kW, which is a small value in the overall heat balance of 568 kW (heat generation inside the debris). Therefore, the value of Kss considered in this paper has little effect.

IX.A.3.d. Convective Heat Transfer Coefficient and Emissivity

As is explained in Sec. B.II, EquationEq. (B.1) can be used for all coefficients, αci, αco, α1, α2, αdg, and αg2.

Emissivities, εTi, εTo, εd, and εc, are also treated in Sec. B.II.

IX.A.3.e. Heat Capacity

The heat capacity of debris Cd is given by EquationEq. (13). The heat capacity of each piece of the debris is Cd/3.

The specific heat cT of SUS-316, which makes up the tank and the internal structural material, is given in the literature.[Citation18] Therefore, the heat capacities of the tank bottom CTb and its top/side CTt (J∙K−1) are expressed as follows:

CTb=cT×13,300
CTt=cT×9100+47,500 =cT×56,600.

The temperature-dependent molar heat capacity of each gas (NO2, NO, O2, and H2O) composing the gas in the tank was obtained from CitationRef. 15. The molar heat capacity of HNO3 vapor was assumed to be the same as that of H2O. The heat capacities of the nitrates (= oxides) and tank (including the internal structural material) are given by EquationEqs. (13) and (Equation14), respectively.

IX.B. Solution Method

IX.B.1. Temperature Inside the Concrete Cell Wall

To obtain a numerical solution from EquationEq. (19) in the late stage, it is necessary to know θC(t,x) at td when the waste temperature reaches 130°C. According to the results of the initial stage, td = 4.64 × 105 s (129 h). Therefore, the various parameters in EquationEq. (19) at = td are obtained from the results at the initial stage (see Sec. X.A).

The boundary condition at = 0 for t > td can be obtained using the fact that the heat flux into the wall due to convection and radiation on the inner wall surface is equal to the conductive heat flux inside the wall:

x= 0,
(33) αCiθaitθCt,0+S1+S4SCεsoσT1t4+S2+S3SCεsoσT2t4S1+S2+S3+S4SCεsoσTCt4=κcTCxx=0.(33)(33)

The temperature of the cell air θAi is given by EquationEq. (32). When obtaining the numerical solution of EquationEq. (33), it was transformed into the following equation:

Λ2ΘθCt,0=κcθCxx=0.

In the above equation, Λ2 and Θ are given as follows:

Λ2=αciα1S1+S4+α2S2+S3α1S1+S4+α2S2+S3+αciSC+S1+S2+S3+S4SCεsoσTCt,03
Θ=αciΛ2α1S1+S4T1+α2S2+S3T2α1S1+S4+α2S2+S3+αciSC+εsoσΛSCS1+S4T14+S2+S3T24.

The boundary condition at L for t > td is the same as that in the initial stage given by EquationEq. (21).

IX.B.2. Calculation of the Heat Balance Equation

The temperatures that need to be obtained in the late stage are the temperatures of the bottom, middle, and upper debris, the temperatures of the tank top/side wall and the bottom wall (the bottom temperature is assumed to be equal to the bottom debris), the gas temperature in the tank, the air temperature in the cell, and the temperature distribution inside the cell wall. Time changes in the debris temperatures and the tank temperatures are obtained by solving numerically the heat balance equations, EquationEqs. (28) through (Equation31), with the values at time td obtained in the initial stage as the initial conditions. The gas temperature and the air temperature can be calculated using the above results and EquationEqs. (31) and (Equation32). The surface temperature of the cell wall is required for radiant heat calculation. This is obtained by numerically solving EquationEq. (19) under the boundary conditions EquationEqs. (33) and (Equation21) at each time step and calculating the temperature distribution of the cell wall. The numerical calculation formulas are the same as those of the initial stage model except for the boundary condition at = 0, and the time step value is also the same (see Sec. VII.C).

X. CALCULATED RESULTS AND DISCUSSION OVER BOTH STAGES

X.A. Transition of the Accident with Time

shows the temperatures of the waste, tank walls (tank top/side and tank bottom, both including the internal structure), tank gas, cell air, and inner and outer cell surfaces as a function of time, obtained by solving the heat balance equations. The temperature of the tank bottom is the same as that of the bottom debris. Elapsed time 0 h corresponds to the time when the cooling function loss occurred. Boiling begins after 14 h. Because Ru volatilization is reported[Citation1] to begin when the waste temperature reaches 119°C, it starts after 111 h. The waste reaches 130°C after 129 h (= td), when the model has changed from the initial stage to the late stage. The temperatures of the upper and bottom debris rise at almost the same rate, but that of the middle debris increases quickly. When those of the upper and bottom debris reach 600°C, the middle debris exceeds 1200°C. This quick temperature increase is explained by the low effective thermal conductivity of the debris (see Sec. IX.A.3.b) and the sandwich effect between the upper and bottom debris. The heat of the upper debris is transferred to the upper part of the tank, and the heat of the lower debris is transferred to the tank bottom. At time td, it can be seen that the temperature of the tank side temporarily drops. This is explained by the abrupt model change, i.e., from the homogeneous liquid state to the heterogeneous dry state.

Fig. 10. Temperatures of the waste, tank walls (tank top/side and the tank bottom), tank gas, cell air, and cell inside and outside wall surfaces versus time curves (temperature of the tank bottom is the same as that of the bottom debris). At time td = 129 h, HLLW temperature reached 130°C, and the stage was changed from the initial stage to the late stage.

Fig. 10. Temperatures of the waste, tank walls (tank top/side and the tank bottom), tank gas, cell air, and cell inside and outside wall surfaces versus time curves (temperature of the tank bottom is the same as that of the bottom debris). At time td = 129 h, HLLW temperature reached 130°C, and the stage was changed from the initial stage to the late stage.

When the bottom debris reaches 180°C after 142 h, the inner cell surface reaches 81°C, and when it reaches 600°C after 456 h, the surface reaches 482°C. The outer cell surface is assumed to be in contact with air at 25°C, and the outer surface temperature increases slowly and reaches 62°C when the bottom debris reaches 600°C.

shows the NO2 and NO generation rates as a function of time, which was calculated by the Arrhenius-type model explained in Sec. V.B. The bottom debris temperature is shown for reference. A small amount of NO2 is generated when boiling begins, but most of it is generated at temperatures of the debris between 150°C and 350°C. The NO generation curve shows several peaks. This is mainly because the temperature of the middle debris is much higher than those of the upper and bottom debris.

Fig. 11. NO2 and NO generation rates and the bottom debris temperature (dotted line for reference) versus time curves.

Fig. 11. NO2 and NO generation rates and the bottom debris temperature (dotted line for reference) versus time curves.

shows the heat absorption rates due to the liquid vaporization and the nitrate decomposition, the heat transfer rate from the tank to the cell, and the sensible heat accumulation rate inside the tank, all as a function of time. The sum of these heat absorption terms should be equal to the value obtained by subtracting the energy release rate of leaked gamma rays from 578 kW due to the decay of radioactive materials. Abrupt discontinuous changes of vaporization and heat accumulation curves at around 129 h are due to the model change from the initial stage to the late stage. The latent heat of vaporization is the main heat absorption source until 142 h (180°C bottom debris temperature) when liquid evaporation from the waste is almost gone, and then, the heat absorption source of nitrate decomposition becomes main until 167 h (280°C bottom debris temperature). After that, the heat transfer to the cell wall becomes the main heat absorption source. The sensible heat accumulation rate curve shows two peaks at 130 to 150 h and 155 to 190 h, and a trough around 150 h. This phenomenon is explained by the fact that much of the decay heat is spent on nitrate decomposition at around 150 h, as can be seen in the figure.

Fig. 12. Vaporization heat absorption rate, nitrate decomposition heat absorption rate, heat transfer rate from the tank, sensible heat accumulation rate in the tank versus time curves (up to 129 h, the waste remains in the liquid state).

Fig. 12. Vaporization heat absorption rate, nitrate decomposition heat absorption rate, heat transfer rate from the tank, sensible heat accumulation rate in the tank versus time curves (up to 129 h, the waste remains in the liquid state).

shows the temperature increasing rate of each debris as a function of time. For reference, the temperature of the bottom debris is indicated by a dotted line. Before boiling, the rate of the liquid waste is 0.06°C∙min−1. After boiling, it decreases to 0.0012°C∙min−1 and then increases gradually as the concentration progresses. When the time exceeds 129 h (130°C), the increasing rates of the upper and bottom debris come within the range of 0.02 to 0.2°C∙min−1 and that of the middle debris within 0.02 to 1.0°C∙min−1. Each rate curve shows two peaks. The rise of the first peak is explained by the substantial end of the vaporization of H2O and HNO3. Then, since the nitrate decomposition generating NO2 begins, the peak decreases forming the first peak. The second peak is formed by the completion of NO2 generation and the increased heat loss mainly due to heat radiation. Humps seen around 200°C are caused by NO generation. After that, the debris temperatures increase under the balance of the internal heat generation and the heat loss through heat radiation, conduction, convection, and sensible heat accumulation.

Fig. 13. Temperature increasing rates of the individual debris and the temperature of the bottom debris (dotted line for reference) versus time curves (up to 129 h, the waste remains in the liquid state).

Fig. 13. Temperature increasing rates of the individual debris and the temperature of the bottom debris (dotted line for reference) versus time curves (up to 129 h, the waste remains in the liquid state).

shows the changes in the released gas composition and the total gas flow rate corresponding to the bottom debris temperature. At the initial stage between 104°C and 130°C, the major component of the gas is H2O vapor. Though it decreases with the increasing temperature, its percentage is still 70% at 180°C, when the waste seems almost solid with a slight nitric acid smell.[Citation9] The figure shows that even at 300°C, the HNO3 vapor is contained in the gas by 17% although the H2O vapor is zero. It should be noted that in this temperature region, these errors in vapor amount can be large. However, the errors do not affect the waste temperature rise because of their small heat absorption rates (see ). Gas flow above 400°C consists of NO and O2 produced by the decomposition of nitrates.

Fig. 14. Changes in the released gas composition and the total gas flow rate corresponding to the bottom debris temperature.

Fig. 14. Changes in the released gas composition and the total gas flow rate corresponding to the bottom debris temperature.

In order to examine the validity of the present model, we attempted to apply it to the experimental results of other papers. However, no comparable results could be found.

X.B. Application of the Results to a Postulated Accident of a Plant

The above results were obtained using the initial waste volume of 120 m3, the decay power density of 4.82 kW∙m−3, and the composition of the initial waste given in . These values will change from time to time in a reprocessing plant. However, information necessary to solve EquationEq. (1) is the decay power density; the initial waste volume; and the initial amounts of H2O, HNO3, and nitrates. Such information will be obtained from the daily HLLW management. Accordingly, the waste temperature and the release rates of the components transferred to the gas phase are predictable as a function of time after the cooling loss using the HLLW conditions in a plant.

The ultimate goal of our study is to predict the release behavior of radioactive aerosol and volatile Ru into the outside atmosphere under accidental conditions. The results obtained above can be used to study the LPF[Citation11] of radioactive materials, especially volatilized Ru, because the Ru and NO2 are absorbed into the condensate of vaporized H2O and HNO3 formed in the release path. Since aerosol release will cease by 130°C[Citation8] and volatile Ru by 300°C,[Citation6] the experimental results up to 300°C shown in this section are important.

In the present study, since Excel calculation was used to obtain the solution, some assumptions and approximations were used for simplicity. For example, although the waste changes gradually from liquid phase to solid phase, it was assumed to change suddenly at 130°C. Although the solid phase was continuous, it was divided into three pieces, and the temperature of each piece was made uniform. The solid-phase thermal conductivity will vary with temperature due to the presence of some residual liquid, nitrate decomposition, and volume change, but it was approximated to be constant. In the future, the use of more a sophisticated modeling and numerical solution will bring these assumptions and approximations closer to reality, enabling more accurate predictions.

XI. CONCLUSIONS

We proposed a method for systematically predicting temperatures of the waste, tank, and cell concrete up to 600°C and individual gas release rates as a function of time under accidental conditions. This method would be applicable to HLLW with different decay power densities, volumes, and compositions.

The results obtained in this study are useful not only for examining the LPF but also for developing measures to reduce it. Further, they will become the basic information to examine the accident management measures and to provide the disaster prevention system since the release of radioactive materials and their carrier gases can be quantitatively predicted as a function of time.

Acknowledgments

The analysis using the MCNP code was performed by Yuri Tajimi, Daisuke Takano, and Tami Mukohara of TEPCO Systems Corporation. The authors are very grateful for their contributions.

Disclosure Statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been corrected with minor changes. These changes do not impact the academic content of the article.

Notes

a The liquid waste changes from a solution state to a dry state with increasing temperature. Hereinafter, “waste” covers both states.

b By 600°C, the generation of vapors and gases that transports radioactive materials outside the building will cease. Further, since it takes more than 10 days to reach 600°C as shown in Sec. X, it is possible to take measures during this time such as introducing water into the cell and submerging the storage tank.

c Sections III, VII, and VIII are applicable to the initial stage, and Sec. IX is applicable to the late stage. Sections IV, V, VI, X, XI, and the Appendixes are applicable to both stages.

d Since H2O and HNO3 are vaporized from nitric acid solution, the interaction between H2O and HNO3 needs to be taken into account. The influence of this effect is explained in Sec. IV.C.3.

e The tank is made of SUS with a thickness of 0.03 m. The thermal conductivity of SUS is 17 W∙m−1∙K−1, and the heat transfer coefficient through the tank wall is 17/0.03 = 570 W∙m−2∙K−1, which is about 100 times larger than the natural convection heat transfer coefficient (2 to 5 W∙m−2∙K−1), and therefore, the heat transfer resistance from the tank to the cell air is controlled by natural convection.

f The meaning of this equal sign is explained in Sec. III.A, No. 5.

g 2440 is obtained from the values in .

h The shape of the top and bottom is slightly mortar-like rather than flat (see ).

i A debris thickness becomes 0.32 m.

References

  • M. PHILIPPE, J. P. MERCIER, and J. P. GUÉ, “Behavior of Ruthenium in the Case of Shutdown of the Cooling System of HLLW Storage Tanks,” Proc. 21st DOE/NRC Nuclear Air Cleaning Conf., NUREG/CP-0116, Vol. 2, p. 831, U.S. Nuclear Regulatory Commission (1991).
  • “Study of the Behavior of Released Radioactive Materials in a Reprocessing Plant,” Management Group on the Study of the Behavior of Released Radioactive Materials (2014) (in Japanese) (available at National Diet Library).
  • K. YOSHIDA and J. ISHIKAWA, “Thermal-Hydraulic Analysis of Boiling Event of Reprocessed Liquid Wastes with MELCOR Code,” JAEA-Research 2012-026, Japan Atomic Energy Agency (2012) (in Japanese).
  • J. ISHIKAWA, K. YOSHIDA, and K. YOSHIDA, “Development of Simulation Tool for Boiling Event of Reprocessed Radioactive Liquid Waste,” Trans. At. Energy Soc. Jpn., 12, 2, 165 (2013); https://doi.org/10.3327/taesj.J12.025 (in Japanese).
  • H. ABE et al., “Investigation of Release Behavior of Volatile Ruthenium Species from Thermal Decomposition of Ruthenium Nitrosyl Nitrate,” JAEA-Research 2014-022, Japan Atomic Energy Agency (2014) (in Japanese).
  • S. TASHIRO et al., “Release Characteristics of Ruthenium from Highly Active Liquid Waste in Drying Step,” Trans. At. Energy Soc. Jpn., 14, 4, 227 (2015); https://doi.org/10.3327/taesj.J14.028 (in Japanese).
  • K. YOSHIDA, J. ISHIKAWA, and H. ABE, “Analysis of Release and Transport of Aerial Radioactive Materials in Accident of Evaporation to Dryness Caused by Boiling of Reprocessed High-Level Liquid Waste,” Trans. At. Energy Soc. Jpn., 14, 4, 213 (2015); https://doi.org/10.3327/taesj.J15.002 (in Japanese).
  • S. TASHIRO et al., “Release of Radioactive Materials from Simulated High-Level Liquid Waste at Boiling Accident in Reprocessing Plant,” Nucl. Technol., 190, 2, 207 (2015); https://doi.org/10.13182/NT14-57.
  • T. KODAMA et al., “Experiments on the Leak Path Factor of Ruthenium Volatilized from High-Level Liquid Waste Tanks in a Reprocessing Plant in Case of the Boiling and Drying Accident,” J. Nucl. Sci. Technol., 52, 467 (2015).
  • T. KODAMA et al., “Boiling and Drying Accident of High-Level Liquid Waste in a Reprocessing Plant: Examination of the NO2 and NO Generation Using the Simulated Waste,” J. Nucl. Sci. Technol., 57, 9, 1101 (2020); https://doi.org/10.1080/00223131.2020.1764405.
  • SCIENCE APPLICATIONS INTERNATIONAL CORPORATION, “Nuclear Fuel Cycle Facility Accident Analysis Handbook,” NUREG/CR-6410, U.S. Nuclear Regulatory Commission (1998).
  • R. O. GAUNTT et al., “MELCOR Computer Code Manuals, Vol. 2: Reference Manuals, Version 1.8.5 May 2000,” NUREG/CR-6119, Vol. 2, Rev. 2, SAND2000-2417/2, Sandia National Laboratories (2000).
  • CHEMICAL SOCIETY OF JAPAN, Handbook of Chemistry: Pure Chemistry, 5th ed., Maruzen Publishing Company (2004) (in Japanese).
  • J. B. JOSHI, V. V. MAHAJANI, and V. A. JUVEKAR, “Invited Review: Absorption of NOx Gases,” Chem. Eng. Commun., 33, 1, 1 (1985); https://doi.org/10.1080/00986448508911161.
  • CRC Handbook of Chemistry and Physics, 51st ed. (2010).
  • SOCIETY OF CHEMICAL ENGINEERS, JAPAN, “Chemical Engineering Handbook,” 6th ed., Maruzen Publishing Company (1999).
  • Handbook on Process and Chemistry of Nuclear Fuel Reprocessing, 3rd ed., JAEA-Review 2015-002, Japan Atomic Energy Agency (2015) (in Japanese).
  • JAPAN SOCIETY OF MECHANICAL ENGINEERS, JSME Data Book: Heat Transfer, 5th ed., Maruzen Publishing Company (2017) (in Japanese).
  • Heat Transfer Engineering, N. SEKI Ed., Morikita Publishing Company (1988) (in Japanese).
  • “2019 Report: Tests on the Migration Behavior of Radioactive Materials in Reprocessing Facilities,” Japan Atomic Energy Agency; https://www.nra.go.jp/data/000317612.pdf (in Japanese).
  • K. WIECZOREK-CIUROWA and A. J. KOZAK, “The Thermal Decomposition of Fe(NO3)3∙9H2O,” J. Thermal Analysis and Colorimetry, 58, 3, 647 (1999); https://doi.org/10.1023/A:1010112814013.
  • K. KAWAMURA et al., “Thermal Stress Analysis of Concrete in Fire,” Technical Report No. 44, p. 39, Taisei Advanced Center of Technology (2011) (in Japanese).
  • X-5 MONTE CARLO TEAM, “MCNP—A General Monte Carlo N-Particle Transport Code, Version 5, LA-UR-03-1987”, Los Alamos National Laboratory (2003).

APPENDIX A

HEAT OF HNO3 GENERATION FROM NITRATES IN THE WASTE

It is reported that Fe nitrate decomposes and releases HNO3 by the following reaction during the temperature increase process[Citation21]:

FeNO33+3/2H2O1/2Fe2O3+3HNO3.

The standard formation enthalpy of Fe(NO3)3 was estimated from that of the chloride compound using EquationEq. (10), and those of the other compounds were cited from the literature.[Citation16] The heat of the reaction is thus obtained as follows:

ΔH=824/2+3×174399×1.17+286×1.5=38kJmol1 .

Therefore, the reaction is exothermic, and the calorific value of HNO3 generation is 13 kJ∙mol−1.

A previous report[Citation10] indicates that 30% of the N in nitrates shown in is released as HNO3 instead of NOx. If the calorific value per mol of HNO3 generation from the nitrates is assumed to be equal to that from Fe nitrate, the calorific value of HNO3 generation from the nitrates in 1 m3 of waste becomes 9500 kJ∙m−3 (= 2440 × 0.30 × 13).Footnoteg Assuming that this exothermic reaction occurred during 100 h beginning from boiling, the exothermic density was as small as 0.026 kW∙m−3, so the heat of HNO3 generation was not considered in the calculation of the heat of NOx generation.

APPENDIX B

PHYSICAL PROPERTY VALUES USED FOR CONVECTIVE AND RADIATIVE HEAT TRANSFER CALCULATIONS

B.I. VALUES RELATED TO THE CELL WALL

The wall material was assumed to be solely concrete, and its physical property values used here are the following:

  1. Thermal conductivity[Citation22] κc at 25°C = 0.00175 kW∙m−1∙°C−1.

  2. Specific heat[Citation18] cc at 25°C = 0.90 kJ∙kg−1∙°C−1.

  3. Density[Citation18] ρc at 25°C = 2400 kg∙m−3.

The thermal diffusivity is ac = κccc−1ρc−1 = 8.1 × 10−7 m2∙s−1. The cell wall is made of reinforced concrete, but it is difficult to estimate the overall thermal properties. Concrete at high temperatures would deteriorate resulting from decomposition of the components or thermal deformation. Analysis of these phenomena is beyond the scope of this paper, and we used the above values as constant for calculation of the wall temperature.

B.II. CONVECTIVE HEAT TRANSFER COEFFICIENT

Convective heat transfer occurs between the tank outer surface and the cell air. Since the length of the tank is as long as several meters, the Rayleigh number becomes 109 or more even if the temperature difference is 1°C, and so, the convection becomes turbulent.[Citation19] Although the heat transfer coefficient differs slightly depending on the direction of the surface (top, side, and bottomFootnoteh ), the difference is small.[Citation19] The following equation of the turbulent natural convection heat transfer coefficient along the vertical plate and the surrounding air[Citation19] was used for all points, α1 and α2. Assuming that the transport properties of the gas in the tank are the same as those of air, the equation is applicable to the tank gas, αdg, and αg2:

(B.1) αT=0.13θθaνaaa13κa,(B.1)

where β = coefficient of thermal expansion of air expressed in (θ + 273.15)−1 (°C−1); νa = kinematic viscosity (m2∙s−1); aa = thermal diffusivity (m2∙s−1); κa = thermal conductivity of air (kW∙m−1∙°C−1); g = gravitational acceleration (m∙s−2). The value of κa was calculated from the following approximate equation, which was derived using the literature values[Citation15]:

κa=3.26×107θai+273.150.773.

The values of νa, aa, and κa used are those at the average temperature of the tank and the cell air. Since the tank temperature changes from 50°C to 600°C, the parameters are also affected by temperature. Using the temperature dependence of νa and aa obtained from the literature,[Citation18] the following equation (in s2∙°C−1∙m−4) was obtained:

β/νaαa=4.66×1017/θTa+273.154.33,

where θTa = average temperature of the tank and the cell air.

Similarly, the heat transfer between the cell air and the cell wall surface is dominated by turbulent natural convection, and the heat transfer coefficient can be handled in the same manner as described above. Therefore, EquationEq. (B.1) is applicable to αci and αco.

B.III. EMISSIVITIES USED FOR THE CALCULATION OF HEAT RADIATION

The following emissivities were assumed from the literature values:

  1. 0.92 for the tank inner surface εTi (assumed from SUS oxidized black surface values).[Citation19]

  2. 0.4 for the tank outer surface εTo (SUS brown surface value due to heating).[Citation19]

  3. 0.98 for the upper debris surface εd (assumed to be equal to that of black leather).[Citation18]

  4. 0.94 for the cell concrete surface εc.[Citation19]

Emissivities above 0.9 were approximated as 1 in the calculations. Therefore, except for the reflection from the outer surface of the tank, the reflection from the other three surfaces was not considered.

APPENDIX C

CALCULATION OF GAMMA-RAY LEAKAGE RATE OUT OF THE TANK AND CORRESPONDING GAMMA HEATING OF THE CELL

The content in this section can be used for both stages.

Of the radiation generated by the decay of radionuclides, all alpha rays and beta rays are absorbed in the tank, but some of the gamma rays leak out of the tank. The gamma-ray leakage rate to the total decay power and the corresponding gamma heating of the concrete cell were examined.

C.I. CALCULATION CONDITIONS

The tank and the cell sizes are given in Sec. VII.A. The amount and the composition of the initial waste are given in Sec. III. According to the progress of the accident, the waste volume changes from 120 to 12 m3. 12 m3 corresponds to the debris (dried waste) volumeFootnotei assuming its density is 1.5 g∙cm−3, which was obtained from the experiment shown in Sec. IV.A. Therefore, the liquid volume in the waste changes from 120 to 0 m3.

In order to calculate the gamma-ray leakage rate and the gamma heating of the cell, the Monte Carlo code MCNP5-1.60 and library MCPLIB84 were used.[Citation23] The energy and generated number of gamma rays + X-rays per unit time were obtained from ORIGEN-2. Bremsstrahlung X-ray was taken into account assuming the surrounding medium was UO2 as the maximum condition, but its influence on the results was less than 3%. Therefore, hereinafter, we do not consider X-rays.

In calculating the gamma heating of the cell wall, we did not distinguish a ceiling, a floor, and four side walls. Therefore, the gamma heating in the depth direction of the cell walls was the same regardless of the position of each wall.

C.II. GAMMA-RAY LEAKAGE RATE FROM THE TANK

Before the start of boiling, the gamma-ray leakage rate was 2.18 kW out of 578 kW. When the waste became an oxide, it was 10.3 kW. For the intermediate time from the boiling start and the dryup, the gamma-ray leakage rate (gamma heating) Rγ(t) (in kilowatts) was approximated by the following expressions:

(C.1) 0ξγ0.66,Rγt=1.64ξγt+2.180.66ξγ1,Rγt=20.8ξγt10.47,(C.1)

where ξγ(t) = mass vaporization ratio of liquid from the waste.

C.III. GAMMA HEATING OF THE CELL

shows the heating rate per unit depth within the cell concrete as a function of depth from the cell surface. It is the average value in all directions. Using this result and EquationEq. (C.2), the heating density qγ(t,x) (in kW∙m−3) corresponding to the progress of the boiling accident is obtained by the following approximate equations:

(C.2) 0ξγ0.66,qγt,x=0.584ξγt+0.0786exp14.2x0.66ξγ1,qγt,x=0.765ξγt0.393exp14.2x.(C.2)

Fig. C.1. Total heating rate per unit depth inside the cell concrete as a function of depth from the cell surface.

Fig. C.1. Total heating rate per unit depth inside the cell concrete as a function of depth from the cell surface.