Anda di halaman 1dari 17

Proceedings of the 2012 9th International Pipeline Conference IPC2012 September 24-28, 2012, Calgary, Alberta, Canada

IPC2012-90691

MODELING THE STATIC COOLING OF WAXSOLVENT MIXTURES IN A CYLINDRICAL VESSEL


Sridhar Arumugam University of Calgary Calgary, Alberta, Canada Adebola S. Kasumu University of Calgary Calgary, Alberta, Canada

Anil K. Mehrotra Department of Chemical & Petroleum Engineering and CEERE University of Calgary Calgary, Alberta, Canada crude oil can be shut down for long periods of time for ABSTRACT maintenance purposes. For such an un-insulated pipeline, when Under subsea conditions, the transportation of waxy the temperature of the surrounding environment is lower than crude oil through pipelines is accompanied by the precipitation the Wax Appearance Temperature (WAT) of the crude oil, the and deposition of higher paraffinic compounds as solids solubility of the paraffinic fractions in the crude oil is (waxes) onto the cooler surfaces of the pipeline. Wax considerably reduced, which leads to their precipitation and deposition is more pronounced during shut down of a pipeline subsequently, deposition onto the cold wall of the pipeline. The since the fluid is held at static conditions. In this study, the yield strength of the deposited solid could be sufficiently high static cooling of waxsolvent mixtures in a cylindrical vessel to reduce the effective pipe diameter available for flow during was modeled as a moving boundary formulation involving the restart of the pipeline. Thus there arises a need to liquidsolid phase transformation. The deposition process understand the physics behind the transient cooling process and during the transient cooling was treated as a partial to predict safe conditions for the start-up of the pipeline. freezing/solidification process. Also, the effect of the mixture composition and the cooling rate on the Wax Precipitation To develop a comprehensive model for the deposit layer Temperature (WPT) or the solubility curve of the waxsolvent formation and growth, several considerations such as mixture was taken into consideration when the bulk liquid thermodynamics and solid-liquid multiphase equilibrium, phase temperature was lowered below the WAT of the initial crystallization, pipeline surface characteristics, fluid dynamics, mixture composition. The predictions for the transient heat transfer, molecular diffusion and rheology are required [1]. temperature profiles in the liquid and the deposit region, and The onset of wax precipitation and the amount of wax the location of the liquiddeposit interface were validated with precipitated can be obtained through thermodynamic recently reported experimental results [19]. The predictions considerations for the waxy mixture considered while wax were also compared with the predictions for the gelling deposition is a transport process which requires a thermal behavior of waxsolvent mixtures under static cooling reported driving force [2]. by Bidmus [19]. The predictions for the temperature profile at seven thermocouple locations and the location of the liquid Many researchers have developed thermodynamic models deposit interface were in agreement with the experimental to predict the onset of wax precipitation. Won [3] studied the results and signified the important role of the solubility curve. solidliquid equilibrium between the precipitating wax and the The mathematical model presented was based on heat transfer waxy mixture and presented thermodynamic correlations for considerations and regarded the deposition process to be predicting the Cloud Point Temperature (CPT) and the amount thermally driven. of wax precipitated assuming the wax phase to be a homogeneous amorphous mixture, which does not undergo second or higher order solid phase transitions. Other INTRODUCTION researchers [4-7] have tried to improve on the model of Won [3] The static cooling of waxsolvent mixture in a cylindrical in one way or another. Hansen et al. [4] applied polymer vessel mimics the transient cooling of crude oil during solution theory for the oil phase while the wax phase was complete shutdown of a pipeline. Pipelines carrying waxy assumed to be an ideal mixture; they used measured cloud

Copyright 2012 by ASME

points to determine the interaction parameters in their model, thus getting a reasonable agreement between the calculated and measured cloud point temperatures. Pedersen et al. [5], using the regular solution theory, improved on the model of Won [3] by increasing the solid phase solubility parameters and making other heat capacity and enthalpy assumptions. Lira-Galeana et al. [6] developed an EOS-based molecular thermodynamic approach for calculating wax precipitation in petroleum mixtures assuming that the wax precipitated consists of several solid phases or the wax precipitation to be a multisolid-phase precipitation process. More recently, Farayola et al. [7] modified the Lira-Galeana et al. [6] model by using the PatelTeja three parameter EOS to model the non-ideality of the liquid phase. Most researchers have used the equilibrium relationship that relates the liquid phase (oil) fugacity and the solid phase (wax) fugacity with thermo-physical properties as the basis for the development of mathematical models, and they make use of the general fugacity equation for solid-liquid equilibrium. Differences, however, exist in the equations used to model the solid and liquid phases and the choice of thermophysical relations. The resulting models are used to predict the WAT values, and in some cases, amount and compositions of wax deposited from waxy or crude mixtures. Several mechanisms have been proposed to predict wax deposition in pipelines such as molecular diffusion, heat transfer, shear dispersion, Brownian diffusion and gravity settling. However, various studies [8-10] looked at the wax deposition from a molecular diffusion perspective and have argued molecular diffusion to be a dominant mechanism. In this approach, when a waxy mixture is transported through a pipeline with its surrounding temperature lower than the WAT of the fluid, the fluid adjacent to the wall can be cooled to a temperature below the WAT. The solubility of the wax being a decreasing function of temperature, more and more wax particles precipitate out of the bulk liquid phase and form a deposit layer on the cold wall. It is assumed that a prevailing temperature gradient between the wall and the bulk of the fluid also creates a concentration gradient necessary for the diffusion of the wax from the region of higher concentration (bulk phase) towards the cold wall where the concentration of dissolved wax is lower. Wax deposition is a thermally driven process and hence heat transfer has been treated as an alternative mechanism in several studies [11-17]. Two different modeling approaches have been suggested namely, steady state and unsteady state modeling of wax deposition. In the steady state modeling approach the thermal resistance offered over to the overall thermal driving force was divided into several individual thermal resistances across the liquid region and the deposit region [11-15]. The amount of wax deposited was related to the fractional temperature difference across the deposit layer. The steady state modeling approach was used to analyze the results of wax deposition in a

flow loop under laminar flow conditions [13] and under turbulent flow conditions [12]. Fong and Mehrotra [12] also suggested the use of a new parameter to define the enrichment of higher paraffins in the deposit during deposition. The unsteady state modeling approach [1, 2, 16, and 17] treats wax deposition to be a moving boundary problem which is characterized by the presence of an interface where the liquidsolid phase transformation of waxy mixture occurs. The formulation was used to predict the extent of wax deposition under static [1], laminar flow [1, 16] and turbulent flow conditions [17]. The model was also extended to incorporate the effect of shear stress on wax deposit characteristics by considering the release of a fraction of liquid phase trapped within the deposit layer through the one-dimensional deformation of a cubical cage [2]. The moving boundary formulation differs from the molecular diffusion approach in the manner it treats the liquid deposit interface temperature. An inherent assumption made in the molecular diffusion approach is that the interface temperature increases from an initial value close to the pipe wall temperature to the WAT at steady state. However, the heat transfer approach assumes that the interface temperature remains constant at the WAT of the liquid phase throughout the deposition process. Experimental results showed that the liquiddeposit interface temperature remained within the range of WAT of the liquid phase throughout the deposit-growth process [18, 19]. The flow of waxsolvent mixtures at a temperature above the WAT is termed as hot flow and the flow of waxsolvent mixtures at a temperature below the WAT of the initial composition, with wax particles suspended in the bulk liquid phase, is termed as cold flow. With the liquiddeposit interface temperature held at the WAT of the liquid phase, the temperature corresponding to the bulk phase precipitation of wax particles is regarded as the Wax Precipitation Temperature (WPT). Under cold flow conditions, the WPT of the liquid phase is lowered due to the precipitation of higher paraffins out of the bulk liquid phase. Hence the effective WPT depends upon the solubility curve for the waxsolvent mixture considered. In this study, the transient cooling of waxsolvent mixtures in a cylindrical vessel was modeled as a moving boundary formulation, taking into consideration the effect of composition and the cooling rate on the WPT (solubility curve) of the wax solvent mixture when the bulk liquid phase temperature was lowered below the WAT. The present modeling approach is a modification of the model presented by Bidmus [19], which assumed that the interface temperature was at the WAT of the initial composition throughout the deposition process. This assumption by Bidmus [19] implies that the bulk liquid phase can never exist below the WAT of the initial mixture composition.

Copyright 2012 by ASME

The model predictions for the transient temperature profiles and the location of the interface were validated with recently reported experimental results [19] for the batch cooling of waxsolvent mixtures in a cylindrical vessel. BATCH DEPOSITION EXPERIMENTS: STATIC COOLING FROM VESSEL WALL The objective of the batch deposition experiments performed by Bidmus [19] was to measure the changes in the liquiddeposit interface temperature during the transient cooling of the waxsolvent mixtures in a cylindrical vessel. Two different multi-component paraffin waxes were used for the experiments namely, Aldrich wax and Parowax. Aldrich wax consisted of n-alkanes ranging from C20-C40 with a mean carbon number of C28 while, Parowax consisted of n-alkanes ranging from C20-C50 with a mean carbon number of C29. These waxes were dissolved in a petroleum solvent (Norpar13) consisting of n-alkanes ranging from C9-C16, with a mean carbon number of 13. Two compositions of 14 mass% and 20 mass% Aldrich waxNorpar13 mixtures and two compositions of 5 mass% and 20 mass% ParowaxNorpar13 mixtures were used in the experiments. The present modeling predictions are however validated with only the experimental results for 14 mass% and 20 mass% Aldrich waxNorpar13 mixtures whose measured values of WAT were 31C and 34C, respectively. The apparatus used for these experiments consisted of a cylindrical vessel made of copper with an inside diameter of 10.2 cm and a height of 15.2 cm. The cylindrical vessel was placed in a temperature regulated bath where the coolant was held at a constant temperature, which was lower than the WAT of the waxsolvent mixture used. All surfaces of the cylindrical vessel, except for the side walls were insulated and hence the inner vessel wall acted as the cold deposition surface. As the mixture was cooled under static conditions, the thermal energy from the mixture was transferred radially outward with the deposit growing radially inward inside the vessel. The transient variation of the mixture temperature was recorded with the help of seven thermocouples labeled from TC1TC7 which were inserted about half way into the cylindrical vessel. The radial locations of the thermocouples inside the cylindrical vessel are listed in Table 1 and the schematic of the batch vessel used for deposition with static cooling from the vessel wall is shown in Figure 1. All of the batch cooling experiments were performed at an initial waxsolvent mixture temperature of WAT+21C while three different cooling temperatures of WAT5C, WAT7C and WAT10C were used. One of the important observations from these experiments was that the temperature of the liquid region decreased to the WAT of the initial waxsolvent mixture even before the liquiddeposit interface reached half way across the cylindrical vessel and the liquid region turned cloudy

due the precipitation of wax particles as the mixture was cooled below the WAT. MOVING BOUNDARY FORMULATION The wax deposition during the static cooling of a multicomponent waxsolvent mixture in a cylindrical vessel is modeled as a moving boundary formulation. The moving boundary formulation accounts for the transient heat transfer involved in phase change problems such as solidification and melting. The phase change problems are characterized by the presence of an interface or moving boundary wherein the liquidsolid phase transformation occurs. This partial solidification of the liquid adjacent to the liquiddeposit interface is treated to be an equilibrium process primarily governed by heat transfer. The phase transformation of pure substances such as in freezing of water occurs at a specified temperature. However, the phase change for a multi-component waxsolvent mixture occurs over a range of temperature and the location/position of the moving boundary is to be determined with time which makes the solution procedure complex. MODEL EQUATIONS: STATIC COOLING OF WAXY MIXTURES IN A CYLINDRICAL VESSEL Consider a homogeneous solution of a waxsolvent mixture held above its WAT in a cylindrical vessel. When the wall of the cylindrical vessel is subjected to a constant temperature lower than the WAT of the waxsolvent mixture by means of a circulating coolant, the temperature difference between the statically held waxsolvent mixture and the cold vessel wall provides the thermal driving force for the wax to precipitate from the liquid phase and deposit on the vessel wall. Thus, there exist two phases in the cylindrical vessel, the deposit phase and the bulk liquid phase, that are separated by an interface. As time progresses, the deposit layer grow radially inward in the cylindrical vessel and the thermal energy (both sensible heat and latent heat of phase change) is transferred radially outward. The deposit growth is treated as a result of the partial freezing/solidification of a layer of liquid adjacent to the deposit until the liquid phase temperature reaches the WAT of initial mixture composition of the waxsolvent mixture considered (hot flow condition: Tl WAT). In the batch deposition experiments performed by Bidmus [19], the temperatures recorded by all the thermocouples, TC1 to TC7, were the same until the liquid cooled down closer to the WAT indicating that there was no radial temperature gradient in the liquid phase. This suggests that the transient cooling of the liquid phase under hot flow conditions can be well approximated by considering heat transfer by natural convection in the liquid region. Thus the energy balance equation for the unsteady state heat transfer by natural convection in the liquid region is given as:

Copyright 2012 by ASME

2h (Tl Td ) k l ( R s)

1
l

Tl , Tl WPT t

0 < r < s (1)

The term df / dt represents the thermal energy released as a result of increased solid phase mass fraction f in the deposit. By the application of chain rule in equation 4:
T 1 r r r r 1 T t f T T , t

where, h is the heat transfer coefficient and s represents the location of the liquiddeposit interface. The initial WPT was approximated to be slightly higher than the WAT (WPT = WAT + 0.1C). This assumption was attributed to the effect of cooling rate on the WPT of waxy mixtures during hot flow. The Nusselt number for determining the heat transfer coefficient was obtained through an expression for heat transfer by natural convection in liquids in a cylindrical vessel as follows [20]:
Nu 0.55(Gr Pr) 0.25

s < r < R (5)

Equation 5 is further rearranged, and the resulting unsteady state energy balance equation for the deposit phase:
T 1 r r r r 1
'

T , t
'

s < r < R (6)

where, the modified thermal diffusivity is defined as:


1
'

in the deposit phase

(2a)

1 k

Equation 2a was modified to include a correction factor, f, in order to scale up the value of the initial heat transfer coefficient, h, used in the calculations. The modified form of equation 2a is given as:
Nu f * 0.55(Gr Pr) 0.25

f , T

(7)

The energy balance governing the movement of the interface and the growth of the deposit layer is given as:

(2b)

T r

h (Tl Td )

fs

ds , dt

r=s

(8)

It is assumed that there is no wax particles suspended in the liquid phase until it cools down to the WPT. When the liquid temperature reaches the WPT, wax particles precipitate in the bulk liquid and become suspended throughout the liquid phase. When the liquid phase temperature is lowered below the WPT, the heat transfer in the liquid phase with wax particles suspended in it is contributed by both conduction and convection. The resulting unsteady state energy balance equation for the liquid phase:
T 1 r l r r r 2h (Tl Td ) k l ( R s) 1
l

where, fs are the equilibrium solid phase fraction at the inner layer of the liquiddeposit interface. In order to solve the partial differential equations 1, 3, 6 and 8, which constitute the moving boundary formulation, a set of initial and boundary conditions were required. The boundary conditions are:

Tl / r 0 ,
Tl t dfl , Tl < WPT k l dt 0 < r < s (3)

r = 0, r = s, r = R,

t>0 t>0 t>0

(9a) (9b) (9c)

Tl = T = Td T = Tw

The term dfl / dt represents the thermal energy released due to the precipitation of wax particles in bulk liquid phase and fl is the mass fraction of wax precipitated. The energy balance equation for unsteady state heat transfer by conduction in the deposit region is as follows:
T 1 r r r r 1 T t df , k dt

The initial conditions for the static cooling of waxy mixtures in the cylindrical vessel are: Tl = Th s = R, t = 0, t = 0, 0rR (10a) (10b)

s < r < R (4)

THERMODYNAMIC CONSIDERATIONS As mentioned previously, the multi-component wax (Aldrich wax) consisted of n-alkanes ranging from C20-C40 with a mean carbon number of 28. It was dissolved in a petroleum solvent (Norpar13) which can be represented by a mean carbon

Copyright 2012 by ASME

number of 13. The waxsolvent mixture was treated as a pseudo binary mixture (C28-C13), represented by their mean carbon numbers to simplify the thermodynamic phase behaviour considerations. It was treated that the C28-C13 would form a eutectic binary mixture. The temperaturecomposition phase diagram for an ideal binary eutectic system can be obtained from the freezing point depression equation.

conditions (equation 9) for the bulk liquid temperature less than the WAT of the initial mixture composition were solved to obtain the temperature profiles for the deposit region, average liquid phase temperature and the movement of the liquid deposit interface with time. The model equations were discretized using an explicit method. The pipe radius (R) was divided into n equally spaced concentric rings such that the radial increment, r = R/n. The time increment, t, and the radial increment, r, were chosen such that to satisfy the stability criterion:

ln x i

( H m )i 1 Rg (TL )i

1 , (T m )i

i = 1, 2 (11)

where, ( H m )i , (TL )i and (T m )i are the enthalpy of melting, liquidus temperature and the melting-point temperature of species i, respectively. The values of the enthalpy of melting and the melting point temperature for C13 and C28 were 28.5 MJ kmol1 and 103.0 MJ kmol1 and 267.8 K and 333.2 K [14]. The liquidus temperature as a function of the composition was estimated from the equilibrium relationship. The estimated solidus temperature for the C28-C13 eutectic mixture was -8.8C. The solid phase mass fraction in deposit (f) held at the equilibrium temperature (T) was obtained by applying the level rule between w28 and w * 28
f ( w 28 w * ) /(1 w * ) , 28 28

t / r 2 1/ 2

(13)

(12)

The input quantities were: the initial mixture composition, w28; initial mixture temperature, T h; the coolant temperature, Tc; and the initial liquid deposit interface temperature, T d. The interface temperature was held at the WAT of the liquid phase at all times. The lever rule (equation 12) was used repeatedly at time, t > 0, to estimate the equilibrium solid phase fraction at all the radial locations (s < r < R) in the deposit region. The liquid phase properties were estimated at its average temperature, (Tl + Td)/2 and small density changes that could result during the partial solidification of the liquid layer adjacent to the deposit was neglected, i.e. l = f = . Further details of the numerical solution for determining the liquid and deposit phase physical and thermal properties of eutectic binary mixtures are available elsewhere [1, 21]. RESULTS AND DISCUSSION The transient cooling of waxsolvent mixtures was considered as a moving boundary problem. Predictions were made for the time dependant cooling of 14 mass% and 20 mass% waxsolvent mixtures in the cylindrical vessel. The predictions for the average liquid phase temperature, the local deposit phase temperatures, and the location of the liquid deposit interface were validated with the experimental results for an inlet mixture temperature of WAT+21C and for a coolant temperature of WAT10C for both compositions. The predictions were also compared with the predictions made by Bidmus [19] where the wax deposition was treated as a partial freezing/solidification problem throughout the deposition process. Temperature profiles during transient cooling The predicted temperature profiles at each thermocouple location, from TC1 to TC7, during the transient cooling of the 14 mass% and 20 mass% waxsolvent mixtures with an inlet mixture temperature of WAT+21C and a coolant temperature of WAT10C are compared with the experimental results and with the predictions of Bidmus [19] in Figures 3 and 4. Note that these predictions were obtained by considering a factor, f, equal to 1.5 to calculate the value of the Nusselts number, Nu, and hence the heat transfer coefficient, h. This implies a 50%

where, w28 is the mass fraction of C28 in the liquid mixture and w * is the equilibrium mass fraction of C28 at T in the deposit. 28 Note that the mass fraction of C28 in the liquid mixture remains constant under hot flow conditions, while the mass fraction of C28 in the liquid mixture is a function of the solubility of the wax in the paraffinic solvent under cold flow conditions. The solubility curve for the given wax in the paraffinic solvent was obtained from the equilibrium relationship given as equation 11. The predicted values of WAT for C28-C13 binary mixtures corresponding to 30 mass% solid of the effective composition and the experimentally measured WAT obtained from previous studies [19, 21] are shown in Figure 2. A good agreement between the predicted and the experimental WAT values was observed for higher compositions while the predicted WAT values are slightly higher than the experimental WAT values for lower compositions of Aldrich waxNorpar13 considered. NUMERICAL SOLUTION METHODOLOGY Equations 1, 6 and 8 along with the boundary and initial conditions (equations 9 and 10) for hot flow conditions (T l WAT) or equations 3, 6 and 8 along with the boundary

Copyright 2012 by ASME

increase in the value of the heat transfer coefficient as that of equation 2a. The temperature profiles are divided into the liquid region temperature profile and the deposit region temperature profile using the WAT of the initial waxsolvent mixture represented as dashed lines. Figures 3a and 4a represent the experimental results; Figures 3b and 4b represent the predictions by the model used in this study, while Figures 3c and 4c represent the predictions by Bidmus [19] where a high effective thermal conductivity of about 3.5 W m1 K1 was used to account for the convective heat transfer in the liquid region. As shown in Figures 3 and 4, the predicted temperature profiles of the present model for the liquid and the deposit phases, match the experimental results well. Also, the radial temperature gradient observed in the liquid region by Bidmus [19] was not observed in the results of this model, since the heat transfer in the liquid region was assumed to be governed by convection alone. For a better understanding of the transient cooling process, the liquid region and the deposit region temperature profiles are discussed in detail in the following sections. Liquid region temperature profiles The predicted liquid region temperature profile for two different values of factor, f, corresponding to 1.0 and 1.5 was compared with the experimental results and with the predictions by Bidmus [19]. The predictions by Bidmus [19], for the liquid region temperature profiles in Figures 3 and 4, showed a radial temperature gradient in the liquid region, despite using a significantly high effective thermal conductivity of 3.5 W m1 K1. The high value of thermal conductivity was not sufficient to offset the radial temperature gradient shown in their predictions. As shown in Figures 3 and 4, no such temperature gradient was observed in the liquid region for the predictions of the present model, and in the experimental results. Hence, the use of natural convection heat transfer for the liquid phase in the present model is a more accurate representation of the heat transfer process occurring in the bulk liquid than the use of an effective thermal conductivity in the conduction equation to account for the convection effects. Also, due to the existence of a radial temperature gradient in the liquid phase, a volume-weighted average method was adopted by Bidmus [19] for each cooling/deposition experiments to calculate the average liquid region temperature, as shown in Figure 5. This procedure was however not required to calculate the average liquid region temperature corresponding to each thermocouple location using the present modeling approach. Figure 5 shows the predicted and experimental average liquid phase temperature profiles for the 14 mass% and 20 mass% waxsolvent mixtures with an inlet mixture temperature of WAT+21C and a coolant temperature of WAT10C. The predicted temperature profile by Bidmus [19] using an effective thermal conductivity of 3.5 W m1 K1 and a volume average

method agrees satisfactorily with the experimental results. The predicted temperature profile by the present model for both the 14 mass% and 20 mass% waxsolvent mixtures agree well with the experimental results as a factor of 1.5 was used to calculate the heat transfer coefficient over a factor of 1.0. The dependence of the convective heat transfer coefficient (h) on the temperature difference between the bulk of the fluid and the interface reduces the value of the heat transfer coefficient as the liquid cools from its initial mixture temperature. Hence the observed cooling rate decreases as the deposition progresses. However, in the case of the predicted results by Bidmus [19], a thermal conductivity of 3.5 W m1 K-1 was assumed throughout the deposition process for predicting the heat transfer in the liquid region. Deposit region temperature profiles The predictions for the local deposit region temperature profiles at the seven thermocouple locations were compared with the experimental results for 14 mass% and 20 mass% waxsolvent mixtures as shown in Figures 6 and 7, respectively. The inlet mixture temperature was WAT+21C and the coolant temperature was WAT10C and a factor of 1.5 was used in all the cases. In Figures 6 and 7, the predictions of the model presented in this work for the deposit region temperature profile compare well with the experimental results for all the thermocouple locations from TC1TC7. However the predictions by Bidmus [19] for the thermocouple locations TC6 and TC7 which were closer to the center of the vessel were higher than the experimental measurements by 1C. Bidmus [19] also observed that their predicted profiles showed a flat region when the local temperature was equal to the WAT (TCWAT = 0). However, a gradual decrease in the temperature was observed in the experimental results and in the model predictions presented here. Growth of the deposit layer Figure 8 shows the comparison between the predicted deposit layer growth using two different values of factor, f, corresponding to 1.0 and 1.5 and the observed location of the liquiddeposit interface at four thermocouple locations, namely from TC1 to TC4 for the 14 mass% and 20 mass% waxsolvent mixtures. For the deposit layer growth recorded at TC1 and TC2 the predictions by Bidmus [19] matches well with the experimental results, while the prediction of the present model using a factor, f, equal to 1.5 satisfactorily matches the deposit layer growth recorded at TC1, TC3 and TC4 experimental observations. CONCLUSIONS The static cooling of waxsolvent mixtures was successfully modeled as a moving boundary problem. The

Copyright 2012 by ASME

assumption of natural convection heat transfer in the liquid region provided satisfactory predictions over the assumption of an effective value of thermal conductivity to account for the convection effects in the liquid region. The inclusion of the effect of solubility curve on the WAT of the liquid phase provided good agreements with experimental observations for the local liquid region and the deposit region profile and for the growth of the deposit layer. It was shown that the inclusion of the solubility curve in the wax deposition process, in addition to being a partial freezing/solidification problem, is a better representation of the transient cooling process. Model predictions of this work gave satisfactory results for the static case, and this approach is extended to solids deposition under various flow conditions in the following chapters. Also, the present model shows that wax deposition from waxy mixtures can be satisfactorily modeled as a thermally-driven process. NOMENCLATURE f = correction factor f = solid phase mass fraction in the deposit fl = solid phase mass fraction suspended in the liquid fs = mass fraction of solid-phase at liquid-deposit interface Gr = Grashof number h = heat transfer coefficient, W m-2 K-1 k = thermal conductivity, W m-1 K-1 Nu = Nusselt number Pr = Prandtl number s = radial location of the liquiddeposit interface, m t = time, s T = temperature, C or K T = deposit layer temperature, C Td = liquiddeposit interface temperature interface, C Tl = average liquid region temperature, C TL = liquidus temperature, C Tm = melting-point temperature, C w28 = mass fraction of C28 in liquid phase w28* = equilibrium mass fraction of C28 in liquid phase x = mole fraction SUBSCRIPTS i = ith component l = liquid phase = deposit phase GREEK LETTERS 1 = liquid phase thermal diffusivity (m2 s1) = deposit phase thermal diffusivity (m2 s1) = apparent thermal diffusivity (m2 s1) = deposit layer thickness (m) = density (kg m3) = heat of solid phase transformation (J kmol1)

ACKNOWLEDGMENTS Financial support from the Natural Sciences and Engineering Research Council of Canada (NSERC) is gratefully acknowledged. Additional financial support was provided by the Centre for Environmental Engineering Research & Education (CEERE) and the Department of Chemical and Petroleum Engineering at the University of Calgary, Calgary, Alberta, Canada. REFERENCES [1] Bhat, N. V., and Mehrotra, A. K., 2005, Modeling of Deposit Formation from Waxy Mixtures via Moving Boundary Formulation: Radial Heat Transfer under Static and Laminar Flow Conditions, Ind. Eng. Chem. Res., 44, pp. 6948-6962. [2] Bhat, N. V., and Mehrotra, A. K., 2008, Modeling the Effect of Shear Stress on the Composition and Growth of the Deposit Layer from Waxy Mixtures under Laminar Flow in a Pipeline, Energy & Fuels, 22, pp. 3237-3248. [3] Won, K. W., 1989, Thermodynamic Calculation of Cloud Point Temperatures and Wax Phase Compositions of Refined Hydrocarbon Mixtures, Fluid Phase Equilibria, 53, pp. 377-396. [4] Hansen, J. H., Fredenslund, A., Pedersen, K. S., and Ronningsen, H. P., 1988, A Thermodynamic Model for Predicting Wax Formation in Crude Oils, AIChE Journal, 34, pp. 1937-1942. [5] Pedersen, K. S., Skovborg, P., and Ronningsen, H. P., 1991, Wax Precipitation from North Sea Crude Oils. 4. Thermodynamic Modeling, Energy & Fuels, 5(6), pp. 924932. [6] Lira-Galeana, C., Firoozabadi, A., and Prausnitz, J. M., 1996, Thermodynamics of Wax Precipitation in Petroleum Mixtures, AIChE Journal, 42, pp. 239-248. [7] Farayola, K. K., Adeboye, Y. B., Adekomaya, O. A., and Olatunde, A. O., 2010, Thermodynamics Prediction of Wax Precipitation Using the PatelTeja Equation of State, SPE Paper No. 136964. [8] Burger, E. D., Perkins, T. K., and Streigler, J. H., 1981, Studies of Wax Deposition in the Trans Alaska Pipeline, SPE Paper No. 8788. [9] Brown, T. S., Niesen, V. G., and Erickson, D. D., 1993, The Effects of Light Ends and High Pressure on Paraffin Formation, SPE Paper No. 28505. [10] Singh, P., Fogler, H. S., and Nagarajan, N., 2000, Formation and Aging of Incipient Thin Film WaxOil Gels, AIChE Journal, 46(5), pp. 1059-1074. [11] Bidmus, H. O., and Mehrotra, A. K., 2004, HeatTransfer Analogy for Wax Deposition from Paraffinic Mixtures, Ind. Eng. Chem. Res., 43, pp. 791-803. [12] Fong, N., and Mehrotra, A. K., 2007, Deposition under Turbulent Flow of Wax-Solvent Mixtures in a BenchScale Flow-Loop Apparatus with Heat Transfer, Energy & Fuels, 21, pp. 1263-1276.

Copyright 2012 by ASME

[13] Parthasarathi, P., and Mehrotra, A. K., 2005, Solids Deposition from Multicomponent WaxSolvent Mixtures in a Benchscale Flow-Loop Apparatus with Heat Transfer, Energy & Fuels, 19, pp. 1387-1398. [14] Bidmus, H. O., and Mehrotra, A. K., 2009, Solids Deposition during Cold Flow of Wax-Solvent Mixtures in a Flow-loop Apparatus with Heat Transfer, Energy & Fuels, 23, pp. 31843194. [15] Tiwary, R., and Mehrotra, A. K., 2009, Deposition from Wax-Solvent Mixtures under Turbulent Flow: Effects of Shear Rate and Time on Deposit Properties, Energy & Fuels, 23, pp. 12991310. [16] Bhat, N.V., and Mehrotra, A. K., 2006, Modeling of Deposition from Waxy Mixtures in a Pipeline Under Laminar Flow Conditions via Moving Boundary Formulation, 2006, Industrial and Engineering Chemistry Research, 45(25), pp. 8728-8737. [17] Bhat, N.V., and Mehrotra, A. K., 2010, Deposition from Waxy Mixtures under Turbulent Flow in Pipelines: Inclusion of a Viscoplastic Deformation Model for Deposit Aging, Energy & Fuels, 24, pp. 2240-2248. [18] Bidmus, H. O., and Mehrotra, A. K., 2008b, Measurement of LiquidDeposit Interface Temperature during Solids Deposition from WaxSolvent Mixtures under Sheared Cooling, Energy & Fuels, 22(6), pp. 4039-4048. [19] Bidmus, H. O., 2009, Static Sheared and Cold Flow Wax Deposition, Ph.D. Thesis, University of Calgary, Calgary, Canada. [20] Evans, L. B., and Stefany, N. E., 1966, Chem. Eng. Progress, Symp. Series, 62(64), 209. [21] Bhat, N. V., 2008, Modeling of Solids Deposition from WaxSolvent mixtures based on the Moving Boundary Problem Approach, Ph.D. Thesis, University of Calgary, Calgary, Canada. [23] Bidmus, H. O., and Mehrotra, A. K., 2008a, Measurement of LiquidDeposit Interface Temperature during Solids Deposition from WaxSolvent Mixtures under Static Cooling Conditions, Energy & Fuels, 22(2), pp. 1174-1182. [24] Mehrotra, A. K., Bidmus, H. O., Bhat, N. V., and Tiwary, R., 2009, Modeling the Gelling Behavior of Wax Solvent Mixtures under Static Cooling,, Trends in Heat and Mass Transfer, Vol. 11, pp. 17-31.

Copyright 2012 by ASME

ANNEX A LIST OF TABLES

Table 1. Radial location of the thermocouples in the cylindrical vessel under static cooling from vessel wall [19].

Thermocouple label TC1 TC2 TC3 TC4 TC5 TC6 TC7

Radial distance from the vessel wall (inch) 0.250 0.375 0.500 0.625 0.750 1.500 2.000

Copyright 2012 by ASME

ANNEX B LIST OF FIGURES Figure 1. Schematic of the batch cooling experiment for static cooling from vessel wall [19].

10

Copyright 2012 by ASME

Figure 2. Predicted and experimental values [19, 21] of WAT for Aldrich waxNorpar 13 mixtures

40

35

temperature (C)

30

25

predicted WAT (Aldrich wax-Norpar13) Measured WAT (Aldrich wax-Norpar13)


20

15

10 5 10 15 20 25

wax concentration (mass%)

11

Copyright 2012 by ASME

Figure 3. Predicted and experimental temperature profiles at TC1 to TC7 radial locations for 14 mass % C 28C13 mixtures (data of Bidmus [19])

20 15 10 5 0 -5 a) experimental (Bidmus [19])

radial location TC1 TC2 TC3 TC4 TC5 TC6 TC7 WAT WAT

temperature [TC - WAT] (C)

-10 20 15 10 5 0 -5 -10 20 16 12 8 4 0 -4 -8 0 20 40 60 80 100 120 140 160 c) predicted (Bidmus [19]) radial location TC7 TC6 TC5 TC4 TC3 TC2 TC1 WAT WAT b) predicted - present model

time (min)

12

Copyright 2012 by ASME

Figure 4. Predicted and experimental temperature profiles at TC1 to TC7 radial locations for 20 mass % C 28C13 mixtures (data of Bidmus [19])

20 15 10 5 0

a) experimental (Bidmus [19])

radial location TC1 TC2 TC3 TC4 TC5 TC6 TC7WAT WAT

temperature [TC - WAT] (C)

-5 -10 20 15 10 5 0 -5 -10 20 16 12 8 4 0 -4 -8 0 20 40 60 80 100 120 140 160 WAT b) predicted - present model

c) predicted (Bidmus [19])

radial location TC7 TC6 TC5 TC4 TC3 TC2 TC1 WAT

time (min)

13

Copyright 2012 by ASME

Figure 5. Comparison of predicted and experimental average liquid region temperature for (a) 14 mass% C28C13 mixtures, and (b) 20 mass % C28C13 mixtures (data of Bidmus [19])

20

(a) 14 mass%

15

average liquid phase temperature [TC - WAT] (C)

10

experimental (Bidmus [19]) predicted (Bidmus [19]) predicted, f = 1.5 predicted, f = 1.0

WAT

20 (b) 20 mass% 15

10

5 WAT

Col 1 vs Col 2 Col 3 vs Col 5 Col 19 vs Col 20 Plot 7


Col 19 vs Col 23

10

20

30

40

time (min)

14

Copyright 2012 by ASME

Figure 6. Comparison of predicted and experimental deposit local temperature profiles for 14 mass % C 28C13 mixtures (data of Bidmus [19])

0 -2 -4 -6 -8 -10 0 -2 -4 -6 -8 -10 0 -2 -4 -6 -8 -10 0 -2 -4 -6 -8 -10 0 100 200 300 (g) TC7 (e) TC5 (c) TC3 (a) TC1

WAT (b) TC2

WAT

experimental (Bidmus [19] ) predicted - (Bidmus [19] ) predicted - present model

deposit local temperature [TC - WAT] (C)

WAT (d) TC4

WAT

WAT (f) TC6

WAT

WAT

100

200

300

400

time (min)

400

time (min)

15

Copyright 2012 by ASME

Figure 7. Comparison of predicted and experimental deposit local temperature profiles for 20 mass % C 28C13 mixtures (data of Bidmus [19])

0 -2 -4 -6 -8 -10 (a) TC1

WAT (b) TC2 experiment (Bidmus [19] ) predicted (Bidmus [19] ) predicted - Present model

WAT

deposit local temperature [TC - WAT] (C)

0 -2 -4 -6 -8 -10 0 -2 -4 -6 -8 -10 0 -2 -4 -6 -8 -10 0 100 200 300 (g) TC7 (e) TC5 (c) TC3

WAT (d) TC4

WAT

WAT (f) TC6

WAT

WAT

100

200

300

400

time (min)

400

time (min)

16

Copyright 2012 by ASME

Figure 8. Comparison of the predicted and experimental deposit layer growth with time for a) 14 mass% C 28-C13 mixture, and b) 20 mass % C28-C13 mixtures (data of Bidmus [19])

0.5 a.) 14 mass% 0.4

0.3

deposit thickness (/R)

0.2

0.1

0.0 0.5 predicted (Bidmus [19]) experimetal (Bidmus [19]) predicted, f = 1.5 predicted, f = 1.0 b. ) 20 mass %

0.4

0.3

0.2

0.1

0.0 0 5 10 15 20 25 30 35

time (min)

17

Copyright 2012 by ASME

Anda mungkin juga menyukai