Next Article in Journal
Simulation Research on Sparse Reconstruction for Defect Signals of Flip Chip Based on High-Frequency Ultrasound
Previous Article in Journal
Truck Handling Stability Simulation and Comparison of Taper-Leaf and Multi-Leaf Spring Suspensions with the Same Vertical Stiffness
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Forced Convection Heat Transfer for Stratospheric Airship Involved Flight State

1
Key Laboratory of Aircraft environment control and life support, MIIT, Nanjing University of Aeronautics & Astronautics, 29 Yudao Street, Nanjing 210016, China
2
College of Energy & Power Engineering, Jiangsu University of Science and Technology, 2 Mengxi, Jingkou, Zhenjiang 212003, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(4), 1294; https://doi.org/10.3390/app10041294
Submission received: 30 December 2019 / Revised: 6 February 2020 / Accepted: 9 February 2020 / Published: 14 February 2020
(This article belongs to the Section Energy Science and Technology)

Abstract

:
Forced convection heat transfer is a significant factor for the thermal control of a stratospheric airship. However, most of researches were conducted without considering the influence of flight state causing serious errors. In order to accurately predict the forced convection heat transfer of the stratospheric airship at an angle of attack, firstly, an empirical correlation of Nusselt number (Nu) as function of Reynolds number (Re) andlength to diameter ratio (e) is developedunder horizontal state based on a validated computational fluid dynamic (CFD) method. Then, a correction factor K, considering its angle of attack (α), is proposed to modify this correlation. The results show that: (1) Nusselt number increases with the increase of Reynolds number, decreases as the length to diameter ratio changes from 2 ~ 6, and increases as the angle of attack changes from 0° ~ 20°. (2) At higher Reynolds number, the calculated results are 30 percent higher than those of previous studies with α = 20°. (3) Compared with α and e, the effect of Re on correction factor K can be ignored, and K is a strong equation of α and e. The efficiency of heat transfer is increased by 6 percent with α = 20°. The findings of this paper provide a technical reference for the thermal control of a stratospheric airship.

1. Introduction

A stratospheric airship is a type of aircraft that takes off by buoyancy. Compared with other aerostats, the stratospheric airship has its unique advantages and is widely used in civil fields such as satellite communications, meteorological measurements, and military fields such as surveillance and defense [1,2,3].
Convection heat transfer process, including natural convection and forced convection, is one of the most common natural phenomena, which is widely used in almost all of the engineering fields, such as heat exchanger [4,5], aircraft [6,7,8,9], electronics [10], airship [11,12,13,14,15,16,17,18,19,20,21,22], and so on.
Thermal performance is one of the most important factors affecting the flight state of a stratospheric airship. The stratospheric airship is filled with a large amount of gas, and it experiences a rough external environment during flight, as the change in temperature affects the buoyancy to a large extent [23]. In the past decades, many research activities have been held on the forced convection heat transfer of a stratospheric airship. In their researches, Kreith and Kreider [11] established a simple numerical model to simulate the average temperature of balloon envelope and lifting gas. Yao [12] proposed a multi-node thermal transient model of stratospheric airship and verified the model using high-altitude flight experiments. Fang [15] built a two-node thermal model to analyze heat source and heat transfer patterns that affect thermal balance and thermal performance of airships flying in stratospheric environment. However, the most existing forced convection heat transfer correlations are only applicable in the case of low Reynolds number, so Dai [17] investigated the steady forced convection heat transfer of an isothermal spherical aerostat with the Reynolds number range from 20 to 108, and a new piecewise correlation was proposed. Later, Shi [21,22] proposed a new fixed-point adjustment method of an airship to solve the problem of height instability due to dramatic daily-temperature swings. The airship membrane was discretized into a triangular element to enhance the computing accuracy, and a multi-node thermodynamic model of the airship was established.
In summary, despite the attainment of considerable knowledge on the forced convection heat transfer of a stratospheric airship (through experimental or numerical work), the most relevant empirical correlations are only applicable to spherical airships or airships flying under horizontal state. So many conclusions are not universal because the consistency of study object and correlation equations geometric model is often neglected. And, no specific correlation is available on the external forced convection heat transfer of an ellipsoidal airship flying with a certain angle of attack, although the literature has pointed out that the efficiency of heat transfer is increased by 7 percent considering the angle of attack [24]. The heat transfer criterion equations need to be improved. The present study investigates the effects of Reynolds number, length to diameter ratio, and angle of attack on the thermal characteristics of an ellipsoidal airship to answer this demand. A new correlation of the average Nusselt number is built based on the data obtained from computational fluid dynamic (CFD) calculations. Moreover, a correction factor K considering angle of attack is proposed to modify this formula.

2. Numerical Method

2.1. Geometric Model

An ellipsoidal airship is taken as the research object in this paper, and its generatrix equation [25] can be written as
x 2 ( L / 2 ) 2 + y 2 ( D / 2 ) 2 = 1       ( L > D > 0 )
where L is the length of the airship, D is the diameter of the airship.
In recent years, the CFD method has been able to successfully simulate the thermal characteristics of an airship, owing to the rapid development of computer technology. Moreover, its precision is sufficient to meet the demands of engineering calculation. However, the scaled-model is only geometrically similar to the actual object, and the Reynolds number is not equal, which lead to a large error due to the scale effect when converting the model data into the actual object [26,27]. Thus, a full-scale airship model with L = 100 m is used in this paper to accurately simulate the forced convection heat transfer over the external surface of an airship. Here, the length to diameter ratio is defined as e = L/D. Five cases with e = 2, 2.5, 3, 4, and 6 are numerically simulated by changing the value of D to change the value of e. The values of L and D are shown in Table 1.
The numerical simulations are based on the CFD software, Version 18.0, and the ICEM software, Version 18.0 is used for the meshing. The research object in this paper is a symmetrical structure without a tail wing; thus, only a 1/2 body structure is used. At the same time, in order to ensure the symmetry of the flow field structure, a “Symmetry” boundary condition is used, and the airflow direction is parallel with the SYM (as shown in Figure 1).

2.2. Control Equations

The phenomena of flow and heat transfer were controlled by the following equations [28]:
Continuity equation
V x x + V y y + V z z = 0
Momentum equation
ρ ( V x V i x + V y V i y + V z V i z ) = P i + ρ g i + μ ( 2 V i x 2 + 2 V i y 2 + 2 V i z 2 )
where i represents x-component, y-component, and z-component, V is the velocity vector, μ is the viscosity of the fluid, P is the pressure of the fluid, ρ is the density of the fluid, g is the gravitational acceleration.
Energy equation
ρ c p ( V x T x + V y T y + V z T z ) = k ( 2 T x 2 + 2 T y 2 + 2 T z 2 )
where c p is the specific heat of the fluid, k is the thermal conductivity and T is the temperature of the fluid.

2.3. Computational Domains, Mesh, and Boundary Condition

The computational domain and the configuration of the airship are illustrated in Figure 1. Figure 1 shows that the boundary conditions of the computational domain are divided into INLET, OUTLET, WALL, FARWALL, and SYM. The INLET boundary is assumed to be a uniform velocity inlet given the magnitude and direction. Different angles of attack are indicated by changing the direction of the velocity. The OUTLET boundary is assumed to be “Opening” with a reference pressure of 0 Pa. The FARWALL boundary is assumed to be “Opening”. The WALL boundary is assumed to be a no-slip wall, and its temperature is 298.15 K. The SYM boundary is assumed to be “Symmetry”. The thermal physical properties of the flow fluid are assumed to be constant at a temperature of 288.15 K, and the pressure of the computational domain is set as 1 atm. The fully implicitly coupled multigrid linear solver is used. The high resolution scheme is used for advection scheme and turbulence numerics.
Given the large physical size of an airship, its Reynolds number may easily increase to a magnitude over 106, which is far beyond the critical Reynolds number. Thus, the flow around an airship is a typical turbulent flow. The computational domain is divided into two separate subdomains, namely, the internal domain and the external domain, to precisely investigate the thermal characteristics over the external surface of an airship. The mesh of the internal domain, includingboundary layers, is fine, and the first layer of the mesh around the surface lies at y+≤1. The mesh of the external domain is relatively coarse. Through convergence analysis, the number of elements is controlled at about 2,000,000. The max aspect ratio is 36.4, the min quality is 0.549, the max value of max angle is 147, and the min value of min angle is 34.2, and the min equiangle skewness is 0.37. The computational mesh is illustrated in Figure 2.

2.4. Turbulence Model

The turbulence model has a significant effect on the results of the numerical simulation [29]. The suitability of the k-ω turbulence model in a study on forced convection heat transfer around an ellipsoidal airship has been proven in the literature [30]. To further verify the applicability of the turbulence model in the present study, a numerical simulation is carried out by using k-ω turbulence model, and the pressure coefficients Cp is plotted in Figure 3. Because experimental results are not available to directly verify the numerical results of an ellipsoidal airship, the pressure coefficients data of a spherical airship is used as a reference to verify the turbulence model and numerical scheme. The size of the model and boundary conditions are same with the literature [31]. The pressure coefficients obtained in the present study are compared with the experimental measurements of Achenbach. In Figure 3, it can be observed that the performance computed with k-ω turbulence model is close to the experimental measurements. Consideringthe smaller discrepancy, thus, k-ω turbulence model is used in the present study.

3. Results and Discussions

Existing studies have shown that the main factors affecting forced convection heat transfer around an ellipsoidal airship are the length to diameter ratio and the Reynolds number under a horizontal state [20,28,32,33,34]. However, an airship generally ascends or descends at a certain angle of attack; thus, estimating its thermal characteristics under this condition is necessary.
The effects of different parameters on forced convection heat transfer are studied in this paper. The values of a single parameter is numbered from 1 to 5, as shown in Table 2.
A reference value is set with Re of 1 × 107, e of 2, and α of 0°. Only the value of the analyzed parameter is changed during the analysis, and the values of the remaining two parameters arethe same as the reference values. A total of 15 cases is conducted in this part to study the effects of different parameters on the forced convection heat transfer around an ellipsoidal airship.

3.1. Effects of Different Parameters on Heat Transfer

Figure 4 shows the relation between the average Nusselt number (Nu) and Re, e, α. Nu can be obtained by N u = ( S N u θ d S ) / S where S is the area of the airship, θ is the angular position.
Figure 4a shows that Nu clearly increases with the increase of Re. According to the definition of the Reynolds number, that is, Re = Ul/ν, where U is the velocity of the fluid, l is the characteristic length of the airship, and ν is the kinematic viscosity of the fluid. When L and ν are constant, URe. The larger the value of Re, the larger the value of U around an airship, which results in the removal of more heat and the increase of Nu.
Figure 4b demonstrates that Nu decreases with the increase of e. According to the definition of e above, the shape of an airship is close to a sphere when e is small. However, the shape of an airship is close to a flat plate as e increases. References [35,36,37] indicate that a vortex occurs in the tail region of a sphere while a uniform flow passes over a sphere, thereby resulting in enhanced convection heat transfer.
Figure 4c indicates that Nu increases with the increase of α. The presence of the angle of attack (α) can cause changes in the flow structure around an airship, thereby resulting in a windward side, which eventually leads to a higher convective heat transfer.
Figure 5 shows the velocity contours and local velocity vectors around the airship with angle of attack of 0° and 20°. The airflow flows from left to right. It is obvious that, at the front region, the velocity of the airflow decreases significantly, and then increases along the airship hull. It reaches a maximum value at the middle region of the airship hull and then gradually decreases. At the tail region, the velocity of the airflow is almost zero due to the separation of the airflow from the surface of the airship. Comparing Figure 5b with Figure 5a, it can be seen that the flow field structure around the airship is changed due to the existing of angle of attack. On the windward side, the position where the velocity of the airflow reaches its maximum moves backward, and in the leeward side, it shows an opposite trend. The change of flow field structure around the airship is the root reason for the change of forced convection heat transfer.

3.2. Formula Fitting

3.2.1. Horizontal State

The horizontal state is the main state in a flight course. Thus, investigating the forced convection heat transfer around an ellipsoidal airship under horizontal state is important. Assume α = 0°, the values of Re and e are determined according to Table 2. The average heat transfer in terms of the Nu around an ellipsoidal airship can be written as a power law equation
N u = c 1 R e c 2 e c 3
where c1, c2 and c3 are constant coefficients.
Based on the data obtained from the aforementioned simulation results, a new correlation of Nu with a determination coefficient (r2) of 0.9996 and a root mean squared error (RMSE) of 2054 is proposed via MATLAB R2018a. The fitting points and fitting surface are shown in Figure 6.
The values of the fitting parameters are listed in Table 3. The value of c1 ~ c3 is introduced into Formula (5) to obtain Formula (6)
N u = 0.0161 R e 0.8543 e 0.0454
where Re ∈ [1 × 107, 2.0 × 108] and e ∈ [2, 6].
The model of the non-fit points is numerically simulated to validate the correctness of Formula (6). Table 4 compares the values of Nu calculated by Formula (6) with those obtained from the simulation results. The calculated results agree well with those of the simulation results.

3.2.2. State Modification

The problem of changes of attack angle of a stratospheric airship is an important subject on the research of the Stratospheric Airship Platform. Li [38] pointed out that the airship reaches its stable attack angle in a short time after released, and that the attack angle is usually large. Then, the airship ascends at its stable attack angle. So, it is necessary to investigate the forced convection heat transfer of the airship at a certain attack angle.
The definition of K is introduced in this part and can be expressed in the form of
K ( R e , e , α ) = N u ( R e , e , α ) / N u ( R e , e )
Figure 7 shows the relation between K and α (Re = 1 × 107). Clearly, K increases with the increase of α regardless of the value of e, the larger the attack angle, the greater the difference. K increases with the increase of the e for a given α, especially at higher α.
Figure 8 shows the relation between K and Re (e = 2). Clearly, K remains nearly the same as the increase of Re regardless of the value of α. For example, K reduces from 1.052 to 1.033 with Re changes from 1 × 107 to 2.0 × 108 when α = 20°. This result means that the effect of Re on K can be ignored. K increases with the increase of α for a given Re. For instance, the value of K at α = 20° and α = 5° is 1.053 and 1.004, respectively, with a difference of 4.8% when Re = 1 × 107.
The correction factor K can be simplified to K (e, α) based on the above analysis. The mean valueof K corresponding to different Re is taken as the final value of K. Based on the data obtained from the above simulation results, a correlation of the correction factor K with a determination coefficient (r2) of 0.9995 and a RMSE of 0.0005286 is proposed via MATLAB R2018a. The formula can be written as
K = 1 0 4 ( 2.17 e 2 + 1.274 α 2 7.653 e 1.01 α + 9991 )
where e ∈ [2, 6] and α ∈ [0°, 20°].
According to the above analysis, the forced convection heat transfer for stratospheric airship involved flight state can be calculated by the following empirical correlation
N u = 0.0161 K R e 0.8543 e 0.0454
The correctness of Formula (9) is likewise validated, and the results are listed in Table 5.
As mentioned above, for the external forced convection heat transfer of an ellipsoidal airship flying with a certain angle of attack, there is no specific correlation available yet. Most of the researchers conducted their researches without considering the effect of flight state. Figure 9 compares the values of the average Nusselt number obtained from Formula (9) with those taken from the literatures [11,30,39]. The geometry models used in these literatures are the same as or close to the one used in present study. They are all ellipse. It can be noted that the results of the present study show a consistent performance trend with literatures. Generally, the present results with α = 20° are higher than those of existing correlations, especially at a higher Reynolds number, where the difference reaches a maximum of about 30 percent compared with Dai. In the thermal design period, the effect of flow state cannot be ignored.

4. Conclusions

A numerical survey is conducted to explore the forced convection heat transfer of a full-scale ellipsoidal airship. The effects of angle of attack, Reynolds number and length to diameter ratio on heat transfer are investigated. An empirical correlation describing the heat transfer of an ellipsoidal airship was developed. Several conclusions were drawn after analyzing the results:
(1)
The change of flow field structure around the airship is the root reason for the change of forced convection heat transfer.
(2)
The average Nusselt number increases with the increase of Re (1 × 107 to 2.0 × 108), decreases with the increase of e (2 to 6), and increases with the increase of α (0 to 20°).
(3)
In the present study, the results show a similar performance trend with the existing correlations, and at higher Reynolds number, the results of the present study are higher by 30 percent than those of Dai with α = 20°.
(4)
K increases with the increase of e and α, and the results with α = 20° have about a 6 percent difference than that without angle of attack. Flight state is an important factor to be considered in the thermal design period of an ellipsoidal airship.

Author Contributions

Conceptualization, Y.J.; Methodology, H.S. and H.P.; Software, B.K; Validation, H.P.; Formal analysis, H.P.; Investigation, H.P.; Resources, B.K.; Data curation, B.K.; Writing—Original draft preparation, B.K.; Writing—Review and editing, H.P.; Visualization, H.P.; Supervision, H.S.; Project administration, Y.J.; Funding acquisition, Y.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Priority Academic Program Development of Jiangsu Higher Education Institutions.

Acknowledgments

I am pleased to acknowledge Tong Zhang for her guidance in English writing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Santapietro, J.J. Persistent wide area surveillance from an airship. IEEE Aerosp. Electron. Syst. Mag. 2012, 27, 11–16. [Google Scholar] [CrossRef]
  2. Izet-Ünsalan, K.; Ünsalan, D. A Low Cost Alternative for Satellites-Tethered Ultra-High Altitude Balloons. In Proceedings of the 5th International Conference on Recent Advances in Space Technologies-RAST2011, Istanbal, Turkey, 9–11 June 2011. [Google Scholar]
  3. Lee, M.; Smith, I.; Androulakakis, S. The High-Altitude Lighter than Air Airship Efforts at the US Army Space and Missile Defense Command/Army Forces Strategic Command. In Proceedings of the 18th AIAA Lighter-than-air Systems Technology Conference, Washington, DC, USA, 4–7 May 2009. [Google Scholar]
  4. Sarafraz, M.M.; Hormozi, F. Intensification of forced convection heat transfer using biological nanofluid in a double-pipe heat exchanger. Exp. Therm. Fluid Sci. 2015, 66, 279–289. [Google Scholar] [CrossRef]
  5. Peyghambarzadeh, S.M.; Sarafraz, M.M.; Vaeli, N.; Ameri, E.; Vatani, A.; Jamialahmadi, M. Forced convective and subcooled flow boiling heat transfer to pure water and n-heptane in an annular heat exchanger. Ann. Nucl. Energy 2013, 53, 401–410. [Google Scholar] [CrossRef]
  6. Moore, D.; Newport, D.; Egan, V.; Lacarac, V.J.A.T.E. Ventilation and internal structure effects on naturally induced flows in a static aircraft wing. Appl. Therm. Eng. 2012, 32, 49–58. [Google Scholar] [CrossRef]
  7. Petersen, D.; Rolfes, R.; Zimmermann, R. Thermo-mechanical design aspects for primary composite structures of large transport aircraft. Aerosp. Sci. Technol. 2001, 5, 135–146. [Google Scholar] [CrossRef]
  8. Storch, D.R. Systems and Methods for A Passive, Forced Convection Cooling System. U.S. Patent Application 12/116786, 12 October 2009. [Google Scholar]
  9. Karabelas, S.J.; Markatos, N.C. Water vapor condensation in forced convection flow over an airfoil. Aerosp. Sci. Technol. 2008, 12, 150–158. [Google Scholar] [CrossRef]
  10. Sarafraz, M.M.; Hormozi, F. Forced convective and nucleate flow boiling heat transfer to alumina nanofluids. Period. Polytech. Chem. Eng. 2014, 58, 37–46. [Google Scholar] [CrossRef] [Green Version]
  11. Kreith, F.; Kreider, F. Numerical Prediction of the Performance of High Altitude Balloons; NCAR Technical Note NCAR-IN/STR-65; NCAR: Boulder, CO, USA, 1974. [Google Scholar]
  12. Yao, W.; Lu, X.; Wang, C.; Ma, R. A heat transient model for the thermal behavior prediction of stratospheric airships. Appl. Therm. Eng. 2014, 70, 380–387. [Google Scholar] [CrossRef]
  13. Wu, J.; Fang, X.; Wang, Z.; Hou, Z.; Ma, Z.; Zhang, H.; Dai, Q.; Xu, Y. Thermal modeling of stratospheric airships. Prog. Aerosp. Sci. 2015, 75, 26–37. [Google Scholar] [CrossRef]
  14. Alam, M.I.; Pant, R.S. A multi-node model for transient heat transfer analysis of stratospheric airships. Adv. Space Res. 2017, 59, 3023–3035. [Google Scholar] [CrossRef]
  15. Fang, X.; Wang, W.; Li, X. A Study of Thermal simulation of stratospheric airships. Spacecr. Recovery Remote Sens. 2007, 28, 5–9. [Google Scholar]
  16. Wang, Y.; Yang, C. A comprehensive numerical model examining the thermal performance of airships. Adv. Space Res. 2011, 48, 1515–1522. [Google Scholar] [CrossRef]
  17. Dai, Q.; Fang, X.; Xu, Y. Numerical study of forced convective heat transfer around a spherical aerostat. Adv. Space Res. 2013, 52, 2199–2203. [Google Scholar] [CrossRef]
  18. Alam, M.I.; Pant, R.S. Modeling Transient Heat Transfer in Stratospheric Airships. In Proceedings of the 15th AIAA Aviation Technology, Integration, and Operations Conference, Washington, DC, USA, 22–26 June 2015. [Google Scholar]
  19. Xia, X.L.; Li, D.F.; Sun, C.; Ruan, L.M. Transient thermal behavior of stratospheric balloons at float conditions. Adv. Space Res. 2010, 46, 1184–1190. [Google Scholar] [CrossRef]
  20. Whitaker, S. Forced convection heat transfer correlations for flow in pipes, past flat plates, single cylinders, single spheres, and for flow in packed beds and tube bundles. Aiche J. 1972, 18, 361–371. [Google Scholar] [CrossRef]
  21. Shi, H.; Zhang, T.; Gao, Z.; Pei, H. Research on fixed point thermal characteristics of a new stratospheric airship. J. Mech. Eng. 2018, 54, 154–161. [Google Scholar] [CrossRef]
  22. Shi, H.; Geng, S.; Qian, X. Thermodynamics analysis of a stratospheric airship with hovering capability. Appl. Therm. Eng. 2019, 146, 600–607. [Google Scholar] [CrossRef]
  23. Wu, J.; Ma, Z.; Hou, Z.; Liu, Z. Numerical research on forced convective heat transfer of stratospheric airship. J. Natl. Univ. Def. Technol. 2016, 38, 177–182. [Google Scholar]
  24. Liu, Q.; Xu, S.; Zhong, W.; Yang, J.; Liu, M. A heat transfer empirical correlation for the stratospheric airship. J. Univ. Sci. Technol. China 2013, 43, 387–392. [Google Scholar] [CrossRef]
  25. Riemann, B. Ueber Die Hypothesen, Welche der Geometriezu Grunde Liegen; University of Gottingen: Gottingen, Germany, 1854. [Google Scholar]
  26. Snyder, W.H. Similarity Criteria for the Application of Fluid Models to the Study of Air Pollution Meteorology. Bound. Layer Meteorol. 1972, 3, 113–134. [Google Scholar] [CrossRef]
  27. Townsend, A.A.R. The Structure of Turbulent Shear Flow, 2rd ed.; Cambridge University Press: Cambridge, UK, 1980; p. 200. [Google Scholar]
  28. Kishore, N.; Gu, S. Momentum and heat transfer phenomena of spheroid particles at moderate Reynolds and Prandtl numbers. Int. J. Heat Mass Transf. 2011, 54, 2595–2601. [Google Scholar] [CrossRef]
  29. Karabelas, S.J.; Koumroglou, B.C.; Argyropoulos, C.D.; Markatos, N.C. High Reynolds number turbulent flow past a rotating cylinder. Appl. Math. Model. 2012, 36, 379–398. [Google Scholar] [CrossRef]
  30. Dai, Q.; Fang, X. Numerical study of forced convective heat transfer around airships. Adv. Space Res. 2016, 57, 776–781. [Google Scholar] [CrossRef]
  31. Amchenbach, E. Experiments on the flow past spheres at very high Reynolds numbers. Fluid Mech. 1972, 54, 565–575. [Google Scholar] [CrossRef]
  32. Hsu, C.J. Heat transfer to liquid metals flowing past spheres and elliptical-rod bundles. Int. J. Heat Mass Transf. 1964, 8, 303–315. [Google Scholar] [CrossRef]
  33. Sideman, S. The equivalence of the penetration and potential flow theories. Ind. Eng. Chem. 1966, 58, 54–58. [Google Scholar] [CrossRef]
  34. Melissari, B.; Argyropoulos, S.A. Development of a heat transfer dimensionless correlation for spheres immersed in a wide range of Prandtl number fluids. Int. J. Heat Mass Transf. 2005, 48, 4333–4341. [Google Scholar] [CrossRef]
  35. Lee, S. A numerical study of the unsteady wake behind a sphere in a uniform flow at moderate Reynolds numbers. Comput. Fluids 1999, 29, 639–667. [Google Scholar] [CrossRef]
  36. Leder, A.; Geropp, D. Unsteady flow structure in the wake of the sphere. Int. Soc. Opt. Eng. 1993, 2052, 119–126. [Google Scholar]
  37. Pao, H.P.; Kao, T.W. Vortex structure in the wake of a sphere. Phys. Fluids 1977, 20, 187. [Google Scholar] [CrossRef]
  38. Li, Y.; Yang, Y.; Zhou, J.; Wang, S. Analysation of changes of pitch angle during formatted launch process of stratospheric airship. Comput. Simul. 2014, 31, 45–49. [Google Scholar]
  39. Holman, J.P. Heat Transfer, 8th ed.; Mc Graw-Hill Book Co.: New York, NY, USA, 1997; pp. 282–283. [Google Scholar]
Figure 1. Computational domain and boundary conditions.
Figure 1. Computational domain and boundary conditions.
Applsci 10 01294 g001
Figure 2. (a) Sketch map of the external domain mesh; (b) Sketch map of the internal domain mesh.
Figure 2. (a) Sketch map of the external domain mesh; (b) Sketch map of the internal domain mesh.
Applsci 10 01294 g002
Figure 3. Comparison of pressure coefficients between literature and the present study.
Figure 3. Comparison of pressure coefficients between literature and the present study.
Applsci 10 01294 g003
Figure 4. The relation of Nu vs. Re, e, and α.
Figure 4. The relation of Nu vs. Re, e, and α.
Applsci 10 01294 g004
Figure 5. Velocity contours and local velocity vectors at α = 0° and α = 20°.
Figure 5. Velocity contours and local velocity vectors at α = 0° and α = 20°.
Applsci 10 01294 g005
Figure 6. Fitting points and fitting surface.
Figure 6. Fitting points and fitting surface.
Applsci 10 01294 g006
Figure 7. Correction coefficient K vs. α (Re = 1.0 × 107).
Figure 7. Correction coefficient K vs. α (Re = 1.0 × 107).
Applsci 10 01294 g007
Figure 8. Correction coefficient K vs. Re (e = 2).
Figure 8. Correction coefficient K vs. Re (e = 2).
Applsci 10 01294 g008
Figure 9. Comparison of values of the Nusselt number from the present study with α = 20°, with those from the literatures.
Figure 9. Comparison of values of the Nusselt number from the present study with α = 20°, with those from the literatures.
Applsci 10 01294 g009
Table 1. Values of L, e, and D.
Table 1. Values of L, e, and D.
L = 100/(m)
E = L/D22.5346
D/(m)504033.332516.67
Table 2. Values of different parameters.
Table 2. Values of different parameters.
ParametersValues
12345
Re (×108)0.10.5751.051.532.0
e22.5345
α/(°)05101520
Table 3. Values of fitting parameters.
Table 3. Values of fitting parameters.
c1c2c3
0.01610.8543−0.0454
Table 4. Validation of Formula (6).
Table 4. Validation of Formula (6).
eReCalculated Results by Formula (6)Numerical Results by CFDError/(%)
2.21 × 10714,83814,9891.01
3.755.75 × 10764,54363,6851.33
4.255.75 × 10764,17763,2021.52
5.32.0 × 108184,29518,49790.37
Table 5. Validation of Formula (9).
Table 5. Validation of Formula (9).
eReα/(°)Calculated Results by Formula (9) Numerical Results by CFDError/(%)
2.21 × 107514,85515,0771.47
3.755.75 × 1071065,25464,4101.31
4.255.75 × 1071064,91565,9231.53
5.32.0 × 10820193,52418,7922.98

Share and Cite

MDPI and ACS Style

Pei, H.; Kong, B.; Jiang, Y.; Shi, H. Forced Convection Heat Transfer for Stratospheric Airship Involved Flight State. Appl. Sci. 2020, 10, 1294. https://doi.org/10.3390/app10041294

AMA Style

Pei H, Kong B, Jiang Y, Shi H. Forced Convection Heat Transfer for Stratospheric Airship Involved Flight State. Applied Sciences. 2020; 10(4):1294. https://doi.org/10.3390/app10041294

Chicago/Turabian Style

Pei, Houju, Benben Kong, Yanlong Jiang, and Hong Shi. 2020. "Forced Convection Heat Transfer for Stratospheric Airship Involved Flight State" Applied Sciences 10, no. 4: 1294. https://doi.org/10.3390/app10041294

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop