Next Article in Journal
Study on Support Mechanisms for Renewable Energy Sources in Poland
Previous Article in Journal
Centralized Multiple Back-Up Protection Scheme with Sharing Data between Adjacent Substations Based on IEC 61850
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transcritical Behavior of Methane in the Cooling Jacket of a Liquid-Oxygen/Liquid-Methane Rocket-Engine Demonstrator

CIRA—Centro Italiano Ricerche Aerospaziali-Via Maiorise, 81043 Capua, Italy
*
Author to whom correspondence should be addressed.
Submission received: 16 May 2022 / Revised: 1 June 2022 / Accepted: 2 June 2022 / Published: 7 June 2022

Abstract

:
The successful design of a liquid rocket engine is strictly linked to the development of efficient cooling systems, able to dissipate huge thermal loads coming from the combustion in the thrust chamber. Generally, cooling architectures are based on regenerative strategies, adopting fuels as coolants; and on cooling jackets, including several narrow axial channels allocated around the thrust chambers. Moreover, since cryogenic fuels are used, as in the case of oxygen/methane-based liquid rocket engines, the refrigerant is injected in liquid phase at supercritical pressure conditions and heated by the thermal load coming from the combustion chamber, which tends to experience transcritical conditions until behaving as a supercritical vapor before exiting the cooling jacket. The comprehension of fluid behavior inside the cooling jackets of liquid-oxygen/methane rocket engines as a function of different operative conditions represents not only a current topic but a critical issue for the development of future propulsion systems. Hence, the current manuscript discusses the results concerning the cooling jacket equipping the liquid-oxygen/liquid-methane demonstrator, designed and manufactured within the scope of HYPROB-NEW Italian Project. In particular, numerical results considering the nominal operating conditions and the influence of variables, such as the inlet temperature and pressure values of refrigerant as well as mass-flow rate, are shown to discuss the fluid transcritical behavior inside the cooling channels and give indications on the numerical methodologies, supporting the design of liquid-oxygen/liquid-methane rocket-engine cooling systems. Validation has been accomplished by means of experimental results obtained through a specific test article, provided with a cooling channel, characterized by dimensions representative of HYPROB DEMO-0A regenerative combustion chamber.

1. Introduction

Technologies based on the oxygen/methane couple in the field of rocket-engine applications are attracting more and more attention in the academic community as well as industrial players in the last twenty years. In fact, there are several well-known reasons, still pushing research and economic efforts into developing and consolidating methane-fuelled space propulsion systems [1,2,3,4]. Although specific impulse is generally lower than hydrogen-based systems, methane exhibits some advantages such as storability, relative ease of handling, and soft cryogenic skills, due to the higher density and critical temperature values, respectively; if compared to rocket-grade kerosene, methane results are cleaner since it does not produce residues (e.g., coking) during the combustion, and this represents also a key factor for developing reusable rockets [5,6,7,8,9,10]. Given these reasons, the world-famous company SpaceX chose methane-based technology, concentrating investments on Raptor’s reusable engines [11].
In summary, methane-based propulsion systems display high specific impulse and thrust-to-weight ratio performances, good capabilities in terms of reusability and throttle-ability, as well as fewer concerns in terms of pollution impact on ground, atmosphere, and space, and reduced costs. Moreover, another strong point has been recently underlined since methane could be “extracted” on Mars in a view of in situ resource utilization framework, and this results into a desirable feature for ascent/descent applications in future planetary-exploration missions [12,13]. In fact, the flexibility of application makes methane an interesting solution for several missions, including access to space (from first stages of launchers to upper stages) and in-space systems (such as service modules, landing or descent vehicles, and ascent stages) [14]. Moreover, Italy is investing in the development and consolidation of methane-based space propulsion systems both on the private and public sides. The Italian Ministry of University and Research is directly funding the HYPROB (HYdrocarbon PROpulsion test Bench) Program, assigned the Italian Aerospace Research Centre (CIRA). The main goals consist of enabling and enriching the national system capabilities for the development of liquid rocket engines (LRE) and subsystems with a specific attention to the liquid-oxygen/liquid-methane couple [15]. The most significant goal is the design, realization and testing of a liquid-oxygen/liquid-methane 30 kN-thrust demonstrator (DEMO-0A), after pursuing an incremental approach strategy, contemplating the test of in-house-designed breadboards conceived to deepen the comprehension of basic physical phenomena, and critical aspects regarding Lox/LCH4 supercritical combustion [16].
The present paper aims at describing the behavior of methane flowing as refrigerant inside a DEMO-0A cooling jacket, representative of typical LRE regenerative cooling systems, by means of 3D numerical simulations. The investigation takes into account different values of inlet conditions in terms of fluid temperature, pressure and mass flow rate in order to analyze the influence on the thermal and fluid-dynamic response of such systems and describe the coolant behavior.

2. Transcritical Behavior of Methane in a LRE Cooling Jacket

The core component of liquid rocket engines (LREs) is represented by thrust chambers that may exhibit combustion-gas temperatures up to 3600 K and pressure values of tens of bar, according to their characteristics in terms of propellants, geometry, and injection strategy. Thus, huge thermal and mechanical stresses must be managed in order to fulfil life-cycle requirements as well as performance goals [17,18]. Given these impressive operative conditions, the design of efficient cooling systems, surrounding the thrust chambers, is mandatory and the development of advanced cooling jackets, made up by narrow axial channels (Figure 1a), is linked with the comprehension of the coolant behavior. In fact, with the exception of kerosene-based engines, most LREs adopt cryogenic propellants as coolant, and this requires a deepening of the analysis of coolant behavior, caused by very different thermophysical properties according to a fixed section of the cooling jacket. Nowadays, although huge efforts have been spent, this kind of investigations remains current, especially for methane-based LREs. Indeed, data under theoretical, numerical and experimental points of view are still needed, if compared to other cryogenic propellants, such as hydrogen, to validate engineering tools and numerical codes [19,20,21]. In the case of methane, evolving in a regenerative cooling jacket, the refrigerant is injected in liquid phase, but due to the huge thermal load from combustion gases, it typically shows a transcritical behavior [22]. According to the regenerative counter-flow cooling strategy given in Figure 1b, the refrigerant (LCH4) is inserted in the cooling channels as a compressed liquid (A), such as at temperature and pressure conditions characterized by values lower and higher than the critical values (Tcr = 190.56 K and Pcr = 4.59 MPa), respectively; then, it is heated by the heat, released by combustion gases, and exhibits a sort of phase “pseudo-change” in the cooling jacket, since both critical temperature and pseudocritical conditions are generally reached and overcome. Thus, moving from the nozzle region, a sort of stratification is gradually observed and the fluid layers (near the bottom part of the channel), behaving as a gas, coexist with fractions of fluid (in the upper part) still remaining in liquid-like conditions (B). Finally, the coolant comes out from the cooling jacket at conditions typically characterized by both temperature and pressure values higher then critical ones, and according to the cooling-system characteristics, even as a supercritical gas. Then, methane is injected into the combustor chamber by means of the injectors (C) [16,23].
Hence, the fluid experiences a sort of “pseudo-change” phase, flowing from the inlet towards the outlet section and changing from a liquid-like condition to a vapor-like one [24]. In other words, thermophysical properties (density, viscosity, specific heat, thermal conductivity, etc.) show impressive variations near the critical conditions (i.e., transcritical or near-critical conditions), as pointed out by Figure 2. However, it is important to remark the differences of behavior with respect to the liquid-to-vapor phase change, observed at subcritical pressure, which is characterized by dramatic sudden abrupt variations [25]. For the comprehension of the transcritical behavior of methane inside the cooling jacket, the pseudo-critical temperature (Tpc > Tcr) should also be considered. It represents the temperature corresponding to the condition, characterized by the maximum value of the specific heat at a specific pressure [26], because can be sometimes associated with a phenomenon called heat-transfer deterioration, strictly linked to the thermal layers occurring in a cross section [26,27,28,29,30,31,32,33]. The pseudocritical temperature value raises as a function of pressure increasing, and in parallel, the specific heat peak tends to decrease. Around the pseudocritical conditions, heat-transfer efficiency may decrease dramatically because a gas-like thin film (featured by low thermal conductivity values), developing near the bottom surface of the channels, reduces the heat transferred from the heated wall to the liquid-like core [30]. Thus, designers used to pay a lot of attention in properly conceiving the cooling passages and took sufficient margins in order to overcome possible undesired problems [31,34]. High values of channel roughness seem to suppress the deterioration mode or to mitigate its effects because of heat-transfer exchange increasing. Moreover, the use of artificial roughness may prevent the heat-transfer deterioration: as a result, wall temperature values decrease but pressure losses may significantly increase [35]. Other authors have also pointed out the effects of different configurations of inlet/outlet manifolds on the behavior of the cooling jacket [36].
It is worth highlighting that the application of the classical semiempirical correlations, adopted to calculate the evaluation of heat-transfer coefficients, does not produce reliable results in the deteriorated mode, and in addition, high values of relative roughness also represent a challenging condition from the numerical-modeling point of view. Thus, CIRA has designed, realized, and tested a dedicated breadboard, the MTP-BB (Methane Thermal Properties Breadboard) in the framework of HYPROB Program [15,37]. Experimental results were useful to collect data on transcritical behavior of methane in a relevant domain and operation, since the breadboard is representative of an LRE cooling channel in terms of geometry and thermal conditions. Moreover, a validated numerical procedure has been set up in view of being adopted to run the thermal simulations on the HYPROB final Demonstrator first unit (DEMO-0A), supporting the detailed design. Further information on the program and on the realized breadboards are available in [14,16,38].

3. Geometrical Description

The final HYPROB demonstrator is a ground engine, including a regeneratively-cooled thrust chamber characterized by a typical counterflow architecture, such as the fuel/coolant entering the cooling system from the nozzle region and exiting in correspondence with the injector head. Figure 3 depicts the assembly with the main components: the most important parts such as the igniter, the injector head, the manifolds, cooling jacket, and combustion chamber. Some details on the engine performances are shown in Table 1. Several narrow axial channels compose the cooling systems. Each channel is generated through the copper-alloy liner, the pure-copper side ribs and the upper surface manufactured by means of electrodeposition. The close-out part is made up of electrodeposited pure nickel.
CIRA has developed a one-dimensional in-house tool with two-dimensional corrections in order to conduct the preliminary design phase of the thrust chamber and related cooling jacket [22]. With reference to Figure 4 and Figure 5, the design choice is based on imposing a constant number of channels and rib width (w) and varying the rib height (h). In this way, the optimization of cooling performances was performed focusing on the three significant sections, highlighted in Figure 5, such as the nozzle (NZ), throat (CT) and cylindrical part (CP). The considered nondimensional geometric parameters, based on reference length L (the overall thrust chamber length), are given below:
  • height of channels (h/L) ranges from 0.0018 to 0.0061;
  • width of channels (b/L) ranges from 0.0019 to 0.0107;
  • width of ribs (w/L) equal to 0.0032;
  • thickness of the liner (h1/L) equal to 0.0020;
  • height of the copper layer (hcu/L) equal to 0.0023;
  • height of the nickel close-out layer (hni/L) equal to 0.0034.
Data reduction was obtained by means of Reynolds number, Nusselt number, Prandtl number, friction factor, and Colburn number. In addition, dimensionless variables were considered, such as temperature, density, viscosity, thermal conductivity, and specific heat. Their formulae are reported in the following:
Re = ρ u a v d h μ
N u = q ˙ d h ( T w T b , f ) λ f
Pr = μ c p λ
T = T T c r
f = 2 Δ P d h L ρ u a v 2
j = N u R e P r 1 / 3
ρ = ρ ρ c r
c p = c p c p , i n
λ = λ λ i n
μ = μ μ i n
where uav is the average velocity, dh is the hydraulic diameter of the channel, and q ˙ is the heat flux evaluated on the channel surface. Moreover, Tw and Tb,f represent the channel surface temperature and the fluid bulk temperature, and ∆P is the pressure difference while Tcr and ρcr are the temperature and density referred to the critical point of methane, respectively. Finally, cp,in, λin and μin are the fluid-specific heat, thermal conductivity, and viscosity at the inlet section.

4. Numerical Modeling

The description of methane transcritical behavior inside a single cooling channel belonging to DEMO was conducted through numerical analyses performed by means of ANSYS Fluent v17 [39]. The governing equations of continuity, momentum, and energy in the 3D form were solved under the hypothesis of steady state, NIST real gas model, and turbulent flow (k-ω sst turbulence model was adopted [40]). The thermo-physical properties of methane have been considered variable and extracted through REFPROP v7.0 database [41]. A steady-state solution and a segregated method were chosen to solve the governing equations. A pressure-based method was considered while a second-order upwind scheme and the SIMPLEC coupling one were chosen for energy and momentum equations and to couple pressure and velocity, respectively [39]. The convergence criteria of 10−6 and 10−8 for the residuals of the velocity components and energy were assumed, respectively. Conduction effects were considered and roughness was imposed to the channel surfaces.
The initialization was accomplished through the inlet-section conditions in terms of fluid temperature, pressure, and single-channel mass-flow rate; their nominal values are 110 K, 16.0 MPa, and 0.02 kg/s (overall mass flow rate is equal to 1.92 kg/s), respectively. In addition, some off-nominal inlet conditions, comprised in a DEMO operating box, were considered to characterize the cooling-jacket behavior at a fluid inlet temperature equal to 100 K and 120 K, inlet pressure equal to 15.7 and 16.3 MPa. Moreover, further mass-flow-rate values, equal to 0.019 and 0.021 kg/s (for each channel), were considered to take into account a possible mass unbalance, due to the not-optimal filling of some channels fed up by the inlet manifold.
The use of the NIST real gas model allowed to manage both the liquid and the vapor phase, and thus describes the transcritical operating conditions of the working fluid. It is important to highlight that a proper use of the real gas also affects the initialization of the simulations and the convergence history. With reference to the aforementioned Figure 4 the considered solid material for the liner part is a copper alloy (CuCrZr) while the electrodeposited layers are made up of pure copper and pure nickel. All the thermophysical properties of such materials were considered temperature-dependent and calculated through a dedicated characterization activity [14]. Results were in line with literature data, as shown in [42]. The computational domain includes both the solid and fluid parts. A half channel was considered to reduce the computational effort, taking advantage of the geometrical and thermo-fluid dynamic symmetry. The following boundary conditions have been assigned, as described by Figure 6:
inlet section: uniform velocity (at a fixed mass flow) and uniform temperature profile;
outlet section: pressure outlet static;
channel surfaces: velocity components equal to zero;
liner surface: input heat flux as depicted by Figure 5;
upper and lateral surfaces: heat flux equal to zero.
The liner surface is heated by the input heat flux, depicted in Figure 5, representing the nominal profile. It derived from CFD reactive simulations on the DEMO combustion chamber side. A constant wall-temperature value of 300 K was imposed [23]. This conservative approach was considered useful to assess the design of the cooling-jacket configurations; then, a weak thermal coupling was also considered for further analyses, as suggested by [22], and to accomplish the comparisons given in the “Results and Discussion” section. Thus, a boundary condition in terms of convective heat-transfer coefficient of the hot-gas side was calculated by properly scaling the “design” nominal heat flux (calculated at Thg = 300 K) through the average hot-gas temperature inside the chamber, evaluated by means of RPA code [43]. The following formula, indeed, has been applied for some significant sections (C, I, O, T), given in Figure 5.
h c , h g = q n o m . / ( T a w T h g )
With Taw equal to 3543 K (outlet section—O), 3538 K (beginning of chamber cylindrical part—C), 3370 K (throat—T), 2456 K (Nozzle exit—I) and Twg = 300 K. Intermediate values between two significant sections were calculated by a linear interpolation according to the values given above.
A mesh-sensitivity analysis was accomplished considering the nominal conditions (nominal input heat flux, Tin = 110 K, Pin = 16.0 MPa and m = 0.01 kg/s—half channel). Three structured mesh distributions, generated by following the mesh suggestions given by Ansys Fluent user’s guide in the case of rough channel walls [39], were considered: they have about 1.3, 2.7, and 5.6 million of nodes, respectively. Finally, the second grid case ensured a good compromise between the machine computational time and the accuracy requirements, as shown by results given in Table 2. Figure 7b allows to give an idea on the nodes’ distribution in the axial and transverse direction and on the attention paid for meshing expansion/contraction sections and ensuring mesh elements’ continuity.

5. Methodology Validation

The validation of the numerical procedure was performed by considering the modeling applied to the MTP-BB numerical rebuilding campaign [37] (Figure 8). That breadboard was specifically designed and tested to collect data on the transcritical behavior of methane and setting up the numerical approach to be used for the DEMO cooling system. The body is made of a copper alloy, and on the top, a rectangular narrow channel, characterized by dimension comparable to DEMO cooling channels, was milled. The heat flux is imposed to the basement by means of ten electrical cartridges (able to supply a maximum power of 12 kW), located in specific slots. Both geometry and operative conditions are representative of typical ones experienced in the LRE cooling system, and in particular, in the DEMO cooling system.
The test campaign was successfully performed, considering the following test parameters:
  • inlet mass flow rate = 0.015, 0.020, 0.025 kg/s;
  • outlet pressure = 8.0, 10.0, 12.0, 15.0 MPa;
  • inlet fluid temperature = 120, 130, 140 K;
  • impressed electrical power = 0 (cold flow), 12 kW.
Results from the numerical rebuilding activity in terms of fluid temperature, pressure drops, and channel wall-temperature values were in good agreement with the experimental data, as also reported in [38]. In fact, the maximum discrepancies in terms of fluid-temperature differences between inlet and outlet sections and pressure drops are equal to about 1.3% and 4.0%, respectively, while differences in terms of the channel’s bottom surface is maximum 1.0%. Further information is available in [37].

6. Results and Discussion

Results for numerical simulations, considering the nominal operating conditions of the cooling jacket (Tin,f = 110 K, Pin = 16.0 MPa and overall m = 1.92 kg/s, corresponding to 0.02 kg/s for each cooling channel), are pointed out. Furthermore, other working points, included in the operating box of HYPROB DEMO, have been considered since the demonstrator could be tested also under the following conditions, determined by the firing test bench characteristics: inlet fluid-temperature values equal to 110 K ± 10 K and inlet pressure equal to 16.0 MPa ± 0.3 MPa. Furthermore, overall mass-flow-rate values equal to 0.020 kg/s ± 0.001 kg/s were considered to analyze the effect of mass unbalance in the filling of the cooling channels, connected to the inlet manifold. In this way, the effects of different inlet conditions on cooling-jacket response may be evaluated. The test matrix is summarized in Table 3. Results include axial profiles for the liner and channel wall temperature, fluid bulk temperature, local and average convective heat-transfer coefficient, local and average Nusselt number, most significant thermophysical properties, and static pressure. Moreover, temperature field plots, including some transversal slices of significant interest, are presented together with some information about thermophysical properties.

6.1. Effects of Fluid Inlet Temperature

Figure 9 depicts the axial profiles of the hot-gas wall, bottom channel, and fluid bulk temperature, considering different values of inlet fluid-temperature values ranging from 100 K to 120 K (runs 1-4). Relative maxima in terms of liner wall temperature and channel bottom surface temperature are attained at the throat zone (x/L = 0.61) and in the region where of hot gases reattach on the combustion chamber wall (x/L = 0.13); if the nominal conditions are considered (Tin,f = 110 K − Tin,f* = 1.00), the corresponding temperature maximum values are equal to about 530 K and 610 K, respectively. Because slight fluctuations in terms of inlet temperature values as well as pressure may be imposed by the firing test bench, Tin,f is reduced to 100 K (Tin,f* = 0.91) or increased to 120 K (Tin,f* = 1.09). The effect is significant, considering the domain from half of the nozzle region, and becomes more evident at the throat zone, where the differences between the nominal conditions and the other operating points may also attain 25 K because of slight changing in fluid thermophysical properties. For what concerns the fluid bulk-temperature profiles, it may be observed that temperature increases more than linearly in the nozzle region, while in the proximity of throat region, the slope tends to change because the methane begins to locally change its conditions from a “liquid-like” to a “gas-like” fluid: in fact, the liquid methane enters the jacket but reaches the critical temperature (190.56 K) approximately in the throat region. From this section, the fluid tends to behave like a highly compressible fluid near the bottom wall of the channel (the hotter one), and as a liquid moving towards the upper wall (the colder one). Moreover, the temperature seems to increase almost linearly. When the fluid moves towards the outlet section, a larger fraction tends to show a vapor-like behavior. Finally, methane traverses the last part of the cooling jacket, such as the cylindrical part, in supercritical gas conditions, since temperature and pressure are much higher than the critical values.
Regarding the static pressure profiles carried out by Figure 10, it is possible to highlight that pressure drops are very low in the nozzle region since the fluid is still in the liquid phase. However, the most significant raise is observed in the convergent and cylindrical parts, because density tends to decrease very rapidly and velocity increases towards the outlet section. This behavior is slightly more evident if the fluid inlet temperature increases.
Figure 11 describes the local and average profiles of convective heat-transfer coefficient and Nusselt number as a function of axial coordinate. The maximum values of the convective heat-transfer coefficients are attained in the throat section, in line with design-phase choices; then, they tend to increase from x/L = 0.36 towards the exit because of the fluid velocity growth, occurring in the cylindrical part. In this case, the inlet temperature has a significant impact in the nozzle region and on the peak value near the throat, which is the highest if Tin,f = 100 K (Tin,f* = 0.91), corresponding to about 10% higher than the case with Tin,f = 120 K (Tin,f* = 1.09); moreover, for Tin,f = 120 K a slight increase in heat transfer coefficients with respect to the other conditions is observed at the beginning of the cylindrical part because of the acceleration of fluid towards the exit, due to the observed low density values. In order to discuss the Nusselt profiles, depicted in Figure 11b, it is worth to highlight that the flow regime is fully turbulent, as exhibited by the Reynolds number profiles depicted by Figure 12. The maximum values are observed in a section slightly moved towards the convergent with respect to the throat because the dimension of the channel is comparable (hence, the heat-transfer coefficients and velocity) but the thermal conductivity begins to rapidly to decrease, as well as the density. In fact, a relative maximum of Reynolds is also detected at x/L = 0.56.
In fact, considering the Reynolds number profiles carried out by Figure 12, they tend to rapidly increase from the inlet section (where fluid is injected as a compressed liquid) towards the throat region, where thermophysical properties sharply change, the channel cross-section area attains the minimum value, and the heat load reaches its peak. Moreover, near-critical conditions are detected from half-convergent, and this means that viscosity decreases more significantly than density, as depicted in the following figures. Then, Reynolds number keeps very high, moving towards the cooling-jacket exit, where supercritical vapor conditions are observed. However, the maximum values of Reynolds numbers are detected for the condition, characterized by the highest fluid inlet temperature. If the Prandtl number profiles are observed, they tend to decrease from typical values, comparable to liquid in the inlet region, to a plateau, exhibited near the throat where critical temperature is on average reached, as shown by the aforementioned figures. The effect of decreasing the inlet temperature on Pr can be observed in the nozzle region, where values keep higher than the other considered cases as the fluid maintains “liquid-like” conditions for larger parts of the cooling jacket. At the beginning of the convergent part, the Prandtl number tends to slightly decrease, exhibiting values typical of gases.
Figure 13 summarizes the aforementioned results in terms of friction factor and Colburn number. Results for both nondimensional numbers remark the strong influences for the effect of entrance region (x/L = 1), the variation of hydraulic diameter along the cooling jacket, the channel roughness and the properties changing along the channel, which does not allow for the establishment of a real fully developed flow, as underlined in [37] since the fluid near the bottom surface of the channel tends to behave like a gas moving towards the exit.
The aforementioned behavior can be deepened by analyzing the axial profiles of the most significant thermophysical properties, such as density, viscosity, specific heat, and thermal conductivity, carried out by Figure 14. Density, viscosity, and thermal conductivity decrease monotonically at about one order of magnitude towards the outlet; and the lower the inlet fluid temperature, the greater the property value at a fixed section. In near-critical conditions, attained in the divergent-throat region, variations are very noticeable, especially from x/L equal to about 0.60. In particular, viscosity and thermal conductivity values reduce very fast in near-critical conditions; then, because the fluid is composed of a supercritical hot gas in the last part of the cooling-jacket channel, they keep almost constant, as reported in [30,31]. The highest values of specific heat, indicating the pseudocritical conditions, are detected at about x/L = 0.50–0.55. If the inlet temperature increases, peaks of specific capacity tend to move towards the throat section and reach for the highest values. Thus, pseudocritical conditions, which are to be taken into consideration to evaluate possible thermal deterioration modes, are attained at about x/L = 0.52 for the nominal point and x/L = 0.49 and 0.54 for Tin,f = 100 K (Tin,f* = 0.91) and 120 K (Tin,f* = 1.09), respectively.
Figure 15 depicts the dimensionless temperature fields, considering the nominal working conditions, and some transversal slices of interest are highlighted. In particular, seven significant sections are chosen, and their position corresponds to the outlet (x/L = 0.00), reattachment point region (x/L = 0.13), a section where the pseudocritical conditions inside the channel are observed on average (x/L = 0.52), the throat one (x/L = 0.61), a section representative of the divergent region (x/L = 0.75), and the inlet (x/L = 1.00). In this way, together with Figure 16, it is possible to appreciate the most stressed zones, which are concentrated in the throat section and in the cylindrical part of the thrust chamber, and more specifically, near the outlet section in correspondence of the reattachment point (x/L = 0.13). Temperature fields show lower temperature values for the close-out walls, made of copper and nickel, respectively, and located on the top of the jacket; and the gradual heating of the working fluid, flowing towards the cooling-jacket exit. Methane is injected as a compressed liquid (x/L = 1.00) and reaches the critical values in the throat region (x/L = 0.61) on average. However, it is possible to observe the transcritical conditions of methane at x/L = 0.75, where some parts of fluid near the bottom wall of the channel have already reached the critical temperature. Thus, a sort of thermal stratification is evident [20,21,30], as pointed out by Figure 16d: very high gradients are observed near the bottom surface of the channel since some parts of the fluid near the bottom walls as well as side walls tend to behave like a “gas”, while in the upper zones of the channel, where temperature values are lower, methane tends to remain in “liquid-like” conditions. From the throat region, larger parts of fluid begin to behave like a “vapor” and after x/L = 0.52, the fluid is fully composed by a supercritical gas because the temperature values overcome the pseudocritical one all over the fluid domain, as also confirmed by density axial profiles, plotted in Figure 14, and the dimensionless fields reported hereinafter in Figure 17. Fluid exhibits very low values of density and tends to accelerate towards the exit (in the cylindrical part of the thrust chamber, ρ* is almost one order of magnitude lower than the inlet section), partially compensating the effects of less favourable conditions for the heat exchange, given by thermophysical properties characterizing a gas in comparison with a liquid.
Figure 17 integrates the description regarding the behavior of methane inside the cooling jacket through the fields of dimensionless density, specific heat, thermal conductivity, and viscosity. From x/L = 0.61, the fluid core is not only characterized by decreasing values of density but specific heat tends to increase. The highest values in the throat section are attained near the channel surfaces, where relative values of thermal conductivity are also locally detected, while at x/L = 0.52, where pseudocritical conditions are detected, a sort of large bubble can be observed. In the middle of the channel the highest values of specific heat are detected and the whole fluid core is characterized by low thermal conductivity since methane can be considered behaving like a supercritical gas. The simultaneous presence of zones, characterized by low thermal conductivity and high specific heat, could determine a deterioration of the fluid thermal performances and consequent local overheating phenomena, which may be difficultly managed.
In fact, as reported in literature [30], a kind of “thermal barrier” between the fluid in proximity of the hot wall and the upper part of the channel may be observed. However, for the present cooling jacket, local wall-temperature profiles do not show relevant overheating, due to the wall roughness values defined in the design phase and fluid pressure conditions adopted, which are sufficiently far from the critical values [31,37]. As a result, temperature peaks, detected between the throat region and the convergent zone, are typical of LRE cooling jackets and definitely far from the maximum allowable values for CuCrZr materials, as pointed out by Figure 9. At the outlet section of the cooling jacket, methane exits as a supercritical vapor, and in fact, very low thermal conductivity and viscosity values, equal to about 0.3 and 0.2 times the inlet values, are observed.

6.2. Effects of Inlet Pressure and Mass-Flow-Rate Variations

In this subsection, the effects of inlet pressure and mass-flow variations with respect to nominal conditions are discussed. Figure 18a shows that the influence of inlet pressure is slight on wall-temperature results and thermal performances. It is possible to point out that the aforementioned variations in terms of inlet temperature of fluid provide more noticeable effects on thermal performances than variables like the inlet pressure, because the temperature margin on critical values is lower in percentage (for example Tin,f/Tcr equal to about 0.57 on average) while inlet pressure is about three times the critical value. The effects of mass-flow rate changes are more evident as expected and the liner temperature difference between the off-nominal conditions is at most 45 K, as observed in the reattachment region of the cylindrical part, while the fluid bulk temperature exhibits ±10 K at most in the outlet section. However, if the minimum allowed mass-flow rate is imposed the maximum temperature reaches about 645 K, a tolerable value for the thrust chamber under material thermomechanical resistance point of view, while in the throat section, a peak of 560 K is detected.
For what concerns the static pressure axial profiles, given in Figure 19a, the consequent pressure drops slightly increases if the inlet pressure is reduced, since there is a slight decrease in terms of critical-condition gain. This kind of behavior is also observed in Figure 19b if mass-flow rate is reduced. In this case, the fluid tends to be hotter than the other cases at a fixed section and the density decrease balances the lower mass-flow rate, and even pressure drops slightly increase, especially in the last part of the cylindrical section where methane is in supercritical vapor conditions.
Figure 20a and Figure 21a remark that the inlet-pressure changes have a slight impact on both convective heat-transfer coefficients and Nusselt number profiles. However, the maximum values are attained for Pin = 16.3 MPa (Pin* = 1.02) because of advantageous property conditions, since the critical conditions are slightly far away with respect to the nominal and low-pressure off-nominal points.
The effects of mass-flow rate variation are more relevant, as expected, and an increase of about 11% in terms of hc,x is observed in the throat section (x/L = 0.61) if an increase of 5% in terms of methane mass, flowing in the cooling jacket, with respect to the nominal operating point. Differences in terms of heat-transfer coefficients are very significant, not only in the throat region but also from the half part of the convergent zone until the half part of the divergent one. If the cylindrical part is considered, differences become less evident moving towards the channel exit since the effects linked to mass are compensated by the increase in fluid temperature and density decrease. Thus, a difference of only 3% is detected if the case of m = 0.019 kg/s (m* = 0.95) is compared to the one at m = 0.021 kg/s (m* = 1.05).
If the Nusselt number axial profiles are considered, it is observed that peaks are detected at the same section of the nominal point and they correspond to an increase of 5.8% and a decrease of 6.6% if mass flow rate is equal to +5% and −5% with respect the nominal conditions. In the cylindrical part, the higher the mass flow rate, the higher Nux results, and variations of ±5% are detected in correspondence with the reattachment-point abscissa.
The discussion on thermophysical-property profiles along the axial coordinates allows us to justify the aforementioned results. If the pressure inlet changes, slight differences are observed in terms of density, viscosity, and thermal conductivity, as pointed out by Figure 22; while specific heat is the most affected property, and its values are higher if pressure decreases, as expected, moving towards the critical conditions.
If different values of inlet mass-flow rate are considered, density, viscosity, and thermal conductivity also show more remarkable variations, as depicted by Figure 23. The higher the mass-flow rate, the lower the fluid temperature, and consequently those properties exhibit the highest values if m = 0.021 kg/s (m* = 1.05). However, when the whole fluid domain behaves like a supercritical vapor (from about x/L = 0.30), the case with m = 0.019 kg/s (m* = 0.95) tends to show the highest viscosity and thermal conductivity average values with respect to the other cases, as expected for those physical conditions. Another effect of changing the inlet mass-flow rate is the shifting of the pseudocritical conditions inside the channel, which moves to about to x/L = 0.48 and 0.54 if m = 0.021 kg/s (m* = 1.05) and 0.019 kg/s (m* = 0.95) are considered, respectively.

7. Conclusions

The present paper describes the thermal and fluid-dynamics investigations, performed to support the development of a typical liquid-oxygen/liquid-methane rocket engine, belonging to the 30-kN thrust class and developed in the context of the HYPROB Program. The presented analyses allow an investigation on refrigerant behavior and the thermal response of the cooling jacket to combustion processes at high pressure, as well supporting verifications under the thermo-structural point of view, which are necessary to evaluate the lifecycle verifications before performing the firing tests.
Investigations were performed considering the nominal operating conditions, defined by the nominal set of inlet parameters, such as fluid temperature, pressure, and mass-flow rate (Tin,f = 110 K, Pin = 16.0 MPa, and m = 0.02 kg/s). Then, the effects of changes of such inlet variables with Tin,f, Pin, and m ranging from 0.91 to 1.09, 0.98 to 1.02, and 0.95 to 1.05 times the nominal inlet parameters, respectively, were also considered. In this way, an extended description of fluid and cooling-jacket behavior before firing tests can be accomplished. Results point out that temperature maxima are attained in the throat section, where the peak of heat flux is located (x/L = 0.61), and in correspondence with the reattachment point region (x/L = 0.13), where the coolant behaves like a supercritical vapor, due to poor thermal properties. Moreover, the coolant, injected as a compressed liquid, tends to reach the critical temperature values on average in the throat region, but some fractions of fluid, near the bottom parts of the cooling channel, are observed to be in near-critical conditions from the half nozzle (x/L = 0.75). Indeed, an evident thermal stratification inside the channel is observed near the bottom surface of the channel, and some parts of the fluid near the bottom and side walls tend to behave like a “gas”, while in the upper zones of the channel, where temperature values are lower, methane tends to remain in “liquid-like” conditions. Moreover, from that region, a dramatic decrease in terms of density, thermal conductivity, and viscosity is observed, as well as just downstream of the throat section, the Nusselt number reaches its peak in accordance with the Reynolds number relative maximum. The pseudocritical conditions are detected at about half a cooling jacket, and then, the fluid is expected to be completely in vapor conditions, as confirmed by the axial profiles and fields, given for the thermophysical properties. At x/L = 0.52, large fractions of the fluid exhibit very high values of specific heat and simultaneously also low values of thermal conductivity. However, no significant overheat phenomena, partly due to the choice of working at pressure levels three times greater than critical values, are observed, neither in that zone nor in other parts of the jacket, where largely tolerable liner temperature values are detected. Finally, methane exits as a supercritical vapor since density values are about one order of magnitude lower than inlet values (with a consequent increase in terms of pressure losses in the last part of the cooling system) as well as for thermal conductivity and viscosity.
If inlet conditions detach from nominal ones—as an effect of test-bench-scattered parameters, for example—the effects connected to changes in terms of fluid inlet temperature are more evident on thermal performances than variations of inlet pressure. The temperature margin on critical values is lower in percentage if compared with pressure ones, and this further influences the thermophysical properties. The effect on the liner-wall temperature results is significant, considering the domain from half of the nozzle region, and becomes more evident at the throat zone, where the differences between the nominal conditions and the other operating points may also attain 25 K because of slight changes in fluid thermophysical properties. For what concerns mass-flow-rate changes, if some imbalance among the channels could happen as a result of a not-optimal filling, their influence is evident and the liner-temperature difference between the off-nominal conditions is at most 45 K, as observed in the reattachment region of the cylindrical part, while the fluid bulk temperature exhibits ±10 K at most in the outlet section. An increase of about 11% in terms of convective heat-transfer coefficient in the throat section (and about 5% in terms of Nusselt number) can be observed if an increase of 5% in terms of methane mass with respect to nominal flow rate is considered. Slight variations are observed in the cylindrical part instead. Finally, the location, where pseudocritical conditions are detected, tends to move upstream if mass-flow rate decreases.

Author Contributions

Conceptualization, D.R., F.B. and M.F.; methodology, D.R. and F.B.; software, D.R. and F.B.; validation, D.R.; formal analysis, D.R. and F.B.; investigation, D.R.; resources, D.R., F.B. and M.F.; data curation, D.R.; writing—original draft preparation, D.R.; writing—review and editing, D.R., F.B and M.F.; visualization, D.R.; supervision, D.R.; project administration, F.B.; funding acquisition, F.B. All authors have read and agreed to the published version of the manuscript.

Funding

HYPROB Project financed by Italian Ministry of University and Research: D.M. 1090/2008.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

This work was accomplished in the framework of the HYPROB Project, financed by the Italian Ministry of University and Research. The authors would like to thank all the involved colleagues, and in particular, Pasquale Natale, Michele Ferraiuolo, and Daniele Cardillo, for their appreciated efforts.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

bWidth of the cooling channel [m]
BBBreadboard
CFDComputational Fluid-Dynamic
cpSpecific heat [J/kg K]
cp*Dimensionless specific heat [-]
CPCylindrical Part of the thrust chamber
CTConvergent section
dDiameter [m]
DEMODemonstrator
fFriction factor [-]
hHeight of the cooling channel [m]
hcConvective heat-transfer coefficient [W/m2 K]
HYPROBHydrocarbon PROpulsion test Bench
IImpulse [s]
jColburn number [-]
kTurbulence kinetic energy [m2/s2]
l or LLength of test articles [m]
LCH4Liquid methane
LOXLiquid oxygen
LRELiquid rocket engine
mMass-flow rate [kg/s]
m*Mass-flow rate referred to nominal operating conditions [-]
MTPMethane Thermal Properties
NISTNational Institute of Standard and Technology
NZNozzle section
O/FMixture ratio (Oxidizer mass/Fuel mass) [-]
PPressure [Pa]
Pin*Inlet pressure referred to nominal operating conditions [-]
PrPrandtl number [-]
qHeat flux [W/m−2]
sstShear stress transport
TTemperature [K]
T*Dimensionless temperature [-]
Tin,f*Fluid inlet temperature referred to nominal operating conditions [-]
wWidth of the rib [m]
x, y, zSpatial coordinates [m]
Greek symbols
λThermal conductivity [W/m K]
λ*Dimensionless thermal conductivity [-]
μViscosity [Pa s]
μ*Dimensionless viscosity [-]
ρDensity [kg/m3]
ωSpecific dissipation rate of turbulence kinetic energy [s−1]
Subscripts
avaverage
awadiabatic wall
bbulk
cccombustion chamber
chchannel
crcritical
cucopper
ffluid
hhydraulic
hghot gas side
ininlet
lliner
maxmaximum
ninickel
nomnominal
outoutlet
pcpseudocritical
sstatic
spspecific
tthroat
wwall

References

  1. Cornelisse, J.W.; Schoyer, H.F.R.; Walker, K.F. Rocket Propulsion and Spaceflight; Butterworth-Heinemann: Oxford, UK, 2007. [Google Scholar]
  2. Liuzzi, D.; Rudnykh, M.; Drigo, D.; Ierardo, N. Architecture Trade-Off for the VEGA-E Upper Stage LOX/CH4 Engine. In Proceedings of the 7th European Conference for Aeronautics and Space Sciences (EUCASS), Milan, Italy, 1–6 July 2017. [Google Scholar] [CrossRef]
  3. Iannetti, A.; Girard, N.; Tchou-kien, D.; Bonhomme, C.; Ravier, N.; Edeline, E. Prometheus, a LOX/LCH4 Reusable Rocket Engine. In Proceedings of the 7th European Conference for Aeronautics and Space Sciences (EUCASS), Milan, Italy, 1–6 July 2017. [Google Scholar] [CrossRef]
  4. Masuda, I. JAXA’s Current Activities for the Research of a LOX/LCH4 (LNG) Engine. In Proceedings of the Space Propulsion 2016, Rome, Italy, 2–6 May 2016. [Google Scholar]
  5. Schuffet, R.; Maier, M.; Sindiy, O.; Ulrich, C.; Fugger, S. Integrated Modeling and Analysis for a LOx/Methane Expander Cycle Engine: Focusing on Regenerative Cooling Jacket Design. In Proceedings of the 42nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibition, Sacramento, CA, USA, 9–12 July 2006. [Google Scholar] [CrossRef]
  6. Brown, C.D. Conceptual Investigations for a Methane-Fueled Expander Rocket Engine. In Proceedings of the 40th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibition, Fort Lauderdale, FL, USA, 11–14 July 2004. [Google Scholar]
  7. Haeseler, D.; Bombelli, V.; Vuillermoz, P.; Lo, R.; Marèe, T.; Caramelli, F. Green Propellant Propulsion Concepts for Space Transportation and Technology Development Needs. In Proceedings of the 2nd International Conference on Green Propellants for Space Propulsion, Cagliari, Italy, 7–8 June 2004. [Google Scholar]
  8. Ierardo, N.; Cuoco, F.; Accettura, A.; Congiunti, C.; Bruno, C. LOX-Methane Systems for High Thrust LRE. In Proceedings of the 4th International Conference Launcher Technology, Liege, Belgium, 3–6 December 2002. [Google Scholar]
  9. Li, S.; Zhang, Y.D.; Li, Y.; Liao, R.Q. Equilibrium Calculation and Technological Parameters Optimization of Natural Gas Liquefaction Process with Mixed Refrigerant. Int. J. Heat Technol. 2015, 33, 123–128. [Google Scholar] [CrossRef] [Green Version]
  10. Boysan, M.E. Analysis of Regenerative Cooling in Liquid Propellant Rocket Engines. Master’s Thesis, Graduate School Natural and Applied Sciences, Middle East Technical University, Ankara, Turkey, 2008. [Google Scholar]
  11. SpaceX. Available online: http://www.nasaspaceflight.com/2016/10/its-propulsion-evolution-raptor-engine/ (accessed on 15 May 2022).
  12. Percy, T.; Polsgrove, T.; Alexander, L.; Turpin, J. Design and Development of a Methane Cryogenic Propulsion Stage for Human Mars Exploration. In Proceedings of the AIAA SPACE 2016, Long Beach, CA, USA, 13–16 September 2016. AIAA 2016-5492. [Google Scholar] [CrossRef] [Green Version]
  13. Hurlbert, E.A.; Attler, M.J.; Melcher, J.C.; Morehead, R.L. Integrated Pressure-Fed Liquid Oxygen/Methane Propulsion Systems—Morpheus Experience, MARE, and Future Applications. In Proceedings of the 52nd AIAA/SAE/ASEE Joint Propulsion Conference, Salt Lake City, UT, USA, 25–27 July 2016. [Google Scholar] [CrossRef]
  14. Ricci, D.; Battista, F.; Ferraiuolo, M.; Natale, P.; Fragiacomo, M. Development of a Liquid Rocket Ground Demonstrator through Thermal Analyses. Heat Transf. Eng. 2020, 41, 1100–1116. [Google Scholar] [CrossRef]
  15. Salvatore, V.; Battista, F.; Votta, R.; Di Clemente, M.; Ferraiuolo, M.; Roncioni, P.; Ricci, D.; Natale, P.; Panelli, M.; Cardillo, D.; et al. Design and Development of a LOX/LCH4 Technology Demonstrator. In Proceedings of the 48th AIAA/ASME Joint Propulsion Conference & Exhibition, Atlanta, GA, USA, 30 July–1 August 2012. [Google Scholar] [CrossRef]
  16. Battista, F.; Ricci, D.; Natale, P.; Cardillo, D.; Fragiacomo, M.; Ferraiuolo, M.; Borrelli, R.; Salvatore, V. The HYPROB Demonstrator Line: Status of the LOX/LCH4 Propulsion Activities. In Proceedings of the 8th European Conference for Aeronautics and Space Sciences (EUCASS), Madrid, Spain, 1–4 July 2019. [Google Scholar] [CrossRef]
  17. Taler, J.; Taler, D.; Kaczmarski, K.; Dzierwa, P.; Trojan, M.; Jaremkiewicz, M. Allowable Rates of Fluid Temperature Variations and Thermal Stress Monitoring in Pressure Elements of Supercritical Boilers. Heat Transf. Eng. 2019, 40, 1430–1441. [Google Scholar] [CrossRef]
  18. Yang, V.; Habiballah, M.; Hulka, J.; Popp, M. Liquid Rocket Thrust Chambers: Aspects of Modeling, Analysis, and Design; American Institute of Aeronautics and Astronautics: Reston, VA, USA, 2004. [Google Scholar]
  19. Rajagopal, M. Numerical Modeling of Regenerative Cooling System for Large Expansion Ratio Rocket Engines. J. Therm. Sci. Eng. Appl. 2015, 7, 011012. [Google Scholar] [CrossRef]
  20. Trejo, A.G.; Garcia, C.; Choudhuria, A. Experimental Investigation of Transient Forced Convection of Liquid Methane in a Channel at High Heat Flux Conditions. Exp. Heat Transf. 2016, 29, 97–112. [Google Scholar] [CrossRef]
  21. Votta, R.; Battista, F.; Gianvito, A.; Smoraldi, A.; Salvatore, V. Experimental Investigation of Methane in Transcritical Conditions. In Proceedings of the 50th AIAA/ASME Joint Propulsion Conference & Exhibition, Cleveland, OH, USA, 28–30 July 2014. [Google Scholar]
  22. Sutton, G.P.; Biblarz, O.O. Rocket Propulsion Elements; John Wiley & Sons: New York, NY, USA, 2010; ISBN 9780470080245. [Google Scholar]
  23. Ferraiuolo, M.; Ricci, D.; Battista, F.; Roncioni, P.; Salvatore, V. Thermo-Structural and Thermo-Fluid Dynamics Analyses supporting the Design of the Cooling System of a Methane Liquid Rocket Engine. In Proceedings of the 2014 ASME International Mechanical Engineering Congress (IMECE2014), Montreal, QC, Canada, 14–20 November 2014. [Google Scholar] [CrossRef]
  24. Kawashima, H.; Okita, K.; Aoki, K.; Azuma, N.; Kumakawa, A.; Onodera, T.; Yoshida, S.S.; Negishi, H.; Manako, H.; Koganezawa, T. Combustion and Regenerative Cooling Characteristics of LOX/Methane Engine. Trans. Space Technol. Jpn. 2009, 7, 7–11. [Google Scholar] [CrossRef] [Green Version]
  25. Kreith, F.; Manglik, R.M.; Bohn, M.S. Principles of Heat Transfer, 7th ed.; Cengage Learning: Boston, MA, USA, 2010. [Google Scholar]
  26. Polyakov, A.F. Heat Transfer under Supercritical Pressures. Adv. Heat Transf. 1991, 21, 1–53. [Google Scholar]
  27. Ruan, B.; Lin, W. Numerical Investigation on Heat Transfer and Flow Characteristics of Supercritical Methane in a Horizontal Tube. Cryogenics 2022, 124, 103482. [Google Scholar] [CrossRef]
  28. Jackson, J.D.; Hall, W.B. Forced Convection Heat Transfer to Fluids at Supercritical Pressure. Turbul. Forced Convect. Channels Bundles 1979, 2, 563–611. [Google Scholar]
  29. Kurganov, V.A.; Ankudinov, V.B. Calculation of Normal and Deterioration Heat Transfer in Tubes with Turbulent Flow of Liquids in the Near-critical and Vapour Region of State. Teploenergetika 1982, 32, 53–57. [Google Scholar]
  30. Pizzarelli, M.; Urbano, A.; Nasuti, F. Numerical Analysis of Deterioration in Heat Transfer to Near-Critical Rocket Propellants. Numer. Heat Transf. 2010, 57, 297–314. [Google Scholar] [CrossRef]
  31. Negishi, H.; Daimon, Y.; Yamanishi, N.; Ohnishi, Y. Numerical Investigation of Supercritical Coolant Flow in Liquid Rocket Engine. In Proceedings of the 46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference, Nashville, TN, USA, 25–28 July 2010. paper AIAA-2010-6888. [Google Scholar] [CrossRef]
  32. Haemisch, J.; Suslov, D.; Oschwald, M. Experimental Study of Methane Heat Transfer Deterioration in a Subscale Combustion Chamber. J. Propuls. Power 2019, 35, 819–826. [Google Scholar] [CrossRef]
  33. Tripathi, P.M.; Basu, S. Insights into the Dynamics of Supercritical Water Flow using a Two-Phase Approach. Phys. Fluids 2021, 33, 043304. [Google Scholar] [CrossRef]
  34. Zhang, M.; Sun, B. Improved Heat-Transfer Correlation for Transcritical Methane Based on a Velocity Profile Correction Term. J. Therm. Sci. Eng. Appl. 2022, 14, 041002. [Google Scholar] [CrossRef]
  35. Zhang, M.; Sun, B. Effect of Artificial Roughness on Flow and Heat Transfer of Transcritical Methane. Int. J. Therm. Sci. 2020, 158, 106528. [Google Scholar] [CrossRef]
  36. Zhang, M.; Sun, B.; Song, J. Effect of Inlet and Outlet Manifolds on Regenerative Cooling in LOX/Methane Thrust Chambers. J. Therm. Sci. 2020, 30, 517–529. [Google Scholar] [CrossRef]
  37. Ricci, D.; Natale, P.; Battista, F. Experimental and Numerical Investigation on the Behaviour of Methane in Supercritical Conditions. Appl. Therm. Eng. 2016, 107, 1334–1353. [Google Scholar] [CrossRef]
  38. Ricci, D.; Battista, F.; Fragiacomo, M. Numerical Investigation on the Thermal Behaviour of a LOx/LCH4 Demonstrator Cooling System. Aerospace 2021, 8, 151. [Google Scholar] [CrossRef]
  39. Ansys Inc. Ansys Fluent User’s Guide; Version 17.0; Ansys Inc.: Canonsburg, PA, USA, 2017. [Google Scholar]
  40. Menter, F.R. Two Equation Eddy-Viscosity Turbulence Models for Engineering Applications. AIAA J. 1994, 32, 1598–1605. [Google Scholar] [CrossRef] [Green Version]
  41. Lemmon, E.W.; McLinden, M.; Huber, M.L. NIST Standard Reference Database 23: REFPROP 7.0; National Institute of Standards & Technology: Gaithersburg, MD, USA, 2002.
  42. De Groh, H.C., III; Ellis, D.L.; Loewenthal, W.S. Comparison of GRCop-84 to Other Cu Alloys with High Thermal Conductivities. J. Mater. Eng. Perform. 2008, 17, 594–606. [Google Scholar] [CrossRef]
  43. Rocket Propulsion Analyses (RPA). User’s Manual, Version 2.1; Rocket Propulsion Analysis: Cologne, Germany, 2014. [Google Scholar]
Figure 1. Cooling-jacket scheme and coolant behavior: (a) Cross section of a typical LRE cooling jacket; (b) regenerative counter-flow scheme with fluid path and typical behavior along the cooling channels.
Figure 1. Cooling-jacket scheme and coolant behavior: (a) Cross section of a typical LRE cooling jacket; (b) regenerative counter-flow scheme with fluid path and typical behavior along the cooling channels.
Energies 15 04190 g001
Figure 2. Thermophysical property variations of methane, as a function of temperature and pressure: (a) density; (b) specific heat; (c) viscosity; (d) thermal conductivity.
Figure 2. Thermophysical property variations of methane, as a function of temperature and pressure: (a) density; (b) specific heat; (c) viscosity; (d) thermal conductivity.
Energies 15 04190 g002
Figure 3. DEMO assembly including the main components and schematics of electroplating process applied to LRE cooling-jacket manufacturing. 1. Igniter; 2. injector head with oxidizer dome; 3. outlet fuel manifold; 4. combustion chamber, including the cooling jacket; 5. inlet fuel manifold.
Figure 3. DEMO assembly including the main components and schematics of electroplating process applied to LRE cooling-jacket manufacturing. 1. Igniter; 2. injector head with oxidizer dome; 3. outlet fuel manifold; 4. combustion chamber, including the cooling jacket; 5. inlet fuel manifold.
Energies 15 04190 g003
Figure 4. Geometry description of DEMO cooling jacket: cross section.
Figure 4. Geometry description of DEMO cooling jacket: cross section.
Energies 15 04190 g004
Figure 5. Chamber profile, hydraulic diameter of cooling channels and input heat-flux profile along the engine.
Figure 5. Chamber profile, hydraulic diameter of cooling channels and input heat-flux profile along the engine.
Energies 15 04190 g005
Figure 6. Schematics of the cooling jacket: cross-section view, including information on materials and adopted boundary conditions.
Figure 6. Schematics of the cooling jacket: cross-section view, including information on materials and adopted boundary conditions.
Energies 15 04190 g006
Figure 7. Numerical model: (a) View of the single cooling channel; (b) Details of transversal and axial mesh distribution in the throat section (left) and inlet section (right).
Figure 7. Numerical model: (a) View of the single cooling channel; (b) Details of transversal and axial mesh distribution in the throat section (left) and inlet section (right).
Energies 15 04190 g007
Figure 8. MTP breadboard: (a) under test; (b) geometrical information.
Figure 8. MTP breadboard: (a) under test; (b) geometrical information.
Energies 15 04190 g008
Figure 9. Axial profiles of hot-gas-wall temperature, temperature of channel bottom wall and fluid bulk temperature, considering different values of inlet fluid temperature (runs 1-4).
Figure 9. Axial profiles of hot-gas-wall temperature, temperature of channel bottom wall and fluid bulk temperature, considering different values of inlet fluid temperature (runs 1-4).
Energies 15 04190 g009
Figure 10. Axial profiles of static pressure, considering different values of inlet fluid temperature (runs 1-4).
Figure 10. Axial profiles of static pressure, considering different values of inlet fluid temperature (runs 1-4).
Energies 15 04190 g010
Figure 11. Axial profiles of local and average convective heat-transfer coefficients (a), local and average Nusselt number (b), considering different values of inlet fluid temperature (runs 1-4).
Figure 11. Axial profiles of local and average convective heat-transfer coefficients (a), local and average Nusselt number (b), considering different values of inlet fluid temperature (runs 1-4).
Energies 15 04190 g011aEnergies 15 04190 g011b
Figure 12. Axial profiles of Reynolds number and Prandtl number, considering different values of inlet fluid temperature (runs 1-4).
Figure 12. Axial profiles of Reynolds number and Prandtl number, considering different values of inlet fluid temperature (runs 1-4).
Energies 15 04190 g012
Figure 13. Axial profiles of friction factor and Colburn number, considering different values of inlet fluid temperature (runs 1-3).
Figure 13. Axial profiles of friction factor and Colburn number, considering different values of inlet fluid temperature (runs 1-3).
Energies 15 04190 g013
Figure 14. Axial profiles of thermophysical properties, considering different values of inlet fluid temperature (runs 1-4): (a) density and viscosity; (b) specific heat and thermal conductivity.
Figure 14. Axial profiles of thermophysical properties, considering different values of inlet fluid temperature (runs 1-4): (a) density and viscosity; (b) specific heat and thermal conductivity.
Energies 15 04190 g014
Figure 15. Nominal operating point: dimensionless temperature distribution, including some transversal slices.
Figure 15. Nominal operating point: dimensionless temperature distribution, including some transversal slices.
Energies 15 04190 g015
Figure 16. Nominal operating point: dimensionless temperature field, transversal slices at: (a) outlet section, (b) x/L = 0.13 (reattachment-point region), (c) x/L = 0.52 (corresponding to maximum specific heat); (d) x/L = 0.61 (throat section and fluid bulk temperature equal to critical values); (e) x/L = 0.75; (f) inlet section.
Figure 16. Nominal operating point: dimensionless temperature field, transversal slices at: (a) outlet section, (b) x/L = 0.13 (reattachment-point region), (c) x/L = 0.52 (corresponding to maximum specific heat); (d) x/L = 0.61 (throat section and fluid bulk temperature equal to critical values); (e) x/L = 0.75; (f) inlet section.
Energies 15 04190 g016
Figure 17. Nominal operating point: dimensionless fields of density, specific heat, thermal conductivity and viscosity at significant transversal slices—fluid zones.
Figure 17. Nominal operating point: dimensionless fields of density, specific heat, thermal conductivity and viscosity at significant transversal slices—fluid zones.
Energies 15 04190 g017
Figure 18. Axial profiles of hot-gas-wall temperature, temperature of channel bottom surface, and fluid bulk temperature (runs 1, 5-8): (a) effects of inlet pressure; (b) effects of mass-flow rate.
Figure 18. Axial profiles of hot-gas-wall temperature, temperature of channel bottom surface, and fluid bulk temperature (runs 1, 5-8): (a) effects of inlet pressure; (b) effects of mass-flow rate.
Energies 15 04190 g018
Figure 19. Axial profiles of static pressure (runs 1, 5-8): (a) effects of inlet pressure; (b) effects of mass flow rate.
Figure 19. Axial profiles of static pressure (runs 1, 5-8): (a) effects of inlet pressure; (b) effects of mass flow rate.
Energies 15 04190 g019
Figure 20. Axial profiles of local and average convective heat-transfer coefficients (runs 1, 5-8): (a) effects of inlet pressure; (b) effects of mass flow rate.
Figure 20. Axial profiles of local and average convective heat-transfer coefficients (runs 1, 5-8): (a) effects of inlet pressure; (b) effects of mass flow rate.
Energies 15 04190 g020
Figure 21. Axial profiles of local and average convective Nusselt number (runs 1, 5-8): (a) effects of inlet pressure; (b) effects of mass-flow rate.
Figure 21. Axial profiles of local and average convective Nusselt number (runs 1, 5-8): (a) effects of inlet pressure; (b) effects of mass-flow rate.
Energies 15 04190 g021
Figure 22. Axial profiles of thermophysical properties, considering the effects of inlet pressure (runs 1, 5, 6): (a) density and viscosity; (b) specific heat and thermal conductivity.
Figure 22. Axial profiles of thermophysical properties, considering the effects of inlet pressure (runs 1, 5, 6): (a) density and viscosity; (b) specific heat and thermal conductivity.
Energies 15 04190 g022
Figure 23. Axial profiles of thermophysical properties, considering the effects of mass-flow rate (runs 1, 7, 8): (a) density and viscosity; (b) specific heat and thermal conductivity.
Figure 23. Axial profiles of thermophysical properties, considering the effects of mass-flow rate (runs 1, 7, 8): (a) density and viscosity; (b) specific heat and thermal conductivity.
Energies 15 04190 g023
Table 1. Summary regarding the main performance parameters of HYPROB DEMO.
Table 1. Summary regarding the main performance parameters of HYPROB DEMO.
O/F3.4PCC5.5 MPa
Reaction efficiency0.98Isp286 s
Thrust30 kNAcc/At4
Table 2. Results of grid-independence analysis.
Table 2. Results of grid-independence analysis.
Type of MeshΔP
[MPa]
Outlet Fluid Bulk Temperature
Tb,f out [K]
Liner Maximum Temperature
Tw,hg max
Channel Bottom Wall Maximum Temperature
Tw,ch max
1—coarse5.051420.1600.4555.4
2—fine5.094420.4610.8562.7
3—finest5.096420.5611.5563.4
Table 3. Test matrix of the numerical campaign.
Table 3. Test matrix of the numerical campaign.
RunInlet TemperatureInlet PressureMass Flow RateImposed Heat Flux
Tin,f
[K]
Tin,f*
(Tin,f/Tin,f−nom)
Pin
[MPa]
Pin*
(Pin/P,nom)
m
[kg/s]
m*
(m/mnom)
11101.0016.01.000.0201.00Nominal
21101.0016.01.000.0201.00Weak Coupling
31000.9116.01.000.0201.00Nominal
41201.0916.01.000.0201.00Nominal
51101.0015.70.980.0201.00Nominal
61101.0016.31.020.0201.00Nominal
71101.0016.01.000.0190.95Nominal
81101.0016.01.000.0211.05Nominal
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ricci, D.; Battista, F.; Fragiacomo, M. Transcritical Behavior of Methane in the Cooling Jacket of a Liquid-Oxygen/Liquid-Methane Rocket-Engine Demonstrator. Energies 2022, 15, 4190. https://0-doi-org.brum.beds.ac.uk/10.3390/en15124190

AMA Style

Ricci D, Battista F, Fragiacomo M. Transcritical Behavior of Methane in the Cooling Jacket of a Liquid-Oxygen/Liquid-Methane Rocket-Engine Demonstrator. Energies. 2022; 15(12):4190. https://0-doi-org.brum.beds.ac.uk/10.3390/en15124190

Chicago/Turabian Style

Ricci, Daniele, Francesco Battista, and Manrico Fragiacomo. 2022. "Transcritical Behavior of Methane in the Cooling Jacket of a Liquid-Oxygen/Liquid-Methane Rocket-Engine Demonstrator" Energies 15, no. 12: 4190. https://0-doi-org.brum.beds.ac.uk/10.3390/en15124190

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