10.2514@6.2020-3767

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Numerical Analysis of Nozzle Heating and Erosion in Hybrid

Rockets and Comparison with Experiments

Daniele Bianchi∗ , Mario Tindaro Migliorino† , and Marco Rotondi‡


Sapienza University of Rome, 00184 Rome, Italy

Landon Kamps§ , and Harunori Nagata¶


Hokkaido University, Sapporo, 060-8628, Japan

The motor performance loss due to nozzle throat erosion in hybrid rocket engines is an
important feature to be considered during the motor design. In fact, a peculiar character-
istic of hybrid rockets operating conditions is a greater concentration of oxidizing species in
the combustion products with respect to solid rockets, which emphasizes the throat erosion
phenomenon. Hence, the need for an accurate physical understanding of this process is funda-
mental for technology advancement. The aim of this work is to investigate the graphite nozzle
erosion in 2kN-class hybrid rockets burning high-density polyethylene (HDPE) and oxygen.
The main focus of the work is the evaluation of nozzle throat erosion and wall temperature as
a function of motor operating conditions by performing numerical simulations from a proven
in-house CFD-solver. The results obtained from a detailed parametric analysis are discussed
and then used to derive a regression law for nozzle throat erosion. The regression law here
proposed is then validated through comparison with the experimental results obtained from
lab-scale 2kN-thrust class rocket firing tests performed at Hokkaido University.

I. Introduction
ocket nozzle throats are typically subjected to a severe thermal environment, hence a proper thermal protection
R system (TPS) is usually employed to protect the underlying structure from the high incoming heat flux. A widespread
solution is to use ablative passive TPSs, which possible erosion can lead to the nozzle throat enlargement and to the
subsequent motor performance loss. The will to reduce nozzle throat consumption, especially concerning hybrid rockets,
generated a renew interest in experimental tests and numerical investigations to study the effects induced on the motor
thrust curve and specific impulse.
Hybrid propellant rockets, defined as such that one propellant is stored in liquid phase and the other in solid phase,
commonly employ a liquid oxidizer and solid fuel [1]. Different combinations of solid fuels and liquid oxidizers
have been proposed in the literature aiming to achieve the following: i) system flexibility compared to solid rockets
(throttling, extinguishing and multiple ignitions); ii) system simplicity compared to liquid rockets (propellant settling
and control systems are halved); iii) higher specific impulse than solid rockets and higher density impulse than liquid
rockets; iv) inherent safety, due to the separation between fuel and oxidizer; v) reduced development and management
costs; vi) lower pollutant emissions with respect to solid rocket motors, due to the absence of traditional oxidizers
such as ammonium perchlorate. Despite their inherent disadvantages like low regression rate, low bulk density and
sometimes low combustion efficiency, as well as O/F shift [2] the above mentioned features give to hybrid rockets a
great potential to become widely used in the future generation of propulsion systems, largely due to the recent emphasis
on propulsion safety, reliability, low development cost, reduced environmental pollution, and greater operability [3].
Nevertheless, hybrid rocket development has not achieved the same level of maturity as traditional solid and liquid
systems [1]. Therefore, a key element for hybrid technology advancement and success is a better understanding of the
major underlying critical phenomena [3–7]. Among them, an aspect that has not been much dealt with in open literature
is that of nozzle erosion, whose minimization and reduction is one of the challenges in hybrid rocket propulsion [8]
∗ Associate Professor, Dipartimento di Ingegneria Meccanica e Aerospaziale, Via Eudossiana 18, Member AIAA, corresponding author,

daniele.bianchi@uniroma1.it
† Research Fellow, Dipartimento di Ingegneria Meccanica e Aerospaziale, Via Eudossiana 18, Member AIAA
‡ Post-Graduate Fellow, Dipartimento di Ingegneria Meccanica e Aerospaziale, Via Eudossiana 18, Member AIAA
§ Specially appointed Assistant Professor, Faculty of Engineering, Department of Mechanical and Space Engineering, Member AIAA
¶ Professor, Faculty of Engineering, Department of Mechanical and Space Engineering, Member AIAA

1
TPS technologies used in the design of solid rocket motors (SRM) (i.e. ablative cooling) can be successfully
extended to hybrid rockets motors (HRM) applications [9]. However, in doing so, it has to be considered that TPS
operating conditions in hybrid rockets are quite different from those in SRMs. In fact, in hybrid rockets, the ablation
process is exacerbated due to the greater concentration than in solid rockets of oxidizing species in the exhaust products,
which largely affect the material thermochemical behaviour [9, 10]. In particular, it has been shown that throat erosion
rate in a hybrid is generally greater than that of a solid propellant system and is a strong function of chamber pressure
and mixture ratio [1, 10–12].
The numerical evaluation of the erosion behavior of nozzle protection materials relies on a proper description
and modeling of the interaction between the combustion gases arising from hybrid fuels and the protective material.
Currently, the predictions of the nozzle thermal protection material erosion rate are heavily based on semi-empirical
correlations [10]. Such approaches are useful during the preliminary design and analysis process, providing a fast and
simple estimation of the nozzle erosion. Anyway, they are based on a simplified model, which cannot account for the
complex physical and chemical phenomena involved. Therefore, they need to be accurately calibrated with existing
experimental data specific for each motor. Indeed, the extension of such models to new motors which can be different in
scale, geometry, etc. is hardly possible without the availability of experimental data for each motor. The development of
advanced numerical models capable of representing more accurately the physico-chemical interactions between the
reacting flow and the protective material is hence required to provide the kind of quantitative data needed for motor
design and optimization.
The primary objective of this work is to investigate the thermochemical erosion behavior as a function of chamber
conditions (both oxidizer-to-fuel mass ratio and chamber pressure) of graphite nozzles in hybrid rocket engines. To
this end, firing tests from a 2 kN thrust-class CAMUI-type hybrid rocket using liquid oxygen (LOX) as oxidizer and
high-density polyethylene (HDPE) as fuel have been used to experimentally investigate the graphite nozzle erosion
process under various experimental conditions, allowing to extend the analysis performed in [12], including new test
conditions (oxidizer to fuel ratio, chamber pressure, initial throat diameter). Numerical simulations have been carried out
by a validated [11–19] approach which is able to treat in detail the interaction between chemically reacting hot gas and
carbon-based ablative materials. The approach relies on a full Navier-Stokes flow solver coupled with a thermochemical
ablation model which takes into account heterogeneous chemical reactions at the nozzle surface, the rate of diffusion of
the species through the boundary-layer, ablation species injection in the boundary layer, the heat conduction inside the
nozzle material, and variable multispecies thermophysical properties. The gas-solid phase coupling is handled through
a customized gas-surface interaction wall boundary condition that is fully integrated within the CFD solver. Firstly, a
detailed parametric numerical analysis is carried out, trying to thoroughly describe all the complex dependencies on the
main motor operating parameters of HRMs throat ablation. Then, numerical results are used to derive a regression
law for nozzle erosion, able to correctly handle chamber pressure, mixture ratio, and nozzle geometrical effects on the
ablation process with reduced computational efforts. The regression law is then validated through comparison with the
experimental data obtained at the University of Hokkaido. The final goal of this work is to increase the understanding
of the nozzle erosion process in hybrid rockets in order to increase our fundamental knowledge of the physics of the
different phenomena that are involved.

II. Theoretical modeling


In the present formulation, the governing equations for the gas phase flow in the nozzle are the reacting turbulent
Reynolds averaged Navier–Stokes equations with variable thermodynamic and transport properties. The thermodynamic
properties of individual species are approximated by seventh-order polynomials of temperature whereas the transport
properties are approximated by fourth order polynomials [20]. Mixture properties for conductivity and viscosity are
derived from Wilke’s rule. The diffusion model is based on the effective diffusion coefficients. The heat conduction
process in the nozzle is treated as being one-dimensional (1-D) and steady state. The basis of the present analysis is
the comprehensive theoretical model for carbon–carbon and graphite nozzle erosion in SRM and HRE developed and
validated in previous studies [11–15].

A. Nozzle thermochemical erosion


The rocket-nozzle material considered in the present study is graphite. Several processes can affect the nozzle throat
erosion rate at high-pressure and high-temperature conditions. These processes include: strong convective and radiative
heat transfer (leading to a steep rise in surface temperature, often to temperatures over 2500 K), turbulent transport of
gas-phase oxidizing chemical species from the flow to the nozzle surface, heterogeneous reactions of these oxidizing

2
species with the exposed carbon surface, injection of ablation and pyrolysis products in the boundary layer, etc. There is
a strong coupling between the above mentioned processes, hence a proper model must be employed.
The thermochemical ablation model used in this work is modeled via surface mass and energy balances and takes
into account heterogeneous chemical reactions at the nozzle surface, rate of diffusion of the species through the boundary
layer and ablation species injection in the boundary layer. If it is assumed that no material is being removed in a
condensed phase (solid or liquid), then the general conservation laws at the gas–solid interface for a non-decomposing
material can be written as [21] (see Fig.1):

Mass: 𝑚¤ w = (𝜌𝑣)w = 𝜌s · 𝑟¤ (1)


𝜕𝑦 𝑖
Species: 𝜌𝐷 + 𝜔¤ 𝑖 = (𝜌𝑣)w · 𝑦 𝑖w (𝑖 = 1, . . . , 𝑁) (2)
𝜕𝜂
𝑁
𝜕𝑇 Õ 𝜕𝑦 𝑖
Energy: 𝑘 + ℎ𝑖 𝜌𝐷 + 𝑚¤ w ℎs,w + 𝑞 radin = 𝑞 c w + 𝑞 radout + (𝜌𝑣)w ℎw (3)
𝜕𝜂 𝑖=1 𝜕𝜂
where 𝑣 w stands for the radial velocity in the gas phase due to ablation products injection, while 𝜔¤ 𝑖 i is the rate of
production of gas-phase species 𝑖 at the nozzle surface due to heterogeneous reactions. The term 𝑚¤ w ℎs,w is the energy
flux entering the surface due to surface recession while the term (𝜌𝑣)w ℎw is the energy flux leaving the surface due to
blowing. The surface energy balance (Eq.3) can be rewritten in a more compact and useful form using Eq.2 multiplied

(a) Surface mass balance.

(b) Surface energy balance.

Fig. 1 Surface mass and energy balances for an ablating material.

for each 𝑖 species’s wall enthalpy, ℎ𝑖,w , summed over all the species.
𝑁
𝜕𝑇 Õ
𝑘 + 𝑞 radin = ℎ𝑖 𝜔¤ 𝑖 + 𝑚¤ w ℎs,w + 𝑞cw + 𝑞 radout (4)
𝜕𝜂 𝑖=1
|{z} |{z} | {z } |{z} |{z}
Convection Radiation in Solid conduction Radiation out
Ablation

Concerning the solid conduction term, it is assumed that the heat conduction is dominant in the direction normal to
the nozzle surface. In fact, although tangential temperature gradients certainly exist along the nozzle wall, they are
generally negligible if compared with those in the normal direction and can be considered to represent a second-order
effect. Assuming a steady-state condition, the solid conduction term can be expressed as
𝜕𝑇s
𝑞cw = 𝑘 s = 𝑚¤ w (ℎs,w − ℎs,0 ) (5)
𝜕𝜂
where the term ℎs,0 represents the enthalpy of the material at the initial or undisturbed temperature. Note that the
steady-state solid heat conduction flux is only affected by the material specific heat (through the enthalpy difference)
and not by its thermal conductivity, that only influences the in-depth solid temperature profile.

3
Table 1 Heterogeneous rate constants and reaction order of carbon with H2 O, CO2 , OH, and O [22].

Surface reaction 𝑗 𝐴𝑗 𝐸 𝑗 , kJ/mol 𝑏𝑗 𝑛𝑗


Cs + H2 O ⇒ CO + H2 1 4.8 ·105 288 0.0 0.5
Cs + CO2 ⇒ 2 CO 2 9.0 ·103 285 0.0 0.5
Cs + H2 O ⇒ CO + H 3 3.61 ·102 0.0 -0.5 1.0
Cs + O ⇒ CO 4 6.655 ·102 0.0 -0.5 1.0

Table 2 Heterogeneous rate constants of carbon with O2 [22].

Surface reaction 𝑗 𝐴𝑗 𝐸 𝑗 , kJ/mol 𝑏𝑗 𝑛𝑗


5 2.4 ·103 125.6 0.0 0.5
Cs + 21 O2 ⇒ CO 6 2.13 ·101 -17.17 0.0 0.5
7 5.35 ·10−1 63.64 0.0 0.5
8 1.81 ·107 406.1 0.0 0.5

The heterogeneous gas-surface chemical reactions taking place at the nozzle surface are modeled by the graphite
oxidation kinetics proposed by Chelliah for non-porous graphite [22] which is a semi-global heterogeneous reaction
mechanism for graphite oxidation consisting of five reactions, listed in Tables 1 and 2. With this mechanism, the
contribution to erosion due to the 𝑖-th species can be expressed in kg/(m2 · s) as:
𝑛
𝑚¤ w𝑖 = 𝑘 𝑗 𝑝 𝑖 𝑗 for species 𝑖 = H2 O, CO2 , OH and O (6)

𝑘 5 𝑝 𝑖𝑌  𝑘 8  −1
𝑚¤ w𝑖 = + 𝑘 7 𝑝 𝑖 (1 − 𝑌 ) where 𝑌 = 1 + for species 𝑖 = O2 (7)
1 + 𝐾6 𝑝 𝑖 𝑘 7 𝑝𝑖
where 𝑝 𝑖 is the partial pressure (in atm) of the oxidizing species i, nj is the overall order of the heterogeneous 𝑗-th
reaction, and 𝑘 𝑗 is the specific rate constant of the 𝑗-th reaction, which can be expressed by an Arrhenius type expression
of the form:
𝑏
𝑘 𝑗 = 𝐴 𝑗 𝑇w 𝑗 exp(−𝐸 𝑗 /𝑅𝑇w ) (8)
where 𝑇w is the wall temperature (in Kelvin) and 𝐴 𝑗 , 𝑏 𝑗 , and 𝐸 𝑗 are the pre-exponential factor, the temperature
exponent, and the activation energy for the 𝑗-th reaction, respectively. The total erosion rate of carbon due to the surface
heterogeneous reactions is finally expressed as:

𝑚¤ w = 𝑚¤ H2 O + 𝑚¤ CO2 + 𝑚¤ OH + 𝑚¤ O + 𝑚¤ O2 (9)

In terms of energy absorbed (or released) by the surface reaction, it is worth noting that the reactions with H2 O,
CO2 , and OH are characterized by a positive heat of reaction, which means that energy is absorbed when carbon is
oxidized by one of these oxidizing species, whereas the reactions with O2 and O are characterized by a negative heat of
reaction, hence energy is released. Hence, it is evident how the oxidant species concentration can alter significantly the
nozzle throat ablation process.

III. Test case and numerical set-up

A. Test case
In this work we consider a series of 2 kN thrust-class CAMUI-type hybrid rocket motor static firing tests performed
at the University of Hokkaido, using liquid oxygen (LOX) as oxidizer and high-density polyethylene (HDPE) as fuel. The
experimental oxidizer-to-fuel mass ratio (O/F) and the throat recession data are obtained from indirect measurements,
using a Nozzle Throat Reconstruction Technique (NTRT) [23–25], as the in-depth temperature profiles, for which the

4
Throat-Temperature Reconstruction Technique (TTRT) proposed in [26] is used. These two methods were shown to
be a useful means to perform low-cost firing tests of HRMs, including throat erosion and heating effects without the
requirement of complex measurement devices.
The NTRT is built upon the governing equations and concepts of the reconstruction techniques introduced originally
in [23] to reconstruct the nozzle throat radius and oxidizer-to-fuel mass ratio histories through data reduction of
commonly measured experimental values: (1) oxidizer mass flow rate; (2) chamber pressure; (3) thrust; (4) overall
fuel mass consumed; and (5) final nozzle throat radius. The basic concept of NTRT is the estimation of the nozzle
erosion by solving the governing equations for compressible flow in converging-diverging ducts, in conjunction with the
governing equations for rocket performance. Solutions obtained in this way from NTRT are independent of the nozzle
material, and thus may be applied to any nozzle shape or material as long as the assumptions behind the governing
equations are satisfied.

TEST CAMPAIGN A:
TEST CAMPAIGN B:
solid lines = ’’nozzle A1’’ (Dth = 27 mm)
dashed lines = ’’nozzle A2’’ (Dth = 23 mm) solid lines = ’’nozzle B1’’ (Dth = 33.4 mm)
dashed lines = ’’nozzle B2’’ (Dth = 27.9 mm)
dash-dotted lines = ’’nozzle B3’’ (Dth = 23.6 mm)
dash-dot-dotted lines = ’’nozzle B4’’ (Dth = 19.8 mm)
10 mm
101 mm
221.3 mm
10 mm
4|5 mm
8.77|11.5 deg 9.9|11.8|13.95|16.7 mm
39.6|47.2|55.8|66.8 mm

66 mm Dth=23|27 mm
48|60 mm 130 mm Dth = 19.8|23.6|27.9|33.4 mm 44.3|52.8 mm

30 deg

58.2|58.9|59.8|60.8 mm

(a) Test campaign A. (b) Test campaign B.

Fig. 2 Schematics of test campaign A (oxidizer rich) and test campaign B (fuel rich) nozzles.

In particular, two different test campaigns carried out at the University of Hokkaido using the NTRT reconstruction
technique are analyzed [24]. The first one is an oxidizer rich test campaign (denoted as test campaign A), which adopts
the nozzle assembly shown in Fig.2(a): two different nozzles throat diameters are foreseen for this test campaign,
resulting in two different possible nozzles, which will be denoted as “nozzle A1” and “nozzle A2”, with respectively a
throat diameter of 27 and 23 mm. The second one is instead a fuel rich test campaign (denoted as test camapign B),
which adopts the nozzle assembly shown in Fig.2(b): in this case, four different nozzles throat diameters have been used,
resulting in four different nozzles, which will be denoted by now as “nozzle B1”, “nozzle B2”, “nozzle B3”, and “nozzle
B4”, with respectively a throat diameter of 33.4, 27.9, 23.6, and 19.8 mm. The material used for both test campaigns A
and B and for all the various nozzles is graphite, with a density of 1850 kg/m3 , for which thermodynamic properties as a
function of temperature (specific heat and enthalpy, hs ) are taken from the literature [20].
The most important measured and reconstructed experimental parameters obtained by the NTRT for both test
campaigns A and B are reported in Figs.3, 4, and 5. In particular, in Fig.3 chamber pressure traces obtained from direct
measurements can be observed. It is worth nothing how test campaign A motor firings are characterized by limited
time durations, approximately of the order of 5 seconds, with the only exception of test A-7 (see Fig.3(a)). Conversely,
test campaign B is characterized by diversified firings durations, which range among 5 and 25 seconds (see Fig.3(b)).
Moreover, test campaign B average chamber pressures show to be higher than test campaign A ones. Conversely to
chamber pressures, mixture ratios distributions over time for the two test campaigns analyzed, here reported in Fig.4,
have been obtained from indirect measurements, employing the NTRT. The reliability of this numerical technique can
be appreciated by looking to Fig.4(b), where the repeatability of the reconstructed O/F for subsequent firings of the same
motor is evident (tests B-1 to B-6). Those reconstructed mixture ratios are here used in order to estimate the average O/F
for each specific firing tests, which are needed in order to build up the test matrix for numerical simulations. Certainly,

5
Test campaign A (ox. rich) Test campaign B (fuel rich)
45 45
Test A-1 Test B-1
Test A-2 Test B-2
40 Test A-3
40
Test B-3
Test A-4 Test B-4
35 Test A-5 35 Test B-5
Test A-6 Test B-6
chamber pressure, bar

chamber pressure, bar


Test A-7 Test B-7
30 30 Test B-8
Test B-9

25 25

20 20

15 15

10 10

5 5

0 0
0 1 2 3 4 5 6 7 8 9 10 0 2.5 5 7.5 10 12.5 15 17.5 20 22.5 25
time, s time, s

(a) Test campaign A (oxidizer rich). (b) Test campaign B (fuel rich).

Fig. 3 Experimental chamber pressure traces from the oxidizer-rich (a) and fuel-rich (b) test campaigns
employing the NTRT reconstruction technique [24].

Test campaign A (ox. rich) Test campaign B (fuel rich)


12 12
Test A-1 Test B-1
11 Test A-2 11 Test B-2
Test A-3 Test B-3
10 Test A-4 10 Test B-4
Test A-5 Test B-5
9 Test A-6 9 Test B-6
Test A-7 Test B-7
mixture ratio (O/F)

mixture ratio (O/F)

8 8 Test B-8
Test B-9
7 7

6 6
ox rich
5 5 ox rich

4 4 O/F st = 3.42
3 O/F st = 3.42 3

2 2 fuel rich
1 fuel rich 1

0 0
0 1 2 3 4 5 6 7 8 9 10 0 2.5 5 7.5 10 12.5 15 17.5 20 22.5 25
time, s time, s

(a) Test campaign A (oxidizer rich). (b) Test campaign B (fuel rich).

Fig. 4 Experimental mixture ratio traces from the oxidizer-rich (a) and fuel-rich (b) test campaigns employing
the NTRT reconstruction technique [24].

the most important information coming out from the NTRT is the time evolution of the nozzle throat radius, here shown
in Fig.5 for both test campaigns A and B. With the exception of the transient period during start-up, the results of
NTRT for both the test campaigns A and B agree closely with the measured values for initial throat radius (instead, the
final throat radius is an input data in the rebuilding technique), and display physically consistent (i.e., continuous and
increasing) values for nozzle-throat erosion history. However, generally the numerically reconstructed initial throat
radius is not always exactly coincident with its measured value, hence in these cases the local average erosion rates,
extrapolated directly from the NTRT data, can lead to some errors with respect to the real nozzle behaviour. Therefore,
it must be identified a certain criterion in order to manage and use the NTRT data, especially in terms of determination
of the nozzle-throat average erosion rate.

6
Test B-1 Test campaign B (fuel rich)
Test B-2
Test campaign A (ox. rich)
16 Test A-1 22 Test B-3
Test A-2 Test B-4
Test A-3 21 Test B-5
Test A-4 Test B-6
Test A-5
20 Test B-7
15
Test A-6 19 Test B-8
Test A-7 Test B-9
18
throat radius, mm

throat radius, mm
14 Rth = 16.7 mm
17

Rth = 13.5 mm 16
13 15
Rth = 13.95 mm
14
13
12
Rth = 11.8 mm
12
Rth = 11.5 mm 11
11 Rth = 9.9 mm
10
9
10 8
0 1 2 3 4 5 6 7 8 9 10 0 2.5 5 7.5 10 12.5 15 17.5 20 22.5 25
time, s time, s

(a) Test campaign A (oxidizer rich). (b) Test campaign B (fuel rich).

Fig. 5 Experimental throat radius traces from the oxidizer-rich (a) and fuel-rich (b) test campaigns employing
the NTRT reconstruction technique [24].

Table 3 Averages of directly measured experimental data (full test duration).

Test tb , s tstart , s nozzle Dth0 , mm pc , bar O/F ¤r, mm/s 𝜂c∗


Test A-1 4.86 0.0 A1 27.0 20.17 (±1.4%) 1.75 (±6.5%) 0.062 (±16.7%) 1.038
Test A-2 4.78 0.0 A1 27.0 21.04 (±1.3%) 2.65 (±4.1%) 0.044 (±24.3%) 0.925
Test A-3 4.70 0.0 A1 27.0 19.52 (±1.4%) 3.38 (±4.1%) 0.136 (±5.6%) 0.947
Test A-4 4.91 0.0 A1 27.0 13.19 (±2.1%) 9.82 (±4.1%) 0.110 (±7.9%) 0.842
Test A-5 4.77 0.0 A1 27.0 16.09 (±1.8%) 5.42 (±3.8%) 0.078 (±36.7%) 0.903
Test A-6 4.82 0.0 A2 23.0 26.24 (±1.1%) 2.87 (±4.4%) 0.243 (±8.5%) 1.035
Test A-7 9.70 0.0 A2 23.0 13.05 (±2.2%) 6.32 (±10.5%) 0.053 (±3.4%) 1.040
Test B-1 4.07 0.0 B3 23.6 32.94 (±0.9%) 1.34 (±8.8%) 0±0.033 0.986
Test B-2 7.04 0.0 B3 23.6 31.46 (±0.9%) 1.53 (±8.5%) 0.055 (±34.7%) 1.030
Test B-3 13.34 0.0 B3 23.6 29.53 (±1.0%) 1.90 (±8.2%) 0.142 (±6.9%) 1.045
Test B-4 19.06 0.0 B3 23.6 27.94 (±1.0%) 2.15 (±8.3%) 0.192 (±4.2%) 1.061
Test B-5 12.44 0.0 B3 23.6 31.04 (±0.9%) 1.84 (±8.2%) 0.103 (±10.6%) 1.071
Test B-6 24.65 0.0 B3 23.6 23.01 (±1.2%) 2.99 (±9.8%) 0.312 (±2.8%) 1.091
Test B-7 10.73 0.0 B4 19.8 27.23 (±1.0%) 1.43 (±17.5%) 0±0.012 0.974
Test B-8 5.35 0.0 B2 27.9 28.53 (±1.0%) 1.72 (±5.4%) 0.094 (±20.9%) 1.067
Test B-9 7.35 0.0 B1 33.4 21.65 (±1.3%) 2.19 (±4.6%) 0.102 (±13.3%) 1.009

In order to carry out a test matrix to be numerically investigated, the NTRT experimental data just presented have
been managed according to two different criteria. The first criterion relies only on the directly measured experimental
data, used in order to get simply the full test duration averaged parameters, here shortly reported in Tab.3 for both test
campaign A and B. In particular, in this case average O/F values does not rely on the reconstructed mixture ratio time
histories, as they come directly from the knowledge of the overall oxidizer and fuel masses consumed. Reconstructed
parameters are not used also in order to evaluate the average nozzle-throat erosion rates, which are simply evaluated
from the knowledge of the initial and final throat radius and of the overall burning times. However, it is worth nothing

7
how in this way the solid transient heating is completely not accounted for, leading to possible underestimations of
the erosion rates, especially in case of short duration firing tests. Hence, a second criterion in order to manage the
experimental data has been defined, which relies on the identification of the instant when the nozzle erosion starts.
The identification of this time instant has been possible thanks to the NTRT reconstructed parameters, in particular
by looking to the throat radius time histories and to their first derivatives. However, it is worth nothing how, once the
nozzle-throat erosion starting time has been identified, the initial throat radius has been always assumed to be equal to
the measured initial radius, even if the NTRT provides a different value. Average oxidizer-to-fuel mass ratios have
been then determined as the ratio between the oxidizer and fuel mass flow ratio time integrals, evaluated by using the
reconstructed O/F time histories and the measured oxidizer flow rates. The final average parameters obtained using this
criterion (denoted as recession start criterion) for both test campaigns A and B are summarized in Tab.4.

Table 4 Averages of experimental data obtained using the NTRT (recession start).

Test tb , s tstart , s nozzle Dth0 , mm pc , bar O/F ¤r, mm/s 𝜂c∗


Test A-1 4.86 3.0 A1 27.0 20.48 (±1.4%) 1.89 (±6.0%) 0.161 (±16.8%) 1.038
Test A-2 4.78 3.8 A1 27.0 19.79 (±1.4%) 3.71 (±3.4%) 0.215 (±24.4%) 0.925
Test A-3 4.70 2.2 A1 27.0 18.94 (±1.5%) 4.16 (±3.9%) 0.256 (±5.6%) 0.947
Test A-4 4.91 2.25 A1 27.0 13.12 (±2.1%) 10.23 (±3.8%) 0.203 (±7.9%) 0.842
Test A-5 4.77 2.0 A1 27.0 16.07 (±1.8%) 5.86 (±3.9%) 0.134 (±36.8%) 0.903
Test A-6 4.82 1.0 A2 23.0 27.47 (±1.0%) 3.05 (±2.9%) 0.306 (±8.5%) 1.035
Test A-7 9.70 1.0 A2 23.0 13.25 (±2.1%) 6.24 (±10.5%) 0.059 (±3.4%) 1.040
Test B-1 4.07 1.0 B3 23.6 34.01 (±0.8%) 1.60 (±8.7%) 0±0.044 0.986
Test B-2 7.04 5.0 B3 23.6 32.89 (±0.9%) 2.12 (±8.5%) 0.188 (±34.7%) 1.030
Test B-3 13.34 6.0 B3 23.6 29.03 (±1.0%) 2.51 (±8.2%) 0.259 (±6.9%) 1.045
Test B-4 19.06 6.0 B3 23.6 26.44 (±1.1%) 2.74 (±8.1%) 0.280 (±4.2%) 1.061
Test B-5 12.44 7.0 B3 23.6 31.78 (±0.9%) 2.60 (±7.9%) 0.236 (±10.6%) 1.071
Test B-6 24.65 4.0 B3 23.6 21.91 (±1.3%) 3.59 (±10.1%) 0.373 (±2.9%) 1.091
Test B-7 10.73 4.0 B4 19.8 29.85 (±0.9%) 1.07 (±20.7%) 0±0.020 0.974
Test B-8 5.35 1.5 B2 27.9 30.38 (±0.9%) 1.44 (±5.2%) 0.130 (±20.9%) 1.067
Test B-9 7.35 2.2 B1 33.4 21.99 (±1.3%) 2.15 (±6.4%) 0.140 (±14.0%) 1.009

B. Numerical set-up and grid convergence


The basic graphite nozzle geometries used in the numerical simulations performed in this work are the ones shown
in Fig.6 and the test conditions of interest are those reported in Tabs.3 and 4. It is worth nothing how the original test
campaign A nozzle geometry (Fig.2(a)) has been slightly modified in the final numerical set-up adopted in this work
(see Fig.6(a)). In addition to a divergent section cut-off, which has been applied to the test campaign B nozzle as well
(Fig.6(b)), the convergent section has been modified, reproducing it by using two 45 deg circular arcs with the same
10 mm radius and with opposed curvature signs, so that it becomes parallel to the nozzle axis at the inlet section. A
constant section portion has been then added prior the convergent zone, in order to keep the overall wall entrance length
of the original nozzle. In fact, the length of the wall prior the throat region is an important parameter because it affects
the boundary-layer thickness, and hence the heat and mass transfer rates, which subsequently alter the ablation process.
By applying these geometrical modifications, the modified test campaign A nozzle shows to have the same throat to
curvature radius ratio and the same wall entrance length of the original one, hence no significant alterations on the throat
ablative response are expected from numerical simulations. With this modified convergent shape arrangement, a purely
axial inlet velocity profile can be assigned and conformal mapping (employed by the CFD solver) can be efficiently
applied to mesh the internal nozzle volume. The use of conformal mapping allows for a grid that is well adapted to the
geometry of the body, which can be of importance when wall parameters are sought for, such as in the present analysis.
Hence, the modified convergent shape nozzle profile as reported in Fig.6(a) has been used in the following numerical

8
TEST CAMPAIGN A: TEST CAMPAIGN B:
mesh resolution: 70x80
mesh resolution: 80x90

Ablative wall boundary

Ablative wall boundary

Subsonic
inlet Supersonic
outlet Subsonic
inlet
Supersonic
outlet

Symmetry Symmetry

(a) Test campaign A nozzle. (b) Test campaign B nozzle.

Fig. 6 Numerical set-up used for the numerical simulations.

analysis for test campaign A nozzles.


In the numerical set-up, an axi-symmetric portion of the nozzles has been considered (see Fig.6). Boundary
conditions other than an ablating wall are enforced as follows: static temperature, static pressure, Mach number, flow
direction and chemical composition at the inflow (supersonic inlet), no assigned conditions at the outflow (supersonic
outlet), and symmetry axis. Combustion chamber conditions, used in the subsonic inlet boundary, have been obtained by
a chemical equilibrium code [20] by varying chamber pressures and O/F over a certain range of interest. Eight species
(CO, CO2 , H2 , H2 O, OH, H, O, and O2 ) have been considered for each equilibrium simulation as they constitute more
than 99.9% of the total combustion gas mass for the LOX-HDPE propellant combination here analyzed.
As shown in Fig.6, a cutted grid has been presented in the numerical set-up for both test campaign A and B nozzles.
In fact, as the throat is the only location where experimental data are available, the numerical domain has been limited
in order to reduce the computational efforts required. Hence, cutted grids, as the ones in Fig.6, have been generated by
terminating both the full length grids immediately after of the nozzle throat. A comparison of results from the complete
grid (full length nozzle) and the cut-off grid is shown in Fig.7 for both test campaign A and B nozzles. In particular,
Fig.7(a) shows the Mach contours for the two grids while Fig.7(c) shows the ablation mass flux distributions for the test
campaign A nozzle. As it can be clearly seen, the results in the throat region are identical for both grids. Similar results
have been obtained for the test campaign B nozzle, as reported in Figs.7(b) and 7(d). Hence, from now on, considering
that the final goal of this work is to accurately simulate only the nozzle throat erosion, the cut-off grids shown in Fig.6
will be used for both test campaign A and B numerical analysis.
Finally, the numerical solutions have been verified by a grid convergence analysis on three grid levels (see Fig.8).
Concerning the test campaign A nozzle, the finest mesh has 140x160 grid points in the axial and radial directions,
respectively. Each mesh coarsening has been then obtained by removing one node out of two in each coordinate
direction. Figure 8(a) shows a comparison between the ablation mass flux distributions along the nozzle wall evaluated
with the three grid levels for the test campaign A nozzle. The quantitative analysis of solutions obtained on three grid
levels confirms that the spatial order of accuracy is, in most regions of the nozzle, close to the formal second-order
value of the scheme. This confirms the asymptotic behavior of the numerical error, and thus gives a good confidence
on the error estimate. The discrepancy between the ablation mass flux evaluated with the medium and fine grids is
roughly 2% at the throat location; therefore, the medium grid (70x80) has been considered sufficiently refined for the
present analysis concerning test campaign A nozzles. A similar result has been obtained for the test campaign B nozzles
(see Fig.8(b)). In particular, in this case, the finest mesh has 120x135 grid points in the axial and radial directions,
respectively. However, the mesh coarsening has not been obtained by halving each time the number of nodes in both
directions, as in this case some problems have been faced to get a solution with the coarser mesh level. This is due to the
high contraction ratio of the nozzle convergent, which leads to low values of the Mach number (even lower than 0.03) in

9
7 12

6 10 Mach: 0.1 0.3 0.5 0.7 0.9 1.1 1.3 1.5 1.7 1.9 2.1 2.3 2.5 2.7

Mach: 0.1 0.3 0.5 0.7 0.9 1.1 1.3 1.5 1.7 1.9 2.1 2.3 2.5 2.7
5 8
radial distance, cm

radial distance, cm
4 6

3 4

2 2

1 0

-2
0
-4
-1
-6
-1 0 1 2 3 4 5 6 7 8 9 0 2 4 6 8 10 12 14 16 18 20 22
axial distance, cm axial distance, cm

(a) Test campaign A mach contours. (b) Test campaign B mach contours.

1 20 0.2
throat throat
9
nozzle 0.9 18 nozzle
full divergent 180x80 full divergent 105x90
8
cut-off divergent 70x80 0.8 cut-off divergent 80x90
ablation mass flux, kg/(s m )

ablation mass flux, kg/(s m )


16
2

2
0.15
7
0.7
radial distance, cm

radial distance, cm

14
6
0.6 12
5
0.5 10 0.1

4 0.4 8

3 0.3 6
0.05
2 0.2 4

1 0.1 2

0 0 0 0
-1 0 1 2 3 4 5 6 7 8 9 0 2 4 6 8 10 12 14 16 18 20 22
axial distance, cm axial distance, cm

(c) Test campaign A nozzle ablation mass flux distributions. (d) Test campaign B nozzle ablation mass flux distributions.

Fig. 7 CFD solution for the oxidizer and fuel rich test campaigns (A and B) nozzle with full length and cut-off
grid at nozzle divergent.

the inlet region, which are not easy to be managed with a reduced number of nodes by using a compressible CFD solver
as the one adopted in this work. Nevertheless, by properly setting a minimum number of grid points for the coarse mesh
level, the numerical solutions from the three different grid levels have been obtained (see Fig.8(b)). The discrepancy
between the ablation mass flux evaluated with the medium and fine grids shows to be roughly 1% at the throat location;
therefore, the medium grid (80x90) has been considered sufficiently refined for the present analysis concerning the test
campaign B nozzles and is used in the following numerical simulations.

IV. Parametric numerical analysis


In this section, the CFD model described in the preceding sections is used to simulate the nozzle erosion in
HRM environments employing the LOX-HDPE propellant combination. The final purpose is to perform an extensive
parametric numerical analysis, including the effect of different operating conditions (i.e., chamber pressure and O/F

10
throat 1 0.2
16
3 nozzle throat
coarse (35x40) 0.9
14 nozzle
medium (70x80)
coarse (56x63)
fine (140x160) 0.8

ablation mass flux, kg/(s m )

ablation mass flux, kg/(s m )


2.5

2
medium (80x90)
fine (120x135) 0.15
12
0.7
radial distance, cm

radial distance, cm
2
0.6 10

0.5 8 0.1
1.5
0.4
6
1 0.3
4 0.05
0.2
0.5
2
0.1

0 0 0
-1 -0.5 0 0.5 1 1.5 2 0 2 4 6 8 10 12 14 16 18
axial distance, cm axial distance, cm

(a) Test campaign A nozzles. (b) Test campaign B nozzles.

Fig. 8 Nozzle ablation mass flux distributions evaluated using three grid levels for the grid convergence analysis
(both A and B test campaign nozzles).

ratio). Nozzle geometries from both test campaign A and B are investigated (see Fig.6), with the purpose to underline
possible differences in the ablative response due to geometrical effects. Reference conditions in terms of chamber
pressure have been chosen to be 20 bar and 30 bar for test campaign A and test campaign B nozzles, respectively, as
these values represent a good average of the experimental pressure traces reported in Fig.3.
First of all, some preliminary numerical analyses have been performed to have a first important insight into the
problem. The first one investigates the effects of the different nozzle throat radius foreseen for both test campaign A
and B nozzles on the throat ablative response (see Fig.9). Concerning test campaign A, “nozzle A1” (Dth = 27 mm)
and “nozzle A2” (Dth = 23 mm) have been numerically investigated using the reference pressure and the stoichiometric
oxidizer-to fuel mass ratio (see Fig.9(a)). As evident, variations of the nozzle geometry have a negligible effect on the
throat ablation mass flux. This is probably due to the limited variation of the nozzle geometrical characteristics, such as

chamber pressure = 20 bar


mixture ratio (O/F) = 3.42 (stoich) mixture ratio (O/F) = 3.42 (stoich)
chamber pressure = 30 bar
Dth = 27 mm 1.2 0.8
11
Dth = 23 mm throat solid lines = ablation rate throat
2 dashed lines = nozzle profile
10 0.7
1
9 Dth 20 mm
ablation mass flux, kg/(s m )

ablation mass flux, kg/(s m )


2

2
Dth 24 mm 0.6
8 Dth 28 mm
radial distance, cm

radial distance, cm

1.5 0.8 Dth 33 mm


7 0.5

6
0.6 0.4
1 5

4 0.3
0.4
3 0.2
0.5
0.2 2
0.1
solid lines = ablation mass flux 1
dashed lines = nozzle profiles
0 0 0 0
0 0.5 1 1.5 2 6 7 8 9 10 11 12 13 14 15 16 17 18
axial distance, cm axial distance, cm

(a) Test campaign A nozzles (b) Test campaign B nozzles

Fig. 9 Throat diameter effects on test campaign A and B nozzles ablation mass flux.

11
contraction ratio, throat to curvature radius ratio, and wall entrance length. Differences can be only reported in the
divergent section, where the divergence angle is changed between the two nozzles. Hence, for test campaign A, no
quantitative differences are expected in terms of throat ablative response by changing the nozzle throat radius. A similar
analysis has been carried out for test campaign B nozzles: “nozzle B1” (Dth = 33.4 mm), “nozzle B2” (Dth = 27.9 mm),
“nozzle B3” (Dth = 23.6 mm), and “nozzle B4” (Dth = 19.8 mm). The reference pressure for test campaign B and the
stoichiometric O/F have been used in these numerical simulations (see Fig.9(b)). Here, an unexpected result has been
obtained. In fact, it is common belief that the convective heat flux (and consequently the ablative response) at nozzle
throat location is increased as the throat radius is reduced, as typically predicted by using semi-empirical laws, such as
the Bartz equation. However, for the test campaign B nozzles the ablation mass flux at throat shows to increase as the
throat radius increases. This is probably due to the gradually reduced contraction ratio when throat radius grows (see
Fig.9(b)), which affects the flowfield in the convergent region, resulting in a different development of the boundary layer
along the nozzle entrance prior the throat section. Therefore, for test campaign B, some differences are expected in
terms of throat ablative response by changing the nozzle throat radius.
From now on, a single nozzle will be analyzed for each experimental test campaign, as nozzle throat radius effects
have been qualitatively understood. In particular, the most common nozzle in each test campaign has been selected as
the reference one (see Tab.3): for test campaign A, “nozzle A1” (27 mm throat radius) has been chosen, while for test
campaign B, “nozzle B3” (23.6 mm throat radius) has been selected. Hence, from now on, test campaign A and B
nozzle will be assumed to be always the reference ones.
A second preliminary numerical analysis has been carried out in order to understand how the different oxidizing
species contribute to the overall ablation mass flux by varying the oxidizer-to-fuel mass ratio. Both test campaign A and
B nozzles have been investigated (see Fig.10). At fuel rich conditions (i.e., O/F=2), for both test campaign A and B
nozzles, water vapor is the dominant oxidizing species, contributing to almost 85% of the total throat ablation mass flux
(see Figs.10(a) and 10(b)). This is due to the high reaction rate of water vapour at high temperatures and its abundance
in the exhaust gas composition at fuel rich conditions. The second most important oxidizing species for both nozzles
appears to be OH, which, although present only in small concentrations in the exhaust gases at fuel rich conditions, has
a very high reaction rate, even at low temperatures. Looking to stoichiometric conditions (i.e., O/F=3.42), water vapour
contribution is reduced from 85% to approximately 60% of the total throat ablation mass flux (see Figs.10(c) and 10(d)).
This is due to the slight increase of mass fractions of the other oxidizing species in the exhaust gases. In particular, due
to the high flame temperature present at stoichiometric conditions, more dissociated species (e.g., OH, O) are produced.
In fact, OH is again the second most important oxidizing species, with an increased contributions to the overall ablation
with respect to fuel rich conditions (30% at O/F=3.42 vs. 15% at O/F=2). Non-negligible contributions at stoichiometric
conditions come from molecular and atomic oxygen as well, for both test campaign A and B nozzles. Lastly, oxidizer-rich
conditions (i.e., O/F=10) have been investigated (see Figs.10(e) and 10(f)). In this case, water vapour contribution to the
overall ablation mass flux is further reduced to less than 50%. Furthermore, an important difference with respect to fuel
rich and stoichiometric conditions is that the second most important oxidizing species is molecular oxygen O2 , with
a contribution to the overall ablation mass flux similar to water vapour. It is worth nothing that the surface reaction
of graphite with O2 is characterized by a negative heat of reaction, hence leading to an exothermic surface oxidation
process, which affects the surface energy balance resulting in increased wall temperatures at oxidizer-rich conditions
even in case of low flame temperatures. Furthermore, at oxidizer-rich conditions, a difference between test campaign
A and test campaign B nozzles can be underlined. In particular, the molecular oxygen contributions to the overall
ablation mass flux in the early convergent zones are quite different between the two nozzles (see Figs.10(e) and 10(f)).
For test campaign B nozzle (Fig.10(f)) molecular oxygen is largely the dominant oxidizing species prior the throat
location, while this is not the case of test campaign A nozzle (Fig.10(e)). This can be justified by the different convergent
geometry, which for test campaign B is characterized by a greater contraction ratio. Instead, near the throat location and
in the divergent region, results from the two nozzles are almost coincident, as there the wall geometries are quite similar.
In the following, parametric numerical analyses, carried out specifically in order to more deeply understand the
combined effects of chamber pressure and oxidizer-to-fuel mass ratio on the nozzle throat ablation in HRMs, are
presented. The reference nozzles of the two experimental test campaigns are analyzed, and the obtained results will
be used in order to derive a regression law for nozzle throat ablation based on numerical data, which will be finally
employed for the numerical rebuilding of the experimental tests (Tabs.3 and 4).

12
chamber pressure = 20 bar
mixture ratio (O/F) = 2 nozzle chamber pressure = 30 bar
overall ablation mixture ratio (O/F) = 2
CO2 0.8 11 0.8
throat
H 2O
nozzle
2 OH 10 overall ablation
O 0.7 0.7
CO2
O2 9 H 2O

ablation mass flux, kg/(s m )

ablation mass flux, kg/(s m )


2

2
OH
0.6 O 0.6
8
radial distance, cm

radial distance, cm
1.5 O2

0.5 7 0.5

6
throat 0.4 0.4
1 5
0.3 4 0.3

3
0.2 0.2
0.5
2
0.1 0.1
1

0 0 0 0
0 0.5 1 1.5 2 6 7 8 9 10 11 12 13 14 15 16 17 18
axial distance, cm axial distance, cm

(a) Fuel rich conditions - Test campaign A. (b) Fuel rich conditions - Test campaign B.

chamber pressure = 20 bar


mixture ratio (O/F) = 3.42 (stoich) nozzle chamber pressure = 30 bar
overall ablation mixture ratio (O/F) = 3.42 (stoich)
CO2 0.8 11 0.8
throat
H 2O
nozzle
2 OH 10 overall ablation
O 0.7 0.7
CO2
O2 9 H 2O
ablation mass flux, kg/(s m )

ablation mass flux, kg/(s m )


2

2
OH
0.6 O 0.6
throat 8
radial distance, cm

radial distance, cm
1.5 O2

0.5 7 0.5

6
0.4 0.4
1 5
0.3 4 0.3

3
0.2 0.2
0.5
2
0.1 0.1
1

0 0 0 0
0 0.5 1 1.5 2 6 7 8 9 10 11 12 13 14 15 16 17 18
axial distance, cm axial distance, cm

(c) Stoichiometric conditions - Test campaign A. (d) Stoichiometric conditions - Test campaign B.

chamber pressure = 20 bar


mixture ratio (O/F) = 10 nozzle chamber pressure = 30 bar
overall ablation mixture ratio (O/F) = 10
CO2 0.8 11 0.8
throat
H 2O
nozzle
2 OH 10 overall ablation
O 0.7 0.7
CO2
O2 9 H 2O
ablation mass flux, kg/(s m )

ablation mass flux, kg/(s m )


2

2
OH
0.6 O 0.6
throat 8
radial distance, cm

radial distance, cm

1.5 O2

0.5 7 0.5

6
0.4 0.4
1 5
0.3 4 0.3

3
0.2 0.2
0.5
2
0.1 0.1
1

0 0 0 0
0 0.5 1 1.5 2 6 7 8 9 10 11 12 13 14 15 16 17 18
axial distance, cm axial distance, cm

(e) Oxidizer rich conditions - Test campaign A. (f) Oxidizer rich conditions - Test campaign B.

Fig. 10 Contributions to the overall ablation mass flux from the different oxidizing species under different
mixture ratios for test campaign A and B reference nozzles.

13
0.8 0.8
0.7 0.7
O/F = 3.42 (stoich) 0.6 O/F = 3.42 (stoich)
0.6 O/F = 10 O/F = 10
throat ablation mass flux, kg/(s m2)

throat ablation mass flux, kg/(s m2)


0.5
0.5 O/F = 2 O/F = 2
0.4
0.4 0.3

0.3 pc0.844 pc1.105


0.2

0.614
0.2 pc
0.869
0.1 pc

pc0.889
pc1.177
0.1

5 10 15 20 25 30 35 40 5 10 15 20 25 30 35 40
chamber pressure, bar chamber pressure, bar

(a) Throat ablation mass flux - Test campaign A. (b) Throat ablation mass flux - Test campaign B.

circles = throat wall temperauture circles = throat wall temperature


O/F = 3.42 (stoich) squares = flame temperature O/F = 3.42 (stoich) squares = flame temperature
O/F = 10 O/F = 10
O/F = 2 O/F = 2
3600 3600

3400 3400

3200 3200
temperature, K

temperature, K

3000 3000

2800 2800

2600 2600

2400 2400

2200 2200

5 10 15 20 25 30 35 40 5 10 15 20 25 30 35 40
chamber pressure, bar chamber pressure, bar

(c) Flame and throat wall temperature - Test campaign A. (d) Flame and throat wall temperature - Test campaign B.

Fig. 11 Nozzle ablation mass flux and temperature distributions at varying chamber pressures for different
mixture ratios for both test campaigns A and B reference nozzles.

A. Effect of chamber pressure


An important parameter that can affect the throat ablation mass flux is the chamber pressure. However, concerning
HRMs, a detailed analysis of its effect on throat erosion, by considering different O/F conditions, can not be retrieved
in literature. Here, three mixture ratio conditions are considered: fuel rich (O/F=2), stoichiometric (O/F=3.42), and
oxidizer rich (O/F=10). The range of chamber pressures analyzed (5 to 40 bar) covers all the pressures obtained during
the two experimental campaigns (see Fig.3).
Figure 11 summarizes the results obtained in terms of throat ablation mass flux and throat wall temperature for
both test campaign A and B nozzles at varying chamber pressures. Fig.11(a) shows how, for test campaign A reference
nozzle, the throat ablation mass flux grows almost linearly with chamber pressure. However, the pressure exponent
decreases monotonically moving towards oxidizer-rich conditions, where it is reduced of approximately 31% with
respect to the fuel-rich one. Hence, chamber pressure effects on throat ablation are gradually reduced shifting towards
oxidizer-rich conditions. As the O/F shift is an intrinsic characteristic of HRM, this effect must be carefully taken into

14
account. Similar results have been obtained for test campaign B reference nozzle (see Fig.11(b)). The main difference is
the absolute value of the pressure exponents, which for test campaign B reference nozzle are higher probably due to the
different nozzle geometry. However, the relative exponent reduction due to the mixture ratio shift is similar to the one
previously reported for the test campaign A nozzle.
Regarding throat wall temperature, (Figs. 11(c) and 11(d)) pressure effects are not as evident. The main reason is
that chamber pressure only slightly affects flame temperatures, while it leaves practically unchanged the exhaust gas
chemical composition. However, it is worth noting that, depending on mixture ratio conditions, throat wall temperature
values change with respect to flame temperature ones. In particular, for both test campaign A and B nozzles, at
oxidizer-rich conditions, throat wall temperature is still as high as at fuel-rich and stoichiometric conditions, even if the
flame temperature is considerably reduced. This is due to the high concentration of molecular oxygen, which leads to
the exothermic oxidation of the nozzle wall, affecting the surface energy balance by increasing the wall temperature.

0.8 0.8
O/Fst pc 5 bar
O/Fst
pc 5 bar
pc 20 bar pc 30 bar
0.7 0.7
throat ablation mass flux, kg/(s m2)

throat ablation mass flux, kg/(s m2)


pc 40 bar pc 40 bar

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10
mixture ratio (O/F) mixture ratio (O/F)

(a) Throat ablation mass flux - Test campaign A. (b) Throat ablation mass flux - Test campaign B.

circles = throat wall temperature circles = throat wall temperature


squares = flame temperature squares = flame temperature
3600 3600

3400 3400

3200 3200

3000 3000
temperature, K

temperature, K

2800 2800

2600 2600

2400 2400

2200 2200

2000 2000
O/Fst pc 5 bar
pc 5 bar O/Fst
pc 30 bar
1800 pc 20 bar 1800
pc 40 bar
pc 40 bar
1600 1600
1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10
mixture ratio (O/F) mixture ratio (O/F)

(c) Flame and throat wall temperature - Test campaign A. (d) Flame and throat wall temperature - Test campaign B.

Fig. 12 Nozzle throat ablation mass flux and temperature distributions at varying mixture ratios for different
chamber pressures for both test campaign A and B reference nozzles.

15
B. Effect of O/F ratio
A peculiar characteristic of HRMs is the shift of the O/F ratio, during both steady-state operation and throttling,
which can yield both fuel-rich and oxidizer-rich conditions. Different works in the literature analyzed HRM nozzles
ablative response at varying mixture ratio conditions [11, 12], however, as already observed in the previous section,
HRM nozzle throat ablation seems to be influenced in a coupled fashion by both O/F and chamber pressure. Hence, a
second ad-hoc parametric numerical analysis is here performed in order to investigate O/F effects on nozzle erosion
under different chamber pressure conditions. The range of mixture ratio analyzed (O/F=1 to 10) has been selected in
order to include all the test conditions from the two experimental campaigns under investigation (see Fig.4). Three
different chamber pressures have been considered, including the reference ones (i.e., 20 bar for test campaign A reference
nozzle, 30 bar for test campaign B reference nozzle), a low chamber pressure value (5 bar), and a high chamber pressure
value (40 bar).
As shown in Fig.12(a), for test campaign A the throat ablation mass flux is peaking at an oxidizer-rich condition for
all the three operating chamber pressure considered, differently from the corresponding peaks of flame temperatures
at slightly fuel rich conditions (see Fig.12(c)). This confirms that the nozzle ablation is primarily influenced by the
exhaust gas composition, and only secondarily by the propellant flame temperature. This, again, is confirmed by the
nozzle throat wall temperature distributions reported in Fig.12(c) for test campaign A reference nozzle, showing, as
well as the ablation mass flux, peaks at oxidizer rich conditions for all the three chamber pressure analyzed. It is also
interesting to note that the throat erosion rate is only slightly reduced moving from stoichiometric towards oxidizer
rich conditions, while it is strongly reduced moving in the fuel rich region. Note that this is a result of a strongly
kinetic-limited oxidation in the fuel-rich region, and not of an absence of oxidizing species. In fact, the nozzle throat
temperature at fuel rich conditions drops drastically to values lower than 2000 K, while at oxidizer-rich condition throat
wall temperature remains approximately as high as at stoichiometric conditions in a wide range of O/F ratios (see
Fig.12(c)). Similar results can be observed for test campaign B nozzle (see. Figs.12(b) and 12(d)). However, at the
same operating conditions (i.e., mixture ratio and chamber pressure) the test campaign B reference nozzle shows lower

1.3 1.3
O/Fst O/Fst
dimensionless throat ablation mass flux

dimensionless throat ablation mass flux

1.2 1.2
1.1 1.1
1 1
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
pc 5 bar pc 5 bar
0.2 pc 20 bar 0.2 pc 30 bar
pc 40 bar pc 40 bar
0.1 0.1
0 0
1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10
mixture ratio (O/F) mixture ratio (O/F)

(a) Test campaign A. (b) Test campaign B.

Fig. 13 Dimensionless nozzle throat ablation mass flux distributions at varying mixture ratios for different
chamber pressures for both test campaign A and B reference nozzles.

ablation mass fluxes (of approximately 30%) than the test campaign A reference nozzle ones. This is a nozzle geometry
effect, due to the greater wall entrance length of nozzle B, which allows the boundary layer to be better developed in
the convergent region, reducing the incoming heat fluxes in the throat section. Instead, by looking to the throat wall
temperature distributions for test campaign B reference nozzle (see Fig.12(d)), maximum wall temperature conditions
are peaking in the oxidizer rich zone for all the chamber pressure considered , as previously observed for nozzle A.
However, in this case the maximum wall temperature conditions are moved further in the oxidizer rich zone, so that for
the lower chamber pressures value considered, the throat wall temperature peak is even beyond the maximum mixture

16
ratio here analyzed.
Dimensionless ablation mass flux distributions with respect to stoichiometric value at varying mixture ratios for
three different chamber pressures are reported in Fig.13, considering both test campaign A and B reference nozzles.
Here, it is possible to immediately appreciate the complex coupled effects of O/F ratio and chamber pressure on HRM
nozzle throat ablative response. As previously reported, ablation mass flux is peaking at oxidizer rich conditions for
all the chamber pressure considered. However, it is worth nothing how the peaking mixture ratio location is modified
by the chamber pressure. In particular, as evident from Fig.13(a), the ablation mass flux peaking point moves further
towards the oxidizer rich region as chamber pressure decreases. Moreover, ablation mass flux reduction with respect to
stoichiometric conditions moving in the oxidizer rich zone shows to be strongly dependent on the operating chamber
pressure (see Fig.13(a)). In particular, for a chamber pressure of 40 bar, the throat ablation mass flux of the test campaign
A nozzle is only reduced by approximately 15% with respect to its value at the stoichiometric condition when the amount
of oxidizer is doubled (O/F=6.84). Such a reduction becomes pratically 0% when the chamber pressure is reduced
from 40 to 5 bar. Hence, by reducing the operating chamber pressure, the nozzle throat ablative response at oxidizer
rich conditions is less and less influenced by mixture ratio variations. Instead, by looking to the fuel rich branch of the
dimensionless ablation mass flux curves shown Fig.13(a), not appreciable effects due to chamber pressure variations can
be reported. Similar results have been obtained for test campaign B nozzle (see Figs.13(b)). However, in this case it is
evident how between the 30 and 40 bar dimensionless ablation mass flux curves there are some overlapping, probably
due to numerical issues linked to the low mach number conditions (less than M=0.03) in the early convergent region
of the test campaign B reference nozzle, which are hard to be captured by a compressible flow solver such as the one
adopted in this work, especially in case of high pressures.
In the following section, numerical results carried out from the parametric numerical analysis just analyzed will
be used to derive a regression law for nozzles throat ablation mass flux of HRMs empoying LOX-HDPE propellant
combination. A first validation of this regression law will be then carried out by comparison with the experimental data
reported in Tabs.3 and 4.

V. Regression laws and comparison with experiments


Regression laws and/or semi-empirical correlations for nozzle throat ablation are useful to rapidly evaluate nozzle
erosion and the corresponding performance losses during engine design. However, such regression laws must be able to
capture all the main features and dependencies of the ablative phenomenon, which can be quite complex especially
concerning HRMs, that characterized by mixture ratio shifts and possible chamber pressure variations. In this context,
this section aims to provide an efficient and versatile regression law for the rapid estimation of nozzle throat ablation for
HRMs employing LOX-HDPE propellants. The regression law here presented has been obtained by fitting the data
computed through the parametric numerical CFD simulations shown in the previous sections. A first validation of the
proposed regression law against experimental data has been performed as well.
The parametric numerical analysis shown in the previous section has brought to the following main conclusions
about HRMs nozzle throat ablation behaviour under different motor operating conditions:
• Throat ablation mass flux increases as chamber pressure grows for a fixed oxidizer-to-fuel mass ratio. However,
such an ablation mass flux growth rate with pressure is reduced when O/F moves towards oxidizer-rich conditions
• For a fixed chamber pressure, the nozzle throat ablation mass flux peak is at oxidizer rich conditions. Furthermore,
drastic ablation mass flux reduction are observed moving from stoichiometric to fuel rich conditions, while moving
towards oxidizer-rich conditions, ablation mass flux is more slowly decreased.
• Nozzle throat ablation mass flux maxima are moved further towards the oxidizer-rich region as chamber pressure
decreases. Moreover, by reducing the chamber pressure, the nozzle throat ablative response at oxidizer-rich
conditions is less and less influenced by mixture ratio variations.
Therefore, an efficient regression law for nozzle throat ablation must be able to capture all these combined effects. A
possible law can be written as:
m¤ w,th = a(O/F) pn(O/F)
c (10)
where n(O/F) is a pressure exponent, here modelled as a mixture ratio function, while a(O/F) represent a pre-exponential
factor, modelled as a mixture ratio function as well. In particular, n(O/F) must be able to take into account all the
coupled effects of chamber pressure and mixture ratio on the nozzle throat ablation, while a(O/F) is used to capture the
asymmetric bell-like shape of the nozzle throat ablation mass flux at varying O/F.

17
In this work, the general regression law just reported in Eq.(10) has been explicitly written in the following form:
c2 2
¤ w,th = O/F c1 10 ( O/F
m + c3 )
pc(c4 O/F + c5 O/F +c6 )
(11)

where c𝑖 (𝑖=1,...,6) are the fitting parameters, which need to be determined. These parameters have been calibrated using
the numerical results from the parametric numerical analysis previously performed in Sec.IV on both test campaign A
and B reference nozzles. In Eq.11 chamber pressure is measured in bar, while the throat ablation mass flux in kg/(s m2 ).
Firstly, two different set of fitting parameters have been extrapolated by using the numerical results obtained for the
two different nozzles analyzed separately (test campaign A and B). Hence, two different regression laws, specific for the
particular nozzle geometry, have been obtained. Results for the two nozzles analyzed are summarized in Tab.5, jointly
with the respective goodness of fit. Extremely high goodness of fit are reported for the two nozzles, with R2 of the order

Table 5 Fitting coefficients and goodness of fit of the throat ablative mass flux regression law for the test
campaigns A and B reference nozzles.

Test campaign A nozzle Test campaign B nozzle


c1 -1.134 -1.119
c2 -2.959 -3.382
c3 -0.04793 -0.4079
c4 0.0006131 0.003269
c5 -0.03541 -0.07628
c6 0.945 1.30
R2 0.989 0.997
R2 −adj 0.988 0.996

Fig. 14 Throat ablation mass flux fitting surface from a data regression law obtained using results from the
numerical parametric analysis on test campaign A and B reference nozzle.

of 0.99 for both nozzles. The high goodness of fit can be also appreciated by looking at Fig.14, where data from Eq.11
for the two nozzles have been plotted separately against the numerical results used for fitting constants calibration.
Clearly, a regression law specifically calibrated on a certain nozzle geometry has a limited field of application. In
fact, more generally, nozzle throat ablation can be heavily affected by geometrical parameters, such as nozzle throat
radius, wall entrance length, throat to curvature radius ratio, contraction ratio etc. This aspect has been clearly underlined
by analyzing the numerical results obtained in Sec.IV for the two extremely different nozzles used in test campaign A
and B. For this reason, a more general regression law is derived by using jointly the numerical results from both the

18
nozzles analyzed for fitting constants calibration. The final correlation is:
2
 −3.157

m¤ w,th = K O/F−1.125 10 ( O/F − 0.1557) pc(0.00174 O/F − 0.05301 O/F + 1.119) (12)

Equation 12 allows to take into account possible geometrical effects, including throat diameter, throat to curvature radius
ratio, convergent contraction ratio, and wall entrance length, by just correctly calibrating the parameter K for the specific
nozzle of interest. In Tab.6, the calibration parameters K for the two nozzle geometries under investigation and the

Table 6 Calibration constant and goodness of fit of the throat ablative mass flux regression law for the two
nozzles analysed.

Test campaign A nozzle Test campaign B nozzle


K 1.005 0.6877
R2 0.969 0.974
2
R −adj 0.967 0.972

corresponding goodness of fit are reported. It is worth nothing how the R2 values are slightly reduced with respect the
specific correlation ones (0.97 vs 0.99), however they remain extremely high, ensuring a high level of reliability for
Eq.12.
The general regression law reported in Eq.12, calibrated by using only numerical results obtained in this work,
has been finally compared to the experimental data in Tabs.3 and 4 for validation. Results are summarized in Fig.15.
Concerning full test duration experimental data (Tab.3), a quantitative good agreement of the regression law results with

Full test duration (measured parameters) Recession start (reconstructed parameters)


OFst OFst
0.45 0.45
Experimental data Experimental data
CFD correlation CFD correlation
0.4 0.4

0.35 0.35
throat erosion rate, mm/s

throat erosion rate, mm/s

0.3 0.3

0.25 0.25

0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05

0 0

-0.05 -0.05
1 2 3 4 5 6 7 8 9 10 11 1 2 3 4 5 6 7 8 9 10 11
mixture ratio (O/F) mixture ratio (O/F)

(a) Full test duration. (b) Recession start.

Fig. 15 Experimental throat erosion rates for test campaign A (ox. rich) and test campaign B (fuel rich) vs
CFD regression law for measured full test duration data and reconstructed data using NTRT info.

the experimental data can be reported (see Fig.15(a)). In particular, throat ablation mass flux confirms to be rapidly
reduced in the fuel rich region. However, for some tests an overestimation of the ablative response can be observed,
mainly due to solid heating transient effects which may be relevant for short time firings, which is the case in the majority
of the oxidizer rich tests (test campaign A) (see Tab.3). By looking to recession start experimental data (Tab.4), again
the regression law results show to perform well in comparison with the experiments (see Fig.15(b)). In particular, the
nozzle throat ablation mass flux behaviour at varying oxidizer-to-fuel mass ratio shows to be correctly captured by the
regression law. However, in this case a general tendency to slightly underestimate the ablative response can be observed.

19
For some tests, this can be referred to wall roughness effects, which may strongly enhance the erosion process [27]. In
fact, experimental evidences of wall roughness have been reported for some tests in the two experimental campaigns
analyzed (see Fig.16). Hence, a future development of this work is an investigation of possible wall roughness effects on
HRM nozzle throat ablative response.

Fig. 16 Example of wall roughness establishment during motor operation: test A-6 pictures after motor firings.

Looking again at Fig.15(b), it is worth nothing the following result for test A-4: representing the most oxidizer-rich
test case (O/F=10.23), test A-4 nozzle throat ablation has been largely underestimated by the regression law proposed in
this work. A possible explanation of this underestimation can be gas-phase reactions effects, which are not accounted
for in the numerical simulations, in which a frozen chemical composition is assumed. The assumption of frozen flow is
particularly good when combustion products show low concentrations of O2 , which is usually the case for composite
solid propellants. In fact, in this case the oxidation reaction of H2 and CO into H2 O and CO2 , respectively, are prevented.
In HRMs, however, it has been shown that the concentration of molecular oxygen can be significant (see Figs.10(e) and
10(f)), especially when oxidizer rich conditions are reached. Hence, at high O/F, gas-phase reactions effects might
have a key role to correctly predict nozzle-throat ablation. Results from past works show how throat ablation mass
flux can be slightly increased by including gas-phase reactions already at stoichiometric conditions [28]. Hence, at
oxidizer-rich conditions, much more pronounced effects are expected, as it seems to be confirmed by experimental
evidences of test A-4. For these reasons, a future development of this work will be certainly a thorough investigation of
gas-phase reactions effects on HRMs nozzle throat ablation, especially concerning oxidizer-rich conditions.
Typically, HRMs can be characterized by lower combustion efficiencies (𝜂c∗ ) than solid rockets due to incomplete
oxidizer–fuel mixing and combustion. However, all the numerical results obtained in the previous sections, as well
as the final throat ablation mass flux regression law just presented, employ a unitary combustion efficiency. Hence, a
preliminary numerical analysis in order to investigate possible combustion efficiency effects on HRMs ablative response
at varying chamber pressure and oxidizer-to-fuel mass ratio has been conducted in the following. Here, combustion
efficiency has been modelled by introducing an adiabaticity defect when combustion chamber conditions are calculated
in the employed chemical equilibrium code [20]. In this way, combustion efficiency will be accounted as function
of both flame temperature and average molecular weight of gaseous mixture (i.e., of combustion products chemical
composition). For combustion efficiency analysis, only test campaign A reference nozzle has been considered. Figure
17(a) shows nozzle throat ablation mass flux for three different combustion efficiencies at varying chamber pressure
under stoichiometric conditions. Aside from a predictable throat ablation mass flux reduction as combustion efficiency
is decreased, it is worth nothing how chamber pressure effect on nozzle response (i.e., chamber pressure exponent)
is practically unaffected by combustion efficiency variations. Figure 17(b) shows nozzle throat ablation mass flux
for three different combustion efficiencies at varying oxidizer-to-fuel mass ratio, using a 20 bar reference chamber
pressure. Qualitatively, the bell-like behaviour of the nozzle erosion with O/F is preserved. However, ablation mass
flux reduction due to combustion efficiency are not equally distributed by varying mixture ratio. In particular, at
stoichiometric conditions, ablation mass flux reduction for an 𝜂c∗ =0.9 is 30% if compared to the unitary combustion
efficiency case. Moving towards fuel rich O/F ratios, such a reduction increase to 47%, reaching the 60% at oxidizer
rich conditions. Hence, the shape of the bell-like curve of nozzle ablative response at varying O/F ratios shows to be
affected by combustion efficiency. Hence, considering the general regression law for HRMs previously described in

20
mixture ratio (O/F) = 3.42 (stoich)
0.8 0.8
0.7 etacstar = 1.0 chamber pressure = 20 bar
0.6 etacstar = 0.9
0.7
thorat ablation mass flux, kg/(s m2)

throat ablation mass flux, kg/(s m2)


0.5 etacstar = 0.8 etacstar = 1.0
etacstar = 0.9
0.4 etacstar = 0.8
0.6
0.3
pc0.919
0.5
0.2
0.4
pc0.933

0.1 0.3
0.963
pc
0.2

0.1

0
5 10 15 20 25 30 35 40 1 2 3 4 5 6 7 8 9 10
chamber pressure, bar mixture ratio (O/F)

(a) Chamber pressure influence. (b) Mixture ratio influence.

Fig. 17 Nozzle throat ablation mass flux at varying mixture ratios and chamber pressures for different com-
bustion efficiencies (test campaign A reference nozzle).

Eq.10, it can be rewritten in a more generic way by including combustion efficiency effects as:

¤ w,th = a(O/F, 𝜂c∗ ) pb(O/F)


m c (13)

Representing an important parameter for HRMs analysis and design, combustion efficiency effects on nozzle throat
ablation will be analyzed more in detail in future works.

VI. Conclusion
An analysis of graphite nozzle erosion behavior has been carried out with a Reynolds-averaged Navier-Stokes
equation solver with a specific application to hybrid rocket motors ablative nozzles. In particular, results obtained from
a detailed parametric numerical analysis have been used in order to derive a general regression law for nozzle throat
erosion in HRMs employing LOX-HDPE propellant combination.
The parametric numerical analysis performed on two different nozzle geometries has shown how the two most
important motor operating parameters (i.e., oxidizer-to-fuel mass ratio and motor chamber pressure) affect nozzle throat
erosion in hybrid rockets. The parametric analysis on motor chamber pressure showed a nozzle throat erosion rate
exhibiting a power exponent dependency, with a pressure exponent which is strongly affected by mixture ratio conditions.
In particular, pressure exponent is gradually reduced moving towards oxidizer rich conditions, independently on the
particular nozzle geometry considered. Numerical results have shown also how, differently from the propellant flame
temperature which is peaking at a slightly fuel rich condition, the nozzle throat erosion rate is peaking at oxidizer rich
conditions, as well as the throat wall temperature. This confirms that the nozzle throat heating and ablation processes are
mainly influenced by the exhaust gas composition and only secondarily by the propellant flame temperature. Moreover,
it is worth nothing how nozzle throat ablation peaking conditions are moved more and more in the oxidizer rich region
as the chamber pressure is decreased.
A regression law for nozzle throat ablation has been derived by a customized fitting of the numerical results obtained
in the parametric numerical analysis. The regression law proposed is able to take into account chamber pressure,
mixture ratio, and nozzle geometrical effects on nozzle ablative response with reduced computational efforts. The
comparison of the regression law with the experimental values has been performed considering separately full test
duration averages (full test measured parameters) and recession start averages (by using information coming from the
NTRT reconstruction technique). In both cases, the regression law perform well against the experiments, showing a
good ability to correctly capture the nozzle throat ablation mass flux behaviour at varying oxidizer-to-fuel mass ratio.
Nozzle geometrical effects are also correctly accounted for by properly tuning a calibration constant for each different

21
nozzle assembly employed in the experimental test campaigns analyzed. First efforts have been finally addressed in
order to include combustion efficiency effects on nozzle throat erosion in the proposed regression law. In particular,
preliminary numerical simulations have underlined how combustion efficiency is expected to an-affect nozzle throat
erosion dependency on chamber pressure, playing a significant role only at varying mixture ratios.
It is worth nothing how, for different test cases, the proposed regression law exhibits a tendency to slightly
underestimate the ablative response. This could be due to different possible parameters not accounted for in the present
analysis. Among them, wall roughness effects, as well as gas-phase reactions effects (i.e., post-combustion processes,
which are expected to be relevant especially at oxidizer rich conditions), will be investigated thoroughly in future works
in order to further increase our fundamental knowledge of the physical phenomena involved in the HRMs nozzle ablation
process.

References
[1] Sutton, G. P., and Biblarz, O., Rocket Propulsion Elements, John Wiley and Sons, Inc., New York, NY, 2001.

[2] Altman, D., and Holzman, A., Overview and History of Hybrid Rocket Propulsion, Fundamentals of Hybrid Rocket Combustion
and Propulsion, Vol. 218, edited by K. Kuo and M. Chiaverini, 2007. https://doi.org/10.2514/5.9781600866876.0001.0036, pp.
1-36.

[3] Kuo, K., and Houim, R., “Theoretical Modeling and Numerical Simulation Challenges of Combustion Processes of Hybrid
Rockets,” 47th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, 2011. https://doi.org/10.2514/6.2011-5608.

[4] Karabeyoglu, A., Zilliac, G., Cantwell, B. J., DeZilwa, S., and Castellucci, P., “Scale-Up Tests of High Regression
Rate Paraffin-Based Hybrid Rocket Fuels,” Journal of Propulsion and Power, Vol. 20, No. 6, 2004, pp. 1037–1045.
https://doi.org/10.2514/1.3340.

[5] Carmicino, C., and Sorge, A. R., “Influence of a Conical Axial Injector on Hybrid Rocket Performance,” Journal of Propulsion
and Power, Vol. 22, No. 5, 2006, pp. 984–995. https://doi.org/10.2514/1.19528.

[6] Galfetti, L., DeLuca, L., Severini, F., Colombo, G., Meda, L., and Marra, G., “Pre and post-burning analysis of nano-aluminized
solid rocket propellants,” Aerospace Science and Technology, Vol. 11, 2007, pp. 26–32. https://doi.org/10.1016/j.ast.2006.08.005.

[7] Casalino, L., and Pastrone, D., “Optimal Design of Hybrid Rocket Motors for Microgravity Platform,” Journal of Propulsion
and Power, Vol. 24, No. 3, 2008, pp. 491–498. https://doi.org/10.2514/1.30548.

[8] Kuo, K. K., and Chiaverini, M., Challenges of Hybrid Rocket Propulsion in the 21st Century, Fundamentals of Hybrid Rocket
Combustion and Propulsion, Vol. 18, edited by K. Kuo and M. Chiaverini, 2007. https://doi.org/10.2514/5.9781600866876.
0593.0638, pp. 593-638.

[9] Davydenko, N., Gollender, R., Gubertov, A., Mironov, V., and Volkov, N., “Hybrid rocket engines: The benefits and prospects,”
Aerospace Science and Technology, Vol. 11, 2007, pp. 55–60. https://doi.org/10.1016/j.ast.2006.08.008.

[10] Bunker, R., and Prince, A., “Hybrid rocket motor nozzle material predictions and results,” 28th Joint Propulsion Conference
and Exhibit, 1992. https://doi.org/10.2514/6.1992-3591.

[11] Bianchi, D., and Nasuti, F., “Numerical Analysis of Nozzle Material Thermochemical Erosion in Hybrid Rocket Engines,”
Journal of Propulsion and Power, Vol. 29, No. 3, 2013, pp. 547–558. https://doi.org/10.2514/1.B34813.

[12] Bianchi, D., Kamps, L. T., Nasuti, F., and Nagata, H., “Numerical and Experimental Investigation of Nozzle Thermochemical
Erosion in Hybrid Rockets,” 53rd AIAA/SAE/ASEE Joint Propulsion Conference, 2017. https://doi.org/10.2514/6.2017-4640.

[13] Bianchi, D., Nasuti, F., Onofri, M., and Martelli, E., “Thermochemical Erosion Analysis for Graphite/Carbon-Carbon Rocket
Nozzles,” Journal of Propulsion and Power, Vol. 27, No. 1, 2011, pp. 197–205. https://doi.org/10.2514/1.47754.

[14] Bianchi, D., and Nasuti, F., “Carbon-Carbon Nozzle Erosion and Shape-Change Effects in Full-Scale Solid-Rocket Motors,”
Journal of Propulsion and Power, Vol. 28, No. 4, 2012, pp. 820–830. https://doi.org/10.2514/1.B34267.

[15] Bianchi, D., and Nasuti, F., “Navier–Stokes Simulation of Graphite Nozzle Erosion at Different Pressure Conditions,” AIAA
Journal, Vol. 53, No. 2, 2015, pp. 356–366. https://doi.org/10.2514/1.J053154.

[16] Bianchi, D., Nasuti, F., and Onofri, M., “Radius of Curvature Effects on Throat Thermochemical Erosion in Solid Rocket
Motors,” Journal of Spacecraft and Rockets, Vol. 52, No. 2, 2015. https://doi.org/10.2514/1.A32944.

22
[17] Bianchi, D., and Neri, A., “Numerical Simulation of Chemical Erosion in Vega Solid-Rocket-Motor Nozzles,” Journal of
Propulsion and Power, Vol. 34, No. 2, 2018. https://doi.org/10.2514/1.B36388.

[18] Bianchi, D., Turchi, A., Nasuti, F., and Onofri, M., “Chemical Erosion of Carbon-Phenolic Rocket Nozzles with Finite-Rate
Surface Chemistry,” Journal of Propulsion and Power, Vol. 29, No. 5, 2013. https://doi.org/10.2514/1.B34791.

[19] Turchi, A., Bianchi, D., Nasuti, F., and Onofri, M., “A Numerical Approach for the Study of the Gas-Surface Interaction in
Carbon-Phenolic Solid Rocket Nozzles,” Aerospace Science and Technology, Vol. 27, No. 1, 2013. https://doi.org/10.1016/j.ast.
2012.06.003.

[20] Gordon, S., and McBride, B. J., “Computer Program for Calculation of Complex Chemical Equilibrium Compositions and
Applications,” Tech. Rep. RP-1311, 1994.

[21] Bianchi, D., Nasuti, F., Martelli, E., and Onofri, M., “A Numerical Approach for High-Temperature Flows over Ablating
Surfaces,” 39th AIAA Thermophysics Conference, 2007. https://doi.org/10.2514/6.2007-4537.

[22] Chelliah, H., Makino, A., Kato, I., Araki, N., and Law, C., “Modeling of graphite oxidation in a stagnation-point flow field
using detailed homogeneous and semiglobal heterogeneous mechanisms with comparisons to experiments,” Combustion and
Flame, Vol. 104, No. 4, 1996. https://doi.org/10.1016/0010-2180(95)00151-4.

[23] Kamps, L., Saito, Y., Kawabata, R., Wakita, M., Totani, T., Takahashi, Y., and Nagata, H., “Method for Determining
Nozzle-Throat-Erosion History in Hybrid Rockets,” Journal of Propulsion and Power, Vol. 33, No. 6, 2017, pp. 1369–1377.
https://doi.org/10.2514/1.B36390.

[24] Kamps, L., Hirai, S., Sakurai, K., Viscor, T., Saito, Y., Guan, R., Isochi, H., Adachi, N., Itoh, M., and Nagata, H., “Investigation
of Graphite Nozzle Erosion in Hybrid Rockets Using Oxygen/High-Density Polyethylene,” Journal of Propulsion and Power,
Vol. 36, No. 3, 2020, pp. 423–434. https://doi.org/10.2514/1.B37568.

[25] Saito, Y., Kamps, L. T., Komizu, K., Bianchi, D., Nasuti, F., and Nagata, H., “The Accuracy of Reconstruction Techniques for
Determining Hybrid Rocket Fuel Regression Rate,” 2018 Joint Propulsion Conference, 2018. https://doi.org/10.2514/6.2018-
4923.

[26] Kamps, L. T., Hirai, S., Sakurai, K., Viscor, T., Saito, Y., Guan, R., Isochi, H., Adachi, N., Itoh, M., and Nagata, H.,
“Investigation of Graphite Nozzle Erosion in Hybrid Rockets Using O2/C2H4,” 2018 Joint Propulsion Conference, 2018.
https://doi.org/10.2514/6.2018-4531.

[27] Turchi, A., Bianchi, D., Thakre, P., Nasuti, F., and Yang, V., “Radiation and Roughness Effects on Nozzle Thermochemical Erosion
in Solid Rocket Motors,” Journal of Propulsion and Power, Vol. 30, No. 2, 2014, pp. 314–324. https://doi.org/10.2514/1.B34997.

[28] Bianchi, D., and Nasuti, F., “Numerical Analysis of Nozzle Material Thermochemical Erosion in Hybrid Rocket Engines,” 48th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference Exhibit, 2012. https://doi.org/10.2514/6.2012-3809.

23

You might also like