Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

AIAA JOURNAL

Vol. 59, No. 9, September 2021

Hot-Air Anti-Icing Heat Transfer and Surface


Temperature Modeling

Ningli Chen,∗ Yaping Hu,† and Honghu Ji‡


Nanjing University of Aeronautics and Astronautics, 210016 Nanjing, People’s Republic of China
and
Min Zhang§
AVIC Chengdu Aircraft Industrial (Group) Company, Ltd.,
610092 Chengdu, People’s Republic of China
https://doi.org/10.2514/1.J059776
An inlet strut with a hot-air anti-icing system is studied experimentally and numerically to understand the heat
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

transfer between impinged supercooled water droplets and solid surfaces. Experiments were performed in an icing
research wind tunnel for dry (without water droplets) and sprayed cases. The results show that the impinged droplets
can decrease the surface temperature, even on areas without droplet impingement or a water film. Theoretically, the
impinged droplets exchange heat with solid surfaces through two heat flow rate terms: Q_ imp , which contains the
kinematic and internal energies of the impinged droplets; and the evaporation energy Q_ evap . Adding these two terms
to the solid surface when simulating the dry case provides a method to predict the surface temperature after spraying.
Calculations for Q_ evap depend on the film coverage that varies due to different operating conditions. However,
calculating Q_ evap for the impinged area or over the entire surface can give the upper and lower bounds of the
prediction for surface temperature. The results show that both methods can give predictions with errors that are
below 5% for all cases studied. The numerical results also show that both Q_ imp and Q_ evap decrease the surface
temperature, whereas the effect of Q_ imp is neglectable when compared with Q_ evap .

Nomenclature I. Introduction
heat capacity of water, J∕kg ⋅ K
Cw
g
K
=
=
=
gravity, m∕s2
momentum transfer coefficient, s−1
D URING flight, supercooled water droplets can impinge on the
front surfaces of aircraft, which may lead to ice accretion [1].
The accreted ice usually has irregular shapes, such as “hornlike ice”
m_ = mass flow rate, kg∕s on the airfoil [2,3] and “needlelike ice” on the engine spinner [4],
p = pressure, Pa which significantly change the aerodynamic shape of the aircraft and
Q_ = heat flow rate, W∕m2 block the flow paths inside the engine. Therefore, ice accretion is
T = temperature, K dangerous for flight operations [5,6]. To avoid adverse effects caused
U0 = incoming velocity, m∕s by ice accretion, different types of anti-icing systems have been
v = velocity, m∕s developed to protect the critical components that include the leading
α = volume fraction edge of the airfoil and the surface of the inlet strut in the engine.
β = local water collection coefficient Thermal anti-icing has been widely applied in the industry using
λ = heat conduction coefficient, W∕m2 ⋅ K different approaches, such as electrothermal anti-icing [7,8], heat
μ = dynamic viscosity, kg∕m ⋅ s pipe anti-icing [9,10], and hot-air anti-icing [11,12]. Hot-air anti-
ρ = density, kg∕m3 icing systems are widely used in turbojets [13,14] because they are
convenient and efficiently use hot air directly from the engine to heat
Subscripts the surfaces. Good knowledge of the heat transfer mechanism is
important to design a hot-air anti-icing system that provides full
a = air protection to the aircraft and ensures efficient operation. Hot-air
d = droplets anti-icing systems experience heat convection between the hot air
evap = evaporation and the inner solid surface, heat conduction inside the solid regime,
imp = impingement and heat convection between the cold air and the outer surface. Heat
w = water transfer between the impinged supercooled water droplets and the
outer surface also plays a critical role [15].
Some studies have considered the heat transfer mechanism in hot-
air anti-icing systems without water impingement on the outer sur-
face (dry cases). Guan et al. [13] experimentally studied the effects of
Received 8 May 2020; revision received 4 February 2021; accepted for
a chevron-shaped nozzle on the heat transfer between hot air and the
publication 4 February 2021; published online 26 March 2021. Copyright © inner conical wall. They also provided methods to enhance the heat
2021 by the American Institute of Aeronautics and Astronautics, Inc. transfer between the hot air and the inner surface through numerical
All rights reserved. All requests for copying and permission to reprint studies [16]. Yu et al. [17] experimentally studied the effect of the size
should be submitted to CCC at www.copyright.com; employ the eISSN of jet holes on the heat transfer between the hot air and the inner wall
1533-385X to initiate your request. See also AIAA Rights and Permissions in anti-icing systems for airfoils without considering the impinge-
www.aiaa.org/randp. ment of water droplets on the outer surface. Saeed [18] numerically
*Lecturer, College of Energy and Power Engineering; also China Aerody-
simulated the hot-air anti-icing system on the leading edge of an
namics Research and Development Center, 621000 Mianyang, People’s
Republic of China; chen04@foxmail.com. airfoil without impinged water droplets and optimized the arrange-

Associate Professor, College of Energy and Power Engineering. ment of the hot-air jet to enhance heat transfer.
‡ Significant efforts have also contributed to the study of anti-icing
Professor, College of Energy and Power Engineering.
§
Engineer, Aero-Engine. systems with water sprays. Dong et al. experimentally studied a
3657
3658 CHEN ET AL.

hot-air anti-icing system for an inlet strut [19], inlet guide vanes [20], provided by an electric heater, and its temperature was controlled by
and a cone wall [21]. Their studies focused primarily on the effects of a programmable logic controller system with an accuracy of 2°C.
the incoming conditions (velocity and temperature of the cold-air and The maximum temperature for the hot air was 450°C, and the maxi-
hot-air mass flow rates) on the surface temperature; however, the heat mum flow rate was about 100 g∕s. Calibrations of the LWC and MVD
transfer between the impinged droplets and the outer surface was not were conducted for different icing conditions before tests using a phase
discussed. Li et al. [22] experimentally studied a hot-air anti-icing Doppler anemometer and an icing grid. During the tests, the uncer-
system for an inlet guide vane with and without water spray. They tainties for the LWC and MVD were controlled within 0.1 g∕m3 and
revealed that without the water spray, heat convection between the 1 μm, respectively, and the cloud uniformity was controlled within
outside cold air and the solid surface accounted for approximately 20% (parameter to describe the distribution of liquid water content in
85% of the energy supplied by the inside hot air. However, the effects the IRWT). The parameters measured in real time during each test were
of the sprayed water droplets on the heat transfer were not discussed the temperature of the freestream airflow, measured by using a T-type
in their study. thermocouple with uncertainties within 0.5°C; the hot-air temper-
Different numerical methods have been proposed to simulate anti- ature, which was controlled by a programmable logic controller system
icing systems with sprayed water droplets: most of which are based on with an accuracy of 2°C; and the flow rate, which was measured
numerical models for ice accretion. The National Aeronautics and using a flow meter with an accuracy of 1%.
Space Administration (NASA) developed an anti-icing model based As depicted in Fig. 1a, the maximum and minimum chord lengths
on Messinger’s icing model [23] and implemented it into the simu- of the strut were 151.2 and 133 mm, respectively; the maximum
lation code LEWICE [24]. Validations have been performed for hot-air
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

thickness was 28.1 mm; and the length along the span direction was
anti-icing systems [25]. Other methods based on Messinger’s icing 54 mm. Seven T-shaped armored thermocouples (D  1.5 mm,
model [23] have been proposed to simulate anti-icing [26,27]; how- range: −200 to 400°C, and error: 0.2°C) were mounted in the
ever, these have not been verified for hot-air anti-icing systems. middle plane of the strut to measure the surface temperature (shown
Pellissier et al. [28] developed an anti-icing model based on their as black dots in Fig. 1a). A photograph of the experimental setup is
proposed shallow water film model icing approach [29] and added
shown in Fig. 1b, and a blackening treatment was performed on the
the model into the simulation code FENSAP. Morency et al. [30,31]
strut surface. The setup was mounted in the test section of the IRWT
added modules to simulate the hot-air anti-icing system into the code
through connections at the side plate as shown in Fig. 1b. The hot air
CANICE, which considers only one-dimensional heat conduction
was pumped into the duct from outside the IRWT to heat the leading
normal to the outer surface. In the FENSAP and CANICE codes, the
edge of the strut and prevent ice accretion. A heat preservation
outside cold-air/droplet flow and the inside hot-air flow are simulated
separately using two sets of grids. Harireche et al. [32] developed an treatment was done outside the hot-air duct to prevent heat loss as
anti-icing model that considers the details of the film flow on the shown in Fig. 1b. In the experiments, a high-speed camera (NAC-HX
surface based on the icing model provided by Myers et al. [33], Myers 1380) was used to observe the water film and ice on the surface.
and Charpin [34], and Myers et al. [35]. The experiments were first performed without a spray of super-
The impingement of the supercooled water droplets can affect the cooled water droplets (dry cases) while the hot air was pumped into
heat transfer of hot-air anti-icing systems, and the underlying mecha- the duct at a constant flow rate and temperature. The surface temper-
nism requires additional research. This paper describes experiments ature was recorded after reaching equilibrium (temperature fluctu-
conducted on a hot-air anti-icing system of an inlet strut, with and ated within 0.1°C). Then, supercooled water droplets were sprayed
without a spray from supercooled water droplets, to investigate the into the IRWT (sprayed cases), which decreased the surface temper-
effect of the heat transfer mechanism between the impinged water ature and caused ice accretion if the supplied heat was insufficient.
droplets and the solid surface. Experimental results and heat transfer Three typical conditions (runs 2, 8, and 11) are chosen for analysis in
analysis are compared to verify the proposed approach. this paper, with each operating under dry and sprayed conditions. The
detailed conditions are shown in Table 1, in which U0 and T cold are the
velocity and temperature of the incoming air/droplets, respectively;
II. Experimental Methods and Results LWC is the liquid water content of the water droplets in the incoming
A. Experimental Setup and Test Conditions flow; MVD is the medianvolume diameter of the droplets; and mhot and
An inlet strut made with aluminum and an inside hot-air duct were T hot are the mass flow rate and temperature of the hot air, respectively.
studied in the ice research wind tunnel (IRWT) of the Wuhan Aviation
Instrument Company (China). The test section of the IRWT was 800 × B. Experimental Results
250 × 350 mm in length, width, and height. In the test section, the Photographs of the strut surface after spray for the three con-
maximum air speed was 200 m∕s, the minimum temperature was ditions are shown in Fig. 2. In run 2, ice accretion was still observed
−25°C, the range of the liquid water content (LWC) of the water because the supplied heat from the hot air was insufficient. A water
droplets was 0.5–3.5 g∕m3 , and the range of the median volume film was also observed on the surface (shown more clearly in
diameter (MVD) of the droplets was 5–50 μm. The hot air was the Supplemental Video), indicating that glaze ice formed on the

a) Profiles of the strut and positions of the b) Photograph of the setup


thermocouples
Fig. 1 Experimental setup.
CHEN ET AL. 3659

Table 1 Experimental conditions for the three considered runs surface. A continuous film existed only on the leading edge, which
changed into rivulets downstream. In run 8, the heat from the hot air
Run U0 , m∕s T cold , °C mhot , g∕s T hot , °C LWC; g∕m3 MVD, μm was sufficient to evaporate all the impinged water droplets and cause a
2 Dry 60.8 −21.8 6.1 149.3 —— —— completely dry surface, as shown in Fig. 2. In run 11, the supplied
Sprayed 60.8 −21.8 6.1 149.3 0.5 20 energy from the hot air was high enough to prevent ice accretion but not
8 Dry 58.9 −11.5 30.5 253.5 —— —— to evaporate all the impinged water. Therefore, a water film flow was
Sprayed 58.9 −11.5 30.5 253.5 1.0 20 observed on the surface, as illustrated in Fig. 2 (shown more clearly in
11 Dry 60.5 −6.2 31.1 149.5 —— —— the Supplemental Video). A continuous film also existed only on the
Sprayed 60.5 −6.2 31.1 149.5 2.0 20 leading edge of the strut and became rivulets downstream.
As depicted in Fig. 3, the temperatures measured using the seven
thermocouples for all three conditions decreased after spraying. They
then reached equilibrium after a certain time (temperature fluctuated
within 0.1°C). All the temperatures discussed in the remainder of this
paper are the equilibrium values. As also shown in Fig. 3, the temper-
ature at the downstream surface requires additional time to reach
equilibrium than that closer to the stagnation point. In the image,
T1–T7 are thermocouple positions as shown in Fig. 1.
The recorded equilibrium surface temperatures under these three
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

conditions for dry and sprayed cases are illustrated in Fig. 4. The solid
lines with squares show the results of dry cases, whereas the solid
lines with triangles show the results of sprayed cases. The temper-
atures over the entire surface decreased after spraying in all three
conditions. The surface temperature in run 2 decreased below 0°C
after spraying, which is inconsistent with the ice accretion observed
in Fig. 2.

III. Simulation of Dry Cases


A. Numerical Methods
The Reynolds-averaged Navier–Stokes (RANS) method can be
used to simulate the flow and heat transfer for the dry cases, which is
done by using the commercial code FLUENT [36]. The continuum
equation for the airflow is
Fig. 2 Photographs of the experimental results under the three consid-
∂ρvai 
ered conditions after spraying. White spots on leading edge in run 2 are 0 (1)
ice, and bright lines downstream of surface in runs 2 and 11 are rivulets. ∂xi

Fig. 3 Temperature changes with time for the different thermocouples after supercooled water droplet spray.
3660 CHEN ET AL.
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

Fig. 6 Comparison of simulated surface temperature with experimental


Fig. 4 Experimental results of the surface temperature before and after (exp.) data for dry cases. Num. represents Numerical, and imp. repre-
spraying for the different cases. sents impingement.

where ρ and vai are the density and velocity of the air, respectively.
The momentum equation of air can be written as
 
∂ ∂p ∂ ∂v
ρvaj vai   −  μ ai − ρvai0 vaj
0  ρgi (2)
∂xj ∂xi ∂xj ∂xj

where −ρvai0 vaj


0
is the Reynolds stress, which is modeled using
the realizable k − ε model [37]. The energy-conservation equation
for air is
 
∂ ∂ ∂T 0 T0  S
ρvai T  λ − ρvaj h (3)
∂xi ∂xj ∂xj

where T is the temperature, λa is the heat conductivity of the air, and


Sh is the source term. The heat conduction in the solid area also
requires simulation as governed by
Fig. 7 Relative error of the numerical results from the experimental
  data.
∂ ∂T
λ  Sh0  0 (4)
∂xj ∂xj The simulated surface temperature is compared with the experimen-
tal data in Fig. 6, which are extracted from the middle section of the strut
where λ is the heat conductivity of the solid, and Sh0 is the volumetric surface (see Appendix B). In the figure, the solid symbols represent the
heat source. Equations (3) and (4) are coupled through the temper- experimental data, whereas the solid lines with open symbols show the
ature boundary conditions. numerical results. The bars on the experimental data are not the uncer-
tainties of the temperature measurements in the experiment but are used
B. Results Analysis to show the deviations of the numerical results from the experimental
data (all bars represent 5°C) throughout this work.
The computational domain used for the simulations was the same
Relative errors error  jT num − T exp j∕T exp are used to quantify
as the test section of the IRWT, as shown in Fig. 5. A structured grid
the results, and T is expressed in degrees Kelvin. The relative errors of
was used (see Appendix A), and four meshes with different numbers the numerical results for the three dry cases are shown in Fig. 7. As
of nodes were tested to determine the independence of the results on seen in the figure, the relative errors for predicted temperature are all
the mesh size. The results of the mesh-independence tests are shown below 3%, which means the simulated results match well with the
in Appendix A, and a mesh of 1,360,450 nodes was chosen for the experimental data.
discussions throughout this paper.

IV. Simulation of Sprayed Cases


A. Water Impingement Calculation
For the sprayed cases, the flow and impingement of the water
droplets can be simulated using the Eulerian method [38]. In this
method, the water droplets are treated as a continuous fluid, and their
density is represented by their volume fraction α. The continuum
equation for the water droplets is


ρ αv   0 (5)
∂xi w i

where vi and ρw are the velocity and density of the water droplets,
Fig. 5 Diagram of the computational domain used for the simulations. respectively. The small volume fraction of the water droplets allows
CHEN ET AL. 3661

Fig. 8 Water collection coefficient on the surface of the guide vane for
run 8.
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

their flow to be treated as inviscid laminar flow. Thus, the momentum


equation can be written as

∂ Fig. 10 Iterative method to calculate the surface temperature after


ρ αv v   ρw αKvai − vi   ρw αgi (6) spray.
∂xj w j i

where K is the momentum transfer coefficient between the air and where m _ evap is the evaporated mass of water that can be calculated
water droplets. using the model provided by NASA [24], and Le is the latent heat for
Equations (5) and (6) can be implemented into FLUENT using the evaporation.
UDF function [36] with four scalar-transfer equations. The velocity If adding these two energy terms (Q_ imp and Q_ evap ) from the impinged
and volume fraction of the droplets are obtained from the simulations, droplets onto the solid domain as the source terms in the numerical
and the local collection coefficient β on the wall is calculated as simulation for the dry cases, the simulated heat transfer is similar to that
of the sprayed cases. Therefore, the calculated surface temperature can
ρw αvn be used to predict the temperatures for the sprayed cases. However, the
β− (7)
LWC ⋅ U0 calculations of Q_ imp and Q_ evap are also dependent on the surface
temperature after spraying, which is unknown. The direct use of the
where vn is the normal velocity of the water droplets on the surface. surface temperature in the dry case for calculations can cause significant
The calculated collection coefficient on the strut surface for run 8 is errors. Therefore, we propose an iterative scheme to calculate the sur-
shown in Fig. 8 (results for runs 2 and 11 are shown in Appendix C). face temperature, as shown in Fig. 10, which is described as follows:
The stagnation point has the largest β, which is close to one, and 1) Simulate the dry cases and the impingement of water droplets to
gradually decreases to zero downstream. Although there is no impinge- obtain the surface temperature T ns (n is the iteration step) and the mass
ment downstream of the strut in run 8 and the surface is dry (see Fig. 2), of impinged water m _ imp.
the temperature downstream still undergoes a significant drop after 2) Use T s  f ⋅ T ns  1 − f ⋅ T n−1
s to calculate Q_ imp and Q_ evap
spraying (see Fig. 4), which implies that the heat conduction inside the (T 0s  273.15 K). Note that f is the relaxation factor that takes a
solid along the chordwise direction plays a critical role. value of 0.5 in this work.
3) Add Q_ imp and Q_ evap into the solid regime as source terms and
B. Surface Temperature Prediction recalculate the dry case to obtain the new surface temperature T n1 s .
As shown in Fig. 9, the impinged water droplets exchange heat 4) If the calculated surface temperature converges, end the calcu-
with the solid surface through two energy terms. One is the energy lation; otherwise, let n  n  1 and repeat steps 2 to 4.
brought in by the impinged water droplets The simulation convergences of all three sprayed cases (see
Table 1) are shown in Fig. 11. The plots show the temperature on
Q_ imp  0.5m
_ imp U2d  cpw m
_ imp T d − T s  (8)

where m_ imp is the mass flow rate of the impinged water droplets, Ud is
the impinged velocity of the droplets, cpw is the heat capacity of
water, T d is the temperature of the droplets, and T s is the surface
temperature. The other is the energy caused from water evaporation,
which can decrease the surface temperature:

Q_ evap  m
_ evap Le (9)

Fig. 9 Diagram for heat transfer in a hot-air anti-icing system. Fig. 11 Convergence tests for simulations of different cases.
3662 CHEN ET AL.

the middle plane of the strut. All three cases need only two iterations smaller surface temperature; together with the errors from the dry
(n  2) to converge. simulations, the results become closer to the experimental data. For
The calculation of Q_ evap depends on the water-covered region and run 8, the simulation of dry cases has smaller errors in the down-
needs additional mathematical models to simulate the film; this is not stream, therefore, the temperature curve with evaporation only in the
discussed further here. Two simple methods are provided: calculating impingement area (same as the film coverage in the experiment) gets
the evaporation only on the impinged area (shown by the solid lines closer to experimental values. Because the aluminum strut has a high
with upward triangles) or calculating the evaporation over the entire heat conductivity, which tries to flatten the surface temperature curve,
surface (shown by the solid lines with downward triangles). These the temperature difference between the two methods seems constant
two methods represent two extreme cases and can give the upper and downstream for the three cases.
lower bounds of the simulations. In the experiments, a film existing Verifications for cases of different air velocities were also done by
on the entire surface was observed for runs 2 and 11 (rivulets down- using experimental data from Ref. [20] (see Appendix D). Therefore,
stream); whereas the entire surface was dry for run 8, which means although the two models are simple in treating the film flow, they can
evaporation occurs only on the impinged area. Therefore, it can be be used for estimating the surface temperature for sprayed cases and
seen in Figs. 12a–12c that calculation of the evaporation on the entire can be extended to a more accurate model for anti-icing systems with
surface provides more accurate results for runs 2 and 11, whereas additional models for thin film flow.
calculation of the evaporation only on the impinged area gives more
accurate results for run 8, especially on the downstream surface. C. Results Analysis
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

However, both methods can give a prediction that has a relative error The impinged water droplets exchange heat with the solid surface
below 5%, as shown in Fig. 12d. Therefore, when the real film through two terms: Q_ imp and Q_ evap . To compare the effects of these
coverage is unknown, one can calculate the evaporation on the whole two terms on the heat transfer, we performed simulations that added
surface or only on the impinged area to give an estimation of the different energy sources into the solid regime: only Q_ imp, only Q_ evap,
surface temperature that has good precision (below 5% for the current and adding both Q_ imp and Q_ evap . The comparisons of the results for
three cases). different conditions are shown in Fig. 13. The evaporation was
It should also be pointed that the error in the simulation of the calculated on the impinged area for run 8 but on the entire surface
sprayed cases comes from both simulations of the dry cases and the
for runs 2 and 11. As shown in Figs. 13a–13c, when only Q_ imp was
methods for prediction of the surface temperature of sprayed cases
added, the calculated surface temperature (solid lines with open
(the simulations of the sprayed cases are based on the simulation of
diamonds) decreased slightly compared with the dry case (line with
dry cases by adding additional evaporation terms). For example, in
open squares) and was much higher than the sprayed experimental
run 8, the temperature predicted by calculating the evaporation on the
data (solid squares). The deviation from the experimental results is
whole surface was closer to the experimental data in the regime of
from about 10 to 35°C, as shown in Fig. 13d. However, when only
0.1 < x∕chord < 0.3. This is because the numerical simulation of dry
cases overpredicts the surface temperature in this region. Therefore, Q_ evap was added into the solid zone, the predicted surface temper-
simulations by calculating the evaporation only on the impinged area ature (lines with upward triangles in Fig. 13d) was much lower; and
(same as the experimental observations) also overpredict the surface very close to the experimental data for all three cases, the deviation
temperature for the sprayed case. However, calculating the evapora- from the experimental dropped within 5°C for most points (see open
tion on the whole surface can cause an overprediction of evaporation symbols with dotted centers in Fig. 13d). Therefore, both Q_ imp and
when compared with the experimental observations and leads to a Q_ evap decrease the surface temperature predictions, whereas the

Fig. 12 Comparison of calculated surface temperature with evaporation over entire surface and on impact area for three different runs. Bars of 5°C on
experimental data are to show deviation of numerical results from experimental data.
CHEN ET AL. 3663
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

Fig. 13 Comparison of heat transfer simulations with impact and evaporation (evap.) energy terms for three different runs. Bars of 5°C on experimental
(exp.) data show deviation of numerical results (num.) from experimental data.

effect of Q_ imp is neglectable when compared to Q_ evap . When both temperature. Different cases have been examined with these two
Q_ imp and Q_ evap are considered in the simulation, a more precise methods, and the results show that both methods can give predictions
with errors that are still typically below 5%. The numerical results
prediction is obtained (see solid symbols in Fig. 13d).
showed that both Q_ imp and Q_ evap caused a decreased surface temper-
ature after spraying, whereas the effect of Q_ imp is neglectable when
V. Conclusions compared to Q_ evap . The proposed model is suitable for quickly
An inlet strut with a hot-air anti-icing apparatus has been exper- estimating the surface temperature for sprayed cases, and it can be
imentally studied in an IRWT, and the heat transfer between the extended to a more accurate model for anti-icing systems with addi-
impinged supercooled water droplets and the solid surface was tional models for thin film flow.
analyzed. The surface temperature was measured using thermocou-
ples for dry and sprayed cases. The temperature over the entire strut
surface decreased after spraying supercooled water droplets, even in Appendix A: Mesh Independence Test
regions where there was no impingement or film due to the heat The structured grid used for the simulation is shown in Fig. A1.
conduction inside the solid. After spraying, the surface temperature Four meshes with different numbers of nodes (1,068,720, 1,360,450,
reached equilibrium over a certain time. 1,812,690, and 2,048,220) were used to test the independence of the
Numerical simulations were also adopted to better understand the numerical results on the mesh size. The test conditions of run 2 were
heat transfer mechanism. The RANS method was used to simulate the selected for the analysis. The surface temperatures (middle section of
dry cases, which predicted the surface temperatures with errors below the strut) for the dry cases are shown in Fig. A2. The results converge
3%. The impingement of the water droplets was simulated using the
Eulerian method. A simple method was proposed to predict the
surface temperature for sprayed cases based on analysis of the heat
transfer. The impinged water droplets introduced two extra energy
terms, Q_ imp and Q_ evap , into the system as compared with the dry
cases. Adding these two terms into the solid as the energy sources to
simulate the dry cases made the simulated heat transfer similar to that
in the sprayed cases; thus, the calculated surface temperature can be
used to predict the sprayed cases. An iterative method was used to
calculate the surface temperature due to the dependencies of Q_ imp and
Q_ evap . The calculation of Q_ evap depends on the water film coverages,
which can vary from only on the impinged area to on the entire
surface due to different anti-icing conditions. The errors of the
predicted surface temperature for all three cases are below 5% when
calculating Q_ evap based on observations of film coverage in the
experiments. However, when the film coverage is unknown, calcu-
lating Q_ evap only on the impinged area or over the entire surface
can give the upper and lower bounds of the prediction for surface Fig. A1 Structured grid used for the simulations.
3664 CHEN ET AL.

Fig. C1 Water collection coefficients on the strut for run 2.


Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

Fig. A2 Results of the mesh independence analysis for run 2 (dry case).

when the total number of nodes is larger than 1,360,450, Therefore, a


mesh size of 1,360,450 was chosen for this study.
Fig. C2 Water collection coefficients on the strut for run 11.
Appendix B: Temperature Results on the Middle Section
of the Strut
The numerical results for the surface temperature of the dry cases
under the conditions of run 8 are shown in Fig. B1. The position of the
middle section is also shown in the figure.

Appendix C: Water Collection Coefficient on the Strut


The simulated water collection coefficients on the strut for runs 2
and 11 are shown in Figs. C1 and C2, respectively. Similar to the
results in Fig. 7, the impingement of the supercooled water droplets
occurs at the leading edge of the strut for all considered cases.

Appendix D: Verifications of the Method for Cases


Fig. D1 Mixed grid used for the simulations.
of Different Air Velocities
Here, we use the experimental data from Ref. [20], which has
different Reynolds numbers of the air, to further verify the method. verification. The mesh used for the simulation is a mixed grid as
The strut in Ref. [20] is different from that in the current paper, which shown in Fig. D1.
has a smaller chord of 75 mm. Three cases (runs 2, 5, and 6) with The simulated surface temperature for these three different cases is
different air velocities (140, 60, and 40 m∕s) were chosen for shown in Fig. D2. Two different methods were used: calculating

Fig. B1 Simulated surface temperature of run 8 before spray (dry case).


CHEN ET AL. 3665
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

Fig. D2 Mixed grid used for the simulations.

evaporation only on the impinged area, and calculating the evaporation [10] Su, Q., Chang, S., Zhao, Y., Zheng, H., and Dang, C., “A Review of Loop
on the whole surface. As can be seen, errors for both methods are within Heat Pipes for Aircraft Anti-Icing Applications,” Applied Thermal
5 deg in run 5, although errors in runs 2 and 6 are a little bit larger than Engineering, Vol. 130, Feb. 2018, pp. 528–540.
5 deg at some points. However, generally, the predictions are accept- https://doi.org/10.1016/j.applthermaleng.2017.11.030
[11] Bu, X., Peng, L., Lin, G., Bai, L., and Wen, D., “Experimental Study of
able in engineering as a quick estimation for anti-icing design.
Jet Impingement Heat Transfer on a Variable-Curvature Concave Sur-
face in a Wing Leading Edge,” International Journal of Heat and Mass
Transfer, Vol. 90, Nov. 2015, pp. 92–101.
References https://doi.org/10.1016/j.ijheatmasstransfer.2015.06.028
[1] Feng, K., Lu, Z., and Yun, W., “Aircraft Icing Severity Analysis Con- [12] Chen, L., Zhang, Y., and Wu, Q., “Heat Transfer Optimization and
sidering Three Uncertainty Types,” AIAA Journal, Vol. 57, No. 4, 2019, Experimental Validation of Anti-Icing Component for Helicopter Rotor,”
pp. 1514–1522. Applied Thermal Engineering, Vol. 127, Dec. 2017, pp. 662–670.
https://doi.org/10.2514/1.J057529 https://doi.org/10.1016/j.applthermaleng.2017.07.169
[2] Janjua, Z. A., Turnbull, B., Hibberd, S., and Choi, K.-S., “Mixed Ice [13] Guan, T., Zhang, J.-Z., Shan, Y., and Hang, J., “Conjugate Heat Transfer
Accretion on Aircraft Wings,” Physics of Fluids, Vol. 30, No. 2, 2018, on Leading Edge of a Conical Wall Subjected to External Cold Flow
Paper 027101. and Internal Hot Jet Impingement from Chevron Nozzle—Part 1:
https://doi.org/10.1063/1.5007301 Experimental Analysis,” International Journal of Heat and Mass Trans-
[3] Chen, N., Ji, H., Cao, G., and Hu, Y., “AThree-Dimensional Mathemati- fer, Vol. 106, March 2017, pp. 329–338.
cal Model for Simulating Ice Accretion on Helicopter Rotors,” Physics
https://doi.org/10.1016/j.ijheatmasstransfer.2016.06.101
of Fluids, Vol. 30, No. 8, 2018, Paper 083602.
[14] Mazumder, I., Saha, R., Gulen, S. C., Devarajan, S., and Kumar, P., “Gas
https://doi.org/10.1063/1.5041896
Turbine Anti-Icing System,” U.S. Patent 9447732, filed 20 Sept. 2016.
[4] Chen, N., Ji, H., Hu, Y., Wang, J., and Cao, G., “Experimental Study of [15] Villalpando, F., Reggio, M., and Ilinca, A., “Prediction of Ice Accretion
Icing Accretion on a Rotating Conical Spinner,” Heat and Mass Trans- and Anti-Icing Heating Power on Wind Turbine Blades Using Standard
fer, Vol. 51, No. 12, 2015, pp. 1717–1729.
Commercial Software,” Energy, Vol. 114, Nov. 2016, pp. 1041–1052.
https://doi.org/10.1007/s00231-015-1536-0
https://doi.org/10.1016/j.energy.2016.08.047
[5] Cao, Y., Wu, Z., Su, Y., and Xu, Z., “Aircraft Flight Characteristics in
[16] Guan, T., Zhang, J.-Z., and Shan, Y., “Conjugate Heat Transfer on
Icing Conditions,” Progress in Aerospace Sciences, Vol. 74, April 2015,
Leading Edge of a Conical Wall Subjected to External Cold Flow and
pp. 62–80.
https://doi.org/10.1016/j.paerosci.2014.12.001 Internal Hot Jet Impingement from Chevron Nozzle—Part 2: Numerical
[6] Cole, J., and Sand, W., “Statistical Study of Aircraft Icing Accidents,” Analysis,” International Journal of Heat and Mass Transfer, Vol. 106,
29th Aerospace Sciences Meeting, AIAA Paper 1991-558, 1991. March 2017, pp. 339–355.
[7] Mohseni, M., and Amirfazli, A., “A Novel Electro-Thermal Anti-Icing https://doi.org/10.1016/j.ijheatmasstransfer.2016.10.048
System for Fiber-Reinforced Polymer Composite Airfoils,” Cold [17] Yu, J., Peng, L., Bu, X., Shen, X., Lin, G., and Bai, L., “Experimental
Regions Science and Technology, Vol. 87, March 2013, pp. 47–58. Investigation and Correlation Development of Jet Impingement Heat
https://doi.org/10.1016/j.coldregions.2012.12.003 Transfer with Two Rows of Aligned Jet Holes on an Internal Surface of a
[8] Ding, L., Chang, S., Yang, S., and Leng, M., “Study on Heating Wing Leading Edge,” Chinese Journal of Aeronautics, Vol. 31, No. 10,
Strategies of Nose Cone Electrothermal Anti-Icing,” Journal of Aircraft, 2018, pp. 1962–1972.
Vol. 55, No. 3, 2018, pp. 1311–1316. https://doi.org/10.1016/j.cja.2018.07.016
https://doi.org/10.2514/1.C034639 [18] Saeed, F., “Numerical Simulation of Surface Heat Transfer from an Array
[9] Lian, W., and Xuan, Y., “Experimental Investigation on a Novel Aero- of Hot-Air Jets,” Journal of Aircraft, Vol. 45, No. 2, 2008, pp. 700–714.
Engine Nose Cone Anti-Icing System,” Applied Thermal Engineering, https://doi.org/10.2514/1.33489
Vol. 121, July 2017, pp. 1011–1021. [19] Dong, W., Zheng, M., Zhu, J., Lei, G., and Zhao, Q., “Experimental
https://doi.org/10.1016/j.applthermaleng.2017.04.160 Investigation on Anti-Icing Performance of an Engine Inlet Strut,” Journal
3666 CHEN ET AL.

of Propulsion and Power, Vol. 33, No. 2, 2017, pp. 379–386. [30] Morency, F., Brahimi, M., Tezok, F., and Paraschivoiu, I., “Hot Air Anti-
https://doi.org/10.2514/1.B36067 Icing System Modelization in the Ice Prediction Code CANICE,” 36th
[20] Dong, W., Zhu, J., Zheng, M., Lei, G., and Zhou, Z., “Experimental Study AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper 1997-
on Icing and Anti-Icing Characteristics of Engine Inlet Guide Vanes,” 0197, 1997.
Journal of Propulsion and Power, Vol. 31, No. 5, 2015, pp. 1330–1337. [31] Morency, F., Tezok, F., and Paraschivoiu, I., “Anti-Icing System Simu-
https://doi.org/10.2514/1.B35679 lation Using CANICE,” Journal of Aircraft, Vol. 36, No. 6, 1999,
[21] Dong, W., Zhu, J., and Zheng, M., “Thermal Analysis and Testing of a pp. 999–1006.
Cone with Leading Edge Hot Air Anti-Icing System,” 52nd Aerospace
https://doi.org/10.2514/2.2541
Sciences Meeting, AIAA Paper 2014-0739, 2014.
[32] Harireche, O., Verdin, P., Thompson, C. P., and Hammond, D. W.,
[22] Li, L., Liu, Y., Tian, L., Hu, H., Hu, H., Liu, X., Hogate, I., and Kohli, A.,
“Explicit Finite Volume Modeling of Aircraft Anti-Icing and De-Icing,”
“An Experimental Study on a Hot-Air-Based Anti-/De-Icing System
for Aero-Engine Inlet Guide Vanes,” Applied Thermal Engineering, Journal of Aircraft, Vol. 45, No. 6, 2008, pp. 1924–1936.
Vol. 167, Feb. 2020, Paper 114778. https://doi.org/10.2514/1.34855
https://doi.org/https://doi/10.1016/j.applthermaleng.2019.114778 [33] Myers, T. G., Charpin, J. P., and Chapman, S. J., “The Flow and
[23] Messinger, B. L., “Equilibrium Temperature of an Unheated Icing Sur- Solidification of a Thin Fluid Film on an Arbitrary Three-Dimensional
face as a Function of Air Speed,” Journal of Aeronautical Sciences, Surface,” Physics of Fluids, Vol. 14, No. 8, 2002, pp. 2788–2803.
Vol. 20, No. 1, 1953, pp. 29–42. https://doi.org/10.1063/1.1488599
https://doi.org/10.2514/8.2520 [34] Myers, T. G., and Charpin, J. P., “A Mathematical Model for Atmospheric
[24] Wright, W., “User’s Manual for LEWICE Version 3.2,” NASA CR- Ice Accretion and Water Flow on a Cold Surface,” International Journal of
214255, 2008. Heat and Mass Transfer, Vol. 47, No. 25, 2004, pp. 5483–5500.
Downloaded by SHANGHAI JIAO TONG UNIVERSITY on July 14, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.J059776

[25] Wright, W., “LEWICE 2.2 Capabilities and Thermal Validation,” 40th https://doi.org/10.1016/j.ijheatmasstransfer.2004.06.037
AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper 2002-0383, [35] Myers, T., Charpin, J., and Thompson, C., “Slowly Accreting Ice due to
2002. Supercooled Water Impacting on a Cold Surface,” Physics of Fluids,
[26] Zhu, J., Dong, W., Zheng, M., Lei, G., and Zhao, Q., “Numerical Vol. 14, No. 1, 2002, pp. 240–256.
Investigation of Heat and Mass Transfer on an Anti-Icing Inlet Cone,” https://doi.org/10.1063/1.1416186
Journal of Propulsion and Power, Vol. 32, No. 3, 2016, pp. 789–797. [36] ANSYS FLUENT, User’s Guide, Customization Manual, Soft Package
https://doi.org/10.2514/1.B35875 Release 17.0, ANSYS, Inc., April 2012.
[27] Bu, X., Lin, G., Shen, X., Hu, Z., and Wen, D., “Numerical Simulation of [37] Lavoie, P., “Modeling of Thin Water Films on Swept Wings in Icing
Aircraft Thermal Anti-Icing System Based on a Tight-Coupling Method,” Condition,” Ph.D. Thesis, École Polytechnique de Montréal, Montreal,
International Journal of Heat and Mass Transfer, Vol. 148, Feb. 2020,
2017.
Paper 119061.
[38] Bourgault, Y., Habashi, W. G., Dompierre, J., and Baruzzi, G. S., “A
https://doi.org/10.1016/j.ijheatmasstransfer.2019.119061
Finite Element Method Study of Eulerian Droplets Impingement Mod-
[28] Pellissier, M., Habashi, W., and Pueyo, A., “Optimization via FENSAP-
ICE of Aircraft Hot-Air Anti-Icing Systems,” Journal of Aircraft, els,” International Journal for Numerical Methods in Fluids, Vol. 29,
Vol. 48, No. 1, 2011, pp. 265–276. No. 4, 1999, pp. 429–449.
https://doi.org/10.2514/1.C031095 https://doi.org/10.1002/(SICI)1097-0363(19990228)29:4<429::AID-
[29] Bourgault, Y., Beaugendre, H., and Habashi, W., “Development of a FLD795>3.0.CO;2-F
Shallow-Water Icing Model in FENSAP-ICE,” Journal of Aircraft,
Vol. 37, No. 4, 2000, pp. 640–646. C. Wen
https://doi.org/10.2514/2.2646 Associate Editor

You might also like