The performance of a turbulent air jet impinging on a flat surface and subjected to cross flow is investigated. The study considers many different parameters, including Reynolds number Re, the normalized orifice-to-target spacing Z/D, and the ratio α of cross-flow to jet Reynolds numbers. As reported by other investigators, the optimum value of Z/D is 6. At this optimum Z/D, when the Reynolds number is raised from 5100 to 23 000, the stagnation Nusselt number Nu increases by 150%. A correlation is proposed to estimate stagnation point heat transfer. Velocity profiles are used to examine the cooling characteristics of jets with and without cross flow. Cooling in both the upstream and downstream regions is estimated and compared with the performance of a jet without cross flow. For Z/D = 6, the stagnation Nu decreases by 84% when α is increased from 0 to 2. The corresponding decrease in the stagnation Nu for Z/D = 2 is 75%. Cross flow significantly affects the flow pattern of fluid near the target surface, and anomalous flow behavior is observed in the stagnation region. Without cross flow, the radial velocity is close to zero at r/D = 0, and the fluid velocity increases in the outward direction and attains a maximum value at r/D = 1. With cross flow, the position of the peak velocity shifts. Average, stagnation point, and peak heat transfer coefficients are discussed in terms of the Stanton number. At low values of α, there is an increase in the peak Nu. With cross flow, the location of the peak velocity changes. The augmentation in cooling in the downstream region and the reduction in cooling in the upstream region depend strongly on Z/D and α. For Z/D = 2, at α = 1, the average heat transfer downstream increases by 15% and the corresponding decrease upstream is 35%.

D

orifice diameter (m)

Dc

cross-flow inlet hydraulic diameter (m)

h

coefficient of heat transfer [W/(m2 · K)]

k

thermal conductivity of air [W/(m · K)]

Nu

Nusselt number

Nu0

stagnation Nusselt number

Pr

Prandtl number

q

heat flux (W/m2)

r

radial distance from center of plate (m)

Re

jet Reynolds number

Rec

cross-flow Reynolds number

St

Stanton number

St0

Stanton number at stagnation point

Sta

average Stanton number

Stp

Stanton number at location of peak heat transfer

Tj

air jet temperature (K)

Tw

target surface temperature (K)

u

radial velocity (m/s)

Vc

cross-flow velocity (m/s)

Vj

jet speed (m/s)

Z

orifice-to-target distance (m)

Greek
α

ratio of cross-flow to jet Reynolds numbers

ν

kinematic viscosity of air (m2/s)

Jet cooling is an attractive cooling method owing to its superior heat transfer capability. An impinging jet produces strong turbulence near the target surface, and the boundary layer becomes narrower. In single-jet cooling, the parameters affecting the heat transfer rate include among others, the Reynolds number Re, the fluid properties, the size and shape of the nozzle, the normalized orifice-to-target spacing Z/D, and the steadiness of the flow. The characteristics of jet cooling are also influenced by the flow from the surroundings. In jet cooling, heat transfer coefficients as high as 100 kW/(m2 · K) are possible with single-phase fluid.1 Lytle and Webb2 studied the influence of Z/D on jet performance and found that secondary peaks occur at low Z/D and high Re, owing to increased turbulence. Correlations have been proposed to calculate the Nusselt number Nu in the three regions near the impinging wall.3 Martin4 presented an extensive overview of jet cooling, with a focus on applications and empirical formulas for predicting heat transfer coefficients. The important variables identified were the Reynolds number, the properties of the cooling medium, the size and shape of the nozzle, and Z/D. Empirical correlations have been proposed for predicting the heat transfer coefficients for a single round nozzle, a slot nozzle, and arrays of these nozzles. Zuckerman and Lior5 presented a comprehensive review of jet impingement cooling. They described the physics of the jet flow, the heat transfer mechanism, and the suitability of various turbulent models for predicting the flow and heat transfer. They listed the various correlations used for predicting heat transfer. Bouchez and Goldstein6 reported that under certain cross-flow conditions, the effectiveness at the stagnation point is as high as 0.6. Metzger and Korstad7 studied the effect of cross flow on multiple jets and found that the cooling is affected by both Z/D and the normalized spacing between jets. They proposed a correlation for the average Stanton number. Studies at Re > 35 000 revealed that at low Z/D, there is a slight improvement in the maximum Nu at low cross flow. However at high Z/D, peak Nu decreases with increase in cross flow.8 Studies of cross flow by Huang et al.9 indicated that changes in heat transfer depend on the cross-flow velocity and direction. Jet impingement studies on a cylinder showed that Nu is influenced by cross flow and Kelvin–Helmholtz instability.10 Wang et al.11 studied the effects of cross flow on an inclined jet and reported that the cross flow decreases the peak Nu but has no influence on wall jet heat transfer. In multiple-jet cooling, a staggered configuration has superior characteristics compared with other arrangements.12 Interaction with neighboring jets and the turbulence produced affect the cooling. Experimental investigations indicate that cross flow improves the peak Nu in comparison with a pure jet.13 In gas turbine blade cooling, diverters can minimize the effect of cross flow.14 Cooling studies have revealed that cross flow considerably affects heat transfer. Higher values of peak Nu are obtained at high Mach number, lower D, and lower Z/D.15 Nonuniformity in temperature due to cross flow can be minimized by increasing the impingement Re and reducing Z/D.16 Masip et al.17 studied heat transfer from a hot surface placed in a channel subjected to jet flow and lateral flow. Lateral flow with jet-to-cross-flow ratios of 1.0 and 1.5 produced maximum cooling. In multiple-jet cooling, the influence of cross flow from adjacent jets can be minimized by providing diverters.18 Cylindrical diverters give the maximum enhancement of heat transfer (5.2%). Choudhari et al.19 studied the combined effect of synthetic jets and cross flows on cooling of a channel and reported that the introduction of cross flow reduces heat transfer by 10%.

The aim of the present study is to investigate the changes in flow field and cooling of a single jet due to lateral flow. A correlation is proposed to calculate Nu without cross flow. This study provides new insight into the cooling performance of jets with cross flow.

Details of the domain considered is shown in Fig. 1. The target plate (stainless steel) has dimensions 160 × 80 × 0.06 mm3.3 The face abfe is the target surface to be cooled, and the heat inflow is from the bottom surface of the target plate. Th face abcd is the cross-flow inlet. All other sides are open (outlets), and pressure boundary conditions are applied. The nozzle diameter is 7.35 mm, and the length-to-diameter ratio is 7. For meshing the computational domain, an unstructured grid system with an inflation layer near the impinging surface is employed (Fig. 2). An appropriate y+ value was obtained by cell refinements in the near-wall regions and is close to 1. This is done to precisely capture the process of heat transfer in the area near the target surface.

FIG. 1.

Computational domain. The face abfe is the surface to be cooled, abcd is the cross-flow inlet, and all other sides are open.

FIG. 1.

Computational domain. The face abfe is the surface to be cooled, abcd is the cross-flow inlet, and all other sides are open.

Close modal
FIG. 2.

Two-dimensional view of the 3D computational domain with mesh generated.

FIG. 2.

Two-dimensional view of the 3D computational domain with mesh generated.

Close modal

The turbulent scheme used is the renormalization group (RNG) k-ε model. This turbulence model is very successful in predicting the thermohydraulic behavior of impinging jets.20,21 For pressure–velocity coupling, the SIMPLE algorithm is used. A second-order upwind method is used for momentum and energy, and first-order equations are employed for the turbulence parameters. The convergence criterion used is <10−6 for residuals of the continuity, momentum, and energy. The temperature of the inlet air is 300 K. At open sides other than inlets, pressure boundary conditions are applied. The flow is assumed to be steady, turbulent, and incompressible. The effects of temperature on the thermophysical properties of air are taken into account.

The equations used are as follows:

Heat transfer coefficient
(1)
Nusselt number
(2)
Jet Reynolds number
(3)
cross-flow Reynolds number
(4)
ratio of cross-flow to jet Reynolds number
(5)
Stanton number
(6)

The influence of Re, Z/D, and α on the heat transfer performance of a turbulent jet impinging on a flat surface is studied with the help of the flow field.

Figure 3 shows the results of the mesh-independence study. The grid is improved by raising inflation and lowering edge sizing. At both values of Re, the optimum number of elements is found to be 3.32 × 106.

FIG. 3.

Mesh-independence analysis for Z/D = 6.

FIG. 3.

Mesh-independence analysis for Z/D = 6.

Close modal

Figure 4 depicts the stream-wise variation of Nu. The Reynolds numbers considered are 5100 and 23 000. The experimental results obtained in some other studies2,3,21 are also shown for comparison. The results from the present numerical study are in fair agreement with experiment at both Reynolds numbers. Under these flow conditions, Nu is maximum at the stagnation point and decreases radially.

FIG. 4.

Local Nu variation for Z/D = 6: (a) Re = 5100; (b) Re = 23 000.24 

FIG. 4.

Local Nu variation for Z/D = 6: (a) Re = 5100; (b) Re = 23 000.24 

Close modal

The Nusselt number distributions at different Reynolds numbers and Z/D values are displayed in Fig. 5. For Z/D = 6 at all Reynolds numbers, the curves are qualitatively similar. For Z/D = 2, at higher Reynolds numbers, there are secondary peaks in the transition region due to the high turbulent kinetic energy.

FIG. 5.

Stream-wise distribution of Nu: (a) Z/D = 2; (b) Z/D = 6.24 

FIG. 5.

Stream-wise distribution of Nu: (a) Z/D = 2; (b) Z/D = 6.24 

Close modal

Figure 6 shows the effect of Reynolds number on stagnation Nusselt number. Nu0 increases with increasing Reynolds number. For Z/D = 6, when the Reynolds number is raised from 5100 to 23 000, the Nu0 increases by 150%. The corresponding increase for Z/D = 2 is 156%. As Re increases, the jet centerline velocity increases and so does the turbulence close to the wall, and both help to increase Nu0.

FIG. 6.

Dependence of stagnation Nusselt number Nu0 on Re.

FIG. 6.

Dependence of stagnation Nusselt number Nu0 on Re.

Close modal
Figure 7 shows the influence of Z/D and Re on the local Nu, with Re being varied from 5100 to 23 000 in steps of 1000, and Z/D being varied from 2 to 8 in steps of 1. Only six cases are shown for clarity in distinguishing the various cases. The curves are like and bell-shaped at low Re. On the basis of the numerical data, a correlation is proposed to describe the dependence of the stagnation Nusselt number on Reynolds number and Z/D. This correlation is based on those available in the literature and is obtained by fitting the numerical data using regression analysis. It is given by
(7)
and is applicable for 5100 ≤ Re ≤ 23 000 and 2 ≤ Z/D ≤ 6, with air as the cooling medium.
FIG. 7.

Dependence of Nu on Re and Z/D.

FIG. 7.

Dependence of Nu on Re and Z/D.

Close modal

Figure 8 shows the dependence of the stagnation heat transfer (based on the correlation proposed) on Z/D. Data from other studies3,22,23 are also shown. Equation (7) is successful in predicting stagnation heat transfer (within 6%).

FIG. 8.

Appraisal of the proposed correlation given by Eq. (7).

FIG. 8.

Appraisal of the proposed correlation given by Eq. (7).

Close modal

Figure 9 shows the variation of the normalized gauge pressure (close to the plate) in the radial direction. Cross flow causes the profile to become asymmetrical, and the peak pressure location shifts radially. Mixing of the cross-flow stream coming inward and the jet stream results in recirculation, which causes creation of a low-pressure region (see Fig. 11).

FIG. 9.

Pressure distribution.

FIG. 9.

Pressure distribution.

Close modal

Figure 10(a) shows the radial velocity profiles without external flow. At the jet axis, i.e., r/D = 0, the radial velocity is close to zero, since the jet is perpendicular to the plate.

FIG. 10.

Distribution of nondimensionalized radial velocity for Re = 5100: (a) α = 0; (b) upstream of the jet, α = 1.5; (c) downstream of the jet, α = 1.5. (Symbols are used to distinguish between various cases and do not represent experimental data.)

FIG. 10.

Distribution of nondimensionalized radial velocity for Re = 5100: (a) α = 0; (b) upstream of the jet, α = 1.5; (c) downstream of the jet, α = 1.5. (Symbols are used to distinguish between various cases and do not represent experimental data.)

Close modal

Near the surface, the flow is accelerated radially and attains its peak velocity at a normalized radial distance r/D = 1. Farther downstream, the velocity decreases because of the thickening of the boundary layer. Higher radial velocities are observed for smaller Z/D. For example, for Z/D = 6, the peak value of the normalized velocity is 0.64, and the corresponding value for Z/D = 2 is 0.91. At very low Z/D, the radial velocity can be more than the jet velocity.

Figure 10(b) illustrates the radial velocity profile upstream (r/D > 0) of jet in the presence of cross flow. For Z/D = 6, the peak radial velocity is obtained at r/D = 0 and it diminishes radially up to r/D = 1. With cross flow, the jet becomes as an inclined jet and the stagnation condition is not attained. In the region r/D > 1, the flow is radially inward. For Z/D = 2, the cross flow has a considerable influence, and in all regions (r/D ≥ 0), the flow is in the cross-flow direction.

Figure 10(c) depicts the radial velocity profile downstream of the jet (r/D < 0) in the presence of an external flow. For both Z/D values, in a small region (−0.5 < r/D < 0) downstream, the flow near the target surface is opposite to the cross-flow, and beyond r/D = −0.5, the flow direction changes. The anomalous flow in the zone −0.5 < r/D < 0.5 is due to mixing of the main stream and the external flow (Fig. 11). The maximum velocity occurs at r/D = −1, and it diminishes in the downstream direction.

FIG. 11.

Velocity vectors in a midplane.

FIG. 11.

Velocity vectors in a midplane.

Close modal

Figure 11 illustrates the velocity vectors in a midplane coinciding with the axis of the jet. The jet is deflected owing to cross flow, and there in mixing of the jet flow and the lateral flow. At the midpoint of the target, the stagnation condition is not achieved.

Figure 12 depicts the influence of α on the local Nu. Cross flow can occur from neighboring jets or other cooling equipment. The cross flow deflects the jet downstream, and the curves are not symmetrical about the centerline. For both values of Z/D, at smaller α, the peak Nu does not vary significantly. In fact, for Z/D = 2, there is a slight increase in the peak Nusselt number, and this is in line with the findings of other investigators. A significant reduction in heat transfer occurs at higher values of α.

FIG. 12.

Radial variation of Nu with cross flow for Re = 5100: (a) Z/D = 6; (b) Z/D = 2.24 

FIG. 12.

Radial variation of Nu with cross flow for Re = 5100: (a) Z/D = 6; (b) Z/D = 2.24 

Close modal

The cooling effect upstream of the jet axis is decrease, while that downstream increases. From Fig. 10, it is clear that owing to cross flow, the fluid upstream decelerates and that downstream accelerates. For example, for Z/D = 2, in the absence of cross flow, the normalized radial velocity at a distance r/D = 1 is 0.9, and with cross flow, it decreases to 0.12. Similarly, without cross flow, the normalized velocity at r/D = −1 is 0.9, and there is acceleration in the presence of cross flow.

Figure 13 shows the effect of cross flow on cooling in terms of the Stanton number at the stagnation point (St0) and at the location of peak heat transfer (Stp). For low Z/D, cross flow with low α helps to improve cooling. At higher values of α, the jet is deflected, and the jet fluid impinges on the plate at an acute angle, and there is also mixing between the two fluid streams.

FIG. 13.

Influence of Z/D on (a) St0 and (b) Stp.

FIG. 13.

Influence of Z/D on (a) St0 and (b) Stp.

Close modal

When the jet impinges orthogonally on the surface, extreme cooling occurs at the stagnation point/center of the plate, and when the jet is inclined owing to cross flow, the peak heat transfer occurs downstream. At higher α, there is a considerable reduction in Sto and Stp. Cross flow has a greater influence at lower Z/D.

Figure 14 shows the ratio of the average Stanton number Sta in the upstream and downstream zones with and without cross flow. Cross flow reduces Sta in the upstream zone irrespective of α. Sta in the downstream zone increases with α to a certain value and then decreases with further increase in α. As α increases, the angle at which the jet impinges decreases and there is more circulation owing to the mixing of the two streams. Improved cooling downstream and reduced cooling upstream depend strongly on Z/D and α. The augmentation in cooling downstream and the reduction in cooling upstream are greater at lower Z/D. For Z/D = 2, at α = 1, the value of Sta downstream increases by 15%, and the corresponding decrease upstream is 35%.

FIG. 14.

Influence of Z/D on average Stanton number.

FIG. 14.

Influence of Z/D on average Stanton number.

Close modal

The cooling characteristics of an air jet exposed to cross flow have been studied numerically under a constant wall heat flux condition. Heat transfer coefficients have been discussed in terms of the Nusselt number and Stanton number. The Stanton number in the presence of cross flow has been normalized by dividing it by the corresponding value for lateral flow. The major findings are as follows:

  1. As reported by other investigators, the optimum Z/D is 6. At this Z/D, when the Reynolds number is raised from 5100 to 23 000, the stagnation Nusselt number increases by 150%.

  2. The proposed correlation is successful in calculating Nu0 in the absence of cross flow.

  3. At higher values of α, cross flow results in a strong interaction between the fluid streams and causes recirculation.

  4. Cross flow significantly affects the heat transfer and flow pattern of fluid near the target surface. For Z/D = 6, the stagnation Nu decreases by 84% when α is increased from 0 to 2. The corresponding decrease for Z/D = 2 is 75%.

  5. With cross flow, anomalous flow behavior is observed in the zone −0.5 < r/D < 0.5, owing to recirculation near the wall resulting from the mixing of the two flows.

  6. Without cross flow, the peak radial velocity occurs at r/D = 1, and cross flow causes the position of this peak velocity to shifts.

  7. At low values of α and Z/D, cross flow results in a marginal improvement in peak Nu.

  8. The augmentation in cooling in the downstream region and the reduction in cooling in the upstream region depend strongly on Z/D and α.

  9. Cross flow has a stronger negative impact at lower Z/D and higher α.

The authors have no conflicts to disclose.

Vivek Mathew Jose: Conceptualization (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Software (lead); Validation (lead); Writing – original draft (lead); Writing – review & editing (lead).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
S. V.
Garimella
and
R. A.
Rice
, “
Confined and submerged liquid jet impingement heat transfer
,”
J. Heat Transfer
117
(
4
),
871
877
(
1995
).
2.
D.
Lytle
and
B. W.
Webb
, “
Air jet impingement heat transfer at low nozzle plate spacing
,”
Int. J. Heat Mass Transfer
28
,
355
363
(
1994
).
3.
V.
Katti
and
S. V.
Prabhu
, “
Experimental study and theoretical analysis of local heat transfer distribution between smooth flat surface and impinging air jet from a circular straight pipe nozzle
,”
Int. J. Heat Mass Transfer
51
,
4480
4495
(
2008
).
4.
H.
Martin
, “
Heat and mass transfer between impinging gas jets and solid surfaces
,”
Adv. Heat Transfer
13
,
1
60
(
1977
).
5.
N.
Zuckerman
and
N.
Lior
, “
Jet impingement heat transfer: Physics, correlations, and numerical modeling
,”
Adv. Heat Transfer
39
,
565
631
(
2006
).
6.
J. P.
Bouchez
and
R. J.
Goldstein
, “
Impingement cooling from a circular jet in a cross flow
,”
Int. J. Heat Mass Transfer
18
(
6
),
719
730
(
1975
).
7.
D. E.
Metzger
and
R. J.
Korstad
, “
Effects of crossflow on impingement heat transfer
,”
J. Eng. Power
94
(
1
),
35
41
(
1972
).
8.
R. J.
Goldstein
and
A. I.
Behbahani
, “
Impingement of a circular jet with and without cross flow
,”
Int. J. Heat Mass Transfer
25
(
9
),
1377
1382
(
1982
).
9.
Y.
Huang
,
S. V.
Ekkad
, and
J. C.
Han
, “
Detailed heat transfer distributions under an array of orthogonal impinging jets
,”
J. Thermophys. Heat Transfer
12
,
73
79
(
1998
).
10.
L.
Yang
,
J.
Ren
,
H.
Jiang
, and
P.
Ligrani
, “
Experimental and numerical investigation of unsteady impingement cooling within a blade leading edge passage
,”
Int. J. Heat Mass Transfer
71
,
57
68
(
2014
).
11.
L.
Wang
,
B.
Sundén
,
A.
Borg
, and
H.
Abrahamsson
, “
Heat transfer characteristics of an impinging jet in crossflow
,”
J. Heat Transfer
133
(
12
),
122202
(
2011
).
12.
S.
Debnath
,
Z. U.
Ahmed
,
M.
Ikhlaq
, and
T.
Khan
, “
Thermal characteristics of arrays of swirling impinging jets: Effect of Reynolds number, impingement distance, and jet-to-jet separation
,”
Heat Transfer
52
,
585
608
(
2023
).
13.
G.
Failla
,
E. H.
Bishop
, and
J. A.
Liburdy
, “
Enhanced jet impingement heat transfer with crossflow at low Reynolds numbers
,”
J. Electron. Manuf.
09
(
02
),
167
178
(
1999
).
14.
D.
Thibault
,
M.
Fénot
,
G.
Lalizel
, and
E.
Dorignac
, “
Experimental study of heat transfer from impinging jet with upstream and downstream crossflow
,”
Heat Transfer Res.
41
(
8
),
889
900
(
2009
).
15.
N.
Madhwesh
,
S.
Purohit
, and
K. V.
Karanth
, “
Computational flow and heat transfer study on impingement cooling in a turbine blade leading edge using an innovative convergent nozzles
,”
J. Braz. Soc. Mech. Sci. Eng.
45
,
86
(
2023
).
16.
D.
Yang
,
M.
Qiu
,
H.
Wu
,
Y.
Li
,
Z.
Jiang
, and
K.
Huang
, “
Temperature uniformity characteristics of array jet impingement cooling with the maximum cross-flow scheme
,”
Int. J. Therm. Sci.
187
,
108161
(
2023
).
17.
Y.
Masip
,
A.
Campo
, and
S. M.
Nuñez
, “
Experimental analysis of the thermal performance on electronic cooling by a combination of cross-flow and an impinging air jet
,”
Appl. Therm. Eng.
167
,
114779
(
2020
).
18.
A. R.
Salem
,
I.
Soliman
, and
R. S.
Aman
, “
Crossflow minimization for jet impingement cooling applications
,” in
ASME Turbo Expo 2023: Turbomachinery Technical Conference and Exposition
,
Boston
,
June 26–30, 2023
.
19.
M.
Chaudhari
,
B.
Puranik
, and
A.
Agrawal
, “
Heat transfer analysis in a rectangular duct without and with cross-flow and an impinging synthetic jet
,”
IEEE Trans. Compon. Packag. Technol.
33
(
2
),
488
497
(
2010
).
20.
S.
Yadav
and
R. P.
Saini
, “
Numerical investigation on the performance of a solar air heater using jet impingement with absorber plate
,”
Sol. Energy
208
,
236
248
(
2020
).
21.
M. A. R.
Sharif
and
K. K.
Mothe
, “
Evaluation of turbulence models in the prediction of heat transfer due to slot jet impingement on plane and concave surfaces
,”
Numer. Heat Transfer, Part B
55
(
4
),
273
294
(
2009
)..
22.
J. W.
Baughn
and
S.
Shimizu
, “
Heat transfer measurement from a surface with uniform heat flux and an impinging jet
,”
J. Heat Transfer
111
,
1096
1098
(
1989
).
23.
N.
Gao
,
H.
Sun
, and
S.
Edwig
, “
Heat transfer to impinging round jets with triangular tabs
,”
Int. J. Heat Mass Transfer
46
,
2557
2569
(
2003
).
24.
V. M.
Jose
, “
Performance assessment of jet impingement cooling with cross flow
,”
IEEE Trans. Compon. Packag. Manuf, Technol.
12
(
6
),
948
955
(
2022
).